Next Article in Journal
Influence of Urban Road Green Belts on Pedestrian-Level Wind in Height-Asymmetric Street Canyons
Next Article in Special Issue
Simulating Atmospheric Organic Aerosol in the Boreal Forest Using Its Volatility-Oxygen Content Distribution
Previous Article in Journal
Accuracy Assessment of a Satellite-Based Rain Estimation Algorithm Using a Network of Meteorological Stations over Epirus Region, Greece
Previous Article in Special Issue
Theoretical Foundation of the Relationship between Three Definitions of Effective Density and Particle Size
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Research Progress on Heterogeneous Reactions of Pollutant Gases on the Surface of Atmospheric Mineral Particulate Matter in China

1
Key Laboratory of Solid Waste Treatment and Resource Recycle, Southwest University of Science and Technology, Mianyang 621010, China
2
College of Science, Xichang University, Xichang 615000, China
3
School of Environment and Resource, Southwest University of Science and Technology, Mianyang 621010, China
*
Author to whom correspondence should be addressed.
Atmosphere 2022, 13(8), 1283; https://doi.org/10.3390/atmos13081283
Submission received: 26 June 2022 / Revised: 9 August 2022 / Accepted: 11 August 2022 / Published: 12 August 2022
(This article belongs to the Special Issue Atmospheric Aging Processes)

Abstract

:
Haze is the phenomenon of visibility degradation caused by extinction effects related to the physicochemical properties of atmospheric particulate matter (APM). Atmosphere heterogeneous reactions can alter the physicochemical properties of APM. Therefore, it is important to understand the atmospheric heterogeneous reactions of APM in order to reveal the cause of haze. Herein, the current situation, developmental trend, source, and composition of APM pollution in China are reviewed. Additionally, we introduce the reaction characteristics and key chemical processes of common inorganic, organic, and mixed pollutant gases on the surface of mineral particles. The effects of mineral particulate matter on aggregation, regulation, and catalysis in the formation of atmospheric aerosols and the synergistic reaction mechanism of SO2, NO2, O3, and VOCs on the surfaces of different mineral particles are summarized. The problems existing in the current research on heterogeneous reactions on the surfaces of mineral particles are also evaluated. This paper aims to gain a deep understanding of the mechanism of mineral particulate matter promoting the formation of secondary aerosols and attempts to provide theoretical support for effective haze control.

1. Introduction

In the 21st century, rapid socioeconomic and industrial developments have caused serious air quality deterioration, and related health problems have gradually attracted significant attention. With the progress of science and technology, the seriousness of the harm to human health caused by haze has become evident, and global efforts have been made to implement relevant systems for haze control. As an example, China has implemented a series of relatively strict pollutant emission standards and pollution control measures since the promulgation and implementation of the Air Pollution Prevention and Control Action Plan in 2013 [1]. Currently, China’s ambient air quality is considered to be significantly improved. The proportion of excessive days with the primary pollutant at a level of PM2.5, as well as the number of days with heavy air pollution, have been significantly reduced, while the number of days with no pollution has significantly increased [2]. However, according to national air quality monitoring data (Figure 1a), despite the fact that the concentrations of SO2, NO2, CO, PM10, and PM2.5 in the air are decreasing significantly, a large proportion of cities still exceed the PM2.5 and PM10 pollution standards (47.2% and 32%, respectively) [2]. Therefore, atmospheric particulate matter (PM2.5, PM10) remains the main air pollutant in cities. Moreover, as the concentrations of other pollutants decline, the pollution of O3 shows an upward trend (Figure 1b). Studies have shown that PM2.5 and O3 generation share a complex link, having common precursors and influencing one another through multiple pathways in the atmosphere [3,4]. Therefore, atmospheric particulate matter (APM) and O3 are important pollutants affecting ambient air quality and have become the major factors restricting its further improvement [5].
APM is a general term for all kinds of solid and liquid particulate matter in the atmosphere. All kinds of APM are evenly dispersed in the air to form a relatively stable suspension system, that is, the aerosol system [6,7]. APM can enter the human respiratory system through inhalation, causing a variety of respiratory and cardiovascular diseases, thus causing harm to human health, especially in the case of PM2.5 [8]. Furthermore, aerosols are able to absorb and scatter solar light, acting as cloud condensation nuclei that alter cloud formation and the atmospheric lifetime. The resulting direct and indirect climate effects are some of the most uncertain factors in global climate change predictions [6,9,10]. Therefore, understanding the migration and transformation of aerosol particles in the atmosphere and their physicochemical properties is of great significance for assessing the health risks, environmental impacts, and climate effects of aerosols.
APM has a very complex composition, with an aerodynamic equivalent diameter of 0.001–100 μm, and approximately 3000–5000 Tg of APM is emitted into the atmosphere every year [11,12]. Its main components include mineral dust, sea salt, organic aerosols, sulfate, nitrate, ammonium salt, and black carbon [13,14]. Owing to the small particle size, large specific surface area, and strong adsorption capacity of APM, heterogeneous reactions tend to occur on its surface or interface [1]. Heterogeneous reactions on the surface of APM in the atmosphere not only involve the physicochemical characteristics of APM, but are also closely related to the type, concentration, and environment of trace gases in the atmosphere [7,15]. The interactions between APM and photo-oxidants such as O3, •OH, and NO2 make the atmospheric chemical processes extremely complex. The reaction of atmospheric photo-oxidants with SO2, NOx, and organic compounds, such as olefins and aromatic hydrocarbons, leads to an increase in the concentration of particulate matter in the atmosphere [16,17,18]. Volatile organic compounds (VOCs) and inorganic gases in the atmosphere form secondary particles by homogeneous reactions and heterogeneous reactions on the surfaces of particles, and affect the toxicity, hygroscopicity, and radiation characteristics of particles [10,19,20,21]. Hence, the transformation of polluted gases in the atmosphere to secondary particles has attracted extensive attention. The formation of secondary particles is related to the precursors, atmospheric oxidizability, adsorption, catalytic capacity, and acid and alkalinity of the particle surface, whereas in heterogeneous reactions, photo-oxidant depletion reactions on the particle surface and photocatalytic oxidation of transitional metals occur, and this also affects the atmospheric oxidizability. These reactions simultaneously make the chemical composition of particulate matter more complex and may have greater effects on human health and climate change [22]. Therefore, the study of atmospheric heterogeneous reactions is of great significance for understanding regional air quality improvements and global climate change. Taking mineral particles as an example, this paper reviews the recent progress in the study of the surface heterogeneous reactions of atmospheric mineral particles.

2. Source and Composition of APM

APM can be divided into primary and secondary particles [12]. Primary particulate matter is directly released into the atmosphere by natural and anthropogenic pollution sources, such as soil particles, sea salt particles, and burning soot. Secondary particulate matter refers to finer particulate matter, including primary gaseous pollutants (e.g., SO2, NOx, VOCs, and NH3) discharged by fuel combustion and industrial and automobile exhaust gas, which are converted into fine particulate matter through gas-particle transformation in the atmosphere [6,12,23]. Based on the composition and morphology of single particles in the atmosphere, APM can also be divided into six types: soot, fly ash, complex secondary, mineral, organic, and metal particles [24,25]. Owing to the different sources and formation processes of APM, its composition can also vary significantly, especially within urban atmospheres, which are affected by a variety of pollution sources. As shown in Figure 2, the chemical composition of PM2.5 is extremely complex and usually contains inorganic substances, such as mineral dust particles and sulfate, nitrate, and ammonium salts; organic matter, such as organic acids, aromatic hydrocarbons, and aerobic organic matter; and trace elements [26].
Atmospheric mineral particulate matter (AMPM) is the most important component of atmospheric aerosols, accounting for approximately 30–60% of the mass concentration of tropospheric atmospheric particles [12,20,27]. These particles mainly originate from ground dust in arid and semi-arid desert areas, and their annual emissions are approximately 1500–4400 Tg [19,28]. Studies on the composition and characteristics of inhalable particulate matter in Beijing, Shanghai, Zhengzhou, Wuhan, and other large cities in China have shown that mineral particles are an important component of urban atmospheric particles, and AMPM accounts for approximately 50% of the mass concentration of APM in dry areas [29]. For example, AMPM in Beijing accounts for approximately 30–70% of the total particulate matter, and while AMPM in Chengdu is lower than in Beijing and some cities in the northwest, where it is still between 34–40% [30]. Clay and diagenetic minerals are the main components of urban particulate matter in dust paths in China, with aluminosilicate particles being the most common [31]. Among them, sodium feldspar, illite, potassium feldspar, anorthosite, hornblende, and chlorite account for 61.59% of the total particulate matter, and calcite and quartz particles account for 13.59%. Dolomite, gypsum, unformed amorphous substances, and other mineral components are also present [31]. Dong et al. [32] found that quartz, clay minerals, and amorphous materials accounted for 24.1%, 28.5%, and 20% of inhalable particles in northern China during dusty weather, respectively. Wang et al. [33] studied the composition of APM during two extremely large sandstorms in Beijing in 2015 and found that AMPM accounted for 85.3% and 95.4% of APM, respectively, among which the clay mineral content was the highest, being more than 50%, followed by quartz, feldspar, and carbonate particles (Figure 3). In India, Spain, Italy, and North Africa, AMPM account for 30–70% of the total particulate matter, and the main mineral phase is the same as in China. However, because of the different geographical locations and pollution situations, the proportion of each mineral phase is different [34].
The inorganic salts in APM are mainly sulfate, nitrate, and ammonium salts from the homogeneous and heterogeneous reactions of SO2, NOx, and NH3 in the atmosphere [7,12,35,36]. The amount of inorganic salts in particulate matter varies depending on source variations, meteorological conditions, and varying atmospheric transformations [15]. Gao et al. [37] measured the concentration of water-soluble ions in Jinan PM2.5, among which the highest SO42− average concentration was 38.33 ± 26.20 μg/m3, accounting for 44.65 ± 11.30% of the total water-soluble ions, while the NH4+ and NO3 average concentrations were 21.16 ± 16.28 and 15.77 ± 12.06 μg/m3, accounting for 17.63 ± 7.61% and 23.07 ± 5.85% of the total ions, respectively. Lai et al. [38] measured water-soluble ions in particulate matter in Guangzhou, Shenzhen, Zhuhai, and Hong Kong and found that SO42−, NH4+, and NO3 together accounted for 59.3–77.7% and 59.3–77.1% of PM2.5 and PM10 water-soluble ions in winter, respectively. In summer, they accounted for 56.5–84.5% and 46.3–79.2% of PM2.5 and PM10, respectively. In particular, the concentration of SO42−, at 6.0–22.0 μg/m3, was higher than all other ions. The contents of sulfate and nitrate in the particulates in Chengdu were 21.55% and 11.20%, respectively, which were lower than those in Beijing, Shanghai, and Guangzhou [39]. Huang et al. [26] measured the chemical composition of PM2.5 in Beijing, Shanghai, Guangzhou, and Xi’an during high pollution events in 2013, among which SO42, NO3, and NH4+ accounted for approximately 8–18%, 7–14%, and 5–10% of the total mass of PM2.5, respectively. Kim et al. showed that the water-soluble ions NH4+, NO3, and SO42− in particulate matter detected at five stations on St. Nicholas Island in the United States accounted for 8–9%, 23–26%, and 6–11% of the total mass of PM10, and 14–17%, 28–41%, and 9–18% of PM2.5, respectively [40]. Rajeev et al. [41] measured the chemical composition of PM2.5 and rainwater in India during the El Niño and Pacific Decadal Oscillation (PDO). Among them, SO42−, NH4+, and NO3 accounted for 33.6%, 23%, and 8.8% of the total amount of water-soluble ions in PM2.5, and 6.9%, 4.8%, and 7.5% in rainwater, respectively. Weagle et al. [42] interpreted the chemical composition and source of global PM2.5 through global chemical transport model (GEOS-Chem) simulation and surface particulate matter network (SPARTAN) site observation. It was found that the secondary inorganic aerosols (SIA, the sum of SO42, NO3-, and NH4+) in each observation site accounted for 15–40% of the total mass of PM2.5, and the content of SO42 was the highest, accounting for 50–80% of the total SIA (Table 1). Moreover, they found that Beijing, Kanpur, and Dhaka all had much higher levels of PM2.5 and SIA than other cities.
Organic pollutants form the main component of APM, accounting for approximately 20–50% of the particulate matter [43,44]. In heavy pollution events, organic matter contributes to more than 50% of atmospheric particulate matter. Due to the influence of molecular weight and saturated vapor pressure, organic pollutants are more likely to attach to fine particles [45]. The organic matter in particulate matter mainly comes from VOCs, of which olefins, aromatic hydrocarbons, alkanes, and other anthropogenic emissions come from industry, agriculture, and automobile exhausts, and fossil, biomass, and solid waste combustion [46]. Isoprene, monoterpenes, and sesquiterpenes are mainly derived from volcanic activity, plant emissions, biological hydrocarbon emissions, and other natural sources [43,47]. VOCs can be produced by photolysis or through reactions with oxidizing agents (such as •OH, •NO3, and O3) in the atmosphere, which can form new free radicals (•HO2, •RO2, •RO) and peroxides on the mineral surface to facilitate VOCs as photochemical smoke promoters [14,48,49,50]. In addition, they promote the formation of organic sulfate, organic nitrate, and other secondary organic aerosol (SOA) owing to their oxygen-containing active functional groups (C=O, COH, COOH) with high reactivity [46,51]. Studies have found that the spatial distributions and temporal variations of typical organic matter concentration levels in APM in China are significantly varied [52]. The concentration levels of organic pollutants in APM in different regions are not only related to energy structure, but also to climatic conditions (such as temperature, humidity, wind speed, and rainfall) and local circulation. The difference in spatial distribution shows that the level of organic pollution in northern cities is much higher than that in southern cities, and that in eastern inland and western cities it is higher than that in eastern coastal cities. The temporal variations in autumn and winter are higher than those in spring and summer [52].
More than 70 types of metal elements exist in APM [53,54,55,56]. Aluminum, Si, Ti, Ca, Mn, Fe, and other crustal elements mainly originate from natural sources and exist as coarse particles, which are related to geological and surface conditions, whereas pollutants such as Pb, As, Cr, Ni, Cu, and Cd are mainly derived from anthropogenic sources and are related to human activities, such as industrial production and fuel combustion [56,57,58]. Zou et al. [59] analyzed air heavy metal pollution in 53 cities in 29 provinces and found that As, Cr, Cd, Ni, Mn, Pb and other heavy metals in China’s atmosphere mainly originate from fossil fuel combustion, metal smelting, and traffic exhaust emissions. Tan et al. [60] compared the differences in heavy metal pollution levels in the atmosphere across northern and southern cities in China and found that the concentrations of Pb, Cr, and Cd in the atmosphere of southern cities were higher than those of northern cities, while the concentrations of V, As, Mn, and Ni in the atmosphere of northern cities were higher than those of southern cities. These results showed that the differences in heavy metal pollution between the northern and southern cities may be related to coal burning, climate characteristics, and the industrial structure in winter. In addition, current studies have found that the volume concentration of heavy metals in coarse particles is higher than that in fine particles, but the mass concentration of heavy metals in fine particles is higher than that in coarse particles, indicating that heavy metals tend to be enriched in fine particles [54]. Most heavy metals present on the surface of APM have different chemical valence states, which act as catalysts in atmospheric chemical reactions, affecting the surface reaction activity of particles and the generation of free radicals, thus accelerating the migration and transformation of pollutants [61].

3. Surface Heterogeneous Reactions of AMPM

Atmospheric heterogeneous reactions refer to gas–solid–liquid three-phase reactions occurring on the surfaces/interfaces of micro-nano atmospheric particles [15]. These surface/interface reactions are involved in all aspects of atmospheric chemical processes and are of great significance for climate change, human health, and ecological balance [62,63]. In the troposphere, aerosol particles and gas composition are complicated by human activities and natural emissions [23]. Heterogeneous chemical reactions can easily occur between pollutant gases and particulate matter, which are coupled with environmental conditions, such as temperature, relative humidity (RH), illumination, and pressure, making the atmospheric environment more complex and affecting a wider area of this environment [7]. Mineral particles usually manifest with a heteromorphic and porous structure, along with a large surface area, strong adsorption properties, and high reactivity, features which enable them to provide a carrier for the adsorption, catalysis, oxidation, and hydrolysis of a variety of gaseous pollutants [12]. Field observation and laboratory simulation studies have confirmed that the heterogeneous reactions on the surfaces of mineral particles have an important impact on the removal of common gas pollutants in the atmosphere [36,50]. Mineral particles are an important source and sink of various gas pollutants, and also produce new secondary components, thus changing the chemical composition of atmospheric particles [44,47,49]. To understand the contribution of heterogeneous reactions involving mineral particulate matter and polluting gases to haze, a large number of laboratory simulations and external field observations have been conducted to reveal the heterogeneous reaction mechanisms, kinetic parameters, and other information related to haze formation. Therefore, the heterogeneous reaction of AMPM surfaces has become an increasingly important research field in atmospheric and environmental science [64].
Research on atmospheric heterogeneous reactions usually adopts the combination technology of the reaction device (Knudsen cell [65,66,67,68] or flow tube [69,70]) and the in situ detection equipment [71,72,73,74,75,76]; diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS); time of flight mass spectrometer (TOF); gas chromatograph–mass spectrometer (GC-MS); micro-Raman spectroscopy) to explore the kinetic parameters (uptake coefficient (γ)) of the reactions of atmospheric particles and atmospheric trace gases that change with humidity, temperature, and time. Thus, the activation energy, entropy change, enthalpy change, and other important physicochemical parameters of the reactions can be obtained. Then, the physicochemical characteristics of the intermediate and final products are characterized by ion chromatography (IC), X-ray photoelectron spectroscopy (XPS), transmission electron microscopy (TEM), scanning electron microscopy (SEM), and energy dispersive spectroscopy (EDS). Finally, based on the abovementioned research results, the mechanisms of the heterogeneous reactions are inferred. However, the samples usually used in the abovementioned technologies are stacked particles, which are very different from the suspended particles that do not come into contact with one another in the actual atmospheric environment. The close contact of stacked samples brings substantial uncertainty to the experiment. Therefore, smog chamber [77,78,79] and individual particle analysis methods [80,81,82,83,84] have become popular in atmospheric heterogeneous reaction research in recent years. Smog chamber and individual particle analysis methods utilize non-contacting or suspended particles, which can truly reflect the heterogeneous reaction process on the surface of particles in the actual atmospheric environment, to obtain accurate kinetic parameters that reflect the real-life conditions of the atmosphere. At the same time, smog chamber and individual particle research, combined with optical microscope or electron microscopy technology, can observe the size, morphology, and phase state changes of particles during the reaction process, which plays an important role in clarifying the reaction process and mechanism.

3.1. Heterogeneous Reactions of Inorganic Gases on the Surface of AMPM

3.1.1. Heterogeneous Oxidation Reaction of SO2 on the Surface of AMPM

The inorganic salts in APM are mainly sulfate, nitrate, and ammonium salts, of which sulfate is the most abundant [77,85,86,87]. Ueda et al. [80] researched the internal structure and composition of single atmospheric dust particles after the Asian dust outflow event in February 2012 using focusing ion beam (FIB) slicing technology and TEM. Detailed microscopic analyses revealed that the dust particles were composed mainly of amorphous silica and calcite, whereas Fe-rich domains were found inside the particles as submicrometer-diameter grains (Figure 4). Interestingly, a sulfur-containing domain that co-exists with calcium was found in a small domain near the particle surface. The authors speculated that the sulfur-containing domains may be the result of the interaction between calcite and SO2 or H2SO4 during the atmospheric transport of mineral dust. However, this mechanism requires further study using actual atmospheric dust particles. To date, a series of advances have been made in the study of the heterogeneous reaction process of SO2 on mineral oxide surfaces. SO2 has been found to form sulfite only at alkaline oxygen vacancies or hydroxyl sites on the mineral oxide surfaces of A12O3, MgO, TiO2, and CaO, and these sulfite species transform into sulfate species only when heated under aerobic conditions or in the presence of other oxidants [63,64,65,66,86,88,89,90]. Using Fourier transform infrared spectroscopy (FTIR), Goodman et al. [65] found that H2O absorbed on the surface of MgO could transform its surface sulfite into sulfate, yet was unable to do so on the surface of A12O3. In contrast to other mineral oxides, Fe(III) oxides can directly convert SO2 to SO42− because of their strong oxidizability. Baltrusaitis et al. [91] used XPS to study the heterogeneous reaction process of SO2 on the surfaces of goethite and hematite under ultra-high vacuum conditions and found that molecular oxygen could be activated on the surfaces of these minerals and convert SO2 to sulfate. Fu et al. [71] further demonstrated by in situ diffuse reflectance infrared spectroscopy and XPS that surface hydroxyl groups and adsorbed oxygen molecules are the main factors for the oxidation of SO2 to sulfate on the iron oxide surface. This conclusion was confirmed by the heterogeneous reaction of SO2 on other typical oxide surfaces observed by Zhang et al. [88]. Dupart et al. [92] studied the reaction of Arizona test dust (ATD) and iron oxide particles with SO2 in aerosol flow tubes and observed the formation of new particles under ultraviolet (UV) irradiation. They believed that light causes the formation of hydroxyl radicals on the surfaces of metal oxides in mineral particles, which then promotes the conversion of SO2 to sulfate. Park and Yu [93,94] measured the kinetic uptake coefficient of SO2 on the surfaces of the Gobi Desert (GDD) and ATD particles and found that the uptake coefficient of SO2 under UV irradiation was significantly higher than that in darkness, and they verified this concept through simulation calculations.

3.1.2. Heterogeneous Reaction of SO2 and O3 on the Surfaces of AMPM

Oxidants such as O3, NOx, and H2O2 in the atmospheric environment have important effects on the heterogeneous reaction process of SO2 on the surfaces of mineral particles. Ozone is the third-largest greenhouse gas after CO2 and CH4. It is an important photo-oxidant in the troposphere and has a significant impact on atmospheric oxidation [63]. Usher et al. [66] studied the uptake process of SO2 on the surfaces of common mineral particles (mineral oxide, CaCO3, and China loess) using FTIR and Knudsen cell. They found that SO2 was reversibly adsorbed on the SiO2 surface, but irreversible adsorption occurred on other particle surfaces, and sulfite species were generated, which could be further oxidized to sulfate by O3. Ullerstam et al. [73] investigated the kinetic process and mechanism of SO2 oxidation by O3 on the Saharan dust surface and demonstrated that, in the absence of an oxidant, only sulfite was produced; however, O3 was able to further convert sulfite to sulfate. Li et al. [74] found that SO2 existed in the form of sulfite on the surface of CaCO3 and was rapidly converted to sulfate by O3. Han et al. [95] investigated the influence of various environmental factors on the formation of sulfate aerosols on the surface of MgO using a technique that combined simulation experiments with long-term real-time field measurements. The results showed that the main product of the SO2 reaction on the surface of MgO was sulfite, and O3 and UV light could promote the conversion of sulfite to sulfate. Under the UV light conditions, the effects of temperature and RH on the surface sulfate formation of MgO exhibited a single-peak mode. The sensitivities of the factors to the sulfate formation followed the order of RH > UV light > temperature > O3 > SO2. Ozone has been widely studied as an oxidant enabling SO2 to produce sulfate, and the roles of other oxidants have gradually attracted attention.

3.1.3. Heterogeneous Reactions of SO2 and NO2 on the Surfaces of AMPM

Nitrogen dioxide is a polluting gas that exists with SO2 in the atmosphere and can participate in a variety of atmospheric chemical reactions, directly or indirectly changing atmospheric oxidation, which has an important impact on air quality [10,72,75]. The results of field observations have shown that mineral particles co-existed with both sulfate and nitrate [11,96]. Moreover, Chen et al. [84] employed the non-destructive surface-enhanced Raman scattering (SERS) technique to investigate the morphology, chemical composition, and mixing state of multiple components of the individual fine particles during a haze episode in Beijing. The results showed that the atmospheric particles were mainly opaque soot during the cleaning period, while during the pollution period, the atmospheric particles were a mixture of opaque (soot) and transparent particles (nitrate and sulfate) and had a core-shell structure. In addition, the abundance of soot was negatively correlated with the PM2.5 mass concentration because, on hazy days, the former declined, while secondary inorganic aerosol (SNA) concentrations increased. This indicates that the formation of sulfate species on the surface of APM is correlated with NO2. Wang et al. [81] studied the heterogeneous reaction of the atmospheric individual particle CaCO3 using micro-Raman spectroscopy. They found that NOx can react with CaCO3 to form Ca(NO3)2, and that NOx and its product Ca(NO3)2 can further react with SO2 and oxidants to form CaSO4. The heterogeneous reaction pathways of CaMg(CO3)2 particles in the atmosphere were similar to those of CaCO3 particles. Ma et al. [75] found that SO2 alone could only exist in the form of sulfite on the γ-Al2O3 surface, while co-existence with NO2 could promote the transformation of sulfite to sulfate. Nitrogen dioxide alone was converted to nitrate through the surface nitrite intermediate, and, when co-existing with SO2, it could oxidize sulfite on the surface of γ-Al2O3, thus forming sulfate. Liu et al. [76] systematically studied the influence of NO2 on the heterogeneous reaction of SO2 on the surfaces of other mineral dust particles and found that the same conclusion was also applicable to other atmospheric particles. They proposed that NO2 can activate surface oxygen to produce N2O4 intermediates, thus promoting the formation of SO42− (Figure 5). The abilities of O3 and NO2 to oxidize SO2 on the surface of α-A12O3 were compared under different humidity conditions by Liu et al. [27]. The sulfate formation rate and uptake coefficient (γ) of SO2 increased with increasing humidity in the two systems. Under the same humidity conditions, the γ of the SO2/O3 system on α-A12O3 particles was higher than that of SO2/NO2, and O3 oxidized SO2 to sulfate more easily than NO2. In addition, the heterogeneous reaction pathway of SO2/NO2 on α-A12O3 varied with humidity. Furthermore, N2O4 replaced nitrite as intermediate in the presence of SO2 under dry conditions. Under high RH conditions, the intermediate was HONO (Figure 6).

3.1.4. Heterogeneous Reactions of SO2 and H2O2 on the Surfaces of AMPM

Moreover, H2O2, as an important oxidant in the atmosphere, has potential effects on the heterogeneous reaction of SO2 on the surfaces of mineral particles. Huang et al. [89] investigated the uptake kinetics of SO2 on α-A12O3 in the absence and presence of H2O2 at different RH values. The results showed that the saturation coverage of SO2 increased with the increase in RH in the absence of H2O2, because the adsorption of water on the surfaces of particles is beneficial to the hydrolysis of SO2 and mitigates the increase in surface acidity. However, the reaction ultimately reached saturation because the produced sulfite/bisulfite could not be further converted to sulfate/bisulfate in the absence of oxidants. In the presence of H2O2, the saturation coverage of SO2, as well as the time required for the reaction to reach saturation, increased with increasing RH. A similar phenomenon has been observed for other types of particulate matter, such as chemically inactive SiO2. Under the same conditions, Huang et al. [69] also investigated the uptake kinetics of SO2 on three authentic mineral dust samples (Asian mineral dust (AMD), Tengger desert dust (TDD), and ATD). It was found that, due to the different mineralogic compositions and aging extents of these dust samples, with an increase in RH, the corrected uptake coefficient (γc) decreased for the AMD particles but increased for the ATD and TDD particles. It was determined that the presence of H2O2 can promote the absorption of SO2 by mineral powders under different humidity conditions and generate •OH radicals on the mineral dust surface through heterogeneous decomposition to promote the formation of sulfate.

3.1.5. Heterogeneous Reactions of NH3/CO2 and SO2 on the Surfaces of AMPM

In addition to these oxidizing gases, Yang et al. [85,86] found that NH3 could coordinate with SO2 to generate sulfate and ammonium in typical mineral dust (α-Fe2O3, MgO, γ-Al2O3, and TiO2). Hydroxyl and oxygen on the surface of mineral dust played major roles in the conversion of SO2 to sulfate. However, NH3 was adsorbed on the Lewis acid site and the hydroxyl groups on the mineral dust surface, enhancing the surface Lewis basicity and leading to more SO2 molecules being adsorbed on the surface of the mineral dust. Surface-adsorbed SO2 interacted with the surface hydroxyl and oxygen atoms and was oxidized to sulfate. In turn, the formation of sulfate contributed to the adsorption of NH3, mainly as NH4+, owing to the enhanced Brønsted acid sites. The synergistic effect between SO2 and NH3 was more significant for acid mineral oxides, such as γ-Al2O3 and α-Fe2O3, than for basic oxides, such as MgO. Further studies have shown that, due to the different affinities of mineral dusts to water, adsorbed water can inhibit the adsorption and transformation of SO2 on α-Fe2O3, despite the presence of NH3. In γ-Al2O3, the increase in RH slightly promoted the formation of sulfate but inhibited the formation of ammonium. Kebede et al. [70] found that surface water from TiO2 particles could catalyze the conversion of surface NH3 into NOx, and that the photo-oxidation of NH3 on the surface of mineral oxides may be one of the sources of NOx in the atmosphere. Carbon dioxide, the most prominent greenhouse gas, is 4–5 orders of magnitude higher than the above trace gas. Liu et al. [97] found that CO2 competed with SO2 on α-A12O3 for adsorption sites and blocked the active groups, thus inhibiting the heterogeneous oxidation of SO2 on α-A12O3. These findings suggest that atmospheric CO2 has a negative effect on sulfate formation, which potentially decreases solar radiation scattering and exacerbates global warming.

3.2. Heterogeneous Reactions of Organic Gases on the Surface of AMPM

As an important component of APM, under different atmospheric temperatures and RH levels, organic matter mainly undergoes heterogeneous chemical reactions with O3, NO3, and •OH in the atmosphere [14,98]. This aging process affects a series of physicochemical properties of the atmospheric particles, such as the chemical composition, phase state, particle size, hygroscopicity, density, and optical properties, thus affecting human health, air quality, and climate change [16,48]. Studies have shown that SOA has multiple formation mechanisms, with three main pathways [6,63]: (1) in the gas phase, VOCs generate semi-volatile organic compounds (SVOCs) through chemical reactions, and the SVOCs are distributed in the gas/particle phase during the nucleation and growth of new aerosol particles [47,78,99]; (2) VOCs and SVOCs enter existing aerosols or cloud droplets via adsorption or absorption [7,43,50]; and (3) heterogeneous reactions. Different reactions can occur on the surface of aerosols or particles, the interface of the gas–liquid phase, and in the interior of aerosols to produce different low and non-volatile organic compounds (NVOCs) and further enrich the composition of aerosols [12,100]. Further study of these three formation mechanisms has shown that the formation of SOA is complex and diverse, causing it to be considered the greatest aspect of uncertainty in global climate model constructions [6]. Therefore, it is important to explore the SOA formation mechanisms.
Xu et al. [82] used TEM and EDS to characterize the morphology, composition, and mixing state of individual aerosol particles in Lin’an, the surrounding area of the Yangtze River Delta (Figure 7). They classified individual aerosol particles into seven types, including sulfur-rich, potassium-rich, organic matter (OM), soot, fly ash, metal, and mineral particles, and found that the fraction of organic-coated S-rich inorganic particles was proportional to sunshine intensity. These findings imply that light can promote the formation of SOA through the heterogeneous reactions of organic matter on the surface of S-rich inorganic particles. They also imply that organic matter can form SOA through heterogeneous reactions on the surfaces of S-containing inorganic particles. However, there is a great deal of uncertainty regarding the origin and formation mechanisms of SOA. This is mainly due to the complexity of organic reactions, the variety of species, and the variability of reaction conditions. However, the adsorption or catalytic oxidation of VOCs on the particulate surface are the two heterogeneous reaction pathways to the formation of SOA.

3.2.1. Heterogeneous Uptake of VOCs on AMPM

1. Effects of humidity and temperature on the heterogeneous reactions
Owing to the different surface properties of minerals and their different ways of interacting with H2O molecules, RH (the surface adsorption of water) affects the adsorption capacity of different minerals to form VOCs. For example, the adsorption of mineral dust on organic acids increases with increasing humidity, whereas the adsorption of some mineral dust decreases with increasing humidity [101,102,103]. Tang et al. [104] investigated the heterogeneous uptake of gaseous acetic acid by different oxides, including γ-Al2O3, CaO, and SiO2, under a range of RH conditions. Under dry conditions, the uptake of acetic acid led to the formation of both acetate and molecularly adsorbed acetic acid on γ-Al2O3 and CaO, and molecularly adsorbed acetic acid alone on SiO2. When the humidity increased from 0% to 15%, the adsorption rate of the gaseous acetic acid molecules by γ-Al2O3 increased by three to five times, whereas for SiO2 particles, acetic acid and water were found to compete for the surface adsorption sites. Fang et al. [105] showed that both HCOOH and HNO3 could reversibly adsorb onto the SiO2 surface under dry conditions. At elevated RH levels, the adsorbed water competed with these acids for surface adsorption sites and promoted their dissociation into hydronium ions and the corresponding anions. The competitive adsorption mechanism of acetic acid and water on kaolinite has been explored in detail by Alstadt et al. [106]. Four chemisorbed acetate species were found on kaolinite after exposure to acetic acid, where acetate bonded through a monodentate or bidentate bridging linkage with an aluminum atom. These species exhibited varying levels of stability after the introduction of water, indicating that water vapor affects the adsorption of organic acids. Specifically, the type of −OH on the surface of mineral dust determines the binding of carboxylic acids to the surface and the stability of these interactions with the addition of water vapor. In atmospheric environments, the ability of H2O to replace chemisorbed molecules may affect the efficiency of mineral dust as a sink for these compounds, thus affecting the atmospheric composition. Romanias et al. [107] studied the heterogeneous interaction between natural Gobi mineral dust and isopropanol at different humidity levels and temperatures. The initial uptake coefficient (γ) was found to be independent of temperature and was negatively correlated with RH. In addition, under UV light, trace TiO2 in natural Gobi dust could oxidize isopropanol to form small amounts of oxygenated VOCs (acetone, formaldehyde, acetic acid, and acetaldehyde).
2. Effects of different mineral dust and VOCs on the heterogeneous reactions
Wu et al. [68] found that α-Al2O3 and CaCO3 could effectively heterogeneously uptake representative unsaturated organic acids (acrylic and methacrylic acids) in the atmosphere. FTIR characterization indicated that these organic acids could interact with the surface OH groups on α-Al2O3, leading to the formation of carboxylate. Conversely, during the heterogeneous uptake of CaCO3, the protons in acrylic acid could react with the HCO3 ions released from Ca(OH)(HCO3) to form carbonic acid. These results suggested that heterogeneous reactions on α-Al2O3 and CaCO3 can act as important sinks for acrylic acid and methacrylic acid, as well as possible contributors to the organic coating found on atmospheric aerosols, especially during high pollution events. Zhang et al. [108] found that the adsorption of VOCs (butyl propionate, 2-heptanone, 1,2,4-trimethylbenzene, and 2-butoxyethanol) on the surface of natural minerals (diatomite, stellerite, and vitric tuff) was positively correlated with the polarity and boiling point of the VOCs, as well as the specific surface area and pore volume of the natural minerals, but negatively correlated with environmental temperature and RH. Romanias et al. [109] studied the heterogeneous interactions of limonene and toluene with dust in six Saharan regions. The results showed that the metal oxides in Saharan dust may affect the dust acidity to enhance the adsorption of limonene, and the SiO2 adsorption activity of toluene was low; therefore, the higher the SiO2 content was, the smaller was the ability of the Saharan sand to adsorb toluene. Because of the influence of SiO2, the six regional mineral dusts could adsorb more limonene than toluene (Figure 8). In addition, the desorption fraction and desorption rate coefficient of the dust samples with a high SiO2 content were significantly reduced, the adsorption of dust samples with a high SiO2 content was mainly irreversible, and the adsorption of dust samples with a low SiO2 content was mainly reversible (Figure 8). Dourtoglou et al. [110,111] removed the metal oxides in natural mordenite and montmorillonite by acid treatment, confirming that the composition of the metal oxides in minerals can regulate their acidity, thus affecting their adsorption and transformation of limonene. Niu et al. [112] also confirmed that the interaction between toluene and SiO2 was very weak at room temperature, and the adsorption of SiO2 to form toluene was much smaller than that of α-Fe2O3 and BS (butlerite and szmolnokite). Wang et al. [113] studied the heterogeneous uptake of formic acid and acetic acid on mineral dust and fly ash in a Knudsen cell reactor in the temperature range of 263–298 K. The study showed that the uptake mechanisms of formic acid and acetic acid in the Inner Mongolian desert sand, ATD, and coal fly ash were mainly ones of physical adsorption, while the uptake process in Xinjiang sierozem was a combination of physical and chemical adsorption, as it contained a considerable amount of calcite. The uptake coefficients of formic and acetic acid on mineral dust increased with the content of elemental Si, aluminosilicate, and/or quartz in the mineralogy, but the uptake coefficients of coal fly ash were smaller than those of mineral dust, which was controlled by the chemical components and microstructure. Compared with formic acid, the uptake coefficients of acetic acid were higher for mineral dust and coal fly ash, especially in low-temperature regions. Iuga et al. [62] studied the adsorption of a single formic acid molecule on different silicate surface models (molecular clusters and a periodic crystal model of the (001) pyrophyllite surface) using methods of quantum mechanics. The silanol groups were found to be the most reactive formic acid adsorption sites. In the case of a periodic system, the silanol groups at the crystal edges were favored. However, the OH groups on the phyllosilicate octahedral sheet were also reactive sites through the tetrahedral cavities. Zeineddine et al. [49] further found that the isoprene uptake in natural Gobi dust was also negatively correlated with ambient temperature and RH. However, unlike isoprene, isopropanol has hydroxyl groups in its structure; therefore, the lifetime of the isopropanol removal by its heterogeneous interaction with the same Gobi dust sample was much longer than that of isoprene. The considerable difference in the lifetime values of the two compounds on the same natural sample indicates that the interactions between VOCs and mineral dust cannot be predicted or inferred from one VOC to another, and they should be determined experimentally to assess the individual importance of VOCs in the atmosphere through different heterogeneous pathways. Therefore, the structure of VOCs limits their interaction mode and pathway on the mineral surface, thus determining their fate and reactivity in heterogeneous reactions, rather than being directly related to their atmospheric abundance.
The aforementioned research has shown that the adsorption of VOCs on the surface of atmospheric mineral dust acts as an important sink of atmospheric VOCs and an important means of forming SOA. In other words, atmospheric mineral dust can be regarded as a reservoir for atmospheric organic matter. In addition to the impacts of environmental temperature and humidity, the physicochemical properties of VOCs (molecular size, group type, polarity, and boiling point), as well as mineral dust types, chemical composition, specific surface area, surface group type, and content, are critical influencing factors.

3.2.2. Heterogeneous Catalytic Oxidation of VOCs on AMPM

The catalytic oxidation of VOCs on the surfaces of mineral particles is another means of forming SOA, which may also cause changes to the atmospheric oxidation and the health risks of atmospheric particles. Styler et al. [114] found that both natural minerals (Mauritanian sand and Icelandic volcanic ash) and metal oxides (Fe2O3 and TiO2) could oxidize oxalate to CO2 under light conditions. The relationship between the oxidation of oxalic acid and the iron and titanium on the surfaces of four types of mineral dust to produce CO2 was studied in the presence and absence of oxygen. Part of this reaction was mediated by iron, and titanium played an important role. Chen et al. [115] studied the oxidation pathways of toluene on TiO2 by combining theoretical calculations and experiments. In general, the homogeneous atmospheric oxidation of xylenes was initiated by •OH, leading to minor H-abstraction and major •OH addition pathways. However, theoretical calculations demonstrated a favorable H-abstraction of the methyl group of xylenes by surface •OH with large exothermic energies, because TiO2 was more easily connected to the methyl in the xylene by hydrogen bonds than it was to the benzene ring (Figure 9). The experimental results highlighted major H-extraction (87.18%) and minor OH-addition (12.82%) pathways to the OH-initiated heterogeneous oxidation of the three xylenes on TiO2 under UV irradiation. Furthermore, the acute risk assessment results indicated that H-extraction products significantly promoted the formation of APM, with a significantly increased risk to human health.
Ponczek et al. [116] found that the adsorption capacity of butanol on ATD was inversely proportional to the environmental temperature and humidity, and the larger the atmospheric mineral dust load was, the smaller was the lifetime of butanol in the atmosphere. Under light conditions, trace TiO2 in the ATD reacted with H2O molecules to produce •OH radicals, thus oxidizing butanol into smaller VOCs. Lv et al. [17] found that isoprene (C5H8) in the atmospheric environment was directly photo-oxidized by iron oxide clusters to generate formaldehyde (CH2O), without the participation of H2O to generate •OH radicals. Under illumination, the photo-oxidation and desorption of isoprene on iron oxide clusters were competitive. Interestingly, both the photo-oxidation and desorption of isoprene on the iron oxide clusters increased with increasing cluster size. However, in the gas-particle interaction system, the desorbed C5H8 molecules were re-adsorbed by the iron oxide clusters, and the large iron oxide clusters could absorb photons and C5H8 molecules more efficiently, thus promoting the formation of CH2O. Theoretical studies have shown that the photo-oxidation of C5H8 to CH2O must overcome a significant barrier, and that light irradiation can accelerate this process. Moreover, compared with UV irradiation, visible irradiation can selectively increase the yield of CH2O.
These studies have shown that atmospheric mineral dust is not inert but acts as a reservoir of organic matter in the atmosphere. Trace metals such as Fe, Ti, and Al in mineral dust can oxidize the adsorbed organic matter on the mineral dust surface, either directly or through •OH radicals generated by the reaction of these metals with water molecules. Therefore, atmospheric mineral dust can not only act as a sink for organic pollutants in the atmosphere, but also transports them into the atmosphere and changes the oxidation and toxicity of atmospheric particles during the migration.

3.3. Heterogeneous Reactions of Mixed Gas on the Surface of AMPM

3.3.1. Heterogeneous Reaction of VOCs and SO2 on the Surfaces of AMPM

Currently, there are few studies on the heterogeneous reactions of organic and inorganic mixed pollution gases on the surface of AMPM. Shao et al. [83] used TEM-EDX to analyze the morphology, composition, and mixing state of single particles collected after the implementation of the “Action Plan for Comprehensive Prevention and Control of Autumn and Winter Air Pollution in Beijing–Tianjin–Hebei and Surrounding Areas 2017–2018” (the Action Plan). After comparing the relative proportions of different particle types and the main pollution sources before and after the implementation of the Action Plan, it was speculated that SO2 and organic matter can react on atmospheric particles (fly ash, mineral dust, and black carbon) to form mixed particles, which may act as an important sink of SO2 and atmospheric organic matter. In the atmospheric environment, VOCs are oxidized to oxygenated volatile organic compounds (OVOCs) by atmospheric oxidants (such as O3, NO3, and •OH) via gas and heterogeneous reactions. OVOCs can form SVOCs and NVOCs through free-radical reactions, thus forming SOA [43]. OVOCs can also be further oxidized and decomposed into small molecules, such as alcohols, aldehydes, ketones, ethers, or acids, and finally oxidized to CO2 [6,117]. There are three main ways by which the effects of small-molecule organic matter on atmospheric heterogeneous reactions are produced:
1. Small-molecule organic matter competes with other polluting gases for reactive oxygen species (ROS) to form aloxy radicals or peroxy alkyl radicals, thereby promoting or inhibiting the reaction of other polluting gases. Chu et al. [118] found that pre-adsorbed propene (C3H6) on TiO2 had no effect on sulfuric acid formation compared to the SO2 reaction on TiO2 alone. This is because formic acid, acetic acid, and other substances produced by the heterogeneous oxidation of C3H6 could be easily removed from the surface of TiO2. However, by introducing both C3H6 and SO2 simultaneously, C3H6 competed with SO2 for ROS, thus inhibiting sulfate formation. In addition, when TiO2 was exposed to NO2 and C3H6 simultaneously, the reaction products occupied the active sites of TiO2, thus suppressing sulfate formation during heterogeneous photo-oxidation.
2. Small-molecule organics inhibit or promote the reactions by binding to or acting as a bridge between the particulate surface and the pollutant gas. Wu et al. [119] found that formic acid and SO2 competed for adsorption on α-Fe2O3, which has a higher affinity to SO2, and only a small amount of formic acid was absorbed on its surface to form formate. However, a small amount of formate on the surface of α-Fe2O3 was catalyzed by SO2 in the gas phase to form CO2, which further promoted the conversion of SO2 in the gas phase to sulfate (Figure 10). This indicates that formic acid and SO2 had a synergistic effect on α-Fe2O3. Li et al. [120] further systematically investigated the effects of dicarboxylic acid salts (including oxalate, malonate, and succinate) on the heterogeneous SO2 transformation with hematite nanoparticles, using a film flow reaction and in situ DRIFTs experiments. It was observed that trace amounts of dicarboxylic acid salts significantly accelerated the sulfate formation under solar light. In addition, the absorption of SO2 increased with the increase in light intensity, and the Fe-oxalate complexation under light irradiation also enhanced the solubility of iron.
3. Small molecule organics compete with other pollutants for reaction sites on the particle surface, thus inhibiting the reaction. Yang et al. [96] found that HAc and SO2 competed for the active sites on α-Al2O3 and CaCO3 to form acetate and sulfite, respectively. However, the gaseous state of HAc was more likely to occupy the -OH and Lewis acid active sites, adsorb to the particulate matter surface, and eventually form acetate. Sulfites could still be generated in the presence of water because water provides abundant -OH active sites, sufficient for HAc and SO2 to be used on the surface of the CaCO3 particles. Wang et al. [72] investigated the influence of NO2 and acetic acid on the heterogeneous reaction of SO2 on the surface of CaCO3 using DRIFTS. It was further confirmed that there was a synergistic interaction between NO2 and SO2, and a competitive interaction between acetic acid and SO2 on the surface of CaCO3 particles. NO2 could promote sulfite oxidation to sulfate, while acetic acid inhibited sulfite generation. Additionally, Zhao et al. [121] found that the pre-adsorption of SO2 on the α-Fe2O3 surface significantly hindered the subsequent heterogeneous oxidation of CH3CHO to produce acetate, whereas the pre-adsorption of CH3CHO significantly inhibited the heterogeneous oxidation reaction of SO2 on the α-Fe2O3 surface.

3.3.2. Heterogeneous Reactions of O3 and VOCs on the Surfaces of AMPM

Ozone in the atmospheric environment also has an important influence on the adsorption and heterogeneous reactions of organic compounds on the surfaces of mineral particles. Wang et al. [122] conducted a comprehensive theoretical study on the heterogeneous reaction processes of four low-molecular-weight unsaturated ketones and O3 on SiO2. They found that not only did SiO2 exhibit a better adsorption performance with these ketones, and a better sedimentation of low-molecular-weight oxygen-containing VOCs in the atmosphere, it also accelerated the reaction between the unsaturated ketones and O3 in the atmosphere. At the same time, the heterogeneous reactions of these ketones on the surfaces of available mineral aerosols may compete with the corresponding gas-phase reactions, especially in arid and semi-arid areas with frequent dust storms. Lian et al. [123] studied the influence of humidity on the heterogeneous reactions of isoprene and O3 on α-Al2O3. The results showed that the presence of O3 led to the rapid transformation of isoprene into carboxylate (COO) ions on the surface of α-Al2O3 particles, whereas the formation of carboxylate ions was significantly inhibited with an increase in humidity. Ma et al. [124] investigated the heterogeneous reaction of O3 with anthracene adsorption on TiO2 and Asian sandstorm particles in the presence and absence of light at room temperature. The adsorption of O3 and anthracene on the Asian sandstorm particles followed the Langmuir–Hinshelwood mechanism. The consumption of anthracene on TiO2 may be caused by volatilization and photodimerization in a flow of dry N2, in the presence of light. In the dark, the anthracene on TiO2 was oxidized to anthraquinone by O3. Under light conditions, the rate constant of the heterogeneous reaction between O3 and anthracene adsorbed on TiO2 was 1.5 times higher than that under dark conditions, and anthraquinone was photo-oxidized further. Moreover, at extremely high O3 concentrations, the anthracene degradation rate in the presence of light was three times higher than that under dark conditions. In addition, the k1,obs values of the heterogeneous reactions of O3 and anthracene adsorbed by the Asian dust storm particles were two orders of magnitude lower than those measured on SiO2, α-Al2O3, and α-Fe2O3 under dark conditions.

3.3.3. Heterogeneous Reactions of NO2 and VOCs on the Surfaces of AMPM

Similarly, Chen et al. [125] found that TiO2, a photoactive material, could promote the reaction between m-xylene and NOx, thus reducing the yield of SOA generated by its interaction in the gas phase. In particular, after the addition of (NH4)2SO4 inorganic aerosol seeds in the gas phase, the yield of SOA decreased. Ma and Wang [126,127] found that the heterogeneous reaction between NO2 and polycyclic aromatic hydrocarbons (PAHs) adsorbed on SiO2, Al2O3, and TiO2 led to the formation of NPAHs. Kameda et al. [21] studied the adsorption and transformation of PAHs on quartz surfaces and reported that the active acid sites on the mineral surfaces were mainly used as electron receptors to participate in the transformation reaction of PAHs and NO2. Ji et al. [18] studied the effect of SiO2 on the heterogeneous reaction between formaldehyde and NO2 using a quantum chemistry method and found that formaldehyde was more easily adsorbed on the SiO2 model than NO2. The presence of SiO2 reduced the reaction barrier and accelerated the reaction rate of formaldehyde and nitrogen dioxide on the SiO2 surface. The reaction produced HONO and altered the oxidation capacity of the atmospheric environment. Chu et al. [128] systematically studied the interaction between NO2, SO2, and toluene in the system of Al2O3 as a seed aerosol. A synergistic effect was observed between NO2, SO2, and toluene, which was enhanced under high concentrations of NH3.
Therefore, very complex heterogeneous reactions can occur under complex air pollution conditions because of the ubiquitous interactions between organic and inorganic species. The relationship between co-existing VOCs and inorganic gases on different mineral surfaces changes due to changes in the mineral and VOCs species and even in the external environment. This makes the oxidation and toxicity of particles more complex through the heterogeneous reaction processes. Therefore, to better explain the causes of haze and evaluate its toxicity, the interactions and influences of VOCs and other inorganic gases and minerals must be considered.

4. Conclusions and Outlook

The environmental and climatic effects of aerosols have received increasing attention, leading to more intensive studies of the heterogeneous reactions on the surface of APM. Although the impacts of mineral dust on the generation of secondary particles are complex, mineral particles may promote atmospheric oxidation and exacerbate haze formation. The exploration of the assemblages of mineral interface reaction products and the significance of their co-evolution can provide a scientific basis for identifying air pollutant sources and pollution control planning.
Firstly, we used China as an example to briefly introduce the current situation and developmental trend of APM pollution. The concentrations of APM in Chinese cities were relatively low before 2010 and then increased annually, with significant seasonal and regional characteristics. After the implementation of a series of air pollution prevention and control policies, especially since the implementation of the Air Pollution Prevention and Control Action Plan in 2013, the emissions of APM, SO2, and NO2 have ehibited a downward trend, the frequency of heavy pollution weather has declined, and the ambient air quality has been improved. However, compared with other countries, the APM concentrations in China remain high, with a large proportion of cities still exceeding the APM pollution standards, and O3 pollution exhibiting an upward trend. APM and O3 are currently important pollutants affecting the environmental air quality in China, and there are intricate correlations between them. Therefore, the coordinated control of APM and O3 is key to the prevention and control of air pollution.
Secondly, we reviewed the sources and characteristics of APM and the surface reaction characteristics of polluted gas and mineral particulate matter. Additionally, we summarized the effects of mineral particulate matter on aggregation, regulation, and catalysis in the formation of atmospheric aerosols. The aim of our analysis was to provide guidance for future studies on the reactions between AMPM and pollutant gases in the atmosphere to form secondary aerosols, and their effects on the chemical composition of the atmosphere. Furthermore, this review is of environmental significance, since it facilitates future studies of both the micro-interface chemistry of the surface characteristics of AMPM in complex pollutants and the combined effects of APM-pollutant aerosol systems in fog-haze formation, transformation, particle production, and blocking behavior.
Finally, having reviewed the existing studies, it can be concluded that a few unresolved issues in the study of atmospheric heterogeneous reactions remain, which will be the focus of future studies. These include the following: (1) The correlation between the atmospheric fine particles and ozone. There is a complex relationship between the formation of atmospheric fine particulate matter and ozone. They not only share common precursors, but also influence one another in the atmosphere in various ways. Therefore, the coordinated control of atmospheric fine particulate matter and ozone is difficult to enact and will be a major challenge for China in its efforts to improve air quality in the next stage. (2) The function of natural mineral dust in haze formation. At present, the role and contribution of mineral dust aerosols to atmospheric aerosol pollution are understood to be mostly based on metal oxides, SiO2, and other active particles. The structure, physics, and physiochemical characteristics of the natural mineral dust, which occupies the main body of the atmospheric particles, and their role in haze formation have been overlooked. Therefore, the synergistic evolution effect mechanism of mineral dust and polluted gas should be explained. (3) Compound effects in heterogeneous reactions. Most studies on the formation process of haze are simplified and involve single laboratory simulation reactions, and few involve the complex action system of multi-component gas. Research on the organic/inorganic mixed gas action system is especially lacking and does not reflect the migration or transformation of multi-component complex gases, nor does it focus on the synergistic or antagonistic effects among various complex or parallel reactions in the process of haze accumulation, which limits our understanding of the formation mechanisms of aerosols. Therefore, it is necessary to study heterogeneous reaction processes when multiple pollutants coexist in a more realistic atmosphere. (4) The role of micro-nano mineral particles. In the future, research should focus on the control mechanisms of the micro-nano mineral surface, micro-area curing, and catalytic transformation of trace pollution gas in the atmosphere from a mineralogical perspective, based on the multiple characteristics of micro-nano mineral particles in haze, such as pollutant sources, carriers, and reaction sites. Additionally, attention should be given to the combined effects of the micro-nano mineral-pollutant aerosol system on fog haze formation, transformation, particle production, and blocking behavior.

Author Contributions

Conceptualization, F.Z. and F.D.; writing—original draft preparation, F.Z.; software, J.Y.; validation, L.Z., X.X. and Y.C.; investigation, F.Z. and Z.L.; data curation, F.Z. and J.X.; writing—review and editing, F.Z., X.L., X.Z. and F.D.; supervision, F.D.; project administration, F.D. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China, grant number 41831285 and 51974261, and the doctoral research initiation project from XiChang University, grant number YBZ202127.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

The authors sincerely thank all those who contributed to this paper and the anonymous reviewers and editors for their helpful comments on the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zheng, B.; Tong, D.; Li, M.; Liu, F.; Hong, C.; Geng, G.; Li, H.; Li, X.; Peng, L.; Qi, J.; et al. Trends in China’s anthropogenic emissions since 2010 as the consequence of clean air actions. Atmos. Chem. Phys. 2018, 18, 14095–14111. [Google Scholar] [CrossRef]
  2. Ministry of Ecology and Environment of the People’s Republic of China. China Ecological and Environment Bulletin. Available online: https://www.mee.gov.cn/hjzl/ (accessed on 2 February 2022).
  3. Xing, J.; Wang, J.; Mathu, R.; Wang, S.; Sarwar, G.; Pleim, J.; Hogrefe, C.; Zhang, Y.; Jiang, J.; Wong, D.C.; et al. Impacts of aerosol direct effects on tropospheric ozone through changes in atmospheric dynamics and photolysis rates. Atmos. Chem. Phys. 2017, 17, 9869–9883. [Google Scholar] [CrossRef]
  4. Xu, J.; Zhang, Y.; Zheng, S.; He, Y. Aerosol effects on ozone concentrations in Beijing: A model sensitivity study. J. Environ. Sci. 2014, 24, 645–656. [Google Scholar] [CrossRef]
  5. Ding, A.; Fu, C.; Yang, X.; Sun, J.; Zheng, F.; Xie, Y.; Herrmann, E.; Nie, W.; Petaja, T.; Kerminen, V.; et al. Ozone and fine particle in the western Yangtze River Delta: An overview of 1 yr data at the SORPES station. Atmos. Chem. Phys. 2013, 13, 5813–5830. [Google Scholar] [CrossRef]
  6. Hallquist, M.; Wenger, J.C.; Baltensperger, U.; Rudich, Y.; Simpson, D.; Claeys, M.; Dommen, J.; Donahue, N.M.; George, C.; Goldstein, A.H.; et al. The formation, properties and impact of secondary organic aerosol: Current and emerging issues. Atmos. Chem. Phys. 2009, 9, 5155–5236. [Google Scholar] [CrossRef]
  7. George, I.J.; Abbatt, J.P. Heterogeneous oxidation of atmospheric aerosol particles by gas-phase radicals. Nat. Chem. 2010, 2, 713–722. [Google Scholar] [CrossRef]
  8. Brook, R.D.; Franklin, B.; Cascio, W.; Hong, Y.; Howard, G.; Lipsett, M.; Luepker, R.; Mittleman, M.; Samet, J.; Smith, S.C.; et al. Air pollution and cardiovascular disease: A statement for healthcare professionals from the Expert Panel on Population and Prevention Science of the American Heart Association. Circulation 2004, 109, 2655–2671. [Google Scholar] [CrossRef]
  9. Ramanathan, V.; Crutzen, P.J.; Kiehl, J.T.; Rosenfeld, D. Aerosols, Climate, and the Hydrological Cycle. Science 2001, 294, 2119–2124. [Google Scholar] [CrossRef] [PubMed]
  10. Tang, M.; Huang, X.; Lu, K.; Ge, M.; Li, Y.; Cheng, P.; Zhu, T.; Ding, A.; Zhang, Y.; Gligorovski, S.; et al. Heterogeneous reactions of mineral dust aerosol: Implications for tropospheric oxidation capacity. Atmos. Chem. Phys. 2017, 17, 11727–11777. [Google Scholar] [CrossRef]
  11. George, C.; Ammann, M.; D’Anna, B.; Donaldson, D.J.; Nizkorodov, S.A. Heterogeneous Photochemistry in the Atmosphere. Chem. Rev. 2015, 115, 4218–4258. [Google Scholar] [CrossRef] [PubMed]
  12. Dong, F.; Shao, L.; Feng, C.; Ju, Y.; Li, J.; Huo, T.; Zhao, Y. Interfacial Reacyion of Atmospheric Micro/Nano Particles and Significance of Mineral Coevolution. Earth Sci. 2018, 43, 1709–1724, (In Chinese with English abstract). [Google Scholar]
  13. Ciere, R.; Querol, X. Solid Particulate Matter in the Atmosphere. Elements 2010, 6, 215–222. [Google Scholar] [CrossRef]
  14. Novak, G.A.; Bertram, T.H. Reactive VOC production from photochemical and heterogeneous reactions occurring at the air–ocean interface. Acc. Chem. Res. 2020, 53, 1014–1023. [Google Scholar] [CrossRef]
  15. Kolb, C.E.; Cox, R.A.; Abbatt, J.P.; Ammann, M.; Davis, E.J.; Donaldson, D.J.; Garrett, B.C.; George, C.; Griffiths, P.T.; Hanson, D.R.; et al. An overview of current issues in the uptake of atmospheric trace gases by aerosols and clouds. Atmos. Chem. Phys. 2010, 10, 10561–10605. [Google Scholar] [CrossRef]
  16. Shi, B.; Wang, W.; Zhou, L.; Li, J.; Wang, J.; Chen, Y.; Zhang, W.; Ge, M. Kinetics and mechanisms of the gas-phase reactions of OH radicals with three C15 alkanes. Atmos. Environ. 2019, 207, 75–81. [Google Scholar] [CrossRef]
  17. Lv, S.; Liu, Q.; Zhao, Y.; Zhang, M.; Jiang, L.; He, S. Formaldehyde Generation in Photooxidation of Isoprene on Iron Oxide Nanoclusters. J. Phys. Chem. C 2019, 123, 5120–5127. [Google Scholar] [CrossRef]
  18. Ji, Y.; Wang, H.; Li, G.; An, T. Theoretical investigation on the role of mineral dust aerosol in atmospheric reaction: A case of the heterogeneous reaction of formaldehyde with NO2 onto SiO2 dust surface. Atmos. Environ. 2015, 103, 207–214. [Google Scholar] [CrossRef]
  19. Zhang, W.; Guo, Z.; Zhang, W.; Ji, Y.; Li, G.; An, T. Contribution of reaction of atmospheric amine with sulfuric acid to mixing particle formation from clay mineral. Sci. Total Environ. 2022, 821, 153336. [Google Scholar] [CrossRef] [PubMed]
  20. Yue, Y.; Cheng, J.; Lee, K.; Stocker, R.; He, X.; Yao, M.; Wang, J. Effects of relative humidity on heterogeneous reaction of SO2 with CaCO3 particles and formation of CaSO4·2H2O crystal as secondary aerosol. Atmos. Environ. 2021, 268, 118776. [Google Scholar] [CrossRef]
  21. Kameda, T.; Azumi, E.; Fukushima, A.; Tang, N.; Matsuki, A.; Kamiya, Y.; Toriba, A.; Hayakawa, K. Mineral dust aerosols promote the formation of toxic nitropolycyclic aromatic compounds. Sci. Rep. 2016, 6, 24427. [Google Scholar] [CrossRef]
  22. Zhu, T. Heterogeneous reactions on the surface of atmospheric particulate matter. Sci. Sin. Chim. 2010, 40, 1729–1730. (In Chinese) [Google Scholar]
  23. Gentner, D.R.; Jathar, S.H.; Gordon, T.D.; Bahreini, R.; Day, D.A.; Haddad, I.E.; Hayes, P.L.; Pieber, S.M.; Platt, S.M.; Gouw, J.A.; et al. A review of urban secondary organic aerosol formation from gasoline and diesel motor vehicle emissions. Environ. Sci. Technol. 2017, 51, 1074–1093. [Google Scholar] [CrossRef] [PubMed]
  24. Huneeus, N.; Schulz, M.; Balkanski, Y.; Griesfeller, J.; Prospero, J.; Kinne, S.; Bauer, S.; Boucher, O.; Chin, M.; Dentener, F.; et al. Global dust model intercomparison in AeroCom phase I. Atmos. Chem. Phys. 2011, 11, 7781–7816. [Google Scholar] [CrossRef]
  25. Ning, C.; Gao, Y.; Zhang, H.; Wang, L.; Yu, H.; Zou, L.; Cao, R.; Chen, J. Molecular chemodiversity of water-soluble organic matter in atmospheric particulate matter and their associations with atmospheric conditions. Sci. Total Environ. 2021, 809, 151171. [Google Scholar] [CrossRef] [PubMed]
  26. Huang, R.; Zhang, Y.; Bozzetti, C.; Ho, K.; Cao, J.; Han, Y.; Daellenbach, K.R.; Slowik, J.G.; Platt, S.M.; Canonaco, F.; et al. High secondary aerosol contribution to particulate pollution during haze events in China. Nature 2014, 514, 218–222. [Google Scholar] [CrossRef]
  27. Liu, W.; He, X.; Pang, S.; Zhang, Y. Effect of relative humidity on O3 and NO2 oxidation of SO2 on α-Al2O3 particles. Atmos. Environ. 2017, 167, 245–253. [Google Scholar] [CrossRef]
  28. Yuan, S.; Liu, S.; Wang, X.; Zhang, H.; Yuan, S. Atomistic insights into heterogeneous reaction of hydrogen peroxide on mineral oxide particles. Appl. Surf. Sci. 2021, 556, 149707. [Google Scholar] [CrossRef]
  29. He, K.; Yang, F.; Ma, Y.; Zhang, Q.; Yao, X.; Chan, C.; Cadle, S.; Chan, T.; Mulawa, P. The characteristics of PM2.5 in Beijing, China. Atmos. Environ. 2001, 35, 4959–4970. [Google Scholar] [CrossRef]
  30. Zhang, X.; Zhang, Y.; Cao, G. Aerosol Chemical Compositions of Beijing PM1 and Its Control Countermeasures. J. Appl. Meteorol. Sci. 2012, 23, 257–264, (In Chinese with English abstract). [Google Scholar]
  31. Shao, L.; Li, W.; Yang, S.; Shi, Z.; Lv, S. Mineralogical characteristics of airborne particles collected in Beijing during a severe Asian dust storm period in spring 2002. Sci. China Ser. D Earth Sci. 2007, 6, 953–959. [Google Scholar] [CrossRef]
  32. Dong, F.; He, X.; Li, G. Study on the basic characteristics of several atmospheric dusts in the norther China. Miner. Pet. 2005, 25, 114–117, (In Chinese with English abstract). [Google Scholar]
  33. Wang, W.; Shao, L.; Zhang, D.; Li, Y.; Li, W.; Liu, P.; Xing, J. Mineralogical similarities and differences of dust storm particles at beijing from deserts in the north and northwest. Sci. Total Environ. 2022, 803, 149980. [Google Scholar] [CrossRef] [PubMed]
  34. Tang, M.; Cziczo, D.J.; Grassian, V.H. Interactions of Water with Mineral Dust Aerosol: Water Adsorption, Hygroscopicity, Cloud Condensation, and Ice Nucleation. Chem. Rev. 2016, 116, 4205–4259. [Google Scholar] [CrossRef] [PubMed]
  35. Shang, D.; Peng, J.; Guo, S.; Hu, M. Secondary aerosol formation in winter haze over the BeijingTianjin-Hebei Region, China. Front. Environ. Sci. Eng. 2021, 15, 34. [Google Scholar] [CrossRef]
  36. Yan, G.; Zhang, P.; Yang, J.; Zhang, J.; Zhu, G.; Cao, Z.; Fan, J.; Liu, Z.; Wang, Y. Chemical characteristics and source apportionment of PM2.5 in a petrochemical city:implications for primary and secondary carbonaceous component. J. Environ. Sci. 2021, 103, 322–335. [Google Scholar] [CrossRef]
  37. Gao, X.; Yang, L.; Cheng, S.; Gao, R.; Zhou, Y.; Xue, L.; Shou, Y.; Wang, J.; Wang, X.; Nie, W.; et al. Semi-continuous measurement of water-soluble ions in PM2.5 in Jinan, China: Temporal variations and source apportionments. Atmos. Environ. 2011, 45, 6048–6056. [Google Scholar] [CrossRef]
  38. Lai, S.; Zou, S.; Cao, J.; Lee, S.; Ho, K. Characterizing ionic species in PM2.5 and PM10 in four Pearl River Delta cities, South China. J. Environ. Sci. 2007, 19, 939–947. [Google Scholar] [CrossRef]
  39. Wang, W.; Fu, G. The Research of Source and Composition of Dust Haze in Chengdu and its Control Countermeasures. Adv. Mater. Res. 2014, 726–731, 2066–2073. [Google Scholar] [CrossRef]
  40. Kim, B.M.; Teffera, S.; Zeldin, M.D. Characterization of PM 25 and PM 10 in the South Coast Air Basin of Southern California: Part 1—Spatial Variations. J. Air Waste Manag. Assoc. 2000, 50, 2034–2044. [Google Scholar] [CrossRef]
  41. Rajeev, P.; Rajput, P.; Gupta, T. Chemical characteristics of aerosol and rain water during an el-nio and pdo influenced indian summer monsoon. Atmos. Environ. 2016, 145, 192–200. [Google Scholar] [CrossRef]
  42. Weagle, C.L.; Snider, G.; Li, C.; Donkelaar, A.V.; Philip, S.; Bissonnette, P.; Burke, J.; Jackson, J.; Latimer, R. Global Sources of Fine Particulate Matter: Interpretation of PM2.5 Chemical Composition Observed by SPARTAN using a Global Chemical Transport Model. Environ. Sci. Technol. 2018, 52, 11670–11681. [Google Scholar] [CrossRef] [PubMed]
  43. Li, X.; Lin, D.; TSONA, N.; GE, M. Anthropogenic Effects on Biogenic Secondary Organic Aerosol Formation. Adv. Atmos. Sci. 2021, 38, 1053–1084. [Google Scholar] [CrossRef]
  44. Wang, J.; Zhao, B.; Wang, S.; Yang, S.; Xing, J.; Morawska, L.D.; Ding, A.; Kulmala, M.; Kerminen, V.M.; Kujansuu, J.; et al. Particulate matter pollution over China and the effects of control policies. Sci. Total Environ. 2017, 584–585, 426–447. [Google Scholar] [CrossRef] [PubMed]
  45. Cheng, M.; Tang, G.; Lv, B.; Li, X.; Wang, Y. Source apportionment of PM2.5 and visibility in Jinan, China. J. Environ. Sci. 2021, 102, 207–215. [Google Scholar] [CrossRef]
  46. Atkinson, R.; Arey, J. Atmospheric degradation of volatile organic compounds. Chem. Rev. 2003, 103, 4605–4638. [Google Scholar] [CrossRef] [PubMed]
  47. Hu, J.; Cheng, Z.; Qin, X.; Dong, P. Reversible and irreversible gas–particle partitioning of dicarbonyl compounds observed in the real atmosphere. Atmos. Chem. Phys. 2022, 22, 6971–6987. [Google Scholar] [CrossRef]
  48. Shi, B.; Wang, W.; Zhou, L.; Sun, Z.; Fan, C.; Chen, Y.; Zhang, W.; Qiao, Y.; Qiao, Y.; Ge, M. Atmospheric oxidation of C10~14 n-alkanes initiated by Cl atoms: Kinetics and mechanism. Atmos. Environ. 2020, 222, 117166. [Google Scholar] [CrossRef]
  49. Zeineddine, M.N.; Romanias, M.N.; Gaudion, V.; Riffault, V.; Thévenet, F. Heterogeneous interaction of isoprene with natural gobi dust. ACS Earth Space Chem. 2017, 1, 236–243. [Google Scholar] [CrossRef]
  50. Shen, X.; Zhao, Y.; Chen, Z.; Huang, D. Heterogeneous reactions of volatile organic compounds in the atmosphere. Atmos. Environ. 2013, 68, 297–314. [Google Scholar] [CrossRef]
  51. Tillmann, R.; Hallquist, M.; Jonsson, Å.M.; Kiendler-Scharr, A.; Saathoff, H.; Iinuma, Y.; Mentel, T.F. Influence of relative humidity and temperature on the production of pinonaldehyde and OH radicals from the ozonolysis of α-pinene. Atmos. Chem. Phys. 2010, 10, 7057–7072. [Google Scholar] [CrossRef]
  52. Mila, A.; Ningbo, G.; Rong, C.; Zhang, H.; Xiuhua, Z.; Jiping, C. Research progress of analytical methods and pollution characteristics of typical organic pollutant in atmospheric particulate matter. Environ. Chem. 2021, 40, 3774–3786, (In Chinese with English abstract). [Google Scholar]
  53. Kyllnen, K.; Vestenius, M.; Anttila, P.; Makkonen, U.; Aurela, M.; Wangberg, I.; Mastromonaco, M.N.; Hakola, H. Trends and source apportionment of atmospheric heavy metals at a subarctic site during 1996–2018. Atmos. Environ. 2020, 236, 117644. [Google Scholar] [CrossRef]
  54. He, A.; Xie, J.; Yuan, C. Heavy Metal Speciation Analysis and Distribution Characteristics in Atmospheric Particulate Matters. Prog. Chem. 2021, 33, 1627–1647, (In Chinese with English abstract). [Google Scholar]
  55. Xu, J.W.; Martin, R.V.; Henderson, B.H.; Meng, J.; Öztanere, Y.B.; Handf, J.L.; Hakamie, A.; Strum, M.; Phillips, S.B. Simulation of airborne trace metals in fine particulate matter over north america. Atmos. Environ. 2019, 214, 116883. [Google Scholar] [CrossRef] [PubMed]
  56. Rajput, J.S.; Trivedi, M.K. Determination and assessment of elemental concentration in the atmospheric particulate matter: A comprehensive review. Environ. Monit. Assess. 2022, 194, 243. [Google Scholar] [CrossRef]
  57. Maia, P.D.; Vieira-Filho, M.; Prado, L.F.; Silva, L.C.; Sodré, F.F.; Ribeiro, H.D.; Ventura, R.S. Assessment of atmospheric particulate matter (PM10) in Central Brazil: Chemical and morphological aspects. Atmos. Pollut. Res. 2022, 13, 101362. [Google Scholar] [CrossRef]
  58. Zhi, M.; Zhang, K.; Lv, W. A review of research advances in the distributions and risk assessments of metal elements in atmospheric particles with different particle sizes. J. Environ. Eng. 2022, 12, 998–1006, (In Chinese with English abstract). [Google Scholar]
  59. Zou, T.; Kang, W.; Zhang, J.; Wang, M.; Fang, X.; Xue, S.; Zhong, B. Concentrations and Distribution Characteristics of Atmospheric Heavy Metals in Urban Areas of China. Res. Environ. Sci. 2015, 28, 1053–1061, (In Chinese with English abstract). [Google Scholar]
  60. Tan, J.; Duan, J. Heavy metals in aerosol in China: Pollution, sources, and control strategies. J. Univ. Chin. Acad. Sci. 2013, 30, 145–155, (In Chinese with English abstract). [Google Scholar]
  61. Wang, M.; Deng, Y.Q.; Dong, F.Q.; He, X.C.; Dai, Q.W.; Sun, S.Y.; Huang, Y.B.; Tang, J.; Chen, W.; Liu, L.Z.; et al. Speciation analysis of cadmium in dust fall about northern China towns. Asian J. Chem. 2013, 25, 7522–7526. [Google Scholar] [CrossRef]
  62. Iuga, C.; Sainz-Díaz, C.I.; Vivier-Bunge, A. Interaction energies and spectroscopic effects in the adsorption of formic acid on mineral aerosol surface models. J. Phys. Chem. C 2012, 116, 2904–2914. [Google Scholar] [CrossRef]
  63. Usher, C.R.; Michel, A.E.; Grassian, V.H. Reactions on mineral dust. Chemical reviews. Chem. Rev. 2003, 103, 4883–4939. [Google Scholar] [CrossRef]
  64. Wang, T.; Liu, Y.; Deng, Y.; Fu, H.; Zhang, L.; Chen, J. Adsorption of SO2 on mineral dust particles influenced by atmospheric moisture. Atmos. Environ. 2018, 191, 153–161. [Google Scholar] [CrossRef]
  65. Goodman, A.L.; Li, P.; Usher, C.R.; Grassian, V.H. Heterogeneous uptake of sulfur dioxide on aluminum and magnesium oxide particles. J. Phys. Chem. A 2001, 105, 6109–6120. [Google Scholar] [CrossRef]
  66. Usher, C.R.; Al-Hosney, H.; Carlos-Cuellar, S.; Grassian, V.H. A laboratory study of the heterogeneous uptake and oxidation of sulfur dioxide on mineral dust particles. J. Geophys. Res. Atmos. 2002, 107, ACH 16-1–ACH 16-9. [Google Scholar] [CrossRef]
  67. Ullerstam, M.; Johnson, M.S.; Vogt, R.; Ljungström, E. DRIFTS and Knudsen cell study of the heterogeneous reactivity of SO2 and NO2 on mineral dust. Atmos. Chem. Phys. 2003, 3, 2043–2051. [Google Scholar] [CrossRef]
  68. Wu, L.; Liu, Q.; Tong, S.; Jing, B.; Wang, W.; Guo, Y.; Ge, M. Mechanism and Kinetics of Heterogeneous Reactions of Unsaturated Organic Acids on α-Al2O3 and CaCO3. ChemPhysChem 2016, 17, 3515–3523. [Google Scholar] [CrossRef]
  69. Huang, L.; Zhao, Y.; Li, H.; Chen, Z. Kinetics of heterogeneous reaction of sulfur dioxide on authentic mineral dust: Effects of relative humidity and hydrogen peroxide. Environ. Sci. Technol. 2015, 49, 10797–10805. [Google Scholar] [CrossRef] [PubMed]
  70. Kebede, M.A.; Varner, M.E.; Scharko, N.K.; Gerber, R.B.; Raff, J.D. Photooxidation of Ammonia on TiO2 as a Source of NO and NO2 under Atmospheric Conditions. J. Am. Chem. Soc. 2013, 135, 8606–8615. [Google Scholar] [CrossRef]
  71. Fu, H.; Wang, X.; Wu, H.; Yin, Y.; Chen, J. Heterogeneous uptake and oxidation of SO2 on iron oxides. J. Phys. Chem. C 2007, 111, 6077–6085. [Google Scholar] [CrossRef]
  72. Wang, R.; Yang, N.; Li, J.; Xu, L.; Tsona, N.T.; Du, L.; Wang, W. Heterogeneous reaction of SO2 on CaCO3 particles: Different impacts of NO2 and acetic acid on the sulfite and sulfate formation. J. Environ. Sci. 2022, 114, 149–159. [Google Scholar] [CrossRef]
  73. Ullerstam, M.; Vogt, R.; Langer, S.; Ljungström, E. The kinetics and mechanism of SO2 oxidation by O3 on mineral dust. Phys. Chem. Chem. Phys. 2002, 4, 4694–4699. [Google Scholar] [CrossRef]
  74. Li, L.; Chen, Z.; Zhang, Y.; Zhu, T.; Li, J.; Ding, J. Kinetics and mechanism of heterogeneous oxidation of sulfur dioxide by ozone on surface of calcium carbonate. Atmos. Chem. Phys. 2006, 6, 2453–2464. [Google Scholar] [CrossRef]
  75. Ma, Q.; Liu, Y.; He, H. Synergistic Effect between NO2 and SO2 in Their Adsorption and Reaction on γ-Alumina. J. Phys. Chem. A 2008, 112, 6630–6635. [Google Scholar] [CrossRef] [PubMed]
  76. Liu, C.; Ma, Q.; Liu, Y.; Ma, J.; He, H. Synergistic reaction between SO2 and NO2 on mineral oxides: A potential formation pathway of sulfate aerosol. Phys. Chem. Chem. Phys. 2012, 14, 1668–1676. [Google Scholar] [CrossRef] [PubMed]
  77. Zhang, S.; Li, D.; Ge, S.; Liu, S.; Wu, C.; Wang, Y.; Chen, Y.; Lv, S.; Wang, F.; Meng, J.; et al. Rapid sulfate formation from synergetic oxidation of SO2 by O3 and NO2 under ammonia-rich conditions: Implications for the explosive growth of atmospheric PM2.5 during haze events in China. Sci. Total Environ. 2021, 17, 9869–9883. [Google Scholar] [CrossRef]
  78. Liu, S.; Huang, D.D.; Wang, Y.Q.; Zhang, S.; Liu, X.D.; Wu, C.; Du, W.; Wang, G.H. Synergetic effects of NH3 and NOx on the production and optical absorption of secondary organic aerosol formation from toluene photooxidation. Atmos. Chem. Phys. 2021, 21, 17759–17773. [Google Scholar] [CrossRef]
  79. Chu, B.; Chen, T.; Liu, Y.; Ma, Q.; Mu, Y.; Wang, Y.; Ma, J.; Zhang, P.; Liu, J.; Liu, C.S.; et al. Application of smog chambers in atmospheric process studies. Natl. Sci. Rev. 2022, 9, 16. [Google Scholar] [CrossRef]
  80. Ueda, S.; Miki, Y.; Kato, H.; Miura, K.; Uematsu, M. Internal structure of asian dust particles over the western north pacific: Analyses using focused ion beam and transmission electron microscopy. Atmosphere 2020, 11, 78. [Google Scholar] [CrossRef]
  81. Wang, M.; Zheng, N.; Zhao, D.; Shang, J.; Zhu, T. Using Micro-Raman Spectroscopy to Investigate Chemical Composition, Mixing States, and Heterogeneous Reactions of Individual Atmospheric Particles. Environ. Sci. Technol. 2021, 55, 10243–10254. [Google Scholar] [CrossRef]
  82. Xu, L.; Lingaswamy, A.P.; Zhang, Y.; Liu, L.; Wang, Y.; Zhang, J.; Ma, Q.; Li, W. Morphology, composition, and sources of individual aerosol particles at a regional background site of the YRD, China. J. Environ. Sci. 2019, 77, 354–362. [Google Scholar] [CrossRef] [PubMed]
  83. Shao, L.; Li, J.; Zhang, M.; Wang, X.; Li, Y.; Jones, T.; Feng, X.; Silva, L.O.; Li, W. Morphology, composition and mixing state of individual airborne particles: Effects of the 2017 Action Plan in Beijing, China. J. Clean. Prod. 2021, 329, 129748. [Google Scholar] [CrossRef]
  84. Chen, H.; Duan, F.; Du, J.; Yin, R.; Zhu, L.; Dong, J.; He, K.; Sun, Z.; Wang, S. Surface-enhanced raman scattering for mixing state characterization of individual fine particles during a haze episode in Beijing, China. J. Environ. Sci. 2021, 104, 216–224. [Google Scholar] [CrossRef]
  85. Yang, W.; He, H.; Ma, Q.; Ma, J.; Liu, Y.; Liu, P.; Mu, Y. Synergistic formation of sulfate and ammonium resulting from reaction between SO2 and NH3 on typical mineral dust. Phys. Chem. Chem. Phys. 2016, 18, 956–964. [Google Scholar] [CrossRef] [PubMed]
  86. Yang, W.; Ma, Q.; Liu, Y.; Ma, J.; Chu, B.; He, H. The effect of water on the heterogeneous reactions of SO2 and NH3 on the surfaces of α-Fe2O3 and γ-Al2O3. Environ. Sci. Nano 2019, 6, 2749–2758. [Google Scholar] [CrossRef]
  87. Dentener, F.J.; Carmichael, G.R.; Zhang, Y.; Lelieveld, J.; Crutzen, P.J. Role of mineral aerosol as a reactive surface in the global troposphere. J. Geophys. Res. Atmos. 1996, 2, 22869–22889. [Google Scholar] [CrossRef]
  88. Zhang, X.; Zhuang, G.; Chen, J.; Wang, Y.; Wang, X.; An, Z.; Zhang, P. Heterogeneous reactions of sulfur dioxide on typical mineral particles. J. Phys. Chem. B 2006, 110, 12588–12596. [Google Scholar] [CrossRef]
  89. Huang, L.; Zhao, Y.; Li, H.; Chen, Z. Hydrogen peroxide maintains the heterogeneous reaction of sulfur dioxide on mineral dust proxy particles. Atmos. Environ. 2016, 141, 552–559. [Google Scholar] [CrossRef]
  90. Shang, H.; Liu, X.; Chneg, N.; Li, M.; Chen, Z.; Ai, Z.; Zhang, L. Research progress on heterogeneous reactions of SO2 on mineral particles. J. Huazhong Agric. Univ. 2020, 39, 9–16, (In Chinese with English abstract). [Google Scholar]
  91. Baltrusaitis, J.; Cwiertny, D.M.; Grassian, V.H. Adsorption of sulfur dioxide on hematite and goethite particle surfaces. Phys. Chem. Chem. Phys. 2007, 9, 5542–5554. [Google Scholar] [CrossRef]
  92. Dupart, Y.; King, S.M.; Nekat, B.; Nowak, A.; Wiedensohler, A.; Herrmann, H.; David, G.; Thomas, B.; Miffre, A.; Rairoux, P.; et al. Mineral dust photochemistry induces nucleation events in the presence of SO2. Proc. Natl. Acad. Sci. USA 2012, 109, 20842–20847. [Google Scholar] [CrossRef] [PubMed]
  93. Park, J.; Jang, M.; Yu, Z. Heterogeneous photo-oxidation of SO2 in the presence of two different mineral dust particles: Gobi and Arizona dust. Environ. Sci. Technol. 2017, 51, 9605–9613. [Google Scholar] [CrossRef] [PubMed]
  94. Yu, Z.; Jang, M.; Park, J. Modeling atmospheric mineral aerosol chemistry to predict heterogeneous photooxidation of SO2. Atmos. Chem. Phys. 2017, 17, 10001–10017. [Google Scholar] [CrossRef]
  95. Han, L.; Liu, X.; Chen, Y.; Xiang, X.; Wang, H. Key factors influencing the formation of sulfate aerosol on the surface of mineral aerosols: Insights from laboratory simulations and acsm measurements. Atmos. Environ. 2021, 253, 118341. [Google Scholar] [CrossRef]
  96. Yang, N.; Tsona, N.T.; Cheng, S.; Li, S.; Xu, L.; Wang, Y.; Wu, L.; Du, L. Competitive reactions of SO2 and acetic acid on α-Al2O3 and CaCO3 particles. Sci. Total Environ. 2020, 699, 134362. [Google Scholar] [CrossRef] [PubMed]
  97. Liu, Y.; Wang, T.; Fang, X.; Deng, Y.; Cheng, H.; Fu, H.; Zhang, L. Impact of greenhouse gas CO2 on the heterogeneous reaction of SO2 on alpha-Al2O3. Chin. Chem. Lett. 2020, 31, 2712–2716. [Google Scholar] [CrossRef]
  98. Wang, Y.; Peng, A.; Chen, Z.; Jin, X.; Gu, C. Transformation of gaseous 2-bromophenol on clay mineral dust and the potential health effect. Environ. Pollut. 2019, 250, 686–694. [Google Scholar] [CrossRef]
  99. Xu, J.; Huang, M.; Feng, Z.; Cai, S.; Zhao, W.; Hu, C.; Gu, X.; Zhang, W. Effects of inorganic seed aerosol on the formationof nitrogen-containing organic compoundsfrom reaction of ammonia with photooxidationproducts of toluene. Pol. J. Environ. Stud. 2020, 29, 909–917. [Google Scholar] [CrossRef]
  100. Hettiarachchi, E.; Grassian, V. Heterogeneous Reactions of α-Pinene on Mineral Surfaces: Formation of Organonitrates and α-Pinene Oxidation Products. J. Phys. Chem. A 2022, 126, 4068–4079. [Google Scholar] [CrossRef]
  101. Rubasinghege, G.; Ogden, S.; Baltrusaitis, J.; Grassian, V.H. Heterogeneous uptake and adsorption of gas-phase formic acid on oxide and clay particle surfaces: The roles of surface hydroxyl groups and adsorbed water in formic acid adsorption and the impact of formic acid adsorption on water uptake. J. Phys. Chem. A 2013, 117, 11316–11327. [Google Scholar] [CrossRef]
  102. Ma, Q.; Liu, Y.; Liu, C.; He, H. Heterogeneous reaction of acetic acid on MgO, α-Al2O3, and CaCO3 and the effect on the hygroscopic behaviour of these particles. Phys. Chem. Chem. Phys. 2012, 14, 8403–8409. [Google Scholar] [CrossRef] [PubMed]
  103. Tong, S.; Wu, L.; Ge, M.; Wang, W.; Pu, Z. Heterogeneous chemistry of monocarboxylic acids on α-Al2O3 at different relative humidities. Atmos. Chem. Phys. 2010, 10, 7561–7574. [Google Scholar] [CrossRef]
  104. Tang, M.; Larish, W.A.; Fang, Y.; Gankanda, A.; Grassian, V.H. Heterogeneous reactions of acetic acid with oxide surfaces: Effects of mineralogy and relative humidity. J. Phys. Chem. A 2016, 120, 5609–5616. [Google Scholar] [CrossRef]
  105. Fang, Y.; Tang, M.; Grassian, V.H. Competition between displacement and dissociation of a strong acid compared to a weak acid adsorbed on silica particle surfaces: The role of adsorbed water. J. Phys. Chem. A 2016, 120, 4016–4024. [Google Scholar] [CrossRef] [PubMed]
  106. Alstadt, V.J.; Kubicki, J.D.; Freedman, M.A. Competitive Adsorption of Acetic Acid and Water on Kaolinite. J. Phys. Chem. A 2016, 120, 8339–8346. [Google Scholar] [CrossRef] [PubMed]
  107. Romanias, M.N.; Zeineddine, M.N.; Gaudion, V.; Lun, X.; Thevenet, F.; Riffault, V. Heterogeneous interaction of isopropanol with natural Gobi dust. Environ. Sci. Technol. 2016, 50, 11714–11722. [Google Scholar] [CrossRef] [PubMed]
  108. Zhang, G.; Liu, Y.; Zheng, S.; Hashisho, Z. Adsorption of volatile organic compounds onto natural porous minerals. J. Hazard. Mater. 2019, 364, 317–324. [Google Scholar] [CrossRef]
  109. Romanías, M.N.; Ourrad, H.; Thévenet, F.; Riffault, V. Investigating the heterogeneous interaction of VOCs with natural atmospheric particles: Adsorption of limonene and toluene on saharan mineral dusts. J. Phys. Chem. A 2016, 120, 1197–1212. [Google Scholar] [CrossRef]
  110. Makarouni, D.; Lycourghiotis, S.; Kordouli, E.; Bourikas, K.; Kordulis, C.; Dourtoglou, V. Transformation of limonene into P-cymene over acid activated natural Mordenite utilizing atmospheric oxygen as a green oxidant: A novel mechanism. Appl. Catal. B Environ. 2018, 224, 740–750. [Google Scholar] [CrossRef]
  111. Lycourghiotis, S.; Makarouni, D.; Kordouli, E.; Bourikas, K.; Kordulis, C.; Dourtoglou, V. Transformation of Limonene into High Added Value Products over Acid Activated Natural Montmorillonite. Catal. Today 2020, 355, 757–767. [Google Scholar] [CrossRef]
  112. Niu, H.; Li, K.; Chu, B.; Su, W.; Li, J. Heterogeneous reactions between toluene and NO2 on mineral particles under simulated atmospheric conditions. Environ. Sci. Technol. 2017, 51, 9596–9604. [Google Scholar] [CrossRef]
  113. Wang, Y.; Zhou, L.; Wang, W.; Ge, M. Heterogeneous uptake of formic acid and acetic acid on mineral dust and coal fly ash. ACS Earth Space Chem. 2020, 4, 202–210. [Google Scholar] [CrossRef]
  114. Styler, S.A.; Donaldson, D.J. Heterogeneous photochemistry of oxalic acid on Mauritanian sand and Icelandic volcanic ash. Environ. Sci. Technol. 2012, 46, 8756–8763. [Google Scholar] [CrossRef] [PubMed]
  115. Chen, J.; Yi, J.; Ji, Y.; Zhao, B.; Ji, Y.; Li, G.; An, T. Enhanced H-abstraction contribution for oxidation of xylenes via mineral particles: Implications for particulate matter formation and human health. Environ. Res. 2020, 186, 109568. [Google Scholar] [CrossRef] [PubMed]
  116. Ponczek, M.; George, C. Kinetics and product formation during the photooxidation of butanol on atmospheric mineral dust. Environ. Sci. Technol. 2018, 52, 5191–5198. [Google Scholar] [CrossRef] [PubMed]
  117. Hu, W.; Liu, D.; Su, S.; Ren, L.; Ren, H.; Wei, L.; Yue, S.; Xie, Q.; Zhang, Z.; Wang, Z.; et al. Photochemical Degradation of Organic Matter in the Atmosphere. Adv. Sustain. Syst. 2021, 51, 2100027. [Google Scholar] [CrossRef]
  118. Chu, B.; Wang, Y.; Yang, W.; Ma, J.; Ma, Q.; Zhang, P.; Liu, Y.; He, H. Effects of NO2 and C3H6 on the heterogeneous oxidation of SO2 on TiO2 in the presence or absence of UV–Vis irradiation. Atmos. Chem. Phys. 2019, 19, 14777–14790. [Google Scholar] [CrossRef]
  119. Wu, L.; Tong, S.; Zhou, L.; Wang, W.; Ge, M. Synergistic Effects between SO2 and HCOOH on α-Fe2O3. J. Phys. Chem. A 2013, 117, 3972–3979. [Google Scholar] [CrossRef] [PubMed]
  120. Li, K.; Fang, X.; Wang, T.; Gong, K.; Tahir, M.A.; Wang, W.; Han, J.; Cheng, H.Y.; Xu, G.J.; Zhang, L.W. Atmospheric organic complexation enhanced sulfate formation and iron dissolution on nano α-Fe2O3. Environ. Sci. Nano 2021, 8, 698. [Google Scholar] [CrossRef]
  121. Zhao, X.; Kong, L.; Sun, Z.; Ding, X.; Cheng, T.; Yang, X.; Chen, J. Interactions between heterogeneous uptake and adsorption of sulfur dioxide and acetaldehyde on hematite. J. Phys. Chem. A 2015, 119, 4001–4008. [Google Scholar] [CrossRef] [PubMed]
  122. Wang, X.; Sun, J.; Han, D.; Bao, L.; Mei, Q.; Wei, B.; An, Z.; He, M.; Yuan, S.; Xie, J.; et al. Gaseous and heterogeneous reactions of low-molecular-weight (LMW) unsaturated ketones with O3: Mechanisms, kinetics, and effects of mineral dust in tropospheric chemical processes. Chem. Eng. J. 2020, 395, 125083. [Google Scholar] [CrossRef]
  123. Lian, H.; Pang, S.; He, X.; Yang, M.; Ma, J.; Zhang, Y. Heterogeneous reactions of isoprene and ozone on α-Al2O3: The suppression effect of relative humidity. Chemosphere 2020, 240, 124744. [Google Scholar] [CrossRef] [PubMed]
  124. Ma, J.; Liu, Y.; Ma, Q.; Liu, C.; He, H. Heterogeneous photochemical reaction of ozone with anthracene adsorbed on mineral dust. Atmos. Environ. 2013, 72, 165–170. [Google Scholar] [CrossRef]
  125. Chen, Y.; Tong, S.; Wang, J.; Peng, C.; Ge, M.; Xie, X.; Sun, J. Effect of titanium dioxide on secondary organic aerosol formation. Environ. Sci. Technol. 2018, 52, 11612–11620. [Google Scholar] [CrossRef] [PubMed]
  126. Ma, J.; Liu, Y.; He, H. Heterogeneous reactions between NO2 and anthracene adsorbed on SiO2 and MgO. Atmos. Environ. 2011, 45, 917–924. [Google Scholar] [CrossRef]
  127. Wang, H.; Hasegawa, K.; Kagaya, S. The nitration of pyrene adsorbed on silica particles by nitrogen dioxide. Chemosphere 2000, 41, 1479–1484. [Google Scholar] [CrossRef]
  128. Chu, B.; Zhang, X.; Liu, Y.; He, H.; Sun, Y.; Jiang, J.; Li, J.; Hao, J. Synergetic formation of secondary inorganic and organic aerosol: Effect of SO2 and NH3 on particle formation and growth. Atmos. Chem. Phys. 2016, 16, 14219–14230. [Google Scholar] [CrossRef]
Figure 1. (a) From 2013 to 2019, the proportion of 337 cities in China whose pollutants exceeded the standard. (b) Quantile concentrations of O3-8 h of 90% in China from 2015 to 2019 (Figure drawn by the authors with the data source from Ref. [2]). (The “2 + 26” city refers to air pollution transmission channels in Beijing–Tianjin–Hebei and the surrounding areas. Yangtze River Delta includes the city of Shanghai and the provinces of Jiangsu, Zhejiang, and Anhui. Fen-wei Plain contains 12 cities and 2 districts in the provinces of Shaanxi, Shanxi, and Henan. Su-Wan-Lu-Yu area contains 22 cities in the provinces of Jiangsu, Anhui, Shandong, and Henan).
Figure 1. (a) From 2013 to 2019, the proportion of 337 cities in China whose pollutants exceeded the standard. (b) Quantile concentrations of O3-8 h of 90% in China from 2015 to 2019 (Figure drawn by the authors with the data source from Ref. [2]). (The “2 + 26” city refers to air pollution transmission channels in Beijing–Tianjin–Hebei and the surrounding areas. Yangtze River Delta includes the city of Shanghai and the provinces of Jiangsu, Zhejiang, and Anhui. Fen-wei Plain contains 12 cities and 2 districts in the provinces of Shaanxi, Shanxi, and Henan. Su-Wan-Lu-Yu area contains 22 cities in the provinces of Jiangsu, Anhui, Shandong, and Henan).
Atmosphere 13 01283 g001
Figure 2. Chemical composition and source analysis of PM2.5 in Beijing, Shanghai, Guangzhou, and Xi’an during high pollution events. (Reprinted with permission from Ref. [26]. Copyright 2014, Macmillan Publishers Limited.).
Figure 2. Chemical composition and source analysis of PM2.5 in Beijing, Shanghai, Guangzhou, and Xi’an during high pollution events. (Reprinted with permission from Ref. [26]. Copyright 2014, Macmillan Publishers Limited.).
Atmosphere 13 01283 g002
Figure 3. Relative number percentages of different particle types (a) and dust particles (b) during (D1-1 and D2-1) and after (D1-2 and D2-2) severe spring dust storm periods (Da–b: (a) for site, (b) for time). (Reprinted with permission from Ref. [33]. Copyright 2021, Elsevier B.V.).
Figure 3. Relative number percentages of different particle types (a) and dust particles (b) during (D1-1 and D2-1) and after (D1-2 and D2-2) severe spring dust storm periods (Da–b: (a) for site, (b) for time). (Reprinted with permission from Ref. [33]. Copyright 2021, Elsevier B.V.).
Atmosphere 13 01283 g003
Figure 4. Deduced internal heterogeneous chemical and mineralogical structures within the cross-sections of Dust Particles 1 (a) and 2 (b) based on the results of TEM, EDS, and SAED analyses [80].
Figure 4. Deduced internal heterogeneous chemical and mineralogical structures within the cross-sections of Dust Particles 1 (a) and 2 (b) based on the results of TEM, EDS, and SAED analyses [80].
Atmosphere 13 01283 g004
Figure 5. Reaction mechanism of (a) the individual reaction of SO2 and NO2 to form MSO3 and MNO3, and (b) the synergistic reaction between SO2 and NO2 to form MSO4 and MNO3 on the surfaces of mineral particles (NO2 = 200 ppmv, SO2 = 200 ppmv). (Reprinted with permission from Ref. [76]. Copyright 2012, Royal Society of Chemistry).
Figure 5. Reaction mechanism of (a) the individual reaction of SO2 and NO2 to form MSO3 and MNO3, and (b) the synergistic reaction between SO2 and NO2 to form MSO4 and MNO3 on the surfaces of mineral particles (NO2 = 200 ppmv, SO2 = 200 ppmv). (Reprinted with permission from Ref. [76]. Copyright 2012, Royal Society of Chemistry).
Atmosphere 13 01283 g005
Figure 6. Effects of RH on O3 and NO2 oxidation of SO2 on α-Al2O3 particles. (Reprinted with permission from Ref. [27]. Copyright 2017, Elsevier Ltd.).
Figure 6. Effects of RH on O3 and NO2 oxidation of SO2 on α-Al2O3 particles. (Reprinted with permission from Ref. [27]. Copyright 2017, Elsevier Ltd.).
Atmosphere 13 01283 g006
Figure 7. TEM images of the mixing states and EDS spectra of individual particles. (a) Mixture of S-rich, fly ash, soot, and coated by organi cmatter; (b) Mixture of S-rich and Fe-rich particles; (c) Mixture of mineral, fly ash, and OM; (d) Mixture of internally and externally mixed Fe-rich particles with S-rich and organic matter; (e) Mixture of S-rich-K, soot, and Pb-rich particles; (f) Mixture of S-rich, mineral, and organic matter. Compositions in the red circle on individual particles were examined by EDS. (Reprinted with permission from Ref. [82]. Copyright 2018, Elsevier B.V.).
Figure 7. TEM images of the mixing states and EDS spectra of individual particles. (a) Mixture of S-rich, fly ash, soot, and coated by organi cmatter; (b) Mixture of S-rich and Fe-rich particles; (c) Mixture of mineral, fly ash, and OM; (d) Mixture of internally and externally mixed Fe-rich particles with S-rich and organic matter; (e) Mixture of S-rich-K, soot, and Pb-rich particles; (f) Mixture of S-rich, mineral, and organic matter. Compositions in the red circle on individual particles were examined by EDS. (Reprinted with permission from Ref. [82]. Copyright 2018, Elsevier B.V.).
Atmosphere 13 01283 g007
Figure 8. Reversibly and non-reversibly adsorbed amounts of (a) limonene and (b) toluene on each Saharan dust sample. (Reprinted with permission from Ref. [109]. Copyright 2016, American Chemical Society.)
Figure 8. Reversibly and non-reversibly adsorbed amounts of (a) limonene and (b) toluene on each Saharan dust sample. (Reprinted with permission from Ref. [109]. Copyright 2016, American Chemical Society.)
Atmosphere 13 01283 g008
Figure 9. Proposed H-abstraction and OH-addition oxidation pathways of m-xylene. (Reprinted with permission from Ref. [115]. Copyright 2020, Elsevier Inc.).
Figure 9. Proposed H-abstraction and OH-addition oxidation pathways of m-xylene. (Reprinted with permission from Ref. [115]. Copyright 2020, Elsevier Inc.).
Atmosphere 13 01283 g009
Figure 10. Mechanism of the co-adsorption of SO2 and HCOOH on the surface of α-Fe2O3. (Reprinted with permission from Ref. [119]. Copyright 2013, American Chemical Society.).
Figure 10. Mechanism of the co-adsorption of SO2 and HCOOH on the surface of α-Fe2O3. (Reprinted with permission from Ref. [119]. Copyright 2013, American Chemical Society.).
Atmosphere 13 01283 g010
Table 1. PM2.5 Mass and Composition at SPARTAN Sites from Measurements (obs) and GEOS-Chem Simulation (GC). Values Are Reported in μg/m−3 for a Laboratory RH of 30−40% and Simulated RH of 35%. (Reprinted with permission from Ref. [42]. Copyright 2018, American Chemical Society.).
Table 1. PM2.5 Mass and Composition at SPARTAN Sites from Measurements (obs) and GEOS-Chem Simulation (GC). Values Are Reported in μg/m−3 for a Laboratory RH of 30−40% and Simulated RH of 35%. (Reprinted with permission from Ref. [42]. Copyright 2018, American Chemical Society.).
SitePM2.5SIASO42NH4+NO3
obsGCobsGCobsGCobsGCobsGC
Beijing, China67.1 ± 9.975.019.7 ± 2.336.311.2 ± 1.413.33.6 ± 0.69.04.9 ± 1.41.4
Bandung, Indonesia30.8 ± 4.520.07.6 ± 0.89.95.6 ± 0.77.21.4 ± 0.32.60.6 ± 0.20.1
Manila, Philippines19.2 ± 2.824.03.0 ± 0.312.02.1 ± 0.39.10.5 ± 0.12.90.4 ± 0.0.0
Rehovot, Israel17.5 ± 2.623.06.4 ± 0.77.74.7 ± 0.65.60.9 ± 0.12.00.8 ± 0.20.1
Dhaka, Bangladesh49.9 ± 7.379.011.3 ± 1.228.07.1 ± 0.915.12.2 ± 0.47.22.0 ± 0.65.7
Buenos Aires, Argentina10.7 ± 1.615.02.5 ± 0.36.21.3 ± 0.24.40.4 ± 0.11.50.8 ± 0.20.3
Ilorin, Nigeria15.8 ± 2.317.52.4 ± 0.21.91.7 ± 0.21.30.5 ± 0.10.50.2 ± 0.10.1
Singapore, Vietnam15.8 ± 2.415.64.0 ± 0.43.53.2 ± 0.42.20.6 ± 0.10.90.2 ± 0.10.4
Kanpur, India71.9 ± 10.694.018.6 ± 1.929.210.2 ± 1.316.64.6 ± 0.17.63.8 ± 1.15.0
Hanoi, Vietnam50.9 ± 7.545.017.2 ± 1.817.110.1 ± 1.310.03.4 ± 0.64.53.7 ± 1.12.6
Pretoria, South Africa17.5 ± 2.630.67.3 ± 0.715.75.3 ± 0.711.31.4 ± 0.23.70.6 ± 0.20.7
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zheng, F.; Dong, F.; Zhou, L.; Chen, Y.; Yu, J.; Luo, X.; Zhang, X.; Lv, Z.; Xia, X.; Xue, J. Research Progress on Heterogeneous Reactions of Pollutant Gases on the Surface of Atmospheric Mineral Particulate Matter in China. Atmosphere 2022, 13, 1283. https://doi.org/10.3390/atmos13081283

AMA Style

Zheng F, Dong F, Zhou L, Chen Y, Yu J, Luo X, Zhang X, Lv Z, Xia X, Xue J. Research Progress on Heterogeneous Reactions of Pollutant Gases on the Surface of Atmospheric Mineral Particulate Matter in China. Atmosphere. 2022; 13(8):1283. https://doi.org/10.3390/atmos13081283

Chicago/Turabian Style

Zheng, Fei, Faqin Dong, Lin Zhou, Yunzhu Chen, Jieyu Yu, Xijie Luo, Xingyu Zhang, Zhenzhen Lv, Xue Xia, and Jingyuan Xue. 2022. "Research Progress on Heterogeneous Reactions of Pollutant Gases on the Surface of Atmospheric Mineral Particulate Matter in China" Atmosphere 13, no. 8: 1283. https://doi.org/10.3390/atmos13081283

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop