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 123 毫秒
1.
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   

2.
Equilibrium gas phase concentration of ammonia in dilute solution has been measured as a function of total ammonia + ammonium concentration (0.002–0.10 M), pH (6–10) and temperature (278.8−290.6 K). Henry's Law is obeyed under these conditions and may be expressed as In KH(M atm−1) = 4092/T −9.70 with a relative standard error of less than 5 %, in good agreement with NBS thermodynamic data. Convenient generation of trace levels of ammonia (1.33 × 10−8–7.77 × 10−4 atm) using a porous membrane tube is described.  相似文献   

3.
A particular agricultural waste, peanut shell, has been used as precursor for activated carbon production by chemical activation with H3PO4. Unoxidized activated carbon was prepared in nitrogen atmosphere which was then heated in air at a desired temperature to get oxidized activated carbon. The prepared carbons were characterized for surface area, surface morphology, and pore volume and utilized for the removal of Cr(VI) from aqueous solution. Batch mode experiments were conducted to study the effects of pH, contact time, particle size, adsorbent dose, initial concentration of adsorbate, and temperature on the adsorption of Cr(VI). Cr(VI) adsorption was significantly dependent on solution pH, and the optimum adsorption was observed at pH 2. Pseudo-first-order, pseudo-second-order, and intraparticle diffusion models were used to analyze the kinetic data obtained at different initial Cr(VI) concentrations. The adsorption kinetic data were described very well by the pseudo-second-order model. Equilibrium isotherm data were analyzed by the Langmuir, Freundlich, and Temkin models. The results showed that the Langmuir adsorption isotherm model fitted the data better in the temperature range studied. The adsorption capacity which was found to increase with temperature showed the endothermic nature of Cr(VI) adsorption. The thermodynamic parameters, such as Gibb’s Free energy change (ΔG°), standard enthalpy change (ΔH°), and standard entropy change (ΔS°) were evaluated.  相似文献   

4.
Sub-cooled liquid vapor pressures (PL 0) of current–use organochlorine and organophosphate pesticides (chlorothalonil, chlorpyrifos methyl, diazinon, fipronil) and selected transformation products (chlorpyrifos oxon, heptachlor epoxide, oxychlordane, 3,5,6-trichloro-2-pyridinol) were determined at multiple temperatures using the gas chromatography retention time technique. Results were utilized to determine vapor pressure-temperature relationships and to calculate enthalpies of vaporization (ΔHvap). While results for chlorothalonil and diazinon were comparable with published values, the measured value for fipronil (1.82 × 10? 6 Pa) is almost an order of magnitude higher than the reported literature value (3.7 × 10? 7 Pa). The availability of vapor pressure temperature relationships for these chemicals will aid in pesticide risk assessment development and improve the effectiveness of mitigation and remediation efforts.  相似文献   

5.
This paper focuses on a detailed analysis of the effects of meteorological factors explaining the variability of rain composition.Inorganic composition of 113 individual rain events was measured from May 2002 to October 2005 at a rural site near Chimay, in the western part of the Belgian Ardennes. Original models were fitted for each studied ion (H+, Mg2+, Ca2+, K+, NH4+, Na+, Cl, NO3 and SO42−) to relate rain event concentration or wet deposition to the rainfall volume (R), the length of the antecedent dry period (ADP), the volume of the previous event (Rprev) as well as to the mean wind speed and the prevailing wind direction during both the dry and the rainy periods. These variables explained from 32% (H+) to 69% (NO3) of rain concentration variability. Concentrations decreased logarithmically with increasing R values except in case of H+ for which a positive effect of rain volume on rain concentration was observed. ADP affected positively rain concentrations of all ions excluding K+ and H+ for which, respectively, a nonsignificant and a negative effect of this variable was observed. Increasing Rprev strengthened the effect of the variable R on H+, Mg2+, Ca2+, Na+, NH4+ and SO42− concentrations while it softened the effect of ADP on NO3 concentrations. Wind speed and direction during dry and rainy periods explained together from 8% (K+) to 38% (Na+) of rain concentration total variability. R2 coefficients of the wet deposition models ranged from 0.51 (K+) to 0.79 (SO42−). For all ions, wet deposition increased significantly with increasing R values while the effects of the other variables were similar to those on concentrations. Wind conditions during dry and rainy periods explained from 4% (H+) to 24% (Na+) of wet deposition total variability. On an annual scale, the total dry period duration, the total rainfall volume as well as the shape of the distributions of the length of the antecedent dry periods and of the rain event volume are important parameters that influence annual wet deposition.  相似文献   

6.
A single ammonium-hydrogen-sulfate particle is levitated in an evacuated quadrupole trap at room temperature and the temperature of an attached tube containing bulk water is slowly cycled introducing then removing water vapor. With increasing pressure the particle dissolves in stages, then grows as a solution droplet by water absorption. With decreasing pressure the droplet supersaturates, crystallizes, then dehydrates completely to return to its initial state. Particle mass, and thus composition, is measured continuously with an electrostatic balance. Twenty-six cycles were studied as solute composition ranged from ammonium bisulfate through letovicite to ammonium sulfate in roughly equal steps. Composition was changed in situ by reaction with ammonia at low partial pressure. With solute composition characterized by x = [NH4]/[SO4], deliquescence was found to occur at water activity aw = 0.394−0.029 (x− 1) for 1 ⩽ x < 1.5 and aw = 0.710−0.023(x−1.5) for 1.5 ⩽ x < 2. Particle growth occurs at deliquescence and subsequently is in excellent agreement with that predicted in a model proposed by Tang for dissolution of a two-component mixed solute. Water activities of the solution droplets are measured up to aw = 0.9. The results are compared with those predicted by the Zdanovskii-Stokes-Robinson method of interpolation from binary data and with those obtained using the mixing rule of Meissner and Kusik. Particle crystallization from supersaturated solution is analyzed thermodynamically using measured water activities, the Gibbs-Duhem equation, and classical nucleation theory. The specific free energy barrier to crystallization, ΔG/n, is found to increase from near zero to 0.04 eV as composition ranges from x = 1 to 2, where n is the number of formula units in the critical nucleus. New phase diagrams are presented and used to discuss the dynamics of mixed sulfate particles in the atmosphere.  相似文献   

7.
Deposition processes of particles with dry diameter larger than about 10 μm are dominated by gravitational settling, while molecular diffusion and Brownian motion predominate the deposition processes of particles smaller than 0.1 μm in dry diameter. Many air pollution derived elements exhibit characteristics common to sub-micron particles. The objective of the present study is to examine the effects of meteorological conditions within the turbulent transfer layer on the deposition velocity of particles with dry diameter between 0.1 and 1 μm. It is for these sub-micron particles that particle growth by condensation in the deposition layer, the broken water surface effect and the enhanced transfer process due to atmospheric turbulence in the turbulent transfer layer play important roles in controlling the particle deposition velocity. Results of the present study show that the `dry air’ assumption of Williams’ model is unrealistic. Effects of ambient air relative humidity and water surface temperature cannot be ignored in determining the deposition velocity over a water surface. Neglecting effects of ambient air relative humidity and water surface temperature will result in defining atmospheric stability incorrectly. It is found that the largest effect of air relative humidity on deposition velocity occurs at an air–water temperature difference corresponding to the point of `displaced neutral stability'. For a given wind speed of U=5 m s−1 the additive effects of water surface temperature, Tw, changes from 5 to 25°C and ambient air relative humidity variations from 85 to 60%, respectively, lead to a maximum difference in vd of about 20%. For a higher wind speed of 10 m s−1, however, the corresponding change in vd reduces to less than 5%. This is further confirmation that wind speed is one of the strongest variables that governs the magnitude of vd. The present study also found that the broken surface transfer coefficient, kbs, given as a multiple of the smooth surface transfer coefficient, kss, is physically more meaningful than assigning it a constant value independent of particle size. The method used in this study requires only a single level of atmospheric data coupled with the surface temperature measurement. The present method is applicable for determining deposition velocity not only at the conventional measurement height of 10 m but also at any other heights that are different from the measurement height.  相似文献   

8.
Airborne particles of diameter > 0.4 μm reaching Dye 3, Greenland during April–May 1983 were highly variable in size and concentration from day to day. Five-day backward air mass trajectories suggest the importance of long-range transport from more northerly latitudes on days with high concentrations; particle sizes were larger on these days. Lower concentrations and smaller particle sizes were associated with transport from the south. It is inferred that Dye 3 may receive material emitted from Eurasian sources and transported over the Pole, similar to inferences for more northern Arctic sites.Elemental analysis of individual particles showed an abundance of crustal material, with many particles also containing sulfur. Bulk chemical analyses of airborne particles and fresh snow, collected during three snowstorms where ice nucleation dominated, provided data which were used to estimate mass-basis scavenging ratios. Average scavenging ratios were in the range ~1000–2000 for the crustal elements Al, Fe, K, Mg, Mn, and Na. Similar values were observed for Cd, Cu and NO3. The corresponding ratios for Pb and SO42− averaged less than 200. These ratios were used with precipitation rate data to estimate wet deposition velocities in the order of ~2 cm s−1 for the first nine species, and ~0.2 cm s−1 for Pb and SO42−. Comparing fresh and older surface snow concentrations gave an average dry deposition velocity of roughly 0.2 cm s−1 for the crustal elements, with the small fraction of large particles (~5–10 μm) dominating deposition; much smaller values were associated with the remaining species. When used with other data in the literature, the results of this study suggest that total deposition velocities of Pb and SO42− may be as small as 0.05 cm s−1 in relatively dry regions of the Arctic.  相似文献   

9.
Wet and dry deposition as collected by a bucket were measured at two sites in southeastern Michigan for two years. The precipitation had an average pH of 4.27 and a SO2−4 to NO3 ratio of 2.0. Particulate dry deposition velocities of 0.6 cm s−1 for SO2−4 and NO3 and > 2 cm s−1 for Cl, Ca2+, Mg2+,Na+ and K+ were calculated. The ambient particle composition, dry bucket collection and wet deposition were compared at two sites, one urban and the other rural. Higher ambient particle concentrations and dry deposition rates were measured at the urban site than the rural site, indicating the influence of local emissions. However, local emissions had no effect on the wet deposition concentrations. The influence of more distant source regions was examined by separating the precipitation events by wind direction. The events from the south and east had the highest SO2−4 to NO3 ratios, which corresponded to the areas with the highest sulfur emissions. NO3 showed no directional dependence.Wet deposition was examined for the effect of storm type and seasonal trends. Contrary to a recent study on Long Island, we found higher concentrations of H+, SO2−4 and NH+4 in winter rain compared to snow. The wet deposition concentrations of H+, SO2−4, and NH+4 were highest in the summer, while only Na+ and Cl concentrations were highest in the winter, presumably due to winter road salting. The total deposition of acidic ions was highest in the summer and lowest in the winter, due both to lower concentrations and lower precipitation volumes in the winter. The dry deposition as collected by a bucket accounted for 1 % of total H+ deposition, 21 % of SO2−4 deposition, 27% of NO3 deposition, 50% of Cl deposition and 61 % of Ca2+ deposition.  相似文献   

10.
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   

11.
The behavior of a chemical species in the atmosphere depends on how it partitions between the gas and aerosol particulate phases. Available evidence indicates that this partitioning can be described by a compound- and temperature-dependent constant K = A(TSP)/F, where A and F are the atmospheric concentrations measured in the gas and particulate phases, respectively (ng m−3). TSP is concentration (in μg m−3) of total suspended particulate matter. K values within some classes of compounds are known to correlate with the corresponding log po values where po is the vapor pressure of the compound. (For a solid compound, the vapor pressure of the sub-cooled liquid should be used.) Such correlations can break down when portions of the compounds are bound inside the particles. This paper develops equations for predicting the magnitude of this effect. The specific case of polycyclic aromatic hydrocarbons is used as an example. Values for log A(TSP)/F are predicted to be significantly undervalued in a variety of circumstances compared with what would be expected at full equilibrium, even when the fraction of non-exchangeable material is only a few per cent. Similarly, φ values [ = F/(A + F)] can be significantly overvalued. A knowledge of the magnitude of this effect will be necessary for accurate predictive modeling of the fates of chemical species in the atmosphere.  相似文献   

12.
A study of deposition velocities to snow was conducted during the 1982–1983 and 1983–1984 winters at the University of Michigan Biological Station in northern Michigan. Weekly measurements were made of dry deposition rates to snow and the atmospheric concentrations of the depositing species. SO2, with an average concentration of 2.2 ppb, was the dominant atmospheric sulfur containing species. NO2, with an average concentration of 1.8 ppb, was the dominant atmospheric nitrogenous species. NO3 deposition was due primarily to HNO3, which averaged 0.2 ppb. The HNO3 deposition velocity averaged 1.4cm s−1. The SO2 deposition velocity varied with temperature, averaging 0.15 cm s−1 for samples with appreciable exposure time above − 3°C, and 0.06 cm s−1 for samples which remained below an ambient temperature of −3°C. Deposition velocities of Ca2+, Mg2+ , Na+, K+ and NH+4 were 2.1, 1.5, 0.44, 0.51 and 0.10cm s−1, respectively. The mass median diameters of these species were 4.4, 2.7, 1.8, 0.9 and 0.46 μm, respectively.  相似文献   

13.
Gas/particle distributions of atmospheric semi-volatile organic compounds (SOCs) are often measured using filter/sorbent samplers. Unfortunately, the adsorption of gaseous SOCs onto a filter can cause positive biases in the measured particle-phase concentrations, and negative biases in the measured gas-phase concentrations. When quartz fiber filters (QFFs) are used, surface-area-normalized gas/quartz partition coefficient (Kp,s, m3 m−2) values will be useful when estimating the magnitudes of these errors. Gas/QFF Kp,s values have been reported in the literature only for polycyclic aromatic hydrocarbons (PAHs) and n-alkanes. Gas/QFF Kp,s values were measured here for a series of polychlorinated dibenzodioxins (PCDDs) and polychlorinated dibenzofurans (PCDFs), and also for a range of PAHs. Within each of the three individual compound classes, plots of log Kp,s vs. log pLo (sub-cooled liquid vapor pressure) were found to be linear with slopes of approximately −1. At relative humidity (RH)=25%, the pooled log Kp,s data at 25°C for the three compound classes were correlated with log pLo nearly as well (r2=0.95) as were the data for the individual compound classes (r2≈0.97). In general, the Kp,s values for the PAHs and PCDD/PCDFs studied were found to be about a factor of 2 larger for partitioning to clean QFFs at RH=25% than for TMFs at RH=21–52%. Backup QFFs used in filter/sorbent sampling in a suburban area yielded Kp,s values for PAHs at RH=37% that were significantly lower than for clean QFFs at the same RH. (This may have been the result of the adsorption of ambient organic compounds that at least partially blocked the direct adsorption of the SOCs to the QFF surface). Therefore, when QFFs are used to separate atmospheric gas- and particle-phase SOCs, corrections for compound-dependent gas adsorption artifacts for QFFs may need to be carried out using Kp,s values that were obtained with ambient backup QFFs.  相似文献   

14.
Measurements of negative chemiions (CI) emitted by a jet engine at the ground were made with an ion trap mass spectrometer. The new instrument offered a high-mass resolution, which led to a first unambiguous identification of negative CI formed by a jet engine. The observed ions are HSO4(H2SO4)a clusters proved by an isotope study. From the mass spectra an efficiency ε for fuel sulfur conversion to SVI of 2%±0.8 could be inferred. In addition thermodynamic properties of the observed cluster ions were inferred from measured ion abundance ratios. An effective free energy ΔGa−1,a0=−14 kcal/mol was calculated (for a=3) and an enthalpy of ΔHa−1,a0=−24 (for a=3) kcal/mol was estimated. This indicates a low stability of HSO4(H2SO4)a (a⩾3) cluster ions against thermal detachment of H2SO4 at the high temperatures of our experiment. However the low temperatures at cruise altitudes around 10–12 km lead to high H2SO4/H2O supersaturation and therefore a rapid growth of HSO4(H2SO4)a cluster ions seems to be possible which is not hindered by thermal H2SO4 detachment.  相似文献   

15.
A new algorithm has been derived for trajectory models to determine the transfer coefficient of each source along or adjacent to a trajectory and to calculate the concentrations of SO2, NOx, sulfate, nitrate, fine particulate matter (PM) and coarse PM at a receptor. The transfer coefficient tf (s m−1) is defined to be the ratio between the contributed concentration ΔC (μg m−3) to the receptor from a ground source and the emission rate of the source q (μg m−2 s−1) at a grid, i.e. tf≡ΔC/q. The model is developed by combining with a backward trajectory scheme and a circuit-type's parameterization. First, the transfer coefficients of grids along or adjacent a back-trajectory are calculated. Then, the contributed concentration of each emission grid is determined by multiplying its emission rate with the transfer coefficient of the grid. Finally, the concentration at the receptor is determined by the summation of all the contributed concentrations within the domain of simulation.  相似文献   

16.
The DWNWND gaussian plume computer code was used to calculate normalized 22.5° sector-averaged dispersant concentrations at distances from 400 to 7000 meters downwind from a 56 m high release point. These results were compared to sector average values obtained in fluorescein release experiments performed at Hanford, WA between 1967 and 1973. Pasquill-Gifford (PG) and Brookhaven National Laboratory (BNL) dispersion parameterization schemes were used, and atmospheric stability classes were determined from temperature gradient (Δr/ΔZ) and wind direction standard deviation (σθ) measurements. Comparisons between calculated values and experimental results were made as functions of changes in dispersion parameterization, stability class determination method, deposition velocity v d , and gravitational fall velocity v g . Values of the correlation coefficient ranged from significant (p < 0.01) to insignificant (p > 0.05). Best agreement between comparisons was found with v g = v d = 1.8 cm/s and Pasquill-Gifford parameters based on stability classes determined using wind direction standard deviation, σθ, rather than temperature gradient, ΔTZ. The best agreement gave a Pearson’s correlation coefficient r = 0.41 (significant at p < 0.01).  相似文献   

17.
Using an aircraft, preliminary investigations have been carried out into the chemical composition of clouds. The emphasis was on testing the sampling devices under several conditions. The results clearly show a considerable spatial distribution in pollutant concentrations that must have been caused by the contribution of many plume sources and the inhomogeneity of cloudwater density as well. Rather high concentrations of H3O+. Cl, NO3, SO2−4, NH+4 and H2O2 were measured. Except for H2O2, these components are scavenged very fast during travel over short distances over source regions. Moreover H3O+, NO3 and SO2−4 are likely to be formed by chemical reactions of gaseous precursors with oxidants in the liquid phase. This may explain the marked decrease in high background H2O2 concentrations (ppm range). Relatively high (molar) NO3 /SO2−4 ratios mainly varying from 1 to 2 were measured over The Netherlands near source areas. The lower values found during a flight over southern Scandinavia are in agreement with data from literature. Occasionally the chloride concentration was found to be much higher than could be expected from rain network data.  相似文献   

18.
Atmospheric dry deposition is an important process for the introduction of aerosols and pollutants to aquatic environments. The objective of this paper is to assess, for the first time, the influence that the aquatic surface microlayer plays as a modifying factor of the magnitude of dry aerosol deposition fluxes. The occurrence of a low surface tension (ST) or a hydrophobic surface microlayer has been generated by spiking milli-Q water or pre-filtered seawater with a surfactant or octanol, respectively. The results show that fine mode (<2.7 μm) aerosol phase PAHs deposit with fluxes 2–3 fold higher when there is a low ST aquatic surface due to enhanced sequestration of colliding particles at the surface. Conversely, for PAHs bound to coarse mode aerosols (>2.7 μm), even though there is an enhanced deposition due to the surface microlayer for some sampling periods, the effect is not observed consistently. This is due to the importance of gravitational settling for large aerosols, rendering a lower influence of the aquatic surface on dry deposition fluxes. ST (mN m−1) is identified as one of the key factor driving the magnitude of PAH dry deposition fluxes (ng m−2 d−1) by its influence on PAH concentrations in deposited aerosols and deposition velocities (vd, cm s−1). Indeed, vd values are a function of ST as obtained by least square fitting and given by Ln(vd)=−1.77 Ln(ST)+5.74 (r2=0.95) under low wind speed (average 4 m s−1) conditions.  相似文献   

19.
The kinetics of two structurally similar unsaturated alcohols, 3-butene-2-ol and 2-methyl-3-butene-2-ol (MBO232), with Cl atoms have been investigated for the first time, as a function of temperature using a relative method. As far as we know, the present work also provides the first value for 3-buten-2-ol. The coefficient at room temperature was also obtained for 2-propene-1-ol (allyl alcohol). The reactions were investigated using a 400 L Teflon reaction chamber coupled with gas chromatograph-coupled with flame-ionization detection (GC-FID) detection. The experiments were performed at atmospheric pressure and at temperatures between 256 and 298 K in air or nitrogen as the bath gas. The obtained kinetic data were used to derive the Arrhenius expressions, kMBO232=(2.83±2.50)×10−14 exp (2670±249)/T, k3-buten-2-ol=(0.65±1.60)×10−15 exp (3656±695)/T (in units of cm3 molecule−1 s−1). Finally, results and atmospheric implications are discussed and compared with the reactivity with OH and NO3 radicals.  相似文献   

20.
Arctic air chemistry observations made in Canada between 1979 and 1984 are discussed. The weekly average concentration of 25 aerosol constituents has been measured routinely at three locations. Anthropogenic pollution typified by SO42− and V has a persistent seasonal cycle. SO42− concentrations are similar at all three locations, although they tend to be somewhat higher at Alert than at Mould Bay and Igloolik. The seasonal variation of an aerosol constituent depends on its source. There are four distinctive seasonal variations for:
  • 1.(i) anthropogenic constituents Cr, Cu, Mn, Ni, Pb, Sr, V, Zn, H+, NH4+, SO42−, NO3,
  • 2.(ii) halogens (excepting Cl) Br, I, F,
  • 3.(iii) sea salt elements Na, Mg, Cl and
  • 4.(iv) soil constituents Al, Ba, Ca, Fe and Ti. In the Arctic winter, the mean concentrations of anthropogenic aerosol constituents, except SO42−, are 2–4 times lower than annual mean concentrations in southern Sweden near a major source region. SO42− concentrations are only 30% lower mainly because of production from SO2. Light scattering (bscat) and SO42− observations indicate that the SO42− fraction of the fine particle mass fluctuates between 3 and 65% during the polluted winter months. Daily mean bsact, at Mould Bay that exceeds 50 × 10−6m−1 is associated with air originating from the northwest. The soluble major ion composition of aerosols during winter varies markedly with particle size. H+, NH4+ and SO42− dominate submicrometre particles while sea-salt ions Mg2+, Na+ and Cl predominate in supermicrometre particles. Winter SO2 concentrations at Mould Bay and Igloolik ranged from 0.2 to 1.5 ppb
  • 5.(v). The fraction of airborne S as SO2 ranged from 20 to 90% and peaked in late December-early January. The concentration of total NO3 (0.025–0.090 ppb(v)) is much lower than that of SO42− (0.3–1.2 ppb (v)).
  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