Efficient Thermal Reactions of Sulfur Dioxide on Ice Surfaces at Low

B3LYP/6-31++G(d)//HF/6-31G(d), 51.0. Previous Studies. H2O–SO2, QCISD(T)/6-31G(d)//MP2/6-31G(d), 128.9. G3 (MP2), 157.7. 2H2O–SO2, G3 (MP2), 100.0...
2 downloads 11 Views 1MB Size
Subscriber access provided by GRIFFITH UNIVERSITY

Article

Efficient Thermal Reactions of Sulfur Dioxide on Ice Surfaces at Low Temperature: A Combined Experimental and Theoretical Study Jaehyeock Bang, Mahbubul Alam Shoaib, Cheol Ho Choi, and Heon Kang ACS Earth Space Chem., Just Accepted Manuscript • DOI: 10.1021/ acsearthspacechem.7b00064 • Publication Date (Web): 06 Sep 2017 Downloaded from http://pubs.acs.org on September 10, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

ACS Earth and Space Chemistry is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 22

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Earth and Space Chemistry

Submitted to ACS Earth and Space Chemistry (revised version)

Efficient Thermal Reactions of Sulfur Dioxide on Ice Surfaces at Low Temperature: A Combined Experimental and Theoretical Study

J. Bang,a M. A. Shoaib,b C. H. Choi,b,* and H. Kang a,*

a

Department of Chemistry, Seoul National University, 1 Gwanak-ro, Seoul 151-747, Republic of

Korea b

Department of Chemistry and Green-Nano Materials Research Center, College of Natural Sciences,

Kyungpook National University, Taegu 702-701, Republic of Korea.

Abstract The interaction of sulfur dioxide (SO2) gas with a crystalline ice surface at low temperature was studied by analyzing the surface species with low energy sputtering (LES) and reactive ion scattering methods, and the desorbing gases with temperature-programmed desorption mass spectrometry. The study gives direct evidence for the occurrence of efficient hydrolysis of SO2 with low energy barriers on the ice surface. Adsorbed SO2 molecules react with the ice surface at temperatures above ~90 K to form anionic molecular species, which were detected by OH–, SO2–, and HSO3– emission signals in the LES experiments. Heating the sample above ~120 K causes the desorption of SO2 gas from the surface-bound hydrolysis products. As a result, the hydrolysis of SO2 on an ice surface is most efficient at 100–120 K. The surface products formed at these temperatures correspond to metastable states, which are kinetically isolated on the cold surface. Quantum mechanical calculations of a model ice system suggest plausible mechanistic pathways for how physisorbed SO2 is transformed into chemisorbed HSO3– species. HSO3– is formed either by direct conversion of physisorbed SO2 or through the formation of a stable H2SO3 surface complex, both involving proton transfers on the ice surface with low energy barriers. These findings suggest the possibility that thermal reactions of SO2 occur efficiently on the ice surface of Jovian satellites, even without bombardment by high-energy radiation.

1

ACS Paragon Plus Environment

ACS Earth and Space Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Keywords : Sulfur Dioxide, Ice, Jovian Satellites, Hydrolysis, Proton Transfer, Reaction Mechanism

1. Introduction Fundamental studies of the interactions of sulfur dioxide (SO2) gas with ice surfaces are important for the understanding of the heterogeneous reactions of SO2 on ice surfaces in the Earth's atmosphere1 or the outer planets of the solar system.2-3 The interaction of SO2 with ice in the temperature range of the Earth's atmosphere (190–273 K) have been extensively studied.1 Ice surfaces are covered with quasiliquid layers at these temperatures.4 The reactivity of an ice surface under these conditions is predominantly influenced by this quasi-liquid layer and may be understood as an extension of the reaction behavior of the surface of liquid water. The present study focuses on the interactions of SO2 with an ice surface at lower temperatures (90–160 K) in which the surface is devoid of a quasi-liquid layer and is stable under vacuum without sublimation. This environment is relevant to Jupiter’s icy satellites, including Europa (86–132 K), Ganymede (~124 K), and Callisto (~115 K),2 where remote sensing has detected surfaces with significant amounts of water ice and minor amounts of SO2, CO2, and H2O2.2-3, 5-7 Several research groups have studied the interactions of SO2 gas with ice surfaces below the temperature of quasi-liquid layer formation.7-11 Devlin and coworkers8-9 examined the interactions of ice nanoparticles with SO2 gas in the 120–140 K temperature range using infrared (IR) spectroscopy. They observed IR absorption bands for SO2 molecular species,8 as well as bands that are similar to those of SO2 in aqueous solution,9 and assigned the latter to anionic products arising from SO2 ionization based on an ab initio normal-mode analysis of HSO3–. Using IR spectroscopy, Loeffler and Hudson7 reported that sulfur oxyanions such as HSO3– and S2O52– are formed in H2O-SO2 mixed films at cryogenic temperatures; S2O52– is formed by the reaction between two HSO3– molecules at high SO2 concentrations. Photoelectron spectroscopic studies10 of SO2 adsorption on ice films at 100 K indicated the presence of molecular SO2 adsorbates as well as sulfuric-acid-like species with distinguishable S(2p) electron binding energies. Kim et al.11 examined the reaction of SO2 with a frozen water film on a Ru(0001) substrate using low energy sputtering (LES) and reactive ion scattering (RIS) techniques; various ions related to SO2 hydrolysis on the ice surface were detected. However, thin frozen-water films undergo a roughening phase transition upon heating,12-13 and this structural change may affect reactions on the sample surface. To avoid such complications, in the present work we prepared a thick crystalline ice film on a Pt(111) substrate, which remains stable during sample heating to ~160 K without undergoing any phase transition. The surface products resulting from the interaction of SO2 gas with the ice surface, as well as the desorbing gases, were examined as functions of temperature by conducting low energy sputtering (LES), reactive ion 2

ACS Paragon Plus Environment

Page 2 of 22

Page 3 of 22

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Earth and Space Chemistry

scattering (RIS), and temperature-programmed desorption (TPD) experiments. Quantum mechanical calculations for a model ice system were performed in order to understand the reaction mechanism of SO2 on the ice surface.

2. Methods 2.1. Experimental section Experiments were conducted in an ultrahigh vacuum (UHV) surface analysis chamber14-15 equipped with instrumentation for RIS, LES, and TPD experiments. An ice film was grown on the (111) face of a Pt single crystal by back-filling the chamber with H2O vapor at a partial pressure of 1.0 × 10–7 Torr. The substrate temperature was maintained at 150 K during the growth of the ice film, and the film was flash-annealed at 165 K. It is known16-19 that this procedure produces a non-porous crystalline ice film with predominant Ih(0001) surfaces. The thicknesses of the ice films were typically 90–110 ML (monolayer; 1 H2O ML = 1.1 × 1019 molecules m-2), as estimated from the water desorption signal in the TPD experiment. The growth of the ice films took about 15 min under these conditions. SO2 (99.99+% purity) gas was introduced into the chamber through a separate leak valve and guided close to the sample surface through a tube doser to minimize chamber wall contamination. SO2 was adsorbed onto the ice film at 95 K. The amount of adsorbed SO2 gas was estimated from TPD measurements. The intensity of the SO2 desorption signal was converted to SO2 coverage through calibration against water desorption signal from H2O monolayer on Pt(111),13 taking into account the ionization cross-section and fragmentation pattern of SO2 and H2O gases. Chemical species present on the ice film surface were analyzed by using RIS and LES methods. The principles of these methods have been described previously.15, 20 In these experiments, a Cs+ beam from a low-energy ion gun (Kimball Physics) collided with the sample surface at an incident energy of 35 eV. The positive and negative ions emitted from the surface were detected by a quadrupole mass spectrometer (ABB Extrel) located at the scattering angle of 55°, with its ionizer disabled. During RIS, neutral species (X) at the surface are picked up by the scattering Cs+ projectiles to form Cs+–neutral clusters (CsX+). During LES, ionic species (Y+ and Z–) at the surface are ejected by Cs+ impact. Thus, the RIS and LES signals reveal the identities of neutral (X) and ionic (Y+ and Z–) species present on the ice film surface. These methods have an ice-surface probing depth of one molecular layer.14 Under these conditions the LES signals mostly represent the preexisting ionic species on the surface, while secondary ion emissions resulting from Cs+ impact become significant only at substantially higher collision energies.20

3

ACS Paragon Plus Environment

ACS Earth and Space Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

2.2. Theoretical methods a) Ice surface model The ice model was constructed to have a crystalline ice (Ih) structure with a (0001) surface plane. The model consisted of bilayers (BLs) of tetrahedrally bonded H2O molecules connected by hydrogen bonds; the oxygen planes were separated by 2.75 Å between bilayers, and 0.92 Å within the bilayer. The Ih ice structure has bilayers in an …ABCABC… stacking sequence. Each O atom was surrounded by four other hydrogen-bonded oxygens that are 2.75 Å away. The H atoms were not centered between two O atoms in the hydrogen bonds but remained off-center by about 0.38 Å in a random distribution that satisfies the Bernal–Fowler structure.4 The (0001) surface of the ice model had “full bilayer” terminations, which is energetically favored (by 30 kJ/mol per unit cell) over the “half-bilayer” terminated surface21 because of the larger number of intermolecular hydrogen bonds. In the “full-bilayer” terminated surface, each of the outermost O atoms in the upper layer is hydrogenbonded to three neighbors of the lower layer. It has been reported13 that the “full bilayer” terminated surface is predominantly formed when an ice film is annealed at ~150 K, the condition employed in the present experiment.

Figure 1. Surface structures depicting (a) an oxygen dangling bond (ODB), and (b) a hydrogen dangling bond (HDB) in clusters of four water molecules.

At the basal surface of the “full bilayer” terminated ice Ih structure, each of the outermost water molecules has one dangling bond, either an oxygen dangling bond (ODB) or a hydrogen dangling bond (HDB), that points outwards, as shown in Figures 1a and 1b, respectively. These dangling bond sites are considered to be the primary reactive places due to their outermost positions on the ice surface. Their statistical distributions on the ice surface reflect the microscopic heterogeneity of the 4

ACS Paragon Plus Environment

Page 4 of 22

Page 5 of 22

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Earth and Space Chemistry

surface, which in turn provide various reaction environments, including multiple interaction sites for adsorbates. We employed a systematic strategy to build the model ice surfaces. The coordinates of a reference ice crystal of 36 hexagons by 15 hexagons and 5 BLs deep were generated by randomizing the hydrogen-bonding distribution to ensure a (near to) zero dipole moment of the crystal. The internal coordinates of the water molecules in the reference ice crystal were set such that the O-O distances, O-H distances, and O-O-O angles matched the experimentally determined values of 2.75 Å, 1.00 Å, and 109.3º, respectively. The specific combination of QM (Quantum Mechanics) and EFP (Effective Fragment Potential) schemes in this paper involves a two-layer model (QM-relaxed region/EFP water), where the coordinates in the QM-relaxed region are fully relaxed during geometry optimization. A large ice cluster consisting of 27 QM waters and 457 EFP waters, referred to as “HDB2(27/457)”, is depicted in Figure 2. The “HDB2” term in this nomenclature system refers to a hexagonal binding site with two HDBs, which are indicated by the arrows in Figure 2a. This model simultaneously includes both the effect of geometry relaxation due to adsorption, and the long-range electrostatic interactions of ice crystals.

ODB HDB

(a)

(b)

Figure 2. (a) QM region (27 waters) and (b) QM+EFP regions (457 waters) of the HDB2(27/457) model cluster. The SO2 adsorption sites are marked by arrows. This hexagonal binding site has two nearby HDBs and one ODB.

5

ACS Paragon Plus Environment

ACS Earth and Space Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

b) Computational details All stationary points for the HDB2(27/0/457) ice model were calculated at the MP2/631++G(d)//HF/6-31G(d) and B3LYP/6-31++G(d)//HF/6-31G(d) levels of theory. The energies obtained using MP2/6-31++G(d)//HF/6-31G(d) are discussed in the sections below. All calculations were performed without any symmetry constraints, unless otherwise specified. The GAMESS22-23 program was used for all calculations.

3. Results 3.1. Experimental results TPD, RIS, and LES measurements were conducted in order to study the reactions of SO2 gas on a crystalline H2O film surface. Figure 3 displays the TPD SO2 spectra of ice films adsorbed with different surface coverages (0.04–0.65 MLE; 1 monolayer equivalent of H2O = 1.1 × 1019 molecules m-2) of SO2. The desorption of SO2 was observed to occur over a broad temperature range (105–140 K), and the desorption profile changed with changes in SO2 coverage. At the lowest coverage (0.04 MLE), the desorption profile exhibits a single peak at temperatures of 120–140 K. At higher coverages, the profiles are broadened toward lower temperatures. These broadened profiles were deconvoluted into two desorption components, one at 120–140 K and the other at 105–130 K. The low-temperature peak grew in proportion to SO2 coverage above ~0.1 MLE, while the growth of the high-temperature peak, with SO2 coverage, was much less pronounced. These peak positions did not significantly change at moderately low SO2 coverage (0.10–0.28 MLE), indicating that they correspond to first-order desorptions. Accordingly, we assign the low-temperature peak to the desorption of physisorbed SO2 species, and the high-temperature peak to ‘chemisorbed’ (strongly bound) SO2 species. The low-temperature peak shifted slightly to a higher temperature at high coverage (0.28–0.65 MLE). This indicates a certain degree of contribution of zeroth order desorption, which may be attributed to SO2 cluster or multilayer formation at high coverage. The hightemperature peak grew by about three times in intensity as SO2 coverage was increased from 0.04 MLE to 0.65 MLE. Deconvolution of the TPD curve at 0.65 MLE into two peaks indicates that ‘chemisorbed’ SO2 species had a coverage of ~0.16 MLE. This coverage, which was near the saturation point, seems reasonable because, as will be shown in Section 3.2, SO2 adsorption sites must have at least one ODB and one HDB, and strong SO2 adsorption typically occupies two HDBs, where the surface density of HDB is 0.25 ML.4 Most of SO2 species desorbed from the surface before the onset of water desorption at ~150 K, with only insignificant overlap observed between the hightemperature SO2-desorption tail and the peak corresponding to water desorption. The well separated 6

ACS Paragon Plus Environment

Page 6 of 22

Page 7 of 22

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Earth and Space Chemistry

desorption peaks observed for SO2 and water indicate that SO2 did not dissolve into the ice sample.

Figure 3. TPD spectra for SO2 (m/z = 64) measured from ice films (90 ML thickness) covered with adsorbed SO2 layers of different coverages (0.04–0.65 MLE). SO2 gas was adsorbed at 95 K. The H2O TPD spectrum from the sample covered by a 0.04-MLE thick adsorbed SO2 layer is shown in gray shading; this spectrum was recorded at m/z = 17 (HO+) to avoid intensity saturation of the mass spectrum. The SO2 TPD curves were deconvoluted by subtracting the high-temperature (120–140 K) desorption peak form the original spectrum, and the remaining portion was regarded as the lowtemperature (105–130 K) peak. The temperature ramping rate was 0.5 K s–1.

Figure 4a shows the RIS mass spectrum of an ice film covered by a small amount (0.04 MLE) of SO2 at 95 K. The large peak at m/z = 133 corresponds to the primary Cs+ ions. RIS signals include Cs(H2O)+ at m/z = 151, due to the pickup of H2O by the Cs+ projectiles, and Cs(SO2)+ at m/z = 197 and 199 (magnified in the inset), due to the pickup of 32SO2 and 34SO2. A very low intensity peak is evident at m/z = 215, which is assigned to Cs(H2O)(SO2)+ produced by the double collection of H2O and SO2. The low intensity of this signal is consistent with the double pickup process. Alternatively, this signal might also correspond to Cs(H2SO3)+, which would indicate the formation of H2SO3 by the reaction of SO2 with the ice surface. This interpretation, however, is not supported by the LES 7

ACS Paragon Plus Environment

ACS Earth and Space Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

measurements, vide infra (Figure 5), which show no evidence of surface H2SO3 species at 95 K.

Figure 4. (a) The RIS mass spectrum of a sample measured at 95 K. The signals at m/z > 190 are magnified in the inset. (b) The LES mass spectrum of the negative ions from the same sample obtained by signal accumulation during a 113 to 118 K temperature scan. The sample consisted of a crystalline ice film (90 ML thickness) covered with a SO2 adsorbed layer (0.04 MLE). The Cs+ beam energy was 35 eV.

Figure 4b depicts the LES mass spectrum of the negative ions measured from the sample described above. In this experiment, the sample temperature was gradually increased from 113 to 118 K during 8

ACS Paragon Plus Environment

Page 8 of 22

Page 9 of 22

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Earth and Space Chemistry

which time the spectrum was acquired through repeated scans. Consequently, the spectrum displays all anionic species appearing on the sample surface in this temperature range. This range was selected because most of the LES signals were very intense over these temperatures; these signals were almost negligible at 90 K. This spectrum shows LES signals corresponding to OH– at m/z = 17, SO2– at m/z = 64, and HSO3– at m/z = 81. These ion emissions indicate that SO2 has reacted to produce negatively charged molecular species on the surface. The LES measurements did not detect any cationic signals on the surface (data not shown). If we interpret this straightforwardly, cationic species are simply not formed on the surface. Alternatively, cationic species are formed, but they are difficult to detect by LES because, for instance, they are formed in a subsurface region or as part of a large molecular structure that is difficult to desorb by Cs+ impact. The latter possibility will be further discussed in Section 4 when the SO2 reaction mechanism is presented with the aid with theoretical calculations. We examined the reaction of SO2 as a function of temperature by conducting temperatureprogrammed LES (TPLES) and RIS (TPRIS) experiments. Figure 5 shows the results of these experiments for the signals of interest over a 90 to 160 K temperature scan. The sample was prepared in identical fashion to that used to generate the results shown in Figure 4. The RIS intensity for Cs(SO2)+ is plotted in terms of its RIS yield, which is defined as the intensity ratio of Cs(SO2)+/ΣiCsXi+, where ΣiCsXi+ is the sum of the intensities of all RIS signals. The RIS yield is a more reliable measure of surface population than the RIS intensity; various circumstantial factors that affect the RIS intensity, in particular the changes in surface morphology, are canceled out by dividing with ΣiCsXi+.15 In support of this rationale, the RIS yield for the H2O signal from the sample surface remained almost constant with increasing temperature, until the point of water desorption was reached at around 150 K (not shown). Beyond the water desorption temperature, the RIS and LES intensity variations were no longer proportional to the surface population changes because of severe surface modification. The TPRIS curve for SO2 shown in Figure 5 indicates that the surface population of SO2 decreases monotonically as the temperature is increased from 90 to 110 K.

9

ACS Paragon Plus Environment

ACS Earth and Space Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 5. TPRIS and TPLES measurements for a 0.04 MLE adsorbed-SO2 ice film. The SO2 RIS yield (black line; left ordinate scale) monitors the population of surface SO2 species. The TPLES curves show the signal intensities of OH–, SO2–, and HSO3– (right ordinate), which indicate the populations of the surface anionic species. The TPD spectra of SO2 and H2O are also shown in gray and red shading, respectively; these indicate the temperature regions of the corresponding gas desorptions. The temperature ramping rate was 0.5 K s–1, and the Cs+ beam energy was 35 eV.

TPLES signals for OH–, SO2–, and HSO3– appeared at 90–100 K and grew stronger with increasing temperature until they reached their maximum values at 110–120 K. The TPLES intensities shown were not calibrated against any internal standard (e.g., externally added halide ions).7 Therefore, the TPLES profiles not only reflect changes in the surface ion populations, but also surface structural changes due to reactions with SO2. At this point, we neglect the latter effect in interpreting the TPLES curves. Two main features are noticeable in Figure 5. Firstly, increases in the intensities of the OH–, SO2–, and HSO3– signals over the 90–115 K range correlate with decreases in the SO2 surface population. This indicates that SO2 adsorbates are hydrolyzed in this temperature region. The emitted anions may correspond to hydrolysis products themselves or originate from the collisional fragmentation of hydrolysis products with larger structure; these two possibilities cannot be 10

ACS Paragon Plus Environment

Page 10 of 22

Page 11 of 22

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Earth and Space Chemistry

distinguished at this point. Nevertheless, the TPLES curves refute the possibility that the anionic signals arise from the secondary ionization of SO2 molecules; this conclusion is based on the negative correlation between the intensities of the anionic signals and the SO2 adsorbate population. Secondly, the hydrolysis of SO2 precedes the desorption of SO2 gas by a substantial temperature gap, as can be seen from the different onset temperatures for SO2 hydrolysis (90–100 K) and SO2 desorption (~110 K). Rather, the increased rate of SO2 desorption at 110–130 K correlates with decreases in the anionic signal intensities. These observations indicate that the SO2 gas is evolved from hydrolysis products on the surface, rather than from physisorbed SO2 species. The TPRIS and TPLES data indicate that physisorbed SO2 molecules on the ice surface are efficiently hydrolyzed at 90–120 K. This hydrolysis efficiency must be of the order of unity since the majority of the physisorbed SO2 molecules “disappeared” without any SO2 gas desorption over this temperature range. A direct estimate of the efficiency for the conversion of SO2 into hydrolysis products, however, is difficult to make because the TPLES intensities for the various ions may not quantitatively reflect their surface populations. Only the results for low SO2 coverage (0.04 MLE) have been discussed so far. When the SO2 coverage was increased to 0.1–0.65 MLE, a few changes were observed in the corresponding spectra. Firstly, as shown in Figure 3, the low-temperature SO2 desorption peak at 105–130 K in the TPD spectra gradually increased in intensity with increasing SO2 coverage. In addition, in the TPRIS experiments, the decay of the SO2 RIS-yield curve shifted to higher temperatures with increasing SO2 coverage (not shown), compared to the original decay observed over the 90–110 K region at a SO2 coverage of 0.04 MLE (Figure 5). The shifted decay of TPRIS curve nicely correlated with the increased intensity of the low temperature peak in TPD spectra (Figure 3); the correlation is based on the relationship that the first derivative of TPRIS curve with temperature (dθ/dT) is proportional to TPD intensity (desorption flux).19 Therefore, under these conditions a large portion of growing populations of physisorbed SO2 species must be directly desorbed from the surface during the sample heating, rather than undergoing reactions. As a result, the hydrolysis efficiency of SO2 will decrease at higher SO2 coverage. The change in SO2 hydrolysis efficiency can be seen from the changing TPD profiles in Figure 3. The hydrolysis yield can be defined as the population ratio of ‘chemisorbed’ SO2 species to all adsorbed SO2 species, which, in turn, is proportional to the areal ratio of the high temperature (120–140 K) peak to the whole desorption profile, if we neglect the dissolved species in the sample that do not desorb. Because the rate for the increase of the high-temperature peak is much slower than that for the low-temperature peak, the hydrolysis efficiency is correspondingly lower at higher SO2 coverage. Based on the results shown in Figures 3–5, we propose the following reaction sequence for the 11

ACS Paragon Plus Environment

ACS Earth and Space Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

hydrolysis of SO2 on an ice surface. Upon adsorption of SO2 gas at 90–120 K, SO2 becomes transformed into anionic molecular species that produce LES signals for OH–, SO2–, and HSO3–. This process occurs efficiently without SO2 desorption at this low temperature. The surface products do not penetrate to the interior of the ice during sample heating, as indicated by the well-separated TPD signals for SO2 and water (Figure 3). Rather, the surface species are transformed back into SO2 gas above ~120 K. Therefore, the surface species formed at 100–120 K are considered to be metastable products of SO2 hydrolysis that are trapped on the cold surface, rather than representing final reaction products that are obtained by reactions in aqueous solution. The desorption of SO2 gas from the ice surface may be kinetically favored over the penetration of SO2 into the ice interior, thereby effectively blocking reaction pathways that lead to solvation and the further hydrolysis of SO2 to energetically more stable products. 3.2. QM study of the interaction of SO2 with an ice surface When SO2 is adsorbed on an ice surface, the ODBs and HDBs of the surface preferentially interact with the sulfur and oxygen atoms of SO2, respectively. Therefore, appropriate SO2 adsorption sites on the surface must have at least one ODB and one HDB. With these considerations in mind, HDB2-type adsorption sites, which are the hexagonal sites with two HDBs and one ODB, were adopted. Figure 6a shows the initial adsorption structure of a physisorbed SO2 molecule on the HDB2(27/457) model ice surface. In this adsorption structure, two hydrogen bonds (O(3)-H(18) = 2.06 Å and O(2)-H(11) = 2.17 Å) are formed between SO2 and surface water molecules. The binding energy of the physisorbed structure is fairly large (97.9 kJ/mol at MP2/6-31++G(d)//HF/6-31G(d)), and is responsible for the efficient entrapment of impinging gaseous SO2 molecules on the ice surface. This adsorption energy can be used for surface chemical transformations of SO2 before quickly dissipating into the lattice. It can be noticed that the binding energy of SO2 is significantly larger than the experimental desorption energy of water (48.25 kJ/mol) from crystalline ice,24 which might appear contradictory to the observation that SO2 desorbs prior to water. Two main features need to be considered before comparing our theoretical binding energy with the desorption energy. First, the theoretical binding energy comes from the most stable physisorbed structure among other possible ones. There are three interfacial binding interactions including two hydrogen bonds, which makes the structure of Figure 6a exceptionally stable. We have also found other physisorbed structures with weaker binding energies. Second, direct SO2 desorption from Figure 6a is unlikely to occur due to its multiple interfacial bindings. Instead, desorption paths through intermediate adsorption stages with weaker binding energies are highly expected, which would significantly reduce the final desorption energy of SO2.

12

ACS Paragon Plus Environment

Page 12 of 22

Page 13 of 22

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Earth and Space Chemistry

Figure 6. Indirect and direct pathways for the formation of HSO3– from physisorbed SO2 on the HDB2(27/0/457) model ice surface. (a) The physisorbed SO2. (b, c) The initial transition states for (a) 13

ACS Paragon Plus Environment

ACS Earth and Space Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 22

indirect (TSIndrect) and (b) direct (TSDirect) pathways. (d, e) Surface-formed H2SO3 and HSO3–, respectively, as marked with blue circles. The surface H3O+ molecule is also marked with a red circle in (e). (f) Transition state (TSInter) connecting (d) and (e). The MP2/6-31++G(d)//HF/6-31G(d)calculated relative energies (kJ/mol) are shown in each figure. The physisorbed SO2 depicted in (a) serves as the reference energy point. B3LYP/6-31++G(d)//HF/6-31G(d)-calculated values are presented in the parentheses, along with hydrogen-bonding distances (Å) calculated at HF/6-31G(d).

Our theoretical investigation focused mainly on determining pathways for the hydrolysis of physisorbed SO2 that produce anionic molecular species that are observed by LES experiments. We found two plausible pathways that lead to the formation of HSO3– species. This structure is negatively charged and could produce OH–, SO2–, and HSO3– ions by either collisional fragmentation or desorption during low energy surface sputtering. SO2  H2SO3  HSO3–

: indirect path

(1a)

SO2  HSO3–

: direct path

(1b)

In reaction 1a (indirect path), a neutral H2SO3 complex is formed first, and subsequent deprotonation leads to HSO3–. The other channel directly converts physisorbed SO2 to HSO3– (reaction 1b). These two mechanisms are discussed in order. Indirect path: In this pathway, the electrophilic attack of S(1) at surface O(4) initiates the transfer of surface proton H(6)  O(7). This also produces a transient hydronium in the transition state TSIndrect (see Figure 6b) in which two hydrogen bonds are slightly shortened (O(3)···H(18) = 1.86Å, and O(2)···H(11) = 1.89Å). The formation of a S-O bond is considered to be the signature of hydrolysis, since a water molecule is added to the physisorbed SO2. TSIndrect links the physisorbed SO2 (Figure 6a) to the surface H2SO3 (Figure 6d) by concerted H(9)  O(10) and H(11)  O(2) proton transfers. The barrier height for the formation H2SO3 from the physisorbed SO2 is calculated to be 51.0 kJ/mol at MP2/6-31++G(d)//HF/6-31G(d). Transition state TSInter (Figure 6f) connects the surface H2SO3 (Figure 6d) and the deprotonated HSO3– (Figure 6e) by transferring H(11) back to O(10), while H(9) remains between O(10) and O(7) in this TS. HSO3– and a surface hydronium ion are formed from TSInter by concerted H(9)  O(7) and H(8)  O(22) proton transfers. Direct path: In this pathway, TSDirect (Figure 6c) directly connects the initial physisorbed SO2 (Figure 6a) and the surface HSO3– (Figure 6e). Electrophilic attack of molecular S(1) at surface O(4) and the synchronous transfer of surface protons H(6) and H(8) are evident in TSDirect. Geometry optimization located HSO3– as an energy minimum at the HF/6-3G(d) level of theory. However, 14

ACS Paragon Plus Environment

Page 15 of 22

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Earth and Space Chemistry

single-point energy calculations using either MP2/6-31++G(d) or B3LYP/6-31++G(d), found that the HSO3– structure is less stable than TSDirect by 3.3 kJ/mol, suggesting that the potential energy surface between the structures in Figures 6c and 6e is almost flat. In this case, the HSO3– is thermodynamically transient structure. A hydronium that can quickly migrate to other locations along the H-bonded water chain is seen in Figure 6e. Due to the limited size of our model, the transfer of protons to other regions could not be studied. However, the high mobility of protons would increase the kinetic stability of HSO3–. TSDirect lies 62.3 kJ/mol above the physisorbed SO2 and is slightly higher in energy, by 11.3 kJ/mol, than TSIndirect. Therefore, a competition exists between the direct and indirect pathways that favors the indirect channel. It should be noted that the hexagonal framework of the ice surface remains nearly intact during the entire surface hydrolysis process. Penetration of the adsorbate through the surface would result in the destruction of the surface bilayer structure and would require much higher energy than surface hydrolysis. Therefore, surface reactions of SO2 are preferred over penetration into the ice. Barrier Height for Hydrolysis: Our calculated reaction barriers for the formation H2SO3 and HSO3– are listed in Table 1 along with previous predictions from the literature. The previous studies predicted prohibitive barriers of 91.6–157.7 kJ/mol. On the other hand, our improved HDB2(27/457) model provides a barrier for the formation of H2SO3 of 51.0 kJ/mol (MP2 result), as measured from physisorbed SO2, which is 40.6 kJ/mol lower than the lowest value previously reported. With a barrier height of similar magnitude (62.3 kJ/mol), hydrolyzed HSO3– can also be formed through the direct pathway. The differences between the present and earlier predictions indicate that a large surface model, such as HDB2(27/457), is required to properly take into account surface relaxation and longrange effects in ice-surface reactions. In summary, unlike the previous theoretical predictions, the hydrolysis of SO2 is feasible on cold ice surfaces due to low reaction barriers that lead to formation of H2SO3 and HSO3–. These low barriers are associated with reaction mechanisms that involve proton transfers that can occur relatively easily on cold ice surfaces.25 The barrier heights for surface reactions are much lower than the energy required for the desorption of physisorbed SO2 (97.9 kJ/mol). Therefore, the hydrolysis of SO2 is energetically favored over the desorption of SO2 at low temperatures, which is consistent with our experimental observation. Reversible Hydrolysis: While physisorbed SO2 requires 51.0–62.3 kJ/mol to overcome the energy barrier for hydrolysis to H2SO3 or HSO3–, the reverse reactions require merely –3.3 to 15.1 kJ/mol for the transformations shown in Figures 6e  6c and 6d  6b, respectively. Therefore, the relative concentration of physisorbed SO2 would be higher than that of the hydrolyzed species. This result seems inconsistent with the efficient conversion of SO2 into hydrolyzed species observed in the 100– 120 K temperature range. However, irreversible desorption of SO2 gas from the hydrolyzed species 15

ACS Paragon Plus Environment

ACS Earth and Space Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 22

can occur, as observed at temperatures above ~110 K in the TPD experiment (Figures 3 and 5). This process shifts the equilibrium further toward the hydrolyzed species. We stress that the present energy calculations, which do not consider entropy effects, are limited in their ability to quantitatively predict the thermodynamic and kinetic behavior of these reactions.

Table 1. Energy barriers (kJ/mol) for the hydrolysis of SO2 at different adsorption sites calculated with different levels of theory. The initial physisorbed SO2 serves as the reference. This Study Model

Method

Energy barrier

TSIndirect, HDB2(27/457)

MP2/6-31++G(d)//HF/6-31G(d)

51.0

B3LYP/6-31++G(d)//HF/6-31G(d)

40.6

MP2/6-31++G(d)//HF/6-31G(d)

62.3

B3LYP/6-31++G(d)//HF/6-31G(d

51.0

QCISD(T)/6-31G(d)//MP2/6-31G(d)

128.9a

G3 (MP2)

157.7b

2H2O-SO2

G3 (MP2)

100.0 b

3H2O-SO2

G3 (MP2)

91.6 b

TSDirect, HDB2(27/457)

Previous Studies H2O-SO2

a

Bishenden, E.; Donaldson, D. J. J. Phys. Chem. A 1998, 102, 4638–4642.

b

Voegele, A. F.;

Tautermann, C. S.; Rauch, C.; Loerting, T; Liedl, K. R. J. Phys. Chem. A 2004, 108, 3859-3864.

4. Discussion The present calculations predict that HSO3– (Figure 6e) can be formed on the ice surface via either direct or indirect pathways involving proton transfers. The barriers for these pathways are sufficiently low such that they can occur even at a low temperature. Therefore, HSO3– offers a reasonable explanation for the identity of the anionic molecular species observed in the LES experiment. Intact desorption of this species upon Cs+ impact results in a signal corresponding to HSO3–. Collisional fragmentation of this structure following Cs+ impact can generate LES signals for SO2– and OH–. The 16

ACS Paragon Plus Environment

Page 17 of 22

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Earth and Space Chemistry

generation of HSO3– is consistent with the results of previous IR spectroscopic studies for H2O-SO2 mixtures prepared in the forms of nanoparticles9 and thin films.7 The relative stabilities of different adsorbates explain the occurrence of SO2 hydrolysis at a lower temperature than that for SO2 desorption. The physisorbed SO2 (Figure 6a) is energetically the most stable state on the ice surface. The energy barriers for its conversion into neutral H2SO3 and anionic HSO3– are 51.0 and 62.3 kJ/mol, respectively, which are substantially lower than the SO2 desorption energy of 97.9 kJ/mol. Consequently, hydrolysis to H2SO3 and HSO3– will first occur upon sample heating. These surface products, however, are only metastable, possibly corresponding to intermediate reaction stages rather than the final products that would be obtained in aqueous solutions. Further SO2 hydrolysis would require extensive structural rearrangement of the adsorbate and the ice surface, which is kinetically prohibitive for a solid ice surface at low temperature. On the other hand, the desorption of SO2 gas from the metastable products is entropically favored and can occur relatively easily at elevated temperatures. Accordingly, above ~120 K, H2SO3 and HSO3– are transformed back into SO2 gas to desorb from the surface. Although the LES experiments revealed molecular anions on the ice surface, their counter cations were not detected. The most likely counter cation candidate is the hydronium (H3O+) ion, which is predicted to form by the theoretical calculations (Figure 6e). SO2 gas dissolved in liquid water at room temperature produces moderate concentrations of HSO3–(aq) and H3O+(aq), with the acid dissociation constant pKa = 1.8. If about the same degree of acid ionization is assumed for the ice surface, then the surface population of H3O+ is expected to be ~0.05 MLE at a SO2 coverage of 0.4 MLE; ions of this surface concentration would readily be detected by LES. However, the counter cations were not detected despite their search under a wide range of experimental condition, including a variety of SO2 coverages (0.04–0.4 MLE) and Cs+ beam energies (up to 50 eV). One possible explanation for the absence of a H3O+ signal is that H3O+ is formed by the reaction but is not ejected from the surface by impact with Cs+. This is possible if H3O+ is very mobile in the ice lattice or exists below the surface.2526

The structure shown in Figure 6(e) is informative in terms of this explanation. Since H3O+ is formed

as part of an extended H-bonded water chain and the excess proton is mobile along the water chain with a very small barrier via the Grotthuss mechanism,25 the impact of Cs+ at low energy may only facilitate the migration of H3O+ along the water chain, making its ejection into a vacuum very unlikely to occur. As such, H3O+ formed on the ice surface will be invisible in the LES experiments. It is worth mentioning that the hydrolysis products observed in this work are somewhat different from those produced by SO2 hydrolysis on a thin (~4 ML) frozen water film reported previously.11 The reaction of SO2 on a thin (~4 ML) D2O film prepared on Ru(0001) produced LES signals corresponding to OD–, SO2–, DSO2–, and DSO3–. If we assume that isotopic difference between H2O 17

ACS Paragon Plus Environment

ACS Earth and Space Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

and D2O films does not influence the reactions, the main change in the present observation is that DSO2– is not formed. To understand this difference, we examined the hydrolysis of SO2 on a D2O monolayer prepared on a Ru(0001) substrate (Supporting Information). This surface produced signals corresponding to OD–, SO2–, DSO2–, SO3–, and DSO3–. Therefore, the ion signals produced from the thin ice films11 are explainable as a superposition of the signals originating from the surfaces of crystalline ice and a water monolayer. It is quite likely that the thin ice films underwent a roughening transition during sample heating, which generated patches of water monolayers and small ice-crystal domains.12-13 As a result, the hydrolysis products from two different surfaces appeared simultaneously. Finally, we wish to mention the implications of these discoveries for SO2 chemistry on Jovian satellites. Discussions about the reactions of SO2 on surface ices of Europa, Ganymede, and Callisto usually emphasize the effects of Jovian magnetospheric radiation, which induces radiolytic and photolytic reactions at the ice surface.2-3, 6 This interpretation, in part, stems from the idea that thermal reactions are difficult on the surface or interior of ice at the low temperatures found on these satellites. There are, however, indications that thermal reactions of SO2 can occur in ice environments at low temperature.7-11 For example, Loeffler and Hudson7 observed that thermal reactions of SO2 occurred in a solid H2O‐SO2 mixture to consume up to 30% of the SO2 molecules at 100 K, and they claimed that similar processes may take place on the ices of Jovian satellites. Our study further supports the occurrence of thermal SO2 reactions at low temperature by demonstrating that SO2 is efficiently hydrolyzed by adsorption on the ice surface, even without full solvation by water molecules as in a H2O‐SO2 mixture. The surface reaction is most efficient in the 100–120 K temperature range, and is significantly retarded at lower temperatures, while SO2 gas desorbs from the surface the higher temperatures. These favorable conditions for SO2 hydrolysis are found on Europa (86–132 K), Ganymede (~124 K), and Callisto (~115 K).2 Furthermore, our theoretical study of the hydrolysis reaction mechanism reveals that a low barrier for the reaction is related to pathways involving proton transfer, which can occur on ice surfaces even at low temperatures.25 These observations support the possibility that efficient heterogeneous reactions of SO2 occur on the ice surfaces of Jovian satellites, even without the assistance of Jovian magnetospheric radiation. One can imagine that these ices contain substantial amounts of hydronium ions and sulfurous anions, as well as other sulfur compounds resulting from unique, low-temperature reactions of SO2 with the ice surface.

5. Summary and conclusions This work reports direct evidence of SO2 hydrolysis on the ice surface through mass spectrometric detection of surface products and TPD measurement of desorbing SO2 gas. In addition, plausible hydrolysis mechanisms are suggested based on quantum chemical calculations. SO2 gas efficiently 18

ACS Paragon Plus Environment

Page 18 of 22

Page 19 of 22

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Earth and Space Chemistry

reacts with a crystalline ice surface at temperatures 100–120 K. The hydrolysis of SO2 leads to the formation of molecular anions on the surface, which is evidenced by signals in LES experiments corresponding to OH–, SO2–, and HSO3–. H3O+ may also form on the surface as the counter cation, although it was invisible to LES. A theoretical investigation into the reaction mechanism reveals that SO2 hydrolysis is aided by proton transfers on the ice surface, leading to small energy barriers that make this process feasible even at low temperatures. There are at least two SO2 hydrolysis channels that lead to HSO3– and H2SO3 surface complexes. The maximum hydrolysis yield observed at 100– 120 K is the result of competition between the thermal activation of the reaction on the ice surface and desorption of SO2 gas. The products observed in this work may correspond to metastable species that are kinetically trapped on the cold ice surface at intermediate stages of the reactions. The experimental observations and theoretically determined reaction mechanisms support the possibility that heterogeneous reactions of SO2 occur efficiently on Jovian satellites via thermal reactions on ice surfaces, even without bombardment by high energy radiation.

AUTHOR INFORMATION Corresponding authors * E-mail: [email protected]. Fax: +82 2 889 8156 (H. Kang) * E-mail: [email protected] (C. H. Choi) Author Contributions J. Bang and M. A. Shoaib contributed equally to this work.

ASSOCIATED CONTENT Supporting Information TPRIS and TPLES measurements for the reaction of SO2 on a water monolayer prepared on a Ru(0001) substrate. The Supporting Information is available free of charge on the ACS Publications website at DOI: xxxx. ACKNOWLEDGMENT We wish to thank Dr. Young-Kwang Kim and Dr. Sun-Kyung Kim for advising the thin film experiment on Ru(0001) and Professor Jin Yong Lee for the discussion with QM calculations. This work was supported by Samsung Science and Technology Foundation (SSTF-BA1301-04). REFERENCES 19

ACS Paragon Plus Environment

ACS Earth and Space Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(1) Huthwelker, T.; Ammann, M.; Peter, T. The Uptake of Acidic Gases on Ice. Chem. Rev. 2006,

106, 1375-1444. (2) Moore, M.; Hudson, R.; Carlson, R. The Radiolysis of SO2 and H2S in Water Ice: Implications for the Icy Jovian Satellites. Icarus 2007, 189, 409-423. (3) Cassidy, T.; Coll, P.; Raulin, F.; Carlson, R.; Johnson, R.; Loeffler, M.; Hand, K.; Baragiola, R. Radiolysis and Photolysis of Icy Satellite Surfaces: Experiments and Theory. Space Sci. Rev. 2010,

153, 299-315. (4) Petrenko, V. F.; Whitworth, R. W. Physics of Ice. Oxford University Press: Oxford ; New York,

1999; p xi, 373 p. (5) Lane, A. L.; Nelson, R. M.; Matson, D. L. Evidence for Sulphur Implantation in Europa's UV Absorption Band. Nature 1981, 292, 38-39. (6) Schriver-Mazzuoli, L.; Schriver, A.; Chaabouni, H. Photo-Oxidation of SO2 and of SO2 Trapped in Amorphous Water Ice Studied by IR Spectroscopy. Implications for Jupiter's Satellite Europa.

Can. J. Phys. 2003, 81, 301-309. (7) Loeffler, M. J.; Hudson, R. L. Thermally‐Induced Chemistry and the Jovian Icy Satellites: A Laboratory Study of the Formation of Sulfur Oxyanions. Geophys. Res. Lett. 2010, 37, L19201. (8) Devlin, J. P.; Buch, V. Evidence for the Surface Origin of Point Defects in Ice: Control of Interior Proton Activity by Adsorbates. J. Chem. Phys. 2007, 127, 091101. (9) Jagoda-Cwiklik, B.; Devlin, J.; Buch, V. Spectroscopic and Computational Evidence for SO2 Ionization on 128 K Ice Surface. Phys. Chem. Chem. Phys. 2008, 10, 4678-4684. (10) Jirsak, T.; Rodriguez, J. A. Chemistry of NO2 and SO2 on Ice Layers and H2O/Zn Interfaces: Photoemission Studies on the Formation of Acid Water and Metal Corrosion. Langmuir 2000, 16, 10287-10293. (11) Kim, Y. K.; Kim, S. K.; Kim, J. H.; Kang, H. Kinetic Isolation of Reaction Intermediates on Ice Surfaces. Precursor States of SO2 Hydrolysis. J. Phys. Chem. C 2009, 113, 16863-16865. (12) Kimmel, G. A.; Petrik, N. G.; Dohnalek, Z.; Kay, B. D. Crystalline Ice Growth on Pt(111): Observation of a Hydrophobic Water Monolayer. Phys. Rev. Lett. 2005, 95, 166102. (13) Hodgson, A.; Haq, S. Water Adsorption and the Wetting of Metal Surfaces. Surf. Sci. Rep.

2009, 64, 381-451. (14) Han, S.; Lee, C.; Hwang, C.; Lee, K.; Yang, M.; Kang, H. Spectrometer for the Study of Angleand Energy-Resolved Reactive Ion Scattering at Surfaces. Bull. Korean Chem. Soc. 2001, 22, 883888. (15) Kang, H. Reactive Ion Scattering of Low Energy Cs+ from Surfaces. A Technique for Surface Molecular Analysis. Bull. Korean Chem. Soc. 2011, 32, 389-398. (16) Kimmel, G. A.; Stevenson, K. P.; Dohnalek, Z.; Smith, R. S.; Kay, B. D. Control of Amorphous Solid Water Morphology Using Molecular Beams. I. Experimental Results. J. Chem. Phys. 2001, 114, 5284. 20

ACS Paragon Plus Environment

Page 20 of 22

Page 21 of 22

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Earth and Space Chemistry

(17) Zimbitas, G.; Haq, S.; Hodgson, A. The Structure and Crystallization of Thin Water Films on Pt(111). J. Chem. Phys. 2005, 123, 174701. (18) Thurmer, K.; Bartelt, N. C. Growth of Multilayer Ice Films and the Formation of Cubic Ice Imaged with STM. Phys. Rev. B 2008, 77, 10. (19) Park, E.; Lee, D. H.; Kim, S.; Kang, H. Transport and Surface Accumulation of Hydroniums and Chlorides in an Ice Film. A High Temperature (140-180 K) Study. J. Phys. Chem. C 2012, 116, 21828-21835. (20) Kang, H. Chemistry of Ice Surfaces. Elementary Reaction Steps on Ice Studied by Reactive Ion Scattering. Acc. Chem. Res. 2005, 38, 893-900. (21) Materer, N.; Starke, U.; Barbieri, A.; VanHove, M. A.; Somorjai, G. A.; Kroes, G. J.; Minot, C. Molecular Surface Structure of Ice(0001): Dynamical Low-Energy Electron Diffraction, Total-Energy Calculations and Molecular Dynamics Simulations. Surf. Sci. 1997, 381, 190-210. (22) Schmidt, M. W.; Baldridge, K. K.; Boatz, J. A.; Elbert, S. T.; Gordon, M. S.; Jensen, J. H.; Koseki, S.; Matsunaga, N.; Nguyen, K. A.; Su, S. General Atomic and Molecular Electronic Structure System.

J. Comput. Chem. 1993, 14, 1347-1363. (23) Fletcher, G.; Schmidt, M.; Gordon, M. Developments in Parallel Electronic Structure Theory.

Adv. Chem. Phys. 1999, 110, 267-294. (24) Speedy, R. J.; Debenedetti, P. G.; Smith, R. S.; Huang, C.; Kay, B. D. The Evaporation Rate, Free Energy, and Entropy of Amorphous water at 150 K. J. Chem. Phys. 1996, 105, 240-244. (25) Park, S. C.; Moon, E. S.; Kang, H. Some Fundamental Properties and Reactions of Ice Surfaces at Low Temperatures. Phys. Chem. Chem. Phys. 2010, 12, 12000-12011. (26) Moon, E.-S.; Kang, H. Metastable Hydronium Ions in UV-Irradiated Ice. J. Chem. Phys. 2012,

137, 204704.

21

ACS Paragon Plus Environment

ACS Earth and Space Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

For TOC Only

22

ACS Paragon Plus Environment

Page 22 of 22