Advanced Search
Article Contents

Profiles and Source Apportionment of Nonmethane Volatile Organic Compounds in Winter and Summer in Xi’an, China, based on the Hybrid Environmental Receptor Model

Fund Project:

This research was supported by the Natural Science Foundation of China (Grant No. 41907188), Natural Science Foundation of Shaanxi Province, China (Grant No. 2019JQ-386), and the China Postdoctoral Science Foundation (Grant No. 2019M653658)


doi: 10.1007/s00376-020-0153-0

  • Summer and winter campaigns for the chemical compositions and sources of nonmethane hydrocarbons (NMHCs) and oxygenated volatile organic compounds (OVOCs) were conducted in Xi’an. Data from 57 photochemical assessment monitoring stations for NMHCs and 20 OVOC species were analyzed. Significant seasonal differences were noted for total VOC (TVOC, NMHCs and OVOCs) concentrations and compositions. The campaign-average TVOC concentrations in winter (85.3 ± 60.6 ppbv) were almost twice those in summer (47.2 ± 31.6 ppbv). Alkanes and OVOCs were the most abundant category in winter and summer, respectively. NMHCs, but not OVOCs, had significantly higher levels on weekends than on weekdays. Total ozone formation potential was higher in summer than in winter (by 50%) because of the high concentrations of alkenes (particularly isoprene), high temperature, and high solar radiation levels in summer. The Hybrid Environmental Receptor Model (HERM) was used to conduct source apportionment for atmospheric TVOCs in winter and summer, with excellent accuracy. HERM demonstrated its suitability in a situation where only partial source profile data were available. The HERM results indicated significantly different seasonal source contributions to TVOCs in Xi’an. In particular, coal and biomass burning had contributions greater than half in winter (53.4%), whereas traffic sources were prevalent in summer (53.1%). This study’s results highlight the need for targeted and adjustable VOC control measures that account for seasonal differences in Xi’an; such measures should target not only the severe problem with VOC pollution but also the problem of consequent secondary pollution (e.g., from ozone and secondary organic aerosols).
  • 加载中
  • Figure 1.  Location of the sampling site and surrounding road distribution.

    Figure 2.  Time series of the concentrations of PAMS NMHCs and OVOCs in the winter and summer campaign (M, A, E and N denote morning, afternoon, evening and night, respectively; the date format is as YY/MM/DD)(data absence on 18/7/23 and 18/7/30 were due to instrument failure).

    Figure 3.  Campaign-average TVOCs and proportions by category.

    Figure 4.  Time series of OFPs by PAMS NMHCs and OVOCs in winter and summer (M, A, E and N denote morning, afternoon, evening and night, respectively; the date format is as YY/MM/DD) (data absence on 18/7/23 and 18/7/30 were due to instrument failure).

    Figure 5.  Time series of SOAPs by PAMS NMHCs and OVOCs in winter and summer (M, A, E and N denote morning, afternoon, evening and night, respectively; the date format is as YY/MM/DD)(data absence on 18/7/23 were due to instrument failure).

    Figure 6.  Hourly averaged top 10 species in (a) winter and (b) summer. The box plots indicate the campaign-averaged VOC concentrations and the minimum, 25th, 50th, 75th, and maximum percentiles. A cumulative frequency curve is fitted to the top 10 species

    Figure 6.  (Continued)

    Figure 7.  Source profiles for HERM inputs and outputs

    Figure 7.  (Continued)

    Figure 8.  Source apportionment results for TVOCs by HERM in winter and summer.

    Table 1.  Average concentrations (Avg) and standard deviations (SD) of OVOCs and NMHCs over weekdays and weekends (ppbv).

    SeasonCategoryWeekdaysWeekends
    AvgSDAvgSD
    SummerOVOCs18.96.219.94.2
    NMHCs24.7a10.934.3b23.9
    WinterOVOCs17.07.917.68.2
    NMHCs56.8c29.673.2d33.8
    Notes: Na, N=8; Nb, N=4; Nc, N=10; Nd, N=4.
    DownLoad: CSV
  • Abeleira, A., I. B. Pollack, B. Sive, Y. Zhou, E. V. Fischer, and D. K. Farmer, 2017: Source characterization of volatile organic compounds in the Colorado Northern Front Range Metropolitan Area during spring and summer 2015. J. Geophys. Res., 122, 3595−3613, https://doi.org/10.1002/2016JD026227.
    An, J. L., B. Zhu, H. L. Wang, Y. Y. Li, X. Lin, and H. Yang, 2014: Characteristics and source apportionment of VOCs measured in an industrial area of Nanjing, Yangtze River Delta, China. Atmos. Environ., 97, 206−214, https://doi.org/10.1016/j.atmosenv.2014.08.021.
    An, J. L., J. X. Wang, Y. X. Zhang, and B. Zhu, 2017: Source apportionment of volatile organic compounds in an urban environment at the Yangtze River Delta, China. Archives of Environmental Contamination and Toxicology, 72, 335−348, https://doi.org/10.1007/s00244-017-0371-3.
    Assan, S., F. R. Vogel, V. Gros, A. Baudic, J. Staufer, and P. Ciais, 2018: Can we separate industrial CH4 emission sources from atmospheric observations?--A test case for carbon isotopes, PMF and enhanced APCA Atmos. Environ., 187, 317−327, https://doi.org/10.1016/j.atmosenv.2018.05.004.
    Bari, M. A., and W. B. Kindzierski, 2018: Ambient volatile organic compounds (VOCs) in Calgary, Alberta: Sources and screening health risk assessment. Science of the Total Environment, 631−632, 627−640, https://doi.org/10.1016/j.scitotenv.2018.03.023.
    Borbon, A., N. Locoge, M. Veillerot, J. C. Galloo, and R. Guillermo, 2002: Characterisation of NMHCs in a French urban atmosphere: Overview of the main sources. Science of the Total Environment, 292, 177−191, https://doi.org/10.1016/S0048-9697(01)01106-8.
    Carter, W. P. L., 2010: Updated maximum incremental reactivity scale and hydrocarbon bin reactivities for regulatory applications. California Air Resources Board Contract 07-339.
    Carter, W. P. L., G. Heo, D. R. Cocker Ⅲ, and S. Nakao, 2012: SOA formation: Chamber study and model development. California Air Resources Board Contract No. 08-326.
    Chen, C. H., Y. C. Chuang, C. C. Hsieh, and C. S. Lee, 2019: VOC characteristics and source apportionment at a PAMS site near an industrial complex in central Taiwan. Atmospheric Pollution Research, 10, 1060−1074, https://doi.org/10.1016/j.apr.2019.01.014.
    Chen, L. W. A., and J. J. Cao, 2018: PM2.5 source apportionment using a hybrid environmental receptor model. Environmental Science & Technology, 52, 6357−6369, https://doi.org/10.1021/acs.est.8b00131.
    Cheng, L. J., S. Wang, Z. Y. Gong, H. Li, Q. Yang, and Y. Wang, 2018: Regionalization based on spatial and seasonal variation in ground-level ozone concentrations across China. Journal of Environmental Sciences, 67, 179−190, https://doi.org/10.1016/j.jes.2017.08.011.
    Dai, W. T., H. Zhong, L. Li, J. J. Cao, Y. Huang, M. Shen, L. Q. Wang, J. G. Dong, X. X. Tie, S. S. H. Ho, K. F. Ho, 2018: Characterization and health risk assessment of airborne pollutants in commercial restaurants in northwestern China: Under a low ventilation condition in wintertime. Science of The Total Environment, 633, 30−316, https://doi.org/10.1016/j.scitotenv.2018.03.124.
    Derwent, R. G., M. E. Jenkin, S. R. Utembe, D. E. Shallcross, T. P. Murrells, and N. R. Passant, 2010: Secondary organic aerosol formation from a large number of reactive man-made organic compounds. Science of the Total Environment, 408, 3374−3381, https://doi.org/10.1016/j.scitotenv.2010.04.013.
    Duan, J. C., J. H. Tan, L. Yang, S. Wu, and J. M. Hao, 2008: Concentration, sources and ozone formation potential of volatile organic compounds (VOCs) during ozone episode in Beijing. Atmospheric Research, 88, 25−35, https://doi.org/10.1016/j.atmosres.2007.09.004.
    Durmusoglu, E., F. Taspinar, and A. Karademir, 2010: Health risk assessment of BTEX emissions in the landfill environment. Journal of Hazardous Materials, 176, 870−877, https://doi.org/10.1016/j.jhazmat.2009.11.117.
    Feng, J. J., S. Gao, Q. Y. Fu, X. J. Chen, X. L. Chen, D. M. Han, and J. P. Cheng, 2019: Indirect source apportionment of methyl mercaptan using CMB and PMF models: A case study near a refining and petrochemical plant. Environmental Science and Pollution Research, 26, 24 305−24 312, https://doi.org/10.1007/s11356-019-05728-4.
    Feng, T., and Coauthors, 2016: Summertime ozone formation in Xi'an and surrounding areas, China. Atmospheric Chemistry and Physics, 16, 4323−4342, https://doi.org/10.5194/acp-16-4323-2016.
    Filella, I., and J. Peñuelas, 2006a: Daily, weekly and seasonal relationships among VOCs, NOxx and O3 in a semi-urban area near Barcelona. Journal of Atmospheric Chemistry, 54, 189−201, https://doi.org/10.1007/s10874-006-9032-z.
    Filella, I., and J. Peñuelas, 2006b: Daily, weekly, and seasonal time courses of VOC concentrations in a semi-urban area near Barcelona. Atmos. Environ., 40, 7752−7769, https://doi.org/10.1016/j.atmosenv.2006.08.002.
    Gu, Y. Y., and Coauthors, 2019: Emission characteristics of 99 NMVOCs in different seasonal days and the relationship with air quality parameters in Beijing, China. Ecotoxicology and Environmental Safety, 169, 797−806, https://doi.org/10.1016/j.ecoenv.2018.11.091.
    Heo, J., B. de Foy, M. R. Olson, P. Pakbin, C. Sioutas, and J. J. Schauer, 2015: Impact of regional transport on the anthropogenic and biogenic secondary organic aerosols in the Los Angeles Basin. Atmos. Environ., 103, 171−179, https://doi.org/10.1016/j.atmosenv.2014.12.041.
    Hernández-Paniagua, I., and Coauthors, 2018: Increasing weekend effect in ground-level O3 in metropolitan areas of Mexico during 1988−2016. Sustainability, 10, 3330, https://doi.org/10.3390/su10093330.
    Ho, K. F., S. C. Lee, H. Guo, and W. Y. Tsai, 2004: Seasonal and diurnal variations of volatile organic compounds (VOCs) in the atmosphere of Hong Kong. Science of the Total Environment, 322, 155−166, https://doi.org/10.1016/j.scitotenv.2003.10.004.
    Ho, K. F., and Coauthors, 2009: Vehicular emission of volatile organic compounds (VOCs) from a tunnel study in Hong Kong. Atmospheric Chemistry and Physics, 9, 7491−7504, https://doi.org/10.5194/acp-9-7491-2009.
    Hopke, P. K., 2016: Review of receptor modeling methods for source apportionment. Journal of the Air & Waste Management Association, 66, 237−259, https://doi.org/10.1080/10962247.2016.1140693.
    Hu, D., Q. J. Bian, T. W. Y. Li, A. K. H. Lau, and J. Z. Yu, 2008: Contributions of isoprene, monoterpenes, β-caryophyllene, and toluene to secondary organic aerosols in Hong Kong during the summer of 2006. J. Geophys. Res., 113, D22206, https://doi.org/10.1029/2008JD010437.
    Huang, R. J., and Coauthors, 2014: High secondary aerosol contribution to particulate pollution during haze events in China. Nature, 514, 218−222, https://doi.org/10.1038/nature13774.
    Huang, Y. S., and C. C. Hsieh, 2019: Ambient volatile organic compound presence in the highly urbanized city: Source apportionment and emission position. Atmos. Environ., 206, 45−59, https://doi.org/10.1016/j.atmosenv.2019.02.046.
    Hwa, M. Y., C. C. Hsieh, T. C. Wu, and L. F. W. Chang, 2002: Real-world vehicle emissions and VOCs profile in the Taipei tunnel located at Taiwan Taipei area. Atmos. Environ., 36, 1993−2002, https://doi.org/10.1016/S1352-2310(02)00148-6.
    Johnson, D., M. E. Jenkin, K. Wirtz, and M. Martin-Reviejo, 2004: Simulating the formation of secondary organic aerosol from the photooxidation of Toluene. Environmental Chemistry, 1, 150−165, https://doi.org/10.1071/EN04069.
    Kelly, J. M., R. M. Doherty, F. M. O'Connor, and G. W. Mann, 2018: The impact of biogenic, anthropogenic, and biomass burning volatile organic compound emissions on regional and seasonal variations in secondary organic aerosol. Atmospheric Chemistry and Physics, 18, 7393−7422, https://doi.org/10.5194/acp-18-7393-2018.
    Kim, E., and P. K. Hopke, 2004: Improving source identification of fine particles in a rural northeastern U.S. area utilizing temperature-resolved carbon fractions. J. Geophys. Res., 109, D09204, https://doi.org/10.1029/2003JD004199.
    Kleindienst, T. E., M. Jaoui, M. Lewandowski, J. H. Offenberg, C. W. Lewis, P. V. Bhave, and E. O. Edney, 2007: Estimates of the contributions of biogenic and anthropogenic hydrocarbons to secondary organic aerosol at a southeastern US location. Atmos. Environ., 41, 8288−8300, https://doi.org/10.1016/j.atmosenv.2007.06.045.
    Lee, S., and A. G. Russell, 2007: Estimating uncertainties and uncertainty contributors of CMB PM2.5 source apportionment results. Atmos. Environ., 41, 9616−9624, https://doi.org/10.1016/j.atmosenv.2007.08.022.
    Li, B. W., and Coauthors, 2017a: Characterizations of volatile organic compounds (VOCs) from vehicular emissions at roadside environment: The first comprehensive study in Northwestern China. Atmos. Environ., 161, 1−12, https://doi.org/10.1016/j.atmosenv.2017.04.029.
    Li, G. H., and Coauthors, 2017b: Widespread and persistent ozone pollution in eastern China during the non-winter season of 2015: Observations and source attributions. Atmospheric Chemistry and Physics, 17, 2759−2774, https://doi.org/10.5194/acp-17-2759-2017.
    Li, N., and Coauthors, 2018: Impacts of biogenic and anthropogenic emissions on summertime ozone formation in the Guanzhong Basin, China. Atmospheric Chemistry and Physics, 18, 7489−7507, https://doi.org/10.5194/acp-18-7489-2018.
    Liu, C. T., C. L. Zhang, Y. J. Mu, J. F. Liu, and Y. Y. Zhang, 2017: Emission of volatile organic compounds from domestic coal stove with the actual alternation of flaming and smoldering combustion processes. Environmental Pollution, 221, 385−391, https://doi.org/10.1016/j.envpol.2016.11.089.
    Liu, W. T., H. C. Hsieh, S. P. Chen, J. S. Chang, N. H. Lin, C. C. Chang, and J. L. Wang, 2012: Diagnosis of air quality through observation and modeling of volatile organic compounds (VOCs) as pollution tracers. Atmos. Environ., 55, 56−63, https://doi.org/10.1016/j.atmosenv.2012.03.017.
    Liu, X. H., and Coauthors, 2010: Understanding of regional air pollution over China using CMAQ, part I performance evaluation and seasonal variation. Atmos. Environ., 44, 2415−2426, https://doi.org/10.1016/j.atmosenv.2010.03.035.
    Liu, Y., M. Shao, L. L. Fu, S. H. Lu, L. M. Zeng, and D. G. Tang, 2008: Source profiles of volatile organic compounds (VOCs) measured in China: Part I. Atmos. Environ., 42, 6247−6260, https://doi.org/10.1016/j.atmosenv.2008.01.070.
    Liu, Y. H., and Coauthors, 2019: Characteristics and sources of volatile organic compounds (VOCs) in Shanghai during summer: Implications of regional transport. Atmos. Environ., 215, 116902, https://doi.org/10.1016/j.atmosenv.2019.116902.
    Louie, P. K. K., and Coauthors, 2013: VOCs and OVOCs distribution and control policy implications in Pearl River Delta region, China. Atmos. Environ., 76, 125−135, https://doi.org/10.1016/j.atmosenv.2012.08.058.
    Lyu, X. P., N. Chen, H. Guo, W. H. Zhang, N. Wang, Y. Wang, and M. Liu, 2016: Ambient volatile organic compounds and their effect on ozone production in Wuhan, central China. Science of the Total Environment, 541, 200−209, https://doi.org/10.1016/j.scitotenv.2015.09.093.
    Maji, K. J., W. F. Ye, M. Arora, and S. M. S. Nagendra, 2019: Ozone pollution in Chinese cities: Assessment of seasonal variation, health effects and economic burden. Environmental Pollution, 247, 792−801, https://doi.org/10.1016/j.envpol.2019.01.049.
    Mazzuca, G. M., X. R. Ren, C. P. Loughner, M. Estes, J. H. Crawford, K. E. Pickering, A. J. Weinheimer, and R. R. Dickerson, 2016: Ozone production and its sensitivity to NOx and VOCs: Results from the DISCOVER-AQ field experiment, Houston 2013. Atmospheric Chemistry and Physics, 16, 14 463−14 474, https://doi.org/10.5194/acp-16-14463-2016.
    Menchaca-Torre, H. L., R. Mercado-Hernández, J. Rodríguez-Rodríguez, and A. Mendoza-Domínguez, 2015: Diurnal and seasonal variations of carbonyls and their effect on ozone concentrations in the atmosphere of Monterrey, Mexico. Journal of the Air & Waste Management Association, 65, 500−510, https://doi.org/10.1080/10962247.2015.1005849.
    Mo, Z. W., M. Shao, and S. Lu, 2016: Compilation of a source profile database for hydrocarbon and OVOC emissions in China. Atmos. Environ., 143, 209−217, https://doi.org/10.1016/j.atmosenv.2016.08.025.
    Ou, J. M., J. Y. Zheng, R. R. Li, X. B. Huang, Z. M. Zhong, L. J. Zhong, and H. Lin, 2015: Speciated OVOC and VOC emission inventories and their implications for reactivity-based ozone control strategy in the Pearl River Delta region, China. Science of the Total Environment, 530−531, 393−402, https://doi.org/10.1016/j.scitotenv.2015.05.062.
    Peng, Y. P., K. S. Chen, C. H. Lai, P. J. Lu, and J. H. Kao, 2006: Concentrations of H2O2 and HNO3 and O3−VOC−NOx sensitivity in ambient air in southern Taiwan. Atmos. Environ., 40, 6741−6751, https://doi.org/10.1016/j.atmosenv.2006.05.079.
    Seco, R., and Coauthors, 2011: Contrasting winter and summer VOC mixing ratios at a forest site in the Western Mediterranean Basin: The effect of local biogenic emissions. Atmospheric Chemistry and Physics, 11, 13 161−13 179, https://doi.org/10.5194/acp-11-13161-2011.
    Shao, P., J. L. An, J. Y. Xin, F. K. Wu, J. X. Wang, D. S. Ji, and Y. S. Wang, 2016: Source apportionment of VOCs and the contribution to photochemical ozone formation during summer in the typical industrial area in the Yangtze River Delta, China. Atmospheric Research, 176−177, 64−74, https://doi.org/10.1016/j.atmosres.2016.02.015.
    Sharkey, T. D., A. E. Wiberley, and A. R. Donohue, 2007: Isoprene emission from plants: Why and how. Annals of Botany, 101, 5−18, https://doi.org/10.1093/aob/mcm240.
    Shen, Z. X., and Coauthors, 2008: Seasonal variations and evidence for the effectiveness of pollution controls on water-soluble inorganic species in total suspended particulates and fine particulate matter from Xi’an, China. Journal of the Air & Waste Management Association, 58, 1560−1570, https://doi.org/10.3155/1047-3289.58.12.1560.
    Shen, Z. X., and Coauthors, 2009: Ionic composition of TSP and PM2.5 during dust storms and air pollution episodes at Xi'an. China. Atmos. Environ., 43, 2911−2918, https://doi.org/10.1016/j.atmosenv.2009.03.005.
    Song, S. K., Z. H. Shon, Y. H. Kang, K. H. Kim, S. B. Han, M. Kang, J. H. Bang, and I. Oh, 2019: Source apportionment of VOCs and their impact on air quality and health in the megacity of Seoul. Environmental Pollution, 247, 763−774, https://doi.org/10.1016/j.envpol.2019.01.102.
    Srivastava, A., 2004: Source apportionment of ambient VOCS in Mumbai city. Atmos. Environ., 38, 6829−6843, https://doi.org/10.1016/j.atmosenv.2004.09.009.
    Sun, J., and Coauthors, 2018: VOCs emission profiles from rural cooking and heating in Guanzhong Plain, China and its potential effect on regional O3 and SOA formation. Atmospheric Chemistry and Physics, https://doi.org/10.5194/acp-2018-36.
    Sun, J., and Coauthors, 2019a: Volatile organic compounds emissions from traditional and clean domestic heating appliances in Guanzhong Plain, China: Emission factors, source profiles, and effects on regional air quality. Environment International, 133, 105252, https://doi.org/10.1016/j.envint.2019.105252.
    Sun, J., and Coauthors, 2019b: Urban VOC profiles, possible sources, and its role in ozone formation for a summer campaign over Xi’an, China. Environmental Science and Pollution Research, 26, 27 769−27 782, https://doi.org/10.1007/s11356-019-05950-0.
    Sun, J., and Coauthors, 2019c: Volatile organic compounds from residential solid fuel burning in Guanzhong Plain, China: Source-related profiles and risks. Chemosphere, 221, 184−192, https://doi.org/10.1016/j.chemosphere.2019.01.002.
    Sun, J., F. K. Wu, B. Hu, G. Q. Tang, J. K. Zhang, and Y. S. Wang, 2016: VOC characteristics, emissions and contributions to SOA formation during hazy episodes. Atmos. Environ., 141, 560−570, https://doi.org/10.1016/j.atmosenv.2016.06.060.
    Suryawanshi, S., A. S. Chauhan, R. Verma, and T. Gupta, 2016: Identification and quantification of indoor air pollutant sources within a residential academic campus. Science of the Total Environment, 569−570, 46−52, https://doi.org/10.1016/j.scitotenv.2016.06.061.
    Tan, Z. F., and Coauthors, 2018: Exploring ozone pollution in Chengdu, southwestern China: A case study from radical chemistry to O3-VOC-NOx sensitivity. Science of the Total Environment, 636, 775−786, https://doi.org/10.1016/j.scitotenv.2018.04.286.
    Tang, G. Q., and Coauthors, 2019: Decreased gaseous carbonyls in the North China plain from 2004 to 2017 and future control measures. Atmos. Environ., 218, 117015, https://doi.org/10.1016/j.atmosenv.2019.117015.
    Verstraeten, W. W., J. L. Neu, J. E. Williams, K. W. Bowman, J. R. Worden, and K. F. Boersma, 2015: Rapid increases in tropospheric ozone production and export from China. Nature Geoscience, 8, 690−695, https://doi.org/10.1038/ngeo2493.
    Wang, H., and Coauthors, 2014: Source profiles of volatile organic compounds from biomass burning in Yangtze River Delta, China. Aerosol and Air Quality Research, 14, 818−828, https://doi.org/10.4209/aaqr.2013.05.0174.
    Wang, H. L., and Coauthors, 2018: Emissions of volatile organic compounds (VOCs) from cooking and their speciation: A case study for Shanghai with implications for China. Science of the Total Environment, 621, 1300−1309, https://doi.org/10.1016/j.scitotenv.2017.10.098.
    Wang, T., L. K, Xue, P. Brimblecombe, Y. F. Lam, L. Li, and L. Zhang, 2017: Ozone pollution in China: A review of concentrations, meteorological influences, chemical precursors, and effects. Science of the Total Environment, 575, 1582−1596, https://doi.org/10.1016/j.scitotenv.2016.10.081.
    Wang, X., Z. X. Shen, J. J. Cao, L. M. Zhang, L. Liu, J. J. Li, S. X. Liu, and Y. F. Sun, 2012: Characteristics of surface ozone at an urban site of Xi'an in Northwest China. Journal of Environmental Monitoring, 14, 116−126, https://doi.org/10.1039/C1EM10541H.
    Watson, J. G., J. C. Chow, E. M. Fujita, 2001: Review of volatile organic compound source apportionment by chemical mass balance. Atmos. Environ., 35, 1567−1584, https://doi.org/10.1016/S1352-2310(00)00461-1.
    Wei, W., S. Y. Cheng, G. H. Li, G. Wang, and H. Y. Wang, 2014: Characteristics of volatile organic compounds (VOCs) emitted from a petroleum refinery in Beijing, China. Atmos. Environ., 89, 358−366, https://doi.org/10.1016/j.atmosenv.2014.01.038.
    Wu, S., and Coauthors, 2020: Vertically decreased VOC concentration and reactivity in the planetary boundary layer in winter over the North China Plain. Atmospheric Research, 240, 104930, https://doi.org/10.1016/j.atmosres.2020.104930.
    Wu, X. M., Y. Wu, S. J. Zhang, H. Liu, L. X. Fu, and J. M. Hao, 2016: Assessment of vehicle emission programs in China during 1998-2013: Achievement, challenges and implications. Environmental Pollution, 214, 556−567, https://doi.org/10.1016/j.envpol.2016.04.042.
    Yuan, B., Y. Liu, M. Shao, S. H. Lu, and D. G. Streets, 2010: Biomass burning contributions to ambient VOCs species at a receptor site in the Pearl River Delta (PRD), China. Environmental Science & Technology, 44, 4577−4582, https://doi.org/10.1021/es1003389.
    Zhang, J., H. Q. Jiang, G. Y. Liu, and W. H. Zeng, 2018a: A study on the contribution of industrial restructuring to reduction of carbon emissions in China during the five Five-Year Plan periods. Journal of Cleaner Production, 176, 629−635, https://doi.org/10.1016/j.jclepro.2017.12.133.
    Zhang, J. F., and K. R. Smith, 2007: Household air pollution from coal and biomass fuels in China: Measurements, health impacts, and interventions. Environmental Health Perspectives, 115, 848−855, https://doi.org/10.1289/ehp.9479.
    Zhang, Q., and Coauthors, 2015: Variations in PM2.5, TSP, BC, and trace gases (NO2, SO2, and O3) between haze and non-haze episodes in winter over Xi'an. China. Atmos. Environ., 112, 64−71, https://doi.org/10.1016/j.atmosenv.2015.04.033.
    Zhang, Q. J., L. Wu, X. Z. Fang, M. Y. Liu, J. Zhang, M. Shao, S. H. Lu, and H. J. Mao, 2018b: Emission factors of volatile organic compounds (VOCs) based on the detailed vehicle classification in a tunnel study. Science of the Total Environment, 624, 878−886, https://doi.org/10.1016/j.scitotenv.2017.12.171.
    Zhang, Z., and Coauthors, 2016: Spatiotemporal patterns and source implications of aromatic hydrocarbons at six rural sites across China's developed coastal regions. J. Geophys. Res., 121, 6669−6687, https://doi.org/10.1002/2016JD025115.
    Zou, Y., E. Charlesworth, C. Q. Yin, X. L. Yan, X. J. Deng, and F. Li, 2019: The weekday/weekend ozone differences induced by the emissions change during summer and autumn in Guangzhou, China. Atmos. Environ., 199, 114−126, https://doi.org/10.1016/j.atmosenv.2018.11.019.
  • [1] HAN Meng, LU Xueqiang, ZHAO Chunsheng, RAN Liang, HAN Suqin, 2015: Characterization and Source Apportionment of Volatile Organic Compounds in Urban and Suburban Tianjin, China, ADVANCES IN ATMOSPHERIC SCIENCES, 32, 439-444.  doi: 10.1007/s00376-014-4077-4
    [2] LIU Dameng, GAO Shaopeng, AN Xianghua, 2008: Distribution and Source Apportionment of Polycyclic Aromatic Hydrocarbons from Atmospheric Particulate Matter PM2.5 in Beijing, ADVANCES IN ATMOSPHERIC SCIENCES, 25, 297-305.  doi: 10.1007/s00376-008-0297-9
    [3] Shuang WU, Guiqian TANG, Yinghong WANG, Rong MAI, Dan YAO, Yanyu KANG, Qinglu WANG, Yuesi WANG, 2021: Vertical Evolution of Boundary Layer Volatile Organic Compounds in Summer over the North China Plain and the Differences with Winter, ADVANCES IN ATMOSPHERIC SCIENCES, 38, 1165-1176.  doi: 10.1007/s00376-020-0254-9
    [4] TAO Jun, CHENG Tiantao, ZHANG Renjian, CAO Junji, ZHU Lihua, WANG Qiyuan, LUO Lei, and ZHANG Leiming, 2013: Chemical composition of PM2.5 at an urban site of Chengdu in southwestern China, ADVANCES IN ATMOSPHERIC SCIENCES, 30, 1070-1084.  doi: 10.1007/s00376-012-2168-7
    [5] Yuepeng PAN, Mengna GU, Yuexin HE, Dianming WU, Chunyan LIU, Linlin SONG, Shili TIAN, Xuemei LÜ, Yang SUN, Tao SONG, Wendell W. WALTERS, Xuejun LIU, Nicholas A. MARTIN, Qianqian ZHANG, Yunting FANG, Valerio FERRACCI, Yuesi WANG, 2020: Revisiting the Concentration Observations and Source Apportionment of Atmospheric Ammonia, ADVANCES IN ATMOSPHERIC SCIENCES, 37, 933-938.  doi: 10.1007/s00376-020-2111-2
    [6] Wang Mingxing, 1985: SOURCE IDENTIFICATION AND APPORTIONMENT FOR ATMOSPHERIC AEROSOL BY FACTOR ANALYSIS, ADVANCES IN ATMOSPHERIC SCIENCES, 2, 469-477.  doi: 10.1007/BF02678745
    [7] Xun ZHU, 2003: Parameterization of the Non-Local Thermodynamic Equilibrium Source Function with Chemical Production by an Equivalent Two-Level Model, ADVANCES IN ATMOSPHERIC SCIENCES, 20, 487-495.  doi: 10.1007/BF02915493
    [8] Fang Juan, Wu Rongsheng, 1998: Influences of Vorticity Source and Momentum Source on Atmospheric Circulation, ADVANCES IN ATMOSPHERIC SCIENCES, 15, 41-46.  doi: 10.1007/s00376-998-0016-6
    [9] WAN Liying, ZHU Jiang, WANG Hui, YAN Changxiang, Laurent BERTINO, 2009: A ``Dressed" Ensemble Kalman Filter Using the Hybrid Coordinate Ocean Model in the Pacific, ADVANCES IN ATMOSPHERIC SCIENCES, 26, 1042-1052.  doi: 10.1007/s00376-009-7208-6
    [10] Yueliang CHEN, Changxiang YAN, Jiang ZHU, 2018: Assimilation of Sea Surface Temperature in a Global Hybrid Coordinate Ocean Model, ADVANCES IN ATMOSPHERIC SCIENCES, 35, 1291-1304.  doi: 10.1007/s00376-018-7284-6
    [11] Wu Aiming, Ni Yunqi, 1999: A Hybrid Coupled Ocean-Atmosphere Model and ENSO Prediction Study, ADVANCES IN ATMOSPHERIC SCIENCES, 16, 405-418.  doi: 10.1007/s00376-999-0019-y
    [12] PENG Guliang, CAI Xuhui, ZHANG Hongsheng, LI Aiguo, HU Fei, Monique Y. LECLERC, 2008: Heat Flux Apportionment to Heterogeneous Surfaces\\[.1cm] Using Flux Footprint Analysis, ADVANCES IN ATMOSPHERIC SCIENCES, 25, 107-116.  doi: 10.1007/s00376-008-0107-4
    [13] Jianfeng WANG, Ricardo M. FONSECA, Kendall RUTLEDGE, Javier MARTÍN-TORRES, Jun YU, 2020: A Hybrid Statistical-Dynamical Downscaling of Air Temperature over Scandinavia Using the WRF Model, ADVANCES IN ATMOSPHERIC SCIENCES, 37, 57-74.  doi: 10.1007/s00376-019-9091-0
    [14] ZHANG Rong-Hua, 2015: A Hybrid Coupled Model for the Pacific Ocean-Atmosphere System. Part I: Description and Basic Performance, ADVANCES IN ATMOSPHERIC SCIENCES, 32, 301-318.  doi: 10.1007/s00376-014-3266-5
    [15] Junya HU, Rong-Hua ZHANG, Chuan GAO, 2019: A Hybrid Coupled Ocean-Atmosphere Model and Its Simulation of ENSO and Atmospheric Responses, ADVANCES IN ATMOSPHERIC SCIENCES, 36, 643-657.  doi: 10.1007/s00376-019-8197-8
    [16] Lu ZHOU, Rong-Hua ZHANG, 2022: A Hybrid Neural Network Model for ENSO Prediction in Combination with Principal Oscillation Pattern Analyses, ADVANCES IN ATMOSPHERIC SCIENCES, 39, 889-902.  doi: 10.1007/s00376-021-1368-4
    [17] Junya HU, Hongna WANG, Chuan GAO, Rong-Hua ZHANG, 2024: Different El Niño Flavors and Associated Atmospheric Teleconnections as Simulated in a Hybrid Coupled Model, ADVANCES IN ATMOSPHERIC SCIENCES.  doi: 10.1007/s00376-023-3082-x
    [18] Zhang Ren, Yu Zhihao, 2000: Numerical and Dynamical Analyses of Heat Source Forcing and Restricting Subtropical High Activity, ADVANCES IN ATMOSPHERIC SCIENCES, 17, 61-71.  doi: 10.1007/s00376-000-0043-4
    [19] B. S. K. REDDY, K. R. KUMAR, G. BALAKRISHNAIAH, K. R. GOPAL, R. R. REDDY, V. SIVAKUMAR, S. Md. ARAFATH, A. P. LINGASWAMY, S. PAVANKUMARI, K. UMADEVI, Y. N. AHAMMED, 2013: Ground-Based In Situ Measurements of Near-Surface Aerosol Mass Concentration over Anantapur: Heterogeneity in Source Impacts, ADVANCES IN ATMOSPHERIC SCIENCES, 30, 235-246.  doi: 10.1007/s00376-012-1234-5
    [20] Boru MAI, Xuejiao DENG, Fang ZHANG, Hao HE, Tian LUAN, Fei LI, Xia LIU, 2020: Background Characteristics of Atmospheric CO2 and the Potential Source Regions in the Pearl River Delta Region of China, ADVANCES IN ATMOSPHERIC SCIENCES, 37, 557-568.  doi: 10.1007/s00376-020-9238-z
  • ESM-200153.pdf

Get Citation+

Export:  

Share Article

Manuscript History

Manuscript received: 30 May 2020
Manuscript revised: 31 August 2020
Manuscript accepted: 09 September 2020
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Profiles and Source Apportionment of Nonmethane Volatile Organic Compounds in Winter and Summer in Xi’an, China, based on the Hybrid Environmental Receptor Model

    Corresponding author: Zhenxing SHEN, zxshen@mail.xjtu.edu.cn
  • 1. Department of Environmental Science and Engineering, Xi’an Jiaotong University, Xi’an 710049, China
  • 2. Key Lab of Aerosol Chemistry & Physics, Institute of Earth Environment, Chinese Academy of Sciences, Xi’an 710049, China
  • 3. Changsha Center for Mineral Resources Exploration, Guangzhou Institute of Geochemistry, Chinese Academy of Sciences, Changsha 410645, China
  • 4. School of Human Settlements and Civil Engineering, Xi’an Jiaotong University, Xi’an 710049, China

Abstract: Summer and winter campaigns for the chemical compositions and sources of nonmethane hydrocarbons (NMHCs) and oxygenated volatile organic compounds (OVOCs) were conducted in Xi’an. Data from 57 photochemical assessment monitoring stations for NMHCs and 20 OVOC species were analyzed. Significant seasonal differences were noted for total VOC (TVOC, NMHCs and OVOCs) concentrations and compositions. The campaign-average TVOC concentrations in winter (85.3 ± 60.6 ppbv) were almost twice those in summer (47.2 ± 31.6 ppbv). Alkanes and OVOCs were the most abundant category in winter and summer, respectively. NMHCs, but not OVOCs, had significantly higher levels on weekends than on weekdays. Total ozone formation potential was higher in summer than in winter (by 50%) because of the high concentrations of alkenes (particularly isoprene), high temperature, and high solar radiation levels in summer. The Hybrid Environmental Receptor Model (HERM) was used to conduct source apportionment for atmospheric TVOCs in winter and summer, with excellent accuracy. HERM demonstrated its suitability in a situation where only partial source profile data were available. The HERM results indicated significantly different seasonal source contributions to TVOCs in Xi’an. In particular, coal and biomass burning had contributions greater than half in winter (53.4%), whereas traffic sources were prevalent in summer (53.1%). This study’s results highlight the need for targeted and adjustable VOC control measures that account for seasonal differences in Xi’an; such measures should target not only the severe problem with VOC pollution but also the problem of consequent secondary pollution (e.g., from ozone and secondary organic aerosols).

1.   Introduction
  • Urbanization and industrialization have increased the global level of ambient volatile organic compounds (VOCs), and VOC pollution has become an urgent environmental issue. VOCs are crucial precursors in the formation of tropospheric ozone (O3) and secondary organic aerosol (SOA) in the particulate phase (Verstraeten et al., 2015; Li et al., 2017b; Wang et al., 2017). Furthermore, VOCs harm human health (Peng et al., 2006; Durmusoglu et al., 2010; An et al., 2014; Song et al., 2019) and the economy (Maji et al., 2019). An extreme example is the photochemical smog in Los Angeles in the 1940s, which was induced by high levels of VOCs in combination with nitrogen oxides under a given set of meteorological conditions (specifically, high-intensity radiation, low wind speed, and low humidity). VOC pollution causes billions of dollars of economic loss and more than 400 deaths annually (Heo et al., 2015).

    Ambient VOC concentrations have been noted to vary with the emission intensities of dominant sources and with meteorological conditions; meteorological conditions, in particular, affect transport and chemical transformation (Filella and Peñuelas, 2006a; Carter, 2010). Studies have reported that urban VOCs are derived from multiple sources—including vehicle exhaust, petrochemical industry emissions, fossil fuel volatilization, the use of chemical solvents, and solid fuel combustion (An et al., 2017; Tan et al., 2018; Sun et al., 2019a) However, VOC compositions and sources have varied significantly between seasons and locations. Specifically, a study observed that aromatic hydrocarbons in northern China originated mainly from the burning of biomass, biofuel, and coal, whereas those in southern China originated mainly from traffic-related emissions (Zhang et al., 2016). An et al. (2014) reported that industry was the primary contributor to summer VOCs in Nanjing, China, whereas the traffic source was the primary contributor in winter.

    Such findings on variations in source contributions may also have been due to the different models used. Due to high uncertainty in source emission data and the lack of a localized VOC source inventory, source inventory-based methods are less popular than the receptor model is in existing source apportionment studies. For instance, positive matrix factorization (PMF) (Abeleira et al., 2017; Assan et al., 2018; Bari and Kindzierski, 2018; Chen et al., 2019) and chemical mass balance (CMB) (Watson et al., 2001; Srivastava, 2004; Feng et al., 2019) have both been widely used in VOC source apportionment. However, PMF requires large datasets, whereas CMB requires accurate localized VOC source profiles (Hopke, 2016). In cases with limited data samples or insufficiently localized source profile information, these two approaches cannot yield reliable source contributions, and a more advanced model is required. The Hybrid Environmental Receptor Model (HERM) is one such model, which was developed to take advantage of both CMB and PMF by allowing the use of incomplete source profiles as the model input (Chen and Cao, 2018).

    Xi’an, a megacity in China with a population of over 10 million people, has frequently had PM2.5 and O3 pollution episodes over the past decade (Shen et al., 2008, 2009; Huang et al., 2014; Zhang et al., 2015). Xi’an’s surrounding basin-like terrain has compounded such pollution to make Xi’an the Chinese city with the worst air quality (Huang et al., 2014). Studies on VOCs in Xi’an have been considerably fewer relative to studies on VOCs in China’s more developed regions—such as the Beijing—Tianjin−Hebei region, Yangtze River Delta, and Pearl River Delta (Duan et al., 2008; Yuan et al., 2010; Wang et al., 2018). Thus, accurate and detailed findings on seasonal VOC concentrations and source apportionment in Xi’an have been lacking, where such findings are essential for pollution control and policymaking. Furthermore, Xi’an’s basin-like terrain minimizes the effect of long-range transport, which makes Xi’an a good candidate for testing the feasibility of HERM (Feng et al., 2016). In this study, nonmethane hydrocarbons (NMHCs) and oxygenated VOC (OVOC) samples, at an hourly temporal resolution, were collected both in summer and winter. This study aimed to (1) characterize the chemical profiles of total VOCs (TVOCs) in summer and winter in Xi’an and (2) conduct the source apportionment of seasonal TVOCs by using HERM.

2.   Methods
  • The sampling site was in the southeastern part of downtown Xi’an, and the sampling height was 15 m above the ground (Fig. 1). The site is a typical urban sampling site, having a superstation for atmospheric monitoring. A residential area, the Xi’an Jiaotong University campus, South Second Ring road, and Xingqing road (both of which are main roads) lie to the north, east, south, and west of the sampling site, respectively (Fig. 1). Four hourly (0800−0900, 1500−1600, 1900−2000, and 2300−2400 LST) ambient NMHC (n = 112) and OVOC (n = 112) samples were collected every day during a summer (16−28 July 2018) and winter (8−21 January 2019) campaign. The four hourly sampling periods were selected to represent the morning rush hour, the period when photochemical reactions are at high levels, the evening rush hour, and the period when human activity is at its lowest, respectively (Ho et al., 2004; Mazzuca et al., 2016).

    Figure 1.  Location of the sampling site and surrounding road distribution.

    The ambient NMHC sampling was conducted using a 3-L silonite-treated stainless steel canister with a flow rate of approximately 50 mL min−1, which was controlled by a sampling valve (Entech Instruments, Simi Valley, CA, USA) (Sun et al., 2019c). The OVOC samples were collected into acidified 2, 4-dintrophenylhydrazine (DNPH) cartridges (Waters Sep-Pak DNPH-silica, Milford, MA, USA) at a flow rate of 30 mL min−1. An O3 scrubber (Waters, Milford, MA) and particle filter were installed before the DNPH cartridges to remove O3 and particulate matter. After the sampling, the canisters were stored in a cool and dark place, and the wrapped DNPH-silica cartridges were stored in a 4°C refrigerator. Continual measurements of gaseous pollutants (CO, O3, SO2, and NOx) were also conducted using instruments listed in Table S1 in the Electronic Supplementary Material (ESM).

  • The NMHC samples were analyzed using a gas chromatography-mass selective detector/flame ionization detector (Agilent Technologies, Santa Clara, CA, USA). The procedures for analysis and temperature programming were identical to those used in the study conducted by Zhang et al. (2016). This study measured 57 NMHCs specified by the VOC monitoring network of photochemical assessment monitoring stations (PAMS), and the minimum detection limits (MDLs) are presented in Table S2 (Liu et al., 2012; Wei et al., 2014). The MDLs for the 57 NMHC species ranged from 4 to 81 pptv.

    The OVOC cartridge was first eluted with 2 mL of acetone-free acetonitrile and then analyzed through high-pressure liquid chromatography (Series 1200, Agilent Technologies). The analysis protocol used in this study was identical to that used by Dai et al. (2018). We quantified 20 target OVOCs (including monocarbonyls and dicarbonyls) in the cartridges. The estimated MDLs of the OVOCs ranged from 0.016 to 0.12 ppbv at the sampling volume of 1.8 m3 (Table S3).

  • This study discusses the source apportionment results obtained using HERM. For HERM, 4 fixed profiles with 57 hydrocarbons and 3 OVOCs were selected for data input (Table S6). HERM was constructed based on the multilinear engine (ME-2) solution to CMB problems, which differs from the current version of CMB software (EPA CMB v.8.2). Specifically, HERM analyzes either one or many samples in a single run, where it solves collinearity problems by inheriting non-negativity constraints (Chen and Cao, 2018):

    Here, Ei,k was the effective variance of pollutant i in source k, i denoted a certain pollutant, j and k denoted two different sources, the δi,j value was set to 0 or 1 when the source profile element Fi,j was specified or unknown/missing in the profiles, respectively, and β is an adjustable factor with a default value of 1; β is included to ensure that the model does not overwrite unspecified profiles due to zero uncertainty information. One crucial difference [Eq. (2)] between the CMB model and HERM is that χ2, χk2, and χi2 are defined in HERM when dealing with missing profiles (Fi,j):

    where Ci,k is the measured concentration of a pollutant i in sample k, Fi,j is the source profile, and Si,j is the source contribution, I, J and K were the total number of pollutant i, source j and k, respectively.

    When no profile information is input, the formulation of HERM becomes akin to that of PMF (${\sigma}_{F{_{i,j}}}$ = 0, δi,j = 1). Moreover, a series of modifications were conducted according to the ME-2 model. These modifications included (1) allowing for an easy switch between the robust and non-robust mode, and (2) assigning a-priori (or random) values for the initial conditions of iteration, which is performed to accommodate HERM requirements. Furthermore, in the case of conventional CMB problems, where a full source profile is always required, HERM uses Eqs. (3) and (4) to estimate the missing profile and uncertainty, respectively:

    where $ {C}_{i,k} $ is the estimated profile and $ {\sigma }_{S,j,k}^{2} $ is the estimated uncertainty, F is the i × j profile matrix, and dEk is an i × i diagonal matrix with diagonal elements (dEk)i,i = Ei,k.

    Therefore, HERM can solve profiles that are either not assigned or have some species missing. HERM algorithms are further detailed in Chen and Cao (2018).

  • We used the conditional function probability (CPF) [Eq. (5)] to estimate the direction of different sources on the monitoring site, doing so by integrating source apportionment results and wind direction data (Chen et al., 2019):

    where mΔθ is the number of valid data points within the wind sector Δθ, and nΔθ is the total number of data points in the single wind sector Δθ. The settings for this function used in this study were identical to those used by Kim and Hopke (2004).

  • The O3 formation potential (OFP) is often used to estimate the capacity of TVOCs in forming O3 through photochemical reactions. The maximum incremental reactivity (MIR) method was detailed by Carter (2010). The method is characterized by the following identity:

    where [VOCx] and MIRx are the concentration and MIR of an individual VOC specie x, respectively. A VOC-limited and high NOx condition in Carter’s study was selected for calculation to match conditions in the Guanzhong area of China.

    Similarly, the SOA formation potential (SOAP) is typically calculated using the toluene-equivalent method, which is characterized as follows:

    where Ix was increment in SOA mass concentration with species x, It was increment in SOA with toluene, SOAP is expressed as an index relative to toluene multiplied by 100. This method has been widely used and verified (Kleindienst et al., 2007; Hu et al., 2008; Derwent et al., 2010). In our study, the capability of toluene in SOA formation was then converted into mass-based values according to our toluene gas-to-particle chamber test results (Johnson et al., 2004). A high NOx condition and low NOx condition were chosen to represent winter and summer conditions in Xi’an, respectively, with the gas-to-particle ratios of 0.06 and 0.20, respectively.

3.   Results and discussion
  • We noted that 54 of the 57 measured PAMS NMHCs and all of the 20 OVOCs had concentrations higher than their individual MDLs. The time series (four data points per day) of the concentrations of TVOCs by category in winter and summer are presented in Fig. 2. The campaign-average TVOC concentration was twice as large in winter (85.3 ± 60.6ppbv) than in summer (47.2 ± 31.6 ppbv). Previous studies have observed similar levels in other locations, which have been mostly attributed to the boundary layer height being lower in winter than in summer (Filella and Peñuelas, 2006b; Liu et al., 2010; Wu et al., 2016). In addition, winter heating in rural areas around Xi’an resulted in greater coal and biomass burning, which resulted in the greater release of more VOCs into the atmosphere (Sun et al., 2018, 2019c). Similarly, Beijing was reported to have a much higher TVOC level in winter (216.05 ± 115.71 ppbv) than in summer (127.01 ± 39.67 ppbv); both of these levels were significantly higher than those in Xi’an (Gu et al., 2019). In winter, alkanes contributed almost 50% of the TVOCs, followed by OVOC (16.8%), acetylene (15.1%), alkenes (11.3%), and aromatics (8.1%). In summer, OVOC was the dominant contributor (47.9%), followed by alkanes (28.6%), alkenes (11.8%), aromatics (7.3%), and acetylene (4.5%). The higher proportion of OVOC in summer than in winter is attributable to secondary formation resulting from the higher photochemical reactivity under stronger UV radiation (Liu et al., 2019).

    Figure 2.  Time series of the concentrations of PAMS NMHCs and OVOCs in the winter and summer campaign (M, A, E and N denote morning, afternoon, evening and night, respectively; the date format is as YY/MM/DD)(data absence on 18/7/23 and 18/7/30 were due to instrument failure).

    Large daily variations of TVOC concentrations (23.4−71.4 and 30.0−112.1 ppbv for summer and winter, respectively) were observed in this study. Table 1 presents a comparison of weekday with weekend data. The average concentrations of PAMS NMHCs were higher on weekends than in weekdays, both during summer and winter (p < 0.05). Alkanes contributed most to the increased PAMS NMHCs on weekends, likely because of higher traffic from weekend activities (e.g., shopping and social gatherings) (Zou et al., 2019). A previous study in Xi’an observed that the vehicle count at South Second Ring road (close to this study’s sampling site) was extremely high on weekends (Li et al., 2017a). These slightly higher NO2 and CO concentrations on weekends than on weekdays also indicated that vehicle count considerably affected PAMS NMHC concentrations (Table S4). Similar trends were also observed for OVOC concentrations, albeit with less significant weekend−weekday differences (p > 0.05) than those of PAMS NMHC. No significant weekend−weekday differences were noted for O3 (p > 0.05). This was because alkanes from vehicle exhausts generally have low photochemical reactivity and will not generate large amounts of OVOC and O3 through photochemical reactions (Hernández-Paniagua et al., 2018). The weekend−weekday difference for average OFP was also statistically nonsignificant (p > 0.05) during both winter and summer. This supported our hypothesis that high TVOC concentrations do not necessarily entail high O3 concentrations (Fig. S1 in the ESM). The weekend−weekday difference for temperature was also statistically nonsignificant (p > 0.05, Table S4). However, the considerably large variance in temperature between the seasons potentially led to large effects on biogenic emissions (i.e., isoprene) and combustion sources (i.e., ethylene), which had the winter-to-summer ratios of 6.5 and 0.02, respectively.

    SeasonCategoryWeekdaysWeekends
    AvgSDAvgSD
    SummerOVOCs18.96.219.94.2
    NMHCs24.7a10.934.3b23.9
    WinterOVOCs17.07.917.68.2
    NMHCs56.8c29.673.2d33.8
    Notes: Na, N=8; Nb, N=4; Nc, N=10; Nd, N=4.

    Table 1.  Average concentrations (Avg) and standard deviations (SD) of OVOCs and NMHCs over weekdays and weekends (ppbv).

    Diurnal variations in ambient TVOCs are controlled by their sources, transport, and chemical reactions (Lyu et al., 2016). The diurnal variations of TVOCs measured in this study are presented in Fig. 3. The TVOC patterns were similar between winter and summer; this pattern was of a single morning peak (M, 0800−0900 LST) before a gradual decline to the lowest value. The increasing planetary boundary layer during the day and the natural degradation of VOCs during the night were largely responsible for this diurnal pattern in summer (Menchaca-Torre et al., 2015; Wu et al., 2020). However, an obvious nighttime rebound (N, 2300−2400 LST) was only observed in winter. Space heating on cold nights potentially explains the slightly different diurnal pattern in winter (Zhang and Smith, 2007; Suryawanshi et al., 2016). In fact, the large increases of alkane, alkene, and aromatic concentrations in winter nights supported this hypothesis, and other studies have also noted this phenomenon (Wang et al., 2014; Cheng et al., 2018).

    Figure 3.  Campaign-average TVOCs and proportions by category.

  • Figure 4 presents the time series of OFPs in winter and summer calculated using the MIR method. The campaign-average OFPs were > 50% higher in summer (average: 304.3 ppbv) than in winter (average: 178.2 ppbv), despite summer having only half the TVOCs compared with winter. Alkenes were the dominant contributor of OFPs in both summer and winter, at 38.7% and 50.5%, respectively. Because the photochemical reactivity difference is not considered in our use of the MIR method, the seasonal differences in alkene-contributed OFPs were mainly caused by the different chemical compositions in this category. For example, biogenic alkenes (i.e., isoprene) had much higher levels in summer (94.8 ppbv) than in winter (0.25 ppbv). The very large seasonal difference in biogenic OFPs was primarily due to the deciduous vegetation in Xi’an; biogenic OFP concentrations were much higher in Guangzhou and exhibited a different pattern probably due to the dominance of evergreen vegetation in Guangzhou (Ou et al., 2015). The order of contributions of other VOC categories to the total OFPs, from high to low, was OVOC > aromatics > alkane > alkyne. OVOCs, especially formaldehyde and methylglyoxal, contributed greatly to OFPs due to their high photochemical reactivity and concentrations (Louie et al., 2013). The relatively low photochemical reactivity of alkanes and aromatics limited their potential in forming O3. The time series of OFPs indicated the ranges of 71.9−746 and 59.8−554 ppbv in summer and winter, respectively. These high calculated OFP values in this study was because the MIR method models an ideal capability of VOCs in forming O3, in which the complex reactions require high temperatures, high levels of solar radiation, and a proper NOx level (Carter, 2010). The higher temperature and stronger solar radiation in summer favor the transfer of OFPs to O3 (Table S4), and the higher OFPs in winter did not result in high O3 concentrations (Seco et al., 2011; Feng et al., 2016).

    Figure 4.  Time series of OFPs by PAMS NMHCs and OVOCs in winter and summer (M, A, E and N denote morning, afternoon, evening and night, respectively; the date format is as YY/MM/DD) (data absence on 18/7/23 and 18/7/30 were due to instrument failure).

    SOAPs in winter and summer are presented in Fig. 5. In toluene-equivalent SOAPs, aromatic hydrocarbons constituted the dominant group because they demonstrated higher photo oxidation reactivity than alkane and alkenes did (Carter et al., 2012; Li et al., 2017a). OVOCs, the second largest contributor to SOAPs, contributed approximately 10% in winter and even less in summer, and benzaldehyde was the major species due to its extremely high toluene equivalence indexes. SOAPs from the group of alkanes and alkenes were negligible, (e.g., one to two orders of magnitude lower than those of the major groups at p < 0.05). According to the results of Huang et al. (2014), SOA occupied an average of 15.5% in Xi’an aerosols, and the campaign-average PM2.5 concentrations in winter and summer were 126.2 μg m−3 and 21.2 μg m−3, respectively (Shaanxi Environmental Monitoring Center, China). The calculated SOAPs had the proportions of 6.6% and 6.7% in SOA for summer and winter, respectively, which were lower than those reported in Beijing by Sun et al. (2016). This difference was possibly due to the SOA estimation methods used and the minimum OC/EC (organic carbon/elemental carbon) method potentially underestimating the actual SOA level (Huang et al., 2014).

    Figure 5.  Time series of SOAPs by PAMS NMHCs and OVOCs in winter and summer (M, A, E and N denote morning, afternoon, evening and night, respectively; the date format is as YY/MM/DD)(data absence on 18/7/23 were due to instrument failure).

  • The statistical results of the measured TVOCs are presented in Table S5a and Table S5b for the winter and summer campaign, respectively. The hourly average TVOCs had the ranges of 58.4−95.5 and 30.8−56.8 ppbv in winter and summer, respectively. The compositions in TVOCs significantly differed between summer and winter. For example, the proportions of acetylene and alkenes were twice as much in winter than in summer and as high as a factor of six during a nighttime period. Short-chain alkenes and alkynes are produced in large quantities from coal and biomass burning (Liu et al., 2017; Sun et al., 2019a), which explained their much higher fractions in winter in Xi’an. By contrast, isoprene exhibited much higher proportions in summer (0.94 ± 0.53 ppbv, 2.0%) than in winter (0.03 ± 0.07 ppbv, 0.04%) because isoprene primarily stems from biogenic emissions. High temperatures in summer favor lush vegetation growth, which resulted in high biogenic VOCs emissions (Li et al., 2018). By contrast, low temperatures in winter suppressed the emissions of isoprene from vegetation (Sharkey et al., 2007).

    Alkanes and aromatics also had higher concentrations and proportions in TVOCs during winter than during summer. The order of the prevailing alkane species, from high to low, was ethane > isopentane > isobutane in summer—which were all related with vehicle exhausts (Hwa et al., 2002)—and ethane > propane > n-butane in winter—which are similar to the chemical composition of combustion products, such as those from the burning of natural gas and biomass (Liu et al., 2008). For aromatics, the most abundant species were toluene and benzene in winter and summer, respectively. These results indicated the main VOC sources in Xi’an to be combustion in winter and vehicle exhausts in summer (Mo et al., 2016). OVOCs, the only category of TVOCs with higher concentrations in summer than winter, contributed approximately 50% of TVOC in summer but only 16.8% in winter, a phenomenon that was potentially due to stronger photochemical reactivity in summer, supported by summer’s high temperatures and atmospheric oxidability (expressed in terms of O3 concentration) (Wang et al., 2012; Shao et al., 2016).

    Because the concentrations of selected VOC species are indicative of pollution sources, the 10 species with the highest campaign-average concentrations in NMHCs are listed in Fig. S2, and their hourly averages are listed in Fig. 6. The ranking of isopentane, a tracer of gasoline vehicle exhaust (Zhang et al., 2018b), decreased from the second in summer to the eighth in winter, indicating that contributions from vehicle exhausts in winter were not as significant as those in summer. Ethylene, which is primarily emitted from coal and biomass burning, ranked seventh in summer but fourth in winter, implying higher contributions from combustion sources in winter than in summer (Yuan et al., 2010). For the hourly average, species rankings and, by implication, relative source contributions, changed very little over time in winter (Fig. 6). By comparison, species rankings varied over time in summer, which indicated variations in the relative source contributions (i.e., traffic) for PAMS.

    Figure 6.  Hourly averaged top 10 species in (a) winter and (b) summer. The box plots indicate the campaign-averaged VOC concentrations and the minimum, 25th, 50th, 75th, and maximum percentiles. A cumulative frequency curve is fitted to the top 10 species

    Figure 6.  (Continued)

    The characteristics of the five selected VOC species (accounting for more than 50% of total PAMS NMHCs) were analyzed using CPF to determine their emission sources, as indicated in Fig. S3. In general, the main sources for the five VOC species were located northwest of the city, which is upwind of the city during winter (Fig. 1). The location map (Fig. 1) clearly indicates a main road (Xi’an Second Ring road) south of the sampling site. Vehicle exhaust with high concentrations of isopentane and short-chain alkanes (ethane and propane) can be brought to the site by southern winds (Li et al., 2017a). The results also indicated considerable sources in the northeast during winter, which potentially stemmed from human activity in the downtown area, that had its pollutants brought to the site by the prevailing wind direction (shown in Fig. S4). For isoprene, the CPF results resembled the wind rose, which indicated that isoprene was mainly affected by wind direction because large concentrations of vegetation (i.e., from the park) were located elsewhere in the city. The CPF results for toluene were unique in that sources differed neither between summer and winter nor between whether they were located in the southwest, northwest, or southeast. Because sources located in these three directions of the sampling site have large and constant toluene emissions (Huang and Hsieh, 2019), the heavy traffic and high population density in these three directions were highly related.

  • HERM needs a certain sample number not less than the input species. In this study, some PAMS species were not detected or were in extremely low concentrations with rare variations, which were invalid inputs to HERM. In this situation, the actual species number input for HERM’s calculations was 46, which was smaller than the sample number (56) and made it possible for the separated source apportionment in summer and winter by HERM. In running HERM, four localized VOC source profiles—namely, biomass burning, coal combustion, gasoline vehicle exhaust, and diesel vehicle exhaust—were selected for model input (Li et al., 2017a; Sun et al., 2019b). Because VOC emissions from biomass burning and coal combustion varied significantly with different fuels and burning devices, two localized profiles from in-situ residential scenarios in Guanzhong plain were selected (Sun et al., 2019a). As biomass open burning has been legally banned by local government, residential use is the predominant source of biomass burning in Xi’an, which makes it representative in source apportionment (http://www.gov.cn/jrzg/2013-06/17/content_2427331.htm). Residential and industrial coal combustion showed similar profiles, and thus a more reliable source profile from residential use was selected and these two sources were not resolved separately due to the colinearly problem (Liu et al., 2008; Mo et al., 2016). By contrast, vehicle emissions exhibited no obvious regional divergence, and we thus selected gasoline and diesel emission profiles from a comprehensive and contemporary study in China (Li et al., 2017a). All 57 PAMS NMHC species and three representative OVOCs, as presented in Table S6, were selected for source apportionment calculation; they comprised 75.6% and 99.0% of the total TVOCs in summer and winter, respectively. After several trials of iterative operations, conducted separately in summer and winter, we decided on a combination of seven source profiles, which had Q > 1000 and R2 > 0.80 (both are crucial parameters for running the HERM). Among these seven source profiles, four were initially provided as model input (as mentioned), and three were newly generated by HERM calculations (Fig. 7). The three new profiles were consistent both for winter and summer. The first profile was characterized by high percentages of m/p-xylene, ethylbenzene, and o-xylene being highly associated with solvent use (Borbon et al., 2002). The second profile was characterized by compositions of isopentane and n-butane, which were determined to be from gasoline vapor (Mo et al., 2016). The third profile had a very high percentage of isoprene, which was likely to be associated with biogenic VOCs (BVOCs) (Kleindienst et al., 2007).

    Figure 7.  Source profiles for HERM inputs and outputs

    Figure 7.  (Continued)

    The percentage contributions from the seven sources are presented in Fig. 8, and the relative contributions from these sources had significant seasonal variations. In winter, combustion-related sources (biomass burning and coal combustion) contributed most to the TVOC concentration (53.4%) because of urban and rural heating, as mentioned. Our previous study demonstrated that residential coal and biomass burning were both crucial sources of ambient TVOCs in Xi’an (Sun et al., 2019b). Traffic-related sources contributed more than one-third (34.3%) to the total TVOC concentration, in which diesel (18.6%) and gasoline (14.5%) vehicle exhausts were dominant and gasoline vapor (1.4%) was negligible. The VOC emission factors were higher for vehicles powered by diesel than by gasoline (Ho et al., 2009). Solvent use, usually in industrial processes (e.g., painting), only contributed 11.9% to the TVOC concentration in winter. By comparison, industrial emissions were the biggest contributor to ambient TVOCs in Xi’an in 2013 (Feng et al., 2016). This indicates that government-led efforts in industrial restructuring have improved air quality in Xi’an (Zhang et al., 2018a). The percentage contribution of BVOCs in winter was as low as 0.2%, which was because deciduous vegetation does not emit BVOCs and because emissions from evergreen vegetation are suppressed by low temperatures in winter (Kelly et al., 2018).

    Figure 8.  Source apportionment results for TVOCs by HERM in winter and summer.

    In summer, traffic-related sources were the largest contributor to TVOCs (53.1%), in which diesel vehicle emissions, gasoline vehicle emissions, and gasoline vapor contributed 26.5%, 21.9% and 4.7%, respectively. However, traffic-related VOC emission intensities changed little between winter and summer (Tang et al., 2019). The contributions from solvent use were comparable between summer (12.8%) and winter (11.9%). Combustion sources contributed less in summer (15.3% in total) than in winter. Biomass burning still contributed 13.2% to the TVOC concentration, which was reflective of the common practices of residential and open biomass burning in Xi’an and surrounding areas. The percentage contribution from BVOCs to TVOCs was much higher in summer (18.7%) than in winter (0.2%) due to strong vegetation emissions in hot weather.

4.   Conclusions
  • To investigate the seasonal characteristics of the chemical compositions and sources of ambient TVOCs in Xi’an, two sampling campaigns (each lasting two weeks) were conducted. Hourly average concentrations of TVOCs in the winter campaign was nearly twice those in the summer campaign. In winter, alkanes contributed most to the TVOC concentration, followed by OVOC, acetylene, alkenes, and aromatics; in summer, OVOCs contributed most, followed by alkanes, alkenes, aromatics, and acetylene. OFPs were estimated to be 50% higher in summer than in winter due to the abundance of active OVOCs. Using four known source profiles as HERM inputs, we identified three more source profiles through internal calculation. These seven source contributions to TVOCs were as follows: coal combustion (29.7%), biomass burning (23.7%), diesel vehicles (18.6%), gasoline vehicles (14.5%), solvent use (11.9%), gasoline vapor (1.4%), and BVOCs (0.2%) in winter and diesel vehicles (26.5%), gasoline vehicles (21.9%), BVOCs (18.7%), biomass burning (13.2%), solvent use (12.8%), gasoline vapor (4.7%), and coal combustion (2.1%) in summer. Although HERM can only resolve primary sources, the model exhibited superior performance in handling partial source profile data for VOC source apportionment. Researchers with such limitations in their data should consider adopting HERM.

    It should be noted that uncertainties in source apportionment widely exist in model execution, data measurement and source profile selection (Lee and Russell, 2007). The advanced HERM used in this study could decrease the uncertainty to some extent by adopting better algorithms and combining the advantages of CMB and PMF models (Chen and Cao, 2018). However, the limited sample number in this study, a mixture of source profiles from local studies and the literature, as well as the absence of certain profiles (i.e., industrial coal combustion), may increase the uncertainty of this study. A comprehensive future study is needed with larger sample sizes and longer sampling periods, along with far more local studies providing representative source profiles to reduce the uncertainty of VOC source apportionment.

    Acknowledgements. This research was supported by the Natural Science Foundation of China (Grant No. 41907188), Natural Science Foundation of Shaanxi Province, China (Grant No. 2019JQ-386), and the China Postdoctoral Science Foundation (Grant No. 2019M653658).

    Electronic supplementary material: Supplementary material is available in the online version of this article at https://doi.org/10.1007/s00376-020-0153-0.

Reference

Catalog

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return