Kinetics of Heterogeneous Reaction of Sulfur Dioxide on Authentic

Aug 17, 2015 - Heterogeneous reaction of SO2 on mineral dust seems to be an important sink for SO2. However, kinetic data about this reaction on authe...
0 downloads 10 Views 1MB Size
Article pubs.acs.org/est

Kinetics of Heterogeneous Reaction of Sulfur Dioxide on Authentic Mineral Dust: Effects of Relative Humidity and Hydrogen Peroxide Liubin Huang, Yue Zhao,† Huan Li, and Zhongming Chen* State Key Laboratory of Environmental Simulation and Pollution Control, College of Environmental Sciences and Engineering, Peking University, Beijing 100871, China

Downloaded by UNIV OF MICHIGAN ANN ARBOR on September 14, 2015 | http://pubs.acs.org Publication Date (Web): August 27, 2015 | doi: 10.1021/acs.est.5b03930

S Supporting Information *

ABSTRACT: Heterogeneous reaction of SO2 on mineral dust seems to be an important sink for SO2. However, kinetic data about this reaction on authentic mineral dust are scarce and are mainly limited to low relative humidity (RH) conditions. In addition, little is known about the role of hydrogen peroxide (H2O2) in this reaction. Here, we investigated the uptake kinetics of SO2 on three authentic mineral dusts (i.e., Asian mineral dust (AMD), Tengger desert dust (TDD), and Arizona test dust (ATD)) in the absence and presence of H2O2 at different RHs using a filter-based flow reactor, and applied a parameter (effectiveness factor) to the estimation of the effective surface area of particles for the calculation of the corrected uptake coefficient (γc). We found that with increasing RH, the γc decreases on AMD particles, but increases on ATD and TDD particles. This discrepancy is probably due to the different mineralogy compositions and aging extents of these dust samples. Furthermore, the presence of H2O2 can promote the uptake of SO2 on mineral dust at different RHs. The probable explanations are that H2O2 rapidly reacts with SO2 on mineral dust in the presence of adsorbed water, and OH radicals, which can be produced from the heterogeneous decomposition of H2O2 on the mineral dust, immediately react with adsorbed SO2 as well. Our results suggest that the removal of SO2 via the heterogeneous reaction on mineral dust is an important sink for SO2 and has the potential to alter the physicochemical properties (e.g., ice nucleation ability) of mineral dust particles in the atmosphere.



INTRODUCTION Mineral dust comprises a significant fraction of atmospheric particulate matter, with 1600 Tg yr−1 being emitted to the atmosphere.1 Particles of large sizes settle quickly near the source regions, but smaller particles can reside in the atmosphere for several days and affect downstream areas through long-range transport.2,3 Mineral dust can provide a medium for the reactions of atmospheric trace gases, and the heterogeneous reaction of trace gases on mineral dust has important impacts on the particle composition and other physicochemical properties of particles.4−6 Sulfur dioxide (SO2) and hydrogen peroxide (H2O2) are of great significance in atmospheric chemistry. It is well-known that SO2 is a significant precursor of sulfate and sulfuric acid, which play critical roles in particle formation in the atmosphere.7,8 Field and model studies have shown that the heterogeneous reaction of SO2 on mineral dust seems to be an important sink for SO2.9−12 The heterogeneous reactions of SO2 on metal oxides have been extensively investigated by laboratory studies.13−19 However, kinetic data about this reaction on authentic mineral dust are limited.20−23 For instance, Usher et al.21 have studied the heterogeneous reaction of SO2 on China Loess. The initial uptake coefficient was measured by Knudsen cell to be (3.0 ± 1.0) × 10−5 and could © XXXX American Chemical Society

be predicted from the reactivity of the single components and their abundance in China loess. Likewise, Adam et al.23 used a flow tube to study the kinetics of SO2 on Saharan dust in the absence and presence of O3. The initial uptake coefficient was determined to be (6.6 ± 0.8) × 10−5 in the absence of O3. Previous studies have been typically performed at low relative humidity (RH). However, water plays a significant role in the heterogeneous reaction of trace gases.24 Given the ubiquity of water vapor in the atmosphere, it is critical to investigate the heterogeneous reaction of SO2 on mineral dust at higher RHs. Hydrogen peroxide (H2O2) is an important oxidant in the atmosphere.25 It has been well-known that H2O2 can efficiently oxidize S(IV) species to sulfate in the aqueous phase.26 Clegg and Abbatt27 found that H2O2 has high reactivity for oxidation of SO2 on the surface of ice. The contribution of this reaction to sulfate production is greater than that of gas-phase OH oxidation in the upper troposphere. Hua et al.28 estimated the contribution of gas and aqueous phase SO2 oxidation to total sulfate formation in a suburban atmosphere in southern China and found that the gas and aqueous phase oxidation processes Received: March 18, 2015 Accepted: August 17, 2015

A

DOI: 10.1021/acs.est.5b03930 Environ. Sci. Technol. XXXX, XXX, XXX−XXX

Article

Environmental Science & Technology

reactor was used to investigate the formation of sulfite and sulfate on mineral dust particles at different RHs. The mass of mineral dust samples used in these experiments was ∼20 mg. The quartz flow reactor and the sample preparation process have been described in detail previously.30 The 400 sccm of SO2-containing synthetic air was directed through the reactor. After 60 min reaction, the products formed on the particles were immediately extracted using 5 mL of 10−1 M H2O2 solution. Generation of Gaseous Reactants. The H2O2 gas was generated by passing a flow of N2 over H2O2 solution (35 wt %, Sigma-Aldrich) maintained at 277 K. The SO2 gas was a standard gas (11 ppmv, National Institute of Metrology, China). The water vapor in the airflow was generated by passing a flow of N2 through a water bubbler. The RH was controlled by adjusting the flow rate of humidified N2, and measured using a RH probe (Vaisala HMT 100). For SO2 uptake kinetics studies (using the filter-based flow reactor), the initial concentrations of SO2 and H2O2 were ∼5 and ∼0.8 ppbv, respectively. For the measurements of sulfite and sulfate formation (using the quartz flow reactor), the initial concentration of SO2 was ∼2 ppmv. All of the experiments were carried out in the dark and at ∼298 K. Mineral Dust Characterization. Three different authentic mineral dusts were used in this study: Asian mineral dust (AMD, collected on the campus Peking University in Beijing during a severe dust storm in April 17, 2006), Tengger Desert dust (TDD, National Institute of Metrology, China), and Arizona test dust (ATD, Powder Technology). In some experiments, AMD particles were washed using ultrapure water to remove the soluble species from the surface of the particles and then dried in a flow of N2 at 343 K (hereafter referred to as “washed AMD”). The BET specific surface area of AMD, washed AMD, TDD, and ATD particles was measured (Micrometritics, ASAP2010) as 6.1 ± 0.21, 12.3 ± 0.47, 16.5 ± 1.12, and 4 ± 0.55 m2 g−1, respectively. The composition of AMD, TDD, and ATD particles was analyzed by X-ray fluorescence spectroscopy (XRF, Thermo Fisher, ARL ADVANT’X), and the results were summarized in Table S1. The typical components of three authentic mineral dusts are similar: major composition is SiO2, with significant amounts of Al2O3, Fe2O3, CaO, and MgO, and smaller amounts of TiO2, Na2O, and K2O. But the abundance of each component in three authentic mineral dusts is different. For instance, ATD particles are composed of approximately 70% SiO2, more than AMD and TDD particles. The size distribution of AMD, TDD, and ATD particles was measured by a laser particle sizer (MasterSizer 2000, Malvern for AMD and TDD particles; Microtrac S3500 for ATD particles). The results are shown in Figure S2. The pore size distribution of AMD, TDD, and ATD particles was measured by a pore size distribution analyzer (Micrometritics, ASAP2010). Figure 2 shows the concentration of water-soluble ions, i.e., Na+, K+, NH4+, Ca2+, Mg2+, SO42−, NO3−, and Cl−, on the surface of AMD, TDD, and ATD particles. Water-soluble ions on mineral dust were extracted with 10 mL of ultrapure water by stirring and shaking for 5 min. Anions were analyzed by an ion chromatograph (IC, Dionex ICS2000), equipped with a Dionex AS11analytical column. Cations were analyzed by another IC (Dionex ICS2500), equipped with a Dionex CS 12A analytical column. Detection of Reactants and Products. H2O2 was determined using high-performance liquid chromatography (HPLC). Details about the method were reported in our

together can only explain 30% of sulfate production. They hence inferred that the interaction between SO2 and H2O2 on aerosol particles may be an additional source of sulfate. It is noted that H2O2 may be able to promote the heterogeneous oxidation of SO2 on mineral dust. However, little is known about this role of H2O2. In the present study, we investigated the uptake kinetics of SO2 on authentic mineral dust in the absence and presence of H2O2 at different RHs using a filter-based flow reactor. The results of this study are helpful to understand the roles of RH and H2O2 in the heterogeneous reaction of SO2 in the atmosphere.



Downloaded by UNIV OF MICHIGAN ANN ARBOR on September 14, 2015 | http://pubs.acs.org Publication Date (Web): August 27, 2015 | doi: 10.1021/acs.est.5b03930

EXPERIMENTAL SECTION Apparatus. A schematic diagram of the experimental setup is shown in Figure 1. A filter-based flow reactor was used to

Figure 1. Schematic diagram of experimental setup. BF, blank filter; DF, dust-deposited filter; SS, stripping solution, i.e., H2O2 solution (10−3 M); IC, ion chromatograph. The coil collector and the diffusion tube were kept in the water bath at 277 K.

investigate uptake kinetics of SO2 on authentic mineral dust. Details about the reactor can be found in a previous study.29 Briefly, the main part of the reactor consists of two perfluoroalkoxy resin filter holders (Savillex Corporation) connected in parallel. A blank PTFE filter (Whatman, 47 mm) or dust-deposited filter was placed in each filter holder. In the experiments, the gas flow containing SO2 or SO2/H2O2 mixture was diluted with dry or humid synthetic air (20% O2 + 80% N2) before entering the reactor. The total 16.7 slpm (standard liters per minute) gas flow was alternatively directed through each filter holder via valves. A portion of gas flow (3 slpm) exiting the filter holder was collected for the measurement of the loss of the SO2 in the gas phase. The difference in SO2 losses between the gas flow passing through the dustdeposited filter and the blank filter corresponds to the amount of SO2 taken up by mineral dust. The dust-deposited filters were prepared using a custom-designed particle resuspension equipment, the details of which are given in the SI. The mineral dust samples deposited on the filter were observed by field emission scan electron microscopy (FESEM, Zeiss, Merlin Compact). The mass of mineral dust (0.9−1.5 mg) on the filters was measured with a high-accuracy mass balance. The filters were sealed and stored in a desiccator before use. It is hard to accurately measure the sulfite and sulfate species formed on the dust particles deposited on the filters due to the low mass loading of dust on the filters. Therefore, a quartz flow B

DOI: 10.1021/acs.est.5b03930 Environ. Sci. Technol. XXXX, XXX, XXX−XXX

Article

Environmental Science & Technology

weight of SO2, kg mol−1; As is the effective surface area of the particles, m2; and ω is the mean molecular velocity, m s−1. Figure S3 shows the amount of SO2 taken up by AMD, TDD, and ATD particles as a function of time under dry conditions. The uptake rate of SO2, d{C}/dt, can be derived by the linear fitting of the uptake data. Three different uptake coefficients of SO2, termed γBET, γext, and γc, were then calculated based on the BET surface area, external surface area (estimated from particle size distributions), and corrected effective surface area (see Results and Discussion below) of particles, respectively.



RESULTS AND DISCUSSION SO2 Uptake under Dry Conditions. It should be noted that the γ calculated using the BET surface area of particles represents the lower limit of the γ. Some explicit models of pore diffusion, for instance KML (Keyser Moore Leu) method, have been employed successfully to correct the γ.32,33 These models are based on the following assumptions: (1) the state of particles accumulated like a sheet, (2) particles themselves are nonporous, and (3) loose packing of the particles makes the sheet highly porous. However, in the present study, the geometric area of the filter is 11.9 cm2, and the external surface area of ATD particles collected on the filter is 6.7 cm2 (the external surface area of the particles is estimated to be 0.67 m2 g−1 and the particle mass is ∼1 mg). Thus, mineral dust particles are expected to be sparsely distributed on the filter, instead of forming an aggregated conglomerate. This was supported by the FESEM measurement of the filter sample (Figure S4). Therefore, the KML method is not applicable to our study. An alternative method34,35 was used to correct the uptake coefficient measured in this study. The assumptions of this method are (1) the particles are spherical and the concentration of SO2 is identical around the particles, and (2) the particles themselves are porous. Assuming that the ATD particles are spherical and have a density of 2.65 g cm−3, according to the particle size distribution, the external surface area of the particles is estimated to be 0.67 m2 g−1, which is a factor of ∼25 smaller than the BET surface area, indicating that ATD particles are porous. Previous studies also found that the authentic mineral dust particles, e.g., Saharan dust particles, are porous.14,22,23 As described in detail in the SI, the porosity of AMD, TDD, and ATD particles was further examined using different methods. The morphology of mineral dust observed by FESEM suggests that the mineral dust really has some pores (Figure S5), and pore size distribution measurements suggest that mineral dust particles are porous and most of those pores are mesopores (Figures S6 and S7). Additionally, the BJH (Barrett−Joyner−Halenda) adsorption accumulative pore area measurements show that the internal surface area of pores accounts for a significant portion of the total surface area of mineral dust particles (Figure S8). For a porous particle, the reactive surface includes both the external surface and the internal surface. The internal surface is provided by the pores in a particle that can be reached by gases via gas-phase diffusion from the exterior of the particle. However, the gas diffusion into the pore of the particle may be hindered, thus influencing the availability of the internal surface area for the reaction.34,35 In this case, the γBET that underestimates the true uptake coefficient because of the limitation of gas-phase diffusion into the internal surface of porous particles can be corrected by eq 4

Downloaded by UNIV OF MICHIGAN ANN ARBOR on September 14, 2015 | http://pubs.acs.org Publication Date (Web): August 27, 2015 | doi: 10.1021/acs.est.5b03930

Figure 2. Concentrations of water-soluble ions on the surface of AMD, TDD, and ATD particles. Error bars represent 1σ.

previous study.28 SO2 was determined by a coil collector collection-IC detection method. The principle of the method is that SO2 can be immediately oxidized to SO42− by H2O2 in the aqueous phase. SO2 or SO2/H2O2-containing air flow after reacting with mineral dust was directed into a coil collector, which was kept in a 277 K water bath. Details of the coil collector were described in our previous study.28 H2O2 solution (10−3 M) used as stripping solution was introduced into the collector by a peristaltic pump to adsorb SO2. SO2 dissolved into the H2O2 solution is rapidly oxidized to SO 4 2− by H 2O 2 . The concentration of SO42− in the solution was measured by IC. Assuming all of the SO2 in the gas phase can be collected and converted to SO42−, the amount of SO42− measured by IC corresponds to that of SO2 in the gas phase. It is noted that the H2O2 stripping solution itself contains dissolved sulfate. However, using a flow rate of 0.15 mL min−1 for the stripping solution, the ratio of background SO42− concentration in stripping solution to total SO42− concentrations after absorbing SO2 could be controlled as 95%. Determination of the Uptake Coefficient. In this study, the uptake coefficient (γ) of SO2 can be calculated using eq 130,31 γ=

d{C}/dt Z

(1)

Z=

1 A s[C ]ω 4

(2)

ω=

8RT πMc

(3)

where {C} is the amount of SO2 taken up by particle surfaces; [C] is the gas concentration of SO2, molecules m−3; Z is the collision frequency, molecules s−1; R is the gas constant, J mol K−1; T is the absolute temperature, K; MC is the molecular C

DOI: 10.1021/acs.est.5b03930 Environ. Sci. Technol. XXXX, XXX, XXX−XXX

Article

Environmental Science & Technology γBETABET = γc(Ae + ηA i )

March 2010).10 Therefore, the hypothesis may not significantly affect the correction of uptake coefficients. The discrepancy in the γc values may be mainly due to different compositions of these mineral dusts. As shown in Table S1, compared to AMD and TDD particles, ATD particles are comprised of more SiO2 but less Fe2O3, CaO, and MgO. A previous study has shown that the reactivity of SiO2 toward SO2 is lower than that of other mineral components (e.g., Al2O3, Fe2O3, CaCO3).21 Therefore, the γc value of SO2 on ATD particles is lowest among the three mineral dusts. Although the γBET represents the lower limit of uptake coefficient, it is still meaningful to report these values in order to compare with literature values. The γBET of SO2 on AMD, TDD, and ATD particles was measured as (3.15 ± 0.33) × 10−5, (3.81 ± 0.25) × 10−5, and (1.31 ± 0.05) × 10−5 under dry conditions, respectively. Recently, the γBET of SO2 on authentic mineral dust has been determined by different methods. Usher et al.21 investigated the heterogeneous uptake of SO2 on China loess using a Knudsen cell reactor, and reported the initial uptake coefficient of (3.0 ± 1.0) × 10−5. Adam et al.20 probed the uptake kinetics of SO2 on Saharan dust using flow tube and the initial uptake coefficient was measured to be (6.6 ± 0.8) × 10−5.20 Using Knudsen cell, Ullerstam et al.22 determined a γBET value of (4.6 ± 0.3) × 10−6 for SO2 on Saharan dust. The γBET values measured in this study are comparable to those reported by Usher et al.21 and Adam et al.,23 but are ∼10 times larger than that by Ullerstam et al.22 This difference can be partly explained by the difference in the preparation of mineral dust samples. In this study, mineral dust particles were highly dispersed on the Teflon filters, whereas the dust sample used by Ullerstam et al.22 was in a highly accumulative state in the sample support of Knudsen cell. The many layers of particles in the latter study will hinder the diffusion of gas into the underlayer particles, resulting in the underestimate of γBET. SO2 Uptake under Humid Conditions. The influence of water vapor on the uptake of SO2 on AMD, TDD, and ATD particles was investigated in the range of 0−90% RH. Table S3 summarizes the γBET and γext of SO2 on AMD, TDD, and ATD particles at different RHs. The value of γext is ∼20 times larger than that of γBET. We also applied the correction method discussed above to correct the uptake coefficients at different RHs. Figure 3 shows the γc values of SO2 on AMD, TDD, and ATD particles as a function of RH. The dependence of γc on

(4)

where γc is the corrected uptake coefficient; ABET is the BET surface area of the particles, m2; Ae is the external surface area of the particles that can be calculated from the particle size distribution (see SI), m2; Ai is the internal surface area of the particles, m2; and η is the effectiveness factor, i.e., the fraction of the internal surface that participates in the reaction. The effectiveness factor can be obtained by eqs 5 and 6 based on the assumption that the reaction is first-order.34,35 η=

1⎡ 1 1 ⎤ − ⎥ ⎢ 3Φ ⎦ Φ ⎣ tan h(3Φ)

Downloaded by UNIV OF MICHIGAN ANN ARBOR on September 14, 2015 | http://pubs.acs.org Publication Date (Web): August 27, 2015 | doi: 10.1021/acs.est.5b03930

Φ=

Rp

k tSgρp

3

De

(5)

(6)

Here, Φ is the Thiele modulus, i.e., the ratio of the surface reaction rate to the gas diffusion rate into the pores; kt is the surface reaction rate constant normalized to the surface area density of particles, m s−1, and can be derived from γcω/4 (see SI); ρp is the density of the particles, g cm−3; Rp is the average radius of the particles, m; Sg is the BET specific surface area of the particles, m2 g−1; and De is the effective diffusion coefficient of gas in the pores, m2 s−1. To evaluate η, De must be determined. In this study, the gasphase diffusion in the pores follows Knudsen diffusion (see SI), so the De can be expressed as eqs 7 and 836

θD k τ

De = Dk =

(7)

2rpω (8)

3 2 −1

where Dk is the Knudsen diffusion coefficient, m s ; rp is the average pore radius, nm; τ is the tortuosity factor; and θ is the porosity. Typically, the τ is 1−4,37 and the value of 2 was used in the present study. The θ and the rp can be given by eqs 9 and 10, respectively34

θ = Vgρp rp =

(9)

2Vg Sg

(10) 3

−1

where Vg is the pore volume of the particles, cm g . Substituting eqs 7−10 into eq 6 gives eq 11. Φ=

R pSg 12Vg

3τγc

(11)

Combining eqs 5 and 11, we can obtain the η of AMD, TDD, and ATD particles under dry conditions. The results are shown in Table S2. The corrected uptake coefficient, γc, of SO2 on AMD, TDD, and ATD particles is estimated to be (2.14 ± 0.23) × 10−4, (2.29 ± 0.15) × 10−4, and (0.35 ± 0.01) × 10−4, respectively, under dry conditions. The ratio of γc to γBET is ∼7, ∼6, and ∼3, respectively. It is known that the shape of mineral dust particles is irregular. Thus, the hypothesis that the particles are spherical will underestimate the external surface area of particles. However, a previous study found that the uncertainty in the external surface area estimation based on the measured size distribution and the spherical particle assumption is less than 5% for Asian mineral dust (collected in a dust storm in

Figure 3. γc of SO2 on AMD, TDD, and ATD particles at different RHs. The initial concentration of SO2 is ∼5 ppbv. Error bars represent 1σ. D

DOI: 10.1021/acs.est.5b03930 Environ. Sci. Technol. XXXX, XXX, XXX−XXX

Article

Environmental Science & Technology

(e.g., OH groups and lattice oxygen) or reacting with surface adsorbed water (i.e., hydration and dissociation).4,15 Adsorbed water can occupy surface active sites and consequently decrease the surface reactivity toward SO2, but it can also accelerate the hydration of SO2. The increase of γc with increasing RH requires that the role of adsorbed water in SO2 hydration prevails over that in depleting surface active sites. Previous studies have proposed that absorbed water can significantly promote the uptake of SO2 on CaCO3 and MgO particles,13,16,38,39 but the influence of absorbed water on other significant components of mineral dust (e.g., alumina and iron oxide) is not clear. Different compositions of mineral dust could affect the dependence of uptake coefficient on RH. However, as shown in Table S1, the chemical composition of AMD particles is similar to that of TDD particles, thus the difference in RH dependence of γc on AMD and TDD particles is not caused by the different compositions of mineral dust. It is noted that AMD was significantly aged during atmospheric transport. As shown in Figure 2, the amount of soluble inorganic salts, such as SO42− and NH4+, coated on the AMD particles is an order of magnitude higher than those on TDD and ATD particles. Some cation ions (e.g., NH4+ and Mg2+) on AMD particles can be hydrolyzed by adsorbed water to produce H+ via the reactions R1 and R2,40 resulting in the increase of acidity of the particle surface.

Downloaded by UNIV OF MICHIGAN ANN ARBOR on September 14, 2015 | http://pubs.acs.org Publication Date (Web): August 27, 2015 | doi: 10.1021/acs.est.5b03930

RH on AMD, TDD, and ATD particles can be empirically described using eqs 12−14, respectively γc,AMD =

2.04 × 10−4 1 + 1.2αH2O2.26

(12)

γc,TDD =

2.13 × 10−4 1 − 0.5αH2O0.74

(13)

γc,ATD =

3.35 × 10−5 1 − 0.7αH2O0.89

(14)

where αH2O is water activity (αH2O = RH/100). It can be seen that the γc is not significantly influenced by the presence of water vapor at lower RHs (80% RH),58 providing an experimental basis for the models studies. Our results show that the uptake coefficient of SO2 is influenced by the composition and aging extent of mineral dust, especially under humid conditions. Given that the mineralogy compositions of mineral dust from different source regions differ significantly and that mineral dust particles can be aged during atmospheric transport,2,63 the uptake coefficients with variable values rather than a single value should be considered in global model studies. In addition, H2O2 can promote the heterogeneous uptake of SO2 on mineral dust and enhance the formation of sulfate. The role of the heterogeneous reaction on mineral dust as a sink of SO2 and a source of sulfate in the atmosphere may be underestimated because of neglecting the enhancement effect of H2O2 in current atmospheric chemical models.



REFERENCES

(1) Andreae, M. O.; Rosenfeld, D. Aerosol-cloud-precipitation interactions. Part 1. The nature and sources of cloud-active aerosols. Earth-Sci. Rev. 2008, 89, 13−41. (2) Formenti, P.; Schutz, L.; Balkanski, Y.; Desboeufs, K.; Ebert, M.; Kandler, K.; Petzold, A.; Scheuvens, D.; Weinbruch, S.; Zhang, D. Recent progress in understanding physical and chemical properties of African and Asian mineral dust. Atmos. Chem. Phys. 2011, 11, 8231− 8256. (3) Maring, H.; Savoie, D. L.; Izaguirre, M. A.; Custals, L.; Reid, J. S. Mineral dust aerosol size distribution change during atmospheric transport. J. Geophys. Res. 2003, 108, 8592. (4) Usher, C. R.; Michel, A. E.; Grassian, V. H. Reactions on mineral dust. Chem. Rev. 2003, 103, 4883−4939. (5) Kolb, C. E.; Cox, R. A.; Abbatt, J. P. D.; Ammann, M.; Davis, E. J.; Donaldson, D. J.; Garrett, B. C.; George, C.; Griffiths, P. T.; Hanson, D. R.; Kulmala, M.; McFiggans, G.; Poschl, U.; Riipinen, I.; Rossi, M. J.; Rudich, Y.; Wagner, P. E.; Winkler, P. M.; Worsnop, D. R.; O'Dowd, C. D. An overview of current issues in the uptake of atmospheric trace gases by aerosols and clouds. Atmos. Chem. Phys. 2010, 10, 10561−10605. (6) Shen, X. L.; Zhao, Y.; Chen, Z. M.; Huang, D. Heterogeneous reactions of volatile organic compounds in the atmosphere. Atmos. Environ. 2013, 68, 297−314. (7) Zhang, R.; Khalizov, A.; Wang, L.; Hu, M.; Xu, W. Nucleation and growth of nanoparticles in the atmosphere. Chem. Rev. 2012, 112, 1957−2011. (8) Finlayson-Pitts, B. J.; Pitts, J. N. Chemistry of the Upper and Lower Atmosphere: Theory, Experiments, and Applications; Academic Press: San Diego, CA, 2000. (9) 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. 1996, 101, 22869−22889. (10) Li, J.; Wang, Z.; Zhuang, G.; Luo, G.; Sun, Y.; Wang, Q. Mixing of Asian mineral dust with anthropogenic pollutants over East Asia: a model case study of a super-duststorm in March 2010. Atmos. Chem. Phys. 2012, 12, 7591−7607. (11) Hanke, M.; Umann, B.; Uecker, J.; Arnold, F.; Bunz, H. Atmospheric measurements of gas-phase HNO3 and SO2 using chemical ionization mass spectrometry during the MINATROC field campaign 2000 on Monte Cimone. Atmos. Chem. Phys. 2003, 3, 417− 436. (12) Tang, Y. H.; Carmichael, G. R.; Kurata, G.; Uno, I.; Weber, R. J.; Song, C. H.; Guttikunda, S. K.; Woo, J. H.; Streets, D. G.; Wei, C.; Clarke, A. D.; Huebert, B.; Anderson, T. L. Impacts of dust on regional tropospheric chemistry during the ACE-Asia experiment: A model study with observations. J. Geophys. Res. 2004, 109, D19S21. (13) 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. (14) Ullerstam, M.; Johnson, M. S.; Vogt, R.; Ljungstrom, E. DRIFTS and Knudsen cell study of the heterogeneous reactivity of SO2 and NO2 on mineral dust. Atmos. Chem. Phys. 2003, 3, 2043− 2051. (15) Crowley, J. N.; Ammann, M.; Cox, R. A.; Hynes, R. G.; Jenkin, M. E.; Mellouki, A.; Rossi, M. J.; Troe, J.; Wallington, T. J. Evaluated kinetic and photochemical data for atmospheric chemistry: Volume V−heterogeneous reactions on solid substrates. Atmos. Chem. Phys. 2010, 10, 9059−9223. (16) Li, L.; Chen, Z. M.; Zhang, Y. H.; Zhu, T.; Li, J. L.; 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. (17) Zhang, X.; Zhuang, G. S.; Chen, J. M.; Wang, Y.; Wang, X.; An, Z. S.; Zhang, P. Heterogeneous reactions of sulfur dioxide on typical mineral particles. J. Phys. Chem. B 2006, 110, 12588−12596. (18) Ma, Q. X.; Liu, Y. C.; He, H. Synergistic effect between NO2 and SO2 in their adsorption and reaction on gamma-alumina. J. Phys. Chem. A 2008, 112, 6630−6635.

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.est.5b03930. Additional material as noted in the text (PDF)



Article

AUTHOR INFORMATION

Corresponding Author

* E-mail: zmchen@ pku.edu.cn; phone: 86-10-62751920. Present Address †

Now at Department of Chemistry, University of California, Irvine, California 92697, USA. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS We gratefully thank the National Natural Science Foundation of China (Grants 41275125, 21190053, 21477002) and the State Key Laboratory of Environment Simulation and Pollution Control (special fund) for financial support. G

DOI: 10.1021/acs.est.5b03930 Environ. Sci. Technol. XXXX, XXX, XXX−XXX

Article

Downloaded by UNIV OF MICHIGAN ANN ARBOR on September 14, 2015 | http://pubs.acs.org Publication Date (Web): August 27, 2015 | doi: 10.1021/acs.est.5b03930

Environmental Science & Technology

(42) Zhao, Y.; Chen, Z.; Shen, X.; Huang, D. Heterogeneous reactions of gaseous hydrogen peroxide on pristine and acidic gasprocessed calcium carbonate particles: effects of relative humidity and surface coverage of coating. Atmos. Environ. 2013, 67, 63−72. (43) Zhao, Y.; Chen, Z.; Shen, X.; Zhang, X. Kinetics and mechanisms of heterogeneous reaction of gaseous hydrogen peroxide on mineral oxide particles. Environ. Sci. Technol. 2011, 45, 3317−3324. (44) Miller, C. M.; Valentine, R. L. Mechanistic studies of surface catalyzed H2O2 decomposition and contanminant degradation in the presence of sand. Water Res. 1999, 33, 2805−2816. (45) Lin, S. S.; Gurol, M. D. Catalytic decomposition of hydrogen peroxide on iron oxide: Kinetics, mechanism, and implications. Environ. Sci. Technol. 1998, 32, 1417−1423. (46) Do, S. H.; Batchelor, B.; Lee, H. K.; Kong, S. H. Hydrogen peroxide decomposition on manganese oxide (pyrolusite): Kinetics, intermediates, and mechanism. Chemosphere 2009, 75, 8−12. (47) Petigara, B. R.; Blough, N. V.; Mignerey, A. C. Mechanisms of hydrogen peroxide decomposition in soils. Environ. Sci. Technol. 2002, 36, 639−645. (48) Romanias, M. N.; El Zein, A.; Bedjanian, Y. Heterogeneous interaction of H2O2 with TiO2 surface under dark and UV Light irradiation conditions. J. Phys. Chem. A 2012, 116, 8191−8200. (49) Wu, Q. Q.; Huang, L. B.; Liang, H.; Zhao, Y.; Huang, D.; Chen, Z. M. Heterogeneous reaction of peroxyacetic acid and hydrogen peroxide on ambient aerosol particles under dry and humid conditions: kinetics, mechanism and implications. Atmos. Chem. Phys. 2015, 15, 6851−6866. (50) Rao, Z. H.; Chen, Z. M.; Liang, H.; Huang, L. B.; Huang, D. Carbonyl compounds over urban Beijing: Concentrations on haze and non-haze days and effects on radical chemistry. Atmos. Environ. 2015, DOI: 10.1016/j.atmosenv.2015.06.050. (51) Yang, Q.; Su, H.; Li, X.; Cheng, Y.; Lu, K.; Cheng, P.; Zhang, Y. Daytime HONO formation in the suburban area of the megacity Beijing, China. Sci. China: Chem. 2014, 57, 1032−1042. (52) Seinfeld, J. H.; Pandis, S. N. Atmospheric Chemistry and Physics: From Air Pollution to Climate Change, 2nd ed.; Wiley: New York, 2006. (53) Bauer, S. E.; Mishchenko, M. I.; Lacis, A. A.; Zhang, S.; Perlwitz, J.; Metzger, S. M. Do sulfate and nitrate coatings on mineral dust have important effects on radiative properties and climate modeling? J. Geophys. Res. 2007, 112, D06307. (54) Levin, Z.; Ganor, E.; Gladstein, V. The effects of desert particles coated with sulfate on rain formation in the eastern Mediterranean. J. Appl. Meteor. 1996, 3, 1511−1523. (55) Archuleta, C. M.; DeMott, P. J.; Kreidenweis, S. M. Ice nucleation by surrogates for atmospheric mineral dust and mineral dust/sulfate particles at cirrus temperatures. Atmos. Chem. Phys. 2005, 5, 2617−2634. (56) Eastwood, M. L.; Cremel, S.; Wheeler, M.; Murray, B. J.; Girard, E.; Bertram, A. K. Effects of sulfuric acid and ammonium sulfate coatings on the ice nucleation properties of kaolinite particles. Geophys. Res. Lett. 2009, 36, L02811. (57) Greenwood, N. N.; Earnshaw, A. Chemistry of the Elements, 2nd ed.; Butterworth-Heinemann: Oxford, UK, 1997. (58) Huang, X.; Song, Y.; Zhao, C.; Li, M. M.; Zhu, T.; Zhang, Q.; Zhang, X. Pathways of sulfate enhancement by natural and anthropogenic mineral aerosols in China. J. Geophys. Res. Atmos. 2014, 119, 14165−14179. (59) Wang, Y.; Zhang, Q.; Jiang, J.; Zhou, W.; Wang, B.; He, K. B.; Xie, Y. Enhanced sulfate formation during China’s severe winter haze episode in January 2013 missing from current models. J. Geophys. Res. Atmos. 2014, 119, 410−425. (60) Song, C. H.; Kim, C. M.; Lee, Y. J.; Carmichael, G. R.; Lee, B. K.; Lee, D. S. An evaluation of reaction probabilities of sulfate and nitrate precursors onto East Asian dust particles. J. Geophys. Res. 2007, 112, D18206. (61) Fairlie, T. D.; Jacob, D. J.; Dibb, J. E.; Alexander, B.; Avery, M. A.; Donkelaar, A. V.; Zhang, L. Impact of mineral dust on nitrate, sulfate, and ozone in transpacific Asian pollution plumes. Atmos. Chem. Phys. 2010, 10, 3999−4012.

(19) Fu, H. B.; Wang, X.; Wu, H. B.; Yin, Y.; Chen, J. M. Heterogeneous uptake and oxidation of SO2 on iron oxides. J. Phys. Chem. C 2007, 111, 6077−6085. (20) Zhou, L.; Wang, W. G.; Gai, Y. B.; Ge, M. F. Knudsen cell and smog chamber study of the heterogeneous uptake of sulfur dioxide on Chinese mineral dust. J. Environ. Sci. 2014, 26, 2423−2433. (21) 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. 2002, 107, 4713. (22) 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. (23) Adams, J. W.; Rodriguez, D.; Cox, R. A. The uptake of SO2 on Saharan dust: a flow tube study. Atmos. Chem. Phys. 2005, 5, 2679− 2689. (24) Rubasinghege, G.; Grassian, V. H. Role(s) of Adsorbed water in the surface chemistry of environmental interfaces. Chem. Commun. 2013, 49, 3071−3094. (25) Reeves, C. E.; Penkett, S. A. Measurements of peroxides and what they tell us. Chem. Rev. 2003, 103, 5199−5218. (26) Calvert, J. G.; Lazrus, A.; Kok, G. L.; Heikes, B. G.; Walega, J. G.; Lind, J.; Cantrell, C. A. Chemical mechanisms of acid generation in the troposphere. Nature 1985, 317, 27−35. (27) Clegg, S. M.; Abbatt, J. P. D. Oxidation of SO2 by H2O2 on ice surfaces at 228 K: a sink for SO2 in ice clouds. Atmos. Chem. Phys. 2001, 1, 73−78. (28) Hua, W.; Chen, Z. M.; Jie, C. Y.; Kondo, Y.; Hofzumahaus, A.; Takegawa, N.; Chang, C. C.; Lu, K. D.; Miyazaki, Y.; Kita, K.; Wang, H. L.; Zhang, Y. H.; Hu, M. Atmospheric hydrogen peroxide and organic hydroperoxides during PRIDE-PRD′06, China: Their concentration, formation mechanism and contribution to secondary aerosols. Atmos. Chem. Phys. 2008, 8, 6755−6773. (29) Zhao, Y.; Huang, D.; Huang, L. B.; Chen, Z. M. Hydrogen peroxide enhances the oxidation of oxygenated volatile organic compounds on mineral dust particles: a case study of methacrolein. Environ. Sci. Technol. 2014, 48, 10614−10623. (30) Zhao, Y.; Chen, Z. M.; Zhao, J. N. Heterogeneous reactions of methacrolein and methyl vinyl ketone on α-Al2O3 particles. Environ. Sci. Technol. 2010, 44, 2035−2041. (31) Molina, M. J.; Molina, L. T.; Golden, D. M. Environmental chemistry (gas and gas-solid interactions): The role of physical chemistry. J. Phys. Chem. 1996, 100, 12888−12896. (32) Keyser, L. F.; Moore, S. B.; Leu, M.-T. Surface reaction and pore diffusion in flow-tube reactors. J. Phys. Chem. 1991, 95, 5496−5502. (33) Underwood, G. M.; Li, P.; Al-Abadleh, H.; Grassian, V. H. A Knudsen cell study of the heterogeneous reactivity of nitric acid on oxide and mineral dust particles. J. Phys. Chem. A 2001, 105, 6609− 6620. (34) Clark, A. The Theory of Adsorption and Catalysis; Academic: New York, 1970. (35) Aris, R. The Mathematical Theory of Diffusion and Reaction in Permeable Catalysts; Clarendon: Oxford, UK, 1975. (36) Satterfield, C. N. Mars Transfer in Heterogeneous Catalysis; MIT Press: Cambridge, MA, 1970. (37) Sattertield, C. N. Heterogeneous Catalysis in Industrial Practice, 2nd ed.; McGraw-Hill: New York, 1991. (38) Al-Hosney, H. A.; Grassian, V. H. Water, sulfur dioxide and nitric acid adsorption on calcium carbonate: A transmission and ATRFTIR study. Phys. Chem. Chem. Phys. 2005, 7, 1266−1276. (39) Prince, A. P.; Kleiber, P.; Grassian, V. H.; Young, M. A. Heterogeneous interactions of calcite aerosol with sulfur dioxide and sulfur dioxide-nitric acid mixtures. Phys. Chem. Chem. Phys. 2007, 9, 3432−3439. (40) Park, J. H.; Ivanov, A. V.; Molina, M. J. Effect of relative humidity on OH uptake by surfaces of atmospheric importance. J. Phys. Chem. A 2008, 112, 6968−6977. (41) Clegg, S. M.; Abbatt, J. P. D. Uptake of gas-phase SO2 and H2O2 by ice surfaces: Dependence on partial pressure, temperature, and surface acidity. J. Phys. Chem. A 2001, 105, 6630−6636. H

DOI: 10.1021/acs.est.5b03930 Environ. Sci. Technol. XXXX, XXX, XXX−XXX

Article

Environmental Science & Technology

Downloaded by UNIV OF MICHIGAN ANN ARBOR on September 14, 2015 | http://pubs.acs.org Publication Date (Web): August 27, 2015 | doi: 10.1021/acs.est.5b03930

(62) Zhang, Y.; Carmichael, G. R. The role of mineral aerosol in tropospheric chemistry in East Asia: A model study. J. Appl. Meteorol. 1999, 38, 353−366. (63) Sullivan, R. C.; Guazzotti, S. A.; Sodeman, D. A.; Prather, K. A. 2007. Direct observations of the atmospheric processing of Asian mineral dust. Atmos. Chem. Phys. 2007, 7, 1213−1236.

I

DOI: 10.1021/acs.est.5b03930 Environ. Sci. Technol. XXXX, XXX, XXX−XXX