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 37 毫秒
1.
Chemical actinometry was used to measure nitrate photolysis rate coefficients, JNO3, on and in snowpack at Summit, Greenland. Sealed glass tubes containing nitrate and a hydroxyl radical trapping system were buried in snow and exposed for between 2 and 24 h. Average JNO3 values for 2-h midday exposures in early June on surface snow were 10–14×10−7 s−1. Averages over 24 h were 3.5–4.5×10−7 s−1. These values reflect the integrated photon flux and also any variation of the nitrate photolysis rate with temperature. Attenuation of JNO3 within the firn was 0.03–0.04 cm−1 for 24-h exposures and 0.08 cm−1 for a 2-h exposure. Different attenuation coefficients may relate to differential light penetration due to changes in sun angle over the course of 24 h.  相似文献   

2.
Sensitivity of ozone (O3) concentrations in the Mexico City area to diurnal variations of surface air pollutant emissions is investigated using the WRF/Chem model. Our analysis shows that diurnal variations of nitrogen oxides (NOx = NO + NO2) and volatile organic compound (VOC) emissions play an important role in controlling the O3 concentrations in the Mexico City area. The contributions of NOx and VOC emissions to daytime O3 concentrations are very sensitive to the morning emissions of NOx and VOCs. Increase in morning NOx emissions leads to decrease in daytime O3 concentrations as well as the afternoon O3 maximum, while increase in morning VOC emissions tends to increase in O3 concentrations in late morning and early afternoon, indicating that O3 production in Mexico City is under VOC-limited regime. It is also found that the nighttime O3 is independent of VOCs, but is sensitive to NOx. The emissions of VOCs during other periods (early morning, evening, and night) have only small impacts on O3 concentrations, while the emissions of NOx have important impacts on O3 concentrations in the evening and the early morning.This study suggests that shifting emission pattern, while keeping the total emissions unchanged, has important impacts on air quality. For example, delaying the morning emission peak from 8 am to 10 am significantly reduced the morning peaks of NOx and VOCs, as well as the afternoon O3 maxima. It suggests that without reduction of total emission, the daytime O3 concentrations can be significantly reduced by changing the diurnal variations of the emissions of O3 precursors.  相似文献   

3.
Studies on the effect of NOx on zinc corrosion are scarce and their results are variable and at times seemingly contradictory. This paper reports laboratory tests involving the dry deposition on zinc surfaces of 800 μg m−3 NO2, alone and in combination with 800 μg m−3 SO2, at temperatures of 35 and 25 °C and relative humidities of 90% and 70%. From the gravimetric results obtained and from the characterisation of the corrosion products by optical microscopy, scanning electron microscopy (SEM/EDX), grazing incidence X-ray diffraction (GIXD) and X-ray photoelectron spectroscopy (XPS), it has been verified that the corrosive action of NO2 alone is negligible compared with SO2. However, an accelerating effect has been observed when NO2 acts in conjunction with SO2 at 25 °C and 90% RH. At 35 °C and 90% RH, the accelerating effect is smaller, and at low relative humidities (70%), the synergistic effect is only slight, which suggests it may be favoured by the presence of moisture. In those cases where an accelerating effect has been observed, a greater proportion of sulphate ions has been found in the corrosion products, and nitrogen compounds have not been detected, indicating that NO2 participates indirectly as a catalyst of the oxidation of SO2 to sulphate.  相似文献   

4.
Boundary layer concentrations of hydroxyl (OH) and hydroperoxyl (HO2) radicals were measured at 1180 m elevation in a mountainous, forested region of north-western Greece during the AEROsols formation from BIogenic organic Carbon (AEROBIC) field campaign held in July–August 1997. In situ measurements of OH radicals were made by laser-induced fluorescence (LIF) at low pressure, exciting in the (0, 0) band of the A–X system at 308 nm. HO2 radicals were monitored by chemical titration to OH upon the addition of NO, with subsequent detection by LIF. The instrument was calibrated regularly during the field campaign, and demonstrated a sensitivity towards OH and HO2 of 5.2×105 and 2.4×106 molecule cm−3, respectively, for a signal integration period of 2.5 min and a signal-to-noise ratio of 1. Diurnal cycles of OH and HO2 were measured on 10 days within a small clearing of a forest of Greek Fir (Abies Borisi-Regis). In total 4165 OH data points and 1501 HO2 data points were collected at 30 s intervals. Noon-time OH and HO2 concentrations were between 4–12×106 and 0.4–9×108 molecule cm−3, respectively. The performance of the instrument is evaluated, and the data are interpreted in terms of correlations with controlling variables. A significant correlation (r2=0.66) is observed between the OH concentration and the rate of photolysis of ozone, J(O1D). However, OH persisted into the early evening when J(O1D) had fallen to very low values, consistent with the modelling study presented in the following paper (Carslaw et al., 2001, OH and HO2 radical chemistry in a forest region of north-western Greece. Atmospheric Environment 35, 4725–4737) that predicts a significant radical source from the ozonolysis of biogenic alkenes. Normalisation of the OH concentrations for variations in J(O1D) revealed a bell-shaped dependence of OH upon NOx (NO+NO2), which peaked at [NOx] ∼1.75 ppbv. The diurnal variation of HO2 was found to be less correlated with J(O1D) compared to OH.  相似文献   

5.
In the United States, fertilized corn fields, which make up approximately 5% of the total land area, account for approximately 45% of total soil NOx emissions. Leaf chamber measurements were conducted of NO and NO2 fluxes between individual corn leaves and the atmosphere in (1) field-grown plants near Champaign, IL (USA) in order to assess the potential role of corn canopies in mitigating soil–NOx emissions to the atmosphere, and (2) greenhouse-grown plants in order to study the influence of various environmental variables and physiological factors on the dynamics of NO2 flux. In field-grown plants, fluxes of NO were small and inconsistent from plant to plant. At ambient NO concentrations between 0.1 and 0.3 ppbv, average fluxes were zero. At ambient NO concentrations above 1 ppbv, NO uptake occurred, but fluxes were so small (14.3±0.0 pmol m−2 s−1) as to be insignificant in the NOx inventory for this site. In field-grown plants, NO2 was emitted to the atmosphere at ambient NO2 concentrations below 0.9 ppbv (the NO2 compensation point), with the highest rate of emission being 50 pmol m−2 s−1 at 0.2 ppbv. NO2 was assimilated by corn leaves at ambient NO2 concentrations above 0.9 ppbv, with the maximum observed uptake rate being 643 pmol m−2 s−1 at 6 ppbv. When fluxes above 0.9 ppbv are standardized for ambient NO2 concentration, the resultant deposition velocity was 1.2±0.1 mm s−1. When scaled to the entire corn canopy, NO2 uptake rates can be estimated to be as much as 27% of the soil-emitted NOx. In greenhouse-grown and field-grown leaves, NO2 deposition velocity was dependent on incident photosynthetic photon flux density (PPFD; 400–700 nm), whether measured above or below the NO2 compensation point. The shape of the PPFD dependence, and its response to ambient humidity in an experiment with greenhouse-grown plants, led to the conclusion that stomatal conductance is a primary determinant of the PPFD response. However, in field-grown leaves, measured NO2 deposition velocities were always lower than those predicted by a model solely dependent on stomatal conductance. It is concluded that NO2 uptake rate is highest when N availability is highest, not when the leaf deficit for N is highest. It is also concluded that the primary limitations to leaf-level NO2 uptake concern both stomatal and mesophyll components.  相似文献   

6.
The photooxidation of methylhydroperoxide (MHP) and ethylhydroperoxide (EHP) was studied in the aqueous phase under simulated cloud droplet conditions. The kinetics and the reaction products of direct photolysis and OH-oxidation were studied for both compounds. The photolysis frequencies obtained were JMHP=4.5 (±1.0)×10−5 s−1 and JEHP=3.8 (±1.0)×10−5 s−1 for MHP and EHP respectively at 6 °C. The rate constants of OH-oxidation of MHP at 6 °C were 6.3 (±2.6)×108 M−1 s−1 and 5.8 (±1.9)×108 M−1 s−1 relative to ethanol and 2-propanol respectively, and the rate constant of OH-oxidation of EHP was 2.1 (±0.6)×109 M−1 s−1 relative to 2-propanol at 6 °C. The reaction products obtained were not only the corresponding aldehydes, but also the corresponding acids, and hydroxyhydroperoxides as primary reaction products. The yields for these products were sensitive to the pH value. The carbon balance was higher than 85% for all experiments, showing that most reaction products were detected. A chemical mechanism was proposed for each reaction, and the atmospheric implications were discussed.  相似文献   

7.
The relative rate method has been used to determine the rate constants for the gas-phase reactions of NO3 radicals with a series of acrylate esters: ethyl acrylate (k1), n-butyl acrylate (k2), methyl methacrylate (k3) and ethyl methacrylate (k4) at 298 ± 1 K and 760 Torr. The obtained rate constants are k1 = (1.8 ± 0.25) × 10?16 cm3 molecule?1 s?1, k2 = (2.1 ± 0.33) × 10?16 cm3 molecule?1 s?1, k3 = (3.6 ± 1.2) × 10?15 cm3 molecule?1 s?1, k4 = (4.9 ± 1.7) × 10?15 cm3 molecule?1 s?1. The experimental rate constants are in good agreement with theoretical rate constants calculated by an algorithm of the correlation between the rate constants and the orbital energies for the reactions of unsaturated VOCs with NO3 radicals. In addition, the atmospheric lifetimes of the compound against NO3 attack are estimated and the results show that NO3 reactions contribute little to the atmospheric losses of acrylate esters except in polluted regions.  相似文献   

8.
Concurrent tropospheric O3 and CO vertical profiles from the Tropospheric Emission Spectrometer (TES) during the MILAGRO/INTEX-B aircraft campaigns over the Mexico City Metropolitan Area (MCMA) and its surrounding regions were used to examine Mexico City pollution outflow on a regional scale. The pollution outflow from the MCMA occurred predominantly at 600–800 hPa as evident in O3, CO, and NOx enhancements in the in situ aircraft observations. TES O3 and CO are sensitive to the MCMA pollution outflow due to their relatively high sensitivities at 600–800 hPa. We examined O3, CO, and their correlation at 600–800 hPa from TES retrievals, aircraft measurements, and GEOS-Chem model results. TES captures much of the spatial and day-to-day variability of O3 seen in the in situ data. TES CO, however, shows much less spatial and day-to-day variability compared with the in situ observations. The ΔO3/ΔCO slope is significantly higher in the TES data (0.43) than the in situ data (0.28) due partly to the lack of variability in TES CO. Extraordinarily high ΔO3/ΔCO slope (0.81) from TES observations at 618 hPa over the Eastern U.S. was previously reported by Zhang et al. [Zhang, L., Jacob, D.J., Bowman, K.W., et al., 2006. Ozone–CO correlations determined by the TES satellite instrument in continental outflow regions. Geophys. Res. Lett. 33, L18804. 10.1029/2006GL026399.]. Thus the application of TES CO–O3 correlation to map continental pollution outflow needs further examination.  相似文献   

9.
Personal exposures, residential indoor, outdoor and workplace levels of nitrogen dioxide (NO2) were measured for 262 urban adult (25–55 years) participants in three EXPOLIS centres (Basel; Switzerland, Helsinki; Finland, and Prague; Czech Republic) using passive samplers for 48-h sampling periods during 1996–1997. The average residential outdoor and indoor NO2 levels were lowest in Helsinki (24±12 and 18±11 μg m−3, respectively), highest in Prague (61±20 and 43±23 μg m−3), with Basel in between (36±13 and 27±13 μg m−3). Average workplace NO2 levels, however, were highest in Basel (36±24 μg m−3), lowest in Helsinki (27±15 μg m−3), with Prague in between (30±18 μg m−3). A time-weighted microenvironmental exposure model explained 74% of the personal NO2 exposure variation in all centres and in average 88% of the exposures. Log-linear regression models, using residential outdoor measurements (fixed site monitoring) combined with residential and work characteristics (i.e. work location, using gas appliances and keeping windows open), explained 48% (37%) of the personal NO2 exposure variation. Regression models based on ambient fixed site concentrations alone explained only 11–19% of personal NO2 exposure variation. Thus, ambient fixed site monitoring alone was a poor predictor for personal NO2 exposure variation, but adding personal questionnaire information can significantly improve the predicting power.  相似文献   

10.
Comparisons were made between the predictions of six photochemical air quality simulation models (PAQSMs) and three indicators of ozone response to emission reductions: the ratios of O3/NOz and O3/NOy and the extent of reaction. The values of the two indicator ratios and the extent of reaction were computed from the model-predicted mixing ratios of ozone and oxidized nitrogen species and were compared to the changes in peak 1 and 8 h ozone mixing ratios predicted by the PAQSMs. The ozone changes were determined from the ozone levels predicted for base-case emission levels and for reduced emissions of volatile organic compounds (VOCs) and oxides of nitrogen (NOx). For all simulations, the model-predicted responses of peak 1 and 8 h ozone mixing ratios to VOC or NOx emission reductions were correlated with the base-case extent of reaction and ratios of O3/NOz and O3/NOy. Peak ozone values increased following NOx control in 95% (median over all simulations) of the high-ozone (>80 ppbv hourly mixing ratio in the base-case) grid cells having mean afternoon O3/NOz ratios less than 5 : 1, O3/NOy less than 4 : 1, or extent less than 0.6. Peak ozone levels decreased in response to NOx reductions in 95% (median over all simulations) of the grid cells having peak hourly ozone mixing ratios greater than 80 ppbv and where mean afternoon O3/NOz exceeded 10 : 1, O3/NOy was greater than 8 : 1, or extent exceeded 0.8. Ozone responses varied in grid cells where O3/NOz was between 5 : 1 and 10 : 1, O3/NOy was between 4 : 1 and 8 : 1, or extent was between 0.6 and 0.8. The responses in such grid cells were affected by ozone responses in upwind grid cells and by the changes in ozone levels along the upwind boundaries of the modeling domains.  相似文献   

11.
We present two years (January 2007–December 2008) of atmospheric SO2, NO2 and NH3 measurements from ten background or rural sites in nine provinces in China. The measurements were made on a monthly basis using passive samplers under careful quality control. The results show large geographical and seasonal variations in the concentrations of these gases. The mean SO2 concentration varied from 0.7 ± 0.4 ppb at Waliguan on Qinghai Plateau to 67.3 ± 31.1 ppb at Kaili in Guizhou province. The mean NO2 concentration ranged from 0.6 ± 0.4 ppb at Waliguan to 23.9 ± 6.9 ppb at Houma in southern Shanxi. The mean NH3 concentration ranged from 2.8 ± 3.0 ppb at Shangdianzi in northeastern Beijing to 13.7 ± 8.4 ppb at Houma. At most sites, SO2 and NO2 peaked in winter and reached minima in summer, while NH3 showed maximum values in summer and lower values in cold seasons. On the whole, the geographical distributions of the observed gas concentrations are consistent with those of emissions. The ground measurements of SO2 and NO2 are contrasted to the SCIAMACHY SO2 and OMI NO2 tropospheric columns, respectively. Although the satellite data can capture the main features of emissions and concentrations of SO2, they do not reflect the variations of SO2 in the surface layer. The situation is better for the case of NO2. The OMI NO2 columns capture the geographical differences in the ground NO2 and correlate fairly well with the ground levels of NO2 at six of the ten sites.  相似文献   

12.
Measurements of atmospheric electricity began at Kew Observatory, London (51°28′N, 0°19′W) in 1843, with recording apparatus installed by Lord Kelvin in 1861. The measured electric potential gradient (PG) at Kew has always been influenced by smoke pollution, causing a December PG maximum and July minimum. Theory links PG variations with aerosol concentrations, and the 20th century smoke measurements made at Kew permit smoke concentrations to be retrieved from 19th century PG data. Absolute calibration of the 1862–1864 PG is achieved by considering changes in the global electric circuit, for which the geomagnetic aa-index is used as a proxy. The mean annual PG in 1863 is estimated as 363±29 V m−1, and the mean smoke concentration found is 0.17±0.05 mg m−3. Diurnal variations in smoke pollution differ between the seasons, and change in their character after the advent of motor traffic.  相似文献   

13.
14.
A method is developed to estimate wet deposition of nitrogen in a 11×14 km (0.125°Lon.×0.125°Lat.) grid scale using the precipitation chemistry monitored data at 10 sites scattered over South Korea supplemented by the routinely available precipitation rate data at 65 sites and the estimated emissions of NO2 and NH3 at each precipitation monitoring site. This approach takes into account the contributions of local NO2 and NH3 emissions and precipitation rates on wet deposition of nitrogen. Wet deposition of nitrogen estimated by optimum regression equations for NO3 and NH4+ derived from annual total monitored wet deposition and that of emissions of NO2 and NH3 is incorporated to normalize wet deposition of nitrogen at each precipitation rate class, which is divided into 6 classes. The optimum regression equations for the estimation of wet deposition of nitrogen at precipitation monitoring sites are developed using the normalized wet deposition of nitrogen and the precipitation rate at 10 precipitation chemistry monitoring sites. The estimated average annual total wet depositions of NO3 and NH4+ are found to be 260 and 500 eq ha−1 yr−1 with the maximum values of 400 and 930 eq ha−1 yr−1, respectively. The annual mean total wet deposition of nitrogen is found to be about 760 eq ha−1 yr−1, of which more than 65% is contributed by wet deposition of ammonium while, the emission of NH3 is about half of that of NO2, suggesting the importance of NH3 emission for wet deposition of nitrogen in South Korea.  相似文献   

15.
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   

16.
A fast response analyzer for HNO3 in highly polluted air is described. The time resolution attainable was 12 s. The method is based on the difference in a technique for HNO3-scrubbed and non-scrubbed air and the reduction of HNO3 to NO with the use of a line of catalytic converters and a method for the subsequent NO-ozone chemiluminescence. A sample air stream, in which particulates are removed with a Teflon filter, is divided into two channels. CH-1 is directly connected to the converter line, and CH-2 contains a HNO3 scrubber packed with a nylon fiber that goes to another converter line. Each converter line is composed of a hot quartz-bead converter (QBC) and a molybdenum converter (MC) in a series. A QBC reduces HNO3 to (NO+NO2), which is called NOx. The MC reduces the NOx to NO.For CH-1, the analyzer detects most compounds that typically comprise NOy (J. Geophys. Res. 91 (1986) 9781). These CH-1 compounds are called NOy′ hereafter (NOy-particulate nitrate) because the particulates are removed by the filter. A difference in the detector signal for the two channels indicates HNO3. For a blank test, atmospheric air in which HNO3 was pre-scrubbed by an extra nylon fiber was introduced to the analyzer. Variations in the blank value were 0.38±0.42 and 0.34±0.55 ppb during the high readings (NOy′-HNO3 ) (called NOy* hereafter) (111±12 ppb, N=180), and low NOy* readings (62±8 ppb, N=180), respectively, indicating that the lowest detection limit of the analyzer is 1.1 ppb (2σ). When the data obtained with the analyzer is compared to the data using the denuder method, a linear correlation with the regression of Y=0.973X+0.077 (r2=0.916 (N=20)) in the range of 0–6.5 ppb HNO3 is obtained, which is an excellent agreement. Atmospheric monitoring was carried out at Kobe. Although the average concentration of HNO3 was 2.6±1.3 ppb, ca.10 ppb for a HNO3 concentration was occasionally observed when the NOy* concentration was high, i.e., more than 100 ppb.  相似文献   

17.
The effect of HNO3 on the atmospheric corrosion of copper has been investigated at varied temperature (15–35 °C) and relative humidity (0–85% RH). Fourier transform infrared (FT-IR) spectroscopy and X-ray diffraction (XRD) confirmed the existence of cuprite and gerhardtite as the two main corrosion products on the exposed copper surface. For determination of the corrosion rate and for estimation of the deposition velocity (Vd) of HNO3 on copper, gravimetry and ion chromatography has been employed. Temperature had a low effect on the corrosion of copper. A minor decrease in the mass gain was observed as the temperature was increased to 35 °C, possibly as an effect of lower amount of cuprite due to a thinner adlayer on the metal surface at 35 °C. The Vd of HNO3 on copper, however, was unaffected by temperature. The corrosion rate and Vd of HNO3 on copper was the lowest at 0% RH, i. e. dry condition, and increased considerably when changing to 40% RH. A maximum was reached at 65% RH and the mass gain remained constant when the RH was increased to 85% RH. The Vd of HNO3 on copper at ⩾65% RH, 25 °C and 0.03 cm s−1 air velocity was as high as 0.15±0.03 cm s−1 to be compared with the value obtained for an ideal absorbent, 0.19±0.02 cm s−1. At sub-ppm levels of HNO3, the corrosion rate of copper decreased after 14 d and the growth of the oxide levelled off after 7 d of exposure.  相似文献   

18.
The night-time tropospheric chemistry of two stress-induced volatile organic compounds (VOCs), (Z)-pent-2-en-1-ol and pent-1-en-3-ol, has been studied at room temperature. Rate coefficients for reactions of the nitrate radical (NO3) with these pentenols were measured using the discharge-flow technique. Because of the relatively low volatility of these compounds, we employed off-axis continuous-wave cavity-enhanced absorption spectroscopy for detection of NO3 in order to be able to work in pseudo first-order conditions with the pentenols in large excess over NO3. The rate coefficients were determined to be (1.53±0.23)×10−13 and (1.39±0.19)×10−14 cm3 molecule−1 s−1 for reactions of NO3 with (Z)-pent-2-en-1-ol and pent-1-en-3-ol. An attempt to study the kinetics of these reactions with a relative-rate technique, using N2O5 as source of NO3 resulted in significantly higher apparent rate coefficients. Performing relative-rate experiments in known excesses of NO2 allowed us to determine the rate coefficients for the N2O5 reactions to be (5.0±2.8)×10−19 cm3 molecule−1 s−1 for (Z)-pent-2-en-1-ol, and (9.1±5.8)×10−19 cm3 molecule−1 s−1 for pent-1-en-3-ol. We show that these relatively slow reactions can indeed interfere with rate determinations in conventional relative-rate experiments.  相似文献   

19.
The interaction of NO2 on carbonaceous aerosol particles in an NO2 concentration range relevant for the troposphere was studied. The adsorption as a function of NO2 concentration (2.5–65 ppb) was investigated along with the dependence on time (1–600 s) and particle concentration. The results exhibit a zero-order process in NO2 for the chemisorption over the measured time and concentration range. The results suggest that the chemisorption reaction is limited by a rapidly established steady-state coverage of a precursor in the form of reversibly adsorbed NO2 which seems to be constant over the whole investigated NO2 concentration range. Within the first 20 s, a chemisorption rate of 2.5×1011 molecules cm-2 s-1 was calculated. To estimate a saturation value for the NO2 adsorption on carbonaceous aerosol particles, bulk experiments were performed where the aerosol was deposited on a filter before exposure to NO2. This gives a lower limit for the total NO2 adsorption of about 1×1014 molecules cm-2 of particle surface area. The measurements show that the concept of the often used sticking coefficient γ (i.e. the number of adsorbed molecules per number of the total gas–surface collisions) is not a useful parameter to describe the chemisorption of NO2 at low ppb concentration on such complex surfaces as carbonaceous aerosol particles.  相似文献   

20.
An aircraft study of air quality in the Hong Kong region during the fall of 1994 has allowed for an estimation of the daytime source strengths for CO and NOy from the Hong Kong metropolitan center. Emission rate estimates for the Hong Kong urban plume for NOy and CO were 5.4×10e(25) molecules s-1 and 1.8×10e(26) molecules s-1 as determined for the case study of 18 October. All emission rate estimates have uncertainties of a factor of 2. On one occasion a distinct plume emanating from Shenzhen in the People’s Republic of China was encountered. While plume delimitation was insufficient for source strength calculations, transect integrals did allow for a CO/NOy ratio of about 16 to be determined. The CO/NOy ratio for the Hong Kong urban plume was about 3.3. The difference in these ratios indicates differences in the overall combustion processes and efficiencies taking place within Hong Kong and the PRC.  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