U.S. flag

An official website of the United States government

NCBI Bookshelf. A service of the National Library of Medicine, National Institutes of Health.

WHO Regional Office for Europe. Review of evidence on health aspects of air pollution – REVIHAAP Project: Technical Report [Internet]. Copenhagen: WHO Regional Office for Europe; 2013.

Cover of Review of evidence on health aspects of air pollution – REVIHAAP Project

Review of evidence on health aspects of air pollution – REVIHAAP Project: Technical Report [Internet].

Show details

CProximity to roads, NO2, other air pollutants and their mixtures

Question C1

There is evidence of increased health effects linked to proximity to roads. What evidence is available that specific air pollutants or mixtures are responsible for such increases, taking into account co-exposures such as noise?

Answer

Motor vehicles are a significant source of urban air pollution. Adverse effects on health due to proximity to roads were observed after adjusting for socioeconomic status and after adjusting for noise. Elevated health risks associated with living in close proximity to roads is unlikely to be explained by PM2.5 mass since this is only slightly elevated near roads. In contrast, levels of such pollutants as ultrafine particles, carbon monoxide, NO2, black carbon, polycyclic aromatic hydrocarbons, and some metals are more elevated near roads. Individually or in combination, these are likely to be responsible for the observed adverse effects on health. Current available evidence does not allow discernment of the pollutants or pollutant combinations that are related to different health outcomes, although association with tailpipe primary PM is identified increasingly.

Exhaust emissions are an important source of traffic-related pollution, and several epidemiological and toxicological studies have linked such emissions to adverse effects on health. Road abrasion, tyre wear and brake wear are non-exhaust traffic emissions that become relatively more important with progressive reductions in exhaust emissions. Toxicological research increasingly indicates that such non-exhaust pollutants could be responsible for some of the observed adverse effects on health.

Rationale

In 2010, the Health Effects Institute published their authoritative report Traffic-related air pollution: a critical review of the literature on emissions, exposure, and health effects, which formed the basis of the current assessment. Motor vehicles emit large quantities of carbon dioxide, carbon monoxide, hydrocarbons, nitrogen oxides, PM, and substances known as mobile source air toxics, such as benzene, formaldehyde, acetaldehyde, 1,3-butadiene and lead (where leaded gasoline is still in use). Furthermore, secondary by-products, such as ozone and secondary aerosols (for example, nitrates and inorganic and organic acids), are formed farther away from roads, but these are not considered here.

Pollutant emissions from vehicles are related to vehicle type (such as light- or heavy-duty vehicles) and age, operating and maintenance conditions, exhaust treatment, type and quality of fuel, wear of parts (such as tyres and brakes), and engine lubricants used. Important non-combustion PM emissions associated with motor vehicles include wear particles from road surfaces, tyres and brakes, as well as resuspended road dust.

Non-combustion emissions contain such chemical compounds as trace metals and organics. Traffic emissions are the principal source of intra-urban variation in the concentrations of air pollutants in many cities, but this can vary both by time and location.

The Health Effects Institute report summarized that measurements of outdoor air quality on roadways indicate that concentrations of ultrafine particles, black carbon, particle-bound polycyclic aromatic hydrocarbons, nitric oxide, NO2, carbon monoxide, benzene, and formaldehyde are high and variable compared with ambient concentrations measured at background locations. Furthermore, concentrations around roadways may represent direct influences from road traffic and from background concentrations. The concentration gradient also seems to be a function of the reactivity of specific pollutants, such as NO2, nitrogen oxides and ozone. Hitchins et al. (2000) reported a 50% decrease in PM2.5 and ultrafine particles within 100–150 m of a road. A decay to background concentrations within as little as 50 m has been described for PM2.5 mass concentration (Tiitta et al., 2002), although PM2.5 tends to be more spatially homogeneous than ultrafine particles. Roorda-Knape et al. (1998) found that concentrations of black smoke, PM2.5, NO2, and benzene decreased to background concentrations within 100–150 m of a roadway (Roorda-Knape et al., 1998).

In an environment with greater volumes of traffic, Zhu et al. (2002) found that ultrafine particles, black carbon, and total PM counts decreased rapidly in the first 150 m and then levelled off. PM2.5 was found to be elevated only modestly (that is, in the range of 20%) near roadways. Zhu et al. (2006) suggested that distance-decay gradients extend to at least 500 m on the downwind side during night-time hours. Some studies concurrently measured such pollutants as NO2 and volatile organic compounds (Roorda-Knape et al., 1998; Weisel et al., 2005) and carbon monoxide (Zhu et al., 2002; Zhang et al., 2005), to assess pollutant mix. Zhu et al. (2002) found that the decay of concentrations with distance on the downwind side of a highway was similar for ultrafine particles, black carbon and carbon monoxide – that is, a 60% to 80% decrease from roadside concentrations within 100 m. Gilbert et al. (2003) also found that NO2 concentrations decayed with distance around a busy highway in Montreal, the greatest decrease occurring within the first 200 m.

In general, distance-decay gradients have different characteristics on upwind and downwind sides of an expressway (Roorda-Knape et al., 1998; Zhu et al., 2002; Gilbert et al., 2003; McConnell et al., 2006b). On the upwind side, concentrations drop off to near background levels within 200 m and, in the case of particles, probably within 100 m or less. On the downwind side, concentrations do not generally reach background levels until 300–500 m. In some studies, this was extended to up to 1500 m for NO2 (Gilbert et al., 2003; Jerrett et al., 2007) and 800 m for ultrafine particle number counts (Reponen et al., 2003).

Zhou & Levy (2007) pooled estimates from more than 30 studies and characterized the decay with distance from the road source for various combinations of reactive and nonreactive pollutants in areas of either high or low background pollution. Further simulations, using dispersion models, were employed to augment the empirical results. Overall, the distance-decay gradients demonstrated a heterogeneity that could be explained by background concentrations, pollutant characteristics, and local meteorological conditions (such as wind speed). Based on dispersion simulations for elemental carbon, the distance-decay gradient was in the range of 100–400 m from the source. For ultrafine particle counts, the gradient was 100–300 m; NO2 had gradients of 200–500 m. Also, metals (Peachey et al., 2009) and polycyclic aromatic hydrocarbons (Schnelle-Kreis et al., 1999) have shown a distance-decay gradient for roads. While this chapter was being prepared, Karner, Eisinger & Niemeier (2010) published a systematic compilation of the proximity measurements of multiple pollutants classified by category, which is a useful addition to this discussion.

In conclusion, there are a number studies showing higher levels of pollutants in proximity to roads. In general PM2.5 does not exhibit the sharp distance-decay gradient evident for carbon monoxide, NO2 or ultrafine particles. The Health Effects Institute Panel identified an exposure zone within a range of up to 300–500 m from a highway or a major road as the area most highly affected by traffic emissions – the range reflecting the variable influence of background pollution concentrations, meteorological conditions, and season. Metals usually attributed to brake and tyre wear, with such metals as copper, iron, antimony, tin, barium and zinc being higher close to roadways, compared with urban background (Querol et al., 2007). These metals were previously only seen in industrialized areas (Lee, Garland & Fox, 1994). Importantly, Ostro et al. (2011) found association between PM2.5 road dust and mortality.

Many studies have shown excess health risks in proximity to roads – after adjustment for a range of possible confounders, including socioeconomic status – for such outcomes as: cardiovascular mortality (Gehring et al., 2006), respiratory mortality and traffic intensity in a 100-m buffer (Beelen et al., 2008a), myocardial infarction (Tonne et al., 2007), cardiovascular disease (Hoffmann et al., 2006), coronary artery calcification (Hoffmann et al., 2007), cardiac function-left ventricular mass index (van Hee et al., 2009), asthma (Morgenstern et al., 2007, 2008; Gauderman et al., 2005; McConnell et al., 2006a; Gordian, Haneuse & Wakefield, 2006; Kim et al., 2008), wheeze (McConnell et al., 2006a; Ryan et al., 2005; Venn et al., 2005; Gauderman et al., 2005; van Vliet et al., 1997), asthma hospitalization (Edwards, Walters & Griffiths, 1994; English et al., 1999; Lin et al., 2002; Wilhelm et al., 2008), lung function reduction (Sekine et al., 2004; Kan et al., 2007; Gauderman et al., 2007; Schikowski et al., 2007), birth weight (Brauer et al., 2008), childhood cancer (Savitz & Feingold, 1989; Pearson, Wachtel & Ebi, 2000), and lung cancer (Beelen et al., 2008b). Therefore, the observed excess risk in proximity to roads cannot solely be explained by socioeconomic status; although associations between traffic proximity and health impacts have been observed in locations where both high and low socioeconomic status occur in close proximity to roads (Généreux et al., 2008), its influence cannot be ruled out.

Some studies have examined the effects of air pollution and noise at the same time. Those who have done so found that excess risks of air pollution in the proximity of roads generally remained after adjustment for noise for cardiovascular mortality (Beelen et al., 2008a; Gan et al., 2012), hypertension and diabetes mellitus (Coogan et al., 2012), hypertension (Fuks et al., 2011; Sørensen et al., 2012), and cognitive performance of primary schoolchildren (van Kempen et al., 2012). Therefore, these studies show effects of air pollution that cannot be explained by noise.

Generally, few epidemiological studies have examined the health effects of multiple air pollutants in proximity to roads. For those studies that have examined multiple air pollutants, it is not clear whether or not these pollutants are coming solely from roads and/or traffic or not. Some studies have examined the effect of multiple pollutants in the proximity of roads, but their small number, the generally high correlation among different pollutants, and the inconsistent results do not provide a good basis to draw firm conclusions.

The only epidemiological study identified that evaluates the short-term effects of multiple air pollutants in proximity to roads and farther away is by Roemer & van Wijnen (2001). These investigators obtained data from a sample of Amsterdam residents (n = 4352) who lived “along roads with more than 10,000 motorized vehicles per day” (actual distance from the roads not specified) from 1987 to 1998, and these were compared with the general population. Ambient-pollutant data from “traffic-influenced” sites and “non-influenced” sites (criteria not specified) were obtained for black smoke, PM10, and gaseous pollutants (carbon monoxide, NO2, SO2 and ozone). They found higher levels of NO2, nitric oxide, carbon monoxide and black smoke at the traffic-influenced measurement sites compared with the background sites, confirming combustion engines as the source of these air pollutants. Black smoke and NO2 were associated with mortality (RR: 1.38 and 1.10, respectively, for an increase of 100 µg/m3 on the previous day). Effect estimates were larger in the summer and in the population living along busy roads. Only 10% of the total Amsterdam population resides along busy roads. Nevertheless, they were still able to show associations between black smoke, NO2, and daily mortality for this subpopulation. These associations were stronger than they were in the total population.

Other studies have examined the effects of multiple pollutants in the proximity of roads for respiratory health and allergic disease outcomes (Brunekreef et al., 1997; van Vliet et al., 1997; Nicolai et al., 2003; Kim J et al., 2004; Gauderman et al., 2005; Morgenstern et al., 2008; Rosenlund et al., 2009a; McConnell et al., 2010; Gehring et al., 2010; Clark et al., 2010; Gruzieva et al., 2012, 2013; Schultz et al., 2012; Willers et al., 2013), birth weight (Brauer et al., 2008), pre-eclampsia and preterm birth (Wu et al., 2011), fatal myocardial infarction (Rosenlund et al., 2006, 2009b), lung cancer (Nyberg et al., 2000) and mortality (Beelen et al., 2008a). However in their analyses these investigators used either a proximity to road measure or a specific pollutant(s), but never the effects of (multiple) pollutants within proximity to roads. Even so, and assuming that often the population may have been near roads, no consistent picture emerged that specific pollutants and/or a mixture may be responsible for the observed health effects.

COMEAP recently concluded that the epidemiological evidence for associations between ambient levels of air pollutants and asthma prevalence at a whole community level was unconvincing; a meta-analysis confirmed a lack of association (Gowers et al., 2012). In contrast, a meta-analysis of cohort studies found an association between asthma incidence and within-community variations in air pollution (largely traffic dominated). Similarly, a systematic review suggested an association between asthma prevalence and exposure to traffic, although only in those living very close to heavily trafficked roads carrying many trucks, suggesting a possible role for diesel exhaust.

A critical review of the literature on the health effects of traffic-related air pollution (HEI, 2010b) included toxicological evidence of the impact of traffic-mixture exposures. Such evidence stems from controlled exposures of animals in areas of high traffic density, real-world exposure design in which subjects spend time in a polluted location (compared with equivalent activities in a location with relatively clean air), and individuals occupationally exposed to traffic and populations (animals or human beings) naturally exposed to polluted urban environments.2 Of the small number of studies reported (compared with the much larger literature on specific components of traffic emissions), the main cardiorespiratory findings in humans were that short-term exposures can bring about decrements in lung function and enhanced responses to allergens in adult subjects with asthma (Svartengren et al., 2000; McCreanor et al., 2007), as well as positive and negative effects on vascular function in healthy subjects (Rundell et al., 2007; Bräuner et al., 2008). On-road animal studies, utilizing compromised or allergic rodents, observed mild pulmonary inflammation (Elder et al., 2004), significant alterations in lung structure and elastic properties (Mauad et al., 2008), and systemic inflammation and effects on vascular function and autonomic control of the heart (Elder et al., 2004, 2007). Effects on reproductive and neurological health – specifically, compromised sperm quality in toll booth employees (de Rosa et al., 2003) and neuropathological lesions in dogs exposed to high concentrations of ambient pollution in Mexico City (Calderón-Garcidueñas et al., 2002) – were interpreted with caution as a result of data limitations. Finally, observations of genotoxic effects were limited to one study that reported higher mutagenicity from total suspended particulates in an area with intense moving traffic than in an area with limited traffic (Bronzetti et al., 1997).

In a recent review of the adverse effects on health of black carbon, the WHO Regional Office for Europe (2012) evaluated the toxicological evidence of effects of diesel exhaust in controlled human exposure experiments. It concluded that there are not enough clinical or toxicological studies to allow an evaluation: of the qualitative differences between the health effects of exposure to black carbon or those of exposure to PM mass (for example, different health outcomes); of a quantitative comparison of the strength of the associations; or of (identifying) any distinctive mechanism of black carbon effects. The review of the results of all available toxicological studies suggested that black carbon (measured as elemental carbon) may not be a major directly toxic component of fine PM, but it may operate as a universal carrier of a wide variety of combustion-derived chemical constituents of varying toxicity to sensitive targets in the human body, such as the lungs, the body’s major defence cells and, possibly, the systemic blood circulation.

Recent noteworthy toxicological evidence on the effects of traffic-mixture exposures include increased respiratory symptoms, decreased peak expiratory flow and an inflammatory response in the upper airways in mild asthmatic adults exposed for 2 hours in a road tunnel (Larsson et al., 2010). Studies of acute (20 minutes to 2 hours) effects of real-life traffic exposure on healthy volunteers have been unremarkable and are limited to a small increase in the percentage of blood neutrophils (Jacobs et al., 2010), modest effects on peak flow, exhaled nitric oxide and airway resistance (Zuurbier et al., 2011a, b). A study by Strak et al. (2012) was specifically designed to evaluate the contribution of different pollutants. They increased exposure contrasts and reduced correlations among pollutants by exposing healthy volunteers at five different locations, including two traffic sites. Changes in particle number concentrations, NO2, and nitrogen oxides during five-hour exposures were associated with increased exhaled nitric oxide and impaired lung function. These associations were robust and insensitive to adjustment for other pollutants. PM mass concentration or other PM characteristics, including elemental carbon and trace metals, were not predictive of the observed responses. Results for several other health end-points, including markers of cardiovascular effects, have not yet been published.

Two toxicological studies have investigated acute cardiovascular health effects in volunteers with type 2 diabetes. Passengers on 90–110 minute car rides on a busy road demonstrated a decrease in high-frequency heart rate variability and an increase in the ratio of low-frequency to high-frequency components compared with pre-ride measurements (Laumbach et al., 2010). Chronic exposure to urban air pollution (in chambers 20 m from a street with heavy traffic in downtown Sao Paulo) exacerbates the susceptibility of low density lipoprotein to oxidation, atherogenesis and vascular remodelling in hyperlipidemic mice (Soares et al., 2009), and in Swiss mice it presents as coronary arteriolar fibrosis and elastosis (Akinaga et al., 2009).

Toxicological reproductive outcomes have been investigated in subjects occupationally exposed to traffic. Findings include abnormal sperm count, mobility and morphology (Guven et al., 2008) and a significantly higher percentage of spermatozoa with damaged chromatin and DNA fragmentation (Calogero et al., 2011) in toll-gate workers. In male traffic policemen, lower free testosterone (Sancini et al., 2011) and higher luteinizing hormone (Tomao et al., 2009) and follicle-stimulating hormone (Tomei et al., 2009) plasma levels were reported. Studies on female traffic police observed significantly higher plasma free testosterone (Tomei et al., 2008) and follicle-stimulating hormone levels during the proliferative phase of the menstrual cycle (Ciarrocca et al., 2011).

Evidence continues to accumulate on the role that oxidative stress has as a mechanism through which traffic-related air pollution causes adverse effects on human health. The validity of urinary excretion of 8-oxo-7,8-dihydro-2-deoxyguanosine (8oxodG) as a biomarker was recently demonstrated in a meta-analysis (Barbato et al., 2010). Oxidative damage to DNA and the formation of bulky adducts are two mechanisms by which traffic-related air pollution could lead to mutagenesis and, ultimately, cause cancer. Bulky DNA adducts have been detected among traffic-exposed workers (Palli et al., 2008) and – together with micronuclei – in cord blood after maternal exposures to traffic-related air pollution, suggesting that transplacental environmental exposures could induce DNA damage in neonates (Pedersen et al., 2009).

Ambient PM – particularly that derived from vehicles – has high oxidative potential (Kelly, 2003), and a clear increment in roadside particulate oxidative potential has been found that appears to be associated with metals arising from engine abrasion (iron, manganese and molybdenum) or brake wear (copper and antimony) (Schauer et al., 2006; Thorpe & Harrison, 2008). The roadside increments of particulate oxidative potential are significant and the metal components identified as determinants of this oxidative activity have established toxicity in human beings (Kelly et al., 2011). These results are potentially important as they highlight the contribution of traffic non-exhaust pollutants that are not regulated currently.

Question C2

Is there any new evidence on the health effects of NO2 that impact upon the current limit values? Are long-term or short-term limit values justified on the grounds that NO2 affects human health directly, or is it linked to other co-emitted pollutants for which NO2 is an indicator substance?

Answer

Many studies, not previously considered, or published since 2004, have documented associations between day-to-day variations in NO2 concentration and variations in mortality, hospital admissions, and respiratory symptoms. Also, more studies have now been published, showing associations between long-term exposure to NO2 and mortality and morbidity. Both short- and long-term studies have found these associations with adverse effects at concentrations that were at or below the current EU limit values, which for NO2 are equivalent to the values from the 2005 global update of the WHO air quality guidelines. Chamber and toxicological evidence provides some mechanistic support for a causal interpretation of the respiratory effects. Hence, the results of these new studies provide support for updating the 2005 global update of the WHO air quality guidelines (WHO Regional Office for Europe, 2006) for NO2, to give: (a) an epidemiologically based short-term guideline value; and (b) an annual average guideline value based on the newly accumulated evidence. In both instances, this could result in lower guideline values.

There is evidence of small effects on inflammation and increased airway hyperresponsiveness with NO2 per se in the range of 380–1880 µg/m3 (0.2–1.0 ppm). The evidence for these effects comes from chamber studies (under a broad range of exposure conditions, with exposure durations of 15 minutes to 6 hours, with some inconsistency in results), with more marked, consistent, responses observed from 1880 µg/m3 (1.0 ppm). New review reports suggest weak to moderate lung cell changes in animal studies at one-hour concentrations of 380–1500 µg/m3 (0.2–0.8 ppm). These concentration ranges are not far from concentrations that occur at roadsides or in traffic for multiple hours. The chamber studies examined small numbers of healthy or mildly asthmatic subjects, whereas the general population will include subjects who are more sensitive and may therefore experience more pronounced effects at lower concentrations.

The associations between NO2 and short-term health effects in many studies remain after adjustment for other pollutants. The pollutants used in the adjustments include PM10, PM2.5, and occasionally black smoke. This does not prove that these associations are completely attributable to NO2 per se, as NO2 in these studies may also represent other constituents (which have adverse effects on health) not represented by currently regulated PM metrics. As there is consistent short-term epidemiological evidence and some mechanistic support for causality, particularly for respiratory outcomes, it is reasonable to infer that NO2 has some direct effects.

It is much harder to judge the independent effects of NO2 in the long-term studies because, in those investigations, the correlations between concentrations of NO2 and other pollutants are often high, so that NO2 might represent the mixture of traffic-related air pollutants. In this case, chamber studies do not apply and toxicological evidence is limited. However, some epidemiological studies do suggest associations of long-term NO2 exposures with respiratory and cardiovascular mortality and with children’s respiratory symptoms and lung function that were independent of PM mass metrics. As with the short-term effects, NO2 in these studies may represent other constituents. Despite this, the mechanistic evidence, particularly on respiratory effects, and the weight of evidence on short-term associations are suggestive of a causal relationship.

Rationale

This question is particularly important at this time as, in Europe, decreasing annual trends in carbonaceous aerosols have generally been observed, probably reflecting the impact of the Euro 4 and 5 vehicle standards in reducing diesel PM emissions; however, NO2 has not been declining in the same way or has even been increasing. This may result in very different ratios of NO2 to organic and elemental carbon emissions (supplemental material available illustrating this difference is upon request) and changes in ratios of concentrations (see discussion). This change in ratio has implications for the interpretation of NO2 as a quantitative proxy for PM vehicle pollution and illustrates the need to understand the effects of NO2 per se.

The text below sets out the short-term epidemiological, chamber study and toxicological evidence and then the long-term epidemiological and toxicological evidence before integrating the evidence and discussing the implications for the guidelines. Although the uncertainty over the long-term guideline has greatest policy importance, the short-term evidence is considered first because it provides support for the plausibility of the long-term effects. It should be emphasized that it was not our remit to review all studies, just those published since 2004, using reviews by others where necessary. Not all areas – for example, birth outcomes – have been reviewed in detail, due to time constraints or lack of implications for the conclusions. Also, our remit was not to actually propose new air quality guidelines, just to advise whether the guidelines needed to be revised in the light of scientific evidence published since the last revision of the guidelines (WHO Regional Office for Europe, 2006). Further detailed consideration will be needed at the guideline setting stage.

1. Short-term guideline

1.1 Time series evidence

The time-series evidence on NO2 has increased since the 2005 global update of the WHO air quality guidelines. Since 2004 (the cut-off date for studies included in the last WHO review), 125 new peer-reviewed ecological time-series studies on NO2 have been published up to April 2011. These were identified using the Air Pollution Epidemiology Database (APED).3 Table 5 shows the total number of studies, according to health outcomes examined and the number of multicity studies available.

Table 5.. Summary of ecological time-series studies of NO2 in APED.

Table 5.

Summary of ecological time-series studies of NO2 in APED.

The new studies were conducted mainly in the WHO Western Pacific Region (which includes China), Europe, the United States and Canada. The majority used 24-hour average concentrations of NO2, measured mainly at urban background locations. A few studies published after the April 2011 cut-off for APED were included in the review, as they contain information on issues of relevance to the Question. These are not reflected in Table 5, but are discussed in the text which follows. A list of all time-series references considered in this project is available upon request. Only a selection of these references is discussed in the text below.

Given the crucial issue of understanding whether NO2 has direct adverse effects on health, emphasis has been placed on time-series studies that investigated whether associations of NO2 are robust to adjustment of concentrations of particles (and other pollutants). More than 65 time-series studies of NO2 that used two-pollutant and/or multipollutant statistical models are available – only a proportion of these included adjustment for a metric of PM, and these are discussed in the sections that follow.

Mortality

In the 2005 global update of the WHO air quality guidelines, WHO concluded that daily concentrations of NO2 are associated significantly with increased daily all-cause, cardiovascular and respiratory mortality within the range of concentrations studied; however, it also noted the reductions in the overall effect estimates in an important meta-analysis, following adjustment for PM (Stieb, Judek & Burnett, 2002, 2003). Since then, comprehensive reviews of the time-series literature on NO2 have emerged with similar conclusions – that is, the short-term associations of NO2 with mortality are suggestive of a direct effect, but there is some uncertainty about the causal nature of these associations (CARB, 2007; EPA, 2008b). In addition to these, a peer-reviewed research report of a comprehensive systematic review and meta-analysis of single pollutant model estimates (Anderson et al., 2007) reported that increases in NO2 concentrations (per 10 µg/m3, 24-hour averages) are associated with increases in all-cause mortality: 0.49% (95% CI: 0.38–0.60%) in all ages and 0.86% (95% CI: 0.50–1.22%) for those older than 65 years of age. Results for maximum 1-hour average concentrations of NO2 were lower: 0.09% (95% CI: -0.01–0.20%) and 0.15% (95% CI: 0.03–0.26%) in all-ages and for those older than 65 years of age, respectively. Increases in daily mortality, for all ages, for cardiorespiratory mortality (0.18% (95% CI: 0.08–0.27%), 24-hour average); cardiovascular mortality (0.34% (95% CI: 0.19–0.48%), maximum 1-hour average, 1.17% (95% CI: 0.82–1.53%), 24-hour average); and respiratory mortality (0.45% (95% CI: 0.21–0.69%), maximum 1-hour average, 1.76% (95% CI: 1.35–2.17%), 24-hour average) were also reported with NO2. Anderson et al. (2007) also compared multipollutant model estimates for NO2 from multicity studies and reported that consistent positive estimates for mortality (and hospital admissions) were found before and after adjustment for co-pollutants, with the size and precision of the estimates not being substantially reduced after such adjustment. The authors also concluded that these findings suggested that the short-term associations between NO2 and health outcomes were unlikely to be confounded by other pollutant measures.

Twenty-four time-series studies of mortality, which used two-pollutant and/or multipollutant models for NO2, have been published since the 2005 global update of the WHO air quality guidelines (Burnett et al., 2004; Dales et al., 2004; Kan, Jia & Chen, 2004a,b; Zeka & Schwartz, 2004; Simpson et al., 2005a; Díaz, Linares & Tobías, 2006; Samoli et al., 2006; Brook et al., 2007; Qian et al., 2007, 2010; Yamazaki et al., 2007; Chen et al., 2008; Hu et al., 2008; Ren Y et al., 2008c; Wong et al., 2008; Breitner et al., 2009; Chen et al., 2010a; López-Villarrubia et al., 2010; Park, Hong & Kim, 2011; Chiusolo et al., 2011; Chen et al., 2012a,b; Faustini et al., 2012; Chen et al., 2013).4 All of these papers included adjustments for a metric of PM; 17 of the 24 papers reported positive, though not always statistically significant, short-term associations of NO2 with mortality for a range of diagnoses and age groups, after adjustment for a PM metric.

Multicity studies (of the aforementioned 24 studies), which included adjustment for particles in two-pollutant models, show robust short-term associations of NO2 with increased all-cause, cardiovascular and respiratory mortality (Table 6), though some evidence of confounding (by black smoke) of the NO2 association with respiratory mortality was identified in the European study, APHEA-2.5 Mainly PM10 was used when controlling for particles in these multicity studies. In contrast, the large American multicity study, NMMAPS, did not find such associations between NO2 and daily mortality: only a small and non-significant association of NO2 with all-cause mortality was reported after adjustment in a two-pollutant model with PM10; no association was found in a multipollutant model, which included PM10, carbon monoxide, SO2 and ozone (Zeka & Schwartz, 2004; see Table 6). Previous NMMAPS analyses of a subset of the 90 cities in the United States in this study found approximately a 0.7% (95% CI: 0.3–1.2%)6 increase in all-cause mortality per 45.12 µg/m3 (24 ppb)7 NO2 at lag 1 (other lagged model results showed that the strongest association between NO2 and all-cause mortality was identified at lag 1) (HEI, 2003). Following adjustment for PM10 or other pollutants (O3 and SO2), the central estimates either increased or were unchanged (based on a plot of the estimates), though they lost statistical significance (HEI, 2003). The reason for the difference between NMMAPS and the other multicity studies is unclear. We note that the method used by Zeka & Schwartz (2004) differed from those used in the other studies, as it sought to deal with measurement error. The shape of the concentration–response function in the multicity studies was often assumed to be linear. Samoli et al. (2006) reported that their assumption of a linear relationship between maximum 1-hour concentrations of NO2 and mortality was based on other results from APHEA-2 (Samoli et al., 2003), which suggested that it was appropriate to make such an assumption. Where tested, the relationship between NO2 and all-cause mortality did in fact appear to be linear (Wong et al., 2008; Chen et al., 2012b).

Table 6.. Summary of two-pollutant model results for NO2 from multicity time-series and case-crossover studies of mortality.

Table 6.

Summary of two-pollutant model results for NO2 from multicity time-series and case-crossover studies of mortality.

Hospital admissions and emergency room visits

Papers on hospital admissions for various diagnoses, published since the 2005 global update of the WHO air quality guidelines, have been identified (an overview of which is given in Table 5). Only studies of hospitalization or emergency room visits for respiratory (all-respiratory, asthma and chronic obstructive pulmonary disease) and cardiovascular and/or cardiac diagnoses have been considered. Many of the new papers have been subject to extensive review – for example Anderson et al, 2007; CARB, 2007; EPA, 2008b) – and therefore have not been reviewed individually. Only studies published since (or not included in) the most recent review – that is, by the EPA (2008b) – are discussed below. A list of all time-series references on NO2, including those on hospital admissions and emergency room visits, is available upon request.

Respiratory hospital admissions and emergency room visits
All respiratory diagnoses and asthma

In the 2005 global update of the WHO air quality guidelines, WHO concluded that the evidence suggested an effect of NO2 on respiratory hospital admissions and emergency room visits, especially for asthma.

This review consists of 41 new papers on respiratory (all-respiratory, asthma and chronic obstructive pulmonary disease) hospital admissions and emergency room visits. Four papers, based on multicity studies, were available (Eilstein et al., 2004; Barnett et al., 2005; Simpson et al., 2005b; Colais et al., 2009).

In 2008, the EPA concluded that there were positive short-term associations of NO2 with increased respiratory hospital admissions and emergency department visits, especially for asthma. The associations were noted to be particularly consistent among children and older adults (more than 65 years of age) for all respiratory diagnoses, and among children and all age groups for asthma admissions. The associations with NO2 were regarded as being generally robust to adjustments for particles and gaseous pollutants. Many of the studies considered in the EPA review used a 24-hour measure of exposure, with a number of them reporting mean concentrations of NO2 within the range of 5.6–94.0 µg/m3 (maximums of 52.6–154.2 µg/m3). Similar conclusions were also reported by the California Environmental Protection Agency Air Resources Board (CARB, 2007). The meta-analysis of single-pollutant model estimates by Anderson et al. (2007) supports the conclusions by the EPA and CARB, reporting positive and mainly statistically significant overall estimates (percentage increase per 10 µg/m3) for respiratory hospital admissions and 24-hour concentrations: 1.80% (95% CI: 1.15–2.45%), 0.82% (95% CI: 0.35–1.29), 1.47% (95% CI: 0.10–2.87), and 0.48% (95% CI: -0.35–1.31) for all ages, children, young adults, and the elderly, respectively. Overall effect estimates for asthma hospital admissions were 1.37% (95% CI: 0.59–2.15) and 2.92% (95% CI: 1.15–4.72) for 24-hour average NO2 in all ages and children, respectively. Pooled estimates for these health outcomes for maximum 1-hour average concentrations of NO2 were smaller than those for 24-hour average concentrations (Anderson et al., 2007).

Papers published since (or not considered in) the EPA’s 2008 review also reported positive (though not always statistically significant) single-pollutant associations for hospital admissions and emergency room visits for: (a) asthma (Bell, Levy & Lin (2008), Colais et al., 2009; Giovannini et al., 2010; Halonen et al., 2008; Jalaludin et al., 2008; Samoli et al., 2011a; Szyszkowicz, 2008; Ueda, Nitta & Odajima, 2010; Villeneuve et al., 2007) and (b) respiratory causes (Colais et al., 2009; Eilstein et al., 2004; Faustini et al., 2013; Granados-Canal et al., 2005; Giovannini et al., 2010; Jayaraman & Nidhi, 2008; Thach et al., 2010; Vigotti et al., 2010).8 Serinelli et al. (2010) and Kim, Kim & Kim (2006) reported negative associations between NO2 and respiratory and asthma hospital admissions, respectively; Kim, Kim & Kim (2006) also found a positive association between NO2 and emergency visits for asthma. The associations reported in the new papers published since EPA (2008b) are based largely on 24-hour average concentrations of NO2 measured at urban background sites. Very few studies used 1-hour measures.

Where confounding by co-pollutants was assessed in the studies since the EPA (2008b) review, some robustness to adjustment was demonstrated for these respiratory health outcomes (Giovannini et al., 2010; Ueda, Nitta & Odajima, 2010; Halonen et al., 2008; Jalaludin et al., 2008; Jayaraman & Nidhi, 2008; Villeneuve et al., 2007). For example, Halonen et al. (2008) found that NO2 was a strong and independent predictor of asthma emergency room visits in children in Finland. The authors examined a range of single day lags from lag 0 to lag 5, and found that for lags of 3–5 days, all particle fractions below 250 nm, NO2 and carbon monoxide were each associated with asthma emergency room visits in children. In two-pollutant model analyses, the association with ultrafine particles (for an interquartile range increase in concentration) was removed after adjustment for NO2 – from 6.6% (95% CI: 2.34–11.00%) to -0.89% (95% CI: -6.11–4.62%) at lag 4. Although the NO2 estimates were not shown in the paper, the authors reported that they were not sensitive to adjustment for other pollutants (PM2.5, nucleation and Aitken mode particle sizes,9 coarse particles and carbon monoxide). These pollutants were not too highly correlated: the highest correlation (0.65) was between NO2 and ultrafine particles. Giovannini et al. (2010) reported estimates for asthma and all-respiratory conditions in children that, respectively, increased (from a RR of 1.002 (1.000–1.004) to a RR of 1.004 (0.993–1.015)) or were negligibly affected (from RR of 1.009 (1.001–1.017) to 1.008 (0.999–1.016)) by adjustment for carbon monoxide. Ueda, Nitta & Odajima (2010) also reported an increase in the estimate for asthma hospitalisation in children (from an odds ratio of 1.112 to an odds ratio of 1.128) following adjustment in a multi-pollutant model, though this lost statistical significance.

Iskandar et al. (2012) also reported statistically significant associations of NO2 with asthma hospital admissions in children in Copenhagen10 following adjustment for several PM metrics. The association was slightly attenuated and remained statistically significant following adjustment for PM10; it increased following adjustment for PM2.5 and ultrafine particles; but was reduced and lost statistical significance following adjustment for nitrogen oxides. With the exception of ultrafine particles, associations with PM10 and PM2.5 and asthma hospital admissions in children remained positive (though attenuated) and statistically significant following adjustment for NO2. Leitte et al. (2011) examined relationships between NO2 and various particle sizes and respiratory emergency room visits in Beijing, China. In two-pollutant model analyses, the most consistent associations were found with NO2 (adjusted for PM10).

Results from APHEA-2 for PM10 and hospital admissions show that after adjustment for NO2, the estimate for asthma in 0–14-year-olds is reduced from a statistically significant increase of 1.2% to 0.1% (95% CI: -0.8–1.0%) per 10 µg/m3 PM10 (Atkinson et al., 2001); similar findings for the 15–64-year age group were reported (reduced from 1.1% to 0.4% (95% CI: -0.5–1.3%). These findings suggest that NO2 had a stronger association with asthma admissions than did PM10.

Not all studies demonstrated robustness. For example, Samoli et al. (2011a) reported a positive (1.10%, per 10 µg/m3 increase in NO2), but a statistically insignificant single-pollutant model association for childhood asthma admissions in Athens. This association was reduced after controlling for PM10 (estimate reduced to 0.54%) and SO2 (estimate reduced to -0.78%) in two-pollutant models. Corresponding results for PM10 showed a small reduction in the central estimate (from 2.54% to 2.28%) and loss of statistical significance after adjustment for NO2. This was a small study, covering 3 years with a median of just two asthma admissions daily. Chen et al. (2010b) did not use two-pollutant models to investigate relationships with respiratory hospital admissions, as no statistically significant associations between NO2 (or PM10 and SO2) and respiratory hospital admission were found in single-pollutant models.

Chronic obstructive pulmonary disease

A total of 13 papers on hospital admissions for chronic obstructive pulmonary disease are available for review. (It should be noted that some of these analysed chronic obstructive pulmonary disease and asthma together.) Eight of the papers formed part of the EPA’s 2008 review, in which they concluded that the limited evidence did not support a relationship between NO2 and admissions for chronic obstructive pulmonary disease. The few remaining papers reported positive and statistically significant associations (Sauerzapf, Jones & Cross, 2009; Thach et al., 2010; Colais et al., 2009), or positive (many insignificant) or negative associations (Halonen et al., 2008; 2009) between NO2 and chronic obstructive pulmonary disease admissions or emergency room visits. These associations were based on single-pollutant models.

Overall, for respiratory outcomes in general, the new studies continue to provide evidence of short-term associations between NO2 and respiratory hospital admissions and emergency room visits, especially for asthma. Many studies have demonstrated that these associations are not confounded by co-pollutants, including PM10 and common gaseous pollutants typically used in two- pollutant and/or multipollutant model analyses. The few data available do not allow firm conclusions to be made about the robustness of these associations to adjustment for ultrafine particles.

Cardiovascular and/or cardiac hospital admissions and emergency room visits

Twenty-two papers on hospital admissions or emergency room visits for cardiovascular and/or cardiac diagnoses (mainly in all ages and the elderly) formed part of this review. The EPA (2008b) reviewed 12 of these, with 10 reporting estimates from two-pollutant and/or multipollutant models. In 2008, the EPA concluded that while positive short-term associations between NO2 and hospital admissions or emergency visits to hospital for cardiovascular-related disorders were identified (at mean 24-hour concentrations in the range 27.6–75.2 µg/m3), most of these were diminished in multipollutant models that also contained carbon monoxide and PM. The 2005 global update of the WHO air quality guidelines concluded that, although positive associations between NO2 and admissions or visits to hospital for cardiovascular and/or cardiac diagnoses had been reported, drawing conclusions about the nature of the relationship was made less clear, since controlling for other pollutants at times lowered the effect estimates and at other times made them lose statistical significance.

The new studies published since (or not considered by) the EPA continue to show positive single-pollutant model associations: Eilstein et al. (2004) in nine French cities; Cakmak, Dales & Judek (2006a) in ten Canadian cities; Larrieu et al. (2007) in eight French cities; Chan et al. (2008) in Taipei, Taiwan, Province of China; Lefranc et al. (2009) in eight French cities; Colais et al. (2009) in nine Italian cities; and Alves, Scotto & Freitas (2010) in Lisbon, Portugal. Although Serinelli et al. (2010) also reported positive associations between NO2 and cardiac hospital admissions for several lagged models, these were all statistically insignificant.

Only three of the new studies published since (or not considered by) the EPA review used two-pollutant and/or multipollutant models. A multicity Canadian study, by Cakmak, Dales & Judek (2006a) reported a positive and statistically significant association for cardiac hospital admissions (5.9% (95% CI: 3.2–8.6%), for a 40.23 µg/m3 concentration change) with NO2 after adjustment for carbon monoxide, SO2 and ozone. Using data from 58 urban counties in the United States for 1999–2005, Bell et al. (2009b) found a 1.30% (95% CI: 0.87–1.73%) increase in cardiovascular hospital admissions in the elderly per 17.7 µg/m3 interquartile range increase in NO2 after adjustment for carbon monoxide and PM2.5. The estimate for 9.8 µg/m3 interquartile range increase in PM2.5 was -0.18% (95% CI: -0.49–0.14%) after adjustment for NO2 and carbon monoxide. Chen et al. (2010b) also reported a robust association for cardiovascular hospital admissions following adjustment for PM10 (from 0.80% (95% CI: 0.10–1.49%) to 0.71% (95% CI: 0.00–1.41%)) but not for SO2 (reduced to 0.28% (95% CI: -0.76–1.32%)).

In summary, the new evidence continues to show positive associations between NO2 and hospital admissions and emergency room visits for cardiovascular and/or cardiac diagnoses. Given the mixed findings from the studies using (or reviewing) two-pollutant and/or multipollutant models published since the 2005 global update of the WHO air quality guidelines and given that none of the short-term studies of other cardiovascular-related endpoints – for example, heart failure, ischaemic heart disease – have been reviewed, it is difficult to comment further on the nature of the relationship between NO2 and hospital admissions or visits for cardiovascular and/or cardiac diseases. The evidence for positive associations could allow quantitative exploration in sensitivity analyses (see Question C4).

1.2 Panel studies

The short-term effects of NO2 on respiratory health in children with asthma were recently reviewed (Weinmayr et al., 2010). The review is based on 36 panel studies published 1992–2006, of which 14 are from the Pollution Effects on Asthmatic Children in Europe (PEACE) project, all from 1998. In the meta-analysis of these studies, NO2 showed statistically significant associations with asthma symptoms when considering all possible lags, but not when similar lags (0, 1 or 0–1) were evaluated from each study. The association with cough was statistically significant, but only when the PEACE studies were not considered (The PEACE study was negative. On the one hand it is the only multicentre study with a uniform protocol; on the other hand, concerns have been raised that the results were influenced by an influenza epidemic during the relatively short observation period (2 months)). The estimated effect on peak expiratory flow was statistically insignificant.

A PubMed search identified 11 new articles: 6 from the Americas (Sarnat et al., 2012; O’Connor et al., 2008; Escamilla-Nuñez et al., 2008; Liu et al., 2009; Castro et al., 2009; Dales et al., 2009), 2 from Europe (Andersen et al., 2008b; Coneus & Spiess, 2012), and 3 from Asia (Min et al., 2008; Yamazaki et al., 2011; Ma et al., 2008). All, except two, studies investigated school-aged asthmatic and/or symptomatic children, while the remaining two considered infants and toddlers (Andersen et al. 2008b; Coneus & Spiess, 2012). Most of the studies observed positive associations between short-term exposure to NO2 (or nitrogen oxides) for different lags and respiratory symptoms, such as coughing and wheezing (O’Connor et al., 2008; Escamilla-Nuñez et al., 2008; Andersen et al., 2008b), as well as for exhaled nitric oxide (Sarnat et al., 2012) and also for pulmonary function decrease (O’Connor et al., 2008; Liu et al., 2009; Castro et al., 2009; Dales et al. 2009; Min et al., 2008; Yamazaki et al., 2011; Ma et al., 2008) in children, although not all of them were statistically significant (Dales et al., 2009). One article reported numerical data only for PM (Min et al., 2008).

The largest of these studies included 861 children with asthma from seven inner-city communities in the United States (O’Connor et al., 2008). A difference of 20 ppb in the 5-day average concentrations of NO2 was associated with an odds ratio (OR) of 1.23 (95% CI: 1.05–1.44), for having a peak expiratory flow less than 90% of personal best. Similar risk elevations were found for decreased FEV1, cough, night-time asthma, slow play and school absenteeism. Multipollutant models showed independent effects for NO2 after adjustment for ozone and PM2.5 for most of these end-points.

In addition, some panel studies were published earlier, but were not included in the meta-analysis mentioned above (Svendsen et al., 2007; Delfino et al., 2006; Trenga et al., 2006; Moshammer et al., 2006; Mar et al., 2005; Rabinovitch et al., 2004; Ranzi et al., 2004; Boezen et al., 1999). These studies are not reviewed here, but are cited to illustrate, combined with the points made earlier, that a new combined analysis would contribute to a more precise quantitative estimation of the effects of short-term fluctuations in outdoor NO2 concentrations on changes in pulmonary function and respiratory symptoms in asthmatic children, given the increased number of studies available. This might be performed by enlarging the existing review (Weinmayr et al., 2010) with the recently published panel studies. A new combined analysis could also consider panel studies performed on indoor exposure and respiratory effects – for example, Marks et al. (2010).

This section has concentrated on panel studies of respiratory health in children. There are also studies on chronic obstructive pulmonary disease in adults (Peacock et al. (2011) found effects of NO2 weakened by control for PM10) and on cardiovascular end-points. Panel studies with end-points relevant to mechanistic questions are discussed in the toxicology section.

Of particular interest is a recent study conducted in the Netherlands that attempted to isolate which component of the ambient aerosol could be related to various acute response endpoints in subjects exposed at five separate microenvironments, of contrasting source profile: an underground train station, two traffic sites, a farm and an urban background site. To date, only the results of the associations with respiratory end-points have been published, but this data suggested that interquartile increases in particle number concentration and NO2 were related to decreased lung function (FVC and FEV1) – associations that were insensitive to adjustment for other gaseous pollutants and an extensive range of PM metrics, including black carbon, particle number concentration, oxidative potential and transition metals (Strak et al., 2012).

1.3 Chamber studies

Since 2005, the literature on human exposure to NO2 has undergone a number of systematic reviews, as part of the EPA integrated scientific assessment for oxides of nitrogen (EPA, 2008b) and later by Hesterberg et al. (2009) and Goodman J et al. (2009). The human chamber study evidence was also reviewed by the California Environmental Protection Agency Air Resources Board (CARB, 2007) in their assessment of the California NO2 standard. Since the publication of these reviews, limited NO2 chamber studies that address lung function and airway inflammation have been performed, and two studies that address cardiovascular end-points have been published (Langrish et al., 2010; Scaife et al., 2012). Therefore, the conclusions that arise from these reviews remain valid for consideration of the current NO2 standards.

The human chamber studies generally only address a single pollutant, but this can be helpful when it is unclear whether the health effects reported in the so-called real world are related to the pollutant itself, or a mixture of pollutants for which NO2 serves as a surrogate. This is particularly true for NO2, where there is an apparent mismatch between the human chamber exposure data (with mixed evidence of inflammation or impaired lung function between 0.26 ppm and 0.60 ppm) and the population-based epidemiology (with effects reported at lower concentrations within the ambient range). The results from the chamber studies are therefore useful in determining whether NO2 should be considered toxic to the population in its own right or should be considered a tracer for local source emissions – that is, traffic in urban areas.

Human clinical studies generally examine healthy subjects, or patients (asthmatics, chronic obstructive pulmonary disease patients, allergic rhinitis patients, and patients undergoing rehabilitation following cardiac events) with relatively stable symptoms exposed to a single NO2 concentration for durations of 1–6 hours. Results are generally compared against a control air exposure, with a range of end-points examined, including lung function, exhaled gases, airway inflammation (either by bronchoscopy or by induced sputum) and airway hyperresponsiveness. The results observed from these acute exposures are usually transient and, though often statistically significant, are seldom clinically so. Healthy subjects have been exposed acutely to NO2 concentrations that range from 0.3 ppm to 2.0 ppm (1–4 hours); sensitive subgroups, including asthmatics and chronic obstructive pulmonary disease patients have been exposed to 0.26–1.00 ppm. Above 1 ppm, clear evidence of inflammation has been observed in healthy subjects in a number of studies (Helleday et al., 1995; Devlin et al., 1999; Pathmanathan et al., 2003; Blomberg et al., 1999), but the picture is less established at concentrations between 0.2 ppm and 1.0 ppm (Vagaggini et al., 1996; Jörres et al., 1995; Gong et al., 2005; Frampton et al., 2002; Riedl et al., 2012), partially because of inconsistent responses in the varying end-points examined.

A number of studies have suggested that NO2 can augment allergen-induced inflammation in asthmatics (Barck et al., 2002, 2005) following short (15–30 minutes) consecutive day exposures at relatively low concentrations (0.26 ppm). However other studies at comparable NO2 concentrations, but at a higher total dose (0.4 ppm for 3 hours), have failed to demonstrate a similar enhancement of airway inflammation or a reduced house-dust-mite allergen provocation dose in allergic asthmatics (Witten et al., 2005). Similarly, Vagaggini et al. (1996) were unable to demonstrate NO2-induced upper airway inflammation in mild asthmatics, by nasal lavage and induced sputum 2 hours after an exposure to 0.3 ppm NO2 for 1 hour. The evidence for inflammation in asthmatics exposed to NO2 concentrations near environmental concentrations is therefore ambiguous.

Few studies have reported acute lung function changes near the current guideline concentration value (Bauer et al., 1986; Jörres et al., 1995) in the absence of a specific or nonspecific challenge (reviewed in Hesterberg et al., 2009; EPA, 2008b). In those studies that reported decrements, exposure concentrations were high (more than 1 ppm) and the magnitude of the observed changes was unlikely to be of clinical significance (Blomberg et al., 1999).

The reported effects of NO2 on nonspecific and specific airway hyperresponsiveness have been reviewed in detail in the EPA integrated science assessment (2008b), as well as by Goodman J et al. (2009) and Hesterberg et al. (2009). In healthy individuals, small increases in nonspecific airway hyperresponsiveness have been reported following short-term (1–3 hour) exposure to NO2 in the range 1.5–2.0 ppm (Mohsenin, 1987; Frampton et al., 1989). In asthmatics, the data on nonspecific hyperresponsiveness to NO2 suggests a sensitizing effect between 0.2 ppm and 0.6 ppm, though the responses, where significant, were small and unlikely to be of clinical significance (Bylin et al., 1985; Mohsenin, 1987; Strand et al., 1996). A limited number of studies have evaluated airway hyperresponsiveness in asthmatics at multiple NO2 concentrations, and the results of these studies do not support a clear concentration–response relationship between 0.1 ppm and 0.5ppm (Bylin et al., 1988; Roger et al., 1990; Tunnicliffe, Burge & Ayres, 1994). It should be noted that, in all of the studies cited, a range of responses were observed; and in many studies, subgroups of responders were identified. Given that chamber studies by their very nature rely on typically small numbers of healthy volunteers, or clinically stable patients, it is likely that they may underestimate the responses of sensitive subgroups within the population, especially in relation to disease severity.

Langrish et al. (2010) failed to demonstrate vascular dysfunction (vascular vasomotor or fibrinolytic function) in healthy volunteers exposed to 4 ppm NO2 or filtered air for 1 hour. This is an important paper, as this result contrasts markedly with the group’s previous findings using diesel exhaust containing high concentrations of NO2 (Mills et al., 2007). A subsequent paper, examining the vascular and prothrombic effects of diesel exhaust (300 µg/m3 for 1 hour), with and without inclusion of a particle trap, demonstrated that a reduction of particle number and mass concentration, in the absence of changes in nitrogen oxides, was associated with reduced adverse cardiovascular outcomes (Lucking et al., 2011). This was interpreted as strongly supporting the view that fine particles, and not NO2, were driving the previously reported cardiovascular effects, especially as NO2 concentrations increased almost five-fold with the particle trap. Further evidence, suggesting the absence of an acute cardiovascular effect of NO2, was reported by Scaife et al. (2012). In this study they found no significant changes in heart variability parameters in 18 heart bypass and myocardial infarction patients following exposure to 400 ppb NO2 for 1 hour.

1.4 Toxicological studies on short-term exposure (hours to days)

The WHO Regional Office for Europe indoor air quality guideline (2010) noted that acute exposures (hours) in the range of 0.04–1.0 ppm were rarely observed to cause effects in animals. The California Environmental Protection Agency Air Resources Board (CARB, 2007) noted acute effects (hours) in rats and mice at 0.2–0.8 ppm NO2 (increased mast cells, quoted from Hayashi & Kohno (1985) in CARB, 2007)*11 and increased synthesis of the carcinogen dimethylnitrosamine (Iqbal, Dahl & Epstein, 1981)* at 0.2 ppm. Also, there were: effects on liver detoxification enzymes at 0.25 ppm (Miller et al., 1980)*; effects on macrophages at 0.3–0.5 ppm (e.g. Robison et al., 1993)*; and increased bronchiolar proliferation at 0.8 ppm (Barth et al., 1994)*. The EPA (2008b) also highlighted the study by Barth et al. (1994). The effects at 0.2–0.25 ppm are harder to interpret (the mast cells may not have been degranulating, dimethylnitrosamine synthesis relied on an additional administered precursor and the effect on liver enzymes may not be adverse), but it is clear that the effects are adverse above 0.3 ppm. Tables in these reports quote other studies that report acute no effect levels from 0.4 ppm to 0.8 ppm.

A literature search, and literature abstracts already held from 2008 onwards (the literature cut-off for WHO Regional Office for Europe (2010)),12 did not indicate any animal studies with short-term exposure to NO2 alone that would change the CARB (2007) or EPA (2008b) view. Urea selective catalytic reduction-treated diesel engine emissions (0.78 ppm NO2 and dilutions) were generally less toxic to rat lungs than conventional diesel engine emissions (0.31 ppm NO2 and dilutions), for various endpoints over durations of 1, 3 or 7 days. However, there were differences in the nature of the oxidative stress produced with either increases in 8-hydroxy-2-deoxyguanosine with conventional diesel engine emissions, or increases in haemoxygenase-1 mRNA expression with urea selective catalytic reduction-treated diesel engine emissions (Tsukue et al., 2010). The latter suggests oxidative stress related to NO2, but is not conclusive, given the mixture of constituents present.

Using exposures lasting from hours to days, recent mechanistic animal studies show: protein S-glutathionylation in the lung at 25 ppm NO2 (Aesif et al., 2009); a decrease in aggregating activity of surfactant-protein D at 10 ppm or 20 ppm (Matalon et al., 2009); and increases in markers of oxidative stress, endothelial dysfunction, inflammation and apoptosis in the hearts of rats from exposures of 2.7 ppm to 10.6 ppm (Li H et al., 2011). Zhu et al. (2012) found that 2.6 ppm NO2 delayed recovery from stroke (slowed reduction in infarct volume) in a rat stroke model and increased behavioural deficits. Dose-related endothelial and inflammatory responses were also found in the range 2.6–10.6 ppm. Channell et al. (2012) found that plasma from healthy adults exposed to 0.5 ppm NO2 for 2 hours was able to activate cultured primary human coronary endothelial cells. This suggested that a circulating factor was mediating, at least in part, the endothelial response that may underly the cardiovascular epidemiological findings. One study in mice (Alberg et al., 2011) found no increased sensitization to intranasal ovalbumin at 5 ppm or 25 ppm NO2 when diesel exhaust particles did show an increase, whereas another study in mice (Hodgkins et al., 2010) at 10 ppm NO2 showed sensitization to inhaled ovalbumin due to increased antigen uptake by antigen-presenting cells. A study in mice suggested that effects in the lung due to 20 ppm NO2 for 10 days are worse with a small increase in vitamin C dose than with a large increase (Zhang et al., 2010), perhaps due to vitamin C increasing NO2 absorption into the epithelial lining fluid (Enami, Hoffmann & Colussi, 2009). All these concentrations (apart from Channell et al., 2012) are far in excess of the ambient concentrations linked to health effects in population studies, and the studies are not designed to show whether the mechanisms extend down to lower concentrations – that is, the mechanisms may or may not be relevant.

Recent studies indicate that the NO2 radical can be directly involved in nitration of tyrosines (Surmeli et al., 2010, for example) and also that it causes a nitration-dependent cis-trans-isomerization to trans-arachidonic acid (linked to microvascular injury), described by the authors as a characteristic process for NO2 (Balazy & Chemtob, 2008). This is not just relevant to the lung. While NO2 itself is not absorbed systemically, as it is likely to react first, its reaction products, nitrite and nitrate ions, are found in the blood after NO2 inhalation (Saul & Archer, 1983). Nitrite and nitrate anions in the blood are now regarded as carriers of nitric oxide (Lundberg, Weitzberg & Gladwin, 2008; Lundberg & Weitzberg, 2010; Weitzberg, Hezel & Lundberg, 2010).

The cycle whereby nitrate is excreted into the saliva and reduced to nitrite by oral bacteria and then converted to nitric oxide in the stomach or in the blood and tissues, after absorption of nitrite, may have a physiological role (to ensure nitric oxide production for vasodilation under hypoxic conditions when oxygen dependent nitric oxide synthases may fail), but may also have adverse consequences (Panesar, 2008). Nitrite can also lead to the NO2 radical via peroxidase catalysed oxidation with hydrogen peroxide, via formation of nitrous acid, which dissociates to nitric oxide and NO2, via formation of peroxynitrite from nitric oxide and superoxide and subsequent breakdown to the NO2 and hydroxyl radicals and, less commonly, direct oxidation of nitric oxide (d’Ischia et al., 2011; Signorelli et al., 2011). In other words, there is an indirect transfer of inhaled NO2 to the NO2 radical in tissues via reactive intermediates.

There is a great deal of literature on reactive nitrogen species (Abello et al., 2009; d’Ischia et al., 2011; Sugiura & Ichinose, 2011), including in literature on atherosclerosis (Upmacis, 2008). The review by Upmacis (2008) describes the occurrence of protein nitrotyrosine in atherosclerotic plaques and also in the bloodstream and describes an emerging view that this is a risk factor for cardiovascular disease. A major proportion of the nitrotyrosine is thought to come via nitric oxide from inducible nitric oxide synthase (induced under inflammatory conditions), but the source of the remainder is unknown. Nitrite and nitrate in the blood from NO2 inhalation provides a potential link with this literature. Human beings exposed to 0.1 ppm NO2 by inhalation have been inferred to form about 3.6 mg nitrite per day, more than the dietary intake of nitrite (Saul & Archer, 1983). More work is needed, however, to understand how significant any contribution of NO2 inhalation to systemic nitrative stress might be in quantitative terms.

Although this is a section on toxicology, some epidemiological studies are considered here as they concentrate on mechanisms. They provide a potential link between the mechanisms in animal studies discussed above and mechanisms operating in human beings, although they lose the advantage of confident specificity for NO2. These studies have mostly addressed cardiovascular end-points: no effects on blood coagulation and inflammatory markers (Zuurbier et al., 2011b; Steinvil et al., 2008) or a non-significant increase (significant for sP-selectin) (Delfino et al., 2008); increases in brachial artery diameter and flow mediated dilatation (perhaps actually due to nitric oxide) (Williams et al., 2012); adverse effects on heart rate variability in heart disease patients (Zanobetti et al., 2010) and in cyclists (Weichenthal et al., 2011) (independent of PM metrics); a non-significant increase in the QT interval in electrocardiograms (significant in diabetics) (Baja et al., 2010); and an increase in lipoprotein-associated phospholipase A2 (linked to inflammation in atherosclerotic plaques) (Brüske et al., 2011).

2. Long-term guideline

2.1 Long-term epidemiological studies

Since the publication of the 2005 global update of the WHO air quality guidelines, a large number of new studies on long-term effects related to NO2 and other traffic pollutants have been published. To prepare a response to the question on new evidence, systematic literature searches were performed for original articles published 2004 to April 2012 on long-term studies (cohort studies, cross-sectional studies, case-control studies) that investigated outdoor NO2 exposure and mortality or diagnosed diseases or lung function. We did not look at other physiological markers and at birth outcomes, as it is not clear whether they are really consequences of long-term exposure, and there were too few studies meeting the criteria. The searches in the literature databases PubMed, ISI Web of Science and Ludok resulted in 160 publications: on mortality or specific death causes (n = 37); on lung function (n = 34); on incidence or prevalence of cardiovascular diseases (n = 19); on diabetes mellitus (n = 6); on asthma (n = 46); and on bronchitis (n = 18).

In addition, two review reports from the California Environmental Protection Agency Air Resources Board and the EPA published in 2007 and 2008, respectively, have been reviewed.

  1. Review of the California Ambient Air Quality Standard for NO2, 2007. The following recommendation was released for a NO2 annual-average standard: establish a new annual average standard for NO2 of 56 µg/m3 (0.030 ppm), not to be exceeded. It was based on evidence from epidemiological studies showing that long-term exposures (that is, one or more years) to NO2 may lead to changes in lung function growth in children, symptoms in asthmatic children, and preterm birth (CARB, 2007).
  2. EPA integrated science assessment for oxides of nitrogen – health criteria, 2008. This document rated the epidemiological and toxicological evidence, examining the effect of long-term exposure to NO2 on respiratory morbidity, as suggestive, but not sufficient to infer a causal relationship. The document also rated the evidence on other morbidity or mortality as inadequate to infer the presence or absence of a causal relationship. It found the relationship between long-term exposure to NO2 and mortality to be inconsistent. Furthermore, when associations were noted, they were not specific to NO2, but also implicated PM and other traffic indicators.

Long-term studies generally investigate geographic differences in NO2 exposure for a period of several years. The spatial distribution of NO2 shows a high variation, as it varies mainly with traffic – the most important NO2 source in European countries. Therefore, individual assessment of exposure is much more important than the assessment of more homogeneously distributed PM2.5 or (often) PM10. So, the spatial resolution of the exposure assessment is crucial. In contrast to particle mass, black smoke (or black carbon) often shows a similar spatial distribution to NO2. The answer to the question about whether long-term effects are due to NO2 per se should ideally be based on studies that include NO2 together with a particle mass indicator (on the one hand) and an indicator such as black carbon or ultrafine particles (on the other hand).

Some of the newly published studies were ecological in nature (lacking individual covariates) or did not give numerical results for NO2; and some had been superseded by more recent analyses with more data and longer follow up. These studies were excluded. As an indicator of long-term traffic exhaust, 31 of the studies modelled exposure to NO2 or nitrogen oxides and did not include a particle measure (14 of them investigating asthma). In these cases, it is not possible to attribute the resulting effects to NO2 per se, so they are not discussed in detail here. A total of 81 studies analysed NO2 and a particle measure in parallel, most of them investigating mortality, asthma, or lung function.

Some of these studies will be highlighted below, according to the following criteria:

  • cohort studies with individual follow-up, as their validity is higher than cross-sectional studies or studies based on so-called ecological units; and
  • studies with a contemporary evaluation of a PM fraction, preferably PM2.5 or black carbon.

Meta-analyses that include studies with and without a contemporary evaluation of a PM fraction and with a comparison with PM are also highlighted. A tabulation of the full set of studies is available on request.

The following discussion focuses on respiratory health in children and long-term mortality studies, as these are the outcomes with the largest number of investigations and with results relevant to possible standard settings.

Long-term effects on lung function development in children

The database available for evaluating the relationship between lung function growth in children and long-term exposures to NO2 has increased. Three large cohort studies have examined this relationship. The California-based Children’s Health Study, examining exposure to various pollutants in children for an 8-year period in 12 communities, demonstrated deficits in lung function growth (Gauderman et al., 2004) (FVC, FEV1, maximal mid-expiratory flow (MMEF)) for NO2, PM2.5, elemental carbon and acid vapour. The average NO2 concentrations during the study period ranged from about 7.5 µg/m3 to 71.4 µg/m3 (4–38 ppb). The effects for NO2 were generally greater than has been found for the other pollutants, although the authors recognize that, as for other studies, they could not discern independent effects of pollutants, because they came from common sources and there is a high degree of inter-correlation between them. A later publication of the Californian children cohort studies by Gauderman et al. (2007), including the results of a second cohort and a focus on traffic, provided almost the same estimates for FEV1 and background NO2. The estimate was not altered by inclusion of the traffic indicator in the model with NO2 and vice versa.

Similar findings for lung function growth have also been observed across 10 areas in Mexico City (Rojas-Martinez et al., 2007), where the levels of pollution were rather high (study area means for 1996–1999 of: NO2: 51–80 µg/m3; PM10: 53–97 µg/m3). The effect of an interquartile range NO2 exposure was slightly larger than that of the effect of PM10 in both boys and girls, with similar results being found in two-pollutant models. In Oslo and Stockholm, Oftedal et al. (2008) and Schultz et al. (2012) reported that the lifetime (Oslo) or first year of life (Stockholm) exposure to NO2 from traffic was associated with decreased lung function at 9–10 (Oslo) and 8 (Stockholm) years. Because of the high correlation with other air pollutants, the effects could not be separated with multipollutant models. Overall, the association with deficits in lung function growth in single-pollutant models noted in the 2005 global update of the WHO air quality guidelines has been confirmed even in cities with low concentrations, and there is now evidence for an effect independent of PM10 and PM2.5 in multipollutant models, at least in a city (Mexico city) with a range of NO2 concentrations at the upper end of the concentration range in Europe (Cyrys et al., 2012).

Long term effects on asthma

All the cohort studies that examined asthma outcomes identified in the search and published up to 2010 were included in a recently conducted systematic review by Anderson, Favarato & Atkinson (2013b). This evaluated the effects of NO2 and PM2.5 on the incidence of asthma and wheeze. Results for NO2 from 13 studies, most conducted in Europe, showed an overall RR estimate of 1.09 (95% CI: 1.05–1.14) per 10 µg/m3, after correcting for the scaling of results for one of the studies (Anderson et al., 2012). The overall effect estimate of five studies that evaluated the association between PM2.5 and the incidence of asthma and wheeze was 1.16 (95% CI: 0.98–1.37) per 10 µg/m3 PM2.5. Given the larger spatial variability of NO2 in comparison with PM2.5 (a factor of 1.5–3.0), the effect estimate for NO2 is comparable, if not larger, than that for PM2.5.

Most of these studies had mean or median (usually annual average) values for NO2 below 40 µg/m3, including some positive studies with the entire concentration range below 40 µg/m3 and some negative studies in locations with higher concentration ranges. But the significance of this for setting guidelines is lessened, as none of the constituent studies performed two-pollutant models with NO2 and particles. The analysis of the California Children’s Health Study (McConnell et al., 2010) on asthma incidence did perform multi-pollutant analyses but did not do a two-pollutant analysis for the fixed site (background) measurements of NO2 and PM, because it did not find a significant association with PM. Instead, it modelled traffic exposure at school and at home, based on nitrogen oxide estimates from modelling in a three-factor model, together with NO2 concentrations from central site monitoring. In this three-factor model, the effects of the centrally monitored NO2 levels (16.4–60.7 µg/m3 (8.7–32.3 ppb), annual average) did not disappear fully, but were reduced to statistical insignificance when including traffic exhaust (estimated by nitrogen oxides) at home and at school. Traffic at school and at home remained significant after including the effects of central site NO2. McConnell et al. (2010) attribute the main effects to traffic exhaust, but note that the lack of statistical significance of the positive association with central-site NO2 controlled for traffic could be due to exposure misclassification.

After the online publication of the Anderson et al. (2013b) meta-analysis, four cohort studies on asthma incidence and long-term average concentrations of NO2 were published and all provide support for an association with NO2. Carlsten et al. (2011a) and Gruzieva et al. (2013) also studied particles (showing an effect for PM2.5, but not for black carbon), but none of these evaluated two-pollutant models with NO2 and PM (Carlsten et al., 2011a; Andersen et al., 2012b; Lee Y et al., 2012; Gruzieva et al., 2013).

The results of prevalence studies of asthma are less clear, as it is uncertain whether they represent a true long-term effect on incidence. We describe a meta-analysis of these studies here, as it is an important body of literature: the studies may include questions on lifetime asthma, and they may be a more suitable basis for quantifying the effects of traffic measures (as asthma can remit over time, and the meta-analysis of asthma incidence did not control for the duration of follow-up period, as it was intended only for hazard assessment). The meta-analyses on the period prevalence of asthma and wheezing during the study period and on lifetime asthma by the same authors (Anderson, Favarato & Atkinson, 2013a), both based on nine studies with pollution gradients mainly between communities,13 did not show any relationship with NO2 or other pollutants. The authors did not exclude an association between the close proximity of traffic pollution and asthma in highly susceptible individuals, which may be diluted in a whole-community study. There are also a range of cross-sectional studies that investigated within-community exposure contrasts, more representative of traffic pollution, which have not been reviewed here, but include many positive associations with NO2, some of which are statistically significant. Three of the recent area studies on asthma prevalence included multipollutant models (Dong et al., 2011; Pan et al., 2010; Sahsuvaroglu et al., 2009). In the study by Dong et al. (2011), the associations were reduced from an OR of 1.19 (95% CI: 1.06–1.34) per 10 µg/m3 in males and an OR of 1.14 (95% CI: 0.99–1.30) in females to an OR of 0.97 for both sexes in a five-pollutant model. In Pan et al. (2010), the effect estimates for current asthma associated with NO2 were strongly reduced and lost statistical significance in a three-pollutant model with total suspended particulates and SO2 (total suspended particulates and NO2 were correlated with r = 0.6). Sahsuvaroglu et al. (2009) investigated asthma prevalence in schoolchildren in Hamilton, Canada. Overall, there was no association of asthma with any pollutant. Asthma was only significantly associated with NO2 estimated with a land use regression model in the subgroup of girls without hay fever. This association showed a larger estimate after including SO2, ozone and PM10 in the same model.

The only study of long-term exposure and respiratory symptoms reviewed for the 2005 global update of the WHO air quality guidelines that included multipollutant models was that by McConnell et al. (2003), as part of the California Children’s Health Study. Due to the importance of multipollutant models to the question of whether NO2 is having an effect per se and the discussion of this study in Question C4, the study is described again here. McConnell et al. (2003) investigated the associations between bronchitis symptoms in children with asthma and particles and gaseous pollutants over a period of four years – between the study communities and with yearly within-community variability of pollution. Across communities, symptoms were associated with PM2.5, elemental carbon and NO2, but as those pollutants were closely correlated, no consistent between-community effects were observed in two-pollutant models. In contrast, the within-community associations were stronger, and the associations of symptoms with the yearly variability of organic carbon and NO2 were, in general, not confounded by other pollutants. The yearly variability was expressed as the annual deviation from the 4-year average and ranged from 2.1 µg/m3 to 24.1 µg/m3 (1.1–12.8 ppb) deviations from 4-year averages of 7.9–71.4 µg/m3 (4.2–38.0 ppb) NO2. No other pollutants were significantly associated with an increase in symptoms in models that included organic carbon or NO2 (McConnell et al., 2003).

McConnell et al were cautious to attribute the associations with organic carbon to diesel exhaust, because there were no associations with elemental carbon, also stemming from diesel engines. They did not expect NO2 to show the strongest associations and suggested the possibility of a smaller error in the measurement of organic carbon and NO2 than in that of other correlated pollutants as being potentially responsible for the effect. No study of quite this design has been published since 2004 but, in general, while studies have continued to report associations between NO2 and respiratory symptoms, in most cases it is neither proven nor disproven that these associations are related to NO2 per se. However, indirect evidence – from short-term studies and long-term studies on lung function, which supports the plausibility that some of these effects are due to NO2 – has increased.

Summary of mortality-related to long-term exposure

A total of 18 cohort studies and two case-control studies published since 2004 provide RR estimates for associations of natural mortality, cardiovascular and respiratory mortality or lung cancer with NO2 and PM. With the exception of studies conducted in China, most investigations were conducted in areas where the average NO2 levels were below 40 µg/m3. Not all of the studies presented correlations between NO2 and other pollutants, but those that did indicated moderate to high correlations; in European studies, the correlation is typically greater than 0.80. We focus mainly on the four cohort studies with multipollutant models and the five European cohort studies that analysed NO2 and a particle metric – but without estimating two-pollutant models – giving details also in Tables 7 and 8.

Table 7.. Cohort studies with multipollutant models for NO2 and a particle measure and mortality risk.

Table 7.

Cohort studies with multipollutant models for NO2 and a particle measure and mortality risk.

Table 8.. Cohort studies from Europe on long-term mortality risk and NO2 and PM without multipollutant models.

Table 8.

Cohort studies from Europe on long-term mortality risk and NO2 and PM without multipollutant models.

The following European studies are relevant to the evaluation.14

Recently, the results from a cohort of more than a million adults in Rome were published (Cesaroni et al., 2013). Long-term exposures to both NO2 and PM2.5 were associated with increased risks of nonaccidental mortality. The strongest association was found for ischaemic heart diseases, but also mortality for cardiovascular diseases and lung cancer were significantly associated with both pollutants. Respiratory mortality was marginally associated with NO2 and not associated with PM2.5. The cohort was built on administrative data and lacked information on such behavioural risk factors as smoking (smoking was available on a subset of the cohort and no confounding from this factor was suggested in a sensitivity analysis). The models were therefore adjusted for pre-existing conditions that share lifestyle risks (diabetes, chronic obstructive pulmonary disease and hypertensive heart disease) and for individual and small-area socioeconomic position. The average exposure estimated for individual address for the follow-up (the mean follow-up duration was 8.3 years) was 43.6 µg/m3 (range: 13.0–75.2 µg/m3) for NO2 and 23.0 µg/m3 (range: 7.2–32.1 µg/m3) for PM2.5. The functional form of the association showed no evidence of deviation from linearity for non-accidental mortality, cardiovascular mortality, respiratory mortality and lung cancer, but it showed some deviation from linearity for the association between NO2 and ischaemic heart disease mortality. In a two-pollutant model, the estimated effect of NO2 on non-accidental mortality was independent of PM2.5.

In a French study (Filleul et al., 2005) a comparison of mortality over a period of 25 years between 18 areas, including both NO2 and black smoke, revealed positive and statistically significant effects of NO2 – concentrations averaged between 12 µg/m3 and 32 µg/m3 over a period of 3 years for natural, cardiopulmonary and lung cancer mortality. The effects of black smoke were lower than those observed for NO2. Six of the twenty-four areas with a NO2 average from 36 µg/m3 to 61 µg/m3 were excluded from the analyses, as the monitors were not considered representative of the population’s exposure (assessed with a high nitrogen oxide-to-NO2 ratio, indicating localized traffic sources). In Germany, Heinrich et al. (2013) followed the vital status of women with baseline NO2 and PM10 exposure values for more than 18 years (SALIA study). Positive effects were found for all-cause and cardiopulmonary mortality in relation to NO2 (median exposure: 41 µg/m3 in the year before baseline investigation, based on the annual average of the next monitoring station)) and for all-cause and cardiopulmonary mortality and for lung cancer mortality in relation to PM10. The association between cardiopulmonary mortality and PM10 was reduced in this extended follow-up period, during which PM10 concentrations (but not NO2 concentrations) were lower, compared with an earlier analysis, after 12 years (Gehring et al., 2006).

In contrast to the analyses of Gehring et al. (2006) and Heinrich et al. (2013), which investigated all-cause and cardiopulmonary mortality, the analyses of the SALIA-study data by Schikowski et al. (2007) investigated cardiovascular mortality risk in relation to NO2 and PM10, and a possible effect modification in women with impaired lung function or pre-existent respiratory diseases. The significantly elevated mortality risk from cardiovascular causes in relation to NO2 (median exposure: 46 µg/m3) was not higher in the susceptible subgroup than in the whole sample. In an analysis of all inhabitants of Oslo, Norway, Naess et al. (2007) reported that the RR estimates for cardiovascular diseases, respiratory diseases and lung cancer were very similar for NO2 and PM2.5 when evaluated across the quartiles of the two pollutants. The Norwegian study also investigated the functional form of the association. The authors give separate results for the two age groups of 51–70 years and 71–90 years. In the younger age group, they observed a threshold for the effect on overall mortality of about 40 µg NO2/m3 due to the results for cardiovascular and lung cancer mortality. Thresholds were also observed for PM2.5 (about 14 µg/m3) and PM10 (about 19 µg/m3). In the older age group, the increase in overall mortality risk was linear in the interval 20–60 µg NO2/m3. For chronic obstructive pulmonary disease, a linear effect was seen for both age groups. The large Dutch study by Beelen et al. (2008a) evaluated several pollutants, including NO2, PM2.5 and black smoke. The effect sizes for natural, cardiovascular, and respiratory mortality were greater for NO2 (population average: 36.9 µg/m3; SD: 8.2 µg/m3) than for PM2.5 (population average: 28.3 µg/m3) and were similar to those for black smoke. The RRs per 30 µg NO2/m3 were 1.08 (95% CI: 1.00–1.16) for natural mortality, 1.07 (95% CI: 0.94–1.21) for cardiovascular mortality and 1.37 (95% CI: 1.00–1.87) for respiratory mortality; the RRs per 10 µg PM2.5/m3 were 1.06 (95% CI: 0.97–1.16) for natural mortality, 1.04 (95% CI: 0.90–1.21) for cardiovascular mortality and 1.07 (95% CI: 0.75–1.52) for respiratory mortality. The spatial correlation between different pollutants was high.

For cardiopulmonary mortality, the large American Cancer Society cohort study found no or a marginally significant association with NO2 (a 1% increase per 18.8 µg NO2/m3 in 1980 (95% CI: 0–2%)) and a marginally increased risk of ischaemic heart disease mortality (a 2% increase (95% CI: 0–3%) at a median exposure of 49 µg NO2/m3). The study compared mortality and pollution between populated regions and was not designed to investigate smaller scale variations in pollution (Krewski et al., 2009). By contrast, a cohort study in Toronto, using land use regression to predict NO2 at the residential address and an interpolation method to predict PM2.5, did find a large effect on natural and cardiovascular mortality for NO2, but not for PM2.5 (Jerrett et al., 2009b). Therefore, the authors modelled NO2 together with traffic proximity (and not with PM2.5), and the effects for NO2 were slightly weakened, but still significant (from 1.40 (95% CI: 1.05–1.86) to 1.39 (95% CI: 1.05–1.85) per interquartile range of 7.6 µg NO2/m3). The median of individually assessed NO2 exposure was 43.05 µg NO2/m3 (25th percentile: 39.1 µg NO2/m3; 75th percentile: 38.46 µg NO2/m3).

A new analysis of the California data from American Cancer Society Cancer Prevention Study II (Jerrett et al., 2011) used several different methods to model exposures to air pollution with high spatial resolution. It found, in a cohort of more than 76 000 adults with 20 432 deaths in the follow up period from 1982 to 2000, consistent associations of PM2.5 with mortality from cardiovascular causes, comparable with the findings from the national American Cancer Society Cancer Prevention Study II. However, the strongest associations were found for NO2. Individual exposure data were derived from a land use regression model of NO2 that predicted local variations in the exposure of participants in the years 1988–2002 (mean of individual level exposure 12.3 ppb or 23.1 µg/m3). The authors stated that NO2 is generally thought to represent traffic sources and concluded that combustion-source air pollution was associated with premature death. The analyses are published as a report for the California Environmental Protection Agency Air Resources Board and are not (yet) published in a peer reviewed journal. The report does not give effect estimates with confidence intervals for the two-pollutant models.

Only a few of the studies reviewed conducted a formal multipollutant analysis designed to differentiate the specific effects of the pollutants. The large registry cohort study in Rome found associations of NO2 with non-accidental mortality, independent of PM2.5 and also independent of traffic density or distance to the next main road (Cesaroni et al., 2013). The analysis of the above-mentioned California data of the American Cancer Society Cancer Prevention Study II observed that, in two-pollutant models of NO2 with PM2.5, both pollutants were associated with significantly elevated effects on mortality from cardiovascular disease and ischaemic heart disease. NO2 was also independently associated with an elevated risk for premature death from all causes and lung cancer (Jerrett et al., 2011), but the report did not provide the adjusted quantitative effect estimates.

Hart et al. (2011) examined the association of ambient residential exposure to PM10, PM2.5, NO2, and SO2 with mortality in 53 814 men in the United States trucking industry. One of the unique features of this study was the ability to model the mortality effects of exposures to multiple pollutants. In the multipollutant models, adverse effects were predominantly seen with exposures to NO2 and SO2, and they were reduced for PM10 and PM2.5. A weakness of the study was that there was no information on other risk factors for mortality, such as cigarette smoking. A subsample with information on smoking showed some correlation with pollution, and a recalculation, adjusting for smoking on that basis, suggested that the hazard ratio would be reduced in size but still remain – that is, it did not affect their qualitative conclusions. Similar findings are available from a Canadian cohort study (Gan et al., 2011), where the effects on mortality were stronger for black carbon and NO2 than for other pollutants; the study used a multipollutant model (with NO2, PM2.5, and black carbon in the same model) that attenuated the NO2 effect, but did not reduce it to null (RR = 1.03, 95% CI: 0.99–1.07). It should be emphasized that the effect estimates for black carbon were robust against adjustment for NO2, whereas the effect estimate for NO2 was reduced to non-significance after adjustment for black carbon.

Overall, the findings suggest that traffic and other sources of fossil fuel combustion are important pollution sources that result in greater overall lung cancer, cardiovascular disease and respiratory disease mortality in the cohorts. In contrast to the cardiovascular mortality risk found to be associated with long-term exposure to NO2 in the above studies, such an association was not found consistently in studies on the incidence or prevalence of cardiovascular disease. Only seven studies with cross-sectional, case-control or cohort designs investigated the relationship of these outcomes with NO2 and particles in parallel. Five of those did not find any association or just a small insignificantly higher risk; two recent publications found an increased risk for ischaemic heart disease and for heart failure, but not for other cardiovascular diagnoses (Beckerman et al., 2012; Atkinson et al., 2013).

In summary, the European cohort studies provide evidence that the NO2 effects on natural and cause-specific mortality are similar to, if not larger than, those estimated for PM. The recent registry cohort study from Italy (Cesaroni et al., 2013) and the American (Jerrett et al., 2011; Hart et al., 2011) and Canadian (Gan et al., 2011) studies have attempted multipollutant models, and they provide support for an effect of NO2 independent from particle mass metrics. In three of these mortality studies with multipollutant models, the major fraction of the populations studied was exposed to NO2 levels lower than 40 µg/m3; in one of them, nearly all participants were exposed to levels lower than 40 µg/m3 (Jerrett et al., 2011). Four of the six European analyses were centred around 40 µg NO2/m3. In the French study, areas with (possibly non-representative) monitor averages above 32 µg NO2/m3 were excluded. The study by Naess et al. (2007) looked at non-linear exposure–response functions and found a possible threshold around 40 µg/m3 for NO2 and also a threshold for particles in the age group of 51–70-year-old people, especially for cardiovascular mortality. In contrast, they found no thresholds in the age group of 71–90-year-old people for all cause and cardiovascular mortality and none for chronic obstructive pulmonary disease mortality in either age group. The Italian study found some evidence for non-linearity in the association between NO2 and ischaemic heart mortality, but not for cardiovascular or non-accidental mortality. All other investigators applied linear exposure–response functions.

As the long-term mortality studies have all included populations exposed in part to annual average NO2 concentrations of well below the current WHO air quality guidelines of 40 µg/m3, or even been conducted over a range almost entirely below the air quality guidelines, it would be wise to consider whether the guideline should be lowered at the next revision of the guidelines.

2.2 Sub-chronic and chronic toxicological studies
Animal studies

Since the 2005 global update of the WHO air quality guidelines, the toxicological evidence up to about 2007 has been described in several reports (EPA, 2008b; CARB, 2007; COMEAP, 2009b; WHO Rgional Office for Europe, 2010). Table 9 summarizes the lowest effect concentrations at the lower end of the dose range, as highlighted in the conclusions sections of the various reports. Such studies as Mercer et al. (1995), highlighted by the EPA, that included spikes of higher concentrations have been excluded. It can be seen that there are now four studies showing effects below 0.34 ppm, including older studies not previously quoted by WHO, marked with an asterisk. Tabacova, Nikiforov & Balabaeva (1985) did not describe maternal toxicity,*15 so it cannot be determined whether the neurobehavioural posture and gait changes are a direct toxic effect of NO2 on the offspring or an indirect effect, via toxicity to the mother. The study has not been replicated by other studies in this dose range, although similar effects have been found at much higher doses. The effects in the studies by Zhu et al. (2012) (reviewed in more detail below) and Takano et al. (2004)* are on risk factors for ischaemia and atherosclerosis, rather than actual occurrence of ischaemia and atherosclerosis themselves. Also, the change in endothelin-1 (Zhu et al., 2012) was in an unexpected direction, a direction that has not yet been replicated in another study. These three studies should not be dismissed, but there is an element of uncertainty in using them to describe a firm lowest effect concentration.

Table 9.. Lowest effect concentrations highlighted by WHO, EPA and CARB at low end of dose rangea.

Table 9.

Lowest effect concentrations highlighted by WHO, EPA and CARB at low end of dose rangea.

Genotoxicity in vitro and in vivo, in animals and humans

Koehler et al. (2011) found weak evidence of DNA fragmentation and increased micronuclei (only after 3 hours) in human nasal epithelial cells at 0.1 ppm NO2.16 This is the first in vitro genotoxicity test in human cells and shows effects at a lower concentration than previously (For context, in vivo mutagenicity tests give both positive and negative results (CARB, 2007; EPA, 2008b)). The Health Effects Institute (HEI, 2012) reported a micronucleus assay, employing flow cytometry performed in mice and rats; they used blood samples from the 3 month bioassay by McDonald et al. (2012), utilizing NO2 within a modern diesel engine emission mixture, as discussed above. While there were some scattered significant findings, these did not form a coherent picture, and it was concluded that the results were negative in both rats and mice. Further studies will be done at longer exposure durations. Another analysis from this bioassay (Hallberg et al., 2012) found no effects of the mixture containing NO2 on strand breaks in lung tissue, assessed with the comet assay, or on 8-hydroxy-2’-deoxyguanosine DNA adducts in the serum of rats or mice.

Another epidemiology study found increased urinary 8-hydroxy-2’-deoxyguanosine (a marker of oxidative DNA damage) in the elderly, with an increased 2- or 3-week moving average for NO2 (daily average for NO2: 17.8 ppb). This was not found for primary traffic pollutants (carbon monoxide, black carbon and elemental carbon), but was found for PM2.5, sulfate and ozone (Ren et al., 2011). A further study found a borderline association between indoor NO2 levels (13–18 µg/m3, averaging time not given) and micronuclei in blood cells of mothers and neonates (Pedersen et al., 2009).

Mechanistic epidemiology studies

In other mechanistic epidemiology studies, NO2 was associated with elevated blood pressure, total cholesterol, fasting glucose, glycosylated haemoglobin and IL-6 in a cross-sectional study, but this was not maintained after adjustment for other pollutants (Chuang et al., 2011). NO2 exposure during pregnancy increased CD-8+ T cells in cord blood. Whether these produced IL-4 was not described – this is of interest, as this type of T cell is higher in atopic asthmatics. In contrast, PM10 reduced regulatory T cells (linked to predisposition towards allergy and asthma) (Baïz et al., 2011). Papers on gene–environment interactions between NO2 and enzymes involved in reducing oxidative stress revealed some interactions, but study limitations (small subject numbers, measurement error, absence of replication) prevented definitive conclusions on the nature of these (Minelli et al., 2011).

Four further studies that reported interactions have been published since. Carlsten et al. (2011b), a small study in a high-risk cohort, showed the GSTP1 Ile105Val polymorphism conferring a weak trend to an expected increased risk of NO2 associated asthma. Adding to the mixed picture on this polymorphism are earlier studies (Castro-Giner et al. (2009) found the interaction not significant; and Melén et al. (2008) found an interaction with allergic sensitization). Tung, Tsai & Lee (2011) found that NO2 exacerbated the increased risk of polymorphisms increasing microsomal epoxide hydrolase activity on lifetime and early-onset asthma, but the biological hypothesis for this is via production of polycyclic aromatic hydrocarbon-derived quinones, rather than a NO2 mechanism. Ungvári et al. (2012) studied polymorphisms in the Nrf2 gene (which coordinates the oxidative stress response) and found an interaction between a NO2 and infection-induced asthma association and a polymorphism with an unknown function (not previously studied). Wenten et al. (2009) showed that the combined CAT–MPO genotype, predicted to protect most against oxidative stress, reduced the risk of respiratory-illness school absence linked to NO2; this has not previously been studied.

This potentially important area of the literature may not be sufficiently mature as yet for firm conclusions. It should be noted that these studies did not control for exposure to PM (or, indeed, exposure to ozone), so it is unclear that these indicate susceptibility to NO2 (per se) or to air pollution (in general), as indicated by NO2.

3. Discussion

We are aware of the possibility that NO2 has no direct effect itself but is, instead, only acting as a marker for primary particles, such as ultrafine particles, and such constituents as metals, polycyclic aromatic hydrocarbons or other organic matter carried on these particles to particular locations in the lung.17 NO2 could also act as a marker for such gases as carbon monoxide or nitric oxide near roads or, as it is a secondary as well as primary pollutant, for regional pollutants such as ozone. Whether this is the case or whether NO2 has a direct effect is a crucially important policy question. The implementation of filter traps in diesel vehicles to meet the Euro 5 emission standards and lowering sulfur in fuel, coupled with the fact that NO2 levels are not being reduced in real-life driving conditions, may have led to important increases in the NO2/ultrafine particle, NO2/black carbon and NO2/elemental carbon plus organic carbon ratios. For example, at a London roadside site, the ratio of nitrogen oxides as NO2 (µg/m3) to particle number (N/cm3) changed more than twofold over 2 years (Jones et al., 2012). Comparison of data from the recent ESCAPE project, which covered the years 2008–2010 (Eeftens et al., 2012b; Cyrys et al., 2012), with the older TRAPCA project, which covered the year 1999 (Hoek et al., 2002a; Lewné et al., 2004),18 suggests that the traffic–urban background contrast for NO2 has increased more over time than for PM2.5 absorbance (by about 10%). The rural–urban background contrasts showed a change of less than 10%. The changes in ratios are likely to be specific to location, site type and time period and would be better defined if there were more widespread and robust measurements of the relevant PM metrics over time.

Given the change in these ratios, now and in the future, and between and within cities, there are clearly policies and behaviours that are changing NO2 independently of other traffic pollution constituents. Unfortunately, there are no means in observational studies to fully test the hypothesis of a direct effect of NO2. Adjustment of NO2 associations for PM10 or PM2.5 may not be sufficient, as there is often a closer correlation between NO2 and traffic pollutants, such as primary PM and its constituents. Correlations with regional pollutants, such as ozone and secondary particles, are usually not as close.

Although the present document does not examine this point in detail, there are several key studies that show effects of NO2 independent of ozone – for example, McConnell et al. (2003) and the review of multicity time-series study multipollutant models by Anderson et al. (2007).

Integration of the epidemiological evidence with chamber studies and toxicological evidence is, therefore, of considerable importance in judging whether there is an effect of NO2 per se.

The discussions below relate mainly to respiratory effects, with a separate paragraph on cardiovascular effects. Effects on the environment are not covered, but are also important (Sutton et al., 2011).

Short-term exposure

In comparing the epidemiological and the toxicological and chamber study evidence for the concentrations at which effects occur, it is important to realize that, in the epidemiological studies, the daily variability at the background sites also reflects the daily variability at hot spots, so that the actual concentrations inducing health effects may be higher than those measured at the background site. Kerbside concentrations can be regularly in the range 380–560 µg/m3 (200–300 ppb) (1-hour average) at some polluted sites (London Air, 2013), and a wider range of sites often exceed the 1-hour limit value of 200 µg/m3 (EEA, 2012). This is within (or approaching) the range of concentrations that result in small effects in the chamber studies (380–1160 µg/m3; 200–600 ppb). Of course, it depends how long people are in those locations but, in addition to such activities as queuing at bus stops, in-vehicle exposure can be similar to the outdoor concentrations in heavy traffic (Chan & Chung, 2003), and car journey durations in congested cities can be considerable. Personal exposures to NO2 reflect the sum of the microenvironment concentrations that an individual moves through over a determined time interval and are, therefore, strongly influenced by high concentration environments.

Concern has been expressed in some studies that ambient concentrations of NO2 are not correlated with personal exposures to NO2 and are, actually, better correlated with personal exposure to PM2.5 (e.g. Sarnat et al., 2001). However, a recent systematic review addressed the issue of the relationship between personal and ambient NO2 exposures, identifying significant correlations between these two exposures estimates, though the strength of the association varied across studies (Meng et al., 2012). Studies not quoted by Meng et al. (2012) also varied, both finding (Rijnders et al., 2001) and not finding (Bellander, Wichmann & Lind, 2012) significant correlations.

It is difficult to extrapolate quantitatively from animal toxicological studies. Modelling (with many assumptions) suggests about a tenfold higher local NO2 concentration in the bronchi of humans than in those of rats, at the same concentration in the air that is being inhaled, and vice versa (much lower) in the alveoli (Tsujino, Kawakami & Kaneko, 2005). Several of the respiratory effects in animals occur in the bronchi, and it is plausible that some of the epidemiological results in human beings (such as asthma admissions) are primarily the result of effects on the bronchi. There may, therefore, be some evidence that supports application of the tenfold safety factors used in traditional toxicology for extrapolation from animals (strictly rats) to human beings. This, together with the distribution of personal exposures and the wide range of susceptibility in the human population, suggests that the epidemiological findings are not necessarily incompatible with the toxicological evidence on NO2 itself. There are too few studies that examine NO2 and particles from defined sources in the same experimental system, to directly compare the toxicological importance of these two pollutants, and their relative toxicological importance may vary by end-point.

The apparent mismatch between the time-series evidence and the lack of apparent responses in the chamber studies at background concentrations may be a consequence of only a small proportion of the population responding at particular times, an effect that could be picked up only in the much larger samples used in time-series studies. Specifically, more sensitive groups, such as severe asthmatics, are not studied in chamber studies because of the risks involved. Thus, the lack of robust effects and/or the lack of evidence at lower doses in chamber studies is insufficient to rule out the reported associations with NO2 in the time-series studies found at the concentrations present in the wider environment. Considering the presence of more sensitive subgroups in the population together with the higher concentrations at microenvironments (such as the kerbsides described above) could explain some of the apparent mismatch, a point also made by Frampton & Greaves (2009).

While the preceding discussion supports the causality of the short-term respiratory effects, the issue is more uncertain for the short-term cardiovascular effects. Positive associations robust to adjustment for PM mass metrics are found in many time-series studies for cardiovascular mortality. The new evidence continues to suggest positive associations between NO2 and cardiovascular hospital admissions, but findings are mixed in terms of robustness of the associations after adjustment for co-pollutants. It is therefore difficult to comment further on the nature of the relationship between NO2 and cardiovascular hospital admissions. The results of panel studies on markers of cardiovascular risk (discussed in the toxicology section) found some interesting results in some cases, but no effects in others. Of the few chamber studies available, most did not find effects on cardiovascular end-points. The few short-term animal studies available did find effects on markers of systemic oxidative stress, inflammation and endothelial dysfunction at above ambient concentrations, but without a defined no-effect level. One study found that plasma from volunteers exposed to NO2 at 0.5 ppm had an effect on coronary epithelial cells in vitro. Although slightly longer term (3 months), the study of modern diesel emissions, more highly dominated by NO2, did not find evidence of cardiovascular effects in healthy young animals. There are theoretical mechanisms by which NO2 inhalation could cause increased nitrative stress in the diseased heart. The overall picture is mainly one of an absence of a sufficient volume of evidence to resolve the mixed results found in the studies available, so it adds uncertainty, rather than challenging the causality of the epidemiological associations.

The current short-term guideline has the advantage of being set on the basis of chamber studies that use NO2 itself. However, there is now a large body of time-series and panel evidence that, in contrast to other pollutants, has not been used so far in determining the level of a short-term guideline. It is recommended that this evidence be included in future considerations of a short-term guideline for NO2.

Long-term exposure

Another challenging issue is whether the effects of long-term exposure are due to NO2 per se.

Human chamber studies are not suitable for investigating long-term exposures. While animal toxicological studies do provide evidence of long-term effects, the degree to which this applies at ambient concentrations is less clear. In studies that compare long- and short-term exposures, the effects of long-term exposures occur at lower concentrations (0.25 ppm for clear adverse effects and some possible effects below this (Table 9)). However, while this concentration can be found at kerbsides, it is less likely that people are exposed to these concentrations several hours a day long-term (the long-term toxicological studies were often not based on exposure for 24 hours a day) than it is for people to be exposed for the shorter durations needed for the short-term effects. The toxicological evidence now includes a few studies that show cardiovascular effects, but the evidence is too limited for firm conclusions. It is much harder to judge the robustness to adjustment in the long-term epidemiological studies than it is in the short-term studies, because the spatial correlations between NO2 and other pollutants are often high. However, there are a few studies that do suggest effects independent of particle mass metrics (including studies on cardiovascular mortality). The existence of effects of short-term exposure provides some plausibility for the effects of long-term exposure – particularly, for respiratory effects.

In summary, there is also some support, to a lesser degree than for the short-term, for a long-term effect of NO2 per se. Again, NO2 may also be capturing the effects of other traffic-related pollutants.

The current long-term guideline developed historically from one based on indoor studies. Now, there are enough outdoor air pollution studies (those described here plus pre-2004 studies) on respiratory effects and on all-cause mortality to consider using them to help set a new long-term guideline, with appropriate caveats about uncertainties. These studies included exposures to concentrations above and below, or all below, the current guideline and, where studied, linear relationships without a threshold have been found for at least some outcomes. This suggests that consideration should be given to lowering the guideline.

How to reflect these uncertainties in regulations is a matter beyond the scope of WHO, but consideration could be given to flagging the guideline to emphasize uncertainty.

Research gaps will be highlighted later in answers to Question C9, but it should be emphasized – particularly for NO2 – where much of the important chamber and toxicological evidence comes from 20–30 years ago. Information on mechanisms is crucial and needs to take advantage of the full range of modern experimental techniques, including systems biology, to capture interacting causal pathways.

Question C3

Based on existing health evidence, what would be the most relevant exposure period for a short-term limit value for NO2?

Answer

The most relevant exposure period based on existing evidence is 1 hour because 1-hour peak exposures in chamber studies have been shown to produce acute respiratory health effects. Toxicological studies also support the plausibility of responses to peak concentrations. Time-series and panel studies have examined associations, using both 24-hour average and 1-hour average NO2 concentrations with similar results. Evidence from these studies would support the development of a 24-hour WHO guideline or a 1 hour guideline but, as there is chamber study and toxicological evidence on, or close to, a 1-hour basis and much less evidence on a 24-hour basis, a 1-hour exposure period is preferred. In urban areas, 1-hour peak concentrations and 24-hour averages were so highly correlated that it should be possible for a 1-hour peak guideline to be derived from studies using 24-hour average NO2, following expert analysis of how these metrics are related in Europe. There is, therefore, no need to develop a 24-hour limit value in addition to a 1-hour guideline based on epidemiological studies.

Rationale

1. Time-series studies

The majority of time-series studies have examined associations using 24-hour averages of NO2 concentrations, with fewer using maximum 1-hour averages. The studies have reported associations that suggest adverse mortality and morbidity effects at concentrations below the current 1-hour WHO air quality guideline for NO2. Consistent findings of positive associations with respiratory and asthma hospital admissions have been reported. These findings are in keeping with the outcomes investigated in chamber studies, which demonstrate direct effects of NO2 over a few hours. Such evidence of direct effects in the chamber studies led to the development of the current short-term guideline. Given the close correlations that exist between 1-hour and 24-hour measures – for example, the range of correlation coefficients between the maximum daily 1- and 24-hour NO2 concentrations across the 29 European cities in Samoli et al. (2006) was 0.80–0.94, with a median of 0.90 – a 1-hour guideline could be converted to a 24-hour one. Samoli et al. (2006) also reported that, in European cities providing both 1-hour and 24-hour average concentrations of NO2, the ratio between the two measurements was 1.64. This could be used to scale a 24-hour average concentration–response function to a 1-hour-average concentration–response function that could then be used to set a time-series-based 1-hour guideline.

2. Panel studies

The panel studies of respiratory effects in asthmatic children have generally used 24-hour averages as the shortest averaging period. As with the time-series studies, this does not necessarily mean that 24-hour continuous exposure is required before effects are found.

3. Chamber studies

The chamber studies on human beings examined exposure intervals over periods of 30 minutes to 6 hours, with 1 hour being a common exposure period. Comparisons of durations within the same experiment were not widely done, and a comparison of the effect of different durations across experiments is difficult, given the weakness of the effect itself (clinically significant effects have only been reported at relatively high concentrations in mild disease – see the discussion for Question C2). There is a great need to know whether NO2 per se has direct effects in severe and hyperreactive asthma. No such chamber studies have been conducted (and would be unlikely to be conducted), and thus the time period for severe asthmatics to respond is unknown. After review of the recent literature – albeit, in mild asthma, but with a range of nonspecific airway hyperresponsiveness – there appears little reason to alter the level or averaging time of the current 1-hour average limit value, from the chamber study perspective.

4. Toxicological studies

It is apparent from the previous three sections that the evidence depends on the averaging time (for the time-series and panel studies) or exposure period (for the chamber studies) chosen for study. For the time-series and panel studies, the averaging time chosen is driven by the monitoring data available, rather than any mechanistic considerations, and this leads the discussion to a comparison of 1-hour and 24-hour averaging times. The question, however, does not relate only to 1-hour and 24-hour average exposure periods. Do toxicological studies that have more flexibility in the exposure periods chosen for study indicate any other exposure period would be more appropriate?

  • a. Is there evidence on the time scale over which NO2 or its reaction products reach potential targets for toxicological effects?

Enami, Hoffmann & Colussi (2009) suggest that the uptake of NO2 gas across the air–liquid interface – via conversion into hydrogen and nitrate ions and nitrous acid – occurs within milliseconds. Absorption into bronchoalveolar lavage fluid via reaction with antioxidants in vitro was already apparent within 30 minutes at environmentally relevant concentrations (Kelly & Tetley, 1997). At 20 ppm, NO2, containing the isotopic nitrogen-15 label, was present in perfused rat lung tissue (soluble and insoluble components) and in the perfusate, probably as nitrite, (analogous to the blood supply) after inhalation for an hour (shorter durations were not tested) (Postlethwait & Bidani, 1989). The NO2 radical (derived from nitrite) had an estimated permeability coefficient of 5 cm per second across lipid membranes, suggesting lipid membranes are not a significant barrier to NO2 transport (Signorelli et al., 2011).

  • b. What was the exposure period for the lowest effect concentrations in short-term toxicology studies described in Question C2?

An increase in mast cell numbers occurred in rat bronchi after 3 hours at 0.2 ppm NO2 (Hayashi & Kohno (1985), quoted in CARB, 2007); altered behaviour of detoxification enzymes (measured by pentobarbital sleeping time) occurred in mouse livers after 3 hours at 0.25 ppm (Miller et al., 1980); biosynthesis of the carcinogen dimethylnitrosamine in dimethylamine-treated mice was detected after 30 minutes at 41.5 ppm or 2 hours at 0.1 ppm (Iqbal, Dahl & Epstein, 1981); and increased proliferation of bronchiolar tissue was found after 24 hours at 0.8 ppm (Barth et al., 1994) These were usually the shortest durations tested, so it is unknown whether shorter durations could have had the same effect.

  • c. Is there evidence that peak concentrations are particularly important?

Defining a guideline on the basis of a shorter exposure period has the effect of controlling peak concentrations more strongly. In toxicological studies, both concentration and time (and, hence, the total amount of chemical delivered) can contribute to the development of effects, but the relative importance of concentration and time may differ for chemicals with different mechanisms. For example, antioxidant defences may handle the same amount of NO2 more easily if presented as a lower sustained exposure, where there is time for induction of antioxidant enzymes to occur, than if presented in a short peak that would overwhelm the antioxidant defences. Although their study was related to long-term exposure and high concentrations, Rombout et al. (1986) investigated the relative importance of peaks of NO2 for morphological changes in the rat lung. The onset of effects on the bronchiolar epithelium occurred earlier and was more serious at 10.6 ppm intermittently (6 hours a day) for 4 weeks than at 2.7 ppm continuously for 28 days (exposures with the same product of concentration and time), so it was concluded that concentration played a more important role than duration. For the influx of macrophages, continuous exposure was more important than intermittent exposure. This was confirmed by Frampton et al. (1989), who found that macrophage activity was affected more in 4 of 9 human volunteers exposed to 0.6 ppm continuously for 3 hours than it was for a background of 0.05 ppm for 3 hours with three 15 minute peaks at 2 ppm (both 108 ppm-minutes).

Miller et al. (1987), however, compared the effect in mice of a continuous 0.2 ppm background exposure 5 days a week for 1 year, and with the same background exposure (but also with a 0.8 ppm spike for an hour twice a day) and found that mortality from infection and the effect on pulmonary function were greater in the presence of spikes. Peaks of 4.5 ppm for 1, 3.5 or 7 hours increased mortality when a Streptococcus challenge was given immediately afterwards; but when given 18 hours later, only the 3.5- and 7-hour duration peaks increased mortality in mice (Graham et al., 1987). This suggests recovery is more likely for the shorter durations. It is not necessarily the case that the importance of peaks compared with duration will be the same for different outcomes. It is also worth noting that many of the longer-term toxicological studies only involved exposures for 6 hours during the day for weeks or months, not 24 hours a day.

A short-term guideline, while set on the basis of short-term exposure studies, may nonetheless have a role in reducing the likelihood of longer-term effects, by controlling peaks. In summary, while the importance of peaks may vary for different outcomes, there is some evidence that peaks are important.

5. Discussion

There is relatively little research aimed directly at assessing the importance of duration of exposure, and many of the toxicological studies that address this are at high doses. Nonetheless, there is some indication that shorter durations may be more important than longer ones, at least for some end-points. There is no strong evidence to argue against using a 1-hour exposure as used in the current guideline.

The question assumes just one short-term limit value. There is an argument for having both a guideline set on the basis of chamber studies, where the toxic agent is known to be NO2, and a further guideline set on the basis of the large body of time-series studies that show the effects at lower concentrations, but with more uncertainty as to the responsible indicator pollutant for health effects. This would make the WHO view on the different types of evidence more transparent. These could subsequently be pooled for regulations. Given the evidence that an hour is sufficient to cause effects, it is not clear what might be the added health benefit of adding a 24-hour average guideline for NO2. This is not to say that the 24-hour average concentration time-series evidence cannot be used. As explained earlier, examination of the relationship between 24-hour and 1-hour average concentrations of NO2 in Europe should enable conversion of concentration–response functions from 24 to 1 hour and/or indicate whether a 24-hour average guideline expressed in 1-hour average terms is tighter than the 1-hour average guideline based on chamber studies. The tighter guideline can then be used as the basis for the standard. Alternatively, the subset of time-series studies that use maximum 1-hour averages could be used to set the time-series-based guideline, although there are fewer studies to choose from.

Question C4

Based on currently available health evidence, what NO2 metrics, health outcomes and concentration–response functions can be used for health impact assessment?

Answer

This answer assumes an application in a health impact assessment of NO2 itself, given that impacts of other pollutants – notably PM mass – are also being quantified. The use of NO2 as an indicator for a health impact assessment of local traffic measures is discussed in the rationale. The evidence base supports quantification of effects of short-term exposure, using the averaging time as in the relevant studies. The strongest evidence is for respiratory hospital admissions, with some support also for all-cause mortality – these are recommended outcomes for use in the core analysis. Cardiovascular hospital admissions can be included as a sensitivity analysis – the evidence is more uncertain than for respiratory admissions. It is recommended to derive concentration–response functions from time-series studies that have provided effect estimates for NO2 adjusted for at least PM mass.

For a core health impact assessment of effects of long-term exposure to NO2, the recommended health outcome is bronchitic symptoms in asthmatic children, with the coefficient adjusted for a PM metric based on the Southern California Children’s Health Study. A health impact assessment using asthma prevalence could also be performed. However, as only estimates from single-pollutant models are currently available for asthma prevalence, this health outcome should only be used in sensitivity analyses that compare results with those of health impact assessments for PM mass.

Cohort studies also show relationships between long-term exposure to NO2 and mortality, but not all are sufficiently robust for use in a core health impact assessment. Therefore, the effect of long-term exposure to NO2 on all-cause mortality is recommended for sensitivity analysis only. Concentration–response functions from cohort studies with effect estimates for NO2 that were adjusted for at least PM mass should be used. In the same way, cardiovascular mortality could also be included in a sensitivity analysis, due to the uncertainty about a mechanistic understanding of cardiovascular effects.

Rationale

1. Context in which the concentration–response functions will be used in health impact assessment

It is important to emphasize that the appropriate concentration–response functions to choose will vary according to the context of the health impact assessment. Some of the factors to be taken into consideration (and the questions that relate to them) are set out below, particularly in terms of whether the concentration–response functions are being used to assess effects of NO2 itself or are being used as an indicator of a mixture associated with NO2.

  1. Is the primary purpose of the health impact assessment to estimate the burden of current air pollution or to assess the health impacts of a change? In general, in a health impact assessment, use of an indicator pollutant is inappropriate for evaluating a change: but such an assessment has less uncertainty in estimating a burden, rather than a change, if based on epidemiological studies from a similar pollution environment – in terms of date and type of area. However, since policy is about change, burden calculations are used more in the context of alerting people to the current problem than for sophisticated future policy analysis. This is because current air pollution is more likely to have a correlation pattern between pollutants similar to that of the situations where the epidemiological associations were derived, whereas a change is more likely to lead to a different correlation pattern between pollutants. This would mean that indicator pollutants would no longer act as indicators in the same way.
  2. If the health impact assessment is about the impact of a change (involving a change in NO2), what exactly is causing the pollution change and what does that imply for pollutants other than NO2? For example, for local traffic measures, is it about reducing local traffic as a whole, or about reducing emissions of NO2 specifically?
  3. What is the spatial scale of the health impact assessment? The role of NO2 as an indicator may vary with spatial scale (close to roads; within a city; and between cities and regions).
  4. What other pollutants are included in the health impact assessment model? There are at least four possibilities available to address this; these possibilities can vary with both context and outcome, as outlined in the following questions and remarks:
    1. modelling of effects of other pollutants, such as PM2.5, with quantification of relationships with NO2 intended to supplement this − that is, is there reasonable confidence in an effect of NO2 itself being additional to other pollutants;
    2. what if scenarios (sensitivity analysis), where there is more confidence in the effect of the other pollutant(s) than in NO2 – that is, “if the possible effect of NO2 is true, what might the additional effect be;
    3. what if scenarios (sensitivity analysis), where it is unclear which pollutant is responsible – that is, if this pollutant were responsible, there could be an effect of x; if NO2 were responsible, there could be an effect of y, so the effect is likely to lie in the range of x to y; and
    4. in circumstances where quantification of effects of, for example, PM2.5, is too difficult (such as those for which measurements are unavailable), or where a traffic measure is not changing the composition of the emissions or the fleet (such as pedestrianization) – NO2 is being used as an indicator.

Coefficients adjusted for other pollutants would be appropriate for (i) and (ii), whereas single-pollutant models would be appropriate for (iii) and (iv).

The main question for this section does not specify the context of the health impact assessment. As our main case, we have assumed that policy measures that may affect NO2 independently are being assessed and that other pollutants are also being quantified (item d(i–iii) above). Other alternatives, which can also be important (item d(iv), will be mentioned at various stages in the text below. A summary in Table 10 at the end of the section will help make the recommendations clear.

Table 10.. Recommended pollutant outcome pairs by health impact assessment context and ranked by uncertainty.

Table 10.

Recommended pollutant outcome pairs by health impact assessment context and ranked by uncertainty.

It is important that the full range of studies should be considered, not just those since 2004 that have been considered in detail in this review. While we have mentioned some earlier studies of which we are aware, we understand that meta-analyses of (or selection of a representative study from) the full range of studies will be performed by another project. We concentrate here on the appropriate outcomes and metrics and on studies or areas of evidence that could be used as sources of concentration–response functions. However, firm recommendations for specific concentration–response functions await further work by this separate project.

2. Effects of short-term exposure

Respiratory hospital admissions. The most consistent evidence comes from short-term epidemiological studies of respiratory morbidity, and this is supported by chamber-study and toxicological evidence (see Question C2) – this can be used in the central analysis (see item d(i) above). We recommend using respiratory hospital admissions for all ages. The averaging times from the relevant studies could be used in health impact assessments, as the majority of studies available used 24-hour average concentrations of NO2. Whether or not to use coefficients adjusted for other pollutants also needs to be considered, and this may affect the choice of metric. Coefficients adjusted for at least PM mass should be considered. There are more adjusted coefficients available for 24-hour average NO2 than for the 1-hour average. Coefficients could be selected either from a representative multicity study or a meta-analysis of all available studies. As well, an existing meta-analysis – for example, Anderson et al. (2007) – could be used, though this would not reflect more recent studies. If there is specific interest in a health impact assessment using maximum 1-hour average concentrations of NO2, consideration could be given to converting the larger body of evidence on associations with 24-hour average NO2 to a maximum 1-hour average concentration–response function (see Question C3). As explained earlier, examination of the relationship between 24-hour and maximum 1-hour average concentrations of NO2 in Europe would be required to enable conversion of concentration–response functions from 24 hours to maximum 1 hour. Whether it is appropriate to convert an adjusted 24-hour coefficient (derived by meta-analysis or selected from a paper) to one for maximum 1-hour is unclear.

Cardiovascular admissions. The epidemiological associations for cardiovascular admissions are not generally robust to adjustment for co-pollutants; there are very few chamber studies or toxicological studies on cardiovascular end-points available, and those that exist are contradictory (see Question C2). The evidence on PM and cardiovascular admissions is stronger. We therefore recommend that concentration–response functions for cardiovascular admissions are used only in sensitivity analysis as a possible effect additional to PM (that is, item d(ii) above) or when using NO2 as an indicator (item d(iv) above).

The same points about metrics and adjusted coefficients, discussed above for respiratory hospital admissions, apply for cardiovascular admissions.

All-cause mortality. Given that respiratory hospital admissions are to be quantified in a central analysis, there is also some plausibility for an effect on respiratory mortality; and, indeed, consistent associations with respiratory mortality have been shown in many cases. Associations have also been shown for cardiovascular mortality but, as mentioned above, the issue of causality is more uncertain. However, since there can be issues about cross-diagnosis between respiratory and cardiovascular deaths and since baseline rates for all-cause mortality are widely available, we recommend use of concentration–response functions for all-cause mortality for all ages for the central analysis (item d(i) above), acknowledging a certain additional level of uncertainty compared with respiratory admissions.

For health impact assessments, the same points about metrics and adjusted coefficients as for respiratory hospital admissions above apply. If it is not possible to do a meta-analysis of the larger body of literature to derive a coefficient, estimates could be based on Samoli et al. (2006), which presents pooled results from 30 cities in Europe using maximum 1-hour average NO2 concentrations with adjustments for PM10 and black smoke. The NO2 estimate adjusted for black smoke (a good indicator of primary combustion particles) would better reflect a possible NO2 per se effect than would a coefficient adjusted for PM10. If a meta-analysis is possible, multiple studies are available on adjusted coefficients for 24-hour average NO2 concentrations, although thought would need to be given to which pollutant-adjusted coefficients to use. Use of a coefficient adjusted for PM is suggested. Multiple studies are also available on 24-hour average single-pollutant models (fewer studies are available for the maximum 1-hour average), for circumstances where this would be appropriate (for example, when other pollutants are not being quantified (item d(iv) above) or where the dangers of double counting are openly acknowledged (for example, item d(iii) above).

3. Effects of long-term exposure

As mentioned in part 1 of Question C4, we have assumed as our main case that policy measures that may affect NO2 independently are being assessed and that other pollutants are also being quantified (item d(i–iii) in part 1 above). Quantifying the effects of NO2 itself is particularly challenging. This will be addressed first. The importance of capturing the health impacts of traffic pollution, which might otherwise be missed, is discussed second – that is, scenarios as described in item d(iv) of part 1 of Question C4. Within each of the following sections, respiratory effects are considered first when discussing the health outcomes – given the greater degree of mechanistic evidence supporting the effect.

3.1 NO2 per se

When both PM metrics and NO2 measures are used in health impact assessments, some of the effects attributed to each of them may overlap. So, estimates from two-pollutant models should be used to attempt to avoid this. Unfortunately, very few long-term studies provide estimates from two-pollutant models (see Question C2), often because NO2 and PM were too closely correlated. In addition, papers may not perform two-pollutant models when one of the pollutants did not show a significant association.

Studies that examine NO2 using exposures based on a few monitoring sites as a broad surrogate for personal exposure may have lower correlations with primary particle metrics. But this advantage may be offset by the exposure measure being less representative of individual personal exposure (measurement error) and, therefore, showing weaker effects in two-pollutant models, independent of a possible so-called true association. Correlations with PM mass metrics may be higher in such studies. This illustrates the point that the spatial scale of the study can influence the interpretation of the concentration–response function to be used in the health impact assessment.

Black carbon, elemental carbon or black smoke, and carbon monoxide, being alternative measures to capture traffic exhaust effects, share with NO2 the aspect of traffic proximity and association with emissions from combustion. They are often more closely associated with NO2 than with particulate mass. It is self-evident that the results of an impact assessment for those indicators should not be added with the results for NO2, if single-pollutant models are being used.

The health outcomes associated most consistently with long-term exposure to NO2 are mortality, respiratory symptoms or asthma, and lung function. Cardiovascular outcomes are more uncertain – not because there is a large body of evidence showing a lack of effects, but because there are relatively few studies and there is less support from mechanistic evidence. The degree to which these associations are a consequence of NO2 per se was discussed in Question C2, and the points relevant to choices of concentration-response functions are highlighted below for each end-point.

Children’s respiratory symptoms and/or asthma symptoms. As discussed in Question C2, there are several cohort studies that show the effects of long-term exposure to NO2 on the respiratory condition of children. Several of these effects were related to lung function, which is a difficult end-point to use in a cost–benefit analysis, as it does not necessarily correspond exactly with symptoms and is therefore hard to give a value in monetary terms (The end-point could be used in cost–effectiveness analysis). Nonetheless, the finding in some studies of an effect on lung function that is stable to adjustment for other pollutants provides some support for the studies that examine respiratory symptoms. Publications from the Southern California Children’s Health Study include a paper on bronchitic symptoms in asthmatic children; the paper used multipollutant models for a variety of PM metrics, as well as NO2 and ozone (McConnell et al., 2003). While the effects of NO2 and organic carbon were difficult to separate, associations with NO2 were found both across communities and across different years within communities, whereas associations with organic carbon were only robust in the latter case. An adjusted coefficient from this study could be used, although the end-point of bronchitic symptoms in asthmatics, as a result of year-to-year variations in annual average NO2, is rather unusual, and there are an insufficient number of studies with adjusted coefficients to perform a meta-analysis. It is difficult to judge whether to use this coefficient for a central analysis. For the evidence currently available, it is the least uncertain of the possible concentration–response functions, as it has an adjusted coefficient and is for a respiratory outcome. Although it is only one study, it is supported in a general way by epidemiological long-term exposure studies that show effects on lung function. For now, before further studies are available, we suggest the use of the adjusted coefficient in a central analysis (item d(i)) above, with acknowledgement of the greater degree of uncertainty (see the summary in section 4 below), compared with other central analysis effects.

There are meta-analyses available for NO2 and asthma incidence (Anderson, Favarato & Atkinson, 2013b) within communities (positive) and for NO2 and asthma prevalence across communities (no associations) (Anderson, Favarato & Atkinson, 2013a). There are also several studies of NO2 and asthma prevalence within communities (most positive) that would be suitable for a meta-analysis. Annual prevalence would be preferable for quantification, particularly as the asthma incidence meta-analysis was intended only for hazard assessment and, thus, did not control for the length of the follow-up period. As correlations were so close, multipollutant models were not available. Nonetheless, these studies could be used to provide coefficients for a sensitivity analysis that compares the possibilities of the effect being entirely due to black smoke, entirely due to NO2, or (more likely) somewhere in between (item d(iii) above).

Mortality. There is now a good body of evidence on NO2 and all-cause mortality in within-community cohort studies. Unfortunately, there is no recent meta-analysis, and only a few of the studies could apply two-pollutant models (see Question C2). A large cohort study investigating multipollutant models with NO2 (Jerrett et al., 2011) did not yet provide exact estimates with confidence intervals for the two-pollutant model NO2–PM2.5. However, the recent large Rome Longitudinal Study (Cesaroni et al., 2013) provided effect estimates for NO2-related natural mortality, while adjusting for PM2.5 in a two pollutant model. In addition, the Canadian study (Gan et al., 2011), which investigated the association of coronary heart disease mortality with the three pollutants PM2.5, black carbon and NO2 together, could (in theory) serve as a cautious approach in sensitivity analysis for calculating independent effects of NO2 from particulate mass and from traffic soot. NO2 had correlation coefficients with PM2.5 and black carbon of 0.47 and 0.39, respectively. It should be noted, however, that this study used a population free of cardiovascular disease at baseline, rather than the general population, so general population baseline rates would not apply. Results from more (including large) cohort studies with results on NO2 in multipollutant models are expected in the near future, together with comprehensive meta-analyses. The additional uncertainty about other supporting evidence for cardiovascular effects would need to be acknowledged. The best option would be to use an adjusted all-cause mortality estimate from the studies expected in sensitivity analysis (item d(ii) above), due to uncertainties about the causality of the cardiovascular component of all-cause mortality. Alternatively, a meta-analysis of single-pollutant models could be used in sensitivity analysis (d(iii)) to compare (but not add) the possible effects, if the relationship is due to NO2, compared with it being due to PM.

3.2 Capturing the effects of traffic pollution and/or small-scale variations in pollution

NO2, particularly primary NO2 from traffic, shows a much larger spatial variation than does PM2.5. With regard to health impact assessments, PM2.5 is not optimally suited to capture long-term effects from small-scale spatial pollution variations (Beelen et al., 2008a; Bemis, Torous & Dertinger, 2012). A health impact assessment that intends to estimate the whole burden of air pollution will miss these effects, if it relies only on PM2.5. The additional burden of small-scale variations in pollution can be estimated by using NO2, as it has been found to be associated with effects not always captured by using PM mass (Jerrett et al., 2009b), in studies where individual exposure to pollution was estimated with more spatially exact methods (such as land use regression models). Especially in Europe, NO2 has therefore been used to investigate the effect from traffic pollution (Brunekreef, 2007; Cesaroni et al., 2013).

A health impact assessment intended to estimate only the long-term health effects of traffic pollution could rely on single-pollutant model concentration–response functions for NO2, alternatively or as a sensitivity analysis to an evaluation of effects associated with PM mass. In that case, estimates from one-pollutant models could be used. The respective results can be compared, but not added. The caveat is that NO2 might not be acting as a marker for traffic in the same way it did at the time of the study − for example, due to the use of particle traps increasing emissions of primary NO2, (see Question C2).

Children’s respiratory symptoms and/or asthma symptoms. The McConnell et al. (2003) study (described in section 3.1, on NO2 per se) is less obviously suitable for a concentration–response function that represents pollution varying on a fine spatial scale, such as traffic pollution, as the study is based on a combination of cross area comparisons and yearly variations. The yearly variation within community component (which may be driven more by local pollution sources) has, however, been used in health impact assessments of local pollution changes (Perez et al., 2009b; Perez et al., 2012; Brandt et al., 2012).

The meta-analyses available for NO2 and asthma incidence (Anderson, Favarato & Atkinson, 2013b) within communities (positive) – or (better (see 3.1)) the studies of NO2 and asthma prevalence within communities (most positive) that have the potential for meta-analysis – are suitable for quantifying the effect of traffic pollution or fine scale varying pollution, in general (that is, d(iv)), where other traffic pollutants are not quantified. To quantify the effect of traffic pollution or fine scale varying pollution, it does not matter that the correlations are too close to perform multipollutant models, as NO2 is being used as an indicator (with an acknowledgement of the caveats).

Mortality. Similarly, a meta-analysis of single-pollutant models of long-term exposure to NO2 and all-cause mortality could be used for the same (item d(iv) above) scenarios. For the reasons given before, we propose concentrating on all cause mortality. If desired, however, there would be less uncertainty in using a cardiovascular mortality concentration–response function than there would in using a NO2 per se calculation. This is not only because there is no need to distinguish between NO2 and PM (or at least traffic PM) in the epidemiology studies, but also because the toxicology that supports an effect of PM on cardiovascular outcomes also supports the epidemiology on a cardiovascular effect of traffic pollution, for which NO2 is being used as a marker.

4. Summary

In a cost–benefit analysis, it can be helpful to group concentration–response functions according to their uncertainty. The cost–benefit analysis can start with a core set of functions and then be rerun with the addition of concentration–response functions in groups of increasing uncertainty. This can illustrate the degree to which the cost–benefit ratio depends on uncertain concentration–response functions. To illustrate the potential for such an exercise, the summary below is ranked on the basis of increasing uncertainty.

The recommendations are also summarized in Table 10, by both health impact assessment context (item d(i–iv) from section 1 above) and the level of uncertainty. The ranking by uncertainty has been applied to the NO2 per se (central or sensitivity analysis) part of the table (three left-hand columns). There is, of course, uncertainty in using NO2 as a marker for traffic, but the criteria for judging greater or lesser uncertainty will be different – for example, the presence or absence of multipollutant models will not be relevant. The uncertainty of assignment of health effects to one pollutant or another may be reduced in the future if biomarkers can be developed that are specific to mechanistic pathways of health effects that differ between pollutants and are convenient for routine measurement in population studies.

The concentration–response functions, listed by increasing uncertainty, are:

least uncertainty:

respiratory hospital admissions, adjusted coefficient, using the averaging times as in the relevant studies;

increased degree of uncertainty:

  • all-cause mortality (short term), adjusted coefficient, using the averaging times as in the relevant studies, increased degree of uncertainty due to the component of cardiovascular mortality, where there is an absence of a solid body of supporting chamber studies and toxicological evidence;
  • bronchitic symptoms in asthmatic children, adjusted coefficient, long-term (year-to-year variations) from McConnell et al. (2003) – only study of long-term exposure to NO2 and respiratory symptoms in children that adjusted for a wide range of other pollutants;

most uncertainty:

  • cardiovascular admissions (short term), adjusted coefficient, using the averaging times as in the relevant studies – sensitivity analysis only, additional to quantification of PM;
  • asthma prevalence, adjusted coefficient unavailable, annual average – for pollutant specific applications, use a sensitivity analysis of comparing effects of either black smoke or NO2 (d(iii));
  • all cause mortality (long term), it may not be possible to find an adjusted coefficient from a suitable study and the element of cardiovascular mortality adds extra uncertainty; for pollutant-specific applications with PM included, use only as a sensitivity analysis.

For the pollutant–outcome pairs in the category “most uncertainty”, there was increased difficulty controlling for confounding by other pollutants, less data or less available supporting clinical or toxicological evidence.

Question C5

Is there any new evidence on the health effects of air emissions of arsenic, cadmium, mercury, lead and nickel (and their compounds) that would impact upon current target values?

Answer

Arsenic. Yes, there is some new evidence on the cancer risk of air emissions of arsenic, but it is contradictory in terms of the direction of risk. This new evidence is insufficient to have an impact on the current EU target value.

Cadmium. Yes, there is new evidence on the health effects of air emissions of cadmium. Reaching the present WHO air quality guidelines and EU target values does not prevent increasing cadmium levels in agricultural soil by air deposition, and thereby contributing to adverse effects on health in the general population. If the WHO air quality guidelines are reviewed, this new evidence should be considered.

Mercury. No, there is no new evidence on the health effects of air emissions of mercury that would have an impact on the current policy.

Lead. Yes, there is definitely new evidence on the health effects of air emissions of lead that would have an impact on the current limit value. This evidence shows that effects on the central nervous system in children and on the cardiovascular system in adults occur at, or below, the present standards in the WHO air quality guidelines and EU.

Nickel. Yes, there is some new evidence on the health effects of air emissions of nickel, but this would probably not have any significant impact on the risk estimate and the present target value.

Rationale

1. Arsenic

Present WHO air quality guidelines

Exposure to arsenic occurs in inorganic and organic forms, and in most cases oral intake predominates. The critical effect of inhalation of inorganic arsenic is considered to be lung cancer. The 2005 global update of the WHO air quality guidelines (WHO Regional Office for Europe, 2006) used the unit risk of 1.5 × 10-3, based on the risk of lung cancer (WHO Regional Office for Europe, 2000). Thus, a lifetime exposure to 6.6 ng/m3 would cause an excess risk of 10-5. For genotoxic carcinogens, such as arsenic, the 2005 global update of the WHO air quality guidelines did not present any guideline level. The evaluation was based mainly on three occupational smelter cohorts (Tacoma and Montana in the United States, and Ronnskar in Sweden). An updated pooled analysis of these cohorts was extrapolated and transformed into unit risk of 1.5 × 10-3 used by the guidelines (Viren & Silvers, 1994). The EU target value for annual average arsenic in PM10 is 6 ng/m3 (EU, 2005). Typical ambient arsenic concentrations in England are about 1 ng/m3 (EPAQS, 2009). Inhalation is a minor part of total exposure.

Later reviews

The United States Agency for Toxic Substances and Disease Registry reviewed the health risks of arsenic (ATSDR, 2007), but the Agency provides minimal risk levels for non-cancer end-points and did not publish any guideline for inorganic arsenic. There is a draft integrated risk information system (IRIS database) document from the EPA (2010b), and it does not propose any guideline limit (inhalation reference concentration, Rfc).

A United Kingdom expert panel evaluated the health risk of inhalation of inorganic arsenic (EPAQS, 2009). As a starting point, it used the midpoint of estimated cumulative exposure (125 µg/m3 multiplied by the number of years) in the lowest stratum in the Swedish smelter study with a significant increase in lung cancer. For a 40-year working life, this gave a lowest-observed-adverse-effect level (LOAEL) of 3 µg/m3. It was divided: by 10, to obtain a presumed no-observed-adverse-effect level (NOAEL); by 10, to obtain a longer exposure time for the general population; and by 10, to obtain the possible susceptible groups. With a factor of 1000, a guideline of 3 ng/m3 was proposed for the PM10 fraction, as an annual average.

Recent studies

We found three studies published after 2005 that are relevant to the risk assessment for inhaled arsenic. The first, by Jones et al. (2007), found no significant association between cumulative arsenic exposure and lung cancer in a United Kingdom smelter, but did find significant associations when less weight was given to exposures that occurred long before the outcome. A statistically significant increase in the RR of lung cancer was found in the stratum with a mean (weighted) cumulative exposure to arsenic of about 1.5 mg/m3 multiplied by the number of years. This is higher than the LOAEL in the Swedish smelter study cited above, but the point estimates of strata with lower exposure are not incompatible with risk estimates in previous reviews.

The second study, by Lubin et al. (2008), reanalysed the Montana cohort and found a higher RR at high-intensity exposure to arsenic (for example, 0.6 mg/m3 for 5 years) compared with low-intensity exposure (for example, 0.3 mg/m3 for 10 years) at the same level of cumulative exposure. This makes sense in view of possible limitations in demethylation and/or detoxification ability. If it is true also for low-level exposure, arsenic in ambient air would be less risky than indicated in previous estimates made from occupational exposure. However, the conclusions are not generally accepted, and the results cannot be directly transformed into a unit risk estimate.

The third study, by Smith AH et al. (2009), compared the RRs of lung cancer of inhaling inorganic arsenic at the Tacoma smelter with the risk at ingestion of inorganic arsenic in drinking water in Chile, using arsenic concentrations in urine. They found that the excess RR of lung cancer versus urinary arsenic level was similar in the two cases. This would indicate that the absorbed dose is carrying the risk, independent of whether it was inhaled or ingested. This is significant. If cancer risk estimates for arsenic in drinking-water (NAS, 2001), calculated as a function of arsenic concentrations in urine, could be transformed into air levels of arsenic, this would be an alternative way of estimating the cancer risk for the general population at low-level arsenic concentrations in ambient air. Using assumptions on absorbed arsenic by inhalation and ingestion, the United States National Academy of Sciences (NAS) estimate of the excess absolute risk of lung cancer would transform into a unit risk of 1 x 10-3, very similar to the 1.5 x 10-3 estimate calculated from completely different data. However, one consequence of this notion is that we must then also consider urinary bladder cancer. It is generally accepted (NAS, 2001) that intake of arsenic in drinking water also increases the risk. The NAS estimates for the United States population for ingesting 10 µg arsenic per day from water for this site (based on Taiwanese data) is 12–23 per 10 000 (lifetime risk).

Evaluation

In the WHO air quality guidelines, a unit risk of 1.5 x 10-3 was proposed, based on extrapolations from cumulative exposure to arsenic in smelter cohorts. In the past decade, studies have suggested that the true unit risk could be lower or higher than that. Another technique, applying an unsafety factor of 1000 to the LOAEL in one of the smelter cohorts resulted in a guideline value of 3 ng/m3. In summary, the new evidence is insufficient to have an impact on the current EU target value.

2. Cadmium

Present WHO air quality guidelines

The main human exposure sources of cadmium are diet (higher uptake at low iron stores, making women usually more exposed than men) and smoking. The most well-known health effects of cadmium are kidney damage and toxic effects on bone tissue (osteomalacia and osteoporosis). Cadmium has been classified by the International Agency for Research on Cancer as a Group 1 human carcinogen, mainly due to the increased risk of lung cancer from occupational exposure to cadmium. In the WHO air quality guidelines (WHO Regional Office for Europe, 2000), the data behind the classification of cadmium as carcinogenic was considered to be complicated to interpret due to concomitant exposure to arsenic. Therefore, no unit risk of lung cancer, based on these studies, could be derived. The 2000 WHO Regional Office for Europe air quality guidelines noted that average kidney cadmium levels in Europe are very close to the critical level for renal effects. A further increase in dietary intake of cadmium, due to accumulation of cadmium in agricultural soils, must be prevented. Therefore, a guideline value of 5 ng/m3 was set for cadmium in air (WHO Regional Office for Europe, 2000). This was also applied as an EU target value (EU, 2005). Present levels of cadmium in air are 0.1–1 ng/m3 in rural areas, 1–10 ng/m3 in urban areas, and higher than 10 ng/m3 in some industrial areas (WHO Regional Office for Europe, 2007). Inhalation is a minor part of total exposure, but ambient levels are important for deposition in soil and, thereby, dietary intake.

Later reviews

WHO working group on long-range transboundary air pollution. A WHO Regional Office for Europe working group published the document Health risks of heavy metals from long-range transboundary air pollution (LRTAP) (WHO Regional Office for Europe, 2007). Emissions, depositions, air levels, and health risks were reviewed, including the impact of other factors (such as fertilizers and sewage). Information on the effects of low-level exposure to cadmium on markers of renal function and bone was updated. The working group mentioned two critical effects at low-level exposure to cadmium: excretion of low molecular weight proteins, due to tubular cell damage, and also osteoporosis. Studies on cadmium balance in topsoil in Europe indicated that the amount of input exceeds that of removal. The working group noted that several European studies in the late 1990s and the beginning of the 2000s showed effects on kidney and/or bone at environmental exposure levels for urinary cadmium as low as 0.5–2.0 µg/g creatinine (µg/gC). The working group proposed a LOAEL of 2 µg/gC. The evidence of lung cancer from inhalation of cadmium was considered to be rather weak. The margin of safety for adverse effects on the kidney and bone is very narrow for the European population and non-existent for sensitive subgroups, such as women with low iron stores. Since food represents more than 90% of the cadmium intake in non-smokers, and no decline has been shown, further efforts should be made to reduce cadmium emissions.

The United States Agency for Toxic Substances and Disease Registry reviewed the health risks of cadmium and recommended a minimal risk level for this hazardous substance (ATSDR, 2012). As a point of departure, the Agency selected the lower confidence limit for cadmium concentrations in urine, resulting in a 10% increase of the excretion of beta-2-microglobulin in one of the European studies – 0.5 µg/gC; it chose the study that found effects at the lowest cadmium concentration in urine. Based on a toxicokinetic model, long-term inhalation of 0.1 µg/m3, combined with the average dietary cadmium intake in the United States population, would result in cadmium concentrations in urine of 0.5 µg/gC. An uncertainty factor of 9 was applied to the cadmium concentration in air, and thus a minimal risk level of 10 ng/m3 was set.

The European Food Safety Authority (EFSA) reviewed cadmium in food (CONTAM, 2009). In a meta-analysis, it found the lower limit of the benchmark dose for a 5% increased prevalence of “elevated” beta-2-microglobulin to be 4 µg/gC for the cadmium concentration in urine. After adjusting for the interindividual variability of the cadmium concentration in urine, a critical concentration of cadmium in urine of 1 µg/gC was derived. EFSA also reviewed data on cadmium effects on bone: “studies summarized indicate a range of U-Cd [cadmium concentrations in urine] for possible effects on bone, starting from 0.5 µg/gC, which is similar to the levels at which kidney damage occurs”. EFSA also reviewed studies on cancer, including those based on environmental exposure (lung cancer: Nawrot et al., 2010; endometrial cancer: Akesson, Julin & Wolk, 2008). The data on hormone-related cancers were considered to need confirmation from other studies.

The Joint WHO/FAO Expert Committee on Food Additives (JECFA) used essentially the same studies on beta-2-microglobulin as did EFSA, but it used a slightly different statistical modelling technique and came to the conclusion that the point estimate for the break point for elevated beta-2-microglobulin was 5.2 µg/gC for cadmium concentrations in urine (JECFA, 2011). This was then transformed to a dietary intake of 0.8 µg/kg/day (lower confidence limit). The effects on bone were not considered.

The International Agency for Research on Cancer recently updated their evaluation of cadmium and cadmium compounds (IARC, 2012). The Agency found epidemiological support for lung cancer in humans from inhalation of cadmium and also found sufficient evidence of lung cancer in animals. Therefore, cadmium and cadmium compounds are carcinogenic to humans (Group 1), mainly based on the increased risk of lung cancer.

Recent studies

The above-mentioned reviews include literature up to 2006–2008. Thereafter, the published studies of major interest are of two types.

  1. Some studies indicate that the associations between low-level exposure to cadmium and excretion of low molecular weight proteins shown in several other studies may not be due to cadmium toxicity. Instead, co-excretion of cadmium and proteins is more likely to be caused by physiological factors, such as varying reabsorption of cadmium and proteins in renal proximal tubules (Chaumont et al., 2012; Akerstrom et al., 2013).
  2. Other published studies showed effects on bone at low-level exposure to environmental cadmium (Gallagher, Kovach & Meliker, 2008; Schutte et al., 2008; Wu, Magnus & Hentz, 2010; Nawrot et al., 2010; Thomas et al., 2011; Engström et al., 2011, 2012), although some did not find positive associations (Rignell-Hydbom et al., 2009; Trzcinka-Ochocka et al., 2010).
Evaluation

Research performed in the new millennium has indicated adverse effects of long-term dietary cadmium on kidney and bone at cadmium concentrations in urine commonly seen in most European countries – about 1 µg/gC. The WHO Regional Office for Europe air quality guidelines for 2000 is still valid; further increase of cadmium in agricultural soils must be prevented. The cadmium input in European agricultural soils is larger than the output, suggesting that the cadmium intake will not decrease. Overall, deposition from cadmium in air contributes typically about half of the cadmium input to soils. Present levels of cadmium in air are too high to obtain a cadmium balance in soils (WHO Regional Office for Europe, 2007). This should be taken into account when deciding whether the WHO air quality guidelines should be reconsidered.

3. Mercury (Hg)
Present WHO air quality guidelines

Humans are exposed to several mercury species, the two most important being elemental mercury vapour (Hg0) and methylmercury (MeHg). Exposure to Hg0 is mainly via inhalation of dental amalgam fillings (about 50% mercury by weight). In subjects without such fillings, exposure occurs by inhalation of ambient air, with a typical level of 1–3 ng/m3 (total mercury, most of which is Hg0), or indoor air, which may have ten times higher levels if occupied by people with amalgam fillings. Since about 80% of inhaled Hg0 is absorbed, 3 ng/m3 will result in an uptake of about 50 ng of mercury a day. The uptake from a person with a dozen amalgam fillings is usually about 100 times higher. The WHO air quality guidelines background document considered this as well as other routes of exposure.

Exposure to MeHg occurs by gastrointestinal absorption (about 90%), from dietary consumption of food – fish, in particular. In people without dental amalgam fillings, MeHg intake from fish is the predominant exposure route. It is well known that long-term occupational exposure to Hg0 may affect the kidney and the central nervous system adversely. According the WHO Regional Office for Europe air quality guidelines for 2000, LOAELs for occupational settings are air levels of 15–30 µg/m3. After correcting for some measurement issues and inhaled volumes of air, an uncertainty factor of 20 was used, and the guideline value for ambient air was set at 1 µg/m3. There is no EU target value for mercury in ambient air.

Later reviews

WHO CICAD. WHO published a review in the Concise International Chemical Assessment Documents (CICADs) series (WHO, 2003). Most information goes back, however, to the WHO International Programme on Chemical Safety document for mercury from 1991 and the Agency for Toxic Substances and Disease review from 1999. As the starting point, the authors of the review considered the subtle effects on the central nervous system of long-term occupational exposure to Hg0 to be the result of about 20 µg/m3 of Hg0. For inhalation by the general public, this corresponds to 5 µg/m3, and an uncertainty factor of 30 resulted in a tolerable concentration of 0.2 µg/m3.

Chapter in Handbook on the toxicology of metals. This handbook (from 2007), edited by Nordberg et al., is the so-called bible of metal toxicology. The 55-page chapter on mercury (Berlin, Zalups & Fowler, 2007) summarizes the information on mercury toxicity in a way similar to that of the CICAD document, but includes some more recent references – for example, a meta-analysis from 2002 on the neurobehavioral effects of exposure to Hg0 in relation to urinary mercury levels (Meyer-Baron, Schaeper & Seeber, 2002). The evaluation of exposure–response is, however, similar to that of the WHO expert groups behind the WHO Regional Office for Europe air quality guidelines for 2000 and the WHO CICAD document from 2003. The Handbook also includes a separate chapter on interactions among metals.

WHO working group on LRTAP. The LRTAP document (WHO Regional Office for Europe, 2007) reviewed the data on emissions, deposition, air levels, and health risks in human beings. For health risks, the working group referred to occupational studies that indicated possible effects on the central nervous system after a long-term exposure to Hg0 of about 20 µg/m3. The working group also discussed the risks of MeHg exposure, concluding that priority should be given to lowering the MeHg levels in fish. Reductions in mercury emissions to air are therefore warranted.

Recent studies

Some additional review papers (such as Clarkson & Magos, 2006) are either similar to the CICAD document or the Handbook chapter, or they are less complete than these. An EU review on the safety of dental amalgam has also been performed (SCENIHR, 2008). A large number of papers were published since the turn of this century, but none of them yields new evidence on exposure–response relationships. Important studies about very low-level exposure to Hg0 include the two large randomized controlled trials of dental amalgam in children, which gave no support for adverse effects on the central nervous system (Bellinger et al., 2006; DeRouen et al., 2006).

Evaluation

The basis for determining a LOAEL for occupationally exposed workers has not changed. With regard to which so-called transformations should be used to go from occupational to environmental exposure, we consider those made by the CICAD document are more justified than those of the WHO air quality guidelines working group. However, there is no new evidence on the health effects of air emissions of mercury that would have an impact on the current policy.

4. Lead
Present WHO air quality guidelines

The WHO Working Group on Air Quality Guidelines noted that cognitive impairment has been shown in children at blood lead levels of 100–150 µg/l and proposed a critical level of 100 µg/l. To assure that at least 98% of schoolchildren have blood lead levels of less than 100 µg/l, the median should not exceed 54 µg/l. The Working Group then assumed a baseline value of the (dietary) contribution to lead in blood of 20 µg/l in uncontaminated areas. In air, 1 µg/m3 of lead was considered to increase the blood lead level by 50 µg/l (19 µg/l directly by inhalation and the rest indirectly). The Working Group aimed at a lead level in air that would not increase blood lead to a level above 50 µg/l, including the baseline; thus, lead in air should contribute no more than 30 µg/l. The target for lead in air was therefore set at 0.5 µg/m3. The same value has been adopted as the EU target value (EU, 2005).

Background levels in Europe are below 10 ng/m3 (CONTAM, 2010), but they may be higher close to certain industrial sources. Levels have declined dramatically in cities after banning lead in gasoline; they were previously often on the order of 0.5–1.0 µg/m3 in large cities (WHO Regional Office for Europe, 2000, 2007). Inhalation of ambient air is a minor part of total exposure, but ambient levels are important for contamination of soil and therefore for children’s exposure.

Later reviews and recent original papers

Several reviews show that the adverse effects of lead in children and adults occur at much lower exposure levels than those that result in a blood lead level of 100 µg/l. The recent reviews – for example, by CONTAM (2010), by JECFA (2011) and by the United States National Toxicology Program (NTP, 2012) – use the pooled analyses by Lanphear et al. (2005). These reviews also consider the effects of exposure to lead on blood pressure and hypertension in adults, but here we put more focus on cognitive effects in children, since this will be the critical effect in deciding on target values.

The most recent review is the one by JECFA (2011). JECFA used a benchmark dose or central estimate of blood lead level of 20 µg/l for an intelligence quotient (IQ) cognitive function decrement of one point in children. The lower confidence limit was 10 µg/l. JECFA transformed blood lead level into dietary intake and chose a bilinear model that yielded a 0.5 IQ point decrease at 12 µg lead/day (0.6 µg/kg/day for a 20 kg child).

If we assume that the relationship in the WHO Regional Office for Europe air quality guidelines for 2000 is correct, lead in air of about 0.2 µg/m3 would increase blood lead levels by about 12 µg/l. Even inhalation alone at this level of lead in air would increase the blood lead level by about 4 µg/l.

For effects on blood pressure in adults, the estimate was an increase of systolic blood pressure of 0.3 mm Hg per increase in blood lead level of 10 µg/l. In the WHO Regional Office for Europe air quality guidelines for 2000, a lead level in air of 1 µg/m3 was assumed to increase the blood lead level by 16 µg/l in adults. Assuming a lead level in air of 0.2 µg/m3, this would transform into an increase (point estimate) of the blood lead level by about 3 µg/l and an increase of systolic blood pressure by about 0.1 mm Hg. We consider the above-mentioned effect on children’s cognitive function to be more important.

Evaluation

It is obvious that the previous evaluation performed by the WHO Working Group on Air Quality Guidelines is not compatible with the evaluations done in later reviews, including those performed by the EU and WHO. The new evidence shows that effects on the central nervous system in children and on the cardiovascular system in adults occur at, or below, the present standards in the WHO air quality guidelines and EU.

5. Nickel
Present WHO air quality guidelines

The WHO Working Group on Air Quality Guidelines reported that ambient levels of nickel in air are about 1–10 ng/m3 in urban areas, but much higher in certain industrial areas. They noted that nickel is a human carcinogen (lung and nasal sinus) and referred to an EPA unit risk estimate of 2.4–4.8 x 10-4, depending on nickel compounds. The Working Group used data on cumulative exposure to nickel in Norwegian refinery workers (Andersen et al., 1996) as the basis for a transformation to a unit risk for environmental exposure of 3.8 x 10-4, corresponding to the excess lifetime lung cancer risk of 10-5 at 25 ng/m3. It is unclear how this calculation was performed. The EU target value is 20 ng/m3 (EU, 2005). Ambient levels are usually below 5 ng/m3, but are higher close to certain metal industries (EPAQS, 2009). Inhalation is a minor part of total exposure.

Later reviews and recent original papers

ATSDR (2005) refers to EPA data from 1986 and to the EPA evaluation of a unit risk of 2.4–4.8 x 10-3 – that is, ten times higher than that of the WHO air quality guidelines.

Although not a formal review, Lippmann et al. (2006) report findings in mice exposed to concentrated ambient particles, as well as their findings when doing new analyses on previous time series analyses of mortality and PM10 in NMMAPS. Moreover, previous literature on the cardiovascular effects of nickel is summarized. Lippmann et al. found effects on heart rate variability related to the content of nickel, vanadium, chromium and iron in ApoE-/- mice exposed to concentrated ambient particles, and of the effects of nickel and vanadium on the risk estimate for PM10 in NMMAPS. In both cases the effect of nickel was the strongest.

The United Kingdom Expert Panel on Air Quality Standards (EPAQS, 2009) used an exposure–response model from Seilkop & Oller (2003) and Norwegian studies by Grimsrud et al. (2002, 2003) and came to the conclusion that there is an increased risk of cancer for occupational exposure for 40 years at a level of 20 µg/m3 of nickel in air. Using an uncertainty factor of 1000, the Panel recommended a guideline value (annual average) of 20 ng/m3.

The International Agency for Research on Cancer (IARC, 2012) updated the epidemiological data of occupational cohorts exposed to nickel (latest reference 2009). The risk estimates are similar to those used by the United Kingdom Expert Panel on Air Quality Standards. According to the International Agency for Research on Cancer, nickel is a Group 1 human carcinogen.

Haney et al. (2012) used the Grimsrud cohort and an older United States cohort to assess the cancer risk and estimated the unit risk to be 1.7 x 10-4.

After the above-mentioned paper by Lippmann et al. (2006), several reports followed up on the contribution of nickel to the adverse effects on health of fine PM and found some support for the hypothesis that fuel combustion in power generation (which usually emits nickel and vanadium) could contribute to the risk of cardiovascular disease (Bell et al., 2009b; Zhou et al., 2011; Ostro et al., 2011; Suh et al., 2011). The strongest indication is from a recent case-crossover study on stroke by Mostofsky et al. (2012), who found nickel to be the element that showed the strongest (although non-significant) association with stroke incidence, surpassed only by black carbon.

As indicated above, the WHO air quality guideline for nickel is based on its carcinogenicity. Nickel is a normal constituent of ambient air and one of the (many) components suspected to carry the risk of fine PM.

Evaluation

There is some updated occupational epidemiology on nickel refinery workers since the review by the WHO Working Group on Air Quality Guidelines for 2000. The impression is, however that this new data will not change the previous unit risk estimate substantially. Data on the effect of ambient nickel levels on cardiovascular risk are yet too limited to permit their use in WHO air quality guideline standards.

Question C6

Is there any new evidence on health effects due to air emissions of polycyclic aromatic hydrocarbons that would impact upon current target values?

Answer

Some polycyclic aromatic hydrocarbons (PAHs) are potent carcinogens, and they are often attached to airborne particles, which may also play a role in their carcinogenicity. As PAHs are carcinogenic by a genotoxic mode of action, their levels in air should be kept as low as possible. There is new evidence linking PAH exposure to cardiovascular end-points, but at present these effects of PAH exposure cannot be separated from the effects of particles and therefore cannot impact on the target values. Studies on early biological effects of PAH exposure based on biomarkers, including PAH-DNA adducts, in general populations of children and adults also suggest a range of potential non-carcinogenic effects. Overall, there is no new evidence from which to propose a new target value. However, it should be noted that, based on previous literature, the existing target value of 1 ng/m3 of benzo[a]pyrene is associated with the lifetime cancer risk of approximately 1 x 10-4.

Rationale

In the context of air pollution, PAHs containing two or three rings are almost entirely present in the vapour phase. Those containing five rings or more (including benzo[a]pyrene) are found predominantly in the particle phase. Four-ring compounds are also particle bound, but have more seasonal variability between phases. The majority of particle-bound PAHs are associated with small particles – that is, smaller than 2.5 µm (EC, 2001).

PM, of which the most studied type is diesel exhaust particles, consists of elemental carbon to which is bound inorganic (such as metals) and organic compounds. The 2005 global update of the WHO air quality guidelines (WHO Regional Office for Europe, 2006) concluded that the health effects of diesel exhaust particles and, possibly, other types of particles are mediated by chemicals adsorbed on to their surfaces, rather than due to the particle core, and that, among these chemicals, the organic constituents – in particular the PAHs or their nitro- or oxy-derivatives – are likely to be toxicologically active.

While a high burden of particles almost completely free of organic mutagens was able to produce tumours in rat lung after chronic inhalation exposure, the carcinogenic potency of particles through non-genotoxic mechanisms at much lower concentrations in ambient air is not known. Under such conditions, the genotoxic action of PAHs and derived mutagenic substances attached to the particles might well be a more significant risk factor (WHO Regional Office for Europe, 2000). However, many studies that have considered exposure to, and the health effects resulting from, PM have not specifically addressed the issue of the concentrations or influence of particle-bound PAHs.

The major anthropogenic emission sources of PAHs are: domestic, mobile, industrial and agricultural. Domestic sources are mainly heating and cooking based on the combustion of fossil fuels. Mobile sources are from transport reliant on combustion engines, either gasoline or diesel fuelled. Catalytic converters for gasoline engines markedly reduce PAH emissions (by up to 90%); equivalent devices for diesel engines also reduce PAH emissions, but not to the same extent as for gasoline engines. For both fuels, an additional source of PAHs in the exhaust is the presence of PAHs in the fuel itself. In some southern European cities, motor scooters with two-stroke engines (fuelled by a mixture of petrol and oil) may represent a significant source of PAH emissions. Industrial sources of PAHs (such as aluminium, steel and coke production; commercial heat and power; waste incineration; creosote, bitumen and asphalt production; and petrochemical and related industries) are comparatively well understood and are being regulated increasingly. Agricultural sources (such as from burning stubble) are more variable, but may nevertheless contribute significantly to PAH levels at certain locations.

Most of the carcinogenic potential of PAHs resides with four- to seven-ringed compounds. The relevant exposure route for the lung is via inhalation of PAHs associated with airborne particles. The unit risk (lifetime exposure to a mixture represented by 1 ng/m3 benzo[a]pyrene), based on a number of occupational studies, is in the range of 80–100 x 10-6. The WHO estimate of a unit risk quoted by the EC as 8.7 x 10-5 (established by WHO in 1987) results in the increased risk associated with benzo[a]pyrene concentrations of 0.01, 0.1, and 1.0 ng/m-3 being 1 x 10-6, 1 x 10-5 and 1 x 10-4, respectively (EC, 2001). This risk estimate (8.7 x 10-5) is as stated by WHO in both 2000 and 2010 (WHO Regional Office for Europe, 2000, 2010), although in both documents the excess lifetime risks of 1 x 10-6, 1 x 10-5 and 1 x 10-4 are given as 0.012, 0.12 and 1.2 ng/m3 benzo[a]pyrene, respectively. The current EU guideline value for benzo[a]pyrene is 1.0 ng/m3, which equates with a lifetime cancer risk of 1 x 10-4 (EC, 2012). In 1999, the United Kingdom Expert Panel on Air Quality Standards recommended a lower value for the air quality standard – namely, 0.25 ng/m3 – as an annual average, deriving this figure from consideration of the lower end of the range of concentrations with observable effects in occupational exposure scenarios (DEFRA, 1999).

Even in the absence of new evidence, the acceptability of the level of risk associated with the current target value should be reviewed and discussed. The current lifetime cumulative risk for benzo[a]pyrene causing cancer (1E-04) that is associated with the current guideline (1 ng/m3) is somewhat high. According to the EPA (EPA Region 8, 2013)

The level of total cancer risk that is of concern is a matter of personal, community, and regulatory judgment. In general, the EPA considers excess cancer risks that are below about 1 chance in 1 000 000 (1×10-6 or 1E-06) to be so small as to be negligible, and risks above 1E-04 to be sufficiently large that some sort of remediation is desirable. Excess cancer risks that range between 1E-06 and 1E-04 are generally considered to be acceptable (see Role of the Baseline Risk Assessment in Superfund Remedy Selection Decisions (Memorandum from D. R. Clay, OSWER 9355.0-30, April 1991), although this is evaluated on a case-by-case basis and EPA may determine that risks lower than 1E-04 are not sufficiently protective and warrant remedial action.

In discussing the use of a single indicator carcinogen (benzo[a]pyrene) to represent the carcinogenic potential of the complex mixture of PAHs, the 2000 WHO air quality guidelines for Europe (WHO Regional Office for Europe, 2000) states:

BaP [benzo[a]pyrene] alone will probably underestimate the carcinogenic potential of airborne PAH mixtures, since co-occurring substances are also carcinogenic (WHO, 2000). Nevertheless, the well-studied common constituent of PAH mixtures, BaP, was chosen as an indicator, although the limitation and uncertainties in such an approach were recognized.

Although the need to analyse the levels of other carcinogenic PAHs has been emphasized (Boström et al., 2002), nevertheless a recent analysis (Delgado-Saborit, Stark & Harrison, 2011) concludes that “the relative contribution of BaP [benzo[a]pyrene] to the PAH overall carcinogenic potency is similar indoors (49%), outdoors (54%) and in the smelter environment (48%)”, suggesting the suitability of benzo[a]pyrene as a marker for the carcinogenic potentials of PAH mixtures, irrespective of the environment.

Dibenzo[a,l]pyrene is one of the most potently carcinogenic PAHs, although it has not been tested for carcinogenicity by inhalation. Estimates of its potency (potency equivalency factor) relative to benzo[a]pyrene vary, according to Delgado-Saborit, Stark & Harrison (2011), from 1 to 100. In most analyses, it is estimated to be the second contributor to the carcinogenicity of PAH mixtures (between 3% and 27%, compared with 45–73% for benzo[a]pyrene), although in one analysis it is estimated to contribute 77%, with benzo[a]pyrene second most important (17%). The United Kingdom PAH Monitoring and Analysis Network has reported that the ratio of dibenzo[a,l]pyrene to benzo[a]pyrene is relatively constant between sites with different dominating sources, at an average of 0.32:1.00 (Conolly, 2009). However, in a recent Italian study (Menichini & Merli, 2012), the ratio was lower (0.022:1.000).

In view of these analyses, there would not appear to be any advantage in diverging from the current policy of using benzo[a]pyrene as the single indicator compound for PAHs.

In 2005, the International Agency for Research on Cancer Working Group on the Evaluation of Carcinogenic Risks to Humans reclassified benzo[a]pyrene as a Group 1 carcinogen (carcinogenic to humans), based on mechanistic evidence summarized as follows (IARC, 2010).

The complete sequence of steps in the metabolic activation pathway of benzo[a]pyrene to mutagenic and carcinogenic diol epoxides has been demonstrated in experimental animals, in human tissues and in humans. Following exposure, humans metabolically activate benzo[a]pyrene to benzo[a]pyrene diol epoxides that form DNA adducts: the anti-benzo[a]pyrene-7,8-diol-9,10-oxide-deoxyguanosine adduct has been measured in populations (e.g. coke-oven workers, chimney sweeps) exposed to PAH mixtures that contain benzo[a]pyrene. The reactive anti-benzo[a]pyrene-7,8-diol-9,10-oxide induces mutations in rodent and human cells. Mutations (G→T transversions) in the K-ras protooncogene in lung tumours from benzo[a]pyrene-treated mice are associated with anti-benzo[a]pyrene-7,8-diol-9,10-oxide-deoxyguanosine adducts. Similar mutations in the K-RAS proto-oncogene and mutations in TP53 were found in lung tumours from nonsmokers exposed to PAH-rich products of coal combustion that are known to contain benzo[a]pyrene (as well as many other PAHs). In an in-vitro study, the codons in the tumour-suppressor gene TP53 that are most frequently mutated in human lung cancer were shown to be targets for DNA adduct formation and mutations induced by benzo[a]pyrene.

In addition, evaluation by the International Agency for Research on Cancer in 2011 of bitumen fumes, which contain PAHs, resulted in the following classifications: occupational exposures to oxidized bitumens and their emissions during roofing are “probably carcinogenic to humans” (Group 2A); occupational exposures to hard bitumens and their emissions during mastic asphalt work are “possibly carcinogenic to humans” (Group 2B); and occupational exposures to straight-run bitumens and their emissions during road paving are “possibly carcinogenic to humans” (Group 2B) (Lauby-Secretan et al., 2011).

Most recently, in June 2012, the International Agency for Research on Cancer evaluated diesel-engine and gasoline-engine exhausts and classified diesel-engine exhaust as “carcinogenic to humans” (Group 1). Thus far, these findings are reported in a brief summary (Benbrahim-Tallaa et al., 2012), which makes no specific evaluation of PAHs other than to mention their presence in the gas and particle phase. Thus, at the present time, it is not possible to derive any specific association from this evaluation of PAHs.

In a study of the relationship between PAH exposure and ischaemic heart disease, a positive correlation was found between mortality from this disease and both cumulative and average exposure indices for benzo[a]pyrene (Burstyn et al., 2005). For average exposures of 273 ng/m3 – that is, occupational exposures considerably higher that environmental levels – the RR was 1.64 (95% CI: 1.13–2.38). PAHs were also associated with increased systemic inflammation, which explained the association with quasi-ultrafine particle mass from traffic emission sources, more so that other organic components of PM (Delfino et al., 2010b). Another study found an association between PM, particle-bound organic compounds (including PAHs) and adverse health symptoms in survivors of myocardial infarctions (Kraus et al., 2011). However in these studies, the effects of PAHs are not fully separated from the effects of the PM to which they are bound.

A number of recent studies have examined the effects of PAH exposure on child development. Levels of PAH-DNA adducts in cord blood have been found to be associated with higher symptom scores of anxiety and depression measured at 4.8 years (Perera et al., 2011). In the same cohort, prenatal exposure to benzo[a]pyrene measured from maternal personal air monitoring (at a median level of 2.27 ng/m3), and also cord blood adduct levels, were associated with these effects, as well as attention problems at age 6–7 years (Perera et al., 2012). Similar findings come from a study of children in Poland (Edwards et al., 2010), where high PAH exposure in utero also restricted fetal growth (Choi et al., 2012). Effects on fetal development were exacerbated by obesity in African-American women (Choi & Perera, 2012).

The carcinogenic and toxicological properties of PAHs have been extensively investigated and reviewed (Luch, 2005). Their mode of action is genotoxic, and their DNA adducts elicit nucleotide excision repair mechanisms. They also induce aberrant gene expression and cell signalling and epigenetic effects that may contribute to their carcinogenic and other toxicological properties.

The utility of biomarkers for monitoring human exposure to PAHs was discussed in the 2000 WHO air quality guidelines for Europe (WHO Regional Office for Europe, 2000). Those biomarkers specific for PAHs include urinary 1-hydroxypyrene and the measurement of PAH-protein and DNA adducts by immunoassay, while the 32P-postlabelling assay for DNA adducts is more sensitive, but less specific. Overall, different biomarkers have been validated to varying extents (Gallo et al., 2008). Measurement of cytogenetic damage, including chromosomal aberrations, is not particularly sensitive for measuring environmental PAH exposure. A recent meta-analysis of occupational exposure to PAHs has concluded that micronucleus formation, chromosomal aberration and sister chromatid exchanges in peripheral blood lymphocytes are all significantly higher in workers, with ranges of exposures (where known) higher than the current target environmental level (Wang et al., 2012).

Although large scale studies have validated chromosomal aberrations as biomarkers of cancer risk (Bonassi et al., 2008), the methods would not be specific if applied to health effects of environmental PAH exposure and would not distinguish other routes of exposure (such as. dietary). Likewise, for DNA adducts, recent studies have validated these as biomarkers of lung cancer risk for smokers (Veglia et al., 2008); however, for exposures to PAHs, dietary exposures would also contribute to the adducts detected. Furthermore, the relationship between adduct levels (such as in white blood cell DNA) and ambient air levels of PAHs is non-linear, with evidence of saturation (that is, plateau) at higher levels of exposure.

Question C7

Is there any new evidence on the health effects of short term (less than 1 day) exposures to SO2 that would lead to changes of the WHO air quality guidelines based on 10 minute and daily averaging periods or the EU’s air quality limit values based on hourly and daily averaging periods?

Answer

There are no new respiratory chamber studies that would change the 10-minute guideline of 500 µg/m3, previously based on these types of studies. However, a reanalysis of the previous literature has found a small difference between responders and non-responders at 572 µg/m3 (0.2 ppm) (not statistically significant after control for multiple comparisons), the starting point for deriving the previous guideline. Thus, while the currently available statistical analysis suggests that the starting point does not need to be changed, a small increase in the safety factor from the current value of 1.15 might be justified when the time comes to reconsider the guideline, as the small (though non-significant) difference between responders and non-responders at this concentration increases the uncertainty as to whether this is a no-effect level or a minimal-effect level. Should further evidence confirm this difference, then the starting point may need to be changed in the future.

The 24-hour average guideline was based on the low end of the concentration ranges used in the time-series studies and on the Hong Kong intervention study. The time-series evidence continues to accumulate and continues to be inconsistent when adjusted for other pollutants for many (but not all) outcomes – for example, it is consistent for asthma admissions. The results of the original Hong Kong intervention study remain as a reduction in mortality for a reduction in pre- and post-intervention exposure to SO2 independent of PM10, although a more recent report suggests more difficulty in disentangling the effects of the reductions in SO2 from reductions in other constituents, such as nickel or vanadium. The new studies are at a similar range of concentrations as the previous studies, so the 24-hour average guideline does not need to be changed if the same method (using a concentration at the low end of the range of concentrations) is followed for setting the guideline.

Rationale

A literature search – for sulfur dioxide or sulphur dioxide, toxicity and health, some author and study searches, consultation of other documents that include reviews and reports (EPA, 2008a) and consultation of APED – indicates that a large number of new studies have been published since 2004 (the literature cut-off date for the 2005 global update of the WHO air quality guidelines (WHO Regional Office for Europe, 2006)). The text below concentrates on the direct health effects of SO2, but indirect effects are also possible. SO2 is an important prerequisite for urban nucleation to form particles smaller than 0.02 microns (see Question C8), although only exceedingly low concentrations are required to allow sufficient formation of sulfuric acid, to initiate nucleation. The health effects of these nanoparticles are poorly understood. The text below also concentrates only on short-term exposures to address the exact question posed.

Chamber study evidence

The WHO guideline value of 500 µg/m3 for a 10-minute average is based on evidence from human chamber studies that show reductions in FEV1 and increases in airway resistance and symptoms (WHO Regional Office for Europe, 2006). A review was recently published (Johns & Linn, 2011) following consideration of chamber study evidence by the EPA in 2008 (EPA, 2008a). The Johns & Linn review only identified two papers published since 2004 (see below) and concluded that the older studies continued to have an integral role in assessing the respiratory effects of SO2. As a result, the key statements in the review reflect very closely those made in the WHO guidelines.

A paper worth close examination by Johns, Svendsgaard & Linn (2010) pooled data on individuals from several key studies to analyse an overall concentration–response relationship in asthmatics, with a particular emphasis on responders and non-responders. The EPA (2008a) estimated that 5–30% of asthmatics during 5–10 minutes of exercise could experience moderate or greater decrements in lung function at 0.2 ppm to 0.3 ppm SO2. The Johns, Svendsgaard & Linn (2010) analysis showed a clear concentration–response relationship between 572 µg/m3 (0.2 ppm) and 2860 µg/m3 (1 ppm) (the maximum concentration examined). This provides a more formal basis for the conclusions in the WHO guidelines. It also provides clearer evidence that the response at lower doses can be split between responders and non-responders (responders being defined at higher doses). Responders showed no significant change in airway resistance and a minor 5% decrease in FEV1 that was not significant after correction for multiple comparisons at 572 µg/m3 (0.2 ppm). This does not suggest a change from the use of 572 µg/m3 (0.2 ppm) as the starting point for setting the current guideline, although further statistical modelling to consider the possible location of a threshold might be helpful.

However, because a separation in the response of responders and non-responders was still apparent at 572 µg/m3 (0.2 ppm), even if not significant with correction for multiple comparisons, it increases the uncertainty as to whether this is a no-effect level or a minimal-effect level. In addition, the studies only involve mild asthmatics. The safety factor applied to the 572 µg/m3 (0.2 ppm) level in the current guideline was only 1.15. Given the uncertainties just raised, it should be considered, in a future reconsideration of the guidelines, whether a larger safety factor would be appropriate. Further studies to give a larger pooled sample size would be needed to confirm whether or not there is a real difference between responders and non-responders at this concentration. If this was confirmed, the starting point for the guideline would need to be changed.

The only other studies are: one showing no pulmonary response in healthy adults below 2 ppm (van Thriel et al., 2010); and one showing a reduction in cardiac vagal control (root mean square of the successive differences (RMSSD)) in 20 normal subjects, but not those with stable angina, 4 hours (but not 1 hour) after exposure to 572 µg/m3 (0.2 ppm) SO2 for 1 hour (Routledge et al., 2006). Baroreflex sensitivity was also reduced. The implications of such changes in healthy subjects are unclear, and the EPA noted that the absolute values of the RMSSD did not differ significantly from subjects exposed to air (although the change from baseline did). Of the stable angina patients, 70% were on beta blockers, which may have protected them from adverse changes. More studies are needed to confirm this finding.

Panel studies

The EPA considered the possibility that there was some evidence for SO2 having an effect on heart rate variability, but that the number of studies was limited. Our search did not pick up any further panel studies on heart rate variability. Also, the EPA considered evidence on arrhythmias to be inconsistent. A study that found a non-significant positive association between SO2 and activation of defibrillators does not change this conclusion (Anderson et al., 2010). The EPA concluded the number of studies was too limited to come to a conclusion about inflammatory markers in the blood. Only one further study, suggesting increased levels of serum C-reactive protein in children, has been published since (Shima, 2007). Goldberg et al. (2008) found positive and statistically significant associations of SO2 with reduced oxygen saturation and increased pulse rate in congestive heart failure patients. Briet et al. (2007) (not considered by the EPA) found that 5-day average SO2 reduced endothelium-dependent brachial artery flow-mediated dilatation in healthy male subjects along with nitric oxide, but not with other pollutants, such as NO2 and PM.

A multicity study by Schildcrout et al. (2006) was highlighted as part of the established evidence for an effect of SO2 on asthma symptoms in children (EPA, 2008a). This showed an effect on asthma symptoms, but not on rescue inhaler use. The risk was increased in joint models with NO2 and carbon monoxide (which were found to be more important pollutants in this study) and unchanged in a joint model with PM10 (with only a marginal loss of statistical significance). PM10 had no effect in this study. The EPA regarded the evidence as more mixed for symptoms in adults and limited for lung function in adults and children. More recent studies have found an effect of SO2, at least in single-pollutant models, for respiratory symptoms (Zhao Z et al., 2008 for indoor not outdoor SO2; Moon et al., 2009) and lung function in children (Chang et al., 2012; Liu et al., 2009), the latter not robust to control for PM2.5. In adults, increased respiratory symptoms were shown in healthy adults returning to an island after a volcanic eruption (Ishigami et al., 2008; Iwasawa et al., 2009) at high concentrations of SO2,19 but increases were statistically insignificant in chronic obstructive pulmonary disease patients (Peacock et al., 2011) at lower concentrations (maximum: 75 ppb for a24-hour average). Possible declines in lung function were found in asthmatic adults (Canova et al., 2010) (maximum: 5 ppb for a 24-hour average), but not in chronic obstructive pulmonary disease patients (Peacock et al., 2011) or healthy adults on the volcanic island (Iwasawa et al., 2009).

These more recent studies do not represent any major shift in evidence since the EPA review.

Time series evidence

The 2005 global update of the WHO air quality guidelines (WHO Regional Office for Europe, 2006) noted that observational time-series studies had reported numerous mortality and morbidity risk estimates for SO2 over the preceding decade. It was considered that the consistency of the association of SO2 with health outcomes appeared to be less than that for PM, but that the magnitude of the estimated risks was often comparable with that of PM.

Short-term exposure and mortality

Since the publication of the 2005 global update of the WHO air quality guidelines, Anderson et al. published a peer-reviewed research report that contains meta-analyses of time-series studies, based on studies up until 2006 (Anderson et al., 2007). In single-city studies, positive and statistically significant associations with SO2 were generally found for all-cause, cardiovascular, cardiac and respiratory mortality in single-pollutant models. While the majority of the single-city studies are from before 2004, there are a few post-2004 studies included in some of the summary estimates described below (Jerrett et al., 2004, Penttinen, Tiittanen & Pekkanen, 2004). Most estimates showed significant heterogeneity, except those for cardiorespiratory mortality and mortality from lower respiratory infections, but there was also significant heterogeneity for the main mortality end-points for PM10. Adjustment for publication bias, where shown (all except cardiorespiratory, cardiac and stroke mortality), reduced the size of the summary estimates, but they remained positive and statistically significant. The size of the estimates was only slightly lower than those for PM10.

There was substantial overlap between the multicity studies reviewed for the 2005 global update of the WHO air quality guidelines (WHO Regional Office for Europe, 2006) and for the Anderson et al. report (Anderson et al., 2007), but the latter covered more studies (none published after 2004). Again there were positive, statistically significant associations in single-pollutant models with all-cause, cardiovascular, cardiac and respiratory mortality, but those associations tested in multipollutant models were reduced by control for other pollutants, often substantially, and often lost statistical significance. Multipollutant models in single-city studies were not reviewed in this report.

The EPA view, published in 2008 (EPA, 2008a), covered much of the same material, but also included the Public Health and Air Pollution in Asia literature review (HEI, 2004), which found that a meta-analysis of time-series studies of SO2 and mortality in Asia gave similar results to those of European multicity studies. Overall, the EPA considered that there was suggestive evidence of a causal relationship (consistent in single-pollutant models, more uncertain after adjustment for co-pollutants). The Hong Kong Intervention study (Hedley et al., 2002) was noted for the effect of a change in SO2 on mortality, in the absence of a change in PM10; also noted was that this change in sulfur in fuel may have included changes in heavy metal content (reported reference: Hedley et al. (2006); now available in journal form: Hedley et al. (2008)).

The Hong Kong Intervention has now been studied in more detail (Wong et al., 2012). The decrease in SO2 concentrations between the pre- and post-intervention periods was accompanied by decreases in NO2 and increases in ozone. There were also decreases in metals (aluminium, iron, manganese, nickel, vanadium, lead and zinc) associated with particles, of which the decreases in nickel and vanadium were most consistent. While nickel and vanadium were associated with mortality in general terms, it was not possible to show a clear link between changes in their concentrations being associated with the intervention and with changes in their effect on mortality after (compared with before) the intervention (although there was a decline). In this more detailed study, SO2 only showed a decrease in the excess risk of respiratory mortality (not statistically significant), but also showed an increase in the excess risk of all-cause (not statistically significant) and cardiovascular mortality after (compared with before) the intervention. In the overall study (irrespective of the intervention), the effects of SO2 on mortality were not stable to adjustment for nickel and vanadium. In summary, while the intervention was still beneficial, a more detailed study suggests that identifying the responsible pollutant is difficult.

APED, used to prepare the review by Anderson et al. (2007), has now been updated to 2009 for papers on SO2. The database uses a comprehensive literature searching strategy, with sifting for study quality criteria. The database indicates seven new studies (Cakmak, Dales & Vidal, 2007; Filleul et al., 2006a; Kowalska et al., 2008; Lee, Son & Cho, 2007b; Tsai et al., 2003; Wong et al., 2008; Yang et al., 2004), for all-cause mortality, all ages, all-year and 24-hour average SO2, that had quantitative estimates and did not overlap with other studies (see Table 11). The literature search identified a further two studies meeting these criteria (Rajarathnam et al., 2011; Rabczenko et al., 2005) and a further paper has been identified since (Chen et al., 2012c). Most of the studies showed positive associations (8 of 10) of which five were statistically significant. The range of the central estimates (-1.4–3% per 10 µg/m3) were all within the range of the 144 previous estimates (-4.63–6.4% per 10 µg/m3). Although two Asian study estimates (Wong et al., 2008, Lee, Son & Cho, 2007b) were higher than the current summary estimate, it is unlikely that this would change an updated summary estimate much, given the large number of studies already in the meta-analysis.

Table 11.. Selected quantitative estimates for SO2 and mortality.

Table 11.

Selected quantitative estimates for SO2 and mortality.

In a study in Santiago, Chile, the SO2 estimate generally remained stable after adjustment for PM10, ozone and carbon monoxide (Cakmak, Dales & Vidal, 2007), and a negative association in Delhi, India, was unchanged by adjustment for PM10, with and without NO2 (Rajarathnam et al., 2011). The HEI report collating four Asian city studies (Wong et al., 2010b) found the estimate remained stable after adjustment for PM10 and ozone, but not for NO2, with the exception of Bangkok, where the estimate was reduced by both PM10 and NO2. In the study of 17 Chinese cities (Chen et al., 2012c), the estimate was reduced, but remained positive and statistically significant when controlled for PM10; the reduction on adjustment for NO2 was substantial, and the estimate lost significance. There were too few multipollutant-model studies in the earlier Asian city meta-analysis to draw an overall conclusion (HEI, 2004). This does not change the previous view (Anderson et al., 2007; WHO Regional Office for Europe, 2006) that estimates for all-cause mortality could be sensitive to adjustment for other pollutants. It is possible that the short sharp peaks of SO2 mean it is more subject to measurement error, which can affect multipollutant models. However, Zeka & Schwartz (2004) found that, using a statistical technique for multipollutant adjustment less subject to measurement error, the SO2 estimate was small and was not statistically significant.

The report on four Asian cities (Wong et al., 2010b) examined the shape of the concentration–response function with mixed results (2 of 4 linear), but none of them examined the shape after control for other pollutants.

Short-term exposure and respiratory hospital admissions

Anderson et al. (2007) found a positive and statistically significant association with respiratory hospital admissions, all ages, all year of 1.51% (95% CI: 0.84–2.18%) per 10 µg/m3 SO2 that showed evidence of heterogeneity. The association remained positive and statistically significant after adjustment for some publication bias. Multicity studies gave a similar result. There were no multicity studies that examined multipollutant models, and the single-city study multipollutant results were not reviewed. A further three studies have been added to APED since (Leem et al., 1998; Jayaraman & Nidhi, 2008; Chang, Hsia & Chen, 2002), another two meeting the criteria (Wong et al., 2010a; Cakmak, Dales & Judek, 2006b) were identified in the literature search, and another one later (Chen et al., 2010b). All six found positive associations with SO2 (two non-significant) that ranged from 0.13% to 8.2% compared with -0.5–22.5% for the previous estimates. Where there was adjustment for other pollutants (Cakmak, Dales & Judek, 2006b; Jayaraman & Nidhi, 2008; Leem et al., 1998), the estimates were reduced and remained significant in only one study (Cakmak, Dales & Judek, 2006b).

Given the chamber study findings, asthma admissions will be described here. Results for other respiratory diagnoses are described in Anderson et al. (2007). The relationship with asthma admissions was significant in children in single and multicity studies and was robust to adjustment in multipollutant models in the multicity studies (unexamined in single-city studies) (Anderson et al., 2007). Further studies published since 2006 on asthma admissions in children are also positive and statistically significant in two studies (Lee, Wong & Lau, 2006; Samoli et al., 2011a), with a range from 1.3% to 6% per 10 µg/m3 SO2 compared with 0.8% to 9% in Anderson et al. (2007). In Samoli et al. (2011a), although control for PM10 resulted in a loss of significance, it also resulted in a smaller reduction (20%) in the coefficient than when PM10 was controlled for SO2 (30% reduction). The SO2 coefficient was stable to adjustment for NO2 and ozone. The coefficient was said to be non-significant (direction not given) in a study in Hong Kong (Ko et al., 2007). Results in other age groups or all ages were more mixed (Anderson et al., 2007; Bell, Levy & Lin, 2008; Ko et al., 2007; Tsai et al., 2006a; Yang et al., 2007).

Anderson et al. (2007) reported a summary estimate per 10 µg/m3 SO2 of a 2.65% (95% CI: 0.39–4.96%) increase in asthma emergency room visits in children, all year. The summary estimate was not significant in adults and the outcome was not examined in multicity studies. The database has been partially updated for emergency room visits, identifying four studies (Jalaludin et al., 2008; Ito, Thurston & Silverman, 2007; Szyszkowicz, 2008; Villeneuve et al., 2007). The latter two found no associations, but the first two found positive and statistically significant associations in children and adults, respectively. In Jalaludin et al. (2008), the association was reduced, but remained significant, on adjustment for other pollutants; the association, however, was not robust to adjustment for NO2 in Ito, Thurston & Silverman (2007) (only the summer association was examined). The literature search identified a few other studies of asthma emergency room visits, notably a large multicity study in Canada (Stieb et al., 2009) that found a non-significant negative association. This study was for all ages. As the Anderson meta-analysis found a greater effect in children, it would have been interesting to see the analysis split by age group. A study in Toronto did find a significant relationship in children across both sexes and in the top and bottom quintiles of socioeconomic position (Burra et al., 2009). A positive and statistically significant association was also found for SO2 and emergency room visits for wheeze in children aged 0–2 years in a multicity study in Italy (Orazzo et al., 2009).

A preliminary study from Taiwan, Province of China, used daily variations in spatially modelled pollutants as the exposure metric (Modelling may be more uncertain for SO2 than for other pollutants, as SO2 concentrations are characterized more by sharp peaks, which are harder to model). The study found a negative association that was not statistically significant in both single and multipollutant models for asthma emergency room visits in all age groups (Chan et al., 2009). Overall, the conclusions are unclear for SO2 and emergency room visits for asthma, as there were mixed results in multipollutant models in the recent studies, but the larger number of earlier studies suggests the association is robust (EPA, 2008a).

Short-term exposure and cardiovascular admissions

Anderson et al. (2007) reported a summary estimate of 0.96% (95% CI: 0.13–1.79%) per 10 µg/m3 for cardiovascular admissions and 24-hour average SO2 for all ages, all year across 5 studies and a summary estimate of 2.26% (95% CI: 1.30–3.22%) for cardiac admissions across 12 studies. A similar result was found for cardiac admissions for a multicity study in eight Italian cities and a lower one in seven European cities. The multicity studies did not include a multipollutant model and the report did not collate multipollutant model estimates for single-city studies. A further multicity study from Spain meeting the same criteria has been published since (Ballester et al., 2006) and reported an estimate of 1.33% (95% CI: 0.21–2.46%). This was not stable to adjustment for carbon monoxide, but was to other pollutants. A study in Shanghai found a positive and statistically significant association with cardiovascular admissions that was stable to adjustment for PM10, but was reduced to some degree and lost statistical significance on adjustment for NO2 (Chen et al., 2010b). The study also reported a positive and statistically significant estimate for cardiac admissions as did the study by Wong et al. (2010a). The latter two estimates were not examined in multipollutant models.

Birth outcomes

The present review concentrates on short-term exposures. Most studies of birth outcomes use longer averaging times (a month or more). Nonetheless, it should be noted that a substantial number of studies have been published in this area. A recent and thorough review concluded that SO2 was associated with preterm birth, but not consistently with low birth weight or small for gestational age births (Shah & Balkhair, 2011). A review by Vrijheid et al. (2011) found some evidence for an association between SO2 and congenital cardiac anomalies. A thorough veterinary epidemiology SO2 study in Western Canada examined whether emissions from the oil and gas industry (including SO2) were associated with effects on the reproduction and health of beef cattle. The study showed that SO2 exposures were not related to: abortion or stillbirth (Waldner, 2009); histopathological lesions in the immune, respiratory or nervous systems in calves that were aborted or died postnatally (Waldner & Clark, 2009); changes in lymphocyte subtype populations in blood samples from neonatal calves or yearling cattle (Bechtel, Waldner & Wickstrom, 2009a,b); or non-pregnancy, risk of disposal in pregnant cows or calving interval (Waldner & Stryhn, 2008). Gestational (but not postnatal) exposure to SO2 above 0.9 ppb was significantly related to pathological lesions in the skeletal or cardiac muscle among calves that died (Waldner & Clark, 2009), and SO2 exposure during the last trimester or across gestation was related to calf mortality in the first 3 months of life (Waldner, 2008). This was unexpected, given that most of the extensive, detailed end-points examined in the study did not show associations.

There are, however, studies on infant mortality that use daily average concentrations as the exposure metric. Dales et al. (2004) found an association between SO2 and sudden infant death syndrome that was independent of adjustment for NO2. Hajat et al. (2007) found a positive and statistically significant association between SO2 and both neonatal and postneonatal deaths. No multipollutant modelling was performed, but there were no significant associations for other pollutants. These results were in line with earlier studies discussed by the authors, but other recent studies have found no association (Tsai et al., 2006b; Woodruff, Darrow & Parker, 2008) or a positive association that was not statistically significant (Son, Cho & Lee, 2008).

Toxicological evidence

The toxicological evidence up to about 2006/2007 has been reviewed by the EPA (2008a), although the majority of studies were pre-2004. For acute exposures and respiratory effects, it was concluded that repeated exposures to SO2, at concentrations as low as 0.1 ppm in guinea pigs, may exacerbate inflammatory and allergic responses in allergic animals. SO2 at a concentration of 10 ppm or less failed to induce airway hyperreactivity, following a nonspecific (rather than allergic) challenge in four different animal models.

For cardiovascular effects, it was noted that, in general, vagally mediated responses in the heart have been observed at lower concentrations of SO2 than have oxidative injuries from SO2 metabolites in the circulation. It was not considered that the limited toxicological evidence provided biological plausibility for an effect on arrhythmias, and the evidence for effects on blood pressure and blood markers of cardiovascular risk was regarded as inconclusive.

Perturbations in potassium-, sodium- and calcium-gated channels in hippocampal or dorsal root ganglion neurons isolated from rats at 0.01–100 µM of SO2 derivatives ex vivo were regarded as of questionable significance, given the high doses needed for effects on the nervous system in vivo.

The evidence from animal toxicological studies was regarded as insufficient to conclude that long-term exposure to ambient SO2 caused prolonged effects on lung morphology, lung function or decrements in lung host defence. There was evidence of oxidation and glutathione depletion in the hearts of rodents exposed by inhalation to SO2 above 5 ppm, but this oxidative injury was not considered relevant to cardiovascular effects seen at ambient levels of SO2.

It was concluded that toxicological studies provided very little biological plausibility for reproductive outcomes related to exposure to SO2.

The literature search and author searches for the present review identified a number of toxicology studies on SO2 published since 2004 that were not considered by the EPA. These studies, but not those reviewed by the EPA, are described below. None of the studies identified examined concentrations below 2 ppm, which exceeds considerably ambient concentrations, except for the veterinary epidemiology study in Canada (see section on “Birth outcomes” above).

Short-term respiratory effects

With a variety of assumptions, modelling of gas transport in theoretical airway models predicted that the local concentration of SO2 in the upper airways of human beings would be 3–4 times higher than in rats or dogs (Tsujino, Kawakami & Kaneko, 2005). Exposure to 50 ppm SO2 1 hour a day for 3 days, followed by ovalbumin, exaggerated chronic allergic airway inflammation and subepithelial fibrosis (Cai et al., 2008). Another study exposed rats to 2 ppm SO2 for 1 hour a day for 7 days prior to or without sensitization with ovalbumin. Prior exposure to SO2 increased the mRNA and protein expression of epidermal growth factor (EGF), epidermal growth factor receptor (EGFR) and cyclooxygenase-2 (COX-2), markers of regulation of mucus hypersecretion, and airway repair and inflammation (Li, Meng & Xie, 2008). A study that developed an animal model for chronic obstructive pulmonary disease found that 5, 10 and 20 ppm SO2 for 3 days resulted in no change in basal mucus secretory activity in trachea preparations and a decrease (not an increase) at 40 ppm and 80 ppm. No changes were found in acetylcholine stimulated secretory activity. Single cell necrosis and loss of cilia occurred at concentrations of 10 ppm and higher (Wagner et al., 2006). SO2 inhalation at l20 ppm for 6 hours a day for 7 days caused significant increases in the proto-oncogenes c-fos and c-jun mRNA and in protein levels in the lungs of rats (Qin & Meng, 2006a) and caused a further increase in the presence of benzo[a]pyrene. It was hypothesized that this might explain the co-carcinogenicity of SO2 and benzo[a]pyrene in hamsters, although the EPA view was that SO2 was not a clear co-carcinogen.

There have also been studies of lung cells in vitro. Human bronchial epithelial (BEP2D) cells were treated with a range of concentrations (0.0001–1 mM) of the SO2 derivatives sodium bisulfite (NaHSO3) and sodium sulfite (Na2SO3) in a 1:3 ratio for various durations up to 24 hours (Li, Meng & Xie, 2007). Expression of EGF, EGFR, intercellular adhesion molecule 1 (ICAM-1) and COX-2 mRNA and protein showed a dose-dependent increase that was greatest at 30 minutes. It was suggested that this would result in mucin overproduction and inflammation, if occurring in vivo. The same dose protocol (with the addition of a 2 mM dose of the SO2 derivatives) for 4 hours was applied to the same cell line, leading to mRNA and protein overexpression of c-fos, c-jun and c-myc at all doses, with other changes in proto-oncogenes and tumour suppressor genes at 0.1–2.0 mM (Qin & Meng, 2009).

Another study, in A549 cells, found significantly reduced cell viability in an air–liquid interface culture, with concentrations from 10 ppm to 200 ppm SO2 (Bakand, Winder & Hayes, 2007). In the same cell line, sodium sulfite (a derivative of SO2) at 1000–2500 µM was shown to enhance interleukin 8 (IL-8) release. IL-8 is a chemical signal involved in neutrophil recruitment and activation. IL-8 release was inhibited by a selection of asthma drugs (Yang et al., 2009). Sulfite oxidation by a mammalian peroxidase–hydrogen peroxide system, resulting in the highly reactive sulfate radical (SO4.-), has been shown for the first time in an experiment using human myeloperoxidase and human neutrophils and (bi)sulfite anions at 20–100 µM (Ranguelova et al., 2012). Ranguelova et al noted that healthy individuals have a mean serum concentration of 5 µM, but it can be raised in disease.

Short-term systemic and cardiovascular effects

A group from China published a series of papers on the systemic effects of SO2. Meng & Liu, 2007, showed morphological changes in various organs in mice after inhalation of 10.6 ppm SO2 and above for 4 hours a day for 7 days. Qin & Meng (2006b) showed a dose-related reduction in activities and mRNA levels of the detoxifying enzymes CYP2B1/2 and CYP2E1 in the lungs, and CYP2B1/2 (but not CYP2E1) in the livers of rats treated with 5.3–21.2 ppm SO2 for 6 hours a day for 7 days. Protein oxidative damage and DNA-protein cross-links were increased in the lungs, liver and heart (in that order) in mice exposed to 5.3–21.2 ppm SO2 for 6 hours a day for 7 days (Xie, Fan & Meng, 2007). Unsurprisingly, 21.2 ppm SO2 for 4 hours a day for 10 days led to oxidative stress in the livers and brains of the mice. This oxidative stress was ameliorated by moderate, but not high, levels of vitamin C and by salicylic acid (Zhao H et al., 2008).

Other than the study at high doses by Xie, Fan & Meng (2007) mentioned above, no other animal studies that reported cardiovascular toxicity were picked up in the search. A review of amino acids as regulators of gaseous signalling (Li et al., 2009) notes that endogenous SO2 (derived from cysteine) can activate guanylyl cyclase and thus elicit a variety of responses, including relaxation of vascular smooth muscle cells. A review on the same subject that is more specific to SO2 is available (Chen S et al., 2011). In a section of the Chen et al. review, on older and newer studies of the effects of exogenous SO2, the authors highlighted a study by Nie & Meng (2007) showing that the SO2 derivatives NaHSO3 and Na2SO3, in a 1:3 ratio at a concentration range of 5–100 µM, inhibited the sodium/calcium exchanger current in rat myocytes and that SO2 derivative concentrations above 10 µM increased intracellular myocyte free calcium. Nie & Meng (2007) also showed that SO2 and its derivatives can lower blood pressure in rats. The Chen et al. review also mentioned that SO2 and its derivatives can act as vasoconstrictors or vasodilators, depending on concentration.

The section of the Chen S et al. (2011) review on endogenous SO2 notes that a decrease in SO2 can protect against vascular structural remodelling in spontaneously hypertensive rats and can protect against pulmonary hypertension in rats with hypoxic pulmonary hypertension. An increase in endogenous SO2 protected against pulmonary hypertension in monocrotaline-induced hypertension (by inhibiting smooth muscle cell proliferation), but mediated myocardial ischaemia reperfusion injury. The concentrations for these studies are not discussed in the review. The authors concluded that endogenous SO2 is involved in regulation of cardiovascular function and that disturbances in this regulation can be found in disease.

Discussion

Although the chamber study evidence has not changed significantly, a pooled analysis of previous data suggested a tendency towards a split response between responders and non-responders that was statistically significant before (but not after) adjustment for multiple comparisons. This might suggest the need for a small increase in the safety factor.

Most of the newer toxicological evidence is at high doses, so it does not have direct implications for the guideline. The new finding of an association between gestational exposure to low levels of SO2 and histopathological lesions in heart or skeletal muscle in beef cattle is hard to put into context, as there are no other studies of this type. It is possible that another unmeasured pollutant present at higher concentrations is actually responsible.

The review of the time-series evidence is based on studies analysed according to current practice, but it needs to be acknowledged that there are many issues that still need further discussion. As many of these issues are shared across all pollutants, they will not be discussed in detail here. These issues include statistical model choice (HEI, 2003; Erbas & Hyndman, 2005; Ito, Thurston & Silverman, 2007) and the challenges of distinguishing the effects of different pollutants in multipollutant models (Kim et al., 2007; Billionnet, Sherrill & Annesi-Maesano, 2012). The low average concentrations of SO2, but with sharp peaks, combined with the fact that, in some studies, SO2 is controlled for PM10 that is measured only once every 6 days means that the presence of measurement error adds uncertainty to the interpretation of the multipollutant model results. More generally, exposure misclassification may be a particular issue for SO2. Sarnat et al. (2007), in a discussion of data from four cities, concluded that ambient SO2 was not well correlated with personal exposures to SO2 in most subjects. It was noted that the concentrations of 24-hour average SO2 personal exposure were very low, leading to the possibility of measurement errors in the personal exposure obscuring the relationship. In addition, the association between peak personal exposures and peak ambient concentrations may be what is of most interest. It is only necessary for these correlations to be present in some susceptible individuals, rather than the whole population, to account for the epidemiological results.

Bearing the above points in mind, the time-series evidence continues to suggest associations with mortality that are not necessarily stable to adjustment for other pollutants. The picture for respiratory hospital admissions is similar, but asthma admissions in children seem to be more stable to adjustment for other pollutants in most cases. A robust effect on asthma admissions ties in with the chamber study evidence, although the fact that associations with asthma admissions are more variable in adults does not.

Associations are also seen with cardiovascular admissions. There are fewer studies that have tested this in multipollutant models. While there is a chamber study, a toxicology study at high doses, and a handful of panel studies on cardiovascular end-points, these recent studies on their own are insufficient to support the time-series finding one way or the other.

As the 24-hour average guideline is partly based on time-series studies, a change in the guideline might be required if none of the outcome associations were stable to adjustment for other pollutants. The present document has not reviewed multipollutant model results on single-city studies published before the Anderson et al. (2007) report. Further work would be needed to do this before coming to overall conclusions as to what outcome associations are stable to adjustment for other pollutants. Currently, the associations with asthma admissions in children seem the most robust. The Hong Kong Intervention study, where SO2 was reduced sharply (but PM10 was not) was also influential in setting the guideline, but more recent work suggests less confidence in allocating the mortality benefit to SO2.

The 24-hour average guideline was influenced by the concentration ranges at which results had been shown in the time-series studies. These have not changed, as the lower end of the ambient concentration range was already very low in the previous studies. It is noted that this means that even quite marked changes in the size of the concentration–response function would have no effect on a guideline set on this basis. An alternative is to specify a small level of acceptable risk and use a concentration–response function (assuming it was robust) to derive a concentration that would minimize risk to this level. This approach should be considered as an option when it comes to the guideline revision stage.

Question C8

Are there important interactions among air pollutants in the induction of adverse health effects that should be considered in developing air quality policy?

Answer

Note. This answer does not consider interactions with host susceptibility behaviour or other factors, with the exception of temperature.

Some interactions among air pollutants change the toxicity of the mixture. These occur as physicochemical interactions in air, as well as biological interactions. In developing air quality policies, the following issues can be considered.

  • There is very little evidence from health studies that the mixture of air pollutants results in significantly more health effects (synergy) than would be expected based on the information for single pollutants. However, this is largely due to a lack of data and methodological limitations.
  • Very few epidemiologic studies have examined the potential of pollutants to interact. This is likely due to their moderate to high correlations. The existence of such pollutant mixtures makes it often difficult, in an uncontrolled setting, to determine either independent or synergistic effects of ambient air pollutants.
  • Synergistic biological effects between ultrafine particles and transition metals and between particles and volatile organic compounds have been shown to indicate a larger combined impact on human health than would be expected from the separate entities.
  • A reduction of emissions of nitrogen oxides without an accompanying abatement of volatile organic compounds may result in no change, or even in an increase of ozone concentrations close to the source.
  • Airborne particles of any kind can carry aeroallergens or toxic condensed vapours, such that their impact can be substantially larger than without particles. There is a trend that the smaller the particles, the stronger the adjuvant effects. Limited evidence has been published suggesting that NO2 can enhance allergic responses.
  • In general, reduction of one component will not result in a significant increase in the health risks associated with other components. The implications for reducing PM, on (semi)volatile organic compound formation, are not evident.
  • There is some evidence of interactions between pollutants and high temperature.
  • Changing the air pollution mixture due to changing fuels may, under certain conditions, lead to more harmful emissions.

Rationale

Definitions of interactions

Interactions among air pollutants can be chemical, physical and biological. A chemical interaction would mean that two or more pollutants result in new components, based on the chemical composition. A well known example is nitrogen oxides and volatile organic compounds that result in the formation of ozone and other products in the presence of sunlight. In physical interactions, solid particles act as absorbers for organic compounds, affecting their transport through air and in the respiratory tract. Biologically, interactions are distinguished by the mode of action: dose addition (similar action), effect or response addition (dissimilar action), and complex interactions (synergistic, potentiating and antagonistic). Dose addition means that the chemicals in the mixture do not affect the toxicity of one another and that each component has different effects – for example, one component with produce lung inflammation, whereas a second component causes rhinitis only. Each of the chemicals in the mixture contributes to the toxicity of the mixture in proportion to its dose. In the case of response or effect addition, the components in a mixture have the same toxicological profile – for example, all lead to inflammation in the lung. Response addition is determined by summing the responses of each toxicant in a mixture.

For interactions, compounds may interact with one another, modifying the magnitude and sometimes the nature of the toxic effect. This modification may make the composite effect stronger or weaker. An interaction might occur in the toxicokinetic phase (processes of uptake, distribution, metabolism and excretion) or in the toxicodynamic phase (effects of chemicals on the receptor, cellular target or organ). In the case of interaction, one cannot predict the toxicity based on exposure concentrations, and dose-response relationships are required to assess whether or not, for example, stronger responses occur than would be expected based on each of the pollutants alone.

Atmospheric chemistry

Reducing tailpipe soot may lead in specific cases to an increase NO2 and ultrafine particles

Reducing the emission of soot particles from motor vehicles has in some cases significantly altered the chemical composition and particle size distributions in specific urban environments (Keuken et al., 2012; Herner et al., 2011). When filter traps were applied without catalysts (urea selective catalytic reduction), NO2 concentrations increased locally at traffic sites, even though a total reduction in concentrations of nitrogen oxides was observed. The regulatory relevance of this shift is clear, although the real health impact of this change in the emission is still under discussion (Keuken et al., 2012). In these specific cases, mass-related emission of PM was significantly reduced with the change of the combustion conditions and the use of a particle filter without urea selective catalytic reduction. The use of particle traps significantly removed the larger particles and, hence, led to formation of a nucleation mode with a significantly increased particle number concentration of about 10 nm at end of the tailpipe (Herner et al., 2011), and a shift of the mode led to a particle number concentration of about 10–30 nm at some distance –for example, that of a pedestrian on a street (Harrison, Beddows & Dall’Osto, 2011; Casati et al., 2007).

Changing reactivity over time or with increasing distance from the source

Particle reactivity is believed to be an important characteristic with direct influence on the hazard potential. Particle reactivity is a general term that includes, for example, the particle intrinsic formation potential of reactive oxygen species and the ability to catalyse redox reactions. Its possible importance for health is currently being discussed (Shiraiwa, Selzle & Pöschl, 2012). Initial studies on spatial and temporal variations of particle reactivity parameters have been conducted (Boogaard et al., 2012a; Künzli et al., 2006). Still missing is the link between particle reactivity and other particle characteristics, the actual mechanisms triggered, and the ageing and/or changes of particle reactivity during atmospheric transport due to chemical reactions. Understanding how particle surface properties are altered during atmospheric transport, physically and chemically, may be one key in linking particle characteristics and emissions to health effects.

Secondary organic aerosol

Although laboratory experiments have shown that organic compounds in both gasoline fuel and diesel engine exhaust can form secondary organic aerosols, the fractional contribution from gasoline and diesel exhaust emissions to ambient secondary organic aerosols in urban environments is poorly understood. Recently, Bahreini et al. (2012) demonstrated that, in Los Angeles, the contribution from diesel emissions to secondary organic aerosol formation is very low and that gasoline emissions dominate diesel exhaust emissions in forming secondary organic aerosol mass. Chamber studies performed in Europe seem to confirm this hypothesis. In a very recent paper, Gentner et al. (2012) reported that diesel exhaust is seven times more efficient at forming aerosol than gasoline exhaust, that both sources are important for air quality and, depending on a region’s fuel use, that diesel is responsible for 65–90% of the vehicle-derived secondary organic aerosols. These conflicting results have important implications for air quality policy, but at present large uncertainties exist.

Verma et al. (2009b) have shown for Los Angeles in summer that both primary and secondary particles possess high redox activity; however, photochemical transformations of primary emissions with atmospheric ageing enhance the toxicological potency of primary particles, by generating oxidative stress and leading to subsequent cell damage.

Studies by Biswas et al. (2009) – using direct exhaust PM emissions from heavy duty vehicles, with and without emission abatement technologies implemented – suggest that the semivolatile fraction of particles are far more oxidative than solid (carbon) particles. It is also possible, in our opinion, that the secondary organic aerosols formed from the condensation of previously volatilized PM are highly oxidative.

Ozone and organics

Reactions of ozone with certain organic molecules that occur indoors at certain concentrations can produce short-lived products that are highly irritating, relative to the reaction precursors, and may also have long-term health effects. Known products of indoor ozone reactions include such compounds as formaldehyde, acetaldehyde and other organic acids. Some of these compounds are known to cause ill health in human beings (Weschler, 2006). The EPA Building Assessment Survey and Evaluation study data was analysed for associations between ambient ozone concentrations and building-related symptom prevalence (Apte, Buchanan & Mendell, 2008). Ambient ozone correlated with indoor concentrations of some aldehydes, a pattern suggesting the occurrence of indoor ozone chemistry. Apte, Buchanan & Mendell (2008) hypothesized that ozone-initiated indoor reactions play an important role in indoor air quality and building occupant health. They also hypothesized that ozone carried along into buildings from the outdoor air is involved in increasing the frequency and the range of upper and lower respiratory, mucosal, and neurological symptoms by as much as a factor of 2 when ambient ozone levels increase from those found in low-ozone regions to those typical of high-ozone regions.

Secondary inorganic aerosols

In regions with high photochemical activity, the reduction of PM mass pollution and the possible increase in frequency of droughts may lead to an increase in midday nucleation episodes with a consequent increase in levels of secondary nano-size or ultrafine particles. Thus, in highly polluted atmospheres, the secondary PM mass grows by condensation on pre-existent particles; however, in cleaner conditions, and especially under high insolation and low relative humidity, new formation on nanoparticles (nucleation) from gaseous precursors may dominate the condensation sink in urban areas (Reche et al., 2011). The nucleation starts from the oxidation of SO2 and the subsequent interaction with ammonia, and these nanoparticles immediately grow – probably by condensation of volatile organic compounds on the nucleated particles (Kulmala & Kerminen, 2008).

Another key atmospheric component is urban ammonia. This is emitted mostly by traffic and other fugitive sources, such as city waste containers and sewage, and is also emitted from animal farming. Ammonia is an alkaline gas and, when emitted in a high NO2 scenario, may enhance the formation of ammonium nitrate (a major component of PM2.5). Furthermore, the levels of ammonium nitrate may also increase due to the marked decrease of SO2 emissions that yielded a marked decrease of ammonium sulfate levels across Europe. This is expected to occur because sulfuric acid is more reactive with ammonia, and most of ammonia is consumed by sulfuric acid; when sulfuric acid decreases and more ammonia is available, more ammonium nitrate can be formed from nitric acid and ammonia.

Inorganic aerosols and metals

Thus, pure ammonium sulfate particles are rare in the atmosphere, and they usually occur as a coating on (or are coated by) other substances. They can be formed from SO2 emissions being converted photochemically into sulfuric acid. This acid coats the outside of other particles, such as metal oxide particles, which can come from the same power plant, from brake wear of cars and trucks, from metal processing, and so on. Alternatively, they may adsorb metal particles on their surface. By internal mixing, the surface components diffuse towards the core of the particle, leading to reactions between acidic sulfates and metals, converting insoluble (and hence weakly toxic) metal oxides into soluble metals. This is critical because transition metals can catalytically induce the production of highly reactive oxygenating compounds in the lung and elsewhere in the body. For example, Ghio et al. (1999) reported that soluble iron concentrations correlate with sulfate concentrations in particle filters and that the ability of soluble extracts from the particles to generate damaging oxidants is directly proportional to the sulfate concentrations. More recently, Rubasinghege et al. (2010) simulated the transformation of non-bioavailable iron to dissolved and (hence) bioavailable iron in atmospheric iron particles in the presence of acids, in both light and dark conditions. The presence of sulfuric acid on the particles results in a dramatic increase in the bioavailable iron.

Metals are not the only case where the presence of sulfates can change the toxicity of other particle components. Popovicheva et al. (2011) showed that the extent of water uptake and modification of elemental carbon particles depended on the sulfate content of the particles. Also, Li W et al. (2011) reported that sulfate aided the ageing of freshly emitted soot particles, which occurred within 200–400 m of major roads.

Wu et al. (2007) examined the effect of ammonium sulfate aerosol on the photochemical reactions of toluene (mostly from cars) and nitrogen oxides to form secondary organic particles. They found that the sulfate particles reduced the time to reach maximum concentrations of secondary organic aerosols, and also increased the total aerosol yield from toluene. That is, in the presence of sulfates, more gaseous emissions from mobile sources will be converted into particles.

Ozone and NO2

It is clear that ozone precursors make an important hemispheric (external to the EU) contribution, but because the high ozone levels are recorded mostly in rural areas, policy pressure on PM and nitrogen oxides is much higher than on ozone. The way of abating ozone levels by local measures that focus on local ozone precursors is very complex. This is because the relationship between ozone and volatile organic compounds and between ozone and NO2 is not linear, so that specific cases of reducing NO2 without compensating for the decrease of volatile organic compounds, or vice versa, may result in ineffective results, or even in an increase in ozone. On the other hand, biogenic volatile organic compounds may also be involved in the process. It is expected, however, that measures applied to reducing nitrogen oxides and industrial volatile organic compounds and also climate measures to abate methane (an ozone precursor) may contribute to abating ambient ozone. But it is also true that it is very difficult to quantify the impact of such abatement.

Interaction due to ultraviolet and/or changes in volatile organic compound and/or particulate organic carbon composition

It is well known that photochemical reactions by, for example, ultraviolet and visible radiation have a significant impact on the gaseous and particulate chemical composition of the atmosphere. One of the components currently believed to be relevant to health is organic carbon, either in the gas phase or the particulate phase. Jimenez et al. (2009) present an overview of the evolution of organic aerosols in the atmosphere. Studies closer to the source – for example, investigations of the photo-oxidation of organic compounds from motor vehicle emissions – also show significant changes (Miracolo et al., 2010). Any changes in the composition of organic carbon compounds in the atmosphere are of great importance, since their relevance to health (independent of gaseous or particle phase) has a huge spectrum, from no health effects to high toxicity. To understand the changes on a small scales – for example, near sources and close to the public – as well as the changes occurring during transport, they have to be monitored more closely.

Toxicology

Already a decade ago, Stone et al. concluded that there was evidence that synergistic interactions occur between ultrafine particles and transition metals, between particles and allergens, and between particles and volatile organic compounds, such that reductions of concentrations of one component will lead to less health effects related to the other (Stone et al., 2003). Participants at a 2007 meeting on combustion by-products (Dellinger et al., 2008) concluded that metals contained in combustion-generated airborne particles mediate the formation of toxic air pollutants, such as polychlorinated dibenzo-p-dioxins and dibenzofurans and persistent free radicals associated with oxidative stress, inflammation and other toxic effects.

PM components and ozone

Some evidence suggests that ambient concentrations of ozone can increase the biological potency of particles. Ozonized diesel exhaust particles may play a role in inducing lung responses to ambient PM (Madden et al., 2000). Similar findings have been observed in clinical studies (Bosson et al., 2008). In terms of vascular and cardiac impairment in rats inhaling ozone and diesel exhaust particles, Kodavanti et al. (2011) reported that the joint effect of exposure to ozone (0.803 mg/m3) and diesel exhaust particles (2.2 mg/m3) was less prominent than exposure to either substance alone. An explanation for this might be found in the duration of the exposure protocol (1 day a week for 16 weeks) that may have led to adaptive responses known to occur for ozone. Also, concentrations as low as 0.423 mg/m3 ozone increased the toxic response of mixtures of carbon and ammonium sulfate particles in rats – including histopathological markers of lung injury, bronchoalveolar lung fluid proteins, and measures of the function of the lung’s innate immunological defences – whereas these effects were not observed with ozone alone (Kleinman et al., 2003).

PM components and nitrogen oxides

As NO2 can cause nitrative stress (and production of nitric oxide via nitrite) and PM can cause oxidative stress (and production of superoxide radicals), the combination of exposure to both might increase production of peroxynitrite over and above either pollutant alone. Peroxynitrite is recognized as a key intermediate, with the potential to affect protein function (Gunaydin & Houk, 2009). It is formed from the reaction of nitric oxide and the superoxide radical. NO2, as is well known, can lead to increased levels of circulating nitrite and nitrate. Nitrite can be reduced back to nitric oxide and the NO2 radical in remote tissues (Lundberg, Weitzberg & Gladwin, 2008). Overall, co-exposure to PM and NO2 may lead to simple additive effects in the lung, and reduction of either one of these components may therefore lead to lower effect estimates in epidemiological studies of the other component. Although the literature on this is very sparse, it seems unlikely that the reduction of PM has a major effect on the health impacts of nitrogen oxides.

PM and other gases and/or vapours

Some papers show that the oxidative potential and toxicity of soot decreases up to 75% after heating and loosing the external organic carbon shell (Biswas et al., 2009). Also, volatile organic compounds are able to form PM (Robinson et al., 2007; Jimenez et al., 2009) In addition, studies of human beings and diesel engine exhaust and clean carbon particles (Mills et al., 2011) strongly suggest that organic chemical compounds on the surface of the carbon particles are responsible for immediate cardiovascular responses. Interestingly, reducing the amount of soot can lead to an increase in NO2, but the net effect is still reduced adverse effects on health (Lucking et al., 2011).

PM: carbon and iron

Animals exposed to soot particles at a concentration of 250 µg/m3 and to iron alone at a concentration of 57 µg/m3 had no adverse respiratory effects, but a synergistic interaction between soot and iron particles, in the combined exposure, was identified with strong inflammatory responses (Zhou et al., 2003).

Ozone and nitrogen oxides

Very little is known about the effects of co-exposures of ozone and NO2. Decades ago, Mautz et al. (1988) were able to show synergistic toxic responses in mice exposed to these pollutants. Since ozone and NO2 will form nitric acid vapour and nitrate radicals, the synergistic effects were explained by chemical interactions (Mautz et al., 1988). Various chemical compounds, as well as allergenic proteins, are efficiently oxygenated and nitrated upon exposure to ozone and NO2, which leads to an enhancement of their toxicity and allergenicity (Shiraiwa, Selzle & Pöschl, 2012). Oxidative stress has been postulated as the underlying mechanism for adverse health outcomes, suggesting a rather unifying and standard response.

Interactions with aeroallergens

The implications of co-exposures to air pollutants and aeroallergens are a rather complex issue, with both antagonistic and synergistic effects observed, depending on the sequence and levels of exposures. Eggleston (2009) concluded that “the same environmental exposures that may cause increased symptoms at one point in time may be protective when the exposure occurs earlier or at high enough levels”.

PM. Co-exposures to aeroallergens (such as grass pollen) and particles result in synergistic effects that lead to much stronger allergic responses in experimental animals and human beings than the sum of the responses to each of these constituents (D’Amato et al., 2005; Steerenberg et al., 2003; Diaz-Sanchez et al., 1999). However, this is different from studies in which the effects of particles have been studied on already allergic subjects. For example, controlled exposures, lasting 2 hours with intermittent exercise, to diluted diesel engine exhaust at a particle mass concentration of 100 µg/m3 did not evoke clear and consistent lower-airway or systemic immunological or inflammatory responses in mildly asthmatic subjects, with or without accompanying challenge with cat allergen (Riedl et al., 2012). Likewise, these diesel engine exhaust exposures did not significantly increase nonspecific or allergen-specific bronchial reactivity. A few isolated statistically significant or near-significant changes were observed during and after exposure to diesel engine exhaust, including increases in nonspecific symptoms (such as headache and nausea) suggestive of subtle, rapid-onset systemic effects. From several inhalation studies with PM2.5 in rat and mice models for allergic asthma, it was concluded that allergic inflammation and other effects can be enhanced by PM exposures for all size ranges of PM10 (Kleinman et al., 2007; Li N et al., 2010; Heidenfelder et al., 2009). However, such an interaction points more towards an enhancement of an existing disease and not to a synergy due to co-exposures.

In summary, ambient particles can, indeed, act as a carrier and adjuvant of aeroallergens, and reductions of PM may therefore result in stronger effects than those based on the concentration–response functions of PM alone, depending on the nature of the co-exposures.

NO2. Alberg et al (2011) were not able to detect an impact of NO2 on allergy induced by ovalbumin, whereas diesel exhaust particles were shown to be a potent adjuvant. In contrast, Bevelander et al. (2007) and Hodgkins et al. (2010) did find a potentiating effect of NO2 within the concentration range used by Alberg et al. (2011) (5–25 ppm) when NO2 exposure occurred immediately beforehand and ovalbumin exposure was by inhalation. Recent in vitro findings suggested that levels of SO2 and NO2 below current EU health based exposure standards can exacerbate pollen allergy on susceptible subjects (Sousa et al., 2012). Moreover, exposure to NO2 significantly enhanced lung inflammation and airway reactivity in animals that were treated with ovalbumin (Layachi et al., 2012). So, there are mixed results in human clinical studies in which NO2 exposure preceded allergen exposure.

Changing fuel composition

There is conflicting evidence about the extent to which biodiesel exhaust emissions present a lower risk to human health when compared with petroleum diesel emissions (Swanson, Madden & Ghio, 2007). German studies have shown significantly increased mutagenic effects, by a factor of 10, of particle extracts from rapeseed oil in comparison with fossil diesel fuel; and the gaseous phase caused even stronger mutagenicity (Bünger et al., 2007). Biodiesel (rapeseed oil methyl ester) has been shown to have a four times higher cytotoxicity than conventional fossil diesel under idling conditions, while no differences were observed for the transient state (Bünger et al., 2000). This was particularly evident for mixtures of rapeseed and fossil diesel, suggesting that a mixture can lead to more harmful particulate emissions. So far, the opposite was found by others: no differences for cytotoxicity with vehicle emissions under idling conditions (Jalava et al., 2010).

Results indicate an elevated mortality risk from short-term exposure to ultrafine particles, highlighting the potential importance of locally produced particles. In an epidemiological study in Erfurt, Germany, decreases in RRs for short-term associations of air pollution were calculated as pollution concentrations decreased and control measures were implemented. However, the mass concentration changes did not explain the variation in the coefficients for NO2, carbon monoxide, ultrafine particles, and ozone (Breitner et al., 2009). The control measures included restructuring of the eastern bloc industries, a changed car fleet (the number of cars with catalytic converters increased over time, as did the number of cars in general) and complete fuel replacement and an exchange from brown coal to natural gas in power plants and in domestic heating (Acker et al., 1998).

Epidemiology

For epidemiological studies, we define a statistical interaction among air pollutants as a case where the exposure to the two pollutants generates an effect that is greater than that observed for either individual pollutant. Generally speaking, epidemiological studies have not extensively examined the potential for statistical interactions among pollutants. This is likely due to the moderate to high correlation among pollutants and the existence of pollutant mixtures, making it often difficult, in a uncontrolled setting, to determine either independent or synergistic effects of ambient air pollutants.

Of the few studies to date that have been undertaken, there is very limited evidence of an interaction among pollutants, and most studies that have tested for interactions have not observed any. For example, in a cohort of 2460 subjects recruited from a pulmonary clinic, researchers analysed the association between chronic exposure to air pollution and the prevalence of ischaemic heart disease (Beckerman et al., 2012). While effects from NO2 were observed, there was no evidence of interactions between it and either ozone or PM2.5. Coogan et al. (2012) examined the association between air pollution and incidence of hypertension and diabetes mellitus in African-American women living in Los Angeles. Again, an effect of NO2 was observed, but there was no evidence of an interaction with PM2.5. In contrast, a study of infant mortality in Mexico City observed that its association with PM10 was heightened when high concentrations of ozone were also present (Carbajal-Arroyo et al., 2011). Finally, in a study of 29 European cities, Katsouyanni et al. (2001) reported that cities with higher long-term average NO2 demonstrated a greater effect of daily PM10 on mortality than did cities with low average NO2. Rather than demonstrating an interaction of pollutants, per se, this observation likely represents a greater contribution of traffic to PM10 in the cities with high levels of NO2.

Several studies have examined whether air pollutants modify or interact with the effects of temperature. Most researchers have focused on PM and ozone, since these pollutants are associated with mortality and are often correlated with higher temperatures. The results to date are mixed, as some investigators reported air pollutants modified the temperature effect, while others reported no interactions, but rather independent effects of pollution and temperature. For example, ozone and temperature were reported to interact during the 2003 heat wave in nine French cities and in 100 cities studied over several summers in the United States (Ren et al., 2008a; Filleul et al., 2006b). In addition, in a new study of nine European cities (Analitis et al., 2013) from the EuroHEAT project, there is evidence that supports interactive effects between heat waves and high ozone and PM10 concentrations. This interaction was more evident and significant in the northern cities, rather than in the Mediterranean ones. In contrast, no interaction between temperature and ozone was reported for cities in Italy and for Toronto, Canada, and only modest evidence for an interaction was observed in a multicity study in Italy (Stafoggia et al., 2008; Rainham & Smoyer-Tomic, 2003). In addition, studies in cities in the United States that examined temperature plus PM2.5, PM10, ozone and NO2 failed to find any interaction (Basu, Feng & Ostro, 2008; Zanobetti & Schwartz, 2008; Ostro, Rauch & Green, 2011).

Finally, a few epidemiological studies considered the impact of exposure to both air pollution and aeroallergens. For example, Anderson et al. (1998) examined the interactive effects of air pollutants and pollen on hospital admissions for asthma in London. While effects were observed for several air pollutants, there was no evidence that exposure to pollen exacerbated the effect of air pollution.

In summary, for policy consideration, based on the limited number of studies that have examined this issue, there is mixed evidence of potential interactions among pollutants or pollution and temperature. The one potential exception may be that related to traffic, where the different mixtures of various pollutants may have a varied impact on the magnitude of the effect on health under investigation. However, it is very difficult in these epidemiological studies to separate the independent effect of individual pollutants in the mixture and thereby determine whether an interaction exists.

Questions A7 & C9

Are there critical data gaps to be filled to help answer A, B and C questions more fully in the future?

Answer

For most air pollutants covered under the REVIHAAP project, several critical data gaps have been identified that prevent a comprehensive and thorough assessment of health hazards and concentration–response functions. More epidemiological studies that contribute to updated exposure–response functions based on meta-analyses for integrated risk assessments will result in a significant reduction in the outstanding uncertainties in risk quantification for the hazards currently identified. The coordinated application of atmospheric science, epidemiological, controlled human exposure and toxicological studies is needed to advance our understanding of the sources responsible for the most harmful emissions, physical–chemical composition of the pollution and biological mechanisms that lead to adverse effects on health. Such studies should include better characterization of the pollution mix, improved exposure assessments and better identification of susceptible groups in the general population. The correlation between many regulated air pollutants is often high, and large uncertainties exist about the effects on human health of short- and long-term exposures to non-regulated components of the air pollution mix, including some size fractions and metrics of PM. The currently regulated pollutants PM, NO2 and ozone, as well as such important particle metrics as black carbon and coarse and ultrafine particles, often have been assessed independently; this is a critical gap. Furthermore, the REVIHAAP review has clearly identified traffic as one of the major air pollution sources that affect health in Europe; however, it remains uncertain whether reducing concentrations of currently regulated pollutants will directly lead to a decrease in the health impacts of traffic-related air pollution.

Air pollution should therefore be considered to be one complex mix, and conditions under which this mix has the largest effect on human health need to be identified. In addition to (or even instead of) studies on single components or metrics, the one-atmosphere concept has been put forward as a novel way to investigate the effects on health of complex mixes. Advances in atmospheric modelling, in conjunction with validation studies that use targeted monitoring campaigns, will provide a more efficient way forward in research on health effects, rather than relying on increasing the number of components measured by routine monitoring networks.

Rationale

The information and data identified in the following text would have allowed us to answer more fully the other questions in sections A–C. Even though some of the text is pollutant-specific, we recommend a comprehensive approach to studying pollutant mixes and to conducting complementary atmospheric science, epidemiological, controlled human exposure and toxicological studies that allow assessment of all causal chains that link pollution emissions to effects on health.

General issues

The following general issues are of relevance when addressing critical data gaps.

  • The amount of literature on the adverse effects on health of air pollution is very large, making its thorough evaluation time consuming. Given the need for a systematic evaluation of various types of evidence, consideration should be given to the development or expansion of resources, to enable regular critical, systematic and quantitative evaluation of the recent literature in relation to science-policy issues of particular relevance to Europe.
  • Future studies should consider air pollution as a complex mix, and conditions need to be identified under which this mix has the largest effect on human health. A novel way to investigate the health effects of complex mixes is the one-atmosphere concept.
  • There is a clear need for more evaluation of the usefulness of two-pollutant and/or multipollutant statistical models in epidemiological studies when pollutants are highly correlated.
  • Collaboration is needed between the health and atmospheric sciences, for both complex monitoring and modelling, especially for exposure to complex pollutant mixes with strong spatial and temporal variability.
  • In coordination with health specialists, more monitoring is needed, both in a regular way and in projects. The use of supersites to perform simultaneous studies with a multipollutant approach that uses the same monitoring and health evaluation approaches across Europe is highly recommended.
  • Further research is needed on the use of health evidence for risk assessment and policy analysis. More specifically, it is needed on: (a) cause-specific PM exposure–response function shape, with a focus on non-linearities at both the high and low ends; (b) the characterization of uncertainty therein and, more broadly; (c) cost–benefit assessment and cost–effectiveness analysis, with recommendations, for example, for further work on the uncertainties in the exposure estimates from the GAINS model, which could be included in a unified estimate of uncertainty due to the major inputs for risk estimates.

PM

1. Health outcomes

This area includes: novel health outcomes; exposure– (concentration–)response functions; chemical composition and sources; distance from major roads; and the health benefits of reducing PM.

  • Novel health outcomes. The literature on the long-term effects of exposure to PM2.5 suggests additional systemic health effects beyond the respiratory and cardiovascular systems – for example, effects on the central nervous system, the progression of Alzheimer’s and Parkinson’s diseases, developmental outcomes in children, and such reproductive health outcomes as low birth weight (Questions A1 and A2). These other health effects are not yet being considered for health impact assessment, because of a lack of sufficient evidence. For this reason, more information on the underlying biological mechanisms should be generated to support a potential causal – that is, explanatory – pathway for these effects.
  • Exposure– (concentration–)response functions. Since additional PM exposure metrics (other than PM10 and PM2.5), such as ultrafine particle number concentration, black carbon or oxidative potential, have been reported, concentration–response functions need to be established for these parameters and for newly identified health outcomes (Questions A2A6). This will also require the generation of large data sets on these exposure metrics.
  • Chemical composition and sources. Although knowledge of the roles of chemical composition and emission sources of particles is accumulating, work to date gives no clear picture of which of these predict the highest hazard within the PM mixture (Question A2). Toxicological and controlled human exposure studies are expected to provide the basic understanding necessary to resolve this critical issue and to open opportunities for developing target reduction strategies. This would require confirmation by real world epidemiological studies based on sufficient chemical composition and source-related exposure data.
  • Distance from major roads. A thorough evaluation of the long-term effects of living near major roads is needed to determine which specific pollutants (including elemental carbon, organic carbon, trace metals, non-tailpipe emissions and NO2) or mixes of them may be responsible and whether the toxicity of pollutants is different near or further away from roads (Question C1). This may include reanalyses of existing epidemiological studies, looking at the relationship between living near roads and specific air pollutants. Improvements in land use regression models would allow a more detailed insight into the spatial variability in health effects associated with various sources of pollutants. This can be used for planning activities – for example, in urban areas.
  • Health benefits of reducing PM. Although very few studies have concluded that strategies for reducing PM may not lead to improved health associated with other pollutants (Question D2), investigations of the implications of interactions among components in the air pollution mix and of the influence of abatement strategies on risk estimates are mostly lacking (Question C8).

2. PM characteristics

A number of PM size fractions and components were identified as relevant to future air quality policy. In many instances however, evidence was missing on health effects, especially those associated with long-term exposures. Comparisons of the hazardousness of the components and an adjustment of their individual effects due to other components have rarely been studied, and (for some components) the spatial heterogeneity provides a challenge. Of special interest are the following pollutants.

  • Coarse particles (PM2.5-10). Several studies available to date have provided evidence for associations between short-term exposures to coarse particles and health. Data from clinical studies are scarce; and toxicological studies report that coarse particles can be equally as toxic as PM2.5 on a mass basis. Studies that assess the long-term health effects of coarse particles and studies that indicate the relative importance of the various sources of coarse particles – including road dust, desert dust, construction dust and volcanic ash, among others– are lacking (QuestionA2). Also, data that help to reduce exposure misclassification are needed, since such misclassification could obscure exposure–response relationships for coarse particles.
  • Ultrafine particles. Critical data gaps include: (a) lack of epidemiological evidence on the effect of ultrafine particles on health, with only a handful of studies published on this topic; (b) insufficient understanding of whether the effects of ultrafine particles are independent of those of PM2.5 and PM10; and (c) evidence of which ultrafine particle physical or chemical characteristics are most significant to health. There is a lack of data on the effects of short-term exposures to ultrafine particles, and there are no epidemiological studies of long-term exposure to ultrafine particles (Question A2).
  • Carbonaceous particles, including black carbon or elemental carbon, and primary and secondary organic aerosols. This gap in evidence is underscored by soot and elemental carbon having been identified as carriers of toxic (semi-)volatile compounds (Questions A2 and C8). The role of organic particles is not well understood, and data are needed on the role of the toxicity of primary or secondary organic aerosols.
  • Particles from different sources and the use of source apportionment tools in epidemiological studies. The main questions about the differences in the health effects of particles originating from different emission sources, including both natural and anthropogenic ones, is the relative contribution of these source-associated particles in comparison with the rest of the pollution mix. Since exposure to particles can also be from indoor sources or other routes of exposures (such as consumer products), the combined effects should be assessed, including possible interactions. As controls for exhaust emissions become more widespread, emissions from non-combustion sources will make up a larger proportion of vehicle emissions. Although traffic-related non-combustion PM emissions are not regulated in the same way as exhaust emissions are, they will need to be considered more closely in future assessments of the effect of motor vehicles on human health (Questions A2 and C1). Furthermore, a category of PM, which is poorly studied in the general environment, is bioaerosols, such as virus particles, bacteria, fungal spores and plant pollen. Primary biological aerosols can range in size from 10 nm (small virus particles) to 100 µm (pollen grains), and some have been associated with infectious diseases, such as Q fever.
  • Secondary inorganic aerosols. The toxicological hazard of secondary inorganic aerosols is classified as relatively low. Yet, epidemiological studies continue to report associations between sulfate or nitrate and human health. It has been suggested that associated metals or adsorbed components, such as organics, play a significant role in these observations or that they represent a mix of certain sources. Specifically, data are lacking on causal constituents or associated toxins that can explain the strong association between sulfates and adverse effects on health (Question A2). Atmospheric chemistry might be able to provide some insights about the morphology and composition of PM.

3. Exposure assessment and monitoring

Time-resolved measures of size fractions, as well as the chemical components of PM, are major gaps in exposure (including personal) studies. Specific gaps include the following.

  • Exposure monitoring. There is a need to better assess how – that is, where, how much and when – people are exposed to health relevant pollutants and, subsequently, to: (a) identify key pollutants; (b) measure and model, with appropriate instruments, the most relevant temporal and spatial resolution; (c) measure and model not only ambient concentrations, but also those in microenvironments; (d) carry out measurements of personal exposure; and (e) collect data on population time-activity patterns (Questions A3 and C10).
  • Monitoring sites. Larger and more specific studies that simultaneously cover a number of cities, regions and long study periods are needed to yield powerful results. The creation of so-called supersites or special sites should be considered, but mobile measurement units may also be employed to complement fixed site measurements in specific situations, designed in collaboration with health researchers. Additional air quality parameters and new instrumentation data – for example, from size-segregated ultrafine particles, online PM speciation measurements with aerosol mass spectrometers or other online PM speciation instruments – and surface area or bioreactivity measurements should be considered. The use of satellite-based estimates should also be increased (Question A2).
  • Modelling. The use of modelling approaches for spatio-temporal variations should be enhanced. Experimental data should also be used to validate models – for example, dispersion models or models using satellite-based or remote sensing data. More needs to be done on improving modelling and on developing low-cost and reliable methods of personal monitoring. The high spatial variability of pollutants, such as coarse PM, should be captured by advances in modelling such pollutants.
  • Internal doses, deposition patterns and distribution for various size fractions of PM. Insight into the dose delivered and the rate of delivery would facilitate the use of in vitro data for risk assessment and reduce uncertainties about extrapolation from in vivo studies to human health. Insight into the relationship between external concentration, exposure, distribution and internal dose would contribute to the evidence on the effects of long-term exposure and would provide evidence to support such novel adverse effects as those on the central nervous system (Question A2).
  • Characterization of exposure to road traffic. Improved techniques on exposure assessment related to traffic are needed; specifically ones that can help discriminate between engine exhaust and non-exhaust traffic-derived emissions (Questions A2 and C1).

Ozone

For ozone, there is a need to better understand: long-term effects; burden of disease estimates; indoor ozone; the effectiveness of abatement measures; threshold levels; and the mechanisms behind its effects on birth outcomes.

  • Long-term effects. There is a need to better understand the long-term effects (mortality effects and also morbidity effects related to the respiratory system and other systems) of ozone. There are very few studies of long-term exposure to ozone with meaningful spatial contrasts in ozone concentrations, without correlated covariates. Given that ozone is a powerful oxidant against which the body is protected to some degree by endogenous antioxidants, a better alignment of toxicological and epidemiological studies, with emphasis on long-term exposure, is needed (Question B1).
  • Burden of disease estimates. The evidence base used in the CAFE Programme to estimate the burden of disease due to the effects of ozone (such as years of life lost due to ozone mortality) should be reassessed (Questions B1B3).
  • Indoor ozone. There is a body of evidence on short-term health effects of ozone indoors due to infiltration from outdoors (Question B4). It seems that indoor reactions in which infiltrating outdoor ozone is involved produce by-products that may affect human health. More studies are needed to support this evidence.
  • Effectiveness of ozone abatement measures. Since important adverse effects on health have been observed, a need has emerged to evaluate the effectiveness of measures for ozone abatement and to increase the understanding of mechanisms that lead to ozone formation related to changes in emission patterns. Regional versus hemispheric ozone origins need more investigation. Studies on the presence or absence of a threshold for ozone effects are strongly recommended (Question B2). If there is a threshold, tackling the peaks in regional ozone is likely to be the most effective policy. If there is no threshold, reducing the hemispheric background becomes a major imperative.
  • Threshold levels. Since no clear threshold level is established for either short-term or long-term effects, the estimated total burden of disease associated with ozone depends very much on the cut-off value selected. Different cut-off values (with combinations of concentration and season) can result in large differences in burden of disease estimates.
  • Mechanisms behind ozone effects on birth outcomes. Some recent studies have reported associations between first trimester ozone and birth outcomes – in particular, preterm birth. Ozone is a potent oxidant known to cause inflammation, and this has been suggested as a possible mechanism behind its effects. Maternal vitamin D levels are important for fetal health, and vitamin D deficiency during pregnancy has been associated with circulating inflammatory proteins and pre-eclampsia. Also, inflammation in early pregnancy is associated with an increased risk of preterm delivery, and the effect of first trimester ozone on risk of preterm birth is larger among asthmatic mothers. More evidence on these novel outcomes is needed.

NO2

This section covers data gaps in relation to: toxicological and controlled human exposure studies; exposure assessment and monitoring; and epidemiological studies.

1. Toxicological and controlled human exposure studies

  • Direct effects of NO2 or NO2 as a representative substance of air pollution from road traffic. It is needed to verify whether it is plausible that NO2 exerts a direct effect on human health at current European ambient levels or whether it simply acts as a representative of other harmful components of the mix that also includes ultrafine particles and other pollutants (Questions C2 and C8). Such data are particularly relevant in areas where NO2 levels are rising while PM levels are decreasing, due to effective emission control strategies. At present, it is not known how this would affect the toxicity of the total air pollution mix and how this would change the (slope of the) concentration–response relationships for NO2.
  • Mechanisms of action. Studies are needed that not only identify biological mechanisms that lead to clinical symptoms and disease, but that also examine whether the mechanisms apply across all concentrations, or only above a certain concentration or threshold level that can be identified for NO2 as a component of a complex mix (Question C2). This data gap would include systemic nitrative stress and effects on the cardiovascular and central nervous systems, as current guidelines do not consider these types of effects, due to a lack of data from epidemiological, toxicological and controlled human exposure studies.
  • Susceptible groups. Previous studies have only considered acute exposure effects in mild asthmatics. With regard to the studies on biological mechanisms, susceptible subgroups need to be identified, to be able to protect the most vulnerable part of the general population (Question C2).

2. Exposure assessment and monitoring

  • Improved assessments of exposure to outdoor versus indoor NO2. Determining separately the effects of indoor and outdoor exposure to NO2 is a major issue. Co-pollutants that accompany NO2 indoors and outdoors are likely to be different because the sources of NO2 are different. There has always been a suspicion that co-pollutants formed alongside NO2 in indoor air may be influential to health, and hence separate evidence from experimental and epidemiological studies of outdoor NO2 exposures is required explicitly on the health effects, especially respiratory effects (Question C10).

3. Epidemiological studies

  • Direct effects of NO2. It was noted in Question C2 that there is a need to better understand whether NO2 per se has direct effects on severe and hyperreactive asthmatics. More studies are therefore needed to verify whether NO2 exerts a direct effect on human health at current ambient levels, acts as an indicator of other harmful components of the pollutant mix, and/or is a combination of these effects (Questions C2 and C8). This should be done by taking into consideration the variability of the mechanisms between population groups – for example, susceptible individuals as opposed to healthy volunteers.
  • Novel approaches. New studies could, for example, take advantage of any changes in the ratio of NO2 to primary PM metrics over time. Evaluation of recent data from epidemiological studies that use such novel approaches could allow the assessment of the relative importance of the adverse effects on health of NO2 and other constituents of the traffic-dominated mix of ambient air pollutants (Questions C1 and C4), as well as the effect of the changing air pollution mix on the risk estimates for NO2 of mortality (Question C4).

A workshop that focused specifically on research needs that relate to NO2 and its effects on health was held in London, United Kingdom, in 2011. The report of the workshop also included a number of recommendations for future research (HPA, 2011).

Metals

The following gaps have been identified for the metals included in Question C5 (arsenic, cadmium, lead and nickel).

  • Arsenic. The estimated cancer risk of inhalation of low levels of arsenic in ambient air is based on extrapolation from high-level exposure in a few occupational cohorts. While this is the standard technique, there is a need for further epidemiological evidence from cohorts at lower exposures, to assess the risk of low levels of arsenic in air.
  • Cadmium. Several studies in the last couple of years have indicated the risk to the general population of atherosclerosis and cardiovascular disease from low levels of exposure to cadmium. If these reports mirror true causal associations, such effects may be as important for public health as effects on kidney and bone, so further epidemiological and experimental studies are needed. The rationale for limiting cadmium levels in air in the air quality guidelines is to decrease deposition of cadmium on soil, to avoid oral exposure, which is predominant. The quantitative association between cadmium in air and human dietary exposure needs to be elaborated.
  • Lead. The causal chain between lead in air and adverse effects on children’s cognition and behaviour needs better data on the estimated increase of lead in blood per unit increase of lead in air at current low levels of lead in air in Europe.
  • Nickel. More epidemiological and experimental studies are needed on the possible association between nickel in ambient air and cardiovascular disease.

Of the metals not included in Question C5, the following may be important in terms of human exposure via ambient air.

  • Hexavalent chromium. Its compounds are carcinogenic. Some studies have suggested that hexavalent chromium in ambient air could contribute substantially to the risk of lung cancer in the general population. The current WHO air quality guideline value was extrapolated from occupational exposure. Further studies are needed to establish whether hexavalent chromium at ambient levels in Europe poses a cancer risk.
  • Cobalt, iron, zinc, and possibly manganese. These have the ability to form reactive oxygen species. The concentration of these elements in ambient air may contribute to PM toxicity. There is a need for more information and evaluation of this issue.
  • Manganese. Exposure to manganese via inhalation is neurotoxic. The present WHO guideline value for ambient air is derived from occupational exposure data and uncertainty factors. Additional studies have emerged in the 2000s on occupational exposure, and there is a need for an evaluation of whether these studies should affect present risk assessment and guidelines.
  • Platinum. For this, no specific guideline value was recommended by WHO in the past. Further studies were recommended, and this is still warranted.
  • Vanadium. The guideline value was based on respiratory symptoms in occupationally exposed workers, after applying an uncertainty factor. An evaluation based on more recent data about effects in the general population would be appropriate.
  • Considering the effects of metals (copper, zinc, iron and manganese) related to oxidative stress, relative to the biological potential of metals in different forms and/or states (such as platinum in catalysts and/or fuel additives; iron and antimony in brake pads), most of the evidence for toxicity related to environmental concentrations is indirect. There is a need to describe the health outcomes related to measurements of oxidative stress.
  • Atmospheric speciation studies of metals, such as hexavalent chromium (chromium (VI)) versus chromium (III), should be carried out to evaluate (more specifically) the health outcomes.

PAHs

The following data gaps have been identified for PAHs.

  • Exposure–response functions. Further research is needed to develop exposure–response functions that can be used to recommend exposure guidelines (Question C6).
  • Relative toxicities. Relative toxicities of different compounds and mixtures (such as those derived from combustion of fossil fuel and biomass) need to be assessed (Question C6).
  • Suitability of benzo[a]pyrene as a marker. Further studies to confirm the suitability of benzo[a]pyrene as a marker of the PAH mixture should be carried out. Occupational studies that are the basis of estimation of health risks do not provide this information.
  • Noncancer outcomes. Noncancer outcomes need to be assessed further, based on existing studies that indicate that a number of noncancer health outcomes (reproductive, cognitive and respiratory) might be the consequence of airborne PAH exposure (Question C6). The effect of exposure to PAHs on gene expression and methylation needs further study.

SO2

Specifically for SO2, the following issues have been raised.

  • Recent reanalysis of chamber study evidence suggests a non-significant difference between responders and non-responders at lower concentrations of SO2 as part of a trend, with more significant differences occurring at higher concentrations. The answer to Question C7 suggests that further research to confirm whether or not this difference applies at lower concentrations would have implications for the 10-minute guideline.
  • Further work is needed to identify the role of SO2 per se at current ambient concentrations in Europe in triggering acute effects on mortality and morbidity
  • Further chamber studies are needed to establish the appropriateness of the current 10-minute air quality guideline of 500 µg/m3 in protecting subjects with severe asthma.

Question C10

What is the contribution of exposure to ambient air pollution to the total exposure of air pollutants covered by the regulations, considering exposures from indoor environments, commuting and workplaces?

Answer

Tobacco smoke, where permitted indoors, dominates the exposure of the individuals to at least PM2.5 (and particle metrics black carbon and ultrafine particles), carbon monoxide, benzene, benzo[a]pyrene and naphthalene, and contributes also to exposure to NO2. Tobacco smoke exposures and risks, however, are targeted in specific policies and not in ambient air policies, and therefore the other answers below refer to conditions free of tobacco smoke.

  • In general, all exposures to air pollution of indoor and occupational origin, as well as exposure from commuting, vary between individuals much more than exposure to air pollution of ambient origin and depend strongly on the microenvironments and behaviour of the individual.
  • Specifically, commuting can increase exposure to PM, NO2, carbon monoxide and benzene, and it is a major contributor to exposure to ultrafine particles, black carbon and some metals – most importantly, iron, nickel and copper in underground rail transport systems.
  • Individual industrial workday exposure levels may be orders of magnitude higher than the average population exposure levels, but as they affect only quite specific and controlled population subgroups and are controlled by occupational (and not ambient air pollution) policies, they are not covered in this chapter.
  • Population exposure to NO2 (where gas appliances are infrequent), PM2.5, black carbon, ozone, carbon monoxide and SO2 (with more limited evidence also concerning inhaled exposures to benzo[a]pyrene, arsenic, cadmium, nickel and lead) comes dominantly from ambient air and outdoor sources.
  • Ambient air, indoor sources and commuting are all important for population exposure to NO2, and (where gas appliances are frequent) benzene and naphthalene are also important.
  • The high end of the individual exposures to PM10-2.5 and naphthalene come from indoor sources and commuting.
  • Solid-fuel-fired indoor fireplaces and stoves, where used under suboptimal conditions, dominate the high end of exposures to PM2.5, black carbon, ultrafine particles, carbon monoxide, benzene and benzo[a]pyrene of the individuals affected.

Rationale

General

On average, active adult urban populations in Europe spend an average 85–90% of their time indoors, 7–9% in traffic and only 2–5% outdoors (Hänninen et al., 2005; Schweizer et al., 2007). The most vulnerable groups, such as infants, toddlers, the elderly and the chronically ill, spend nearly 100% of their time indoors. By sheer time allocation, therefore, exposures indoors dominate total air pollution exposures. In the absence of indoor smoking and the burning of solid fuel, however, indoor exposures to most of the EU-regulated air pollutants are primarily due to outdoor sources from which the pollutants disperse via ambient air and penetrate into indoor spaces by air exchange, or are carried indoors as dust on shoes and clothes. For some pollutants, indoor and outdoor sources are of similar importance, and for yet others, the highest exposures and health risks arise from indoor sources, even though population average exposures arise mainly from ambient air.

The assessments in the current chapter refer to total exposure via inhalation to each of the contaminants through a population’s daily activities and microenvironments, regardless of the source for (or location of) the exposure. For two reasons, the so-defined total exposure may not always be the most relevant metric for risk assessment or control: (a) although PM epidemiology demonstrates remarkable consistency for very different PM source mixtures and compositions, the same PM2.5 mass originating from different sources and of different composition is hardly identical in toxicity; and (b) the role of a pollutant as an indicator of a complex mixture is likely to be different when it originates from different sources, such as NO2, indicating outdoor traffic exhaust rather than indoor gas appliances.

Tobacco smoke

Where smoking tobacco occurs indoors, it alone dominates the exposure of non-smoking individuals to PM2.5, carbon monoxide, benzene, naphthalene and benzo[a]pyrene, and contributes also to their exposure to NO2. Keeping this in mind, all of the following pollutant-by-pollutant assessments apply only to tobacco-smoke-free indoor environments.

Summary table

For each of EU-regulated air pollutant, Table 12 summarizes: (a) the most important and common non-ambient sources; (b) their significance for high end individual exposure (and respective risks) and relative population-level contributions (of); (c) indoor and (non-industrial) occupational sources; (d) commuting exposure; (e) population exposure to ambient air; (f) the proportion of population exposure influenced by ambient air regulation; and (g) the reduction in population exposure due to a given reduction in the amount of ambient air pollution.

Table 12.. Most important and common non-ambient sources of air pollutants.

Table 12.

Most important and common non-ambient sources of air pollutants.

For ozone, for example, Table 12 shows that its indoor sources are rare (such as ozonators) or weak (laser printers), and ambient air dominates the population exposure, of which almost all is, therefore, influenced by ambient air policies. However, because ozone infiltration from ambient to indoor air is generally low, a 10 µg/m3 reduction in ambient air ozone concentration reduces the average population exposure concentration by less, 2–7 µg/m3.

The table generalizes large quantities of often inconsistent data compiled from all of the articles cited. It does not necessarily apply to all individual conditions and should, consequently, be interpreted with caution.

PM2.5 and PM10-2.5

Indoor exposure to PM of ambient origin and commuting exposure (excess exposure relative to outdoor air while in transit) dominate the population exposure to PM2.5. On average, outdoor air PM2.5 is responsible for 40–70% of the total population exposure to PM2.5. Ambient air PM2.5 policies affect 60–80% of the urban population exposure, which consists of exposure to PM2.5 of ambient origin that penetrates indoors, the part of commuting exposure influenced by ambient air policies, plus exposure during the time spent in outdoor environments. An ambient air PM2.5 reduction of 10 µg/m3 reduces the average population exposure concentration by 5–8 µg/m3 – less than the 10 µg/m3 ambient reduction because, on average, only 40–70% of ambient air PM2.5 penetrates into indoor spaces where people spend an overwhelming proportion of their time. The average PM2.5 infiltration into buildings decreases steadily as new, sealed and air-conditioned buildings replace the older building stock (Chen & Zhao 2011; Hänninen et al., 2004a,b; Koistinen et al., 2004; Lai et al., 2004; Lanki et al., 2007; Johannesson et al., 2007; Fromme et al., 2008; Wichmann et al., 2010; Hänninen et al., 2011; Gariazzo et al., 2011; Oeder et al., 2012).

The mean excess PM2.5 exposure levels, while commuting, range from negligible in modern cars, buses and trams with intake air filtration, up to 20–30 µg/m3 when exposed directly to busy street air in vehicles with open windows, at bus stops, in metro stations and tunnels, or when walking or biking. The contribution of commuting to the total daily exposure therefore depends on the means, time and route. No studies representative of the population were found that would report commuting exposures in the context of total exposure and ambient air concentrations (Adams et al., 2001; Riediker et al., 2003; Seaton et al., 2005; Aarnio et al., 2005; Fondelli et al., 2008; Asmi et al., 2009; Grass et al., 2009).

On average, about half of ambient air PM10 is PM2.5. Ambient PM10 policies reduce population exposures mainly via their impacts on the fine PM fraction, but have a smaller impact on the exposure to the coarse PM fraction, of which a large proportion originates from indoor sources (Chen & Zhao, 2011 Hänninen et al., 2011; Gariazzo et al., 2011).

Ozone

Indoor ozone sources are infrequent (ozonators, old electrostatic air cleaners) or weak (laser printers). Indoors as well as outdoors, therefore, ozone of ambient origin is responsible for almost all of the population exposure, and ambient air ozone policies affect nearly all of the urban population exposures. Ozone is the most reactive of the EU-regulated air pollutants and, therefore, much of the ozone is lost in air exchange, in ventilation systems and in reactions with indoor surfaces and co-pollutants. Consequently, the average indoor air ozone levels are only 20–50% of the ambient air levels, and a given ambient air ozone reduction (µg/m3) results in a much smaller population exposure reduction.

The broader air pollution exposure impact of ozone policies may be more significant for reducing some irritating and toxic products of atmospheric ozone chemistry, as well as reactions with indoor air co-pollutants and dust on air filters in ventilation systems (Blondeau et al., 2005; Weschler, 2006; Bekö et al., 2006; Baxter et al., 2007).

NO2

Where present, indoor sources of NO2, most importantly unvented gas appliances, may significantly increase individual exposures. Indoors, NO2 is a moderately reactive gas and, consequently, the indoor concentration stays substantially below the ambient air concentration, except when emissions from gas appliances increase them to (and even above) the ambient air concentration level. Without indoor sources, the population exposure to NO2 is dominated primarily by NO2 of ambient origin and secondarily by commuting exposure. Ambient NO2 policies affect from 50% (with gas appliances) to 100% (no gas appliances) of the NO2 exposures and reduce the exposure by 50–80% (depending on region and season) of the reduction in the ambient concentration (Monn, 2001; Kousa et al., 2001; Lee et al., 2002; Lai et al., 2004; Baxter et al., 2007; Kornartit et al., 2010).

Carbon monoxide

Indoor sources and unvented, faulty and/or incorrectly operated combustion equipment are responsible for (almost) all high level exposures to carbon monoxide. Otherwise, almost all of the population exposure to carbon monoxide originates from commuting – which was much more significant in the past – and from ambient air. In Milan, Italy, in 1996–1997, the average exposure concentration of 2.1 mg/m3 was attributed to ambient air (84%), residential indoor sources (4%), occupational indoor sources (3%), excess concentrations during transport (8%), and other sources (1%) (Bruinen de Bruin et al., 2004). Current ambient air carbon monoxide policies affect almost all of the total urban population exposure and reduce the exposure by 100% of the ambient air concentration reduction.

Current urban ambient air carbon monoxide levels are an order of magnitude below the EU air quality standards. The still frequent and often lethal carbon monoxide intoxications and the indoor sources and levels of carbon monoxide that cause them, however, are not related to ambient air carbon monoxide concentrations or affected by policies. (Alm et al., 2001; Braubach et al., 2013)

SO2

Indoor sources, residential coal burning, unvented paraffin lamps and heaters still dominate some exposures of some individuals to SO2. In current European urban environments, however, indoor sources that would significantly influence the population exposure to SO2 are rare. Consequently, ambient air SO2 policies are effective in reducing the indoor air concentrations and exposures.

Benzene

Benzene-containing organic solvents, in interior materials and household chemicals, are being increasingly restricted by regulations, but still remain significant for exposures in some European regions. Airborne benzene of indoor – most importantly from attached garages – and outdoor-origin and commuting exposures are all relevant for population exposures. Ambient air benzene policies affect from 50% (with indoor sources) to 90% (no known indoor sources) of the average urban population exposure and reduce the exposure by almost all of the ambient air concentration reduction (Cocheo et al., 2000; Edwards & Jantunen; 2001; Ilacqua & Jantunen, 2003; Lai et al., 2004; Pérez-Ballesta et al., 2006; Kotzias et al., 2009; Delgado-Saborit et al., 2009; Sarigiannis et al., 2011).

PAHs: naphthalene and benzo[a]pyrene

Most of the naphthalene in air is in the vapour phase, and all benzo[a]pyrene, instead, is in the particle phase. Solid fuel combustion remains a significant indoor source of both benzo[a]pyrene and naphthalene. Naphthalene has also other indoor sources, such as naphthalene mothballs and coal tar based waterproofing. All high-end exposures are caused by indoor sources.

In the absence of indoor solid fuel combustion, all exposure to benzo[a]pyrene is of ambient origin. The contribution of ambient air to the exposure of the population to naphthalene is 30-60%. Benzo[a]pyrene infiltration follows the infiltration of PM2.5. Policy impacts should, therefore, also be similar in terms of the reduction of exposure due to the reduction of ambient air concentration. Naphthalene infiltration is similar to the infiltration of benzene, and the impact of policies should also be similar (Jantunen et al., 1999; Gustafson, Ostman & Sällsten, 2008; Ravindra, Sokhi & Van Grieken, 2008; Delgado-Saborit, 2009; Jia & Batterman, 2010; Sarigiannis et al., 2011; Yazar, Bellander & Merritt, 2011).

Arsenic, cadmium and lead

In relation to ambient air, there appear to be significant indoor sources of arsenic, cadmium and lead that contribute to indoor air, because their indoor levels often exceed the ambient air concentrations. Infiltration of these elements from ambient to indoor air is likely to follow closely the infiltration of PM2.5. Policy impacts should, therefore, also be similar in terms of the reduction of exposure and the reduction of ambient air concentration (Hänninen et al., 2004a; Lai et al., 2004; Komarnicki, 2005).

These elements enter from the outdoor to the indoor environment not only in airborne particles, but also with dust, which may be a more significant exposure pathway – for toddlers, in particular – and it warrants a different regulatory approach.

Mercury

Mercury in indoor air is mainly in the vapour phase. The greatest sources of mercury in indoor air are broken mercury-filled thermometers or barometers. The common indoor sources are broken fluorescent tubes (short peak exposures) and amalgam fillings (long-term low-level exposures). The exposure contribution of these common indoor sources to long-term exposure may be of the same order as the contribution from ambient air.

Footnotes

2

Note. It is not possible, in such experimental settings, to separate traffic-related pollutants from those derived from other sources, and this may be the case, in particular, in the field studies of Mexico City, where ozone is the most important pollutant in terms of frequency of occurrence of high levels, persistence and spatial distribution.

3

APED contains peer-reviewed ecological time-series studies published up to April 2011 − that is, the cut-off date for the last update of its literature search. A description of APED can be found in Anderson et al. (2007).

4

Chiusolo et al. (2011), Chen et al. (2012a,b), Faustini et al. (2012) and Chen et al., 2013 were published after the last update of APED in April 2011 and are not presented in Table 5. Two further papers (Koop & Tole (2004) and Roberts & Martin(2005)), which explored approaches to estimating the health effects of multiple air pollutants, are also available and are included in the information presented in Table 5.

5

Faustini et al. (2012) compared deaths from all-causes and specific causes in the general population without chronic obstructive pulmonary disease with a chronic obstructive pulmonary disease cohort in Rome, Italy. Associations for all-cause and respiratory mortality were much stronger for NO2 than for PM10, with larger estimates in the chronic obstructive pulmonary disease cohort, especially for respiratory mortality.

6

Estimate taken from CARB (2007).

7

1ppb = 1.88 µg/m3 – this has been used throughout the review of time-series studies.

8

Faustini et al. (2013) was published after the last update of APED in April 2011 and is not presented in Table 5. In two-pollutant models, with both NO2 and PM10, the associations for both pollutants for respiratory hospital admissions remained but were lower and not statistically significant.

9

Particle definitions provided by Halonen et al. (2008): nucleation mode (< 0.03 µm), Aitken mode (0.03−0.1 µm), ultrafine particles (< 0.1 µm), accumulation (0.1−0.29 µm) mode, coarse particles (PM10-PM2.5).

10

Iskandar et al. (2012) and Leitte et al. (2011) were published after the last update of APED in April 2011 and are not presented in Table 5.

11

Studies followed by asterisks (*) are older studies not referenced in earlier guidelines.

12

Literature abstracting service was provided by the Institute of Environment and Health at Cranfield until 2009; it was funded by the United Kingdom Department of Health.

13

One study was published since 2004.

14

The averaging times in this section are generally for the period of measurement (years) as set out in Tables 7 and 8.

15

* This indicates older studies not referenced in earlier guidelines.

16

The evidence is clearest for the olive tail moment in the Comet assay, but less clear for % DNA in the tail or tail length.

17

It may even be acting as a marker for nitrogen oxides, which are highest near roads before the proportion of NO2 is increased with distance as nitric oxide is oxidized to NO2. However, nitric oxide is generally regarded as less toxic than NO2.

18
19

Above a 0.1 ppm hourly average or above a 2 ppm 1-minute average within each hour (Ishigami et al., 2008); 4–7.5% of hourly average exceeding 0.1 ppm; maximum 5-minute average of 5–17 ppm (Iwasawa et al., 2009).

© World Health Organization 2013.

All rights reserved. The Regional Office for Europe of the World Health Organization welcomes requests for permission to reproduce or translate its publications, in part or in full.

Bookshelf ID: NBK361807

Views

  • PubReader
  • Print View
  • Cite this Page
  • PDF version of this title (2.6M)

Recent Activity

Your browsing activity is empty.

Activity recording is turned off.

Turn recording back on

See more...