You are on page 1of 12

Atmospheric Research 218 (2019) 257–268

Contents lists available at ScienceDirect

Atmospheric Research
journal homepage: www.elsevier.com/locate/atmosres

Temporal distribution and source apportionment of PM2.5 chemical T


composition in Xinjiang, NW-China
Yusan Turapa, Dilinuer Talifua, , Xinming Wangb, Abulikemu Abulizia, Mailikezhati Maihemutia,

Yalkunjan Tursuna, Xiang Dingb, Tuergong Aierkena, Suwubinuer Rekefua


a
Key Laboratory of Coal Clean Conversion & Chemical Engineering Process, College of Chemistry and Chemical Engineering, Xinjiang University, Urumqi 830046, China
b
State Key Laboratory of Organic Geochemistry, Guangzhou Institute of Geochemistry Chinese Academy of Sciences, Guangzhou 510640, China

ARTICLE INFO ABSTRACT

Keywords: Daily fine particulate matter samples were collected in Dushanzi district within four months from September
Fine particulate matter 2015 to August 2016 and represent the four seasons. The samples were determined for major chemical com-
Chemical composition ponents in PM2.5, including elements, water-soluble ions (WSIs) and the organic/elemental carbon (OC/EC). The
Hysplit trajectory model results indicated that the annual mean PM2.5 concentration was 62.85 ± 43.5 μg m−3 in the Dushanzi district,
Sources apportionment
with the highest seasonal average in winter (95.47 ± 61.7 μg m−3) and the lowest in summer
(33.22 ± 17.7 μg m−3). The crustal elements were the most abundant elements and accounted for 96.51% of
the total analyzed elements. Carcinogenic metals, such as Cr, Pb, As and Cd, originated from human activity,
especially during winter. The highest total WSI concentration was 68.99 μg m−3 in winter, followed by autumn
(16.32 μg m−3), spring (10.23 μg m−3) and summer (7.06 μg m−3). SO42−, NO3− and NH4+ were the most
abundant WSIs in Dushanzi. Ion balance calculations showed that PM2.5 in winter was acidic; in autumn and
spring alkaline; and in summer nearly neutral. Total carbonaceous aerosol (TCA) accounted for 34% of the
PM2.5. The chemical mass closure (CMC) indicated that minerals and WSIs were the major fraction, accounting
for 33.58% and 23.17% of PM2.5 mass concentration, respectively. Dushanzi was controlled by four major air
masses, and the relative contributions of these air masses differ by season. Positive matrix factorization (PMF)
analysis identified six sources including vehicle emission, biomass burning, coal combustion, industrial pollu-
tion, secondary aerosols and soil dust, with annual mean contributions of 9.43%, 10.86%, 18.45%, 12.15%,
18.26% and 30.85%, respectively. Moreover, the relative contributions of these identified sources varied sig-
nificantly with the changing seasons.

1. Introduction formation mechanism of PM2.5 in some megacities of China, such as


Beijing, Nanjing, Shanghai and Guangzhou, have been investigated
Atmospheric particulate matter (PM) pollution has become a serious (Huang et al., 2014; Ming et al., 2017; Wang et al., 2006). However,
environmental problem, especially in developing countries, owing to its very few investigations of PM2.5 have been conducted in northwest
severe threat to human health and climate stability (Cesari et al., 2016; China, especially in the Xinjiang Uygur Autonomous Region. Xinjiang is
Ming et al., 2017). Particularly, airborne PM2.5 (mean aerodynamic located in the central post of the largest continent, the Eurasian con-
diameter ≤ 2.5 μm) is extremely harmful because the tiny particles are tinent; it is an important strategic area of the Belt and Road and is going
composed of various carcinogenic chemicals that can enter into the through rapid economic and industrial development. In the last decade,
human respiratory system and even penetrate into the circulatory Xinjiang has suffered from a high level of air pollution. The source of
system, leading to cell dysfunction, inflammation and cardio-pul- PM in Xinjiang gradually shifted from dust storms to a mixture of ve-
monary disease (Semmler et al., 2008; Yu et al., 2016). Numerous hicle emissions, industrial emissions, biomass combustion and dust
epidemiological and toxicological studies found that 3.45 million pre- storms.
mature deaths worldwide were related to PM2.5 pollution in 2007 Dushanzi district is an enclave of Karmay city, which is a central city
(Zhang et al., 2017). in Xinjiang and is in an upstream area of Asia dust. It is located in the
Earlier studies, chemical compositions, potential sources and the northwestern margin of China's second largest basin, Junggar basin,


Corresponding author.
E-mail address: dilnurt@xju.edu.cn (D. Talifu).

https://doi.org/10.1016/j.atmosres.2018.12.010
Received 4 February 2018; Received in revised form 14 November 2018; Accepted 10 December 2018
Available online 11 December 2018
0169-8095/ © 2018 Elsevier B.V. All rights reserved.
Y. Turap et al. Atmospheric Research 218 (2019) 257–268

with a total residential area of 26 km2 and a population of 90 thousand; equipment was installed on the roof of a residential building (20 m
most of the area is covered by desert and Gobi (Li et al., 2016). above ground level). The topography around the site was flat and open.
Dushanzi district faces the Gurbantongute desert in the east and Ta- PM2.5 was collected by a high-volume sampler (Wuhan Tianhong
kelamakan desert in the south. The geomorphological type of the Instruments Co., Ltd., TH-1000) at a controllable flow rate of
Dushanzi district is predominantly the Gobi desert; the district has a 1.05 m3 min−1. The filters were changed every 22 h (from 10:00 a.m. to
typical temperate continental arid desert climate. The ratio of eva- 8:00 a.m. the next day). Field blank filters were also collected without
poration to precipitation of rainwater is > 27.6 (Pan et al., 2012). The ambient air passing through the instrument during the sampling per-
poor ecological environment is uniquely susceptible to atmospheric iods. In total, 83 PM2.5 samples were collected during the following
pollution. In addition, Dushanzi district is an industrial area char- periods: September 14 to October 5, 2015 (autumn, 20 samples), De-
acterized by large-scale petrochemical oil refinery industries and power cember 5 to 28, 2015 (winter, 22 samples), April 19 to May 16, 2016
plants. In recent years, Dushanzi has undergone rapid economic de- (spring, 22 samples), and July 1 to 22, 2016 (summer, 23 samples).
velopment and population growth, which have increased the emissions PM2.5 samples were collected on quartz-fiber filters (Whatman,
of air pollutants, thus raising widespread health concerns in local re- 203 × 254 mm), which were pre-heated at 450 °C for 4 h before use,
sidents. However, there have been few reports on air pollutants in with the aim to reduce the residual carbon levels that were associated
Dushanzi. Systematic knowledge of chemical composition and sources with the filters. The filters were then equilibrated at a temperature
apportionment of PM2.5 is still lacking for Dushanzi compared to other between 20 ± 2 °C, and the relative humidity was between 35% and
cities. 45% for 48 h. Each filter was wrapped in Al-foils, sealed in polyethylene
The objectives of this study were to elucidate the seasonal variations bags and stored in a refrigerator at −20 °C until analysis.
and potential sources of PM2.5 by continuously collecting the daily
PM2.5 samples across Dushanzi for four months. The primary aims of
2.2. Chemical analysis
this study are to critically: (1) identify the chemical species in the PM2.5
samples in an arid and petrochemical industrialized area of Xinjiang;
The weighing of simples and blank filters were performed at least
(2) analyze the seasonal variations of the chemical components and
three times both before and after sampling using an electronic micro-
source contributions; and (3) quantitatively assess the potential sources
balance (Metller Toledo AB204-S) with a precision of 0.1 mg, and the
of PM2.5 using a PMF model coupled with chemical mass closure and a
differences of the filter weights were divided by the sampling volume to
Hysplit Trajectory Model. The results gained from this study will be
obtain PM2.5 mass concentrations.
useful for understanding the temporal distribution of PM2.5 chemical
An area of 5.06 cm2 of each filter, including a blank sample, was cut
compounds in northwest China and will contribute to the development
to analyze elements. First, each filter was soaked in a mixture acid
of more effective control strategies for PM2.5 in arid and petrochemical
(4 mL 67–70% HNO3 and 1 mL 47–51% HF) and was extracted by mi-
areas.
crowave digestion (XT-9912, Shanghai). Then, the sample was heated
by an electric stove and was evaporated until 1 mL of residual was left.
2. Materials and methods After cooling to room temperature, the sample was transferred to a test
tube, and 12 mL ultrapure water was added. Subsequently, inductively
2.1. PM2.5 sampling coupled plasma-mass spectrometry (ICP-MS) was used to determine the
concentrations of the elements. The major elements including Na, Mg,
Air samples were collected during typical periods in a residential Al, K, Ca, V, Cr, Mn, Fe, Co, Ni, Cu, Zn, As, Rb, Sr, Cd, Ba, Tl and Pb
and commercial area of the Dushanzi district (44°19′33″N, 84°53′0″E) were measured. The method detection limit (MDL) of the elements
(Fig. 1), which was located in the southwestern part of the petro- ranged from 0.1 to 1 ng m−3, and the uncertainties were < 5%.
chemical refinery industries and thermal power plant. The sampling A small portion of each filter (area of 5.06 cm2) was cut to analyze the

Fig. 1. Location of the sampling site.

258
Y. Turap et al. Atmospheric Research 218 (2019) 257–268

water-soluble ions. The filters were extracted ultrasonically using approxi- TEO = 1.3 × [0.5 × (Sr + Ba + Mn + Co + Rb + Ni + V)
mately 15 g ultrapure water (specific resistivity =18.5 MΩ cm−1) for
+ 1.0 × (Cu + Zn + Mo + Cd + Tl + Pb + As + Se)] (3)
20 min and were centrifugalized for 10 min. Then, 4 mL supernatant solu-
tion was taken in a clean bottle. This process was carried out twice. Next,
1 mL solution was drawn out by using a syringe and was injected into the 2.5. Sources identification methods
ion chromatograph (BIC-plus-883). Before analysis of the water-soluble
ions, the standard solutions were detected three times and low relative Backward trajectories with a Hybrid Single Particle Lagrangian
standard deviations were observed. Finally, water-soluble ions were iden- Integrated Trajectory (HYSPLIT) model were used to determine the
tified by the solution flowing through the cation and anion columns. Field transport trajectory of air parcels at the sampling site. We calculated
blanks were used to correct the concentrations of ionic species. The detec- 48 h air mass back trajectories arriving at sampling site during the
tion limits of these ions were within the range of 0.01 to 0.03 μg m−3, and sampling period by using the National Oceanic and Atmospheric
the uncertainties were < 5%. Administration HYSPLIT 4 (Li et al., 2017). The six hourly final archive
A 0.5 cm2 punch area from each filter was measured by DRI Model data were generated from the National Center for Environmental Pre-
2015 (thermal/optical carbon analyzer) for OC and EC, following the diction's Global Data Assimilation System (GDAS) wind field reanalysis.
Interagency Monitoring of Protected Visual Environment (IMPROVE). More details about the HYSPLIT model can be found at https://ready.
The analysis consisted of two stages: In the first stage, a non-oxidizing arl.noaa.gov/archives.php (NOAA Air Resources Laboratory). The
helium atmosphere of four OC fractions were produced. In the second model was run four times per day at starting times of 04:00, 10:00,
stage, the samples were heated in helium containing 2% oxygen, which 16:00, and 22:00 UTC (12:00, 18:00,00:00, and 06:00 LT - local time,
yielded three EC fractions. In this process, pyrolyzed carbon was de- respectively). The arrival level was set at 100 m above ground level
termined by the reflected laser light and returned to the initial value (AGL).
after oxygen was added to the analysis atmosphere. The detection limits The PMF, which was developed by the Environmental Protection
were 0.18 ± 0.06 and 0.04 ± 0.01 μg m−3 for OC and EC, respec- Agency, is an effective source apportionment receptor model that does
tively, with uncertainties < 5%. not require source profiles prior to analysis and has no limitation on the
numbers of source (Lee et al., 2003; Zhang et al., 2013). The principles
of PMF can be found elsewhere, in detail (Chen et al., 2017; Hsu et al.,
2.3. Quality assurance and quality control
2016; Paatero and Tapper, 1994; Yu, 2013). In this study, PMF 5.0 was
employed to measure components, including: PM2.5, OC, EC, Cl−,
All analytical procedures were monitored with strict quality assur-
NO3−, SO42−, NH4+, Na,Mg, Al, K, Ca, V, Cr, Mn, Fe, Co, Ni, Cu, Zn,
ance (QA) and quality control (QC) measures. Field and laboratory
As, Rb, Sr, Cd, Ba, Tl and Pb. WSIs such as Na+, K+, Mg2+ and Ca2+
blanks were analyzed in the same manner as the filter samples. Both
were excluded due to their duplication in elemental analysis. The un-
pre- and post-sampling and component analysis strictly guaranteed that
certainty (Unc) is calculated by using the following equation, which is
the quartz fiber filter was intact and undamaged, and the cracked filters
based on the error fraction and the method detection limit (MDL). If the
were excluded. The background contamination was regularly mon-
concentration is greater than the MDL provided, the calculation is based
itored by blank tests, which were used to validate and correct the
on Eq. (4), by contrast, if it is less than or equal to MDL, the Unc is
corresponding data. Blanks and duplicate sample analyses were con-
calculated using Eq. (5). In this study, all species were categorized as
ducted for nearly 10% of the samples. Certified reference materials
strong due to the signal–to-noise ratio (S/N) > 2, and PM2.5 was spe-
(CRMs, produced by National Research Center for Certified Reference
cified as a total variable (Wang et al., 2016).
Materials, China) were used for QA/QC. The pretreatment procedure,
chemical analysis and QA/QC were described in detail in other related Unc = (Error Fraction × Concentration) 2 + (0.5 × MDL)2 (4)
studies (Liu et al., 2016; Zhang et al., 2015; Zhao et al., 2013).
Unc = 5 6 MDL (5)
2.4. Chemical mass closure

In this study, PM2.5 mass was reconstructed by taking into account 3. Results and discussion
SO42−, NO3−, NH4+, particulate organic matter (POM), EC, trace ele-
ment oxide (TEO) and mineral dust. SO42−, NO3−, and NH4+ were the 3.1. The characteristics of PM2.5 and chemical compositions
major WSIs and their concentrations were directly measured (Chen
et al., 2017; Tao et al., 2013). The POM mass was estimated by the 3.1.1. PM2.5 mass concentration
organic carbon (OC) mass concentration multiplied by a conversion Continuous meteorological parameters, including wind speed, wind
factor of 1.6, as in the following equation: direction, temperature (T) and the relative humidity (RH), were re-
corded simultaneously at the sampling site, which is also the environ-
[POM] = 1.6 × OC (1) mental observation station that was set by the Dushanzi environmental
monitoring station. Fig. 2 shows the time series of the meteorological
The factor has been adopted in many other investigations, because
parameters during the sampling period. The annual wind speed was
the latest result shows an OM/OC ratio that averaged at 1.59 ± 0.18 in
0.73 m s−1, with the highest wind speed occurring in spring
PM2.5 across the China (Wang et al., 2016; Zhang et al., 2013).
(0.96 m s−1) and the lowest occurring in winter (0.39 m s−1). The
The mineral dust component is often estimated through several
temperature was highest in summer (26.51 °C) and lowest in winter
formulas (Zhang et al., 2013). A straightforward method conventionally
(−8.7 °C); in contrast, the relative humidity was highest in winter
used in estimating dust aerosols from Al was used in the present study,
(93.10%) and lowest in summer (43.51%). The annual average con-
as follows:
centration of PM2.5 was 62.85 ± 43.5 μg m−3 at Dushanzi, which was
[Mineral] = Al/0.07 (2) a factor of 2 higher than the Chinese National Ambient Air Quality
Standards' (NAAQS) Grade II of 35 μg m−3, implying serious PM2.5
where 0.07 is the average Al content (7%). A similar estimation had pollution in Dushanzi. The seasonal variations were evident: PM2.5 level
been applied previously (Hsu et al., 2010; Zhang et al., 2013). was highest in winter (95.47 ± 61.7 μg m−3), followed by autumn
For TEO, the contribution of heavy metals as metal oxides was es- (70.22 ± 25.3 μg m−3), spring (56.14 ± 37.1 μg m−3) and summer
timated by the following equation (Chen et al., 2017): (33.22 ± 17.7 μg m−3) (Fig. 2 and Table 1). The seasonality of PM2.5

259
Y. Turap et al. Atmospheric Research 218 (2019) 257–268

Fig. 2. Time series of meteorological records and PM2.5 during the sampling time.

in Dushanzi was similar to the typical pollution character of northwest that is, heterogeneous reactions occurred only on the particle surfaces
China, with the maxima in winter and the minima in summer (Chen (Maria et al., 2004; Virtanen et al., 2010).
et al., 2017; Wang et al., 2016). The difference in the mean con- However, 62.85 ± 43.5 μg m−3 was higher than that of most mega-
centrations between winter and summer was 62.45 μg m−3, which cities in southern China that are located in the subtropical monsoon climate
suggested large seasonal variability. This probably was caused by a zone, such as Guangzhou (42.4 μg m−3), Zhuhai (33.8 μg m−3), and
combination of factors including favorable wind direction and stronger Shenzhen (32.0 μg m−3) (Ming et al., 2017), as well as that of other
atmospheric vertical mixing in the summer and higher emissions from countries, such as Korea (Choi et al., 2012), and Greece (Samara et al.,
coal-fired thermal plants and stronger stable atmospheric boundary 2014). The higher PM2.5 concentration was likely caused by the high level
layer in the winter. of precursor gas from regional industrial (such as power plants and petro-
The annual mean concentration of PM2.5 in Dushanzi was approxi- chemical oil refinery industries) and transportation sources. The other
mately 32% lower than that in Lanzhou (93.7 μg m−3) (Wang et al., reason was probably that high frequency of dust storms could have con-
2016) and 49% lower than in Delhi (122 μg m−3) (Sharma et al., 2016) tributed markedly to the PM2.5 concentration in Dushanzi due to the
because the lower humidity (the annual average humidity was 57%) southwest and northwest wind (Fig. 2) brought dust from the Gobi, Gur-
does not favor the growth of aerosol particuls and PM2.5 formation. The bantongute and Takelamakan deserts (Fig. 1) to Dushanz. The composition
electrical low-pressure impactor (ELPI) and scanning electron micro- and morphology of the dust particles are subject to transformation by ad-
scope (SEM) measurements showed that the atmospheric particulates sorbing gaseous species, surface reactions, and coagulation with other
were an amorphous solid state at low humidity (lower than 70%) and particles during transport and will lead to high levels of geological aerosols;
that the chemical reactions were impeded in glassy or highly viscous the same results are demonstrated in other studies (Di et al., 2016; Wang
aerosol particles because the mass transport of reactants within the et al., 2016; Wei et al., 2005; Zhang et al., 2015). Therefore, local emissions
aerosol particle bulk became a rate-limiting step. For the same reasons, and poor ecological conditions (most of the area is covered by the desert
the chemical reactions in solid particles were typically surface-limited; and Gobi) could both play important roles in the PM2.5 level in Dushanzi.

Table 1
The seasonal and annual mean concentration of PM2.5 and metal elements.
Species Autumn(n = 19) Winter(n = 21) Spring(n = 21) Summer(n = 22) Annual(n = 83)

μg m−3
PM2.5 70.22 ± 25.3 95.47 ± 61.7 56.14 ± 37.1 33.22 ± 17.7 62.85 ± 43.5
Na 1.54 ± 1.0 3.61 ± 1.5 2.09 ± 0.2 3.12 ± 1.1 1.67 ± 0.8
Mg 0.85 ± 0.5 0.55 ± 0.2 1.12 ± 0.6 0.88 ± 0.3 0.84 ± 0.5
Al 3.08 ± 1.7 0.51 ± 0.3 3.24 ± 1.6 1.59 ± 0.8 2.1 ± 1.2
K 1.46 ± 0.9 0.26 ± 0.1 1.14 ± 1.0 0.56 ± 0.3 0.78 ± 0.4
Ca 2.25 ± 1.1 0.41 ± 0.2 1.98 ± 1.7 1.33 ± 0.9 1.51 ± 1.3
Fe 2.24 ± 1.2 0.13 ± 0.8 2.97 ± 1.6 1.21 ± 0.8 1.45 ± 1.4

ng m−3
V 6.39 ± 3.8 0.59 ± 0.3 5.58 ± 3.8 3.12 ± 2.2 4.02 ± 3.0
Cr 5.28 ± 2.8 3.17 ± 2 11.47 ± 8.6 9.4 ± 6.2 7.26 ± 5.7
Mn 58.84 ± 39.4 14.58 ± 9.9 59.84 ± 43.6 39.7 ± 22.6 44.05 ± 37.2
Co 0.92 ± 0.5 0.47 ± 0.3 0.83 ± 0.7 0.52 ± 0.3 0.7 ± 0.5
Ni 10.44 ± 9.6 1.29 ± 0.8 17.61 ± 14.8 16.34 ± 12.6 12.11 ± 9.9
Cu 6.73 ± 2.6 6.22 ± 4.1 5.97 ± 3.0 4.24 ± 3.6 6.07 ± 3.5
Zn 50.91 ± 28.5 24.73 ± 12.7 55.31 ± 30.1 21.45 ± 10.5 38.63 ± 24.8
As 1.91 ± 0.8 4.14 ± 2.6 1.56 ± 0.7 1.48 ± 1.1 2.39 ± 1.9
Rb 5.79 ± 3.4 1.02 ± 0.7 4.8 ± 3.4 2.22 ± 1.8 3.58 ± 2.1
Sr 29.88 ± 15.8 3.48 ± 1.8 27.95 ± 25.3 17.04 ± 11.1 20.85 ± 16.4
Cd 0.27 ± 0.1 0.35 ± 0.1 0.29 ± 0.1 0.11 ± 0.1 0.24 ± 0.1
Ba 37.6 ± 20.8 3.84 ± 2.0 30.09 ± 21.3 17.4 ± 11.0 22.75 ± 18.2
Tl 0.25 ± 0.1 0.18 ± 0.1 0.11 ± 0.1 0.16 ± 0.1 0.15 ± 0.1
Pb 15.59 ± 8 17.51 ± 10.1 14.54 ± 6.2 15.28 ± 11.3 15.28 ± 10.5
SUM/PM2.5 0.16 ± 0.06 0.04 ± 0.03 0.21 ± 0.05 0.23 ± 0.06 0.16 ± 0.09

260
Y. Turap et al. Atmospheric Research 218 (2019) 257–268

3.1.2. Metal elements (which could have occurred from the deposition during the early years)
The annual and seasonal mean concentrations of airborne metal ele- on the ground of Xinjiang (including road dust and the agriculture
ments are listed in Table 1. For the annual mean concentration, the ele- field), which could be transported into the atmosphere, and found that
ments contributed to 16.01 ± 9.0% of the PM2.5 mass, higher than the surface soil contains a higher concentration of Pb, As and Cd
Lanzhou (8.8 ± 2.9%) (Wang et al., 2016) and Heze (12.5%) (Liu et al., compared to their crustal abundance, which indicated that these ele-
2017). The most abundant elements are crustal elements (e.g., Na, Mg, Al, ments in the aerosols might come not only from the re-suspended road
K, Ca and Fe), which account for 96.51 ± 1.3% of the total analyzed dust but also the from outside transported soil from the Jungger Basin.
elements, and this percentage was higher than that found for Lanzhou In this study, trace elements including Cd, Pb, As and Zn, have the
(93.6 ± 4.5%) (Wang et al., 2016), and Xinxiang (90.8%) (Feng et al., highest EF values in each season, especially in winter (276.30, 129.81,
2016). Crustal elements waxed in the spring and autumn and waned in the 57.72 and 54.24), which was probably due to coal combustion for
winter. Such levels of seasonal mean concentrations and seasonality are domestic heating promoted the emission of metals (such as Pb, As, Sb
consistent with those that were observed in previous studies (Wang et al., and Cr) in winter. In other seasons, it might have originated from
2016; Xu et al., 2016), and are related to dust storms and anthropogenic mixing the local anthropogenic aerosol with the transported soil dust.
fugitive dust. The trace elements (V, Cr, Mn, Ni, Cu, Zn, As, Rb, Sr, Ba and Hence, these elements in PM2.5 in Dushanzi were related to human
Pb) are present at levels between 2.4 and 44.0 ng m−3. The seasonal levels activities, such as coal combustion, vehicle emission, biomass burning
of elements decreased in the following order: nearly equal in spring and industry, and were also impacted by dust particles transported from
(11.68 ± 8.8 μg m−3) and autumn (11.63 ± 6.0 μg m−3) > summer the desert.
(9.54 ± 3.4 μg m−3) > winter (2.51 ± 1.0 μg m−3), and contributing to
the PM2.5 mass concentrations were 21.12 ± 5.0%, 23.21 ± 6. 1%, 3.1.3. Water-soluble ions
16.06 ± 6.0% and 4.14 ± 3.2% for spring, summer, autumn and winter, Table 2 shows the mean concentrations of the WSIs in four seasons
respectively. Shi et al., 2017 reported that the crustal elements mainly in Dushanzi. The mass concentrations of them followed the order
originated from re-suspension, and snow can reduce the re-suspension. In SO42− > NO3− > NH4+ > Ca2+ > Na+ > Cl− > K+ > Mg2+
this study, the wind speed was the lowest in winter (Fig. 2), and Na and during all seasons, and the total WSIs accounted for 20.12%, 64.20%,
Cocker, 2009 reported that dry, hot weather could facilitate re-suspension 21.17% and 16.18% of PM2.5 in autumn, winter, spring and summer,
of soil-related material by wind; and crustal element concentrations were respectively. It is worth noting that the WSIs contributions in winter
higher by approximately a factor of 2 for September and October compared were approximately 3 times higher than that in other seasons. This
with December and January. Therefore, the re-suspension was reduced and might be in north China where coal-based heating systems and others
the concentrations of crustal elements were the lowest in winter. However, combustion in winter have become the dominant sources of SO2, NOx
heavy metals including As, Pb, Cu and Zn were the highest in winter due to and primary PM (Song et al., 2015), and the domestic heating in Xin-
coal burning and thermal power plants, which demonstrated that coal jiang occurs earlier (usually from 15 October to 15 March the following
burning was an important source of As and Pb (Wang et al., 2016). The year) than in other cities in China, and the six months of heating ac-
other trace elements in PM2.5 in Dushanzi can be traced back to vehicle cumulated the precursor gas of PM. The stable atmosphere (low wind
exhaust, industrial emissions and dust storms (due to the large standard speed) and low temperature (Fig. 2) in winter might also increase the
deviation). nucleation and growth, as well as accelerate the adsorption of aerosol in
Enrichment factor (EF) is a quantity that distinguishes anthro- Dushanzi, these results were consistent with those of earlier studies
pogenic sources from natural process and assesses the degree of ele- (Chen et al., 2017; Ming et al., 2017; Zhang et al., 2013).
ments pollution; it is estimated by the Eqs (Hsu et al., 2016; Shao et al., Among these, WSIs such as NH4+, NO3− and SO42− were the
2016). dominant species and accounted for 78.55 ± 13.3% of the total ionic
concentration in Dushanzi. The annual mean concentration of NH4+,
EF = (Ci /Cr )sample /(Ci /Cr )crust (6)
NO3− and SO42+ were 5.03 ± 8.9, 6.75 ± 10.3 and
where (Ci/Cr)sample represent the ratio of the target element to the re- 11.94 ± 21.1 μg m−3, respectively, which constituted 5.67 ± 6.0%,
ference element mass concentration in aerosol and (Ci/Cr)crust represent 8.07 ± 7.6% and 14.0 ± 12.6% of the PM2.5 mass respectively, which
the ratio of the target element to the reference element in the upper was lower than in Beijing (Zhang et al., 2013), Shanghai (Ming et al.,
continental crust (Clemens Reimann and Caritat‡, 2000; Taylor and 2017) and Chongqing (Chen et al., 2017). However, the levels of NH4+,
Mclennan, 1995). The crustal element concentration was extracted NO3− and SO42− in winter were factors of 8 to 10 higher than those in
from previous studies that investigated a region near the Dushanzi other seasons. Other WSIs, such as Ca2+, Mg2+ and Na+, were ob-
district (Luo et al., 2012; Zhao and Fan, 1994). The typical reference served in higher concentrations in spring and autumn than in other
elements include Al, Fe, Ti and Mg, and there is still no well-established seasons, due to dust storms frequently occurring in spring and autumn
rule for the choice of a reference element (Khare and Baruah, 2010). In in Dushanzi (Wang et al., 2016; Wei et al., 2005; Xu et al., 2016). The
this study, Al was chosen for the reference element due to its abundance highest K+ and Cl− concentrations were observed in the autumn,
in the earth's crust and its stability to most anthropogenic contaminants which was likely due to agricultural burning after the harvest around
(Hsu et al., 2016). Dushanzi, which was also found by some previous studies in Beijing (He
Fig. 3 shows the seasonal EF values; elements such as Al, Fe, Mn, Rb, et al., 2017; Zhang et al., 2013). Generally, the NO3−/SO42− ratios
K, Ga were lower than 10 during the sampling period, which indicated were used as a relative measure of the importance of mobile versus
that these elements originated mostly from mineral dust. In contrast, in stationary pollution emission sources in many studies (Xu et al., 2016).
autumn, the EF values for Tl, Zn, Pb and Cd were above 10. The EF The high NO3−/SO42− ratio was found in vehicle-predominant regions.
values were higher than 10 in winter for Cr, Cu, Tl, Zn, As Pb and Cd, in However, lower ratios < 1 have been found, usually in China, because
spring for Cr, Pb, Ni, Zn and Cd and in summer for As, Tl, Ni, Cu, Cr, Zn, of the wide use of sulfur-containing coal (Feng et al., 2009). In the
Pb and Cd. On average, trace metals such as Zn, Pb and Cd have high EF present study, the NO3−/SO42− ratios ranged from 0.14 to 1.12, and
values in all seasons, which suggested that they were mostly from an- the annual mean ratio was 0.61 ± 0.4, which was lower than that in
thropogenic sources (Hsu et al., 2016). Generally, EF values could be cities with heavy traffic density, such as Guangzhou (0.79) (Tan et al.,
used to source tracers (Hsu et al., 2016). For instance, high EF values of 2009), but higher than that in coal-dominated cities, such as Guiyang
Cr, Pb, As and Cd were associated with coal combustion, high levels of (0.14) and Qingdao (0.35) (He et al., 2017), suggesting that the
Cu and Zn have demonstrated a link to vehicle emission and abundant stronger influences come from stationary pollution emission sources,
Ni and V originated from oil combustion (Zhai et al., 2014; Zhao et al., such as coal combustion; however, the mobile sources should not be
2015). Moreover, Li et al., 2008 collected the surface soil samples ignored.

261
Y. Turap et al. Atmospheric Research 218 (2019) 257–268

Fig. 3. Enrichment factors of elements in four seasons.

Understanding the acidity of the PM is significantly important due


to it being a key parameter that affects heterogeneous reactions, hy-
groscopic growth, and toxicity of atmospheric PM, as well as the neu-
tralizing process of acid rain (Shi et al., 2017; Tao et al., 2013).
Ion balance calculations were useful for studying the acid-base
balance of aerosol. The ratios of AE (anion equivalent) and CE (cation
equivalent) could also be used to indicate the acidity of atmospheric
aerosol. The calculations of particulate anion and cation equivalent are
as follows (Shi et al., 2017; Tao et al., 2013):

AE = [NO3 ]/62 + [SO4 2 ]/48 + [Cl ]/35.5 (7)

CE = [NH 4+]/18 + [Ca2 +]/20 + [K+]/39 + [Mg2 +]/12 + [Na+]/23 (8)

where [NO3−], [SO42−], [Cl−], [NH4+], [Ca2+], [K+], [Mg2+] and


[Na+] represent the mass concentrations of these ion species in the
PM2.5 samples. Previous studies on an AE/CE ratio that was larger than
1 had been used to indicate the acidic particle condition, and a ratio
lower than 1 suggests an alkaline condition (Shi et al., 2017; Tao et al.,
2013), Fig. 4. Scatter plots of AE (anion equivalent) vs CE (cation equivalent) in four
In the present study, strong correlations (R2 = 0.98) between AE seasons.
and CE were found for all of the samples (Fig. 4), and the average AE/

Table 2
The seasonal and annual mean concentration of water-soluble ions (μg m−3).
Species Autumn(n = 19) Winter(n = 21) Spring(n = 21) Summer(n = 22) Annual(n = 83)

Na+ 0.48 ± 0.2 0.51 ± 0.3 0.34 ± 0.2 0.34 ± 0.2 0.47 ± 0.5
NH4+ 1.73 ± 1.2 15.32 ± 12.1 1.15 ± 0.8 0.47 ± 0.2 5.03 ± 8.9
K+ 0.30 ± 0.2 0.17 ± 0.2 0.11 ± 0.1 0.10 ± 0.02 0.17 ± 0.1
Mg2+ 0.13 ± 0.1 0.02 ± 0 0.08 ± 0.01 0.06 ± 0.03 0.07 ± 0.1
Ca2+ 2.02 ± 0.9 0.09 ± 0.06 1.27 ± 1.0 0.74 ± 0.5 1.02 ± 1
Cl− 0.75 ± 0.6 0.05 ± 0.1 0.25 ± 0.3 0.10 ± 0.1 0.29 ± 0.5
NO3− 4.45 ± 3.5 17.97 ± 14.1 2.47 ± 2.1 0.58 ± 0.3 6.75 ± 10.3
SO42− 4.38 ± 1.8 33.63 ± 31.5 4.25 ± 2.6 2.47 ± 0.8 11.94 ± 21.1
NO3−/SO42− 0.92 ± 0.4 0.66 ± 0.3 0.65 ± 0.5 0.24 ± 0.1 0.61 ± 0.4
SUM/PM2.5 0.20 ± 0.04 0.64 ± 0.2 0.21 ± 0.1 0.16 ± 0.03 0.31 ± 0.2

262
Y. Turap et al. Atmospheric Research 218 (2019) 257–268

Table 3
The seasonal and annual mean concentration of carbonaceous species in PM2.5.
μg m−3 Autumn(n = 18) Winter(n = 19) Spring(n = 18) Summer(n = 18) Annual(n = 73)

OC 5.67 ± 2.1 8.76 ± 4.3 6.46 ± 2.6 7.53 ± 3.0 7.13 ± 3.3
EC 1.73 ± 0.8 1.37 ± 0.6 0.90 ± 0.4 0.97 ± 0.5 1.25 ± 0.7
TCA 10.81 ± 3.9 15.39 ± 7.2 11.23 ± 4.4 13.01 ± 5.0 12.55 ± 5.6
OC/EC 3.48 ± 0.8 6.27 ± 1.2 7.47 ± 1.5 8.45 ± 2.5 6.44 ± 2.5
SOC 0.98 ± 0.8 3.37 ± 2.1 1.87 ± 1.0 2.19 ± 1.2 2.0 ± 1.6
SOC/OC 0.25 ± 0.1 0.35 ± 0.2 0.30 ± 0.1 0.32 ± 0.2 0.31 ± 0.2
TCA/PM2.5 0.16 ± 0.03 0.35 ± 0.1 0.29 ± 0.1 0.44 ± 0.2 0.34 ± 0.2

CE for all samples was 0.87 ± 0.2, which indicated that the collected
PM was slightly alkaline. This finding agreed with that of other studies
in northern China (Shi et al., 2017; Tao et al., 2013). On the other hand,
the anion and cation equivalents were strongly correlated in four sea-
sons (autumn: R2 = 0.91, winter: R2 = 0.99, spring: R2 = 0.94 and
summer: R2 = 0.95). The AE/CE ration revealed a typical seasonality
with larger than 1 for winter and lower than 1 for other seasons. For
autumn and spring, the AE/CE ration was almost equal to the 1:1 line,
and the particles were closer to neutral because of the significant in-
crease in alkaline compounds from the sand storms. In summer, almost
all of the samples were under the 1:1 line, which might have been re-
lated to carboxylic and phosphate that were not detected in this study,
and would be higher with higher air temperature and a stronger pho-
tochemical reaction (He et al., 2017). In winter, 74% samples were
above the 1:1 line, which illustrated that the particles were acidic as a
result of the highest levels of acid compounds (NO3− and SO42−).

3.1.4. Carbonaceous aerosols


Table 3 provides a summary of the seasonal and annual average level of Fig. 5. Scatter plot between OC and EC in four seasons in Dushanzi.
carbonaceous aerosols in Dushanzi. The annual mean concentrations of OC
and EC in PM2.5 were 7.13 ± 3.3 and 1.25 ± 0.7 μg m−3, which account- indirect method to estimate the SOC, the EC tracer method formation
for PM2.5 17.47 ± 10% and 2.72 ± 2.7%, respectively. The levels of OC with the following equation proposed by Turpin and Huntzicker (1995),
and EC were consistent with Shanghai (9.89 ± 8.9 μg m−3 for OC and Castro et al. (1999):
1.63 ± 1.5 μg m−3 for EC) (Ming et al., 2017). Seasonal levels of OC con-
centrations decreased in the following order: winter (8.76 ± 4.3 μg m−3) SOC = OC (OC/EC) pri × EC (9)
> summer (7.53 ± 3.0 μg m−3) > spring (6.46 ± 2.6 μg m−3) > autumn
(5.67 ± 2.1 μg m−3). However, the seasonal variation of EC was different, The minimum OC/EC ratio during the sampling periods is often
assumed as the primary OC/EC ratio ((OC/EC)pri). The estimated SOC
the EC concentration was the highest for autumn (1.73 ± 0.8 μg m−3), and
the levels were similar for spring (0.90 ± 0.4 μg m−3) and summer concentrations in this study are presented in Table 3. The annual mean
concentration of SOC in PM2.5 was 2.0 ± 1.6, with the seasonal var-
(0.97 ± 0.5 μg m−3). The seasonal variation of EC was different from other
cities in China, such as Lanzhou (the highest for winter) (Tan et al., 2017), iation in winter > summer > spring > autumn. The seasonality may be
governed by the variability in emission strengths (OC was the lowest
possibly due to the big farm in southwest Dushanzi and the biomass burning
after the harvest in autumn influencing the EC emission. It was demonstrated and EC was the highest in autumn) and meteorology. The SOC/OC
ratios were higher in winter (0.35 ± 0.2) and summer (0.32 ± 0.2)
that OC can either release primary OC or secondary origins (formed through
photochemical conversion), while EC is essentially a primary pollutant from than in spring (0.30 ± 0.1) and autumn (0.25 ± 0.1). These results
were consistent with previous observations in other cities in China,
combustion (Tan et al., 2017). Therefore, the discrepancy of OC and EC in
Dushanzi had much to do with their seasonal sources, chemical transforma- which showed high values of SOC and SOC/EC in winter, due to the
coal combustion accelerating the emission of carbon aerosols and the
tions and meteorology.
A number of scholars have investigated the OC/EC ratio as a useful low temperatures increasing the adsorption of volatile organic com-
pounds (VOCs) and SOC yields (Tao et al., 2013; Wang et al., 2017;
tool for identifying the sources of carbonaceous aerosols (Chen et al.,
2017; Koch, 2001; Wang et al., 2016; Watson et al., 2001). As shown in Wang et al., 2016).
The total carbonaceous aerosol (TCA) was calculated by the sum of
Table 3, the OC/EC ratios in Dushanzi were higher in summer
(8.45 ± 2.5) than those in spring (7.47 ± 1.5) winter (6.27 ± 1.2) EC and organic matter, which was estimated by multiplying the amount
of OC by 1.6 to assess the contribution of carbonaceous aerosols to the
and autumn (3.84 ± 10.5), indicating the source variations of carbo-
naceous aerosols were vague because the ratios partially overlapped. PM mass (Ming et al., 2017). TCA accounted for an average of 34% of
PM2.5, with the seasonal rank of winter > summer > spring > autumn.
Strong correlation, however, were shown between OC and EC (Fig.5),
with an R2 ranging from 0.80 to 0.88, which suggested that OC and EC The ratios were higher than those of other cities, such as Tianjin (16%),
Zhejiang (17%), Guangdong (23%) and Sichuan (20%) (Zhou et al.,
had common sources, such as coal combustion, vehicle emission and
2016). It was suggested that carbonaceous aerosol was an important
biomass burning. Chow et al., 1996 demonstrated that OC/EC ratios
fraction of PM2.5 in Dushanzi.
exceeding 2.0 may indicate the presence of secondary organic matter
(SOC). Interestingly, in this study, the ratios were higher than the
standard. It is difficult to quantify and separate primary and secondary 3.2. Chemical mass closure
OC due to the vast number of organic compounds with a variety of
chemical and physical properties. However, it is possible to use an The PM2.5 mass was reconstructed on a seasonal and annual basis by

263
Y. Turap et al. Atmospheric Research 218 (2019) 257–268

Fig. 6. Correlation between reconstructed and measured and chemical mass closure of PM2.5 in four seasons and annually.

the methods described in Sect. 2.3. The gravimetric PM2.5 mass con- (Wang et al., 2016), which indicated more dust pollution in Dushanzi.
centration was compared with the reconstructed PM2.5 mass con- The relatively high level of mineral dust was likely, because it was
centration, as shown in Fig. 6(a), which show a good correlation surrounded by the desert and Gobi, and a low vegetation cover. Dust
(higher than 0.90) with one another in each season and throughout the was lower in the winter because of the reduction of dust emissions by
year. The proportions of all components in PM2.5 together with the snow (Shi et al., 2017). The secondary inorganic species (SO42−, NO3−
unidentified constituents as a whole are schematically illustrated by and NH4+) have the largest proportion in winter (53.25%) and lower
five pie charts for the four seasons and for the whole year. percentage during other seasons. The maximal value of SO42−, NO3−
Collectively, PM2.5 in the Dushanzi district was composed of mi- and NH4+ were all noted in winter 26.24%, 14.58% and 12.43%, re-
neral dust 33.58%, TEO 0.20%, POM 14.87%, EC 1.56%, NH4+ 4.58%, spectively. The POM fraction largely varied as follows: summer
NO3− 6.18%, SO42− 11.07% and an unidentified constituent 27.95% (25.40%) > winter (12.97%) > spring (12.70%) > autumn (9.26%).
(Fig. 6(f)). The major components were mineral dust, secondary in- EC and TEO exhibited the largest proportion 2.02% and 0.28% in
organic aerosols (a combination of SO42+, NO3− and NH4+) and POM, summer, which were lower than those in Beijing and Chongqing (Chen
each of which was > 15%, albeit with seasonal variations. Specifically, et al., 2017; Zhang et al., 2013). The unidentified components reached
the proportions of mineral dust showed a notable seasonal change, with 27.95% of the total PM2.5. They showed seasonal variability with the
the maximum in spring (50.17%), the minimum in winter (only 7.14%) smallest (22.64%) in summer and the largest (32.20%) in autumn. Such
and an intermediate (40.90% and 41.56%) in autumn and summer, a high unidentified component was demonstrated in autumn in Beijing
which were much higher than those in Beijing (23.5%) (Zhang et al., (33.9%) due to the volatilization of some components of PM2.5, espe-
2013), Chongqing (8.2%) (Chen et al., 2017) and Lanzhou (30.3%) cially in autumn and summer during the storage of the weighted

264
Y. Turap et al. Atmospheric Research 218 (2019) 257–268

Fig. 7. Backward trajectory clusters and their chemical components by season, at 100 m height located at 44.19°N, 84.53°E.

samples before extraction, which may have led to negative biases in the concentration. All of the chemical compounds except mineral had ap-
specific components (Zhang et al., 2013). proximately similar distributions in each air mass.
Above all, the seasonality of increased proportions of PM2.5 in In winter, three major air masses influenced the Dushanzi district,
Dushanzi was as follows: mineral in autumn and spring; secondary in- and they originated mainly from two areas, including the northwest,
organic species in winter; POM, TEO and EC in summer. south and southwest. Air mass 1 carried the highest concentrations of
PM2.5 (138.16 μg m−3), including a high proportion of SO42−, NO3−
3.3. Sources apportionment analysis and NH4+ (which probably originated from coal combustion and ve-
hicle emission), and arrived via a short pathway from Kuitun city. Air
Previous studies have reported that the regional sources and transport of mass 2 reached the sampling site via a long pathway from the south
air pollutants exert a profound impact on the local air quality (Wang et al., direction, thus loading the highest SO42−. Air mass 3 originated from
2004). To address this issue, the Hybrid Single Particle Lagrangian In- the neighboring countries, and the concentration of loaded PM2.5 was
tegrated Trajectory (HYSPLIT) model was used to determine the transport 78.57 μg m−3, which suggested that neighboring countries also influ-
trajectory of air parcels at the sampling site. The 48 h back trajectories were enced the air quality of the Dushanzi district.
conducted every 6 h, at 100 m above the ground. Fig. 7 shows that the In spring, the trajectories originated from two areas of the Dushanzi
PM2.5 concentrations were related to the size of the pie charts and the re- district. The majority arrived on air mass 1 (consisting of 86% the
lative chemical components by season. trajectories), which originated from the southwest direction. Air mass 2
In autumn, six major air masses influenced the Dushanzi district. Air started from Kazakhstan, which loading the highest mineral con-
masses 1 and 3 originated from the southwest (SW) direction and air centration (54%) and was consistent with that of autumn. Other com-
mass 3 consisted of 71% of the trajectories. The SW cluster was further pounds that were in proportion in this air mass was similar to those of
differentiated into two types; i. e., fast (SWf, air mass 1) and slow (SWs, air mass 1, which indicated that the concentrations of elements in PM2.5
air mass 3), according to the distance of air parcels. Air masses 2 and 6 in the Dushanzi district were greatly impacted by dust particles that has
arrived from the northwest direction, from Kazakhstan. Among them, been transported from the desert and were in line with Section 3.1.2.
air mass 6 carried the highest PM2.5 mass concentration (83.24 μg m−3) In summer, three air masses impacted the Dushanzi district atmo-
and the highest mineral concentration (60.79%), which demonstrated sphere. The highest PM2.5 concentration (36.04 μg m−3) was loaded by
that the northwest might be a potential origin of the crustal dust and air mass 1. Air mass 2 and air mass 3 started from northwest and
the long pathway promotes the absorptivity of PM2.5. Air masses 4 and southwest direction, respectively. The PM2.5 concentrations loaded by
5 arrived from north direction, providing the lowest PM2.5 three air masses were less than or close to the NAAQS limit, indicating

265
Y. Turap et al. Atmospheric Research 218 (2019) 257–268

that the transport air masses were relatively clean or loaded low pol- Table 4
lution compounds. In addition the three air masses were approximately Relative contributions from six identified sources of PM2.5 in Dushanzi within
similar in their distributions of chemical components. the one-year and four-season periods.
In short, the Dushanzi district atmosphere was influenced by four air Source Autumn Winter Spring Summer Annual
masses (SWf, SWs, W and NW), which accounted for 38.17%, 24.73%,
19.35% and 17.75%, respectively. Kazakhstan (via long pathway vehicular 4.76% 13.94% 9.21% 8.22% 9.43%
biomass 18.85% 8.26% 18.93% 15.20% 10.86%
transport) impacted the mineral distribution, and Kuitun city impacted
coal 21.35% 36.67% 17.59% 7.79% 18.45%
the WSI in the PM2.5. Most of the air mass came from the desert and industry 15.43% 5.47% 9.24% 18.35% 12.15%
Gobi areas, which indicated that the desert significantly influenced the secondary 20.68% 30.90% 6.62% 23.45% 18.26%
PM2.5 concentration in Dushanzi, and part of the air masses via a short soil 18.94% 4.81% 38.23% 26.98% 30.85%
pathway transport arrived at the sampling site, which further indicated
that local pollutant emissions mixed with the mineral aerosols on the
pathway were the main source of the PM in Dushanzi. natural gas engines in Xinjiang. It was also worth noting that the con-
By utilizing the PMF model with the obtained full data set as the centrations of Ca and Mn were relatively high, which suggested that
input date, six main sources were identified: vehicular emission, bio- road dust could be an important contributor to this factor due to the
mass burning, coal combustion, industrial pollution, secondary aerosols extremely dry climate in Dushanzi. Therefore, the factor can be re-
and soil dust. Their mean contributions to PM2.5 in the Dushanzi district garded as a mixed source. The contribution of vehicular emission to
were 9.43%, 10.86%, 18.45%, 12.15%, 18.26% and 30.85%, respec- PM2.5 was 9.43%, with a maximum in winter (13.94%) and a minimum
tively. Fig. 8 shows the source profiles, and Table 4 summarizes the in autumn (4.79%). The result is similar to the contribution of traffic
source apportionment results of the relative contributions of the in- emissions to the PM2.5 mass in Chongqing (Chen et al., 2017).
dividual source to the PM2.5 on both seasonal and annual basis in the The second factor represented biomass burning characterized by a
Dushanzi district. high proportion of K and Tl. K is an excellent tracer of biomass burning
The first factor was characterized by high proportions of Cu, Zn, Cd, aerosols (Zhang et al., 2013). On average, this factor accounted for
Pb, Mg and OC in PM2.5 featured by vehicular emission. Previous stu- 10.86% of the PM2.5 mass concentration in the Dushanzi district, which
dies have reported that traffic-generated aerosol particles are rich in was higher than Lanzhou (Wang et al., 2016). Table 4 shows that bio-
OC, Cu, Zn and Pb (Chen et al., 2017). OC is the main pollutants from mass burning has higher contributions in autumn, spring and summer
gasoline and diesel combustion, Zn is used as an additive in lubricating than in winter. Zhang et al., 2010 reported that the contribution of
oil and Pb is linked to metal brake wear (Yao et al., 2016). In this factor, biomass varied with seasons; however, they were based on data for only
the proportion of Cd and Mg were relatively high, and may be emitted a few months at a limited number of sites and were further restricted by
from natural gas engine combustion due to many cars whit gasoline/ uncertainties in biomass burning emissions and tracer concentrations.

Fig. 8. Factor profiles (% of species total) obtained from the PMF analysis in Dushanzi.

266
Y. Turap et al. Atmospheric Research 218 (2019) 257–268

In this study, the contribution of biomass was not different in autumn, in arid and semi-arid areas (Ganor and Foner, 2001; Ganor et al., 2010),
spring and summer, which was probably due to open field burning of and dust is a major source of global atmospheric particulate matter
agricultural waste after harvesting in autumn and cultivation in spring (PM) (Shao et al., 2011). In this study, the seasonal contribution of soil
and summer enhanced the unorganized burning in agriculture (Zhang dust was the highest in spring (38.23%), which was consistent with the
et al., 2013). Moreover, there is a large farm in the southwest sampling variation of mineral in Sec. 3.2 and the results of the backward tra-
site, and the transport trajectories (Fig. 7) also indicated that the jectory clusters of air mass originated from the desert direction and
southwest area was a potential source region. frequent occurrence of dust storms in Xinjiang.
The third factor was coal combustion, which was characterized by
high level of Na, V, As, OC and Cl−, and it was reported that As was 4. Conclusions
proposed to be a more useful tracer of coal combustions (Hsu et al.,
2016). In addition, there was a large coal-fired power plant in Dush- Ambient PM2.5 and its major components were measured in four
anzi, and a large consumption coal every year strengthened the atmo- seasons during the period 2015–2016 in Dushanzi, northwest China.
spheric As and V emission. Different seasons have significantly different The annual mean concentration of PM2.5 was 62.85 ± 43.5 μg.m−3,
contributions of coal combustion to PM2.5, which ranged from 7.79 to which was much higher than that established for Grade II (35 μg.m−3)
36.67%, with a mean contribution of 18.45%. The portion of coal of the Chinese National Ambient Air Quality Standards' (NAAQS). The
combustion was higher in winter and lower in summer because do- seasonal averages of PM2.5 in winter were a factor of 3 higher than
mestic heating would aggravate the emission in wintertime. Such sea- those in summer. Distinctive seasonality occurred for various aerosol
sonality was also found by some previous studies (Chen et al., 2017; species. The highest and lowest concentrations of dominant chemical
Wang et al., 2016). compositions were observed in winter and summer, and the highest
The fourth factor was characterized by the high level of Ni Co and mineral elements were observed in spring, summer and autumn. The
OC. Some previous studies found that Ni could be from oil burning in enrichment factor showed that elements such as Ni, Mg, Cr, Cu, Ba, Zn,
industrial oil boilers or no-ferrous smelters (Chen et al., 2017; Dall'Osto As, Pb and Cd were attributable to anthropogenic emission. SO42−,
et al., 2013; Hsu et al., 2016). Although Co and OC represented the coal NO3−and NH4+ were the dominant species, which accounted for
combustion and vehicle emission, Zhang et al. (2013) reported that coal 78.55% of the total ionic concentration, and the ion balance indicated
was the primary energy source that is commonly used industrially in that the collected PM in Dushanzi was slightly alkaline. Carbonaceous
China, and carbonaceous aerosols were also a vital source of the in- aerosol was an important component in PM2.5 in Dushanzi, with the
dustrial pollution. There are large oil refineries and petrochemical in- TCA accounting for 34.12% of PM2.5. CMCs were successfully con-
dustries in Dushanzi; therefore, industrial pollution was a main factor. structed on a seasonal basis, although with an unidentified fraction that
On average, industrial pollution accounted for 12.5% of the PM2.5 mass, was averaged at 27.95%. The major aerosol speciation that was con-
with a higher percentage in summer (18.35%) due to the volatilization sidered in this study included mineral dust, POM, and SIA. The tra-
of chemical solvent in high temperature. Moreover, the trajectories jectory clustering method and the PMF model showed that the mixing
(Fig.7) showed that the airflow originated from the north oil refinery of the anthropogenic aerosol sources (vehicle emission, biomass
and petrochemical industry area. burning, coal combustion, industrial pollution, secondary aerosols)
The fifth factor was recognized as secondary aerosols, as it had a with the transported soil dust were the main sources of PM in Dushanzi.
high content of SO42−, NO3−, NH4+ and OC. Secondary products were
formed by photochemical or other chemical reaction processes. Their Acknowledgments
levels were affected by precursor gases (e.g., SO2 and NOx) and atmo-
spheric conditions (e.g., humidity, temperature, solar radiation and This work was supported by the National Natural Science
insolation). This factor contributed to 18.26% during the sampling Foundation of China (No. 41465007) and State Key Laboratory of
period. Seasonality percentages of the secondary aerosols were the Organic Geochemistry, GIGCAS (No. SKLOG-2016201624).
highest in wintertime (30.90%) and were the lowest in springtime
(6.62%). These finding were distinct from those of other previous stu- References
dies, in which summer often had the highest percentage of secondary
aerosols (Chen et al., 2017; Zhang et al., 2013). However, the results Castro, L.M., Pio, C.A., Harrison, R.M., Smith, D.J.T., 1999. Carbonaceous aerosol in
were consistent with Sec. 3.1.3, which had the highest concentration of urban and rural European atmospheres: estimation of secondary organic carbon
concentrations. Atmos. Environ. 33, 2771–2781.
NH4+, NO3− and SO42+ in winter due to the higher emission of pre- Cesari, D., Donateo, A., Conte, M., Merico, E., Giangreco, A., Giangreco, F., Contini, D.,
cursor gas in winter. 2016. An inter-comparison of PM2.5 at urban and urban background sites: Chemical
The sixth factor was relevant to soil dust, which was typically characterization and source apportionment. Atmos. Res. 174-175, 106–119.
Chen, Y., Xie, S.-d., Luo, B., Zhai, C.-z, 2017. Particulate pollution in urban Chongqing of
characterized by remarkable crustal elements, such as Al, Ca, Mg, Fe southwest China: Historical trends of variation, chemical characteristics and source
and K. The Ca and Mn contents were higher than Al, which indicated a apportionment. Sci. Total Environ. 584-585, 523–534.
Ca- and Mn-rich dust. It was reported that Ca was the indicator element Choi, J.K., Heo, J.B., Ban, S.J., Yi, S.M., Zoh, K.D., 2012. Chemical characteristics of PM
2.5 aerosol in Incheon. Korea. Atmos. Environ. 60, 583–592.
for construction dust (Zhang and Iwasaka, 1999), and construction Chow, J.C., Watson, J.G., Lu, Z., Lowenthal, D.H., Frazier, C.A., Solomon, P.A., Thuillier,
activities were prevalent in Dushanzi in recent years. Interestingly, R.H., Magliano, K., 1996. Descriptive analysis of PM 2.5 and PM 10 at regionally
there was also the presence of V, Cr, Rb, Sr, Ba and Cl−, of which the representative locations during SJVAQS/AUSPEX. Atmos. Environ. 30, 2079–2112.
Clemens Reimann, A., Caritat‡, P.D., 2000. Intrinsic flaws of element enrichment factors
contents were close to 30%. Both Rb and Sr with low EF values were
(EFs) in environmental geochemistry. Environ. Sci. Technol. 34, 5084–5091.
identified as nature sources in Sec. 3.1.2 (Fig. 4). Because the location Dall'Osto, M., Querol, X., Amato, F., Karanasiou, A., 2013. Hourly elemental concentra-
of our study was near to the Gurbantongute desert and Takelamakan tions in PM2.5 aerosols sampled simultaneously at urban background and road site
desert and the geomorphological type of the Dushanzi district is pre- during SAPUSS – diurnal variations and PMF receptor modelling. Atmos. Chem. Phys.
13, 4375–4392.
dominantly Gobi desert, the heavy wind-blown soil dust from the desert Di, A., Xue, Y., Yang, X., Leys, J., Jie, G., Mei, L., Wang, J., She, L., Hu, Y., He, X., 2016.
likely contributed to these crustal elements during the sampling period. Dust aerosol optical depth retrieval and dust storm detection for xinjiang region using
Therefore, this source possibly was mixed with resuspended road dust, indian national satellite observations. Remote Sensing 8, 702.
Feng, Y., Chen, Y., Guo, H., Zhi, G., Xiong, S., Li, J., Sheng, G., Fu, J., 2009.
fugitive dust, anthropogenic construction dust and desert/Gobi dust. On Characteristics of organic and elemental carbon in PM2.5 samples in Shanghai.
average, this factor accounted for 30.85% of the PM2.5 mass con- China. Atmos. Res. 92, 434–442.
centration in the Dushanzi district, which was higher than in other ci- Feng, J., Yu, H., Su, X., Liu, S., Li, Y., Pan, Y., Sun, J.-H., 2016. Chemical composition and
source apportionment of PM 2.5 during Chinese Spring Festival at Xinxiang, a heavily
ties in China (Chen et al., 2017; Wang et al., 2016; Zhang et al., 2013). polluted city in North China: Fireworks and health risks. Atmos. Res. 182, 176–188.
Previous studies reported that dust storms are a common phenomenon

267
Y. Turap et al. Atmospheric Research 218 (2019) 257–268

Ganor, E., Foner, H.A., 2001. Mineral dust concentrations, deposition fluxes and de- Guangzhou. Atmos. Res. 94, 238–245.
position velocities in dust episodes over Israel. J. Geophys. Res. 106, 18431–18438. Tan, J., Zhang, L., Zhou, X., Duan, J., Li, Y., Hu, J., He, K., 2017. Chemical characteristics
Ganor, E., Osetinsky, I., Stupp, A., Alpert, P., 2010. Increasing trend of African dust, over and source apportionment of PM2.5 in Lanzhou. China. Sci. Total Environ. 601–602.
49 years, in the eastern Mediterranean. J. Geophys. Res. Atmos. 115, 2232. Tao, J., Zhang, L., Engling, G., Zhang, R., Yang, Y., Cao, J., Zhu, C., Wang, Q., Luo, L.,
He, Q., Yan, Y., Guo, L., Zhang, Y., Zhang, G., Wang, X., 2017. Characterization and 2013. Chemical composition of PM2.5 in an urban environment in Chengdu, China:
source analysis of water-soluble inorganic ionic species in PM 2.5 in Taiyuan city. Importance of springtime dust storms and biomass burning. Atmos. Res. 122,
China. Atmos. Res. 184, 48–55. 270–283.
Hsu, S.C., Liu, S.C., Arimoto, R., Shiah, F.K., Gong, G.C., Huang, Y.T., Kao, S.J., Chen, J.P., Taylor, S.R., Mclennan, S.M., 1995. The geochemical evolution of the continental crust.
Lin, F.J., Lin, C.Y., 2010. Effects of acidic processing, transport history, and dust and Rev. Geophys. 33, 293–301.
sea salt loadings on the dissolution of iron from Asian dust. J. Geophys. Res. Atmos. Turpin, B.J., Huntzicker, J.J., 1995. Identification of secondary aerosol episodes and
115, 1485–1490. quantification of primary and secondary organic aerosol concentrations during
Hsu, C.-Y., Chiang, H.-C., Lin, S.-L., Chen, M.-J., Lin, T.-Y., Chen, Y.-C., 2016. Elemental SCAQS. Atmos. Environ. 29, 3527–3544.
characterization and source apportionment of PM10 and PM2.5 in the western Virtanen, A., Joutsensaari, J., Koop, T., Kannosto, J., Yli-Pirila, P., Leskinen, J., Makela,
coastal area of central Taiwan. Sci. Total Environ. 541, 1139–1150. J.M., Holopainen, J.K., Poschl, U., Kulmala, M., Worsnop, D.R., Laaksonen, A., 2010.
Huang, G., Cheng, T., Zhang, R., Tao, J., Leng, C., Zhang, Y., Zha, S., Zhang, D., Li, X., Xu, An amorphous solid state of biogenic secondary organic aerosol particles. Nature
C., 2014. Optical properties and chemical composition of PM2.5 in Shanghai in the 467, 824–827.
spring of 2012. Particuology. 13, 52–59. Wang, Y.Q., Zhang, X.Y., Arimoto, R., Cao, J.J., Shen, Z.X., 2004. The transport pathways
Khare, P., Baruah, B.P., 2010. Elemental characterization and source identification of PM and sources of PM10 pollution in Beijing during spring 2001, 2002 and 2003.
2.5 using multivariate analysis at the suburban site of North-East India. Atmos. Res. Geophys. Res. Lett. 31, L14110.
98, 148–162. Wang, X., Bi, X., Sheng, G., Fu, J., 2006. Chemical composition and sources of PM10 and
Koch, D., 2001. Transport and direct radiative forcing of carbonaceous and sulfate PM2.5 aerosols in Guangzhou. China. Environ. Monit. Assess. 119, 425–439.
aerosols in the GISS GCM. J. Geophys. Res. Atmos. 106, 20311–20332. Wang, Y., Jia, C., Tao, J., Zhang, L., Liang, X., Ma, J., Gao, H., Huang, T., Zhang, K., 2016.
Lee, P.K., Brook, J.R., Dabek-Zlotorzynska, E., Mabury, S.A., 2003. Identification of the Chemical characterization and source apportionment of PM 2.5 in a semi-arid and
major sources contributing to PM2.5 observed in Toronto. Environ. Sci. Technol. 37, petrochemical-industrialized city. Northwest China. Sci. Total Environ. 573,
4831. 1031–1040.
Li, J., Zhuang, G., Huang, K., Lin, Y., Xu, C., Yu, S., 2008. Characteristics and sources of Wang, Q., Jiang, N., Yin, S., Li, X., Yu, F., Guo, Y., Zhang, R., 2017. Carbonaceous species
air-borne particulate in Urumqi, China, the upstream area of Asia dust. Atmos. in PM 2.5 and PM 10 in urban area of Zhengzhou in China: Seasonal variations and
Environ. 42, 776–787. source apportionment. Atmos. Res. 191, 1–11.
Li, Y., Yao, N., Sahin, S., Appels, W.M., 2016. Spatiotemporal variability of four pre- Watson, J.G., Chow, J.C., Houck, J.E., 2001. PM2.5 chemical source profiles for vehicle
cipitation-based drought indices in Xinjiang. China. Theor. Appl. Climatol. 1–18. exhaust, vegetative burning, geological material, and coal burning in Northwestern
Li, M., Hu, M., Du, B., Guo, Q., Tan, T., Zheng, J., Huang, X., He, L., Wu, Z., Guo, S., 2017. Colorado during 1995. Chemosphere 43, 1141.
Temporal and spatial distribution of PM 2.5 chemical composition in a coastal city of Wei, W., Zhou, H., Shi, Y., Abe, O., Kai, K., 2005. Climatic and Environmental Changes in
Southeast China. Sci. Total Environ. 605–606. the Source Areas of Dust Storms in Xinjiang, China, during the Last 50 Years. Water,
Liu, B.S., Song, Na, Dai, Q.L., Mei, R.B., Sui, B.H., Bi, X.H., Feng, Y.C., 2016. Chemical Air, & Soil Pollution: Focus 5, 207–216 (In Chinese).
composition and source apportionment of ambient PM2.5 during the non-heating Xu, H., Cao, J., Chow, J.C., Huang, R.J., Shen, Z., Chen, L.W.A., Ho, K.F., Watson, J.G.,
period in Taian. China. Atmos. Res. 170, 23–33. 2016. Inter-annual variability of wintertime PM 2.5 chemical composition in Xi'an,
Liu, B.S., Wu, J., Zhang, J., Wang, L., Yang, J., Liang, D., Dai, Q., Bi, X., Feng, Y., Zhang, China: Evidences of changing source emissions. Sci. Total Environ. 545-546,
Y., Zhang, Q., 2017. Characterization and source apportionment of PM 2.5 based on 546–555.
error estimation from EPA PMF 5.0 model at a medium city in China. Environ. Pollut. Yao, L., Yang, L., Yuan, Q., Yan, C., Dong, C., Meng, C., Sui, X., Yang, F., Lu, Y., Wang, W.,
222, 10–22. 2016. Sources apportionment of PM2.5 in a background site in the North China Plain.
Luo, Y., Zheng, C., Jiang, P., Xu, Y., Wu, H., 2012. Assessment of Ecological Risk of Heavy Sci. Total Environ. 541, 590–598.
Metals in Soils in Kuitun. Xinjiang. Chinese J. Soil Sci. 43, 1247–1252 (In Chinese). Yu, L., 2013. Characterization and Source Apportionment of PM2.5 in an Urban
Maria, S.F., Russell, L.M., Gilles, M.K., Myneni, S.C., 2004. Organic Aerosol Growth Environment in Beijing. Aerosol Air Qual. Res. 13, 574–583.
Mechanisms and Their Climate-Forcing Implications. Science 306, 1921–1924. Yu, Q., Gao, B., Li, G., Zhang, Y., He, Q., Deng, W., Huang, Z., Ding, X., Hu, Q., Huang, Z.,
Ming, L., Jin, L., Li, J., Fu, P., Yang, W., Liu, D., Zhang, G., Wang, Z., Li, X., 2017. PM2.5 Wang, Y., Bi, X., Wang, X., 2016. Attributing risk burden of PM 2.5 -bound polycyclic
in the Yangtze River Delta, China: Chemical compositions, seasonal variations, and aromatic hydrocarbons to major emission sources: Case study in Guangzhou, south
regional pollution events. Environ. Pollut. 223, 200–212. China. Atmos. Environ. 142, 313–323.
Na, K., Cocker, D.R., 2009. Characterization and source identification of trace elements in Zhai, Y., Liu, X., Chen, H., Xu, B., Zhu, L., Li, C., Zeng, G., 2014. Source identification and
PM2.5 from Mira Loma, Southern California. Atmos. Res. 93, 793–800. potential ecological risk assessment of heavy metals in PM 2.5 from Changsha. Sci.
Paatero, P., Tapper, U., 1994. Positive matrix factorization: A non-negative factor model Total Environ. 493, 109–115.
with optimal utilization of error estimates of data values. Environmetrics 5, 111–126. Zhang, D., Iwasaka, Y., 1999. Nitrate and sulfate in individual Asian dust-storm particles
Pan, F., Zhou, W., Chen, F., 2012. Evaluation of ecological sensitivity in Karamay, in Beijing, China in Spring of 1995 and 1996. Atmos. Environ. 33, 3213–3223.
Xinjiang. China. J. Geogr. Sci. 22, 329–345. Zhang, X., Hecobian, A., Zheng, M., Frank, N.H., Weber, R.J., 2010. Biomass burning
Samara, C., Voutsa, D., Kouras, A., Eleftheriadis, K., Maggos, T., Saraga, D., Petrakakis, impact on PM 2.5 over the southeastern US during 2007: integrating chemically
M., 2014. Organic and elemental carbon associated to PM10 and PM 2.5 at urban speciated FRM filter measurements, MODIS fire counts and PMF analysis. Atmos.
sites of northern Greece. Environ. Sci. Pollut. R. 21, 1769–1785. Chem. Phys. 10, 6839–6853.
Semmler, M., Seitz, J., Erbe, F., Mayer, P., Heyder, J., Oberdörster, G., Kreyling, W.G., Zhang, R., Jing, J., Tao, J., Hsu, S.C., 2013. Chemical characterization and source ap-
2008. Long-term clearance kinetics of inhaled ultrafine insoluble iridium particles portionment of PM2.5 in Beijing: seasonal perspective. Atmos. Chem. Phys. 13,
from the rat lung, including transient translocation into secondary organs. Inhal. 7053–7074.
Toxicol. 16, 453–459. Zhang, J., Tian, W., Long, X., Tian, H., Huang, Q., Pingping, X.U., Yang, Q., Zhang, J.,
Shao, Y., Wyrwoll, K.H., Chappell, A., Huang, J., Lin, Z., Mctainsh, G.H., Mikami, M., 2015. Fact and simulation of dust aerosol transported to stratosphere during a strong
Tanaka, T.Y., Wang, X., Yoon, S., 2011. Dust cycle: an emerging core theme in earth dust storm in south Xinjiang. Plateau Meteorology 14, 313–318.
system science. Aeolian Res. 2, 181–204. Zhang, Q., Jiang, X., Tong, D., Davis, S.J., Zhao, H., Geng, G., Feng, T., Zheng, B., Lu, Z.,
Shao, L., Hou, C., Geng, C., Liu, J., Hu, Y., Wang, J., Jones, T., Zhao, C., Bérubé, K., 2016. Streets, D.G., Ni, R., Brauer, M., van Donkelaar, A., Martin, R.V., Huo, H., Liu, Z., Pan,
The oxidative potential of PM 10 from coal, briquettes and wood charcoal burnt in an D., Kan, H., Yan, Y., Lin, J., He, K., Guan, D., 2017. Transboundary health impacts of
experimental domestic stove. Atmos. Environ. 127, 372–381. transported global air pollution and international trade. Nature 543, 705–709.
Sharma, S.K., Mandal, T.K., Jain, S., 2016. Source apportionment of PM2.5 in Delhi, India Zhao, G., Fan, Z., 1994. Study on soil Element Background Values of the Taklamakan
Using PMF Model. B. Environ. Contam. Tox. 97, 1–8. Desert. Arid Zone Research 2, 35–40 (In Chinese).
Shi, G., Xu, J., Peng, X., Xiao, Z., Chen, K., Tian, Y., Guan, X., Feng, Y., Yu, H., Nenes, A., Zhao, X.J., Zhao, P.S., Xu, J., Meng, W., Pu, W.W., Dong, F., He, D., Shi, Q.F., 2013.
Russell, A.G., 2017. pH of aerosols in a polluted atmosphere: source contributions to Analysis of a winter regional haze event and its formation mechanism in the North
highly acidic Aerosol. Environ. Sci. Technol. 51, 4289–4296. China Plain. Atmos. Chem. Phys. 13, 5685–5696.
Song, W., Chang, Y., Liu, X., Li, K., Gong, Y., He, G., Wang, X., Christie, P., Zheng, M., Zhao, M., Huang, Z., Qiao, T., Zhang, Y., Xiu, G., Yu, J., 2015. Chemical characterization,
Dore, A.J., Tian, C., 2015. A multiyear assessment of air quality benefits from China's the transport pathways and potential sources of PM 2.5 in Shanghai: Seasonal var-
emerging shale gas revolution: Urumqi as a case study. Environ. Sci. and Technol. 49, iations. Atmos. Res. 158–159, 66–78.
2066–2072. Zhou, J., Xing, Z., Deng, J., Du, K., 2016. Characterizing and sourcing ambient PM2.5
Tan, J.H., Duan, J.C., Chen, D.H., Wang, X.H., Guo, S.J., Bi, X.H., Sheng, G.Y., He, K.B., over key emission regions in China I: Water-soluble ions and carbonaceous fractions.
Fu, J.M., 2009. Chemical characteristics of haze during summer and winter in Atmos. Environ. 135, 20–30.

268

You might also like