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 140 毫秒
1.
Twelve hours integrated fine particles (PM2.5) and 24-h average size-segregated particles were collected to investigate the chemical characteristics and to determine the size distribution of ionic species during October–December 1999 in three cities of different urban scale; Chongju, Kwangju, and Seoul, Korea. Concentrations of 5-min PM2.5 black carbon (BC) and hourly criteria air pollutants (PM10, CO, NOx, SO2, and O3) were also measured using the Aethalometer and ambient air monitoring system, respectively.Highest PM2.5 mass concentrations at Chongju, Kwangju, and Seoul sites were 63.0, 77.9, and 143.7 μg m−3, respectively. For the time period when highest PM2.5 mass occurred, BC level out of PM2.5 chemical species was highest at both Chongju and Kwangju, and highest NO3 (23.6 μg m−3) followed by BC (23.1 μg m−3) were observed at Seoul site, indicating that highest PM2.5 pollution is closely associated with the traffic emissions. Strong relationships of Fe with BC and Zn at both Kwangju and Seoul sites support that the Fe and Zn measured there are originated partly from same source as BC, i.e. diesel traffics. However, it is suggested that the Fe measured at Chongju is most likely derived from dispersion of soil dust.The size distributions of SO42−, NO3, and NH4+ ionic species indicated similar unimodal distributions at all sampling sites. However, different unimodal patterns in the accumulation mode size range with a peak in the smaller size (0.28–0.53 μm, condensation mode) in both Kwangju and Seoul, and in the relatively larger size (0.53–1.0 μm, droplet mode) in Chongju, were found. The potassium ion under the study sites dominates in the fine mode, and its size distribution showed unimodal character with a maximum in the size range 0.56–1.0 μm.  相似文献   

2.
Aircraft measurements of air pollutants were made to investigate the characteristic features of long-range transport of sulfur compounds over the Yellow Sea for the periods of 26–27 April and 7–10 November in 1998, and 9–11 April and 19 June in 1999, together with aerosol measurements at the Taean background station in Korea. The overall mean concentrations of SO2, O3 and aerosol number in the boundary layer for the observation period ranged 0.1–7.4 ppb 32.1–64.1 ppb and 1.0–143.6 cm−3, respectively. It was found that the air mass over the Yellow Sea had a character of both the polluted continental air and clean background air, and the sulfur transport was mainly confined in the atmospheric boundary layer. The median of SO2 concentration within the boundary layer was about 0.1–2.2 ppb. However, on 8 November, 1998, the mean concentrations of SO2 and aerosol number increased up to 7.4 ppb and 109.5 cm−3, respectively, in the boundary layer, whereas O3 concentration decreased remarkably. This enhanced SO2 concentration occurred in low level westerly air stream from China to Korea. Aerosol analyses at the downstream site of Taean in Korea showed 2–3 times higher sulfate concentration than that of other sampling days, indicating a significant amount of SO2 conversion to non sea-salt sulfate during the long-range transport.  相似文献   

3.
A highly sensitive cavity ring-down spectrometer (CRDS) was used to monitor the aerosol extinction coefficient at 532 nm. The performance of the spectrometer was evaluated using measurements of nearly monodisperse polystyrene particles with diameters between 150 and 500 nm. By comparing the observed results with those determined using Mie theory, the accuracy of the CRDS instrument was determined to be >97%, while the upper limit for the precision of the instrument was estimated to be 0.6–3.5% (typically 2%), depending on the particle number concentration, which was in the range of 30–2300 particles cm?3. Simultaneous measurements of the extinction (bext), scattering (bsca) and absorption (babs) coefficients of ambient aerosols were performed in central Tokyo from 14 August to 2 September 2007 using the CRDS instrument, two nephelometers and a particle/soot absorption photometer (PSAP), respectively. The value of bext measured using the CRDS instrument was compared with the sum of the bsca and babs values measured with a nephelometer and a PSAP, respectively. Good agreement between the bext and bsca + babs values was obtained except for data on days when high ozone mixing ratios (>130 ppbv) were observed. During the high-O3 days, the values for bsca + babs were ~7% larger than the value for bext, possibly because the value for babs measured by the PSAP was overestimated due to interference from coexisting non-absorbing aerosols such as secondary organic aerosols.  相似文献   

4.
Airborne measurements of the growth of the marine accumulation mode after multiple cycles through stratocumulus cloud are presented. The nss-sulphate cloud residual mode was log-normal in spectral shape and it’s mode radius was observed to progressively increase in size from 0.78 to 0.94 μm over 155 min of air parcel evolution through the cloudy marine boundary layer. The primary reason for this observed growth was thought to result from aqueous phase oxidation of SO2 to aerosol sulphate in activated cloud drops. An aqueous phase aerosol–cloud-chemistry model was used to simulate this case study of aerosol growth and was able to closely reproduce the observed growth. The model simulations illustrate that aqueous phase oxidation of SO2 in cloud droplets was able to provide enough additional sulphate mass to increase the size of activated aerosol. During a typical cloud cycle simulation, ≈4.6 nmoles kg-1air (0.44 μg m-3) of sulphate mass was produced with ≈70% of sulphate production occurring in cloud droplets activated upon sea-salt nuclei and ≈30% occurring upon nss-sulphate nuclei, even though sea-salt nuclei contributed less than 15% to the activated droplet population. The high fraction of nss-sulphate mass internally mixed with sea-salt aerosol suggests that aqueous phase oxidation of SO2 in cloud droplets activated upon sea-salt nuclei is the dominant nss-sulphate formation mechanism and that sea-salt aerosol provides the primary chemical sink for SO2 in the cloudy marine boundary layer.  相似文献   

5.
Seawater, atmospheric dimethylsulfide (DMS) and aerosol compounds, potentially linked with DMS oxidation, such as methanesulfonic acid (MSA) and non-sea-salt sulfate (nss-SO42?) were determined in the North Yellow Sea, China during July–August, 2006. The concentrations of seawater and atmospheric DMS ranged from 2.01 to 11.79 nmol l?1 and from 1.68 to 8.26 nmol m?3, with average values of 6.20 nmol l?1 and 5.01 nmol m?3, respectively. Owing to the appreciable concentration gradient, DMS accumulated in the surface water was transferred into the atmosphere, leading to a net sea-to-air flux of 6.87 μmol m?2 d?1 during summer. In the surface seawater, high DMS values corresponded well with the concurrent increases in chlorophyll a levels and a significant correlation was observed between integrated DMS and chlorophyll a concentrations. In addition, the concentrations of MSA and nss-SO42? measured in the aerosol samples ranged from 0.012 to 0.079 μg m?3 and from 3.82 to 11.72 μg m?3, with average values of 0.039 and 7.40 μg m?3, respectively. Based on the observed MSA, nss-SO42? and their ratio, the relative biogenic sulfur contribution was estimated to range from 1.2% to 11.5%, implying the major contribution of anthropogenic source to sulfur budget in the study area.  相似文献   

6.
Fine particulate matter (PM2.5) was sampled at 5 Spanish locations during the European Community Respiratory Health Survey II (ECRHS II). In an attempt to identify and quantify PM2.5 sources, source contribution analysis by principal component analysis (PCA) was performed on five datasets containing elemental composition of PM2.5 analysed by ED-XRF. A total of 4–5 factors were identified at each site, three of them being common to all sites (interpreted as traffic, mineral and secondary aerosols) whereas industrial sources were site-specific. Sea-salt was identified as independent source at all coastal locations except for Barcelona (where it was clustered with secondary aerosols). Despite their typically dominant coarse grain-size distribution, mineral and marine aerosols were clearly observed in PM2.5. Multi-linear regression analysis (MLRA) was applied to the data, showing that traffic was the main source of PM2.5 at the five sites (39–53% of PM2.5, 5.1–12.0 μg m−3), while regional-scale secondary aerosols accounted for 14–34% of PM2.5 (2.6–4.5 μg m−3), mineral matter for 13–31% (2.4–4.6 μg m−3) and sea-salt made up 3–7% of the PM2.5 mass (0.4–1.3 μg m−3). Consequently, despite regional and climatic variability throughout Spain, the same four main PM2.5 emission sources were identified at all the study sites and the differences between the relative contributions of each of these sources varied at most 20%. This would corroborate PM2.5 as a useful parameter for health studies and environmental policy-making, owing to the fact that it is not as subject to the influence of micro-sitting as other parameters such as PM10. African dust inputs were observed in the mineral source, adding on average 4–11 μg m−3 to the PM2.5 daily mean during dust outbreaks. On average, levels of Al, Si, Ti and Fe during African episodes were higher by a factor of 2–8 with respect to non-African days, whereas levels of local pollutants (absorption coefficient, S, Pb, Cl) showed smaller variations (factor of 0.5–2).  相似文献   

7.
The aerosol scattering properties were investigated at two continental sites in northern China in 2004. Aerosol light scattering coefficient (σsp) at 525 nm, PM10, and aerosol mass scattering efficiencies (α) at Dunhuang had a mean value of 165.1±148.8 M m−1, 157.6±270.0 μg m−3, and 2.30±3.41 m2 g−1, respectively, while these values at Dongsheng were, respectively, 180.2±151.9 M m−1, 119.0±112.9 μg m−3, and 1.87±1.41 m2 g−1. There existed a seasonal variability of aerosol scattering properties. In spring, at Dunhuang PM10, σsp, and α were 184.1±211.548 μg m−3, 126.3±89.6 M m−1, and 1.05±0.97 m2 g−1, respectively, and these values at Dongsheng were 146.4±142.1 μg m−3, 183.4±81.7 M m−1, and 1.98±1.52 m2 g−1, respectively. However, in winter at Dunhuang PM10, σsp, and α were 158.1±261.4 μg m−3, 303.3±165.2 M m−1, and 3.17±1.93 m2 g−1, respectively, and these values at Dongsheng were 155.7±170.1 μg m−3, 304.4±158.1 M m−1, and 2.90±1.72 m2 g−1, respectively. σsp and α in winter were higher than that in spring at both the sites, which coincides with the characteristics of dust aerosol and pollution aerosol. Overall, the dominant aerosol types in spring and winter at both sites in northern China are dust aerosol and pollution aerosol, respectively.  相似文献   

8.
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   

9.
The gas-phase ozonolysis of (E)-β-farnesene was investigated in a 3.91 m3 atmospheric simulation chamber at 296 ± 2 K and relative humidity of around 0.1%. The relative rate method was used to determine the reaction rate coefficient of (4.01 ± 0.17) × 10?16 cm3 molecule?1 s?1, where the indicated errors are two least-squares standard deviations and do not include uncertainties in the rate coefficients for the reference compounds (γ-terpinene, cis-cyclooctene and 1,5-cyclooctadiene). Gas phase carbonyl products were collected using a denuder sampling technique and analyzed with GC/MS following derivatization with O-(2,3,4,5,6-pentafluorobenzyl) hydroxylamine (PFBHA). The reaction products detected were acetone, 4-oxopentanal, methylglyoxal, 4-methylenehex-5-enal, 6-methylhept-5-en-2-one, and (E)-4-methyl-8-methylenedeca-4,9-dienal. A detailed mechanism for the gas-phase ozonolysis of (E)-β-farnesene is proposed, which accounts for all of the products observed in this study. The results of this work indicate that the atmospheric reaction of (E)-β-farnesene with ozone has a lifetime of around 1 h and is another possible source of the ubiquitous carbonyls, acetone, 4-oxopentanal and 6-methylhept-5-en-2-one in the atmosphere.  相似文献   

10.
Real-world emissions of a traffic fleet on a transit route in Austria were determined in the Tauerntunnel experiment in October 1997. The total number of vehicles and the average speed was nearly the same on both measuring days (465 vehicles 30 min−1 and 76 km h−1 on the workday, 477 and 78 km h−1 on Sunday). The average workday fleet contained 17.6% heavy-duty vehicles (HDV) and the average Sunday fleet 2.8% HDV resulting in up to four times higher emission rates per vehicle per km on the workday than on Sunday for most of the regulated components (CO2, CO, NOx, SO2, and particulate matter-PM10). Emission rates of NMVOC accounted for 200 mg vehicle−1 km−1 on both days. The relative contributions of light-duty vehicles (LDV) and HDV to the total emissions indicated that aldehydes, BTEX (benzene, toluene, ethylbenzene, xylenes), and alkanes are mainly produced by LDV, while HDV dominated emissions of CO, NOx, SO2, and PM10. Emissions of NOx caused by HDV were 16,100 mg vehicle−1 km−1 (as NO2). Produced by LDV they were much lower at 360 mg vehicle−1 km−1. Comparing the emission rates to the results that were obtained by the 1988 experiment at the same place significant changes in the emission levels of hydrocarbons and CO, which accounted 1997 to only 10% of the levels in 1988, were noticed. However, the decrease of PM has been modest leading to values of 80 and 60% of the levels in 1988 on the workday and on Sunday, respectively. Emission rates of NOx determined on the workday in 1997 were 3130 mg vehicle−1 km−1 and even higher than in 1988 (2630 mg vehicle−1 km−1), presumable due to the increase of the HD-traffic.  相似文献   

11.
Ammonia-nitrogen flux (NH3-N=(14/17)NH3) was determined from six anaerobic swine waste storage and treatment lagoons (primary, secondary, and tertiary) using the dynamic chamber system. Measurements occurred during the fall of 1998 through the early spring of 1999, and each lagoon was examined for approximately one week. Analysis of flux variation was made with respect to lagoon surface water temperature (∼15 cm below the surface), lagoon water pH, total aqueous phase NHx(=NH3+NH4+) concentration, and total Kjeldahl nitrogen (TKN). Average lagoon temperatures (across all six lagoons) ranged from approximately 10.3 to 23.3°C. The pH ranged in value from 6.8 to 8.1. Aqueous NHx concentration ranged from 37 to 909 mg N l−1, and TKN varied from 87 to 950 mg N l−1. Fluxes were the largest at the primary lagoon in Kenansville, NC (March 1999) with an average value of 120.3 μg N m−2 min−1, and smallest at the tertiary lagoon in Rocky Mount, NC (November 1998) at 40.7 μg N m−2 min−1. Emission rates were found to be correlated with both surface lagoon water temperature and aqueous NHx concentration. The NH3-N flux may be modeled as ln(NH3-N flux)=1.0788+0.0406TL+0.0015([NHx]) (R2=0.74), where NH3-N flux is the ammonia flux from the lagoon surface in μg N m−2 min−1, TL is the lagoon surface water temperature in °C, and [NHx] is the total ammonia-nitrogen concentration in mg N l−1.  相似文献   

12.
Fifty-five seasonal PM2.5 samples were collected March 2003–January 2004 at Changdao, a resort island located at the demarcation line between Bohai Sea and Yellow Sea in Northern China. Changdao is in the transport path of the continental aerosols heading toward the Pacific Ocean in winter and spring due to the East Asia Monsoon. Solvent-extractable organic compounds (SEOC), organic carbon (OC), elemental carbon (EC) and water-soluble organic carbon (WSOC) were analyzed for source identification based on molecular markers. This data set provides useful information for the downstream site researchers of the Asian continental outflow. Total carbon (TC, OC+EC) was ∼18 μg m−3 in winter, ∼9 μg m−3 in spring and autumn and a large part of the TC was WSOC (33% in winter, >45% in the other seasons). Winter and spring were the high SEOC seasons with n-fatty acids the highest at ∼290 and ∼170 ng m−3, respectively, followed by n-alkanes at ∼210 and ∼90 ng m−3, and polycyclic aromatic hydrocarbons (PAHs) were also at high at ∼120 and ∼30 ng m−3. High WSOC/TC, low C18:1/C18 of fatty acids, and low concentrations of labile PAHs such as benzo(a)pyrene, together with back trajectory analysis suggested that the aerosols were aged and transported. PAHs, triterpane and sterane distributions provided evidence that coal burning was the main source of the continental outflow. The detection of levoglucosan and β-sitosterol in nearly all the samples showed the impact of biomass burning.  相似文献   

13.
A bimolecular rate constant, kOH+Benzyl alcohol, of (28 ± 7) × 10?12 cm3 molecule?1 s?1 was measured using the relative rate technique for the reaction of the hydroxyl radical (OH) with benzyl alcohol, at (297 ± 3) K and 1 atm total pressure. Additionally, an upper limit of the bimolecular rate constant, kO3+Benzyl alcohol, of approximately 6 × 10?19 cm3 molecule?1 s?1 was determined by monitoring the decrease in benzyl alcohol concentration over time in an excess of ozone (O3). To more clearly define part of benzyl alcohol's indoor environment degradation mechanism, the products of the benzyl alcohol + OH were also investigated. The derivatizing agents O-(2,3,4,5,6-pentafluorobenzyl)hydroxylamine (PFBHA) and N,O-bis(trimethylsilyl) trifluoroacetamide (BSTFA) were used to positively identify benzaldehyde, glyoxal and 4-oxopentanal as benzyl alcohol/OH reaction products. The elucidation of other reaction products was facilitated by mass spectrometry of the derivatized reaction products coupled with plausible benzyl alcohol/OH reaction mechanisms based on previously published volatile organic compound/OH gas-phase reaction mechanisms.  相似文献   

14.
Due to the high temporal and spatial variability of N2O fluxes, estimates of N2O emission from temperate forest ecosystems are still highly uncertain, particularly at larger scales. Although highest N2O emissions with up to 7.0 kg N ha−1 yr−1 were mainly reported for soils affected by stagnant water, most of the reported gas flux measurements were performed at forest sites with well-aerated soils yielding mostly to low mean annual emission rates less than 1.0 kg N ha−1 yr−1. This study compares N2O fluxes from upland (Cambisols) and temporally water-logged (Gleysols, Histosols) soils of the Central Black Forest (South-West Germany) over a period of 2 yr. Mean annual N2O fluxes from investigated soils ranged between 0.2 and 3.9 kg N ha−1 yr−1. The fluxes showed a large variability between the different soil types. Emissions could be clearly ranked in the following order: Cambisols (0.26–0.75 kg N ha−1 yr−1)<Gleysols (1.37–2.68 kg N ha−1 yr−1)<Histosol (3.66–3.95 kg N ha−1 yr−1). Although the Cambisols cover two-thirds of the investigated area, only about half of the overall N2O is emitted from this soil type. Therefore, regional or national N2O fluxes from temperate forest soils are underestimated if soils characterised by intermediate aeration conditions are disregarded.  相似文献   

15.
The emissions of VOC from freshly cut and shredded Grevillea robusta (Australian Silky Oak) leaves and wood have been measured. The VOC emissions from fresh leaf mulch and wood chips lasted typically for 30 and 20 h respectively, and consisted primarily of ethanol, (E)-2-hexenal, (Z)-3-hexen-1-ol and acetaldehyde. The integrated emissions of the VOCs were 0.38±0.04 g kg−1 from leaf mulch, and 0.022±0.003 g kg−1 from wood chips. These emissions represent a source of VOCs in urban and rural air that has previously been unquantified and is currently unaccounted for. These VOCs from leaf mulch and wood chips will contribute to both urban photochemistry and secondary organic aerosol formation. Any CH4 emissions from leaf mulch and wood chips were <1×10−11 g g dry mass−1 s−1.  相似文献   

16.
Dry deposition modelling typically assumes that canopy resistance (Rc) is independent of ammonia (NH3) concentration. An innovative flux chamber system was used to provide accurate continuous measurements of NH3 deposition to a moorland composed of a mixture of Calluna vulgaris (L.) Hull, Eriophorum vaginatum L. and Sphagnum spp. Ammonia was applied at a wide range of concentrations (1–100 μg m−3). The physical and environmental properties and the testing of the chamber are described, as well as results for the moorland vegetation using the ‘canopy resistance’ and ‘canopy compensation point’ interpretations of the data.Results for moorland plant species demonstrate that NH3 concentration directly affects the rate of NH3 deposition to the vegetation canopy, with Rc and cuticular resistance (Rw) increasing with increasing NH3 concentrations. Differences in Rc were found between night and day: during the night Rc increases from 17 s m−1 at 10 μg m−3 to 95 s m−1 at 80 μg m−3, whereas during the day Rc increases from 17 s m−1 at 10 μg m−3 to 48 s m−1 at 80 μg m−3. The lower resistance during the day is caused by the stomata being open and available as a deposition route to the plant. Rw increased with increasing NH3 concentrations and was not significantly different between day and night (at 80 μg m−3 NH3 day Rw=88 s m−1 and night Rw=95 s m−1). The results demonstrate that assessments using fixed Rc will over-estimate NH3 deposition at high concentrations (over ∼15 μg m−3).  相似文献   

17.
Currently, in operational modelling of NH3 deposition a fixed value of canopy resistance (Rc) is generally applied, irrespective of the plant species and NH3 concentration. This study determined the effect of NH3 concentration on deposition processes to individual moorland species. An innovative flux chamber system was used to provide accurate continuous measurements of NH3 deposition to Deschampsia cespitosa (L.) Beauv., Calluna vulgaris (L.) Hull, Eriophorum vaginatum L., Cladonia spp., Sphagnum spp., and Pleurozium schreberi (Brid.) Mitt. Measurements were conducted across a wide range of NH3 concentrations (1–140 μg m−3).NH3 concentration directly affects the deposition processes to the vegetation canopy, with Rc, and cuticular resistance (Rw) increasing with increasing NH3 concentration, for all the species and vegetation communities tested. For example, the Rc for C. vulgaris increased from 14 s m−1 at 2 μg m−3 to 112 s m−1 at 80 μg m−3. Diurnal variations in NH3 uptake were observed for higher plants, due to stomatal uptake; however, no diurnal variations were shown for non-stomatal plants. Rc for C. vulgaris at 80 μg m−3 was 66 and 112 s m−1 during day and night, respectively. Differences were found in NH3 deposition between plant species and vegetation communities: Sphagnum had the lowest Rc (3 s m−1 at 2 μg m−3 to 23 at 80 μg m−3), and D. cespitosa had the highest nighttime value (18 s m−1 at 2 μg m−3 to 197 s m−1 at 80 μg m−3).  相似文献   

18.
For over one year, the Environmental Protection Commission of Hillsborough County (EPCHC) in Tampa, Florida, operated two dichotomous sequential particulate matter air samplers collocated with a manual Federal Reference Method (FRM) air sampler at a waterfront site on Tampa Bay. The FRM was alternately configured as a PM2.5, then as a PM10 sampler. For the dichotomous sampler measurements, daily 24-h integrated PM2.5 and PM10–2.5 ambient air samples were collected at a total flow rate of 16.7 l min−1. A virtual impactor split the air into flow rates of 1.67 and 15.0 l min−1 onto PM10–2.5 and PM2.5 47-mm diameter PTFE® filters, respectively. Between the two dichotomous air samplers, the average concentration, relative bias and relative precision were 13.3 μg m−3, 0.02% and 5.2% for PM2.5 concentrations (n=282), and 12.3 μg m−3, 3.9% and 7.7% for PM10–2.5 concentrations (n=282). FRM measurements were alternate day 24-h integrated PM2.5 or PM10 ambient air samples collected onto 47-mm diameter PTFE® filters at a flow rate of 16.7 l min−1. Between a dichotomous and a PM2.5 FRM air sampler, the average concentration, relative bias and relative precision were 12.4 μg m−3, −5.6% and 8.2% (n=43); and between a dichotomous and a PM10 FRM air sampler, the average concentration, relative bias and relative precision were 25.7 μg m−3, −4.0% and 5.8% (n=102). The PM2.5 concentration measurement standard errors were 0.95, 0.79 and 1.02 μg m−3; for PM10 the standard errors were 1.06, 1.59, and 1.70 μg m−3 for two dichotomous and one FRM samplers, respectively, which indicate the dichotomous samplers have superior technical merit. These results reveal the potential for the dichotomous sequential air sampler to replace the combination of the PM2.5 and PM10 FRM air samplers, offering the capability of making simultaneous, self-consistent determinations of these particulate matter fractions in a routine ambient monitoring mode.  相似文献   

19.
Conductometry was used to study the kinetics of the oxidation of hydrogen sulfite, HSO3, by hydrogen peroxide in aqueous non-buffered solution at the low concentration level of 10−5–10−6 M, typically found in cloud water. The kinetic data confirm that the rate law reported for the pH range 3–6 at higher concentration levels, rate=kH·[H+]·[HSO3]·[H2O2], is valid at the low concentration level and at low ionic strength Ic. At 298 K and Ic=1.5×10−4 M, third-order rate constant kH was found to be kH=(9.1±0.5)×107 M−2 s−1. The temperature dependence of kH led to an activation energy of Ea=29.7±0.9 kJ mol−1. The effect of the ionic strength (adjusted with NaCl) on rate constant kH was studied in the range Ic=2×10−4–5.0 M at pH=4.5–5.2 by conductometry and stopped-flow spectrophotometry. The dependence of kH on Ic can be described with a semi-empirical relationship, which is useful for the purpose of comparison and extrapolation. The kinetic data obtained are critically compared with those reported earlier.  相似文献   

20.
Uptake of aromatic hydrocarbons (AH) by ice crystals during vapor deposit growth was investigated in a walk-in cold chamber at temperatures of 242, 251, and 260 K, respectively. Ice crystals were grown from ambient air in the presence of gaseous AH namely: benzene (C6H6), toluene (methylbenzene, C7H8), the C8H10 isomers ethylbenzene, o-, m-, p-xylene (dimethylbenzenes), the C9H12 isomers n-propylbenzene, 4-ethyltoluene, 1,3,5-trimethylbenzene (1,3,5-TMB), 1,2,4-trimethylbenzene (1,2,4-TMB), 1,2,3-trimethylbenzene (1,2,3-TMB), and the C10H14 compound tert.-butylbenzene. Gas-phase concentrations calculated at 295 K were 10.3–20.8 μg m−3. Uptake of AH was detected by analyzing vapor deposited ice with a very sensitive method composed of solid-phase micro-extraction (SPME), followed by gas chromatography/mass spectrometry (GC/MS).Ice crystal size was lower than 1 cm. At water vapor extents of 5.8, 6.0 and 8.1 g m−3, ice crystal shape changed with decreasing temperatures from a column at a temperature of 260 K, to a plate at 251 K, and to a dendrite at 242 K. Experimentally observed ice growth rates were between 3.3 and 13.3×10−3 g s−1 m−2 and decreased at lower temperatures and lower value of water vapor concentration. Predicted growth rates were mostly slightly higher.Benzene, toluene, ethylbenzene, and xylenes (BTEX) were not detected in ice above their detection limits (DLs) of 25 pg gice−1 (toluene, ethylbenzene, xylenes) and 125 pg gice−1 (benzene) over the entire temperature range. Median concentrations of n-propylbenzene, 4-ethyltoluene, 1,3,5-TMB, tert.-butylbenzene, 1,2,4-TMB, and 1,2,3-TMB were between 4 and 176 pg gice−1 at gas concentrations of 10.3–10.7 μg m−3 calculated at 295 K. Uptake coefficients (K) defined as the product of concentration of AH in ice and density of ice related to the product of their concentration in the gas phase and ice mass varied between 0.40 and 10.23. K increased with decreasing temperatures. Values of Gibbs energy (ΔG) were between −4.5 and 2.4 kJ mol−1 and decreased as temperatures were lowered. From the uptake experiments, the uptake enthalpy (ΔH) could be determined between −70.6 and −33.9 kJ mol−1. The uptake entropy (ΔS) was between −281.3 and −126.8 J mol−1 K−1. Values of ΔH and ΔS were rather similar for 4-ethlytoluene, 1,3,5-TMB and tert.-butylbenzene, whereas 1,2,3-TMB showed much higher values.  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