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 78 毫秒
1.
We report here direct observation by differential optical absorption spectroscopy (DOAS) of the formation of ppb levels of gaseous nitrous acid (MONO) from the reaction of ppm levels of nitrogen dioxide (NO2) with water vapor, in an indoor environment. The rate of formation of HONO displayed first order kinetics with respect to NO2 with a rate of (0.25 ±0.04) ppb min−1 per ppm of NO2 present. Assuming a lifetime of l h with respect to both physical and chemical removal processes for HONO, this leads to an estimated steady state concentration of ~ 15 ppb of HONO per ppm of NO2 present. This relatively high level of HONO associated with NO2-air mixtures raises new questions concerning the health implications of elevated NO2 concentrations in indoor environments e.g. HONO is a respirable nitrite known to convert secondary amines in vitro to carcinogenic nitrosamines.  相似文献   

2.
Gaseous nitrogen dioxide (NO2) represents an oxidant that is present in relatively high concentrations in various indoor settings. Remarkably increased NO2 levels up to 1.5 ppm are associated with homes using gas stoves. The heterogeneous reactions of NO2 with adsorbed water on surfaces lead to the generation of nitrous acid (HONO). Here, we present a HONO source induced by heterogeneous reactions of NO2 with selected indoor paint surfaces in the presence of light (300 nm?<?λ?<?400 nm). We demonstrate that the formation of HONO is much more pronounced at elevated relative humidity. In the presence of light (5.5 W m?2), an increase of HONO production rate of up to 8.6?·?109 molecules cm?2 s?1 was observed at [NO2]?=?60 ppb and 50 % relative humidity (RH). At higher light intensity of 10.6 (W m?2), the HONO production rate increased to 2.1?·?1010 molecules cm?2 s?1. A high NO2 to HONO conversion yield of up to 84 % was observed. This result strongly suggests that a light-driven process of indoor HONO production is operational. This work highlights the potential of paint surfaces to generate HONO within indoor environments by light-induced NO2 heterogeneous reactions.  相似文献   

3.
A denuder technique for sampling and analysing nitrous acid at sub ppb levels in air is described. After sampling, the sodium carbonate coated denuder is leached in water, and the NO2 concentration is determined spectrophotometrically. Field tests show that PAN is partly sorbed and hydrolized to nitrite in the sodium carbonate layer. It seems as HNO2 also can be formed by heterogeneous reactions between NO and NO2 at the denuder wall. These sampling artifacts were overcome by sampling with two or three denuders in series. The presence of PAN deteriorates the detection limit, which during optimal conditions is about 0.5 nmole m−3 (0.01 ppb). The method is therefore not recommended for measurements in background air, where HNO2 concentrations normally are low compared to PAN concentrations.  相似文献   

4.
5.
Dinitrogen pentoxide (N2O5), which is present in equilibrium with NO3 radicals and NO2, has been recognized for some time as an intermediate in the NOx chemistry of night-time atmospheres. However, until the advent of long pathlength spectroscopic techniques for the measurement of atmospheric NO3 radical concentrations, no reliable method for estimating N2O5 concentrations has been available. We have calculated maximum night-time N2O5 concentrations from the available experimentally determined concentrations of the NO3 radical and NO2 in the U.S. and Germany, and find that N2O5 concentrations as high as ~ 15 ppb can occur. We have also estimated removal rates for N2O5 and for NO3 radicals during these nights. From data obtained under conditions devoid of point sources of NOx, upper limit estimates of the homogeneous rate constant for the reaction of N2O5 with water vapor are obtained, leading to the conclusion that the homogeneous gas phase rate constant for this reaction is ⩽ 1 × 10−21 cm3 molecule−1 s−1 at 298 K, consistent with recent environmental chamber data.  相似文献   

6.
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   

7.
Simultaneous measurements of nitrous acid (HONO) and nitrogen dioxide (NO2) using a differential optical absorption spectroscopy system, nitrogen oxide (NO) by an in situ chemiluminescence analyser and carbon dioxide (CO2) by a gas chromatographic technique were carried out in the Wuppertal Kiesbergtunnel. At high traffic density HONO concentrations of up to 45 ppbV were observed. However, at low traffic density unexpectedly high HONO concentrations of up to 10 ppbV were measured caused by heterogeneous HONO formation on the tunnel walls. In addition to the tunnel campaigns, emission measurements of HONO, NO2, NO and CO2 from different single vehicles (a truck, a diesel and a gasoline passenger car) were also performed. For the correction of the HONO emission data, the heterogeneous HONO formation on the tunnel walls was quantified by two different approaches (a) in different NO2 emission experiments in the tunnel without traffic and (b) on tunnel wall residue in the laboratory. The HONO concentration corrected for heterogeneous formation on the tunnel walls, in relation to the CO2 concentration can be used to estimate the amount of HONO, which is directly emitted from the vehicle fleet. From the measured data, emission ratios (e.g. HONO/NOx) and emission indices (e.g. mg HONO kg−1 fuel) were calculated. The calculated emission index of 88±18 mg HONO kg−1 fuel allows an estimation of the HONO emission rates from traffic into the atmosphere. Furthermore, the heterogeneous formation of HONO from NO2 on freshly emitted exhaust particles is discussed.  相似文献   

8.
Measurements of ammonia and particulate ammonium were made in the daytime (1200–1500) at a urban site in Yokohama during the 5-year period, 1982–1986. Diurnal NH3 concentrations showed a distinct seasonal trend with a maximum in summer. The diurnal monthly average concentrations were above 10 ppb during the late spring and summer months, while the concentrations during the winter months were between 1 and 5 ppb. The seasonal variation was found to be very similar to that of the average air temperature and showed a periodic pattern over 1 year. A good correlation was observed between diurnal NH3 concentrations and average air temperatures during the 5-year period. The annual mean concentrations were in the range of 6.6–7.6 ppb with only a minor deviation. The diurnal monthly average concentrations of particulate NH4+ were between 1 and 4 μg m−3 and no significant seasonal variations were seen. As a short-term study, simultaneous measurements of NH3, HNO3 and particulate NO3 were made. The diurnal mean concentrations of NH3 and HNO3 were 7.6 and 0.8 ppb, respectively. The concentration of particulate NO3 ranged from 0.3 to 6μg−3. Both HNO3 and particulate NO3 concentrations were relatively low and constant. Thus, NH3 and HNO3 levels did not agree with the concentrations predicted from the NH4NO3 equilibrium constant.  相似文献   

9.
During the 1999 summer field season at Summit, Greenland, we conducted several series of experiments to follow up on our 1998 discovery that NOx is released from the sunlit snowpack. The 1999 experiments included measurements of HONO in addition to NO and NO2, and were designed to confirm, for Greenland snow, that the processes producing reactive nitrogen oxides in the snow are largely photochemical. Long duration experiments (up to 48 h) in a flow-through chamber and in the natural snowpack revealed sun-synchronous diurnal variations of all three reactive nitrogen oxides. In a second set of experiments we alternately shaded or exposed snow (again in the natural snowpack and in the chamber) to ambient sunlight for short periods to reduce any temperature changes during variations in light intensity. All three N oxides increased (decreased) very rapidly when sunlit (shaded). In all experiments NO2 was approximately 3-fold more abundant than NO and HONO (which were at similar levels). Higher concentrations of NO3 in the snow resulted in higher mixing ratios of HONO, NO and NO2 in the snow pore air, consistent with our hypothesis that photolysis of NO3 is the source of the reactive N oxides.  相似文献   

10.
The temporal behavior of HONO and NO2 was investigated at an urban site in Guangzhou city, China, by means of a DOAS system during the Pearl River Delta 2006 intensive campaign from 10 to 24 July 2006. Within the whole measurement period, unexpected high HONO mixing ratios up to 2 ppb were observed even during the day. A nocturnal maximum concentration of about 8.43 ± 0.4 ppb was detected on the night of 24 July 2006. Combining the data simultaneously observed by different instruments, the coupling of HONO–NO2 and the possible formation sources of HONO are discussed. During the measurement period, concentration ratios of HONO to NO2 ranged from (0.03 ± 0.1) to (0.37 ± 0.09), which is significantly higher than previously reported values (0.01–0.1). Surprisingly, in most cases a strong daytime correlation between HONO and NO2 was found, contrary to previous observations in China. Aerosol was found to have a minor impact on HONO formation during the whole measurement period. Using a pseudo steady state approach for interpreting the nocturnal conversion of NO2 to HONO suggests a non-negligible role of the relative humidity for the heterogeneous HONO formation from NO2.  相似文献   

11.
Much rain and strong winds caused by a cold front occurred in Beijing during the period of Sep. 27 to Oct. 4, 2004 and led to sharp drops in maximum and mean concentrations of HONO, HCHO, O3, and NO2, i.e., the maximum concentrations were reduced by 5.9, 21.3, 45.6, and 44.4 ppb, respectively, and the mean concentrations were decreased by 4.0, 5.5, 30.3, and 32.3 ppb, respectively. For daily HOx production rates HONO photolysis was the largest contributor and over 90% contributions were from photolysis of HONO and HCHO. Large number and area percentages of soot aggregate from PM10, and high correlations between concentrations of PM10 and chemical formation of HONO suggested that heterogeneous reactions of NO2 on surfaces of soot aggregate could be a key source of HONO in the heavy traffic areas of Beijing during the night and should be considered in air quality simulations for such areas.  相似文献   

12.
A wet effluent denuder - aerosol collector (WEDD/AC) system coupled to ion chromatography for the measurement of atmospheric HONO, HNO3 and particulate nitrite, nitrate and sulfate is described. Several experiments were performed to outline its performance. The main features are low detection limits and a fast response to concentration changes which enables measurements with high time resolution. In contrast to highly soluble gases, the collection efficiency of less soluble gases is shown to depend on the Henry’s law constant rather than on the uptake kinetics. To improve the collection efficiency for HONO under simultaneous presence of acidifying gases, NaHCO3 was added to the effluent solution. The system was tested in a field campaign in the suburban area of Zürich, Switzerland. Elevated concentrations of nitrous acid up to 3.2 ppb were detected during the measurement campaign. The diurnal variation of the HONO to NO2 ratio clearly points to a fast and persistent process producing HONO in the atmosphere. The correlation with NOx and black carbon suggests a heterogeneous formation of HONO, and is consistent with a reaction on soot aerosol particle surfaces postulated from previous laboratory results.  相似文献   

13.
Nitrous acid (HONO), nitric acid (HNO3), and organic aerosol were measured simultaneously atop an 18-story tower in Houston, TX during August and September of 2006. HONO and HNO3 were measured using a mist chamber/ion chromatographic technique, and aerosol size and chemical composition were determined using an Aerodyne quadrupole aerosol mass spectrometer. Observations indicate the potential for a new HONO formation pathway: heterogeneous conversion of HNO3 on the surface of primary organic aerosol (POA). Significant HONO production was observed, with an average of 0.97 ppbv event?1 and a maximum increase of 2.2 ppb in 4 h. Nine identified events showed clear HNO3 depletion and well-correlated increases in both HONO concentration and POA-dominated aerosol surface area (SA). Linear regression analysis results in correlation coefficients (r2) of 0.82 for HONO/SA and 0.92 for HONO/HNO3. After correction for established HONO formation pathways, molar increases in excess HONO (HONOexcess) and decreases in HNO3 were nearly balanced, with an average HONOexcess/HNO3 value of 0.97. Deviations from this mole balance indicate that the residual HNO3 formed aerosol-phase nitrate. Aerosol mass spectral analysis suggests that the composition of POA could influence HONO production. Several previously identified aerosol-phase PAH compounds were enriched during events, suggesting their potential importance for heterogeneous HONO formation.  相似文献   

14.
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   

15.
Snow chamber and snow-pile experiments performed during the ‘Alert 2000’ campaign show significant release of NO, NO2, and HONO in steady ratios under the influence of irradiation. Both light and a minimal degree of heating are required to produce this effect. We suggest diffusion and re-distribution of NO3 in the form of HNO3 as an important step in the mechanism of active nitrogen release from the snowpack.  相似文献   

16.
Simultaneous measurements of gaseous species and fine-mode, particulate inorganic components were performed at the University of Seoul, Seoul in Korea. In the simultaneous measurements, a certain level of nitrous acid (HONO) was observed in the gas-phase, indicating possible heterogeneous HONO production on the surface of the ambient aerosols. On the other hand, high particulate nitrite (NO2?) concentrations of 1.41(±2.26) μg/m3 were also measured, which sometimes reached 18.54 μg/m3. In contrast, low HONO-to-NO2 ratios of 0.007(±0.006) were observed in Seoul. This indicates that a significant fraction of HONO is dissolved in atmospheric aerosols. Around the Seoul site, sufficient alkalinity may have been provided to the atmospheric aerosols from the excessive presence of NH3 in the gas-phase. Due to the alkaline particulate conditions (defined in this study as a particle pH >~3.29), the HONO molecules produced at the surface of the atmospheric aerosols appeared to have been converted into particulate nitrite, thereby preventing their further participation in the atmospheric O3/NOy/HOx photochemical cycles. It was estimated that a minimum average of 65% of HONO was captured by alkaline, anthropogenic, urban particles in the Seoul measurements.  相似文献   

17.
A chamber placed in a constant temperature freezing room was used to study the surface resistance during deposition of HNO3 to a snow surface. The resistance decreased with increasing temperature from larger than 5 s mm−1 at − 18°C to about l s mm−1 at −3°C. Measurements of gaseous and particulate nitrate concentrations during winter at a rural site in south central Sweden gave concentrations in the range of 0.4–5 μg HNO3 m−1 and 0.3–3 μg NO3 m−3 with a mean value of 1.3 μg HNO3 m−3 and 0.7 μg NO3 m−3, respectively. The results indicate that for periods with temperatures below − 2°C estimated dry deposition of HNO3 to snow is at most 4 % of measured wet deposition of nitrate in the area.  相似文献   

18.
Nitrous acid (HONO) and formaldehyde (HCHO) are important precursors for radicals and are believed to favor ozone formation significantly. Traffic emission data for both compounds are scarce and mostly outdated. A better knowledge of today's HCHO and HONO emissions related to traffic is needed to refine air quality models. Here the authors report results from continuous ambient air measurements taken at a highway junction in Houston, Texas, from July 15 to October 15, 2009. The observational data were compared with emission estimates from currently available mobile emission models (MOBILE6; MOVES [MOtor Vehicle Emission Simulator]). Observations indicated a molar carbon monoxide (CO) versus nitrogen oxides (NOx) ratio of 6.01 ± 0.15 (r 2 = 0.91), which is in agreement with other field studies. Both MOBILE6 and MOVES overestimate this emission ratio by 92% and 24%, respectively. For HCHO/CO, an overall slope of 3.14 ± 0.14 g HCHO/kg CO was observed. Whereas MOBILE6 largely underestimates this ratio by 77%, MOVES calculates somewhat higher HCHO/CO ratios (1.87) than MOBILE6, but is still significantly lower than the observed ratio. MOVES shows high HCHO/CO ratios during the early morning hours due to heavy-duty diesel off-network emissions. The differences of the modeled CO/NOx and HCHO/CO ratios are largely due to higher NOx and HCHO emissions in MOVES (30% and 57%, respectively, increased from MOBILE6 for 2009), as CO emissions were about the same in both models. The observed HONO/NOx emission ratio is around 0.017 ± 0.0009 kg HONO/kg NOx which is twice as high as in MOVES. The observed NO2/NOx emission ratio is around 0.16 ± 0.01 kg NO2/kg NOx, which is a bit more than 50% higher than in MOVES. MOVES overestimates the CO/CO2 emission ratio by a factor of 3 compared with the observations, which is 0.0033 ± 0.0002 kg CO/kg CO2. This as well as CO/NOx overestimation is coming from light-duty gasoline vehicles.
Implications: Nitrous acid (HONO) and formaldehyde (HCHO) are important precursors for radicals that ultimately contribute to ozone formation. There still exist uncertainties in emission sources of HONO and HCHO and thus regional air quality modeling still tend to underestimate concentrations of free radicals in the atmosphere. This paper demonstrates that the latest U.S. Environmental Protection Agency (EPA) traffic emission model MOVES still shows significant deviations from observed emission ratios, in particular underestimation of HCHO/CO and HONO/NOx ratios. Improving the performance of MOVES may improve regional air quality modeling.  相似文献   

19.
We have identified and quantified 2-nitrofluoranthene (2-NO2-FL) and 2-nitropyrene (2-NO2-PY), both strong, direct mutagens in paniculate organic matter (POM) samples collected in polluted ambient air in southern California. Samples were collected during four consecutive 6-h periods on 18–19 September 1984, during which the ambient concentrations of gaseous co-pollutants were characterized by long pathlength spectroscopic techniques and conventional analyzers. Concentrations ranged up to 0.3 ng m−3 for 2-NO2-FL and 0.02 ng m−3 for 2-NO2-PY. The 2-nitro isomers of fluoranthene and pyrene have not been observed in direct emissions of POM (e.g. diesel exhaust and wood smoke). However, we recently observed these isomers from laboratory reactions involving N2O5 or OH radicals and the parent compounds. Thus the identification of 2-NO2-FL and 2-NO2-PY in ambient POM suggests that chemical transformations of fluoranthene and pyrene may take place in polluted atmospheres.  相似文献   

20.
Size-resolved fog drop chemical composition measurements were obtained during a radiation fog campaign near Davis, California in December 1998/January 1999 (reported in Reilly et al., Atmos. Environ. 35(33) (2001) 5717; Moore et al., Atmos. Environ. this issue). Here we explore how knowledge of this size-dependent drop composition—particularly from the newly developed Colorado State University 5-Stage cloud water collector—helps to explain additional observations in the fog environment. Size-resolved aerosol measurements before and after fog events indicate relative depletion of large (>2 μm in diameter) particles during fog accompanied by a relative increase in smaller aerosol particle concentrations. Fog equivalent air concentrations suggest that entrainment of additional particles and in-fog sedimentation contributed to observed changes in the aerosol size distribution. Calculated deposition velocities indicate that sedimentation was an important atmospheric removal mechanism for some species. For example, nitrite typically has a larger net deposition velocity than water and its mass is found preferentially in the largest drops most likely to sediment rapidly. Gas–liquid equilibria in fog for NO3/HNO3, NH4+/NH3, and NO2/HONO were examined. While these systems appear to be close to equilibrium or relative equilibrium during many time periods, divergences are observed, particularly for low liquid water content (<0.1 g m−3) fogs and in different drop sizes. Knowledge of the drop size-dependent composition provided additional data useful to the interpretation of these deviations. The results suggest that data from multi-stage cloud water collectors are useful to understanding fog processes as many depend upon drop size.  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