首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
A mathematical model is used to study the fate of nitrogen oxides (NOx) emissions and the reactions responsible for the formation of nitric acid (HNO3). Model results indicate that the majority of the NOx inserted into an air parcel in the Los Angeles basin is removed by dry deposition at the ground during the first 24 h of travel, and that HNO3 is the largest single contributor to this deposition flux. A significant amount of the nitric acid is produced at night by N2O5 hydrolysis. Perturbation of the N2O5 hydrolysis rate constant within the chemical mechanism results in redistribution of the pathway by which HNO3 is formed, but does not greatly affect the total amount of HNO3 produced. Inclusion of NO3-aerosol and N2O5-aerosol reactions does not affect the system greatly at collision efficiencies, α, of 0.001, but at α = 0.1 or α = 1.0, a great deal of nitric acid could be produced by heterogeneous chemical processes.Ability to account for the observed nitrate radical (NO3) concentrations in the atmosphere provides a key test of the air quality modeling procedure. Predicted NO3 concentrations compare well with those measured by Platt et al. (Geophys. Res. Lett.7, 89–92, 1980). Analysis shows that transport, deposition and emissions, as well as chemistry, are important in explaining the behavior of NO3 in the atmosphere.  相似文献   

2.
In situ measurements of nitric acid (HNO3), reactive nitrogen (NOy), nitric oxide (NO), and ozone (O3) made in the upper troposphere (UT) and lower stratosphere (LS) between 29° and 33°N latitudes during September 1999 are used to examine NOy partitioning and correlations between the measured species in these regions. The fast-response (1 s) HNO3 measurements are acquired with a new autonomous CIMS instrument. In the LS, HNO3 accounts for the majority of NOy, and the sum of HNO3 and NOx accounts for approximately 90% of NOy. In the UT, the sum of HNO3 and NOx varies between 40% and 100% of NOy. Both HNO3 and NOy are strongly positively correlated with O3, with larger correlation slopes in the UT than in the LS. In the UT at low values of the quantity (NOy–NOx–HNO3), it is uncorrelated with O3, while at higher values, a positive correlation with O3 is found. Of these two air mass types, those with higher (NOy–NOx–HNO3) mixing ratios are likely associated with the presence of peroxyacetyl nitrate (PAN) that is produced by NOx-hydrocarbon chemistry.  相似文献   

3.
A wintertime episode during the 2000 California Regional PM Air Quality Study (CRPAQS) was simulated with the air quality model CMAQ–MADRID. Model performance was evaluated with 24-h average measurements available from CRPAQS. Modeled organic matter (OM) was dominated by emissions, which were probably significantly under-represented, especially in urban areas. In one urban area, modeled daytime nitrate concentrations were low and evening concentrations were high. This diurnal profile was not explained by the partition of nitrate between the gas and particle phases, because gaseous nitric acid concentrations were low compared to PM nitrate. Both measured and simulated nitrate concentrations aloft were lower than at the surface at two tower locations during this episode. Heterogeneous reactions involving NO3 and N2O5 accounted for significant nitrate production in the model, resulting in a nighttime peak. The sensitivity of PM nitrate to precursor emissions varied with time and space. Nitrate formation was on average sensitive to NOx emissions. However, for some periods at urban locations, reductions in NOx caused the contrary response of nitrate increases. Nitrate was only weakly sensitive to reductions in anthropogenic VOC emissions. Nitrate formation tended to be insensitive to the availability of ammonia at locations with high nitrate, although the spatial extent of the nitrate plume was reduced when ammonia was reduced. Reductions in PM emissions caused OM to decrease, but had no effect on nitrate despite the role of heterogeneous reactions. A control strategy that focuses on NOx and PM emissions would be effective on average, but reductions in VOC and NH3 emissions would also be beneficial for certain times and locations.  相似文献   

4.
Abstract

The ozone (O3) sensitivity to nitrogen oxides (NOx, or nitric oxide [NO] + nitrogen dioxide [NO2]) versus volatile organic compounds (VOCs) in the Mexico City metropolitan area (MCMA) is a current issue of scientific controversy. To shed light on this issue, we compared measurements of the indicator species O3/NOy (where NOy represents the sum of NO + NO2 + nitric acid [HNO3] + peroxyacetyl nitrate [PAN] + others), NOy, and the semiempirically derived O3/NOz surrogate (where NOz surrogate is the derived surrogate NOz, and NOz represents NOx reaction products, or NOy – NOx) with results of numerical predictions reproducing the transition regimes between NOx and VOC sensitivities. Ambient air concentrations of O3, NOx, and NOy were measured from April 14 to 25, 2004 in one downwind receptor site of photo-chemically aged air masses within Mexico City. MCMA-derived transition values for an episode day occurring during the same monitoring period were obtained through a series of photochemical simulations using the Multiscale Climate and Chemistry Model (MCCM). The comparison between the measured indicator species and the simulated spatial distribution of the indicators O3/NOy, O3/NOz surrogate, and NOy in MCMA suggest that O3 in this megacity is likely VOC-sensitive. This is in opposition to past studies that, on the basis of the observed morning VOC/NOx ratios, have concluded that O3 in Mexico City is NOx-sensitive. Simulated MCMA-derived sensitive transition values for O3/NOy, hydrogen peroxide (H2O2)/HNO3, and NOy were found to be in agreement with threshold criteria proposed for other regions in North America and Europe, although the transition crossover for O3/NOz and O3/HNO3 was not consistent with values reported elsewhere. An additional empirical evaluation of weekend/weekday differences in average maximum O3 concentrations and 6:00- to 9:00-a.m. NOx and NO levels registered at the same site in April 2004 indirectly confirmed the above results. A preliminary conclusion is that additional reductions in NOx emissions in MCMA might cause an increase in presently high O3 levels.  相似文献   

5.
Under the National Ambient Air Quality Standards (NAAQS), put in place as a result of the Clean Air Amendments of 1990, three regions in the state of Utah are in violation of the NAAQS for PM10 and PM2.5 (Salt Lake County, Ogden City, and Utah County). These regions are susceptible to strong inversions that can persist for days to weeks. This meteorology, coupled with the metropolitan nature of these regions, contributes to its violation of the NAAQS for PM during the winter. During January–February 2009, 1-hr averaged concentrations of PM10-2.5, PM2.5, NOx, NO2, NO, O3, CO, and NH3 were measured. Particulate-phase nitrate, nitrite, and sulfate and gas-phase HONO, HNO3, and SO2 were also measured on a 1-hr average basis. The results indicate that ammonium nitrate averages 40% of the total PM2.5 mass in the absence of inversions and up to 69% during strong inversions. Also, the formation of ammonium nitrate is nitric acid limited. Overall, the lower boundary layer in the Salt Lake Valley appears to be oxidant and volatile organic carbon (VOC) limited with respect to ozone formation. The most effective way to reduce ammonium nitrate secondary particle formation during the inversions period is to reduce NOx emissions. However, a decrease in NOx will increase ozone concentrations. A better definition of the complete ozone isopleths would better inform this decision.

Implications: Monitoring of air pollution constituents in Salt Lake City, UT, during periods in which PM2.5 concentrations exceeded the NAAQS, reveals that secondary aerosol formation for this region is NOx limited. Therefore, NOx emissions should be targeted in order to reduce secondary particle formation and PM2.5. Data also indicate that the highest concentrations of sulfur dioxide are associated with winds from the north-northwest, the location of several small refineries.  相似文献   


6.
This study investigates several factors that could influence ozone chemistry occurring in subsonic aircraft plumes in the upper troposphere. The study focuses on uncertainties in gas-phase rate parameters, but also examines the influence of selected heterogeneous reactions, the rate of expansion of the plume, ambient and initial plume concentrations, and the time of emissions. Monte Carlo analysis with Latin hypercube sampling was applied to an expanding box model of an aircraft plume, in order to estimate the sensitivities of O3 perturbations (ΔO3) to uncertainties in rate constants in the RADM2 chemical mechanism. The resulting coefficient of variation in ΔO3 at the end of a 36 h simulation was about 50%. Influential uncertainties in gas-phase rate parameters include those for photolysis of NO2 and HCHO, O3+NO, HO2+NO, and formation of PAN and HNO3. With high background concentrations of non-methane hydrocarbons, uncertainties in rate parameters of reactions involving peroxy radicals from ethene and propene oxidation were also influential. The coefficient of variation for ΔO3 due to uncertainties in emission indices of NOx, CO, and organic compounds was less than 15%. The effects of the heterogeneous reaction of N2O5 leading to HNO3 formation, and hypothesized reactions of HNO3 and NO2 on soot, were also investigated. The results suggest that the latter two reactions could be influential for ΔO3 if published estimates of reaction probabilities and high estimates of soot concentrations in plumes are realistic.  相似文献   

7.
Abstract

Air quality data collected in the California Regional PM10/PM2.5 Air Quality Study (CRPAQS) are analyzed to qualitatively assess the processes affecting secondary aerosol formation in the San Joaquin Valley (SJV). This region experiences some of the highest fine particulate matter (PM2.5) mass concentrations in California (≤188 μg/m3 24-hr average), and secondary aerosol components (as a group) frequently constitute over half of the fine aerosol mass in winter. The analyses are based on 15 days of high-frequency filter and canister measurements and several months of wintertime continuous gas and aerosol measurements. The phase-partitioning of nitrogen oxide (NOx)-related nitrogen species and carbonaceous species shows that concentrations of gaseous precursor species are far more abundant than measured secondary aerosol nitrate or estimated secondary organic aerosols. Comparisons of ammonia and nitric acid concentrations indicate that ammonium nitrate formation is limited by the availability of nitric acid rather than ammonia. Time-resolved aerosol nitrate data collected at the surface and on a 90-m tower suggest that both the daytime and nighttime nitric acid formation pathways are active, and entrainment of aerosol nitrate formed aloft at night may explain the spatial homogeneity of nitrate in the SJV. NOx and volatile organic compound (VOC) emissions plus background O3 levels are expected to determine NOx oxidation and nitric acid production rates, which currently control the ammonium nitrate levels in the SJV. Secondary organic aerosol formation is significant in winter, especially in the Fresno urban area. Formation of secondary organic aerosol is more likely limited by the rate of VOC oxidation than the availability of VOC precursors in winter.  相似文献   

8.
The role of the wall of a smog chamber as a radical source has been investigated in several ways. From data in the literature evidence is obtained that HNO3, present on the reactor wall, may react with NO in the gas phase according to the reaction HNO3(wall) + 2NO + H2O → 3HNO2 to give nitrous acid. Nitrous acid may subsequently photolyze to give hydroxyl radicals.Experimental evidence about the occurrence of this reaction was obtained by u.v.-irradiation of propane-NOx mixtures with and without NH3. The presence of NH3 resulted in a drastically reduced photochemical reactivity, suggesting that neutralization of nitric acid prevented the reaction of HNO3 with NO. Inclusion of the reaction mentioned above into a computer-model gave a good agreement between experimental and calculated concentration profiles of NO, NO2 and O3 in experiments with CO-NOx and propane-NOx, mixtures. The results of our findings and those from others are discussed.  相似文献   

9.
This paper reports the results of over 2 years of measurements of several of the species comprising atmospheric SOx (=SO2+SO42−) and NOy (=NO+NO2 + PAN + HNO3+NO3+ organicnitrates + HONO + 2N2O5 …) at Whiteface Mountain, New York. Continuous real-time measurements of SO2 and total gaseous NOy provided data for about 50% and 65% of the period, respectively, and 122 filter pack samples were obtained for HNO3, SO2 and aerosol SO42−, NO3, H+ and NH4+. Concentrations of SO2 and NOy were greatest in winter, whereas concentrations of the reaction products SO42− and HNO3were greatest in summer. The seasonal variation in SO42− was considerably more pronounced than that of HNO3and the high concentrations of SO42− aerosol present in summer were also relatively more acidic than SO42− aerosol in other seasons. As a result, SO42− aerosol was the predominant acidic species present in summer, HNO3was predominant in other seasons. Aerosol NO3 concentrations were low in all seasons and appeared unrelated to simultaneous NOy and HNO3concentrations. These data are consistent with seasonal variations in photochemical oxidation rates and with existing data on seasonal variations in precipitation composition. The results of this study suggest that emission reductions targeted at the summer season might be a cost-effective way to reduce deposition of S species, but would not be similarly cost-effective in reducing deposition of N species. kwAcid deposition, seasonal variation, sulfate, nitrate, nitric acid, sulfur dioxide, oxides of nitrogen, hydrogen peroxide, ozone, air pollution, Adirondack Mountains  相似文献   

10.
Increased reactive nitrogen (Nr) deposition due to expansion of agro-industry was investigated considering emission sources, atmospheric transport and chemical reactions. Measurements of the main inorganic nitrogen species (NO2, NH3, HNO3, and aerosol nitrate and ammonium) were made over a period of one year at six sites distributed across an area of ∼130,000 km2 in southeast Brazil. Oxidized species were estimated to account for ∼90% of dry deposited Nr, due to the region’s large emissions of nitrogen oxides from biomass burning and road transport. NO2-N was important closer to urban areas, however overall HNO3-N represented the largest component of dry deposited Nr. A simple mathematical modeling procedure was developed to enable estimates of total Nr dry deposition to be made from knowledge of NO2 concentrations. The technique, whose accuracy here ranged from <1% to 29%, provides a useful new tool for the mapping of reactive nitrogen deposition.  相似文献   

11.
In this study, we present ∼1 yr (October 1998–September 1999) of 12-hour mean ammonia (NH3), ammonium (NH4+), hydrochloric acid (HCl), chloride (Cl), nitrate (NO3), nitric acid (HNO3), nitrous acid (HONO), sulfate (SO42−), and sulfur dioxide (SO2) concentrations measured at an agricultural site in North Carolina's Coastal Plain region. Mean gas concentrations were 0.46, 1.21, 0.54, 5.55, and 4.15 μg m−3 for HCl, HNO3, HONO, NH3, and SO2, respectively. Mean aerosol concentrations were 1.44, 1.23, 0.08, and 3.37 μg m−3 for NH4+, NO3, Cl, and SO42−, respectively. Ammonia, NH4+, HNO3, and SO42− exhibit higher concentrations during the summer, while higher SO2 concentrations occur during winter. A meteorology-based multivariate regression model using temperature, wind speed, and wind direction explains 76% of the variation in 12-hour mean NH3 concentrations (n=601). Ammonia concentration increases exponentially with temperature, which explains the majority of variation (54%) in 12-hour mean NH3 concentrations. Dependence of NH3 concentration on wind direction suggests a local source influence. Ammonia accounts for >70% of NHx (NHx=NH3+NH4+) during all seasons. Ammonium nitrate and sulfate aerosol formation does not appear to be NH3 limited. Sulfate is primarily associated ammonium sulfate, rather than bisulfate, except during the winter when the ratio of NO3–NH4+ is ∼0.66. The annual average NO3–NH4+ ratio is ∼0.25.  相似文献   

12.
The impact of ship emissions on air quality in Alaska National Parks and Wilderness Areas was investigated using the Weather Research and Forecasting model inline coupled with chemistry (WRF/Chem). The visibility and deposition of atmospheric contaminants was analyzed for the length of the 2006 tourist season. WRF/Chem reproduced the meteorological situation well. It seems to have captured the temporal behavior of aerosol concentrations when compared with the few data available. Air quality follows certain predetermined patterns associated with local meteorological conditions and ship emissions. Ship emissions have maximum impacts in Prince William Sound where topography and decaying lows trap pollutants. Along sea-lanes and adjacent coastal areas, NOx, SO2, O3, PAN, HNO3, and PM2.5 increase up to 650 pptv, 325 pptv, 900 pptv, 18 pptv, 10 pptv, and 100 ng m?3. Some of these increases are significant (95% confidence). Enhanced particulate matter concentrations from ship emissions reduce visibility up to 30% in Prince William Sound and 5–25% along sea-lanes.  相似文献   

13.
Hourly average concentrations of up to 15 ppbv of PAN were measured during the summer 1982 a few km downwind of the chemical industries in southern Telemark, Norway, in sea breeze situations. The O3/PAN ratio was as low as 6 by volume for the highest PAN concentrations. The chemical industries are emitters of, among other gases, C12, NOx, SO2, and hydrocarbons. A model for the chemistry and dilution of the plume from the main industrial complex is described. The emission of C12 seems to be the cause of the photochemical activity. The release of atomic chlorine through the rapid photodissociation of C12 is calculated to give maximum hydroxyl concentration close downwind of the main industrial complex where also the peak concentrations of SO2 and NOx are found, giving rise to rapid nitric acid and sulphate formation. A reduction in the NOx emissions would increase the photochemical activity, while it is calculated that reducing the C12 emissions would reduce the formation of photochemical oxidants. It is shown that PAN is a much better indicator of the photochemical activity than O3.  相似文献   

14.
The formation of chemical oxidants, particularly ozone, in Mexico City were studied using a newly developed regional chemical/dynamical model (WRF-Chem). The magnitude and timing of simulated diurnal cycles of ozone (O3), carbon monoxide (CO) and nitrogen oxides (NOx), and the maximum and minimum O3 concentrations are generally consistent with surface measurements. Our analysis shows that the strong diurnal cycle in O3 is mainly attributable to photochemical variations, while diurnal cycles of CO and NOx mainly result from variations of emissions and boundary layer height. In a sensitivity study, oxidation reactions of aromatic hydrocarbons (HCs) and alkenes yield highest peak O3 production rates (20 and 18 ppbv h−1, respectively). Alkene oxidations, which are generally faster, dominate in early morning. By late morning, alkene concentrations drop, and oxidations of aromatics dominate, with lesser contributions from alkanes and CO. The sensitivity of O3 concentrations to NOx and HC emissions was assessed. Our results show that daytime O3 production is HC-limited in the Mexico City metropolitan area, so that increases in HC emissions increase O3 chemical production, while increases in NOx emissions decrease O3 concentrations. However, increases in both NOx and HC emissions yield even greater O3 increases than increases in HCs alone. Uncertainties in HC emissions estimates give large uncertainties in calculated daytime O3, while NOx emissions uncertainties are less influential. However, NOx emissions are important in controlling O3 at night.  相似文献   

15.
The threshold values of indicator species and ratios delineating the transition between NOx and VOC sensitivity of ozone formation are assumed to be universal by various investigators. However, our previous studies suggested that threshold values might vary according to the locations and conditions. In this study, threshold values derived from various model simulations at two different locations (the area of Switzerland by UAM Model and San Joaquin Valley of Central California by SAQM Model) are examined using a new approach for defining NOx and VOC sensitive regimes. Possible definitions for the distinction of NOx and VOC sensitive ozone production regimes are given. The dependence of the threshold values for indicators and indicator ratios such as NOy, O3/NOz, HCHO/NOy, and H2O2/HNO3 on the definition of NOx and VOC sensitivity is discussed. Then the variations of threshold values under low emission conditions and in two different days are examined in both areas to check whether the models respond consistently to changes in environmental conditions. In both cases, threshold values are shifted similarly when emissions are reduced. Changes in the wind fields and aging of the photochemical oxidants seem to cause the day-to-day variation of the threshold values. O3/NOz and HCHO/NOy indicators are predicted to be unsatisfactory to separate the NOx and VOC sensitive regimes. Although NOy and H2O2/HNO3 provide a good separation of the two regimes, threshold values are affected by changes in the environmental conditions studied in this work.  相似文献   

16.
Abstract

Ambient air quality data were analyzed to empirically evaluate the effects of reductions of volatile organic compounds (VOCs) and oxides of nitrogen (NOx) emissions on weekday and weekend levels of ozone (O3; 1991–1998) and particulate NO3 - (1980–1999) in southern California. Despite significantly lower O3 precursor levels on weekends, 20 of 28 South Coast Air Basin (SoCAB) sites (28 of all 78 southern California sites) showed statistically significant higher mean O3 levels on Sundays than on weekdays (p < 0.01); 49 of the remaining 50 sites showed no significant differences between mean weekday and Sunday peak O3 levels. We also observed no statistically significant differences between mean weekday and weekend concentrations of particulate NO3 - or nitric acid (HNO3, the precursor of particulate NO3 -). Averaged over sites, the mean Sunday NOx and nonmethane hydrocarbon concentrations were 25–41% and 16–30% lower, respectively, than on weekdays. Site-to-site differences between weekend and weekday mean peak hourly O3 levels were related to whether O3 formation was limited by the availability of NOx. A thermodynamic equilibrium model predicts that particulate NO3 - levels would decrease in response to a reduction of HNO3, and that particulate ammonium NO3 - formation was not limited by the availability of ammonia. The similarity of mean weekday and weekend levels of NO3 - therefore did not result from limitations on the formation of particulate NO3 - from its precursor, HNO3.  相似文献   

17.
Radical chemistry in the nocturnal urban boundary layer is dominated by the nitrate radical, NO3, which oxidizes hydrocarbons and, through the aerosol uptake of N2O5, indirectly influences the nitrogen budget. The impact of NO3 chemistry on polluted atmospheres and urban air quality is, however, not well understood, due to a lack of observations and the strong impact of vertical stability of the boundary layer, which makes nocturnal chemistry highly altitude dependent.Here we present long-path DOAS observations of the vertical distribution of the key nocturnal species O3, NO2, and NO3 during the TRAMP experiment in Summer 2006 in Houston, TX. Our observations confirm the altitude dependence of nocturnal chemistry, which is reflected in the concentration profiles of all trace gases at night. In contrast to other study locations, NO3 chemistry in Houston is dominated by industrial emissions of alkenes, in particular of isoprene, isobutene, and sporadically 1,3-butadiene, which are responsible for more than 70% of the nocturnal NO3 loss. The nocturnally averaged loss of NOx in the lowest 300 m of the Houston atmosphere is ~0.9 ppb h?1, with little day-to-day variability. A comparison with the daytime NOx loss shows that NO3 chemistry is responsible for 16–50% of the NOx loss in a 24-h period in the lowest 300 m of the atmosphere. The importance of the NO3 + isoprene/1,3-butadiene reactions implies the efficient formation of organic nitrates and secondary organic aerosol at night in Houston.  相似文献   

18.
ABSTRACT

A thermodynamic equilibrium model was used to investigate the response of aerosol NO3 to changes in concentrations of HNO3, NH3, and H2SO4. Over a range of temperatures and relative humidities (RHs), two parameters provided sufficient information for indicating the qualitative response of aerosol NO3. The first was the excess of aerosol NH4 + plus gas-phase NH3 over the sum of HNO3, particulate NO3, and particulate SO4 2- concentrations. The second was the ratio of particulate to total NO3 concentrations. Computation of these quantities from ambient measurements provides a means to rapidly analyze large numbers of samples and identify cases in which inorganic aerosol NO3 formation is limited by the availability of NH3. Example calculations are presented using data from three field studies. The predictions of the indicator variables and the equilibrium model are compared.  相似文献   

19.
The regional-scale transport, chemistry and deposition of acidifying compounds, photochemical oxidants, and their precursors are analyzed using a second-generation Eulerian model. The important atmospheric processes are incorporated using chemical, dynamical and thermodynamical parameterizations having sufficient detail to accommodate boundary layer-free troposphere exchange in cloudy and cloud-free environments, and in-cloud and below-cloud wet removal and chemistry. Forty-one species are considered, many of which are also present in the liquid-drop phases. In the regional scale transport, the advected species are NO, NO2, SO2, SO−24, O3, HNO3, NH3, PAN, H2O2, HCHO, alkanes, C2H4, other olefins, aromatics, RCHO, ROOH, HNO2, RONO2 and RO2NO2. The model capabilities are illustrated by showing simulations in which non-precipitating clouds are present to absorb gas-phase species, chemically alter these, and then release them to the atmosphere.  相似文献   

20.
We evaluated the effect of a 20% reduction in the rate constant of the reaction of the hydroxyl radical with nitrogen dioxide to produce nitric acid (OH+NO2→HNO3) on model predictions of ozone mixing ratios ([O3]) and the effectiveness of reductions in emissions of volatile organic compounds (VOC) and nitrogen oxides (NOx) for reducing [O3]. By comparing a model simulation with the new rate constant to a base case scenario, we found that the [O3] increase was between 2 and 6% for typical rural conditions and between 6 and 16% for typical urban conditions. The increases in [O3] were less than proportional to the reduction in the OH+NO2 rate constant because of negative feedbacks in the photochemical mechanism. Next, we used two different approaches to evaluate how the new OH+NO2 rate constant changed the effectiveness of reductions in emissions of VOC and NOx: first, we evaluated the effect on [O3] sensitivity to small changes in emissions of VOC (d[O3]/dEVOC) and NOx (d[O3]/dENOx); and secondly, we used the empirical kinetic modeling approach to evaluate the effect on the level of emissions reduction necessary to reduce [O3] to a specified level. Both methods showed that reducing the OH+NO2 rate constant caused control strategies for VOC to become less effective relative to NOx control strategies. We found, however, that d[O3]/dEVOC and d[O3]/dENOx did not quantitatively predict the magnitude of the change in the control strategy because the [O3] response was nonlinear with respect to the size of the emissions reduction. We conclude that model sensitivity analyses calculated using small emissions changes do not accurately characterize the effect of uncertainty in model inputs (in this case, the OH+NO2 rate constant) on O3 attainment strategies. Instead, the effects of changes in model inputs should be studied using large changes in precursor emissions to approximate realistic attainment scenarios.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号