PUBLICATIONS

2020
Yuan W, Huang RJ, Yang L, Guo J, Chen Z, Duan J, Wang T, Ni H, Han Y, Li Y, et al. Characterization of the light-absorbing properties, chromophore composition and sources of brown carbon aerosol in Xi'an, northwestern China. Atmospheric Chemistry and Physics. 2020;20:5129-5144.
Duan J, Huang RJ, Li Y, Chen Q, Zheng Y, Chen Y, Lin C, Ni H, Wang M, Ovadnevaite J, et al. Summertime and wintertime atmospheric processes of secondary aerosol in Beijing. Atmospheric Chemistry and Physics. 2020;20:3793-3807.
2019
Wang M, Huang R-J, Cao J, Dai W, Zhou J, Lin C, Ni H, Duan J, Wang T, Chen Y, et al. Determination of n-alkanes, polycyclic aromatic hydrocarbons and hopanes in atmospheric aerosol: evaluation and comparison of thermal desorption GC-MS and solvent extraction GC-MS approaches. Atmospheric Measurement Techniques. 2019;12:4779-4789.
Huang G, Liu Y, Shao M, Li Y, Chen Q, Zheng Y, Wu Z, Liu Y, Wu Y, Hu M, et al. Potentially Important Contribution of Gas-Phase Oxidation of Naphthalene and Methylnaphthalene to Secondary Organic Aerosol during Haze Events in Beijing. Environmental Science & Technology. 2019;53:1235-1244.
Huang R-J, Wang Y, Cao J, Lin C, Duan J, Chen Q, Li Y, Gu Y, Yan J, Xu W, et al. Primary emissions versus secondary formation of fine particulate matter in the most polluted city (Shijiazhuang) in North China. Atmospheric Chemistry and Physics. 2019;19:2283-2298.
Duan J, Huang R-J, Lin C, Dai W, Wang M, Gu Y, Wang Y, Zhong H, Zheng Y, Ni H, et al. Distinctions in source regions and formation mechanisms of secondary aerosol in Beijing from summer to winter. Atmospheric Chemistry and Physics. 2019;19:10319-10334.
Sun JJ, Liang MJ, Shi ZH, Shen FZ, Li JY, Huang L, Ge XL, Chen Q, Sun YL, Zhang YL, et al. Investigating the PM2.5 mass concentration growth processes during 2013-2016 in Beijing and Shanghai. Chemosphere. 2019;221:452-463.Abstract
The North China Plain and the Yangtze River Delta are the two of the most heavily polluted regions in China. Observational studies revealed that 'explosive' PM2.5 mass concentration growths frequently occurred in the two regions. This study analyzed all the PM2.5 mass concentration growth processes from clean condition (i.e., <35 mu g m(-3)) to heavy pollution condition (i.e., >150 mu g m(-3)) in Beijing (BJ) and Shanghai (SH), two representative cities of the two regions, using hourly monitored PM2.5 concentrations during 2013-2016. 173 and 76 growth processes were identified in BJ and SH, respectively. PM2.5 rising rates (PMRR) and dynamic growth durations were calculated to illustrate the characteristics of the growth processes. Hourly particulate chemical composition data and meteorological data in BJ and SH were further analyzed. The 4-year averaged PMRR of PM2.5 total mass were similarly of 7.11 +/- 9.82 mu g m(-3) h(-1) in BJ and 6.71 +/- 6.89 mu g m(-3) h(-1) in SH. A decreasing trend was found for the PM2.5 growth processes in two cities from 2013 to 2016, reflecting the effectiveness of emission controls implemented in the past years. The contributions of particulate components to the PM2.5 total mass growth were different in BJ and SH. Average PMRR value of PM1 organic aerosols (OA), SO24-, NO3-, and NH4+ in BJ was 1.90, 0.95, 0.82, and 0.53 mu g m(-3) h(-1), respectively. Average PMRR of PM2.5 OA, SO42-, NO3-, and NH4+ in SH was 1.70, 1.18, 1.99 and 1.14 mu g m(-3) h(-1), respectively. Based on the contributions of different components, the PM2.5 mass concentration growth processes in BJ and SH were proposed to be classified into 'other components-dominant growth processes', 'all components-contributing growth processes', 'one or more explosive secondary components-dominant growth processes', and 'mixed-factor growth processes'. Potential source contribution function analysis and the meteorological condition analysis showed that source origins and prevailing wind for the two cities during different categories of growth processes had substantial difference. The important source areas included Hebei and Shandong for BJ, and Jiangsu and Anhui for SH. The dominant wind directions during growth processes were northeast, south and southwest in BJ, and were west to north in SH. The results suggested the contributing components, the prevailing wind conditions, and the formation processes were substantially different in the two cities, despite the similar PMRR of PM2.5 total mass during the growth processes between BJ and SH. Future research is needed to study the detailed formation mechanisms of the different PM2.5 mass concentration growth processes in the two cities. (C) 2019 Elsevier Ltd. All rights reserved.
2018
Huang R-J, Yang L, Cao J, Chen Y, Chen Q, Li Y, Duan J, Zhu C, Dai W, Wang K, et al. Brown Carbon Aerosol in Urban Xi'an, Northwest China: The Composition and Light Absorption Properties. Environmental Science & Technology. 2018;52:6825-6833.
An J, Chen Y, Qu Y, Chen Q, Zhuang B, Zhang P, Wu Q, Xu Q, Cao L, Jiang H, et al. An Online-coupled Unified Air Quality Forecasting Model System. Advance in Earth Sciences. 2018;33:445-454.
Huang R-J, Cao J, Chen Y, Yang L, Shen J, You Q, Wang K, Lin C, Xu W, Gao B, et al. Organosulfates in atmospheric aerosol: synthesis and quantitative analysis of PM2.5 from Xi’an, northwestern China. Atmos. Meas. Tech. 2018;11:3447-3456.Abstract
The sources, formation mechanism and amount of organosulfates (OS) in atmospheric aerosol are not yet well understood, partly due to the lack of authentic standardsfor quantification. In this study, we report an improved robust procedure for the synthesis of organosulfates with different functional groups. Nine authentic organosulfate standards were synthesized and four standards (benzyl sulfate, phenyl sulfate, glycolic acid sulfate, and hydroxyacetone sulfate) were used to quantify their ambient concentrations. The authentic standards and ambient aerosol sampleswere analyzed using an optimized ultra performance liquid chromatography–electrospray ionization-tandem mass spectrometric method (UPLC–ESI–MS/MS). The recovery ranged from 80.4 to 93.2 %, the limits of detection and limits of quantification obtained were 1.1–16.7 and 3.4– 55.6 pgm-3, respectively. Measurements of ambient aerosol particle samples collected in winter 2013/2014 in urban Xi’an, northwestern China, show that glycolic acid sulfate (77.349.2 ngm-3/ is the most abundant species of the identified organosulfates followed by hydroxyacetone sulfate (1.30.5 ngm-3/, phenyl sulfate (0.140.09 ngm-3/,and benzyl sulfate (0.040.01 ngm-3/. Except for hydroxyacetone sulfate, which seems to form mainly from biogenic emissions in this region, the organosulfates quantified during winter in Xi’an show an increasing trend with an increase in the mass concentrations of organic carbon indicating their anthropogenic origin.
Zhou W, Zhao J, Ouyang B, Mehra A, Xu W, Wang Y, Bannan TJ, Worrall SD, Priestley M, Bacak A, et al. Production of N2O5 and ClNO2 in summer in urban Beijing, China. Atmos. Chem. Phys. 2018;18:11581-11597.
Huang R-J, Cheng R, Jing M, Yang L, Li Y, Chen Q, Chen Y, Yan J, Lin C, Wu Y, et al. Source-specific health risk analysis on particulate trace elements: Coal combustion and traffic emission as major contributors in wintertime Beijing. Environ. Sci. Technol. 2018:Availabe online.
2017
Wang H, Lu K, Chen X, Zhu Q, Chen Q, Guo S, Jiang M, Li X, Shang D, Tan Z, et al. High N2O5 concentrations observed in urban Beijing: Implications of a large nitrate formation pathway. Environ. Sci. Technol. Lett. 2017;4(10):416–420.Abstract
The heterogeneous hydrolysis of dinitrogen pentoxide (N2O5) is important to understanding the formation of particulate nitrate (pNO3–). Measurements of N2O5 in the surface layer taken at an urban site in Beijing are presented here. N2O5 was observed with large day-to-day variability. High N2O5 concentrations were determined during pollution episodes with the co-presence of large aerosol loads. The maximum value was 1.3 ppbv (5 s average), associated with an air mass characterized by a high level of O3. N2O5 uptake coefficients were estimated to be in the range of 0.025–0.072 using the steady-state lifetime method. As a consequence, the nocturnal pNO3– formation potential by N2O5 heterogeneous uptake was calculated to be 24–85 μg m–3 per night and, on average, 57 μg m–3 during days with pollution. This was comparable to or even higher than that formed by the partitioning of HNO3. The results highlight that N2O5 heterogeneous hydrolysis is vital in pNO3– formation in Beijing.
Chen Q, Fu T-M, Hu J, Ying Q, Zhang L. Modelling secondary organic aerosols in China. National Science Review [Internet]. 2017:10.1093/nsr/nwx143. Link
2016
Fu PQ, Aggarwal SG, Chen J, Li J, Sun YL, Wang ZF, Chen HS, Liao H, Ding AJ, Umarji GS, et al. Molecular markers of secondary organic aerosol in Mumbai, India. Environmental Science & Technology. 2016;50:4659-4667.Abstract
Biogenic secondary organic aerosols (SOA) are generally considered to be more abundant in summer than in winter. Here, polar organic marker compounds in urban background aerosols from Mumbai were measured using gas chromatography-mass spectrometry. Surprisingly, we found that concentrations of biogenic SOA tracers at Mumbai were several times lower in summer (8-14 June 2006; wet season; n = 14) than in winter (13-18 February 2007; dry season; n = 10). Although samples from less than 10% of the season are extrapolated to the full season, such seasonality may be explained, by the predominance of the southwest summer monsoon, which brings clean marine air masses to Mumbai. While heavy rains are an important contributor to aerosol removal during the monsoon season, meteorological data (relative humidity and T) suggest no heavy rains occurred during our sampling period. However, in winter, high levels of SOA and their day/night differences suggest significant contributions of continental aerosols through long-range transport together with local sources. The winter/summer pattern of SOA loadings was further supported by results from chemical transport models (NAQPMS and GEOS-Chem). Furthermore, our study suggests that monoterpene- and sesquiterpene-derived secondary organic carbon. (SOC) were more significant than those of isoprene- and toluene-SOC at Mumbai.
Liu J, Mauzerall DL, Chen Q, Zhang Q, Song Y, Peng W, Klimont Z, Qiu XH, Zhang SQ, Hu M, et al. Air pollutant emissions from Chinese households: A major and underappreciated ambient pollution source. Proceedings of the National Academy of Sciences of the United States of America. 2016;113:7756-7761.Abstract
As part of the 12th Five-Year Plan, the Chinese government has developed air pollution prevention and control plans for key regions with a focus on the power, transport, and industrial sectors. Here, we investigate the contribution of residential emissions to regional air pollution in highly polluted eastern China during the heating season, and find that dramatic improvements in air quality would also result from reduction in residential emissions. We use the Weather Research and Forecasting model coupled with Chemistry to evaluate potential residential emission controls in Beijing and in the Beijing, Tianjin, and Hebei (BTH) region. In January and February 2010, relative to the base case, eliminating residential emissions in Beijing reduced daily average surface PM2.5 (particulate mater with aerodynamic diameter equal or smaller than 2.5 micrometer) concentrations by 14 +/- 7 mu g.m(-3) (22 +/- 6% of a baseline concentration of 67 +/- 41 mu g.m(-3); mean +/- SD). Eliminating residential emissions in the BTH region reduced concentrations by 28 +/- 19 mu g.m(-3) (40 +/- 9% of 67 +/- 41 mu g.m(-3)), 44 +/- 27 mu g.m(-3) (43 +/- 10% of 99 +/- 54 mu g.m(-3)), and 25 +/- 14 mu g.m(-3) (35 +/- 8% of 70 +/- 35 mu g.m(-3)) in Beijing, Tianjin, and Hebei provinces, respectively. Annually, elimination of residential sources in the BTH region reduced emissions of primary PM2.5 by 32%, compared with 5%, 6%, and 58% achieved by eliminating emissions from the transportation, power, and industry sectors, respectively. We also find air quality in Beijing would benefit substantially from reductions in residential emissions from regional controls in Tianjin and Hebei, indicating the value of policies at the regional level.
2015
Chen Q, Heald CL, Jimenez JL, Canagaratna MR, Zhang Q, He L-Y, Huang X-F, Campuzano-Jost P, Palm BB, Poulain L, et al. Elemental composition of organic aerosol: The gap between ambient and laboratory measurements. Geophysical Research Letters. 2015;42:4182-4189.Abstract
A large data set including surface, aircraft, and laboratory observations of the atomic oxygen-to-carbon (O:C) and hydrogen-to-carbon (H:C) ratios of organic aerosol (OA) is synthesized and corrected using a recently reported method. The whole data set indicates a wide range of OA oxidation and a trajectory in the Van Krevelen diagram, characterized by a slope of -0.6, with variation across campaigns. We show that laboratory OA including both source and aged types explains some of the key differences in OA observed across different environments. However, the laboratory data typically fall below the mean line defined by ambient observations, and little laboratory data extend to the highest O:C ratios commonly observed in remote conditions. OA having both high O:C and high H:C are required to bridge the gaps. Aqueous-phase oxidation may produce such OA, but experiments under realistic ambient conditions are needed to constrain the relative importance of this pathway.
Hu WW, Campuzano-Jost P, Palm BB, Day DA, Ortega AM, Hayes PL, Krechmer JE, Chen Q, Kuwata M, Liu YJ, et al. Characterization of a real-time tracer for isoprene epoxydiols-derived secondary organic aerosol (IEPOX-SOA) from aerosol mass spectrometer measurements. Atmospheric Chemistry and Physics. 2015;15:11807-11833.Abstract
Substantial amounts of secondary organic aerosol (SOA) can be formed from isoprene epoxydiols (IEPOX), which are oxidation products of isoprene mainly under low-NO conditions. Total IEPOX-SOA, which may include SOA formed from other parallel isoprene oxidation pathways, was quantified by applying positive matrix factorization (PMF) to aerosol mass spectrometer (AMS) measurements. The IEPOX-SOA fractions of organic aerosol (OA) in multiple field studies across several continents are summarized here and show consistent patterns with the concentration of gas-phase IEPOX simulated by the GEOS-Chem chemical transport model. During the Southern Oxidant and Aerosol Study (SOAS), 78% of PMF-resolved IEPOX-SOA is accounted by the measured IEPOX-SOA molecular tracers (2-methyltetrols, C5-Triols, and IEPOX-derived organosulfate and its dimers), making it the highest level of molecular identification of an ambient SOA component to our knowledge. An enhanced signal at C5H6O+ (m/z 82) is found in PMF-resolved IEPOX-SOA spectra. To investigate the suitability of this ion as a tracer for IEPOX-SOA, we examine fC(5)H(6)O (fC(5)H(6)O = C5H6O+ / OA) across multiple field, chamber, and source data sets. A background of similar to 1.7 +/- 0.1 parts per thousand (parts per thousand = parts per thousand) is observed in studies strongly influenced by urban, biomass-burning, and other anthropogenic primary organic aerosol (POA). Higher background values of 3.1 +/- 0.6 parts per thousand are found in studies strongly influenced by monoterpene emissions. The average laboratory monoterpene SOA value (5.5 +/- 2.0 parts per thousand) is 4 times lower than the average for IEPOX-SOA (22 +/- 7 parts per thousand), which leaves some room to separate both contributions to OA. Locations strongly influenced by isoprene emissions under low-NO levels had higher fC(5)H(6)O (similar to 6.5 +/- 2.2 parts per thousand on average) than other sites, consistent with the expected IEPOX- SOA formation in those studies. fC(5)H(6)O in IEPOX- SOA is always elevated (12-40 parts per thousand) but varies substantially between locations, which is shown to reflect large variations in its detailed molecular composition. The low fC(5)H(6)O (< 3 parts per thousand) reported in non-IEPOX-derived isoprene-SOA from chamber studies indicates that this tracer ion is specifically enhanced from IEPOX- SOA, and is not a tracer for all SOA from isoprene. We introduce a graphical diagnostic to study the presence and aging of IEPOX- SOA as a triangle plot of f(CO2) vs. fC(5)H(6)O. Finally, we develop a simplified method to estimate ambient IEPOX- SOA mass concentrations, which is shown to perform well compared to the full PMF method. The uncertainty of the tracer method is up to a factor of similar to 2, if the fC(5)H(6)O of the local IEPOX- SOA is not available. When only unit mass-resolution data are available, as with the aerosol chemical speciation monitor (ACSM), all methods may perform less well because of increased interferences from other ions at m/z 82. This study clarifies the strengths and limitations of the different AMS methods for detection of IEPOX- SOA and will enable improved characterization of this OA component.
Canagaratna MR, Jimenez JL, Kroll JH, Chen Q, Kessler SH, Massoli P, Ruiz LH, Fortner E, Williams LR, Wilson KR, et al. Elemental ratio measurements of organic compounds using aerosol mass spectrometry: characterization, improved calibration, and implications. Atmospheric Chemistry and Physics. 2015;15:253-272.Abstract
Elemental compositions of organic aerosol (OA) particles provide useful constraints on OA sources, chemical evolution, and effects. The Aerodyne high-resolution time-of-flight aerosol mass spectrometer (HR-ToF-AMS) is widely used to measure OA elemental composition. This study evaluates AMS measurements of atomic oxygen-to-carbon (O : C), hydrogen-to-carbon (H : C), and organic mass-to-organic carbon (OM : OC) ratios, and of carbon oxidation state ((OS) over bar (C))for a vastly expanded laboratory data set of multifunctional oxidized OA standards. For the expanded standard data set, the method introduced by Aiken et al. (2008), which uses experimentally measured ion intensities at all ions to determine elemental ratios (referred to here as "Aiken-Explicit"), reproduces known O: C and H: C ratio values within 20% (average absolute value of relative errors) and 12%, respectively. The more commonly used method, which uses empirically estimated H2O+ and CO+ ion intensities to avoid gas phase air interferences at these ions (referred to here as "Aiken-Ambient"), reproduces O: C and H: C of multifunctional oxidized species within 28 and 14% of known values. The values from the latter method are systematically biased low, however, with larger biases observed for alcohols and simple diacids. A detailed examination of the H2O+, CO+, and CO2+ fragments in the high-resolution mass spectra of the standard compounds indicates that the Aiken-Ambient method underestimates the CO C and especially H2O+ produced from many oxidized species. Combined AMS-vacuum ultraviolet (VUV) ionization measurements indicate that these ions are produced by dehydration and decarboxylation on the AMS vaporizer (usually operated at 600 degrees C). Thermal decomposition is observed to be efficient at vaporizer temperatures down to 200 degrees C. These results are used together to develop an "Improved-Ambient" elemental analysis method for AMS spectra measured in air. The Improved-Ambient method uses specific ion fragments as markers to correct for molecular functionality-dependent systematic biases and reproduces known O : C (H : C) ratios of individual oxidized standards within 28% (13 %) of the known molecular values. The error in Improved-Ambient O : C (H : C) values is smaller for theoretical standard mixtures of the oxidized organic standards, which are more representative of the complex mix of species present in ambient OA. For ambient OA, the Improved-Ambient method produces O : C (H : C) values that are 27% (11 %) larger than previously published Aiken-Ambient values; a corresponding increase of 9% is observed for OM : OC values. These results imply that ambient OA has a higher relative oxygen content than previously estimated. The (OS) over bar (C) values calculated for ambient OA by the two methods agree well, however (average relative difference of 0.06 (OS) over bar (C) units). This indicates that (OS) over bar (C) is a more robust metric of oxidation than O : C, likely since (OS) over bar (C) is not affected by hydration or dehydration, either in the atmosphere or during analysis.
Chen Q, Farmer DK, Rizzo LV, Pauliquevis T, Kuwata M, Karl TG, Guenther A, Allan JD, Coe H, Andreae MO, et al. Submicron particle mass concentrations and sources in the Amazonian wet season (AMAZE-08). Atmospheric Chemistry and Physics. 2015;15:3687-3701.Abstract
Real-time mass spectra of the non-refractory species in submicron aerosol particles were recorded in a tropical rainforest in the central Amazon Basin during the wet season from February to March 2008, as a part of the Amazonian Aerosol Characterization Experiment (AMAZE-08). Organic material accounted on average for more than 80% of the non-refractory submicron particle mass concentrations during the period of measurements. There was insufficient ammonium to neutralize sulfate. In this acidic, isoprene-rich, HO2-dominant environment, positive-matrix factorization of the time series of particle mass spectra identified four statistical factors to account for the 99% of the variance in the signal intensities of the organic constituents. The first factor was identified as associated with regional and local pollution and labeled "HOA" for its hydrocarbon-like characteristics. A second factor was associated with long-range transport and labeled "OOA-1" for its oxygenated characteristics. A third factor, labeled "OOA-2," was implicated as associated with the reactive uptake of isoprene oxidation products, especially of epoxydiols to acidic haze, fog, or cloud droplets. A fourth factor, labeled "OOA-3," was consistent with an association with the fresh production of secondary organic material (SOM) by the mechanism of gasphase oxidation of biogenic volatile organic precursors followed by gas-to-particle conversion of the oxidation products. The suffixes 1, 2, and 3 on the OOA labels signify ordinal ranking with respect to the extent of oxidation represented by the factor. The process of aqueous-phase oxidation of water-soluble products of gas-phase photochemistry might also have been associated to some extent with the OOA-2 factor. The campaign-average factor loadings had a ratio of 1.4 : 1 for OOA-2 : OOA-3, suggesting the comparable importance of particle-phase compared to gas-phase pathways for the production of SOM during the study period.

Pages