这是用户在 2024-7-31 10:48 为 https://pubs.acs.org/doi/10.1021/jacs.0c06858 保存的双语快照页面,由 沉浸式翻译 提供双语支持。了解如何保存?

Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

You’ve supercharged your research process with ACS and Mendeley!

STEP 1:
Click to create an ACS ID

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

MENDELEY PAIRING EXPIRED
Your Mendeley pairing has expired. Please reconnect
ACS Publications. Most Trusted. Most Cited. Most Read
Photoinduced Oxidation Reactions at the Air–Water Interface
CONTENT TYPES

Figure 1Loading Img
  • Subscribed
Perspective

Photoinduced Oxidation Reactions at the Air–Water Interface
Click to copy article link
点击复制文章链接
Article link copied!
化学TOP简介JCI 2.66EI检索SCI升级版 化学1区SCI Q1IF 14.4

Open PDF

Journal of the American Chemical Society

Cite this: J. Am. Chem. Soc. 2020, 142, 38, 16140–16155
Click to copy citationCitation copied!
https://doi.org/10.1021/jacs.0c06858
Published August 24, 2020
Copyright © 2020 American Chemical Society

Abstract

Click to copy section link
点击复制版块链接
Section link copied!

Chemistry on water is a fascinating area of research. The surface of water and the interfaces between water and air or hydrophobic media represent asymmetric environments with unique properties that lead to unexpected solvation effects on chemical and photochemical processes. Indeed, the features of interfacial reactions differ, often drastically, from those of bulk-phase reactions. In this Perspective, we focus on photoinduced oxidation reactions, which have attracted enormous interest in recent years because of their implications in many areas of chemistry, including atmospheric and environmental chemistry, biology, electrochemistry, and solar energy conversion. We have chosen a few representative examples of photoinduced oxidation reactions to focus on in this Perspective. Although most of these examples are taken from the field of atmospheric chemistry, they were selected because of their broad relevance to other areas. First, we outline a series of processes whose photochemistry generates hydroxyl radicals. These OH precursors include reactive oxygen species, reactive nitrogen species, and sulfur dioxide. Second, we discuss processes involving the photooxidation of organic species, either directly or via photosensitization. The photochemistry of pyruvic acid and fatty acid, two examples that demonstrate the complexity and versatility of this kind of chemistry, is described. Finally, we discuss the physicochemical factors that can be invoked to explain the kinetics and thermodynamics of photoinduced oxidation reactions at aqueous interfaces and analyze a number of challenges that need to be addressed in future studies.

This publication is licensed under the terms of your institutional subscription. Request reuse permissions.

Copyright © 2020 American Chemical Society

1. Introduction

Click to copy section link
点击复制版块链接
Section link copied!

Chemical processes are sensitive to the molecular environments where they take place, including the solvent, the protein active site, or the surface of a material. The environment influences the relative stability of the species along the reaction path. In the case of liquids, the solvation power and the nature of the reaction system determine how different species are stabilized, and the differential stability may retard or accelerate the reaction process. Moreover, environmental factors can also modify the regioselectivity or stereoselectivity, the endo-/exothermicity, or even the mechanism. (1)
When reactions occur at the surface of liquid water or at the interface of liquid water with a hydrophobic medium, the asymmetry of the interface, the presence of free water OH groups, and the distinctive properties of interfacial water often lead to unexpected solvation effects on chemical reactions, and the resulting properties are not necessarily intermediate between the two bulk phases. (2) For instance, a variety of thermal reactions occurring at aqueous interfaces have been found to be significantly accelerated compared to the bulk reactions; this effect is referred to as “on-water catalysis”. (3) The reaction rate acceleration can stem from an increase in reactant concentration (molecular crowding) or the stabilization of a transition state (or a combination of both); however, a full understanding of this mechanism remains lacking. The enhancement of redox reactions has been broadly studied in this context, with prominent examples including H-abstraction by OH, (4) ozonolysis (5−11) and ozonations, (12) Fenton and Fenton-like reactions, (13) Dakin and Baeyer–Villiger oxidations, (14) halide anion oxidation by OH and O3, (15) and the spontaneous formation of hydrogen peroxide (14) and spontaneous reduction of organic molecules (16) at the neat water surface.
This Perspective deals with the special case of photoinduced oxidation reactions at aqueous interfaces. Photoredox chemistry covers a broad range of processes with great atmospheric and environmental relevance. (17,18) These processes are also key in the fast-developing field of photocatalysis, (19) which in a nonrestrictive sense connects multiple disciplines such as synthetic organic chemistry to biochemistry, electrochemistry, water purification technologies, and solar energy conversion. A review of the literature related to all these fields obviously goes beyond the scope of the present report; however, the roles of liquid water interfaces in reaction kinetics, dynamics, and thermodynamics are highly relevant in all the above fields.
In the atmosphere, gas-phase photochemistry is fundamentally associated with photooxidation by OH radicals and other reactive oxygen species (ROS). (20,21) The main source of tropospheric OH is the photolysis of ozone via the absorption of ultraviolet (UV) light in the presence of water vapor, although several other routes can lead, directly or indirectly, to the formation of OH radicals. (22) ROS react with other compounds to initiate and propagate most tropospheric chemical reactions. A simplified scheme illustrating the complexity of the tropospheric chemistry of ROS in the gas phase is shown in Figure 1. It shows the set of coupled reactions that constitute the sources and sinks of reactive oxygen species and the crucial role played by ozone and hydroxyl radicals. On the other hand, the aqueous-phase chemistry of ROS is of cross-cutting importance in biology, environmental science, and water treatment, for instance. (23)

Figure 1

Figure 1. Gas-phase atmospheric reactions involving reactive oxygen species (ROS). Reproduced with permission from ref (23). Copyright 2015 American Chemical Society.
图 1.涉及活性氧 (ROS) 的气相大气反应。经参考文献(23)许可转载。版权所有 2015 美国化学学会。

Interest in the direct photochemistry of organic compounds in the atmosphere has grown significantly in recent years in parallel with the development of photocatalysis. Most compounds directly emitted to the atmosphere (e.g., saturated hydrocarbons, monoterpenes, and alcohols) do not absorb sunlight available in the troposphere (i.e., beyond the cutoff of 290 nm) and therefore do not undergo significant direct photochemical reactions. However, their oxidation products can be photochemically active. For instance, methyl hydroperoxide (CH3OOH) is an important trace gas in Earth’s atmosphere that is mainly produced by the oxidation of methane, the ozonolysis of alkenes, and biomass burning. CH3OOH has an electronic transition band that is centered at 200 nm but extends beyond the tropospheric cutoff, leading to photodissociation into OH and CH3O radicals; the latter can react with O2 to form HO2 and formaldehyde. (24,25) Other species produced via the oxidation of primary organic compounds [e.g., pyruvic acid (PYA) and certain imidazole derivatives] and compounds emitted to the atmosphere through combustion processes (e.g., polycyclic aromatic hydrocarbons and aromatic ketones) are also photoactive. The excited states of these species can initiate chemical processes through interactions with other organic compounds that are not themselves photoactive. In these photoinduced processes, a chromophore called a photosensitizer absorbs light and transfers the excess energy to a nearby molecule that is not naturally photoactive, eventually initiating the oxidation chain. Commonly, the photosensitizer absorbs light to a singlet excited state, which then evolves to a triplet state via intersystem crossing (ISC) with a large quantum yield. The higher lifetime of the triplet compared to the singlet facilitates encounters and reaction with other molecules. Overall, two main types of photosensitization processes occur in the atmosphere (18) depending on how the excited photosensitizer is quenched. The quencher can be an organic molecule (Type I) or triplet oxygen (Type II), as shown in Figure 2. In the first case, an electron-transfer or H-abstraction reaction results in the creation of a new organic radical that propagates the oxidation process. In the second case, the photosensitizer reacts with In the first case, an electron-transfer or H-abstraction reaction results in the creation of a new organic radical that propagates the oxidation process. In the second case, the photosensitizer reacts with 3O2, either by transferring the excess energy to form the reactive excited singlet , either by transferring the excess energy to form the reactive excited singlet 1O2 or by electron transfer to form the superoxide anion Oor by electron transfer to form the superoxide anion O2. In principle and according to the IUPAC definition, a photosensitization process is one in which the photosensitizer is not consumed in the reaction. From this point of view, photosensitization and photocatalysis can be considered synonyms. However, there are processes in which the photosensitizer molecule is also a reactant; in such a case, the molecule is referred to as a photoinitiator.. In principle and according to the IUPAC definition, a photosensitization process is one in which the photosensitizer is not consumed in the reaction. From this point of view, photosensitization and photocatalysis can be considered synonyms. However, there are processes in which the photosensitizer molecule is also a reactant; in such a case, the molecule is referred to as a photoinitiator.

Figure 2

Figure 2. (a) Schematic representation of a Type I sensitization process involving the degradation of phenols via a reaction with an excited carbonyl compound as photosensitizer. (b) Comparison of Type I and Type II photosensitization processes. Reproduced with permission from ref (18). Copyright 2012 American Chemical Society.
图2.(a) I型敏化过程的示意图,该过程涉及通过与作为光敏剂的激发羰基化合物反应来降解酚类物质。(b) I型和II型光敏过程的比较。经参考文献(18)许可转载。版权所有 2012 美国化学学会。

While all the above processes can occur in the gas phase, the importance of multiphase chemistry involving surfaces and interfaces of different natures (e.g., mineral dust, liquid water and ice surfaces, aerosols, and urban surfaces) has also been established. (17,18,26,27) This Perspective first presents some studies that found enhancements in these photoinduced processes at the air–water interface, which is illustrated by some selected examples. We then discuss the different factors that can lead to modifications of the reaction properties. Finally, we present some perspectives in the field.

2. Photochemistry of OH Precursors

Click to copy section link
点击复制版块链接
Section link copied!

In this section, we present a few select examples to describe the photochemistry of simple precursors that generate OH radicals (and possibly other ROS), which then trigger the oxidation of more complex compounds present in the environment. These reaction pathways are key in atmospheric and environmental chemistry, and the examples given below refer essentially to these fields. The production of OH originates from the direct photolysis of the precursor (hydroperoxides, HNO3) or from hydrogen abstraction from water molecules, either by an appropriate precursor excited state (SO2*, NO2*) or one of its photolytic products (O3).

2.1. ROS

Ozone, hydrogen peroxide, and methyl hydroperoxide are among the main ROS. (23) They are strong oxidants with relevance in atmospheric chemistry, biological processes, and wastewater treatment in peroxone chemistry. (17,21,23,24,28−37) In the stratosphere, ozone absorbs solar radiation, protecting living organisms from harmful UV radiation; in the troposphere, ozone acts as a greenhouse gas and a pollutant, damaging the respiratory tissues of animals. (38) Hydrogen peroxide and methyl hydroperoxide are prototypical hydroperoxides that are mainly formed via the self-reaction of HO2 radicals and the oxidation of methane, respectively. (24,25,39) These hydroperoxides are reservoirs of hydroxyl radicals and are found in rainwater and aerosols. (40−42)
One of the main roles of tropospheric ozone is to serve as an important source of hydroxyl radicals. Ozone photolysis produces O(1D), which reacts with water vapor to form two OH radicals (reactions 1 and 2). (43) In the gas phase, the measured photolysis rate constant for these reactions is 3.8 × 10–5 s–1, (44) while calculations produced a value of 3.2 × 10–5 s–1. (45) It should be noted that both of these values assume sunlight actinic flux, as will be considered from now on in this work.
(1)
(2)
(3)
Anglada and co-workers (27,45) carried out quantum mechanics/molecular mechanics molecular dynamics (QM/MM-MD) simulations to investigate the spectrum of ozone at the air–water interface and found a red shift and shape broadening compared to the gas phase. They also predicted the photolysis rate constant to be 1 order of magnitude greater at the interface than in the gas phase ((1.1–7.6) × 10–4 s–1) when the contribution of reaction 3 was taken into account. (45) Moreover, simulations (45−47) have shown that ozone pile up at the air–water interface. For instance, for a typical gas-phase concentration of 7.5 × 1011 molecules·cm–3, an average concentration of 9.6 × 1012 molecules·cm–3 in the interfacial layer is estimated. This corresponds to a predicted increase of approximately 4 orders of magnitude in the rate of OH radical formation at the interface, which has important atmospheric implications. (45)
In contrast to ozone, the photolytic process of hydroperoxides (reactions 4 and 5) produces OH radicals directly. In the case of alkyl derivatives like CH3OOH, OH radicals are also produced by the reaction of alkoxy radicals with water (reaction 6).
(4)
(5)
(6)
The absorption cross-sections and quantum yields of H2O2 and CH3OOH have been reported in the gas phase, (48−52) in aqueous solution, and in snow ices. (53−58) The reported photolysis rate constants for H2O2 are 4.8 × 10–6 s–1 in the gas phase and 1.5 × 10–5 s–1 in aqueous solution. (59) For CH3OOH, the rate constants are 1.3 × 10–6 and 1.1 × 10–6 s–1 in the gas phase (51) and 4.5 ± 1.0 × 10–5 s–1 in aqueous solution. (58)
Hydrogen peroxide is much more soluble in water than methyl hydroperoxide (the Henry’s law coefficients are 1.0 × 105 and 3.0 × 102 mol·dm–3·atm–1, respectively (59)). However, according to molecular dynamics (MD) and quantum mechanics/molecular mechanics (QM/MM) calculations, (46,60−62) both species accumulate at the air–water interface, leading to an increase in concentration that can enhance the importance of interfacial photochemistry. (46,62) However, few works have focused on the photochemistry of these species at the air–water interface, (27,62,63) and the results are contrasting. On one hand, Nissenson and co-workers (63) studied the photochemistry of H2O2 in water with a focus on cage effects on the photolysis quantum yields upon the addition of 2-propanol as an OH scavenger. The authors estimated the photolysis rate constant at the surface to be 9.3 × 10–8 s–1, much smaller than the experimental values reported by other authors for the gas phase and in the bulk, (59) although no discussion on this issue was provided. On the other hand, Martins-Costa and co-workers (62) conducted QM/MM MD calculations to study the photolysis of CH3OOH at the air–water interface. The calculated UV–visible spectra reveal a significant blue shift of the main absorption band corresponding to the interface (0.42 eV). This blue shift is attributed to the hydrogen-bond interactions between methyl hydroperoxide and the water molecules. While one would expect this blue shift to significantly decrease the photolytic rate constant, the calculations also show a broadening of the absorption band corresponding to the air–water interface that results in comparable cross-sections in the red part of the band compared to the gas phase (Figure 3Figure 3b), the photolysis rate constant at the interface (JCH3OOH = 0.76 × 10–6 s–1) is only slightly smaller than that in the gas phase (JCH3OOH = 1.05 × 10–6 s–1) All in all, the rate of OH production rate has been shown to be much higher at the interface than in the gas phase, primarily due to the higher concentration of methyl hydroperoxide in this environment and, to a minor extent, the contribution of reaction 6.
过氧化氢比甲基氢过氧化物更易溶于水(亨利定律系数分别为 1.0 × 105 和 3.0 × 102 mol·dm–3·atm–1 (59))。然而,根据分子动力学(MD)和量子力学/分子力学(QM/MM)计算,(46,60−62)两种物质都在空气-水界面处积累,导致浓度增加,从而增强了重要性界面光化学。 (46,62) 然而,很少有工作关注这些物种在空气-水界面的光化学,(27,62,63) 并且结果是相反的。一方面,Nissenson 和同事 (63) 研究了水中 H2O2 的光化学,重点关注添加 2-丙醇作为 OH 清除剂后笼效应对光解量子产率的影响。作者估计表面的光解速率常数为 9.3 × 10–8 s–1,远小于其他作者报告的气相和本体的实验值,(59),尽管没有对此问题进行讨论假如。另一方面,Martins-Costa 和同事 (62) 进行了 QM/MM MD 计算,以研究 CH3OOH 在空气-水界面的光解作用。计算的紫外-可见光谱显示与界面相对应的主吸收带发生显着的蓝移(0.42 eV)。这种蓝移归因于甲基氢过氧化物和水分子之间的氢键相互作用。虽然人们预计这种蓝移会显着降低光解速率常数,但计算还显示,与空气-水界面相对应的吸收带变宽,导致与气体相比,吸收带红色部分的横截面相当阶段(图3a)。结果(见图3b),界面处的光解速率常数(CH3OOH = 0.76 × 10–6 s–1)仅略小于气相中的光解速率常数(CH3OOH = 1.05 × 10–6 s–1)总而言之,界面处的 OH 生成速率比气相中的要高得多,这主要是由于该环境中甲基氢过氧化物的浓度较高,并且在较小程度上是由于反应 6 的贡献。关闭

Figure 3

Figure 3. (a) Calculated absorption cross-sections of CH3OOH in the gas phase and at the air–water interface along with different experimental values reported for the gas phase. (b) Calculated partial photolysis rates for CH3OOH. Integrated total rates are indicated. Reproduced from ref (62) with permission from the PCCP Owner Societies.

2.2. Reactive Nitrogen Species (RNS)

RNS are key compounds in atmospheric and biological chemistry. Like ozone, nitrogen dioxide is a target compound for measuring atmospheric pollution worldwide. Nitrogen dioxide may cause serious health problems by reducing immunity to lung infections. Reactions 717 and Figure 1 summarize the most relevant redox processes involving oxygenated nitrogen species; these processes are strongly interwoven. (22,25,43,64)
(7)
(8)
(9)
(10)
(11)
(12)
(13)
(14)
(15)
(16)
(17)
Reaction 7 generates atomic oxygen, which reacts with molecular oxygen (reaction 8) producing atmospheric ozone in gas phase, while reaction 9 regenerates NO2. Reactions 10 and 11 are precursors of OH radicals in the gas phase, which is formed via reaction 12. (65) The hydrolysis of NO2 (reaction 13) has been suggested to occur via heterogeneous processes (64) that produce HONO and HNO3, and the photolysis of nitrous acid is a major source of hydroxyl radicals in the atmosphere (reaction 14). (22) The reaction of nitrogen dioxide with peroxy radicals (reaction 15) produces RO2NO2, an atmospheric reservoir of NO2 that can be transported and released by photolysis through reaction 16. (66) Finally, nitrogen dioxide can react with hydroxyl radicals to produce nitric acid (reaction 17). (25)
The photochemistry of nitrogen dioxide adsorbed on different interfaces has been studied. QM/MM MD simulations at the air–water interface predict an excess concentration of approximately ten times relative to the gas phase (67,68) along with a significant enhancement in the absorption cross-section, (68) as plotted in Figure 4a). The calculated photolysis rate constants for reaction 7 are J7 = 10.8 × 10–3 s–1 in the gas phase (the experimental value (69) is 8.8 × 10–3 s–1) and J7 = 15.3 × 10–3 s–1 at the air–water interface. For reaction 10, the calculated photolysis rate constants are similar for both environments (J10 = 2.2 × 10–2 s–1 in the gas phase and 2.5 × 10–2 s–1 at the air–water interface), and the estimated OH production rate at the air–water interface as a function of the NO2 concentration (Figure 4b) is approximately 2 orders of magnitude larger than in the gas phase (for details, see ref (68)).
研究了吸附在不同界面上的二氧化氮的光化学。空气-水界面处的 QM/MM MD 模拟预测相对于气相 (67,68) 的浓度过量约十倍,同时吸收截面显着增强 (68),如图 4a) 所示。计算得出的反应 7 的气相光解速率常数为 7 = 10.8 × 10–3 s–1(实验值(69)为 8.8 × 10–3 s–1)和 7 = 15.3 × 10–3 s–1 1 空气-水界面。对于反应 10,两种环境下计算的光解速率常数相似(气相中 10 = 2.2 × 10–2 s–1,空气-水界面处 2.5 × 10–2 s–1),并且估计的 OH空气-水界面处的生成速率与 NO2 浓度的关系(图 4b)比气相中的生成速率大约大 2 个数量级(详细信息请参见参考文献(68))。关闭

Figure 4

Figure 4. (a) Experimental UV spectrum of NO2 in the gas phase and calculated spectra in the gas phase and at the air–water interface. (b) Estimated upper and lower limits of the OH production rate (molecule·cm–3·s–1) for different gas-phase concentrations of NO2 (molecule·cm–3). Calculations in the gas phase (light gray) and at the air–water interface (dark gray) using k12 from ref (70). The plain and dashed red lines respectively correspond to the gas-phase and interfacial values using k12 from ref (71). The gas-phase values assume a relative humidity of 20%. The horizontal plain line represents a typical OH production rate from ozone photolysis. Reproduced with permission from ref (68).
图4.(a) NO2在气相中的实验紫外光谱以及在气相和空气-水界面处计算的光谱。(b) 不同气相浓度NO2(分子·cm–3)下OH产速率(molecule·cm–3·s–1)的估计上限和下限。在气相(浅灰色)和空气-水界面(深灰色)使用参考文献(70)中的k12进行计算。红线和红色虚线分别对应于参考文献(71)中使用k12的气相值和界面值。气相值假设相对湿度为 20%。水平素线代表臭氧光解产生的典型 OH 率。经参考文献(68)许可转载。

Further interesting results are available regarding the photochemistry of NO2 adsorbed on other surfaces. Adsorbed NO2 releases significant amounts of HONO. Thus, the irradiation of NO2 in the range of 300–700 nm on organic surfaces containing phenols, aromatic ketones, humic acid, aromatic and polycyclic aromatic hydrocarbons, or even urban grime (72−77) results in HONO formation in both the UV-A spectral region and in the visible region under atmospheric conditions. (73) The observed flux densities of HONO in these processes range between 3.2 × 109 and 5 × 1010 molecules·cm–2·s–1. (73,77) In these processes, radiation is not directly absorbed by NO2. Instead, the radiation is absorbed by a photosensitizer (e.g., 4-benzoylbenzoic acid), which then reacts with an organic compound on the surface (e.g., phenols) to form a reductant (e.g., phenolate) capable of converting NO2 to NO2. Finally, the protonation of NO2 produces HONO. (72,73)
The photochemistry of nitric acid at the air–water interface is also an interesting process that has raised many questions. In the troposphere, HNO3 has long been considered as a sink of NOx species; however, recent studies have stressed its role in chemical (78) and photochemical (79−86) renoxification processes that generate HONO, NO, and NO2. Thus, nitric acid may be a reservoir of nitrogen oxides as well as a sink. Here, we focus on its photolysis, which directly produces hydroxyl radicals in the gas phase (reaction 18). When nitric acid is dissolved in water, it mainly exists in the form of nitrate, and its photolysis produces O radicals, which react with water to form OH (reactions 1921).
(18)
(19)
(20)
(21)
The rate constants for HNO3 photolysis in the gas phase (reaction 18) and nitrate photolysis in bulk water (reactions 19 and 20) are very small (J18 = 8.2 × 10–7 s–1, (69)J19 = 8.5 × 10–7 s–1, J20 = (4.5–8.5) × 10–8 s–1). (53,54,69,83,87,88) Note that the NO2 formed in reaction 20 can either be protonated, releasing HONO, or photolyzed (J21 = 1 × 10–4 s–1). (63,87,89,90) However, photolysis at different interfaces is far more important. On one hand, field observations have shown that the photolysis of HNO3 adsorbed on the forest canopy surface is an important source of HONO. (82) On the other hand, experiments in thin water films containing nitrate, chlorine, and bromine salts have shown that the photolysis of NO3 occurs much faster at interfaces than in the bulk, (63,81,90−94) with J values ranging between 1.2 × 10–3 and 3 × 10–6 s–1, depending on the features of the interface. (63,83,95,96) A potential explanation for this enhancement is associated with the solvation of nitrate anion. In bulk water, nitrate is surrounded by a solvent cage that facilitates both the recombination of the NO2 and OH radicals produced by reaction 19 and their deactivation by collision with solvent molecules, thus decreasing the photolysis quantum yield. (63,87,89,90) MD simulations have shown that when halide anions (especially bromide) are present, nitrate ions are dragged closer to the interface so that the water solvent cage surrounding them is reduced, (90,97) arguably making the escape of NO2 to the gas phase easier. (63,81,90−94) Nevertheless, despite the large amount of work done on nitrate photochemistry, open issues remain that are not yet completely understood. For instance, it was recently shown that an important part of nitric acid remains undissociated at the air–water interface, (81,98−106) and the relevance of the corresponding photochemistry is still undetermined.

2.3. Sulfur Dioxide (SO2)

The ground-state chemistry of SO2 has been broadly studied. (107−110) Generally, SO2 is oxidized in the atmosphere to higher sulfur oxides or sulfuric acid. SO2 is emitted by the burning of fossil fuels, industrial processes, ships and other vehicles, and volcanic eruptions. It is eliminated by dry and wet deposition, which contributes to particulate matter and aerosol formation along with acid rain. Exposure to SO2 or particulate matter harms the human respiratory system and makes breathing difficult. (69,111−113) The atmospheric concentrations of SO2 are extremely variable, with mixing ratios ranging from 20 ppt to over 1 ppb in continental background air, 20–50 ppt in the unpolluted marine boundary layer, several hundreds of ppb in urban areas, and much higher in volcanic gases. (111)
Recent studies have shown that SO2 can act as an oxidizer via the photochemistry of its triplet excited state. Upon the UV irradiation of a mixture of SO2 and water vapor without any other oxidants, Donaldson and co-workers (114) observed the formation of new particles, which Kroll and co-workers (115) ascribed to the formation of HOSO and OH radicals according to reactions 2226. UV light excites SO2 into its first singlet electronic state that evolves to a triplet electronic state via ISC. This triplet electronic state is quite reactive and abstracts one hydrogen atom from water, producing HOSO and OH radicals. In this mechanism, SO2 behaves as a photoinitiator.
(22)
(23)
(24)
(25)
(26)
The reaction between the excited triplet electronic states of SO2 and H2O (reaction 24) may follow two different reaction paths, (116) for which the potential energy surfaces are schematized in Figure 5a. These pathways correspond to either a conventional hydrogen atom transfer (HAT) or a proton-coupled electron transfer (PCET), which requires lower activation energy.

Figure 5

Figure 5. (a) Schematic of the free energy surface at 298 K for the reaction of SO2(a 3B1) with one water molecule. (b) Schematic free energy surface at 298 K for the reaction of SO2(a 3B1) with a cluster of four water molecules. Reproduced from ref (116) with permission from the PCCP Owner Societies.
图5.(a) SO2(a 3B1)与一个水分子反应的298 K自由能表面示意图。(b) SO2(a 3B1)与四个水分子簇反应时在298 K下的自由能表面示意图。转载自参考文献(116),经PCCP业主协会许可。

The most relevant electronic features describing these processes are collected in Figure 6. The PCET mechanism is described by a two-center three-electron structure, where the single-occupied 8a1 orbital of SO2 interacts with the lone pair [lp(O)] of water in a way that the electron density in the doubly occupied orbital 20a is shared between the oxygen atom of the water moiety and the terminal oxygen atom of SO2. In contrast, in the single-occupied orbital 22a, the antibonding combination is formed. This situation depicts an electron transfer from the oxygen atom of water to one oxygen atom of SO2; at the same time, one proton jumps from water to the other oxygen atom of SO2.

Figure 6

Figure 6. Orbital diagram for the PCET and HAT mechanisms for the reaction SO2(a 3B1) + H2O. Reproduced from ref (116) with permission from the PCCP Owner Societies.
图6.反应 SO2(a 3B1) + H2O. 的 PCET 和 HAT 机理的轨道图,经 PCCP 所有者协会许可,转载自参考文献 (116)。

In the HAT mechanism, the single-occupied 8a1 orbital of SO2 interacts with one of the σ(O–H) bonds of water to form the doubly occupied orbital 20a with bonding character and the single-occupied orbital 22a with antibonding character. This situation also corresponds to a three-center three-electron system; however, in this case, the electronic density lies over the O–H–O moiety, indicating the simultaneous breaking and forming of the covalent bonds (OS)O–H–O(H). Indeed, elementary reactions occurring through a PCET mechanism in the oxidation of organic and inorganic species by radicals have been found to be more favorable than oxidation processes occurring through the HAT mechanism. (117−127)
The reaction of SO2 (3B1) with up to four water molecules (116) (Figure 5b) sheds light on the microsolvation effect on this process and can be considered as a model for reactivity at the air–water interface, where, as for the other species discussed above, SO2 accumulates. (128−133) Indeed, the free energy barrier of the PCET mechanism is quite small (0.70 kcal·mol–1), suggesting that the reaction could be spontaneous at the interface, making this photochemical mechanism a significant source of OH radicals, as confirmed by further QM/MM simulations. (133) For instance, for a relative humidity of 100% and a gas-phase SO2 concentration of 1012 molecules·cm–3, which is typical of urban areas, the rate of formation of OH radicals is estimated to be approximately 105 molecules·cm–3·s–1 in the gas phase. In comparison, the rate reaches 7 × 108 molecules·cm–3·s–1 when the process occurs at the air–water interface for the same gas-phase concentration. (133)
Beyond the formation of OH radicals, another outcome of the interfacial photochemistry of SO2 is the acidification of the medium because HOSO is quite acidic (pKa = −1) and undergoes fast ionic dissociation at the interface, as shown by QM/MM MD simulations (see Figure 7). (134) Moreover, the formed SO2 ions can be further oxidized by H2O2, O3, OH, or HO2 to produce sulfuric acid. (134)

Figure 7

Figure 7. Different steps of the photo-oxidation of SO2 leading to sulfate. QM/MM MD simulations at the air–water interface reveal that as HOSO radical is formed, it rapidly ionizes. This is shown in the bottom part of the figure by the time evolution of OH distances in the H2O···HOSO system (bottom, right). The structure of the transition structure for proton transfer is also displayed (bottom, left). Adapted with permission from ref (134).
图7.SO2光氧化导致硫酸盐的不同步骤。在空气-水界面的QM/MM MD模拟表明,随着HOSO自由基的形成,它迅速电离。这在图的底部通过H2O···HOSO系统(右下)。还显示了质子转移的过渡结构的结构(底部,左侧)。经参考文献(134)许可改编。

3. Direct Photooxidation of Organic Species

Click to copy section link
点击复制版块链接
Section link copied!

The photooxidation of organic species at the aqueous interface represents a vast field of research and turns out to be much more complex than the chemistry of OH precursors. To some extent, the role of organic radicals formed in the early stages of the photochemical process can be compared to the role played by the oxidizing OH radicals in overall atmospheric chemistry. As noted in the introduction, this photochemistry generally involves the presence of photoinitiators or photosensitizers. Such reactions are also enormously important in atmospheric and environmental chemistry because organic matter accumulates at the surfaces of cloud water droplets and aerosols as well as on surface waters, lakes, rivers, and, most importantly, oceanic surface microlayers, which are enriched with organic matter such as humic acid, fatty acids, amino acids, proteins, lipids, and other biogenic materials. (135−139) To illustrate these processes, we have chosen two examples that have received considerable attention in recent publications.

3.1. Pyruvic Acid (PYA), a Versatile Photoinitiator

An intriguing case study is PYA, which has major biological significance and is also relevant in atmospheric chemistry as a product of the atmospheric oxidation of isoprene. (140) PYA is detected in clouds and aerosols (141) and is important in prebiotic chemistry. (139) The photochemistry of PYA has been investigated in different environments as a model of α-dicarbonyl compounds. A unique feature of PYA is that its photochemistry is totally different in the gas phase than in aqueous environments.
In the gas phase, the main photolytic process starts with absorption to the first excited singlet electronic state (n→π* transition, λmax = 350 nm), which induces PYA cleavage into carbon dioxide and methyl hydroxycarbene (reaction 27) via hydrogen transfer from the acidic group to the keto oxygen atom. (142,143) Excitation at higher energies (up to λ = 300 nm) leads to the formation of acetyl and carboxyl radicals (reaction 28). (144)
(27)
(28)
In the aqueous phase, the photolysis is far more intricate than in the gas phase, and the mechanisms are totally different. A first challenge is the coexistence of different forms in aqueous solution, namely PYA and its conjugated base pyruvate (pKa = 2.45), the corresponding hydrated gem-diols (PYT), and oligomeric species such as parapyruvic acid (PPA) and zymonic acid (ZYA; see Figure 8). PYA may also undergo keto–enol equilibria. (145)

Figure 8

Figure 8. Chemical structures of pyruvic acid, pyruvate, the corresponding gem-diols, parapyruvic acid, and zymonic acid.
图8.丙酮酸、丙酮酸、相应的 gem-diols、副丙酮酸和酶酸的化学结构。

Several works have investigated PYA photochemistry in water under highly acidic conditions (pH = 1–2.5). (139,146−156) Absorption to the single excited state [1S(n,π*)] is blue-shifted (λmax = 320 nm) with respect to gas phase. (139) The excited singlet then undergoes ISC with the lowest-lying triplet excited state [3T(n,π*)], which becomes the reactive species (see Figure 2). This triplet state can react with another PYA or PYT molecule in its ground electronic state through a mechanism that involves either the abstraction of the acidic hydrogen atom via proton-coupled electron transfer (146) or a concerted hydrogen atom abstraction. (147) This initiates a chain of reactions that cause decarboxylation, the formation of oligomers, dimethyltartaric acid, acetoin, acetic and lactic acids, and oxo-C8 and C7 products. (26,146,147,157) However, controversy exists regarding the entire mechanism and the formation of several products. (147−152) As the pH increases, the gem-diol form diminishes, and the pyruvate concentration grows, resulting in an increase in absorption intensity and a moderate blue shift (λmax = 316 nm). (158) Another mechanism then becomes competitive: the triplet excited state of PYA abstracts one of the hydrogen atoms of the methyl group (instead of the acidic group) from the ground electronic state of another PYA molecule. The main oligomeric species formed in this mechanism are PPA and ZYA. (153) Some of the newly formed compounds (PPA in particular) can act as secondary photosensitizers that contribute to or initiate the formation of oligomers (26,153,158) and participate in cross-reactions with glyoxylic acids, fatty acids, and fatty alcohols. (153,154,159) Interestingly, PYA has been shown to behave as a photosensitizer, affecting the atmospheric multiphase reactions of ozone with oxalic acid. (160)
The enormous versatility of this bulk photochemistry is also expected at the air–water interface. However, the interfacial mechanisms remain poorly understood. Using vibrational sum-frequency spectroscopy, Gordon and co-workers (161) identified the oligomers PPA and ZYA at the air–water interfaces of PYA solutions, demonstrating that these compounds are surface active. Additionally, they performed MD simulations in the bulk and at the air–water interface; the corresponding density profiles are collected in Figure 9. Interestingly, the results indicate the depletion of PYT and the accumulation of PYA, PPA, and ZYA at the air–water interface, which suggests that the enhanced photoreactivity of the interface might boost oligomerization reactions. Moreover, Eugen and co-workers (162) demonstrated a reduction of 1.8 units in the effective pKa of PYA at the air–water interface, which would increase the undissociated fraction of PYA and further enhance the photoreactivity at the air–water interface.

Figure 9

Figure 9. Density profiles from MD simulations of a mixed aqueous solution containing PYA (red), PYT (blue), PPA (green), and ZYA (yellow). The concentration of each species was ∼0.25 M. The vertical dotted gray line indicates the approximate interface boundary. Reproduced with permission from ref (161).

Very recently, Fu and co-workers (163) employed time-of-flight secondary ion mass spectrometry (SIMS) to investigate the products of PYA photolysis at the air–water interface and in bulk water. The SIMS analyses were carried out with dry samples, and inevitable drying effects may have influenced the results of the study. However, the authors established a reaction flowchart, which is displayed in Figure 10. The radical, decarboxylation, esterification, and anhydride formation reactions have been identified. In addition, at the interface, oxygen from the air forms other oxidation products. As a rule, it is observed that radical reactions and the formation of high-molecular-weight compounds are enhanced at the air–water interface.

Figure 10

Figure 10. Reaction flowchart showing reactions in bulk water and at the air–water interface. Numbers on arrows indicate reaction types: (1) radical reactions, (2) decarboxylation reactions, (3) anhydride formation, and (4) esterification. Red squares indicate new products identified by the authors, and chemical formulas correspond to undetermined or to too-large structures. Reproduced with permission from ref (163).
图 10.反应流程图显示了在散装水中和空气-水界面上的反应。箭头上的数字表示反应类型:(1)自由基反应,(2)脱羧反应,(3)酸酐形成和(4)酯化反应。红色方块表示作者确定的新产品,化学式对应于未确定或太大的结构。经参考文献(163)许可转载。

3.2. Fatty Acids

Another illustrative example is the photoinduced reactivity of fatty acids and fatty alcohols initiated by photosensitizers in chromophoric dissolved organic matter (CDOM). (164−166) Imidazole 2-carboxaldehyde, 4-benzoylbenzoic acid, and humic acids have been widely studied as suitable proxies for CDOM photosensitizers. This represents an important environmental photooxidation chemistry, especially at the sea surface microlayer and interfacial bubbles at the sea surface, which are enriched by many amphiphilic organic compounds with biogenic origins. Fatty acids have been identified in these organic layers, and photosensitizers such as those cited above show a propensity for partitioning to them. (165)
The starting mechanism involves the formation of a triplet electronic state via ISC from the excited singlet state reached by the absorption of sunlight by the photosensitizer. The triplet state then reacts with the organic species at the air–water interface, usually via the abstraction of one hydrogen atom or through electron transfer. In this way, two radicals are formed and initiate a chain of complex chemical reactions that has been thoroughly described for nonanoic acid as a fatty acid proxy. (165−167) The proposed mechanism with humic acid as the photosensitizer is displayed in Figure 11. (166)

Figure 11

Figure 11. Proposed mechanism for the photochemical degradation of nonanoic acid at the air–water interface in the presence of humic acid as photosensitizer. Reproduced with permission from ref (166). Copyright 2016 American Chemical Society.
图 11.提出了在腐殖酸作为光敏剂存在下壬酸在空气-水界面上光化学降解的机理.经参考文献(166)许可转载。版权所有 2016 美国化学学会。

Figure 11 shows that the triplet excited state of humic acid (3HA) abstracts one hydrogen atom of nonanoic acid via the photosensitizer, unleashing a chain of processes involving dimerization, disproportionation, cross-reactions, and decarboxylation. Moreover, the availability of O2 from the gas phase and water molecules at the air–water interface promotes reactions such as the addition of O2 to form peroxide radicals, allowing the formation of other oxidized products such as hydroxyacids.
Despite organic acids does not absorb light in the actinic range, recent experiments (167) have indicated the presence of radical reactions when a monolayer of nonanoic acid at the water surface is irradiated in the region of 280–330 nm in the absence of any additional photosensitizer. Although a low rate was observed, this study suggests that nonanoic acid itself acts as a photosensitizer at the air–water interface. (167) According to this study, the absorption of radiation by nonanoic acid leads to a triplet state that can dissociate into radicals or react with another fatty acid molecule to from a diol radical. However, other studies on this reaction with different experimental setups could not confirm this mechanism and concluded that if this mechanism exists, it is of minor relevance compared to processes triggered by photoinitiators. (138,159,164−166)

4. Discussion

Click to copy section link
点击复制版块链接
Section link copied!

In the preceding section, we briefly described a few representative examples of photooxidation reactions at the air–water interface. These and other studies (17,18,27,168) on related systems reveal that the photochemical reactivity at the interface may differ considerably from those in both the gas phase and the bulk. Despite this large body of work, a full understanding of the solvation effects on interfacial photochemistry is not yet available. Hereafter, we briefly discuss some of the thermodynamic and kinetic factors (see Scheme 1) that can be advanced to explain the observed effects but that still require verification through further experiments and calculations.

Scheme 1

Scheme 1. Factors Affecting Photoinduced Reactions at the Air–Water Interface
方案 1.影响空气-水界面光致反应的因素
The prominent role played by aqueous interfaces in atmospheric and environmental photochemistry has several explanations. The most obvious one is the amphiphilic character of a large fraction of organic matter and chromophoric compounds. The biogenic surfactants mentioned earlier are an emblematic example as they are ubiquitous in environmental interfaces. (137,168,169) Indeed, the photochemistry of microlayers at the ocean surface is a major source of volatile organic compounds, which compete with emissions from marine biology. (170) In addition, the accumulation of organic matter at the ocean’s surface acts as a solar filter that changes the intensity and spectral distribution of the radiation that penetrates into the inner water layers, thereby modifying the photochemistry there. (168) However, accumulation at aqueous interfaces is not reserved to surfactants, and both simulations and experiments have clearly demonstrated that many small polar molecules and even soft anions, including important oxidants from the ROS/RNS chemical families, display affinity for interfaces. (45,46,171) The accumulation of organic matter, chromophores, and oxidants in the same location therefore represents an important driving force of photooxidation chemistry at environmental aqueous interfaces.
Nevertheless, the solvation power of the air–water interface is still incompletely understood. Attempts to establish scales of interfacial polarity (172) have faced challenging problems (173) due to the asymmetry of the interfaces and preferential stabilization of some solute orientations, resulting in quite different solvent responses for compounds bearing similar groups but having different stereochemistries. (174) The enthalpic and entropic contributions to the interfacial free energy of solvation have been discussed for some organic molecules showing that the stability at the surface is mainly driven by the enthalpic component (Figure 12). (175) Similar studies have been carried out for ions. (176) However, in this case, the interface affinity remains incompletely understood, as discussed in many previous reviews. (177−186) Briefly, hard and multiply charged ions are repelled from the interface, while soft and large polarizable anions (e.g., I and Br) tend to accumulate at the air–water interface. (177,178) In any case, the difference in solvation energy between the interface and the bulk has direct consequences for the equilibrium constants of conformational, tautomeric, and acid/base equilibria, which remains a subject of intense debate in the literature (see, for instance, refs (187and188)). Changes in equilibrium constants imply changes in speciation and thus changes in photochemical behavior. The photochemistries of nitric acid and PYA described above are good examples illustrating these issues. For similar reasons, one expects variations in the effective redox potentials at the interface with respect to bulk values, particularly for ROS/RNS couples. (171,189) Estimations for the O2/O2 couple indicate a decrease as large as 0.32 V due to the poor solvation of the anion at the interface. (171)
然而,空气-水界面的溶剂化能力仍不完全清楚。由于界面的不对称性和某些溶质取向的优先稳定,建立界面极性尺度(172)的尝试面临着具有挑战性的问题(173),导致具有相似基团但具有不同立体化学的化合物的溶剂响应截然不同。 (174) 对于某些有机分子,已经讨论了焓和熵对溶剂化界面自由能的贡献,表明表面稳定性主要由焓分量驱动(图 12)。 (175) 对离子也进行了类似的研究。 (176) 然而,在这种情况下,界面亲和力仍然不完全被理解,正如许多以前的评论中所讨论的那样。 (177−186) 简而言之,硬离子和多电荷离子被界面排斥,而软离子和大的可极化阴离子(例如 I- 和 Br-)倾向于在空气-水界面积聚。 (177,178) 无论如何,界面和本体之间溶剂化能的差异对构象、互变异构和酸/碱平衡的平衡常数有直接影响,这仍然是文献中激烈争论的主题(参见,例如,参考文献(187和188))。平衡常数的变化意味着形态的变化,从而导致光化学行为的变化。上述硝酸和 PYA 的光化学是说明这些问题的好例子。出于类似的原因,人们预计界面处的有效氧化还原电位相对于体积值会发生变化,特别是对于 ROS/RNS 配对。 (171,189) 对 O2/O2- 电对的估计表明,由于界面处阴离子的溶剂化不良,电压下降幅度高达 0.32 V。 (171)关闭

Figure 12 图12

Figure 12. Free energy profiles for water accommodation of six polar and apolar organic compounds (A), and the corresponding enthalpic (B) and entropic components (C) as a function of position in a water droplet. The gray dashed line shows the Gibbs dividing surface. Reproduced with permission from ref (175).

Interface solvation not only controls the concentration of the species, it also affects the electronic properties (i.e., the reactivity). This happens in the initiation steps (absorption, photolysis, and ISC) as well as in the subsequent propagation steps (electron transfer, hydrogen transfer, and reaction with oxygen). As for bulk solvation, interfacial solvation may be the origin of solvatochromism effects, producing blue shifts (hydroperoxides) or red shifts (ozone) in the gas-phase absorption bands. On the other hand, large fluctuations of the solvation shell influence the shapes of the bands, and both effects (band shifts and broadening) eventually modify the photochemical rate constants. (45) Moreover, solvation effects modulate ISC in photosensitizers, as shown by experiments (190,191) and calculations, (192,193) and in the more general case of conical intersections. (193−197) In contrast, interfacial solvent effects on quantum yields have not yet been fully investigated. The case of PYA represents a limiting case in which the photochemistry at the aqueous interface differs from that in the gas phase. (146−148,153) The case of nitrate illustrates the change in quantum yield due to solvent cage effects. (63,87,89,90)
界面溶剂化不仅控制物质的浓度,还影响电子性质(即反应性)。这发生在起始步骤(吸收、光解和 ISC)以及随后的传播步骤(电子转移、氢转移和与氧反应)中。至于本体溶剂化,界面溶剂化可能是溶剂变色效应的起源,在气相吸收带中产生蓝移(氢过氧化物)或红移(臭氧)。另一方面,溶剂化壳层的动会影响能带的形状,这两种效应(能带位移和展宽)最终都会改变光化学速率常数。(45) 此外,溶剂化效应调节光敏剂中的 ISC,如实验 (190,191) 和计算 (192,193) 以及更一般的锥形交叉点情况所示。(193−197) 相比之下,界面溶剂对量子产率的影响尚未得到充分研究。PYA的情况代表了一种极限情况,其中水界面的光化学与气相中的光化学不同。(146−148,153)硝酸盐的情况说明了由于溶剂笼效应导致的量子产率变化。(63,87,89,90)
After the photochemical formation of the initial radicals, reactions proceed via energy or electron transfer, HAT or PCET mechanisms, reaction with oxygen, etc. The relaxation of the photosensitizer triplet state via phosphorescence is also conceivable and may compete with other processes. Electron transfer at interfaces has a long history that is justified by a broad applicability ranging from electrochemistry to biology and solar energy technologies (see ref (198) for a recent review on these processes). Likewise, in bulk solution, Marcus theory (199−202) can account for the observed solvent effects on electron transfer at interfaces. (203,204) A study of photoinduced electron transfer at water/dimethylaniline interfaces (205,206) concluded that the interfacial processes are faster than the bulk processes, possibly due to faster solvation dynamics at the interface. (207) The stability of solvated electrons at aqueous interfaces has also been studied, (208−210) and simulations have predicted that the water surface electron affinity is ∼0.6 eV higher than in the bulk, and that the conduction band edge is deeper in energy. Electron transfer is also relevant in PCET mechanisms and can be significantly affected by aqueous solvation, as shown for the 3SO2 + H2O reaction (116) and the H2CO + HO2 reaction. (189) One should note that the kinetics of such bimolecular reactions are often controlled by differences in intermolecular energy differences between the highest occupied and lowest unoccupied molecular orbitals. At the interface, these differences depend strongly on the electric field created by the polarized water surface, which in turn varies with the proton donor or acceptor character of the reactants. (189)
在初始自由基的光化学形成之后,反应通过能量或电子转移、HAT 或 PCET 机制、与氧反应等进行。通过磷光使光敏剂三重态的松弛也是可以想象的,并且可能与其他过程竞争。电子在界面上的转移有着悠久的历史,从电化学到生物学和太阳能技术的广泛适用性证明了这一点(参见参考文献(198)关于这些过程的最新综述)。同样,在本体溶液中,Marcus理论(199−202)可以解释观察到的溶剂对界面处电子转移的影响。(203,204) 一项关于水/二甲基苯胺界面光生电子转移的研究 (205,206) 得出结论,界面过程比本体过程更快,可能是由于界面处的溶剂化动力学更快。(207)还研究了溶剂化电子在水界面上的稳定性,(208−210)和模拟预测,水表面电亲和力比本体高∼0.6 eV,并且导带边缘的能量更深。电子转移与PCET机制也有关,并且可能受到水溶剂化的显着影响,如3SO2 + H2O反应(116)和H2CO + HO2反应所示。(189) 人们应该注意到,这种双分子反应的动力学通常由最高占据和最低未占据的分子轨道之间的分子间能量差异控制。在界面上,这些差异很大程度上取决于极化水表面产生的电场,而极化水表面又随反应物的质子供体或受体特性而变化。(189)

5. Conclusions 5. 结论

Click to copy section link
点击复制版块链接
Section link copied!

Similarly to thermal reactions, many photoinduced oxidation reactions at the air–water interface are significantly enhanced in comparison with the corresponding gas-phase or bulk processes. However, despite a considerable amount of experimental and theoretical work, the precise mechanisms by which this effect is achieved remain incompletely understood. Enthalpic and entropic factors may result in the accumulation of photoactive species at the interface, where partial hydration modify the electronic absorption energies, as well as the reactivity of the radicals formed in the early stages of the photo-oxidation processes. Interfacial dynamic effects are important as well. They have been found to be responsible for band broadening and the increase of photolytic rate constants. Likewise, solvent cage effects reducing the quantum yields turn out to be less important at the interface that in the bulk and this also contributes to enhance the photoreactivity. Various other effects have been briefly presented in the discussion above: changes in equilibrium constants (tautomeric, acid/base), effective redox potentials, rate constants for electron of hydrogen transfer, etc. All of them deserve further examination through a combination of new experiments and theoretical simulations in order to achieve a deeper understanding of these challenging phenomena, which have implications in many areas. For instance, incorporation of photo-oxidation mechanisms on cloud water surfaces would represent a considerable improvement in atmospheric models. Accounting for enhanced interfacial photochemistry is also relevant to design novel applications in water treatment and disinfection technologies, which already use the photolysis of O3 and H2O2 to trigger the oxidation of organic matter. Finally, these processes also open avenues in synthetic organic chemistry. On-water reactions are now broadly studied in microdroplets, thin films and microfluidic systems thanks to the development of innovative setups, and in many cases processes have been found to be dramatically accelerated in comparison to bulk reactions. Photoinduced oxidation on-water is likely to experience similar success in the coming years.
与热反应类似,与相应的气相或本体过程相比,空气-水界面处的许多光生氧化反应显着增强。然而,尽管进行了大量的实验和理论工作,但实现这种效应的确切机制仍然不完全清楚。焓因子和熵因子可能导致光活性物质在界面处的积累,其中部分水合改变了电子吸收能,以及在光氧化过程的早期阶段形成的自由基的反应性。界面动态效应也很重要。已发现它们负责谱带展宽和光解速率常数的增加。同样,降低量子产率的溶剂笼效应在界面上不如在本体界面上重要,这也有助于增强光反应性。在上面的讨论中已经简要介绍了各种其他效应:平衡常数(互变异构、酸/碱)、有效氧化还原电位、氢转移电子的速率常数等的变化。所有这些都值得通过新的实验和理论模拟的结合来进一步研究,以便更深入地理解这些具有挑战性的现象,这些现象在许多领域都有影响。例如,在云水面上加入光氧化机制将代表大气模型的相当大的改进。考虑到增强的界面光化学也与设计水处理和消毒技术中的新应用有关,这些技术已经利用 O3 和 H2O2 的光解来触发有机物的氧化。 最后,这些过程也为合成有机化学开辟了道路。由于创新装置的发展,现在在微滴、薄膜和微流体系统中对水上反应进行了广泛的研究,在许多情况下,与本体反应相比,水上反应被发现大大加快了速度。未来几年,水上光诱导氧化可能会取得类似的成功。

Author Information 作者信息

Click to copy section link
点击复制版块链接
Section link copied!

Acknowledgments 确认

Click to copy section link
点击复制版块链接
Section link copied!

J.M.A. thanks the Generalitat de Catalunya (Grant No. 2017SGR348) for financial support. M.F.R.-L. and M.T.C.M.-C. are grateful to the French CINES (project lct2550) for providing computational resources.
J.M.A. 感谢加泰罗尼亚政府(拨款编号:2017SGR348)的财政支持。M.F.R.-L.和 M.T.C.M.-C.感谢法国CINES(lct2550项目)提供的计算资源。

References 引用

Click to copy section link
点击复制版块链接
Section link copied!

This article references 210 other publications.
本文引用了210篇其他文献。

  1. 1
    Reichardt, C. Solvents and Solvent Effects in Organic Chemistry, 3rd ed.; Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim, 2003.
  2. 2
    Ruiz-López, M. F.; Francisco, J. S.; Martins-Costa, M. T. C.; Anglada, J. M. Molecular reactions at aqueous interfaces. Nat. Rev. Chem. 2020,  DOI: 10.1038/s41570-020-0203-2
  3. 3
    Narayan, S.; Muldoon, J.; Finn, M. G.; Fokin, V. V.; Kolb, H. C.; Sharpless, K. B. ″On water″: Unique reactivity of organic compounds in aqueous suspension. Angew. Chem., Int. Ed. 2005, 44, 32753279,  DOI: 10.1002/anie.200462883
  4. 4
    Enami, S.; Hoffmann, M. R.; Colussi, A. J. Extensive H-atom abstraction from benzoate by OH-radicals at the air-water interface. Phys. Chem. Chem. Phys. 2016, 18, 3150531512,  DOI: 10.1039/C6CP06652F
  5. 5
    Enami, S.; Colussi, A. J. Efficient scavenging of Criegee intermediates on water by surface-active cis-pinonic acid. Phys. Chem. Chem. Phys. 2017, 19, 1704417051,  DOI: 10.1039/C7CP03869K
  6. 6
    Enami, S.; Colussi, A. J. Reactions of Criegee Intermediates with Alcohols at Air-Aqueous Interfaces. J. Phys. Chem. A 2017, 121, 51755182,  DOI: 10.1021/acs.jpca.7b04272
  7. 7
    Enami, S.; Hoffmann, M. R.; Colussi, A. J. Criegee Intermediates React with Levoglucosan on Water. J. Phys. Chem. Lett. 2017, 8, 38883894,  DOI: 10.1021/acs.jpclett.7b01665
  8. 8
    Qiu, J. T.; Ishizuka, S.; Tonokura, K.; Colussi, A. J.; Enami, S. Reactivity of Monoterpene Criegee Intermediates at Gas-Liquid Interfaces. J. Phys. Chem. A 2018, 122, 79107917,  DOI: 10.1021/acs.jpca.8b06914
  9. 9
    Qiu, J. T.; Ishizuka, S.; Tonokura, K.; Enami, S. Reactions of Criegee Intermediates with Benzoic Acid at the Gas/Liquid Interface. J. Phys. Chem. A 2018, 122, 63036310,  DOI: 10.1021/acs.jpca.8b04995
  10. 10
    Qiu, J. T.; Ishizuka, S.; Tonokura, K.; Enami, S. Interfacial vs Bulk Ozonolysis of Nerolidol. Environ. Sci. Technol. 2019, 53, 57505757,  DOI: 10.1021/acs.est.9b00364
  11. 11
    Qiu, J. T.; Ishizuka, S.; Tonokura, K.; Sato, K.; Inomata, S.; Enami, S. Effects of pH on Interfacial Ozonolysis of alpha-Terpineol. J. Phys. Chem. A 2019, 123, 71487155,  DOI: 10.1021/acs.jpca.9b05434
  12. 12
    Mmereki, B. T.; Donaldson, D. J.; Gilman, J. B.; Eliason, T. L.; Vaida, V. Kinetics and products of the reaction of gas-phase ozone with anthracene adsorbed at the air-aqueous interface. Atmos. Environ. 2004, 38, 60916103,  DOI: 10.1016/j.atmosenv.2004.08.014
  13. 13
    Enami, S.; Sakamoto, Y.; Colussi, A. J. Fenton chemistry at aqueous interfaces. Proc. Natl. Acad. Sci. U. S. A. 2014, 111, 623628,  DOI: 10.1073/pnas.1314885111
  14. 14
    Gao, D.; Jin, F.; Lee, J. K.; Zare, R. N. Aqueous microdroplets containing only ketones or aldehydes undergo Dakin and Baeyer-Villiger reactions. Chem. Sci. 2019, 10, 1097410978,  DOI: 10.1039/C9SC05112K
  15. 15
    Thomas, J. L.; Jimenez-Aranda, A.; Finlayson-Pitts, B. J.; Dabdub, D. Gas-phase molecular halogen formation from NaCl and NaBr aerosols: When are interface reactions important?. J. Phys. Chem. A 2006, 110, 18591867,  DOI: 10.1021/jp054911c
  16. 16
    Lee, J. K.; Samanta, D.; Nam, H. G.; Zare, R. N. Micrometer-sized water droplets induce spontaneous reduction. J. Am. Chem. Soc. 2019, 141, 1058510589,  DOI: 10.1021/jacs.9b03227
  17. 17
    George, C.; Ammann, M.; D’Anna, B.; Donaldson, D. J.; Nizkorodov, S. A. Heterogeneous Photochemistry in the Atmosphere. Chem. Rev. 2015, 115, 42184258,  DOI: 10.1021/cr500648z
  18. 18
    Gomez Alvarez, E.; Wortham, H.; Strekowski, R.; Zetzsch, C.; Gligorovski, S. Atmospheric Photosensitized Heterogeneous and Multiphase Reactions: From Outdoors to Indoors. Environ. Sci. Technol. 2012, 46, 19551963,  DOI: 10.1021/es2019675
  19. 19
    Kozlowski, M.; Yoon, T. Editorial for the Special Issue on Photocatalysis. J. Org. Chem. 2016, 81, 68956897,  DOI: 10.1021/acs.joc.6b01717
  20. 20
    Finlayson-Pitts, B. J.; Pitts, J. N. Chemistry of the upper and lower atmosphere: theory, experiments, and applications; Academic Press: San Diego, CA, 2000.
  21. 21
    Gligorovski, S.; Strekowski, R.; Barbati, S.; Vione, D. Environmental Implications of Hydroxyl Radicals (OH). Chem. Rev. 2015, 115, 1305113092,  DOI: 10.1021/cr500310b
  22. 22
    Monks, P. S. Gas-phase radical chemistry in the troposphere. Chem. Soc. Rev. 2005, 34, 376395,  DOI: 10.1039/b307982c
  23. 23
    Anglada, J. M.; Martins-Costa, M.; Francisco, J. S.; Ruiz-Lopez, M. F. Interconnection of Reactive Oxygen Species Chemistry across the Interfaces of Atmospheric, Environmental, and Biological Processes. Acc. Chem. Res. 2015, 48, 575583,  DOI: 10.1021/ar500412p
  24. 24
    Zhang, X.; He, S.; Chen, Z.; Zhao, Y.; Hua, W. Methyl hydroperoxide (CH3OOH) in urban, suburban and rural atmosphere: ambient concentration, budget, and contribution to the atmospheric oxidizing capacity. Atmos. Chem. Phys. 2012, 12, 89518962,  DOI: 10.5194/acp-12-8951-2012
  25. 25
    Jacob, D. J. In Handbook of Weather, Climate and Water: Atmospheric Chemistry, Hydrology and Societal Impacts; Potter, T. D., Colman, B. R., Eds.; Wiley-Interscience: Hoboken, NJ, 2003.
  26. 26
    Reed Harris, A. E.; Pajunoja, A.; Cazaunau, M.; Gratien, A.; Pangui, E.; Monod, A.; Griffith, E. C.; Virtanen, A.; Doussin, J. F.; Vaida, V. Multiphase Photochemistry of Pyruvic Acid under Atmospheric Conditions. J. Phys. Chem. A 2017, 121, 33273339,  DOI: 10.1021/acs.jpca.7b01107
  27. 27
    Zhong, J.; Kumar, M.; Anglada, J. M.; Martins-Costa, M. T. C.; Ruiz-Lopez, M. F.; Zeng, X. C.; Francisco, J. S. Atmospheric Spectroscopy and Photochemistry at Environmental Water Interfaces. Annu. Rev. Phys. Chem. 2019, 70, 4569,  DOI: 10.1146/annurev-physchem-042018-052311
  28. 28
    Lee, M. H.; Heikes, B. G.; O’Sullivan, D. W. Hydrogen peroxide and organic hydroperoxide in the troposphere: A review. Atmos. Environ. 2000, 34, 34753494,  DOI: 10.1016/S1352-2310(99)00432-X
  29. 29
    Khan, M. A. H.; Cooke, M. C.; Utembe, S. R.; Xiao, P.; Morris, W. C.; Derwent, R. G.; Archibald, A. T.; Jenkin, M. E.; Percival, C. J.; Shallcross, D. E. The global budgets of organic hydroperoxides for present and pre-industrial scenarios. Atmos. Environ. 2015, 110, 6574,  DOI: 10.1016/j.atmosenv.2015.03.045
  30. 30
    Halliwell, B. Reactive species and antioxidants. Redox biology is a fundamental theme of aerobic life. Plant Physiol. 2006, 141, 312322,  DOI: 10.1104/pp.106.077073
  31. 31
    Langlais, B.; Reckhow, D. A.; Brink, D. R. Ozone in Water Treatment. Application and Engineering; Lewis Publishers: Chelsea, MI, 1991.
  32. 32
    Penkett, S. A. Atmospheric Chemistry: Hydrogen peroxide in cloudwater. Nature 1986, 319, 624624,  DOI: 10.1038/319624a0
  33. 33
    Wang, C. X.; Chen, Z. M. Effect of CH3OOH on the atmospheric concentration of OH radicals. Prog. Nat. Sci. 2006, 16, 11411149,  DOI: 10.1080/10020070612330121
  34. 34
    von Sonntag, C.; von Gunten, U. Chemistry of Ozone in Water and Wastewater Treatment: From Basic Principles to Applications; IWA Publishing, 2012.
  35. 35
    Staehelin, J.; Hoigne, J. Decomposition of ozone in water - rate of initiation by hydroxide ions and hydrogen-peroxide. Environ. Sci. Technol. 1982, 16, 676681,  DOI: 10.1021/es00104a009
  36. 36
    Herrmann, H.; Hoffmann, D.; Schaefer, T.; Bräuer, P.; Tilgner, A. Tropospheric Aqueous-Phase Free-Radical Chemistry: Radical Sources, Spectra, Reaction Kinetics and Prediction Tools. ChemPhysChem 2010, 11, 37963822,  DOI: 10.1002/cphc.201000533
  37. 37
    Bianco, A.; Passananti, M.; Brigante, M.; Mailhot, G. Photochemistry of the Cloud Aqueous Phase: A Review. Molecules 2020, 25, 423,  DOI: 10.3390/molecules25020423
  38. 38
    Iriti, M.; Faoro, F. Oxidative stress, the paradigm of ozone toxicity in plants and animals. Water, Air, Soil Pollut. 2007, 187, 285301,  DOI: 10.1007/s11270-007-9517-7
  39. 39
    Jaeglé, L.; Jacob, D. J.; Brune, W. H.; Wennberg, P. O. Chemistry of HOx radicals in the upper troposphere. Atmos. Environ. 2001, 35, 469489,  DOI: 10.1016/S1352-2310(00)00376-9
  40. 40
    Hewitt, C. N.; Kok, G. L. Formation and Occurence of Organic Hydroperoxides in the Troposphere: Laboratory and Field Observations. J. Atmos. Chem. 1991, 12, 181194,  DOI: 10.1007/BF00115779
  41. 41
    Hellpointner, E.; Gäb, S. Detection of methyl, hydroxymethyl and hydroxyethyl hydroperoxides in air and precipitation. Nature 1989, 337, 631634,  DOI: 10.1038/337631a0
  42. 42
    Jacob, D. J. Heterogeneous chemistry and tropospheric ozone. Atmos. Environ. 2000, 34, 21312159,  DOI: 10.1016/S1352-2310(99)00462-8
  43. 43
    Finlayson-Pitts, B. J.; Pitts, J. N., Jr. Atmospheric Chemistry: Fundamental and Experimentals Techniques; John Wiley and Sons: New York, 1986.
  44. 44
    Frost, G.; Vaida, V. Atmospheric Implications of the Photolysis of the Ozone-Water Weakly-Bound Complex. J. Geophys. Res. 1995, 100, 1880318809,  DOI: 10.1029/95JD01940
  45. 45
    Anglada, J. M.; Martins-Costa, M.; Ruiz-López, M. F.; Francisco, J. S. Spectroscopic signatures of ozone at the air–water interface and photochemistry implications. Proc. Natl. Acad. Sci. U. S. A. 2014, 111, 1161811623,  DOI: 10.1073/pnas.1411727111
  46. 46
    Vácha, R.; Slavíček, P.; Mucha, M.; Finlayson-Pitts, B. J.; Jungwirth, P. Adsorption of Atmospherically relevant Gases at the Air/Water Interface: Free Energy Profiles of Aqueous Solvation of N2, O2, O3, H2O, HO2, and H2O2. J. Phys. Chem. A 2004, 108, 1157311579,  DOI: 10.1021/jp046268k
  47. 47
    Vieceli, J.; Roeselova, M.; Potter, N.; Dang, L. X.; Garrett, B. C.; Tobias, D. J. Molecular dynamics simulations of atmospheric oxidants at the air-water interface: Solvation and accommodation of OH and O3. J. Phys. Chem. B 2005, 109, 1587615892,  DOI: 10.1021/jp051361+
  48. 48
    Vaghjiani, G. L.; Ravishankara, A. R. Photodissociation of H2O2 and CH3OOH at 248 nm and 298 K: Quantum yields for OH, O(3P) and H(2S). J. Chem. Phys. 1990, 92, 9961003,  DOI: 10.1063/1.458081
  49. 49
    Vaghjiani, G. L.; Turnipseed, A. A.; Warren, R. F.; Ravishankara, A. R. Photodissociation of H2O2 at 193 and 222 nm: Products and quantum yields. J. Chem. Phys. 1992, 96, 58785886,  DOI: 10.1063/1.462684
  50. 50
    Blitz, M. A.; Heard, D. E.; Pilling, M. J. Wavelength dependent photodissociation of CH3OOH: Quantum yields for CH3O and OH, and measurement of the OH+CH3OOH rate coefficient. J. Photochem. Photobiol., A 2005, 176, 107113,  DOI: 10.1016/j.jphotochem.2005.09.017
  51. 51
    Roehl, C. M.; Marka, Z.; Fry, J. L.; Wennberg, P. O. Near-UV photolysis cross-sections of CH3OOH and HOCH2OOH determined via action spectroscopy. Atmos. Chem. Phys. 2007, 7, 713720,  DOI: 10.5194/acp-7-713-2007
  52. 52
    Matthews, J.; Sinha, A.; Francisco, J. S. The Importance of Weak Absorption Features in Promoting Tropospheric Radical Production. Proc. Natl. Acad. Sci. U. S. A. 2005, 102, 74497452,  DOI: 10.1073/pnas.0502687102
  53. 53
    Herrmann, H. On the photolysis of simple anions and neutral molecules as sources of O/OH, SOx and Cl in aqueous solution. Phys. Chem. Chem. Phys. 2007, 9, 39353964,  DOI: 10.1039/B618565G
  54. 54
    Zellner, R.; Exner, M.; Herrmann, H. Absolute OH quantum yields in the laser photolysis of nitrate, nitrite and dissolved H2O2 at 308 and 351 nm in the temperature-range 278–353 K. J. Atmos. Chem. 1990, 10, 411425,  DOI: 10.1007/BF00115783
  55. 55
    Anastasio, C.; Robles, T. Light absorption by soluble chemical species in Arctic and Antarctic snow. J. Geophys. Res. 2007, 112, 2217,  DOI: 10.1029/2007JD008695
  56. 56
    Kamboures, M. A.; Nizkorodov, S. A.; Gerber, R. B. Ultrafast photochemistry of methyl hydroperoxide on ice particles. Proc. Natl. Acad. Sci. U. S. A. 2010, 107, 66006604,  DOI: 10.1073/pnas.0907922106
  57. 57
    Epstein, S. A.; Shemesh, D.; Tran, V. T.; Nizkorodov, S. A.; Gerber, R. B. Absorption Spectra and Photolysis of Methyl Peroxide in Liquid and Frozen Water. J. Phys. Chem. A 2012, 116, 60686077,  DOI: 10.1021/jp211304v
  58. 58
    Monod, A.; Chevallier, E.; Durand Jolibois, R.; Doussin, J. F.; Picquet-Varrault, B.; Carlier, P. Photooxidation of methylhydroperoxide and ethylhydroperoxide in the aqueous phase under simulated cloud droplet conditions. Atmos. Environ. 2007, 41, 24122426,  DOI: 10.1016/j.atmosenv.2006.10.006
  59. 59
    Warneck, P. The relative importance of various pathways for the oxidation of sulfur dioxide and nitrogen dioxide in sunlit continental fair weather clouds. Phys. Chem. Chem. Phys. 1999, 1, 54715483,  DOI: 10.1039/a906558j
  60. 60
    Martins-Costa, M. T. C.; Ruiz-Lopez, M. F. Highly accurate computation of free energies in complex systems through horsetail QM/MM molecular dynamics combined with free-energy perturbation theory. Theor. Chem. Acc. 2017, 136, 50,  DOI: 10.1007/s00214-017-2078-y
  61. 61
    Martins-Costa, M. T. C.; Ruiz-Lopez, M. F. Reaching Multi-Nanosecond Timescales in Combined QM/MM Molecular Dynamics Simulations through Parallel Horsetail Sampling. J. Comput. Chem. 2017, 38, 659668,  DOI: 10.1002/jcc.24723
  62. 62
    Martins-Costa, M. T. C.; Anglada, J. M.; Francisco, J. S.; Ruiz-Lopez, M. F. Impacts of cloud water droplets on the OH production rate from peroxide photolysis. Phys. Chem. Chem. Phys. 2017, 19, 3162131627,  DOI: 10.1039/C7CP06813A
  63. 63
    Nissenson, P.; Dabdub, D.; Das, R.; Maurino, V.; Minero, C.; Vione, D. Evidence of the water-cage effect on the photolysis of and FeOH2+. Implications of this effect and of H2O2 surface accumulation on photochemistry at the air–water interface of atmospheric droplets. Atmos. Environ. 2010, 44, 48594866,  DOI: 10.1016/j.atmosenv.2010.08.035
  64. 64
    Finlayson-Pitts, B. J.; Wingen, L. M.; Sumner, A. L.; Syomin, D.; Ramazan, K. A. The heterogeneous hydrolysis of NO2 in laboratory systems and in outdoor and indoor atmospheres: An integrated mechanism. Phys. Chem. Chem. Phys. 2003, 5, 223242,  DOI: 10.1039/b208564j
  65. 65
    Crowley, J. N.; Carl, S. A. OH formation in the photoexcitation of NO2 beyond the dissociation threshold in the presence of water vapor. J. Phys. Chem. A 1997, 101, 41784184,  DOI: 10.1021/jp970319e
  66. 66
    Morel, O.; Simonaitis, R.; Heicklen, J. Ultraviolet absorption spectra of HO2NO2, CCl3O2NO2, CCl2FO2NO2, and CH3O2NO2. Chem. Phys. Lett. 1980, 73, 3842,  DOI: 10.1016/0009-2614(80)85197-9
  67. 67
    Murdachaew, G.; Varner, M. E.; Phillips, L. F.; Finlayson-Pitts, B. J.; Gerber, R. B. Nitrogen dioxide at the air-water interface: trapping, absorption, and solvation in the bulk and at the surface. Phys. Chem. Chem. Phys. 2013, 15, 204212,  DOI: 10.1039/C2CP42810E
  68. 68
    Martins-Costa, M. T. C.; Anglada, J. M.; Francisco, J. S.; Ruiz-Lopez, M. F. Theoretical Investigation of the Photoexcited NO2+H2O reaction at the Air-Water Interface and Its Atmospheric Implications. Chem. - Eur. J. 2019, 25, 1389913904,  DOI: 10.1002/chem.201902769
  69. 69
    Warneck, P.; Williams, J. The Atmospheric Chemist’s Companion; Springer: The Netherlands, 2012.
  70. 70
    Dillon, T. J.; Crowley, J. N. Reactive quenching of electronically excited NO2* and NO3* by H2O as potential sources of atmospheric HOx radicals. Atmos. Chem. Phys. 2018, 18, 1400514015,  DOI: 10.5194/acp-18-14005-2018
  71. 71
    Li, S.; Matthews, J.; Sinha, A. Atmospheric hydroxyl radical production from electronically excited NO2 and H2O. Science 2008, 319, 16571660,  DOI: 10.1126/science.1151443
  72. 72
    George, C.; Strekowski, R. S.; Kleffmann, J.; Stemmler, K.; Ammann, M. Photoenhanced uptake of gaseous NO2 on solid-organic compounds: a photochemical source of HONO?. Faraday Discuss. 2005, 130, 195210,  DOI: 10.1039/b417888m
  73. 73
    Stemmler, K.; Ammann, M.; Donders, C.; Kleffmann, J.; George, C. Photosensitized reduction of nitrogen dioxide on humic acid as a source of nitrous acid. Nature 2006, 440, 195198,  DOI: 10.1038/nature04603
  74. 74
    Stemmler, K.; Ndour, M.; Elshorbany, Y.; Kleffmann, J.; D’Anna, B.; George, C.; Bohn, B.; Ammann, M. Light induced conversion of nitrogen dioxide into nitrous acid on submicron humic acid aerosol. Atmos. Chem. Phys. 2007, 7, 42374248,  DOI: 10.5194/acp-7-4237-2007
  75. 75
    Ammann, M.; Rossler, E.; Strekowski, R.; George, C. Nitrogen dioxide multiphase chemistry: Uptake kinetics on aqueous solutions containing phenolic compounds. Phys. Chem. Chem. Phys. 2005, 7, 25132518,  DOI: 10.1039/b501808k
  76. 76
    Cazoir, D.; Brigante, M.; Ammar, R.; D’Anna, B.; George, C. Heterogeneous photochemistry of gaseous NO2 on solid fluoranthene films: A source of gaseous nitrous acid (HONO) in the urban environment. J. Photochem. Photobiol., A 2014, 273, 2328,  DOI: 10.1016/j.jphotochem.2013.07.016
  77. 77
    Liu, J. P.; Li, S.; Mekic, M.; Jiang, H. Y.; Zhou, W. T.; Loisel, G.; Song, W.; Wang, X. M.; Gligorovski, S. Photoenhanced Uptake of NO2 and HONO Formation on Real Urban Grime. Environ. Sci. Technol. Lett. 2019, 6, 413417,  DOI: 10.1021/acs.estlett.9b00308
  78. 78
    Knipping, E. M.; Dabdub, D. Modeling surface-mediated renoxification of the atmosphere via reaction of gaseous nitric oxide with deposited nitric acid. Atmos. Environ. 2002, 36, 57415748,  DOI: 10.1016/S1352-2310(02)00652-0
  79. 79
    Rivera-Figueroa, A. M.; Sumner, A. L.; Finlayson-Pitts, B. J. Laboratory Studies of Potential Mechanisms of Renoxification of Tropospheric Nitric Acid. Environ. Sci. Technol. 2003, 37, 548554,  DOI: 10.1021/es020828g
  80. 80
    Handley, S. R.; Clifford, D.; Donaldson, D. J. Photochemical loss of nitric acid on organic films: A possible recycling mechanism for NOx. Environ. Sci. Technol. 2007, 41, 38983903,  DOI: 10.1021/es062044z
  81. 81
    Finlayson-Pitts, B. J. Reactions at surfaces in the atmosphere: integration of experiments and theory as necessary (but not necessarily sufficient) for predicting the physical chemistry of aerosols. Phys. Chem. Chem. Phys. 2009, 11, 77607779,  DOI: 10.1039/b906540g
  82. 82
    Zhou, X.; Zhang, N.; TerAvest, M.; Tang, D.; Hou, J.; Bertman, S.; Alaghmand, M.; Shepson, P. B.; Carroll, M. A.; Griffith, S.; Dusanter, S.; Stevens, P. S. Nitric acid photolysis on forest canopy surface as a source for tropospheric nitrous acid. Nat. Geosci. 2011, 4, 440443,  DOI: 10.1038/ngeo1164
  83. 83
    Baergen, A. M.; Donaldson, D. J. Photochemical Renoxification of Nitric Acid on Real Urban Grime. Environ. Sci. Technol. 2013, 47, 815820,  DOI: 10.1021/es3037862
  84. 84
    Ye, C.; Zhou, X.; Pu, D.; Stutz, J.; Festa, J.; Spolaor, M.; Tsai, C.; Cantrell, C.; Mauldin, R. L.; Campos, T.; Weinheimer, A.; Hornbrook, R. S.; Apel, E. C.; Guenther, A.; Kaser, L.; Yuan, B.; Karl, T.; Haggerty, J.; Hall, S.; Ullmann, K.; Smith, J. N.; Ortega, J.; Knote, C. Rapid cycling of reactive nitrogen in the marine boundary layer. Nature 2016, 532, 489491,  DOI: 10.1038/nature17195
  85. 85
    Zhou, X.; Gao, H.; He, Y.; Huang, G.; Bertman, S. B.; Civerolo, K.; Schwab, J. Nitric acid photolysis on surfaces in low-NOx environments: Significant atmospheric implications. Geophys. Res. Lett. 2003, 30, 2217  DOI: 10.1029/2003GL018620
  86. 86
    Schuttlefield, J.; Rubasinghege, G.; El-Maazawi, M.; Bone, J.; Grassian, V. H. Photochemistry of adsorbed nitrate. J. Am. Chem. Soc. 2008, 130, 1221012211,  DOI: 10.1021/ja802342m
  87. 87
    Scharko, N. K.; Berke, A. E.; Raff, J. D. Release of Nitrous Acid and Nitrogen Dioxide from Nitrate Photolysis in Acidic Aqueous Solutions. Environ. Sci. Technol. 2014, 48, 1199112001,  DOI: 10.1021/es503088x
  88. 88
    Dubowski, Y.; Colussi, A. J.; Hoffmann, M. R. Nitrogen dioxide release in the 302 nm band photolysis of spray-frozen aqueous nitrate solutions. Atmospheric implications. J. Phys. Chem. A 2001, 105, 49284932,  DOI: 10.1021/jp0042009
  89. 89
    Richards, N. K.; Wingen, L. M.; Callahan, K. M.; Nishino, N.; Kleinman, M. T.; Tobias, D. J.; Finlayson-Pitts, B. J. Nitrate Ion Photolysis in Thin Water Films in the Presence of Bromide Ions. J. Phys. Chem. A 2011, 115, 58105821,  DOI: 10.1021/jp109560j
  90. 90
    Richards-Henderson, N. K.; Callahan, K. M.; Nissenson, P.; Nishino, N.; Tobias, D. J.; Finlayson-Pitts, B. J. Production of gas phase NO2 and halogens from the photolysis of thin water films containing nitrate, chloride and bromide ions at room temperature. Phys. Chem. Chem. Phys. 2013, 15, 1763617646,  DOI: 10.1039/c3cp52956h
  91. 91
    Yu, Y.; Ezell, M. J.; Zelenyuk, A.; Imre, D.; Alexander, L.; Ortega, J.; Thomas, J. L.; Gogna, K.; Tobias, D. J.; D’Anna, B.; Harmon, C. W.; Johnson, S. N.; Finlayson-Pitts, B. J. Nitrate ion photochemistry at interfaces: a new mechanism for oxidation of alpha-pinene. Phys. Chem. Chem. Phys. 2008, 10, 30633071,  DOI: 10.1039/b719495a
  92. 92
    Wingen, L. M.; Moskun, A. C.; Johnson, S. N.; Thomas, J. L.; Roeselova, M.; Tobias, D. J.; Kleinman, M. T.; Finlayson-Pitts, B. J. Enhanced surface photochemistry in chloride-nitrate ion mixtures. Phys. Chem. Chem. Phys. 2008, 10, 56685677,  DOI: 10.1039/b806613b
  93. 93
    Richards-Henderson, N. K.; Anderson, C.; Anastasio, C.; Finlayson-Pitts, B. J. The effect of cations on NO2 production from the photolysis of aqueous thin water films of nitrate salts. Phys. Chem. Chem. Phys. 2015, 17, 3221132218,  DOI: 10.1039/C5CP05325K
  94. 94
    Finlayson-Pitts, B. J. Introductory lecture: atmospheric chemistry in the Anthropocene. Faraday Discuss. 2017, 200, 1158,  DOI: 10.1039/C7FD00161D
  95. 95
    Ye, C. X.; Gao, H. L.; Zhang, N.; Zhou, X. L. Photolysis of Nitric Acid and Nitrate on Natural and Artificial Surfaces. Environ. Sci. Technol. 2016, 50, 35303536,  DOI: 10.1021/acs.est.5b05032
  96. 96
    Ye, C. X.; Zhang, N.; Gao, H. L.; Zhou, X. L. Matrix effect on surface-catalyzed photolysis of nitric acid. Sci. Rep. 2019, 9, 10,  DOI: 10.1038/s41598-018-37973-x
  97. 97
    Nissenson, P.; Knox, C. J. H.; Finlayson-Pitts, B. J.; Phillips, L. F.; Dabdub, D. Enhanced photolysis in aerosols: evidence for important surface effects. Phys. Chem. Chem. Phys. 2006, 8, 47004710,  DOI: 10.1039/b609219e
  98. 98
    Yang, H. S.; Finlayson-Pitts, B. J. Infrared spectroscopic studies of binary solutions of nitric acid and water and ternary solutions of nitric acid, sulfuric acid, and water at room temperature: Evidence for molecular nitric acid at the surface. J. Phys. Chem. A 2001, 105, 18901896,  DOI: 10.1021/jp004224f
  99. 99
    Shamay, E. S.; Buch, V.; Parrinello, M.; Richmond, G. L. At the water’s edge: Nitric acid as a weak acid. J. Am. Chem. Soc. 2007, 129, 1291012911,  DOI: 10.1021/ja074811f
  100. 100
    Bianco, R.; Wang, S. Z.; Hynes, J. T. Theoretical study of the dissociation of nitric acid at a model aqueous surface. J. Phys. Chem. A 2007, 111, 1103311042,  DOI: 10.1021/jp075054a
  101. 101
    Hirokawa, J.; Kato, T.; Mafune, F. Uptake of Gas-Phase Nitrous Acid by pH-Controlled Aqueous Solution Studied by a Wetted Wall Flow Tube. J. Phys. Chem. A 2008, 112, 1214312150,  DOI: 10.1021/jp8051483
  102. 102
    Miller, Y.; Thomas, J. L.; Kemp, D. D.; Finlayson-Pitts, B. J.; Gordon, M. S.; Tobias, D. J.; Gerber, R. B. Structure of Large Nitrate-Water Clusters at Ambient Temperatures: Simulations with Effective Fragment Potentials and Force Fields with Implications for Atmospheric Chemistry. J. Phys. Chem. A 2009, 113, 1280512814,  DOI: 10.1021/jp9070339
  103. 103
    Wang, S.; Bianco, R.; Hynes, J. T. Dissociation of nitric acid at an aqueous surface: Large amplitude motions in the contact ion pair to solvent-separated ion pair conversion. Phys. Chem. Chem. Phys. 2010, 12, 82418249,  DOI: 10.1039/c002299n
  104. 104
    Lewis, T.; Winter, B.; Stern, A. C.; Baer, M. D.; Mundy, C. J.; Tobias, D. J.; Hemminger, J. C. Does Nitric Acid Dissociate at the Aqueous Solution Surface?. J. Phys. Chem. C 2011, 115, 2118321190,  DOI: 10.1021/jp205842w
  105. 105
    Moussa, S. G.; Stern, A. C.; Raff, J. D.; Dilbeck, C. W.; Tobias, D. J.; Finlayson-Pitts, B. J. Experimental and theoretical studies of the interaction of gas phase nitric acid and water with a self-assembled monolayer. Phys. Chem. Chem. Phys. 2013, 15, 448458,  DOI: 10.1039/C2CP42405C
  106. 106
    Kido Soule, M. C.; Blower, P. G.; Richmond, G. L. Nonlinear Vibrational Spectroscopic Studies of the Adsorption and Speciation of Nitric Acid at the Vapor/Acid Solution Interface. J. Phys. Chem. A 2007, 111, 33493357,  DOI: 10.1021/jp0686994
  107. 107
    Stockwell, W. R.; Calvert, J. G. The mechanism of the HO-SO2 reaction. Atmos. Environ. 1983, 17, 22312235,  DOI: 10.1016/0004-6981(83)90220-2
  108. 108
    Jayne, J. T.; Pöschl, U.; Chen, Y.-m.; Dai, D.; Molina, L. T.; Worsnop, D. R.; Kolb, C. E.; Molina, M. J. Pressure and temperature dependence of the gas-phase reaction of SO3 with H2O and the heterogeneous reaction of SO3 with H2O/H2SO4 surfaces. J. Phys. Chem. A 1997, 101, 1000010011,  DOI: 10.1021/jp972549z
  109. 109
    Morokuma, K.; Muguruma, C. Ab initio Molecular orbital Study of the Mechanism of the Gas Phase Reaction SO3 + H2O: Importance of the secons water molecule. J. Am. Chem. Soc. 1994, 116, 1031610317,  DOI: 10.1021/ja00101a068
  110. 110
    Loerting, T.; Liedl, K. R. Toward elimination of discrepancies between theory and experiment: The rate constant of the atmospheric conversion of SO3 to H2SO4. Proc. Natl. Acad. Sci. U. S. A. 2000, 97, 88748878,  DOI: 10.1073/pnas.97.16.8874
  111. 111
    Seinfeld, J. H.; Pandis, S. N. Atmospheric Chemistry and Physics: From Air Pollution to Climate Change, 2nd ed.; Wiley, 2006.
  112. 112
    Mirabel, P.; Clavelin, J. L. Application of nucleation to the study of the gas-phase photooxidation of sulfur-dioxide. J. Chem. Phys. 1979, 70, 57675772,  DOI: 10.1063/1.437405
  113. 113
    Gurjar, B. R.; Jain, A.; Sharma, A.; Agarwal, A.; Gupta, P.; Nagpure, A. S.; Lelieveld, J. Human health risks in megacities due to air pollution. Atmos. Environ. 2010, 44, 46064613,  DOI: 10.1016/j.atmosenv.2010.08.011
  114. 114
    Donaldson, D. J.; Kroll, J. A.; Vaida, V. Gas-phase hydrolysis of triplet SO2: A possible direct route to atmospheric acid formation. Sci. Rep. 2016, 6, 30000,  DOI: 10.1038/srep30000
  115. 115
    Kroll, J. A.; Frandsen, B. N.; Kjaergaard, H. G.; Vaida, V. Atmospheric Hydroxyl Radical Source: Reaction of Triplet SO2 and Water. J. Phys. Chem. A 2018, 122, 44654469,  DOI: 10.1021/acs.jpca.8b03524
  116. 116
    Anglada, J. M.; Martins-Costa, M. T. C.; Francisco, J. S.; Ruiz-López, M. F. Triplet State Promoted Reaction of SO2 with H2O by Competition Between Proton Coupled Electron Transfer (pcet) and Hydrogen Atom Transfer (hat) Processes. Phys. Chem. Chem. Phys. 2019, 21, 97799784,  DOI: 10.1039/C9CP01105F
  117. 117
    Olivella, S.; Anglada, J. M.; Sole, A.; Bofill, J. M. Mechanism of the hydrogen transfer from the OH group to oxygen-centered radicals: Proton-coupled electron-transfer versus radical hydrogen abstraction. Chem. - Eur. J. 2004, 10, 34043410,  DOI: 10.1002/chem.200305714
  118. 118
    Anglada, J. M. Complex mechanism of the gas phase reaction between formic acid and hydroxyl radical. Proton coupled electron transfer versus radical hydrogen abstraction mechanisms. J. Am. Chem. Soc. 2004, 126, 98099820,  DOI: 10.1021/ja0481169
  119. 119
    Anglada, J. M.; Olivella, S.; Sole, A. Hydrogen transfer between sulfuric acid and hydroxyl radical in the gas phase: Competition among hydrogen atom transfer, proton-coupled electron-transfer, and double proton transfer. J. Phys. Chem. A 2006, 110, 19821990,  DOI: 10.1021/jp056155g
  120. 120
    Gonzalez, J.; Anglada, J. M. Gas Phase Reaction of Nitric Acid with Hydroxyl Radical without and with Water. A Theoretical Investigation. J. Phys. Chem. A 2010, 114, 91519162,  DOI: 10.1021/jp102935d
  121. 121
    Anglada, J. M.; Gonzalez, J. Different Catalytic Effects of a Single Water Molecule: The Gas-Phase Reaction of Formic Acid with Hydroxyl Radical in Water Vapor. ChemPhysChem 2009, 10, 30343045,  DOI: 10.1002/cphc.200900387
  122. 122
    Jorgensen, S.; Jensen, C.; Kjaergaard, H. G.; Anglada, J. M. The gas-phase reaction of methane sulfonic acid with the hydroxyl radical without and with water vapor. Phys. Chem. Chem. Phys. 2013, 15, 51405150,  DOI: 10.1039/c3cp44034f
  123. 123
    Anglada, J. M.; Olivella, S.; Solé, A. Unexpected Reactivity of Amidogen Radical in the Gas Phase Degradation of Nitric Acid. J. Am. Chem. Soc. 2014, 136, 68346837,  DOI: 10.1021/ja501967x
  124. 124
    Anglada, J. M.; Olivella, S.; Sole, A. Atmospheric formation of the NO3 radical from gas-phase reaction of HNO3 acid with the NH2 radical: proton-coupled electron-transfer versus hydrogen atom transfer mechanisms. Phys. Chem. Chem. Phys. 2014, 16, 1943719445,  DOI: 10.1039/C4CP02792B
  125. 125
    Anglada, J. M.; Solé, A. The Atmospheric Oxidation of HONO by OH, Cl, and ClO Radicals. J. Phys. Chem. A 2017, 121, 96989707,  DOI: 10.1021/acs.jpca.7b10715
  126. 126
    Anglada, J. M.; Crehuet, R.; Sole, A. The gas phase oxidation of HCOOH by Cl and NH2 radicals. Proton coupled electron transfer versus hydrogen atom transfer. Mol. Phys. 2019, 117, 14301441,  DOI: 10.1080/00268976.2018.1554829
  127. 127
    Anglada, J. M.; Crehuet, R.; Adhikari, S.; Francisco, J. S.; Xia, Y. Reactivity of hydropersulfides toward the hydroxyl radical unraveled: disulfide bond cleavage, hydrogen atom transfer, and proton-coupled electron transfer. Phys. Chem. Chem. Phys. 2018, 20, 47934804,  DOI: 10.1039/C7CP07570G
  128. 128
    Zhong, J.; Zhu, C.; Li, L.; Richmond, G. L.; Francisco, J. S.; Zeng, X. C. Interaction of SO2 with the Surface of a Water Nanodroplet. J. Am. Chem. Soc. 2017, 139, 1716817174,  DOI: 10.1021/jacs.7b09900
  129. 129
    Jayne, J.; Davidovits, P.; Worsnop, D.; Zahniser, M.; Kolb, C. Uptake of sulfur dioxide (g) by aqueous surfaces as a function of pH: the effect of chemical reaction at the interface. J. Phys. Chem. 1990, 94, 60416048,  DOI: 10.1021/j100378a076
  130. 130
    Donaldson, D.; Guest, J. A.; Goh, M. C. Evidence for adsorbed SO2 at the aqueous-air interface. J. Phys. Chem. 1995, 99, 93139315,  DOI: 10.1021/j100023a002
  131. 131
    Yang, H.; Wright, N. J.; Gagnon, A. M.; Gerber, R. B.; Finlayson-Pitts, B. J. An upper limit to the concentration of an SO2 complex at the air–water interface at 298 K: infrared experiments and ab initio calculations. Phys. Chem. Chem. Phys. 2002, 4, 18321838,  DOI: 10.1039/b108907b
  132. 132
    Shamay, E. S.; Johnson, K. E.; Richmond, G. L. Dancing on Water: The Choreography of Sulfur Dioxide Adsorption to Aqueous Surfaces. J. Phys. Chem. C 2011, 115, 2530425314,  DOI: 10.1021/jp2064326
  133. 133
    Martins-Costa, M. T. C.; Anglada, J. M.; Francisco, J. S.; Ruiz-López, M. F. Photochemistry of SO2 at the Air–Water Interface: A Source of OH and HOSO Radicals. J. Am. Chem. Soc. 2018, 140, 1234112344,  DOI: 10.1021/jacs.8b07845
  134. 134
    Ruiz-Lopez, M. F.; Martins-Costa, M. T. C.; Anglada, J. M.; Francisco, J. S. A New Mechanism of Acid Rain Generation from HOSO at the Air-Water Interface. J. Am. Chem. Soc. 2019, 141, 1656416568,  DOI: 10.1021/jacs.9b07912
  135. 135
    Ciuraru, R.; Fine, L.; van Pinxteren, M.; D’Anna, B.; Herrmann, H.; George, C. Photosensitized production of functionalized and unsaturated organic compounds at the air-sea interface. Sci. Rep. 2015, 5, 12741  DOI: 10.1038/srep12741
  136. 136
    Ciuraru, R.; Fine, L.; van Pinxteren, M.; D’Anna, B.; Herrmann, H.; George, C. Unravelling New Processes at Interfaces: Photochemical Isoprene Production at the Sea Surface. Environ. Sci. Technol. 2015, 49, 1319913205,  DOI: 10.1021/acs.est.5b02388
  137. 137
    Brüggemann, M.; Hayeck, N.; Bonnineau, C.; Pesce, S.; Alpert, P. A.; Perrier, S.; Zuth, C.; Hoffmann, T.; Chen, J.; George, C. Interfacial photochemistry of biogenic surfactants: a major source of abiotic volatile organic compounds. Faraday Discuss. 2017, 200, 5974,  DOI: 10.1039/C7FD00022G
  138. 138
    Shrestha, M.; Luo, M.; Li, Y.; Xiang, B.; Xiong, W.; Grassian, V. H. Let there be light: stability of palmitic acid monolayers at the air/salt water interface in the presence and absence of simulated solar light and a photosensitizer. Chem. Sci. 2018, 9, 57165723,  DOI: 10.1039/C8SC01957F
  139. 139
    Rapf, R. J.; Vaida, V. Sunlight as an energetic driver in the synthesis of molecules necessary for life. Phys. Chem. Chem. Phys. 2016, 18, 2006720084,  DOI: 10.1039/C6CP00980H
  140. 140
    Grosjean, D.; Williams, E. L.; Grosjean, E. Atmospheric chemistry of isoprene and of its carbonyl products. Environ. Sci. Technol. 1993, 27, 830840,  DOI: 10.1021/es00042a004
  141. 141
    Kawamura, K.; Tachibana, E.; Okuzawa, K.; Aggarwal, S. G.; Kanaya, Y.; Wang, Z. F. High abundances of water-soluble dicarboxylic acids, ketocarboxylic acids and alpha-dicarbonyls in the mountaintop aerosols over the North China Plain during wheat burning season. Atmos. Chem. Phys. 2013, 13, 82858302,  DOI: 10.5194/acp-13-8285-2013
  142. 142
    Reed Harris, A. E.; Doussin, J.-F.; Carpenter, B. K.; Vaida, V. Gas-Phase Photolysis of Pyruvic Acid: The Effect of Pressure on Reaction Rates and Products. J. Phys. Chem. A 2016, 120, 1012310133,  DOI: 10.1021/acs.jpca.6b09058
  143. 143
    Chang, X. P.; Fang, Q.; Cui, G. L. Mechanistic photodecarboxylation of pyruvic acid: Excited-state proton transfer and three-state intersection. J. Chem. Phys. 2014, 141, 154311  DOI: 10.1063/1.4898085
  144. 144
    Reed Harris, A. E.; Cazaunau, M.; Gratien, A.; Pangui, E.; Doussin, J.-F.; Vaida, V. Atmospheric Simulation Chamber Studies of the Gas-Phase Photolysis of Pyruvic Acid. J. Phys. Chem. A 2017, 121, 83488358,  DOI: 10.1021/acs.jpca.7b05139
  145. 145
    Chiang, Y.; Kresge, A. J.; Pruszynski, P. Keto-enol equilibria in the pyruvic acid system: determination of the keto-enol equilibrium constants of pyruvic acid and pyruvate anion and the acidity constant of pyruvate enol in aqueous solution. J. Am. Chem. Soc. 1992, 114, 31033107,  DOI: 10.1021/ja00034a053
  146. 146
    Guzman, M. I.; Colussi, A. J.; Hoffmann, M. R. Photoinduced oligomerization of aqueous pyruvic acid. J. Phys. Chem. A 2006, 110, 36193626,  DOI: 10.1021/jp056097z
  147. 147
    Griffith, E. C.; Carpenter, B. K.; Shoemaker, R. K.; Vaida, V. Photochemistry of aqueous pyruvic acid. Proc. Natl. Acad. Sci. U. S. A. 2013, 110, 1171411719,  DOI: 10.1073/pnas.1303206110
  148. 148
    Eugene, A. J.; Xia, S.-S.; Guzman, M. I. Negative production of acetoin in the photochemistry of aqueous pyruvic acid. Proc. Natl. Acad. Sci. U. S. A. 2013, 110, E4274E4275,  DOI: 10.1073/pnas.1313991110
  149. 149
    Griffith, E. C.; Carpenter, B. K.; Shoemaker, R. K.; Vaida, V. Reply to Eugene et al.: Photochemistry of aqueous pyruvic acid. Proc. Natl. Acad. Sci. U. S. A. 2013, 110, E4276E4276,  DOI: 10.1073/pnas.1316367110
  150. 150
    Eugene, A. J.; Guzman, M. I. Reactivity of Ketyl and Acetyl Radicals from Direct Solar Actinic Photolysis of Aqueous Pyruvic Acid. J. Phys. Chem. A 2017, 121, 29242935,  DOI: 10.1021/acs.jpca.6b11916
  151. 151
    Vaida, V.; Reed Harris, A. E.; Rapf, R. J.; Perkins, R. J.; Carpenter, B. K. Comment on “Reactivity of Ketyl and Acetyl Radicals from Direct Solar Actinic Photolysis of Aqueous Pyruvic Acid”. J. Phys. Chem. A 2017, 121, 87388740,  DOI: 10.1021/acs.jpca.7b06018
  152. 152
    Eugene, A. J.; Guzman, M. I. Reply to “Comment on ‘Reactivity of Ketyl and Acetyl Radicals from Direct Solar Actinic Photolysis of Aqueous Pyruvic Acid’”. J. Phys. Chem. A 2017, 121, 87418744,  DOI: 10.1021/acs.jpca.7b08273
  153. 153
    Rapf, R. J.; Perkins, R. J.; Carpenter, B. K.; Vaida, V. Mechanistic Description of Photochemical Oligomer Formation from Aqueous Pyruvic Acid. J. Phys. Chem. A 2017, 121, 42724282,  DOI: 10.1021/acs.jpca.7b03310
  154. 154
    Xia, S. S.; Eugene, A. J.; Guzman, M. I. Cross Photoreaction of Glyoxylic and Pyruvic Acids in Model Aqueous Aerosol. J. Phys. Chem. A 2018, 122, 64576466,  DOI: 10.1021/acs.jpca.8b05724
  155. 155
    Leermakers, P. A.; Vesley, G. F. Photochemistry of alpha-keto acids and alpha-keto esters 0.1. Photolysis of pyruvic acid and benzoylformic acid. J. Am. Chem. Soc. 1963, 85, 37763779,  DOI: 10.1021/ja00906a013
  156. 156
    Closs, G. L.; Miller, R. J. Photo-reduction and photodecarboxylation of pyruvic-acid. Applications of CIDNP to mechanistic photochemistry. J. Am. Chem. Soc. 1978, 100, 34833494,  DOI: 10.1021/ja00479a033
  157. 157
    Eugene, A. J.; Guzman, M. I. Production of Singlet Oxygen (1O2) during the Photochemistry of Aqueous Pyruvic Acid: The Effects of pH and Photon Flux under Steady-State O2(aq) Concentration. Environ. Sci. Technol. 2019, 53, 1242512432,  DOI: 10.1021/acs.est.9b03742
  158. 158
    Rapf, R. J.; Dooley, M. R.; Kappes, K.; Perkins, R. J.; Vaida, V. pH Dependence of the Aqueous Photochemistry of α-Keto Acids. J. Phys. Chem. A 2017, 121, 83688379,  DOI: 10.1021/acs.jpca.7b08192
  159. 159
    Rapf, R. J.; Perkins, R. J.; Dooley, M. R.; Kroll, J. A.; Carpenter, B. K.; Vaida, V. Environmental Processing of Lipids Driven by Aqueous Photochemistry of alpha-Keto Acids. ACS Cent. Sci. 2018, 4, 624630,  DOI: 10.1021/acscentsci.8b00124
  160. 160
    Grgić, I.; Nieto-Gligorovski, L. I.; Net, S.; Temime-Roussel, B.; Gligorovski, S.; Wortham, H. Light induced multiphase chemistry of gas-phase ozone on aqueous pyruvic and oxalic acids. Phys. Chem. Chem. Phys. 2010, 12, 698707,  DOI: 10.1039/B914377G
  161. 161
    Gordon, B. P.; Moore, F. G.; Scatena, L. F.; Richmond, G. L. On the Rise: Experimental and Computational Vibrational Sum Frequency Spectroscopy Studies of Pyruvic Acid and Its Surface Active Oligomer Species at the Air-Water Interface. J. Phys. Chem. A 2019, 123, 1060910619,  DOI: 10.1021/acs.jpca.9b08854
  162. 162
    Eugene, A. J.; Pillar, E. A.; Colussi, A. J.; Guzman, M. I. Enhanced Acidity of Acetic and Pyruvic Acids on the Surface of Water. Langmuir 2018, 34, 93079313,  DOI: 10.1021/acs.langmuir.8b01606
  163. 163
    Fu, Y.; Zhang, Y.; Zhang, F.; Chen, J.; Zhu, Z.; Yu, X.-Y. Does interfacial photochemistry play a role in the photolysis of pyruvic acid in water?. Atmos. Environ. 2018, 191, 3645,  DOI: 10.1016/j.atmosenv.2018.07.061
  164. 164
    Fu, H. B.; Ciuraru, R.; Dupart, Y.; Passananti, M.; Tinel, L.; Rossignol, S.; Perrier, S.; Donaldson, D. J.; Chen, J. M.; George, C. Photosensitized Production of Atmospherically Reactive Organic Compounds at the Air/Aqueous Interface. J. Am. Chem. Soc. 2015, 137, 83488351,  DOI: 10.1021/jacs.5b04051
  165. 165
    Tinel, L.; Rossignol, S.; Bianco, A.; Passananti, M.; Perrier, S.; Wang, X.; Brigante, M.; Donaldson, D. J.; George, C. Mechanistic Insights on the Photosensitized Chemistry of a Fatty Acid at the Air/Water Interface. Environ. Sci. Technol. 2016, 50, 1104111048,  DOI: 10.1021/acs.est.6b03165
  166. 166
    Bernard, F.; Ciuraru, R.; Boreave, A.; George, C. Photosensitized Formation of Secondary Organic Aerosols above the Air/Water Interface. Environ. Sci. Technol. 2016, 50, 86788686,  DOI: 10.1021/acs.est.6b03520
  167. 167
    Rossignol, S.; Tinel, L.; Bianco, A.; Passananti, M.; Brigante, M.; Donaldson, D. J.; George, C. Atmospheric photochemistry at a fatty acid-coated air-water interface. Science 2016, 353, 699702,  DOI: 10.1126/science.aaf3617
  168. 168
    Donaldson, D. J.; Vaida, V. The Influence of Organic Films at the Air-Aqueous Boundary on Atmospheric Processes. Chem. Rev. 2006, 106, 14451461,  DOI: 10.1021/cr040367c
  169. 169
    Andreae, M. O.; Crutzen, P. J. Atmospheric aerosols: Biogeochemical sources and role in atmospheric chemistry. Science 1997, 276, 10521058,  DOI: 10.1126/science.276.5315.1052
  170. 170
    Bruggemann, M.; Hayeck, N.; George, C. Interfacial photochemistry at the ocean surface is a global source of organic vapors and aerosols. Nat. Commun. 2018, 9, 3222  DOI: 10.1038/s41467-018-05687-3
  171. 171
    Martins-Costa, M. T. C.; Anglada, J. M.; Francisco, J. S.; Ruiz-Lopez, M. Reactivity of Atmospherically Relevant Small Radicals at the Air–Water Interface. Angew. Chem., Int. Ed. 2012, 51, 54135417,  DOI: 10.1002/anie.201200656
  172. 172
    Wang, H. F.; Borguet, E.; Eisenthal, K. B. Generalized interface polarity scale based on second harmonic spectroscopy. J. Phys. Chem. B 1998, 102, 49274932,  DOI: 10.1021/jp9806563
  173. 173
    Sen, S.; Yamaguchi, S.; Tahara, T. Different Molecules Experience Different Polarities at the Air/Water Interface. Angew. Chem., Int. Ed. 2009, 48, 64396442,  DOI: 10.1002/anie.200901094
  174. 174
    Martins-Costa, M. C.; Ruiz-Lopez, M. Solvation effects on electronic polarization and reactivity indices at the air–water interface: insights from a theoretical study of cyanophenols. Theor. Chem. Acc. 2015, 134, 17  DOI: 10.1007/s00214-014-1609-z
  175. 175
    Hub, J. S.; Caleman, C.; van der Spoel, D. Organic molecules on the surface of water droplets - an energetic perspective. Phys. Chem. Chem. Phys. 2012, 14, 95379545,  DOI: 10.1039/c2cp40483d
  176. 176
    Caleman, C.; Hub, J. S.; van Maaren, P. J.; van der Spoel, D. Atomistic simulation of ion solvation in water explains surface preference of halides. Proc. Natl. Acad. Sci. U. S. A. 2011, 108, 68386842,  DOI: 10.1073/pnas.1017903108
  177. 177
    Jungwirth, P.; Tobias, D. J. Specific ion effects at the air/water interface. Chem. Rev. 2006, 106, 12591281,  DOI: 10.1021/cr0403741
  178. 178
    Petersen, P. B.; Saykally, R. J. On the nature of ions at the liquid water surface. Annu. Rev. Phys. Chem. 2006, 57, 333364,  DOI: 10.1146/annurev.physchem.57.032905.104609
  179. 179
    Jungwirth, P.; Winter, B. Ions at aqueous interfaces: From water surface to hydrated proteins. Annu. Rev. Phys. Chem. 2008, 59, 343366,  DOI: 10.1146/annurev.physchem.59.032607.093749
  180. 180
    Levin, Y.; dos Santos, A. P.; Diehl, A. Ions at the Air-Water Interface: An End to a Hundred-Year-Old Mystery?. Phys. Rev. Lett. 2009, 103, 257802  DOI: 10.1103/PhysRevLett.103.257802
  181. 181
    Otten, D. E.; Shaffer, P. R.; Geissler, P. L.; Saykally, R. J. Elucidating the mechanism of selective ion adsorption to the liquid water surface. Proc. Natl. Acad. Sci. U. S. A. 2012, 109, 701705,  DOI: 10.1073/pnas.1116169109
  182. 182
    Tobias, D. J.; Stern, A. C.; Baer, M. D.; Levin, Y.; Mundy, C. J. Simulation and Theory of Ions at Atmospherically Relevant Aqueous Liquid-Air Interfaces. Annu. Rev. Phys. Chem. 2013, 64, 339359,  DOI: 10.1146/annurev-physchem-040412-110049
  183. 183
    Levin, Y.; dos Santos, A. P. Ions at hydrophobic interfaces. J. Phys.: Condens. Matter 2014, 26, 203101,  DOI: 10.1088/0953-8984/26/20/203101
  184. 184
    Sun, L.; Li, X.; Tu, Y. Q.; Agren, H. Origin of ion selectivity at the air/water interface. Phys. Chem. Chem. Phys. 2015, 17, 43114318,  DOI: 10.1039/C4CP03338H
  185. 185
    Tong, Y.; Zhang, I. Y.; Campen, R. K. Experimentally quantifying anion polarizability at the air/water interface. Nat. Commun. 2018, 9, 1313,  DOI: 10.1038/s41467-018-03598-x
  186. 186
    Wise, P. K.; Ben-Amotz, D. Interfacial Adsorption of Neutral and Ionic Solutes in a Water Droplet. J. Phys. Chem. B 2018, 122, 34473453,  DOI: 10.1021/acs.jpcb.7b10488
  187. 187
    Agmon, N.; Bakker, H. J.; Campen, R. K.; Henchman, R. H.; Pohl, P.; Roke, S.; Thamer, M.; Hassanali, A. Protons and Hydroxide Ions in Aqueous Systems. Chem. Rev. 2016, 116, 76427672,  DOI: 10.1021/acs.chemrev.5b00736
  188. 188
    Saykally, R. J. Air/water interface: Two sides of the acid–base story. Nat. Chem. 2013, 5, 8284,  DOI: 10.1038/nchem.1556
  189. 189
    Martins-Costa, M. T. C.; Anglada, J. M.; Francisco, J. S.; Ruiz-Lopez, M. F. Reactivity of Volatile Organic Compounds at the Surface of a Water Droplet. J. Am. Chem. Soc. 2012, 134, 1182111827,  DOI: 10.1021/ja304971e
  190. 190
    Sitzmann, E. V.; Langan, J.; Eisenthal, K. B. Intermolecular effects on intersystem crossing studied on the picosecond time scale - the solvent polarity effect on the rate of singlet to triplet intersystem crossing of diphenylcarbene. J. Am. Chem. Soc. 1984, 106, 18681869,  DOI: 10.1021/ja00318a069
  191. 191
    Kellmann, A. Intersystem crossing and internal conversion quantum yields of acridine in polar and nonpolar solvents. J. Phys. Chem. 1977, 81, 11951198,  DOI: 10.1021/j100527a014
  192. 192
    Munoz Losa, A.; Fdez. Galvan, I.; Sanchez, M. L.; Martin, M. E.; Aguilar, M. A. Solvent effects on internal conversions and intersystem crossings: The radiationless de-excitation of acrolein in water. J. Phys. Chem. B 2008, 112, 877884,  DOI: 10.1021/jp075706v
  193. 193
    Sanchez-Rodriguez, J. A.; Mohamadzade, A.; Mai, S.; Ashwood, B.; Pollum, M.; Marquetand, P.; Gonzalez, L.; Crespo-Hernandez, C. E.; Ullrich, S. 2-Thiouracil intersystem crossing photodynamics studied by wavelength-dependent photoelectron and transient absorption spectroscopies. Phys. Chem. Chem. Phys. 2017, 19, 1975619766,  DOI: 10.1039/C7CP02258A
  194. 194
    Toniolo, A.; Olsen, S.; Manohar, L.; Martinez, T. Conical intersection dynamics in solution: the chromophore of green fluorescent protein. Faraday Discuss. 2004, 127, 149163,  DOI: 10.1039/B401167H
  195. 195
    Burghardt, I.; Cederbaum, L. S.; Hynes, J. T. Environmental effects on a conical intersection: A model study. Faraday Discuss. 2004, 127, 395411,  DOI: 10.1039/b315071b
  196. 196
    Yamazaki, S.; Kato, S. Locating the lowest free-energy point on conical intersection in polar solvent: Reference interaction site model self-consistent field study of ethylene and CH2NH2+. J. Chem. Phys. 2005, 123, 114510  DOI: 10.1063/1.2038867
  197. 197
    Spezia, R.; Burghardt, I.; Hynes, J. T. Conical intersections in solution: non-equilibrium versus equilibrium solvation. Mol. Phys. 2006, 104, 903914,  DOI: 10.1080/00268970500417895
  198. 198
    Benjamin, I. Reaction Dynamics at Liquid Interfaces. Annu. Rev. Phys. Chem. 2015, 66, 165188,  DOI: 10.1146/annurev-physchem-040214-121428
  199. 199
    Marcus, R. A. On the theory of electron-transfer reactions. VI. Unified treatment for homogeneous and electrode reactions. J. Chem. Phys. 1965, 43, 679701,  DOI: 10.1063/1.1696792
  200. 200
    Marcus, R. Reorganization free energy for electron transfers at liquid-liquid and dielectric semiconductor-liquid interfaces. J. Phys. Chem. 1990, 94, 10501055,  DOI: 10.1021/j100366a005
  201. 201
    Marcus, R. Theory of electron-transfer rates across liquid-liquid interfaces. J. Phys. Chem. 1990, 94, 41524155,  DOI: 10.1021/j100373a051
  202. 202
    Marcus, R. Theory of electron-transfer rates across liquid-liquid interfaces. 2. Relationships and application. J. Phys. Chem. 1991, 95, 20102013,  DOI: 10.1021/j100158a023
  203. 203
    Eugster, N.; Fermín, D. J.; Girault, H. H. Photoinduced electron transfer at liquid/liquid interfaces. Part VI. On the thermodynamic driving force dependence of the phenomenological electron-transfer rate constant. J. Phys. Chem. B 2002, 106, 34283433,  DOI: 10.1021/jp015533o
  204. 204
    Eugster, N.; Fermín, D. J.; Girault, H. H. Photoinduced electron transfer at liquid| liquid interfaces: Dynamics of the heterogeneous photoreduction of quinones by self-assembled porphyrin ion pairs. J. Am. Chem. Soc. 2003, 125, 48624869,  DOI: 10.1021/ja029589n
  205. 205
    McArthur, E. A.; Eisenthal, K. B. Ultrafast excited-state electron transfer at an organic liquid/aqueous interface. J. Am. Chem. Soc. 2006, 128, 10681069,  DOI: 10.1021/ja056518q
  206. 206
    Rao, Y.; Xu, M.; Jockusch, S.; Turro, N. J.; Eisenthal, K. B. Dynamics of excited state electron transfer at a liquid interface using time-resolved sum frequency generation. Chem. Phys. Lett. 2012, 544, 16,  DOI: 10.1016/j.cplett.2012.05.054
  207. 207
    Cooper, J. K.; Benjamin, I. Photoinduced excited state electron transfer at liquid/liquid interfaces. J. Phys. Chem. B 2014, 118, 77037714,  DOI: 10.1021/jp409541u
  208. 208
    Sagar, D. M.; Bain, C. D.; Verlet, J. R. R. Hydrated electrons at the water/air interface. J. Am. Chem. Soc. 2010, 132, 69176919,  DOI: 10.1021/ja101176r
  209. 209
    Siefermann, K. R.; Liu, Y. X.; Lugovoy, E.; Link, O.; Faubel, M.; Buck, U.; Winter, B.; Abel, B. Binding energies, lifetimes and implications of bulk and interface solvated electrons in water. Nat. Chem. 2010, 2, 274279,  DOI: 10.1038/nchem.580
  210. 210
    Gaiduk, A. P.; Pham, T. A.; Govoni, M.; Paesani, F.; Galli, G. Electron affinity of liquid water. Nat. Commun. 2018, 9, 247  DOI: 10.1038/s41467-017-02673-z

Cited By

Click to copy section linkSection link copied!
Citation Statements
  • Supporting
    Supporting1
  • Mentioning
    Mentioning55
  • Contrasting
    Contrasting0
Explore this article's citation statements on scite.ai

This article is cited by 47 publications.

  1. Alison Bain, Lara Lalemi, Nathan Croll Dawes, Rachael E. H. Miles, Alexander M. Prophet, Kevin R. Wilson, Bryan R. Bzdek. Surfactant Partitioning Dynamics in Freshly Generated Aerosol Droplets. Journal of the American Chemical Society 2024, 146 (23) , 16028-16038. https://doi.org/10.1021/jacs.4c03041化学TOP简介JCI 2.66EI检索SCI升级版 化学1区SCI Q1IF 14.4
  2. Josep M. Anglada, Marilia T. C. Martins-Costa, Joseph S. Francisco, Manuel F. Ruiz-López. Triplet State Radical Chemistry: Significance of the Reaction of 3SO2 with HCOOH and HNO3. Journal of the American Chemical Society 2024, 146 (20) , 14297-14306. https://doi.org/10.1021/jacs.4c03938化学TOP简介JCI 2.66EI检索SCI升级版 化学1区SCI Q1IF 14.4
  3. Genevieve Duché, John M Sanderson. The Chemical Reactivity of Membrane Lipids. Chemical Reviews 2024, 124 (6) , 3284-3330. https://doi.org/10.1021/acs.chemrev.3c00608化学TOP简介JCI 4.32EI检索SCI升级版 化学1区SCI Q1IF 51.4
  4. Wei Wang, Yangyang Liu, Tao Wang, Qiuyue Ge, Kejian Li, Juan Liu, Wenbo You, Longqian Wang, Lifang Xie, Hongbo Fu, Jianmin Chen, Liwu Zhang. Significantly Accelerated Photosensitized Formation of Atmospheric Sulfate at the Air–Water Interface of Microdroplets. Journal of the American Chemical Society 2024, 146 (10) , 6580-6590. https://doi.org/10.1021/jacs.3c11892化学TOP简介JCI 2.66EI检索SCI升级版 化学1区SCI Q1IF 14.4
  5. Marilia T. C. Martins-Costa, Manuel F. Ruiz-López. Reactivity of Monoethanolamine at the Air–Water Interface and Implications for CO2 Capture. The Journal of Physical Chemistry B 2024, 128 (5) , 1289-1297. https://doi.org/10.1021/acs.jpcb.3c06856简介JCI 0.46EI检索SCI升级版 化学2区SCI Q3IF 2.8
  6. Thufail M. Ismail, Deepak Patkar, Pookkottu K. Sajith, Milind M. Deshmukh. Interplay of Hydrogen, Pnicogen, and Chalcogen Bonding in X(H2O)n=1–5 (X = NO, NO+, and NO–) Complexes: Energetics Insights via a Molecular Tailoring Approach. The Journal of Physical Chemistry A 2023, 127 (49) , 10360-10374. https://doi.org/10.1021/acs.jpca.3c04181简介JCI 0.61EI检索SCI升级版 化学2区SCI Q2IF 2.7
  7. Zepeng Rao, Ye-Guang Fang, Yishuai Pan, Wanchao Yu, Baoliang Chen, Joseph S. Francisco, Chongqin Zhu, Chiheng Chu. Accelerated Photolysis of H2O2 at the Air–Water Interface of a Microdroplet. Journal of the American Chemical Society 2023, 145 (45) , 24717-24723. https://doi.org/10.1021/jacs.3c08101化学TOP简介JCI 2.66EI检索SCI升级版 化学1区SCI Q1IF 14.4
  8. Tao Wang, Carmen Kalalian, Daniel Fillion, Sébastien Perrier, Jianmin Chen, Florent Domine, Liwu Zhang, Christian George. Sunlight Induces the Production of Atmospheric Volatile Organic Compounds (VOCs) from Thermokarst Ponds. Environmental Science & Technology 2023, 57 (45) , 17363-17373. https://doi.org/10.1021/acs.est.3c03303环境科学与生态学TOP简介JCI 1.56EI检索SCI升级版 环境科学与生态学1区SCI Q1IF 10.8
  9. Longqian Wang, Kejian Li, Yangyang Liu, Kedong Gong, Juan Liu, Jianpeng Ao, Qiuyue Ge, Wei Wang, Minbiao Ji, Liwu Zhang. Significantly Accelerated Hydroxyl Radical Generation by Fe(III)–Oxalate Photochemistry in Aerosol Droplets. The Journal of Physical Chemistry A 2023, 127 (1) , 250-260. https://doi.org/10.1021/acs.jpca.2c05919简介JCI 0.61EI检索SCI升级版 化学2区SCI Q2IF 2.7
  10. Shendong Tan, Xiaomeng Zhang, Yongjian Lian, Xi Chen, Shi Yin, Lin Du, Maofa Ge. OH Group Orientation Leads to Organosulfate Formation at the Liquid Aerosol Surface. Journal of the American Chemical Society 2022, 144 (37) , 16953-16964. https://doi.org/10.1021/jacs.2c05807化学TOP简介JCI 2.66EI检索SCI升级版 化学1区SCI Q1IF 14.4
  11. Caleb J. C. Jordan, Eleanor A. Lowe, Jan R. R. Verlet. Photooxidation of the Phenolate Anion is Accelerated at the Water/Air Interface. Journal of the American Chemical Society 2022, 144 (31) , 14012-14015. https://doi.org/10.1021/jacs.2c04935化学TOP简介JCI 2.66EI检索SCI升级版 化学1区SCI Q1IF 14.4
  12. Belaid Malek, Wenchao Lu, Prabhu Prasad Mohapatra, Niluksha Walalawela, Shakeela Jabeen, Jianbo Liu, Alexander Greer. Probing the Transition State-to-Intermediate Continuum: Mechanistic Distinction between a Dry versus Wet Perepoxide in the Singlet Oxygen “Ene” Reaction at the Air–Water Interface. Langmuir 2022, 38 (19) , 6036-6048. https://doi.org/10.1021/acs.langmuir.2c00279简介JCI 0.67EI检索SCI升级版 化学2区SCI Q2IF 3.7
  13. Tatsuya Ishiyama, Tahei Tahara, Akihiro Morita. Why the Photochemical Reaction of Phenol Becomes Ultrafast at the Air–Water Interface: The Effect of Surface Hydration. Journal of the American Chemical Society 2022, 144 (14) , 6321-6325. https://doi.org/10.1021/jacs.1c13336化学TOP简介JCI 2.66EI检索SCI升级版 化学1区SCI Q1IF 14.4
  14. Javier Carmona-García, Antonio Francés-Monerris, Carlos A. Cuevas, Tarek Trabelsi, Alfonso Saiz-Lopez, Joseph S. Francisco, Daniel Roca-Sanjuán. Photochemistry and Non-adiabatic Photodynamics of the HOSO Radical. Journal of the American Chemical Society 2021, 143 (29) , 10836-10841. https://doi.org/10.1021/jacs.1c05149化学TOP简介JCI 2.66EI检索SCI升级版 化学1区SCI Q1IF 14.4
  15. Alexandra M. Deal, Rebecca J. Rapf, Veronica Vaida. Water–Air Interfaces as Environments to Address the Water Paradox in Prebiotic Chemistry: A Physical Chemistry Perspective. The Journal of Physical Chemistry A 2021, 125 (23) , 4929-4942. https://doi.org/10.1021/acs.jpca.1c02864简介JCI 0.61EI检索SCI升级版 化学2区SCI Q2IF 2.7
  16. Marilia T. C. Martins-Costa, Josep M. Anglada, Joseph S. Francisco, Manuel F. Ruiz-López. The Aqueous Surface as an Efficient Transient Stop for the Reactivity of Gaseous NO2 in Liquid Water. Journal of the American Chemical Society 2020, 142 (50) , 20937-20941. https://doi.org/10.1021/jacs.0c10364化学TOP简介JCI 2.66EI检索SCI升级版 化学1区SCI Q1IF 14.4
  17. Bo Tang, Wenyan Zhang, Wenting Chen, Wenting Tan, Guoqiang Shi, Hong Qi, Guohong Deng. Influence of aeration-induced air–water interfaces on pollutant degradation in water treatment: A theoretical and experimental study. Separation and Purification Technology 2024, 347 , 127595. https://doi.org/10.1016/j.seppur.2024.127595工程技术TOP简介JCI 1.36EI检索SCI升级版 工程技术1区SCI Q1IF 8.1
  18. Cheng-Xin Chen, Shan-Shan Yang, Ji-Wei Pang, Lei He, Ya-Ni Zang, Lan Ding, Nan-Qi Ren, Jie Ding. Anthraquinones-based photocatalysis: A comprehensive review. Environmental Science and Ecotechnology 2024, 22 , 100449. https://doi.org/10.1016/j.ese.2024.100449JCI 2.09SCI升级版 环境科学与生态学1区SCI Q1IF 14.0
  19. Petersen-Sonn Emma A, Jespersen Malte F, Johnson Matthew S, Mikkelsen Kurt V. Mechanistic Insights into UV Spectral Changes of Pyruvic Acid and Pyruvate Part 1: Interaction with Water Molecules. International Journal of Physics Research and Applications 2024, 7 (2) , 100-107. https://doi.org/10.29328/journal.ijpra.1001092
  20. Marilia T. C. Martins‐Costa, Manuel F. Ruiz‐López. The Structure of Carbon Dioxide at the Air‐Water Interface and its Chemical Implications. Chemistry – A European Journal 2024, 29 https://doi.org/10.1002/chem.202400825简介JCI 0.66EI检索SCI升级版 化学2区SCI Q2IF 3.9
  21. Lynn E. Krushinski, Kathryn J. Vannoy, Jeffrey E. Dick. Single Liquid Aerosol Microparticle Electrochemistry on a Suspended Ionic Liquid Film. Small 2024, 20 (28) https://doi.org/10.1002/smll.202308637材料科学TOP简介JCI 1.89EI检索SCI升级版 材料科学2区SCI Q1IF 13.0
  22. Thufail M. Ismail, Deepak Patkar, Pookkottu K. Sajith, Milind M. Deshmukh. Hydrogen bond strengths in microhydrated clusters of HNO and HONO: energetic insights via a molecular tailoring approach. New Journal of Chemistry 2024, 48 (25) , 11591-11605. https://doi.org/10.1039/D4NJ02163K简介JCI 0.53EI检索SCI升级版 化学3区SCI Q2IF 2.7
  23. María de los A. Garavagno, Federico J. Hernández, Rafael A. Jara-Toro, Gustavo A. Pino. Understanding the active role of water in laboratory chamber studies of reactions of the OH radical with alcohols of atmospheric relevance. Physical Chemistry Chemical Physics 2024, 26 (16) , 12745-12752. https://doi.org/10.1039/D3CP05667H简介JCI 0.69EI检索SCI升级版 化学3区SCI Q1IF 2.9
  24. Kathryn J. Vannoy, Myles Quinn Edwards, Christophe Renault, Jeffrey E. Dick. An Electrochemical Perspective on Reaction Acceleration in Droplets. Annual Review of Analytical Chemistry 2024, https://doi.org/10.1146/annurev-anchem-061622-030919简介JCI 2.44EI检索SCI升级版 化学2区SCI Q1IF 5.9
  25. Qin Dai, Xiaoyu Zhang, Jingyi Lin, Tao Cui, Wenbo Wang, Guangfei Yu, Hongbin Cao, He Zhao. Photoconverting waste amphiphiles at the mild air–water interface into easily separable value-added products. Green Chemistry 2024, 26 (4) , 2102-2113. https://doi.org/10.1039/D3GC03599A化学TOP简介JCI 1.41EI检索SCI升级版 化学1区SCI Q1IF 9.3
  26. Manuel F. Ruiz-López, Marilia T.C. Martins-Costa. Chemical Processes at Aqueous Interfaces. 2024, 559-577. https://doi.org/10.1016/B978-0-12-821978-2.00085-4
  27. Tao Wang, Yangyang Liu, Shengqian Zhou, Guochen Wang, Xiansheng Liu, Longqian Wang, Hongbo Fu, Jianmin Chen, Liwu Zhang. Key Factors Determining the Formation of Sulfate Aerosols Through Multiphase Chemistry—A Kinetic Modeling Study Based on Beijing Conditions. Journal of Geophysical Research: Atmospheres 2023, 128 (20) https://doi.org/10.1029/2022JD038382地球科学TOP简介JCI 0.91SCI升级版 地球科学2区SCI Q2IF 3.8
  28. Elise Lorre, Federica Bianchi, Irma Vybernaite-Lubiene, Jovita Mėžinė, Mindaugas Zilius. Phthalate esters delivery to the largest European lagoon: Sources, partitioning and seasonal variations. Environmental Research 2023, 235 , 116667. https://doi.org/10.1016/j.envres.2023.116667环境科学与生态学TOP简介JCI 1.98EI检索SCI升级版 环境科学与生态学2区SCI Q1IF 7.7
  29. Wen Zhang, Lu Xu, Haofei Zhang. Recent advances in mass spectrometry techniques for atmospheric chemistry research on molecular‐level. Mass Spectrometry Reviews 2023, 40 https://doi.org/10.1002/mas.21857简介JCI 1.15EI检索SCI升级版 化学2区SCI Q1IF 6.9
  30. Juliana S. A. Carneiro, Giada Innocenti, Hyun June Moon, Yoseph Guta, Laura Proaño, Carsten Sievers, Miles A. Sakwa‐Novak, Eric W. Ping, Christopher W. Jones. Insights into the Oxidative Degradation Mechanism of Solid Amine Sorbents for CO 2 Capture from Air: Roles of Atmospheric Water. Angewandte Chemie International Edition 2023, 62 (24) https://doi.org/10.1002/anie.202302887化学TOP简介JCI 2.70EI检索SCI升级版 化学1区SCI Q1IF 16.1
  31. Juliana S. A. Carneiro, Giada Innocenti, Hyun June Moon, Yoseph Guta, Laura Proaño, Carsten Sievers, Miles A. Sakwa‐Novak, Eric W. Ping, Christopher W. Jones. Insights into the Oxidative Degradation Mechanism of Solid Amine Sorbents for CO 2 Capture from Air: Roles of Atmospheric Water. Angewandte Chemie 2023, 135 (24) https://doi.org/10.1002/ange.202302887
  32. Li Lin, Sheng Zhang, Lei Dong, Yueqi Cao, Wei Zhang, Xiong Pan, Yi Li, Chi Zhang, Jingxiang Tao, Di Jia, John Crittenden. Photodegradation behavior and mechanism of dibutyl phthalate in water under flood discharge atomization. Science of The Total Environment 2023, 871 , 161822. https://doi.org/10.1016/j.scitotenv.2023.161822环境科学与生态学TOP简介JCI 1.62EI检索SCI升级版 环境科学与生态学1区SCI Q1IF 8.2
  33. Avinashkumar V. Karre, Kalliat T. Valsaraj, Vivek Vasagar. Review of air-water interface adsorption and reactions between trace gaseous organic and oxidant compounds. Science of The Total Environment 2023, 873 , 162367. https://doi.org/10.1016/j.scitotenv.2023.162367环境科学与生态学TOP简介JCI 1.62EI检索SCI升级版 环境科学与生态学1区SCI Q1IF 8.2
  34. Marilia T. C. Martins-Costa, Manuel F. Ruiz-López. Probing solvation electrostatics at the air–water interface. Theoretical Chemistry Accounts 2023, 142 (3) https://doi.org/10.1007/s00214-023-02969-y简介JCI 0.27SCI升级版 化学4区SCI Q4IF 1.6
  35. Robin Schulte, Marco Löcker, Heiko Ihmels, Maximilian Heide, Carsten Engelhard. Pushing Photochemistry into Water: Acceleration of the Di‐π‐Methane Rearrangement and the Paternó‐Büchi Reaction “On‐Water”. Chemistry – A European Journal 2023, 29 (9) https://doi.org/10.1002/chem.202203203简介JCI 0.66EI检索SCI升级版 化学2区SCI Q2IF 3.9
  36. Conor J. Bready, Vance R. Fowler, Leah A. Juechter, Luke A. Kurfman, Grace E. Mazaleski, George C. Shields. The driving effects of common atmospheric molecules for formation of prenucleation clusters: the case of sulfuric acid, formic acid, nitric acid, ammonia, and dimethyl amine. Environmental Science: Atmospheres 2022, 2 (6) , 1469-1486. https://doi.org/10.1039/D2EA00087CJCI 0.65IF 2.8
  37. Huifan Deng, Pascale S. J. Lakey, Yiqun Wang, Pan Li, Jinli Xu, Hongwei Pang, Jiangping Liu, Xin Xu, Xue Li, Xinming Wang, Yuzhong Zhang, Manabu Shiraiwa, Sasho Gligorovski. Daytime SO 2 chemistry on ubiquitous urban surfaces as a source of organic sulfur compounds in ambient air. Science Advances 2022, 8 (39) https://doi.org/10.1126/sciadv.abq6830综合性期刊TOP简介JCI 2.76EI检索SCI升级版 综合性期刊1区SCI Q1IF 11.7
  38. Kevin A. Wokosin, Emma L. Schell, Jennifer A. Faust. Emerging investigator series: surfactants, films, and coatings on atmospheric aerosol particles: a review. Environmental Science: Atmospheres 2022, 2 (5) , 775-828. https://doi.org/10.1039/D2EA00003BJCI 0.65IF 2.8
  39. Binbin Wu, Chong Zhou, Guoqiang Zhao, Jingyi Wang, Hengyi Dai, Tian Liu, Xiaoshan Zheng, Baoliang Chen, Chiheng Chu. Enhanced photochemical production of reactive intermediates at the wetland soil-water interface. Water Research 2022, 223 , 118971. https://doi.org/10.1016/j.watres.2022.118971环境科学与生态学TOP简介JCI 2.14EI检索SCI升级版 环境科学与生态学1区SCI Q1IF 11.4
  40. Kejian Li, Kedong Gong, Juan Liu, Lukas Ohnoutek, Jianpeng Ao, Yangyang Liu, Xi Chen, Guanjun Xu, Xuejun Ruan, Hanyun Cheng, Jin Han, Guodong Sui, Minbiao Ji, Ventsislav K. Valev, Liwu Zhang. Significantly accelerated photochemical and photocatalytic reactions in microdroplets. Cell Reports Physical Science 2022, 3 (6) , 100917. https://doi.org/10.1016/j.xcrp.2022.100917JCI 1.46SCI升级版 综合性期刊2区SCI Q1IF 7.9
  41. Pablo Corral Arroyo, Grégory David, Peter A. Alpert, Evelyne A. Parmentier, Markus Ammann, Ruth Signorell. Amplification of light within aerosol particles accelerates in-particle photochemistry. Science 2022, 376 (6590) , 293-296. https://doi.org/10.1126/science.abm7915综合性期刊TOP简介JCI 9.91SCI升级版 综合性期刊1区SCI Q1IF 44.7
  42. Marilia T. C. Martins-Costa, Josep M. Anglada, Joseph S. Francisco, Manuel F. Ruiz-López. Photosensitization mechanisms at the air–water interface of aqueous aerosols. Chemical Science 2022, 13 (9) , 2624-2631. https://doi.org/10.1039/D1SC06866K化学TOP简介JCI 1.41EI检索SCI升级版 化学1区SCI Q1IF 7.6
  43. Ximei Liang, Fangling Guan, Zhiyou Ling, Honghong Wang, Yunwen Tao, Elfi Kraka, Huajun Huang, Chenglong Yu, Danping Li, Jinbao He, Hansun Fang. Pivotal role of water molecules in the photodegradation of pymetrozine: New insights for developing green pesticides. Journal of Hazardous Materials 2022, 423 , 127197. https://doi.org/10.1016/j.jhazmat.2021.127197环境科学与生态学TOP简介JCI 1.85EI检索SCI升级版 环境科学与生态学1区SCI Q1IF 12.2
  44. Yu Wang, Marcello Brigante, Gilles Mailhot, David Talaga, Yanlin Wu, Wenbo Dong, Sophie Sobanska. Toward a better understanding of ferric-oxalate complex photolysis: The role of the aqueous/air interface of droplet. Chemosphere 2022, 289 , 133127. https://doi.org/10.1016/j.chemosphere.2021.133127环境科学与生态学TOP简介JCI 1.58EI检索SCI升级版 环境科学与生态学2区SCI Q1IF 8.1
  45. Manuel F. Ruiz-Lopez. Midair transformations of aerosols. Science 2021, 374 (6568) , 686-687. https://doi.org/10.1126/science.abl8914综合性期刊TOP简介JCI 9.91SCI升级版 综合性期刊1区SCI Q1IF 44.7
  46. Manuel F. Ruiz-López, Marilia T. C. Martins-Costa, Joseph S. Francisco, Josep M. Anglada. Tight electrostatic regulation of the OH production rate from the photolysis of hydrogen peroxide adsorbed on surfaces. Proceedings of the National Academy of Sciences 2021, 118 (30) https://doi.org/10.1073/pnas.2106117118简介
  47. Inna A. Nemirovskaya, Vyacheslav V. Gordeev. Features of suspended matter distribution at the atmosphere-water boundary in the Atlantic and Southern Oceans. Russian Journal of Earth Sciences 2021, 21 (5) , 1-16. https://doi.org/10.2205/2021ES000777JCI 0.23IF 0.7
Open PDF

Journal of the American Chemical Society

Cite this: J. Am. Chem. Soc. 2020, 142, 38, 16140–16155
Click to copy citationCitation copied!
https://doi.org/10.1021/jacs.0c06858
Published August 24, 2020
Copyright © 2020 American Chemical Society

Article Views

5215

Altmetric

19

Citations

Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Figure 1

    Figure 1. Gas-phase atmospheric reactions involving reactive oxygen species (ROS). Reproduced with permission from ref (23). Copyright 2015 American Chemical Society.

    Figure 2

    Figure 2. (a) Schematic representation of a Type I sensitization process involving the degradation of phenols via a reaction with an excited carbonyl compound as photosensitizer. (b) Comparison of Type I and Type II photosensitization processes. Reproduced with permission from ref (18). Copyright 2012 American Chemical Society.

    Figure 3

    Figure 3. (a) Calculated absorption cross-sections of CH3OOH in the gas phase and at the air–water interface along with different experimental values reported for the gas phase. (b) Calculated partial photolysis rates for CH3OOH. Integrated total rates are indicated. Reproduced from ref (62) with permission from the PCCP Owner Societies.

    Figure 4

    Figure 4. (a) Experimental UV spectrum of NO2 in the gas phase and calculated spectra in the gas phase and at the air–water interface. (b) Estimated upper and lower limits of the OH production rate (molecule·cm–3·s–1) for different gas-phase concentrations of NO2 (molecule·cm–3). Calculations in the gas phase (light gray) and at the air–water interface (dark gray) using k12 from ref (70). The plain and dashed red lines respectively correspond to the gas-phase and interfacial values using k12 from ref (71). The gas-phase values assume a relative humidity of 20%. The horizontal plain line represents a typical OH production rate from ozone photolysis. Reproduced with permission from ref (68).

    Figure 5

    Figure 5. (a) Schematic of the free energy surface at 298 K for the reaction of SO2(a 3B1) with one water molecule. (b) Schematic free energy surface at 298 K for the reaction of SO2(a 3B1) with a cluster of four water molecules. Reproduced from ref (116) with permission from the PCCP Owner Societies.

    Figure 6

    Figure 6. Orbital diagram for the PCET and HAT mechanisms for the reaction SO2(a 3B1) + H2O. Reproduced from ref (116) with permission from the PCCP Owner Societies.

    Figure 7

    Figure 7. Different steps of the photo-oxidation of SO2 leading to sulfate. QM/MM MD simulations at the air–water interface reveal that as HOSO radical is formed, it rapidly ionizes. This is shown in the bottom part of the figure by the time evolution of OH distances in the H2O···HOSO system (bottom, right). The structure of the transition structure for proton transfer is also displayed (bottom, left). Adapted with permission from ref (134).

    Figure 8

    Figure 8. Chemical structures of pyruvic acid, pyruvate, the corresponding gem-diols, parapyruvic acid, and zymonic acid.

    Figure 9

    Figure 9. Density profiles from MD simulations of a mixed aqueous solution containing PYA (red), PYT (blue), PPA (green), and ZYA (yellow). The concentration of each species was ∼0.25 M. The vertical dotted gray line indicates the approximate interface boundary. Reproduced with permission from ref (161).

    Figure 10

    Figure 10. Reaction flowchart showing reactions in bulk water and at the air–water interface. Numbers on arrows indicate reaction types: (1) radical reactions, (2) decarboxylation reactions, (3) anhydride formation, and (4) esterification. Red squares indicate new products identified by the authors, and chemical formulas correspond to undetermined or to too-large structures. Reproduced with permission from ref (163).

    Figure 11

    Figure 11. Proposed mechanism for the photochemical degradation of nonanoic acid at the air–water interface in the presence of humic acid as photosensitizer. Reproduced with permission from ref (166). Copyright 2016 American Chemical Society.

    Scheme 1

    Scheme 1. Factors Affecting Photoinduced Reactions at the Air–Water Interface

    Figure 12

    Figure 12. Free energy profiles for water accommodation of six polar and apolar organic compounds (A), and the corresponding enthalpic (B) and entropic components (C) as a function of position in a water droplet. The gray dashed line shows the Gibbs dividing surface. Reproduced with permission from ref (175).

  • References


    This article references 210 other publications.

    1. 1
      Reichardt, C. Solvents and Solvent Effects in Organic Chemistry, 3rd ed.; Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim, 2003.
    2. 2
      Ruiz-López, M. F.; Francisco, J. S.; Martins-Costa, M. T. C.; Anglada, J. M. Molecular reactions at aqueous interfaces. Nat. Rev. Chem. 2020,  DOI: 10.1038/s41570-020-0203-2
    3. 3
      Narayan, S.; Muldoon, J.; Finn, M. G.; Fokin, V. V.; Kolb, H. C.; Sharpless, K. B. ″On water″: Unique reactivity of organic compounds in aqueous suspension. Angew. Chem., Int. Ed. 2005, 44, 32753279,  DOI: 10.1002/anie.200462883
    4. 4
      Enami, S.; Hoffmann, M. R.; Colussi, A. J. Extensive H-atom abstraction from benzoate by OH-radicals at the air-water interface. Phys. Chem. Chem. Phys. 2016, 18, 3150531512,  DOI: 10.1039/C6CP06652F
    5. 5
      Enami, S.; Colussi, A. J. Efficient scavenging of Criegee intermediates on water by surface-active cis-pinonic acid. Phys. Chem. Chem. Phys. 2017, 19, 1704417051,  DOI: 10.1039/C7CP03869K
    6. 6
      Enami, S.; Colussi, A. J. Reactions of Criegee Intermediates with Alcohols at Air-Aqueous Interfaces. J. Phys. Chem. A 2017, 121, 51755182,  DOI: 10.1021/acs.jpca.7b04272
    7. 7
      Enami, S.; Hoffmann, M. R.; Colussi, A. J. Criegee Intermediates React with Levoglucosan on Water. J. Phys. Chem. Lett. 2017, 8, 38883894,  DOI: 10.1021/acs.jpclett.7b01665
    8. 8
      Qiu, J. T.; Ishizuka, S.; Tonokura, K.; Colussi, A. J.; Enami, S. Reactivity of Monoterpene Criegee Intermediates at Gas-Liquid Interfaces. J. Phys. Chem. A 2018, 122, 79107917,  DOI: 10.1021/acs.jpca.8b06914
    9. 9
      Qiu, J. T.; Ishizuka, S.; Tonokura, K.; Enami, S. Reactions of Criegee Intermediates with Benzoic Acid at the Gas/Liquid Interface. J. Phys. Chem. A 2018, 122, 63036310,  DOI: 10.1021/acs.jpca.8b04995
    10. 10
      Qiu, J. T.; Ishizuka, S.; Tonokura, K.; Enami, S. Interfacial vs Bulk Ozonolysis of Nerolidol. Environ. Sci. Technol. 2019, 53, 57505757,  DOI: 10.1021/acs.est.9b00364
    11. 11
      Qiu, J. T.; Ishizuka, S.; Tonokura, K.; Sato, K.; Inomata, S.; Enami, S. Effects of pH on Interfacial Ozonolysis of alpha-Terpineol. J. Phys. Chem. A 2019, 123, 71487155,  DOI: 10.1021/acs.jpca.9b05434
    12. 12
      Mmereki, B. T.; Donaldson, D. J.; Gilman, J. B.; Eliason, T. L.; Vaida, V. Kinetics and products of the reaction of gas-phase ozone with anthracene adsorbed at the air-aqueous interface. Atmos. Environ. 2004, 38, 60916103,  DOI: 10.1016/j.atmosenv.2004.08.014
    13. 13
      Enami, S.; Sakamoto, Y.; Colussi, A. J. Fenton chemistry at aqueous interfaces. Proc. Natl. Acad. Sci. U. S. A. 2014, 111, 623628,  DOI: 10.1073/pnas.1314885111
    14. 14
      Gao, D.; Jin, F.; Lee, J. K.; Zare, R. N. Aqueous microdroplets containing only ketones or aldehydes undergo Dakin and Baeyer-Villiger reactions. Chem. Sci. 2019, 10, 1097410978,  DOI: 10.1039/C9SC05112K
    15. 15
      Thomas, J. L.; Jimenez-Aranda, A.; Finlayson-Pitts, B. J.; Dabdub, D. Gas-phase molecular halogen formation from NaCl and NaBr aerosols: When are interface reactions important?. J. Phys. Chem. A 2006, 110, 18591867,  DOI: 10.1021/jp054911c
    16. 16
      Lee, J. K.; Samanta, D.; Nam, H. G.; Zare, R. N. Micrometer-sized water droplets induce spontaneous reduction. J. Am. Chem. Soc. 2019, 141, 1058510589,  DOI: 10.1021/jacs.9b03227
    17. 17
      George, C.; Ammann, M.; D’Anna, B.; Donaldson, D. J.; Nizkorodov, S. A. Heterogeneous Photochemistry in the Atmosphere. Chem. Rev. 2015, 115, 42184258,  DOI: 10.1021/cr500648z
    18. 18
      Gomez Alvarez, E.; Wortham, H.; Strekowski, R.; Zetzsch, C.; Gligorovski, S. Atmospheric Photosensitized Heterogeneous and Multiphase Reactions: From Outdoors to Indoors. Environ. Sci. Technol. 2012, 46, 19551963,  DOI: 10.1021/es2019675
    19. 19
      Kozlowski, M.; Yoon, T. Editorial for the Special Issue on Photocatalysis. J. Org. Chem. 2016, 81, 68956897,  DOI: 10.1021/acs.joc.6b01717
    20. 20
      Finlayson-Pitts, B. J.; Pitts, J. N. Chemistry of the upper and lower atmosphere: theory, experiments, and applications; Academic Press: San Diego, CA, 2000.
    21. 21
      Gligorovski, S.; Strekowski, R.; Barbati, S.; Vione, D. Environmental Implications of Hydroxyl Radicals (OH). Chem. Rev. 2015, 115, 1305113092,  DOI: 10.1021/cr500310b
    22. 22
      Monks, P. S. Gas-phase radical chemistry in the troposphere. Chem. Soc. Rev. 2005, 34, 376395,  DOI: 10.1039/b307982c
    23. 23
      Anglada, J. M.; Martins-Costa, M.; Francisco, J. S.; Ruiz-Lopez, M. F. Interconnection of Reactive Oxygen Species Chemistry across the Interfaces of Atmospheric, Environmental, and Biological Processes. Acc. Chem. Res. 2015, 48, 575583,  DOI: 10.1021/ar500412p
    24. 24
      Zhang, X.; He, S.; Chen, Z.; Zhao, Y.; Hua, W. Methyl hydroperoxide (CH3OOH) in urban, suburban and rural atmosphere: ambient concentration, budget, and contribution to the atmospheric oxidizing capacity. Atmos. Chem. Phys. 2012, 12, 89518962,  DOI: 10.5194/acp-12-8951-2012
    25. 25
      Jacob, D. J. In Handbook of Weather, Climate and Water: Atmospheric Chemistry, Hydrology and Societal Impacts; Potter, T. D., Colman, B. R., Eds.; Wiley-Interscience: Hoboken, NJ, 2003.
    26. 26
      Reed Harris, A. E.; Pajunoja, A.; Cazaunau, M.; Gratien, A.; Pangui, E.; Monod, A.; Griffith, E. C.; Virtanen, A.; Doussin, J. F.; Vaida, V. Multiphase Photochemistry of Pyruvic Acid under Atmospheric Conditions. J. Phys. Chem. A 2017, 121, 33273339,  DOI: 10.1021/acs.jpca.7b01107
    27. 27
      Zhong, J.; Kumar, M.; Anglada, J. M.; Martins-Costa, M. T. C.; Ruiz-Lopez, M. F.; Zeng, X. C.; Francisco, J. S. Atmospheric Spectroscopy and Photochemistry at Environmental Water Interfaces. Annu. Rev. Phys. Chem. 2019, 70, 4569,  DOI: 10.1146/annurev-physchem-042018-052311
    28. 28
      Lee, M. H.; Heikes, B. G.; O’Sullivan, D. W. Hydrogen peroxide and organic hydroperoxide in the troposphere: A review. Atmos. Environ. 2000, 34, 34753494,  DOI: 10.1016/S1352-2310(99)00432-X
    29. 29
      Khan, M. A. H.; Cooke, M. C.; Utembe, S. R.; Xiao, P.; Morris, W. C.; Derwent, R. G.; Archibald, A. T.; Jenkin, M. E.; Percival, C. J.; Shallcross, D. E. The global budgets of organic hydroperoxides for present and pre-industrial scenarios. Atmos. Environ. 2015, 110, 6574,  DOI: 10.1016/j.atmosenv.2015.03.045
    30. 30
      Halliwell, B. Reactive species and antioxidants. Redox biology is a fundamental theme of aerobic life. Plant Physiol. 2006, 141, 312322,  DOI: 10.1104/pp.106.077073
    31. 31
      Langlais, B.; Reckhow, D. A.; Brink, D. R. Ozone in Water Treatment. Application and Engineering; Lewis Publishers: Chelsea, MI, 1991.
    32. 32
      Penkett, S. A. Atmospheric Chemistry: Hydrogen peroxide in cloudwater. Nature 1986, 319, 624624,  DOI: 10.1038/319624a0
    33. 33
      Wang, C. X.; Chen, Z. M. Effect of CH3OOH on the atmospheric concentration of OH radicals. Prog. Nat. Sci. 2006, 16, 11411149,  DOI: 10.1080/10020070612330121
    34. 34
      von Sonntag, C.; von Gunten, U. Chemistry of Ozone in Water and Wastewater Treatment: From Basic Principles to Applications; IWA Publishing, 2012.
    35. 35
      Staehelin, J.; Hoigne, J. Decomposition of ozone in water - rate of initiation by hydroxide ions and hydrogen-peroxide. Environ. Sci. Technol. 1982, 16, 676681,  DOI: 10.1021/es00104a009
    36. 36
      Herrmann, H.; Hoffmann, D.; Schaefer, T.; Bräuer, P.; Tilgner, A. Tropospheric Aqueous-Phase Free-Radical Chemistry: Radical Sources, Spectra, Reaction Kinetics and Prediction Tools. ChemPhysChem 2010, 11, 37963822,  DOI: 10.1002/cphc.201000533
    37. 37
      Bianco, A.; Passananti, M.; Brigante, M.; Mailhot, G. Photochemistry of the Cloud Aqueous Phase: A Review. Molecules 2020, 25, 423,  DOI: 10.3390/molecules25020423
    38. 38
      Iriti, M.; Faoro, F. Oxidative stress, the paradigm of ozone toxicity in plants and animals. Water, Air, Soil Pollut. 2007, 187, 285301,  DOI: 10.1007/s11270-007-9517-7
    39. 39
      Jaeglé, L.; Jacob, D. J.; Brune, W. H.; Wennberg, P. O. Chemistry of HOx radicals in the upper troposphere. Atmos. Environ. 2001, 35, 469489,  DOI: 10.1016/S1352-2310(00)00376-9
    40. 40
      Hewitt, C. N.; Kok, G. L. Formation and Occurence of Organic Hydroperoxides in the Troposphere: Laboratory and Field Observations. J. Atmos. Chem. 1991, 12, 181194,  DOI: 10.1007/BF00115779
    41. 41
      Hellpointner, E.; Gäb, S. Detection of methyl, hydroxymethyl and hydroxyethyl hydroperoxides in air and precipitation. Nature 1989, 337, 631634,  DOI: 10.1038/337631a0
    42. 42
      Jacob, D. J. Heterogeneous chemistry and tropospheric ozone. Atmos. Environ. 2000, 34, 21312159,  DOI: 10.1016/S1352-2310(99)00462-8
    43. 43
      Finlayson-Pitts, B. J.; Pitts, J. N., Jr. Atmospheric Chemistry: Fundamental and Experimentals Techniques; John Wiley and Sons: New York, 1986.
    44. 44
      Frost, G.; Vaida, V. Atmospheric Implications of the Photolysis of the Ozone-Water Weakly-Bound Complex. J. Geophys. Res. 1995, 100, 1880318809,  DOI: 10.1029/95JD01940
    45. 45
      Anglada, J. M.; Martins-Costa, M.; Ruiz-López, M. F.; Francisco, J. S. Spectroscopic signatures of ozone at the air–water interface and photochemistry implications. Proc. Natl. Acad. Sci. U. S. A. 2014, 111, 1161811623,  DOI: 10.1073/pnas.1411727111
    46. 46
      Vácha, R.; Slavíček, P.; Mucha, M.; Finlayson-Pitts, B. J.; Jungwirth, P. Adsorption of Atmospherically relevant Gases at the Air/Water Interface: Free Energy Profiles of Aqueous Solvation of N2, O2, O3, H2O, HO2, and H2O2. J. Phys. Chem. A 2004, 108, 1157311579,  DOI: 10.1021/jp046268k
    47. 47
      Vieceli, J.; Roeselova, M.; Potter, N.; Dang, L. X.; Garrett, B. C.; Tobias, D. J. Molecular dynamics simulations of atmospheric oxidants at the air-water interface: Solvation and accommodation of OH and O3. J. Phys. Chem. B 2005, 109, 1587615892,  DOI: 10.1021/jp051361+
    48. 48
      Vaghjiani, G. L.; Ravishankara, A. R. Photodissociation of H2O2 and CH3OOH at 248 nm and 298 K: Quantum yields for OH, O(3P) and H(2S). J. Chem. Phys. 1990, 92, 9961003,  DOI: 10.1063/1.458081
    49. 49
      Vaghjiani, G. L.; Turnipseed, A. A.; Warren, R. F.; Ravishankara, A. R. Photodissociation of H2O2 at 193 and 222 nm: Products and quantum yields. J. Chem. Phys. 1992, 96, 58785886,  DOI: 10.1063/1.462684
    50. 50
      Blitz, M. A.; Heard, D. E.; Pilling, M. J. Wavelength dependent photodissociation of CH3OOH: Quantum yields for CH3O and OH, and measurement of the OH+CH3OOH rate coefficient. J. Photochem. Photobiol., A 2005, 176, 107113,  DOI: 10.1016/j.jphotochem.2005.09.017
    51. 51
      Roehl, C. M.; Marka, Z.; Fry, J. L.; Wennberg, P. O. Near-UV photolysis cross-sections of CH3OOH and HOCH2OOH determined via action spectroscopy. Atmos. Chem. Phys. 2007, 7, 713720,  DOI: 10.5194/acp-7-713-2007
    52. 52
      Matthews, J.; Sinha, A.; Francisco, J. S. The Importance of Weak Absorption Features in Promoting Tropospheric Radical Production. Proc. Natl. Acad. Sci. U. S. A. 2005, 102, 74497452,  DOI: 10.1073/pnas.0502687102
    53. 53
      Herrmann, H. On the photolysis of simple anions and neutral molecules as sources of O/OH, SOx and Cl in aqueous solution. Phys. Chem. Chem. Phys. 2007, 9, 39353964,  DOI: 10.1039/B618565G
    54. 54
      Zellner, R.; Exner, M.; Herrmann, H. Absolute OH quantum yields in the laser photolysis of nitrate, nitrite and dissolved H2O2 at 308 and 351 nm in the temperature-range 278–353 K. J. Atmos. Chem. 1990, 10, 411425,  DOI: 10.1007/BF00115783
    55. 55
      Anastasio, C.; Robles, T. Light absorption by soluble chemical species in Arctic and Antarctic snow. J. Geophys. Res. 2007, 112, 2217,  DOI: 10.1029/2007JD008695
    56. 56
      Kamboures, M. A.; Nizkorodov, S. A.; Gerber, R. B. Ultrafast photochemistry of methyl hydroperoxide on ice particles. Proc. Natl. Acad. Sci. U. S. A. 2010, 107, 66006604,  DOI: 10.1073/pnas.0907922106
    57. 57
      Epstein, S. A.; Shemesh, D.; Tran, V. T.; Nizkorodov, S. A.; Gerber, R. B. Absorption Spectra and Photolysis of Methyl Peroxide in Liquid and Frozen Water. J. Phys. Chem. A 2012, 116, 60686077,  DOI: 10.1021/jp211304v
    58. 58
      Monod, A.; Chevallier, E.; Durand Jolibois, R.; Doussin, J. F.; Picquet-Varrault, B.; Carlier, P. Photooxidation of methylhydroperoxide and ethylhydroperoxide in the aqueous phase under simulated cloud droplet conditions. Atmos. Environ. 2007, 41, 24122426,  DOI: 10.1016/j.atmosenv.2006.10.006
    59. 59
      Warneck, P. The relative importance of various pathways for the oxidation of sulfur dioxide and nitrogen dioxide in sunlit continental fair weather clouds. Phys. Chem. Chem. Phys. 1999, 1, 54715483,  DOI: 10.1039/a906558j
    60. 60
      Martins-Costa, M. T. C.; Ruiz-Lopez, M. F. Highly accurate computation of free energies in complex systems through horsetail QM/MM molecular dynamics combined with free-energy perturbation theory. Theor. Chem. Acc. 2017, 136, 50,  DOI: 10.1007/s00214-017-2078-y
    61. 61
      Martins-Costa, M. T. C.; Ruiz-Lopez, M. F. Reaching Multi-Nanosecond Timescales in Combined QM/MM Molecular Dynamics Simulations through Parallel Horsetail Sampling. J. Comput. Chem. 2017, 38, 659668,  DOI: 10.1002/jcc.24723
    62. 62
      Martins-Costa, M. T. C.; Anglada, J. M.; Francisco, J. S.; Ruiz-Lopez, M. F. Impacts of cloud water droplets on the OH production rate from peroxide photolysis. Phys. Chem. Chem. Phys. 2017, 19, 3162131627,  DOI: 10.1039/C7CP06813A
    63. 63
      Nissenson, P.; Dabdub, D.; Das, R.; Maurino, V.; Minero, C.; Vione, D. Evidence of the water-cage effect on the photolysis of and FeOH2+. Implications of this effect and of H2O2 surface accumulation on photochemistry at the air–water interface of atmospheric droplets. Atmos. Environ. 2010, 44, 48594866,  DOI: 10.1016/j.atmosenv.2010.08.035
    64. 64
      Finlayson-Pitts, B. J.; Wingen, L. M.; Sumner, A. L.; Syomin, D.; Ramazan, K. A. The heterogeneous hydrolysis of NO2 in laboratory systems and in outdoor and indoor atmospheres: An integrated mechanism. Phys. Chem. Chem. Phys. 2003, 5, 223242,  DOI: 10.1039/b208564j
    65. 65
      Crowley, J. N.; Carl, S. A. OH formation in the photoexcitation of NO2 beyond the dissociation threshold in the presence of water vapor. J. Phys. Chem. A 1997, 101, 41784184,  DOI: 10.1021/jp970319e
    66. 66
      Morel, O.; Simonaitis, R.; Heicklen, J. Ultraviolet absorption spectra of HO2NO2, CCl3O2NO2, CCl2FO2NO2, and CH3O2NO2. Chem. Phys. Lett. 1980, 73, 3842,  DOI: 10.1016/0009-2614(80)85197-9
    67. 67
      Murdachaew, G.; Varner, M. E.; Phillips, L. F.; Finlayson-Pitts, B. J.; Gerber, R. B. Nitrogen dioxide at the air-water interface: trapping, absorption, and solvation in the bulk and at the surface. Phys. Chem. Chem. Phys. 2013, 15, 204212,  DOI: 10.1039/C2CP42810E
    68. 68
      Martins-Costa, M. T. C.; Anglada, J. M.; Francisco, J. S.; Ruiz-Lopez, M. F. Theoretical Investigation of the Photoexcited NO2+H2O reaction at the Air-Water Interface and Its Atmospheric Implications. Chem. - Eur. J. 2019, 25, 1389913904,  DOI: 10.1002/chem.201902769
    69. 69
      Warneck, P.; Williams, J. The Atmospheric Chemist’s Companion; Springer: The Netherlands, 2012.
    70. 70
      Dillon, T. J.; Crowley, J. N. Reactive quenching of electronically excited NO2* and NO3* by H2O as potential sources of atmospheric HOx radicals. Atmos. Chem. Phys. 2018, 18, 1400514015,  DOI: 10.5194/acp-18-14005-2018
    71. 71
      Li, S.; Matthews, J.; Sinha, A. Atmospheric hydroxyl radical production from electronically excited NO2 and H2O. Science 2008, 319, 16571660,  DOI: 10.1126/science.1151443
    72. 72
      George, C.; Strekowski, R. S.; Kleffmann, J.; Stemmler, K.; Ammann, M. Photoenhanced uptake of gaseous NO2 on solid-organic compounds: a photochemical source of HONO?. Faraday Discuss. 2005, 130, 195210,  DOI: 10.1039/b417888m
    73. 73
      Stemmler, K.; Ammann, M.; Donders, C.; Kleffmann, J.; George, C. Photosensitized reduction of nitrogen dioxide on humic acid as a source of nitrous acid. Nature 2006, 440, 195198,  DOI: 10.1038/nature04603
    74. 74
      Stemmler, K.; Ndour, M.; Elshorbany, Y.; Kleffmann, J.; D’Anna, B.; George, C.; Bohn, B.; Ammann, M. Light induced conversion of nitrogen dioxide into nitrous acid on submicron humic acid aerosol. Atmos. Chem. Phys. 2007, 7, 42374248,  DOI: 10.5194/acp-7-4237-2007
    75. 75
      Ammann, M.; Rossler, E.; Strekowski, R.; George, C. Nitrogen dioxide multiphase chemistry: Uptake kinetics on aqueous solutions containing phenolic compounds. Phys. Chem. Chem. Phys. 2005, 7, 25132518,  DOI: 10.1039/b501808k
    76. 76
      Cazoir, D.; Brigante, M.; Ammar, R.; D’Anna, B.; George, C. Heterogeneous photochemistry of gaseous NO2 on solid fluoranthene films: A source of gaseous nitrous acid (HONO) in the urban environment. J. Photochem. Photobiol., A 2014, 273, 2328,  DOI: 10.1016/j.jphotochem.2013.07.016
    77. 77
      Liu, J. P.; Li, S.; Mekic, M.; Jiang, H. Y.; Zhou, W. T.; Loisel, G.; Song, W.; Wang, X. M.; Gligorovski, S. Photoenhanced Uptake of NO2 and HONO Formation on Real Urban Grime. Environ. Sci. Technol. Lett. 2019, 6, 413417,  DOI: 10.1021/acs.estlett.9b00308
    78. 78
      Knipping, E. M.; Dabdub, D. Modeling surface-mediated renoxification of the atmosphere via reaction of gaseous nitric oxide with deposited nitric acid. Atmos. Environ. 2002, 36, 57415748,  DOI: 10.1016/S1352-2310(02)00652-0
    79. 79
      Rivera-Figueroa, A. M.; Sumner, A. L.; Finlayson-Pitts, B. J. Laboratory Studies of Potential Mechanisms of Renoxification of Tropospheric Nitric Acid. Environ. Sci. Technol. 2003, 37, 548554,  DOI: 10.1021/es020828g
    80. 80
      Handley, S. R.; Clifford, D.; Donaldson, D. J. Photochemical loss of nitric acid on organic films: A possible recycling mechanism for NOx. Environ. Sci. Technol. 2007, 41, 38983903,  DOI: 10.1021/es062044z
    81. 81
      Finlayson-Pitts, B. J. Reactions at surfaces in the atmosphere: integration of experiments and theory as necessary (but not necessarily sufficient) for predicting the physical chemistry of aerosols. Phys. Chem. Chem. Phys. 2009, 11, 77607779,  DOI: 10.1039/b906540g
    82. 82
      Zhou, X.; Zhang, N.; TerAvest, M.; Tang, D.; Hou, J.; Bertman, S.; Alaghmand, M.; Shepson, P. B.; Carroll, M. A.; Griffith, S.; Dusanter, S.; Stevens, P. S. Nitric acid photolysis on forest canopy surface as a source for tropospheric nitrous acid. Nat. Geosci. 2011, 4, 440443,  DOI: 10.1038/ngeo1164
    83. 83
      Baergen, A. M.; Donaldson, D. J. Photochemical Renoxification of Nitric Acid on Real Urban Grime. Environ. Sci. Technol. 2013, 47, 815820,  DOI: 10.1021/es3037862
    84. 84
      Ye, C.; Zhou, X.; Pu, D.; Stutz, J.; Festa, J.; Spolaor, M.; Tsai, C.; Cantrell, C.; Mauldin, R. L.; Campos, T.; Weinheimer, A.; Hornbrook, R. S.; Apel, E. C.; Guenther, A.; Kaser, L.; Yuan, B.; Karl, T.; Haggerty, J.; Hall, S.; Ullmann, K.; Smith, J. N.; Ortega, J.; Knote, C. Rapid cycling of reactive nitrogen in the marine boundary layer. Nature 2016, 532, 489491,  DOI: 10.1038/nature17195
    85. 85
      Zhou, X.; Gao, H.; He, Y.; Huang, G.; Bertman, S. B.; Civerolo, K.; Schwab, J. Nitric acid photolysis on surfaces in low-NOx environments: Significant atmospheric implications. Geophys. Res. Lett. 2003, 30, 2217  DOI: 10.1029/2003GL018620
    86. 86
      Schuttlefield, J.; Rubasinghege, G.; El-Maazawi, M.; Bone, J.; Grassian, V. H. Photochemistry of adsorbed nitrate. J. Am. Chem. Soc. 2008, 130, 1221012211,  DOI: 10.1021/ja802342m
    87. 87
      Scharko, N. K.; Berke, A. E.; Raff, J. D. Release of Nitrous Acid and Nitrogen Dioxide from Nitrate Photolysis in Acidic Aqueous Solutions. Environ. Sci. Technol. 2014, 48, 1199112001,  DOI: 10.1021/es503088x
    88. 88
      Dubowski, Y.; Colussi, A. J.; Hoffmann, M. R. Nitrogen dioxide release in the 302 nm band photolysis of spray-frozen aqueous nitrate solutions. Atmospheric implications. J. Phys. Chem. A 2001, 105, 49284932,  DOI: 10.1021/jp0042009
    89. 89
      Richards, N. K.; Wingen, L. M.; Callahan, K. M.; Nishino, N.; Kleinman, M. T.; Tobias, D. J.; Finlayson-Pitts, B. J. Nitrate Ion Photolysis in Thin Water Films in the Presence of Bromide Ions. J. Phys. Chem. A 2011, 115, 58105821,  DOI: 10.1021/jp109560j
    90. 90
      Richards-Henderson, N. K.; Callahan, K. M.; Nissenson, P.; Nishino, N.; Tobias, D. J.; Finlayson-Pitts, B. J. Production of gas phase NO2 and halogens from the photolysis of thin water films containing nitrate, chloride and bromide ions at room temperature. Phys. Chem. Chem. Phys. 2013, 15, 1763617646,  DOI: 10.1039/c3cp52956h
    91. 91
      Yu, Y.; Ezell, M. J.; Zelenyuk, A.; Imre, D.; Alexander, L.; Ortega, J.; Thomas, J. L.; Gogna, K.; Tobias, D. J.; D’Anna, B.; Harmon, C. W.; Johnson, S. N.; Finlayson-Pitts, B. J. Nitrate ion photochemistry at interfaces: a new mechanism for oxidation of alpha-pinene. Phys. Chem. Chem. Phys. 2008, 10, 30633071,  DOI: 10.1039/b719495a
    92. 92
      Wingen, L. M.; Moskun, A. C.; Johnson, S. N.; Thomas, J. L.; Roeselova, M.; Tobias, D. J.; Kleinman, M. T.; Finlayson-Pitts, B. J. Enhanced surface photochemistry in chloride-nitrate ion mixtures. Phys. Chem. Chem. Phys. 2008, 10, 56685677,  DOI: 10.1039/b806613b
    93. 93
      Richards-Henderson, N. K.; Anderson, C.; Anastasio, C.; Finlayson-Pitts, B. J. The effect of cations on NO2 production from the photolysis of aqueous thin water films of nitrate salts. Phys. Chem. Chem. Phys. 2015, 17, 3221132218,  DOI: 10.1039/C5CP05325K
    94. 94
      Finlayson-Pitts, B. J. Introductory lecture: atmospheric chemistry in the Anthropocene. Faraday Discuss. 2017, 200, 1158,  DOI: 10.1039/C7FD00161D
    95. 95
      Ye, C. X.; Gao, H. L.; Zhang, N.; Zhou, X. L. Photolysis of Nitric Acid and Nitrate on Natural and Artificial Surfaces. Environ. Sci. Technol. 2016, 50, 35303536,  DOI: 10.1021/acs.est.5b05032
    96. 96
      Ye, C. X.; Zhang, N.; Gao, H. L.; Zhou, X. L. Matrix effect on surface-catalyzed photolysis of nitric acid. Sci. Rep. 2019, 9, 10,  DOI: 10.1038/s41598-018-37973-x
    97. 97
      Nissenson, P.; Knox, C. J. H.; Finlayson-Pitts, B. J.; Phillips, L. F.; Dabdub, D. Enhanced photolysis in aerosols: evidence for important surface effects. Phys. Chem. Chem. Phys. 2006, 8, 47004710,  DOI: 10.1039/b609219e
    98. 98
      Yang, H. S.; Finlayson-Pitts, B. J. Infrared spectroscopic studies of binary solutions of nitric acid and water and ternary solutions of nitric acid, sulfuric acid, and water at room temperature: Evidence for molecular nitric acid at the surface. J. Phys. Chem. A 2001, 105, 18901896,  DOI: 10.1021/jp004224f
    99. 99
      Shamay, E. S.; Buch, V.; Parrinello, M.; Richmond, G. L. At the water’s edge: Nitric acid as a weak acid. J. Am. Chem. Soc. 2007, 129, 1291012911,  DOI: 10.1021/ja074811f
    100. 100
      Bianco, R.; Wang, S. Z.; Hynes, J. T. Theoretical study of the dissociation of nitric acid at a model aqueous surface. J. Phys. Chem. A 2007, 111, 1103311042,  DOI: 10.1021/jp075054a
    101. 101
      Hirokawa, J.; Kato, T.; Mafune, F. Uptake of Gas-Phase Nitrous Acid by pH-Controlled Aqueous Solution Studied by a Wetted Wall Flow Tube. J. Phys. Chem. A 2008, 112, 1214312150,  DOI: 10.1021/jp8051483
    102. 102
      Miller, Y.; Thomas, J. L.; Kemp, D. D.; Finlayson-Pitts, B. J.; Gordon, M. S.; Tobias, D. J.; Gerber, R. B. Structure of Large Nitrate-Water Clusters at Ambient Temperatures: Simulations with Effective Fragment Potentials and Force Fields with Implications for Atmospheric Chemistry. J. Phys. Chem. A 2009, 113, 1280512814,  DOI: 10.1021/jp9070339
    103. 103
      Wang, S.; Bianco, R.; Hynes, J. T. Dissociation of nitric acid at an aqueous surface: Large amplitude motions in the contact ion pair to solvent-separated ion pair conversion. Phys. Chem. Chem. Phys. 2010, 12, 82418249,  DOI: 10.1039/c002299n
    104. 104
      Lewis, T.; Winter, B.; Stern, A. C.; Baer, M. D.; Mundy, C. J.; Tobias, D. J.; Hemminger, J. C. Does Nitric Acid Dissociate at the Aqueous Solution Surface?. J. Phys. Chem. C 2011, 115, 2118321190,  DOI: 10.1021/jp205842w
    105. 105
      Moussa, S. G.; Stern, A. C.; Raff, J. D.; Dilbeck, C. W.; Tobias, D. J.; Finlayson-Pitts, B. J. Experimental and theoretical studies of the interaction of gas phase nitric acid and water with a self-assembled monolayer. Phys. Chem. Chem. Phys. 2013, 15, 448458,  DOI: 10.1039/C2CP42405C
    106. 106
      Kido Soule, M. C.; Blower, P. G.; Richmond, G. L. Nonlinear Vibrational Spectroscopic Studies of the Adsorption and Speciation of Nitric Acid at the Vapor/Acid Solution Interface. J. Phys. Chem. A 2007, 111, 33493357,  DOI: 10.1021/jp0686994
    107. 107
      Stockwell, W. R.; Calvert, J. G. The mechanism of the HO-SO2 reaction. Atmos. Environ. 1983, 17, 22312235,  DOI: 10.1016/0004-6981(83)90220-2
    108. 108
      Jayne, J. T.; Pöschl, U.; Chen, Y.-m.; Dai, D.; Molina, L. T.; Worsnop, D. R.; Kolb, C. E.; Molina, M. J. Pressure and temperature dependence of the gas-phase reaction of SO3 with H2O and the heterogeneous reaction of SO3 with H2O/H2SO4 surfaces. J. Phys. Chem. A 1997, 101, 1000010011,  DOI: 10.1021/jp972549z
    109. 109
      Morokuma, K.; Muguruma, C. Ab initio Molecular orbital Study of the Mechanism of the Gas Phase Reaction SO3 + H2O: Importance of the secons water molecule. J. Am. Chem. Soc. 1994, 116, 1031610317,  DOI: 10.1021/ja00101a068
    110. 110
      Loerting, T.; Liedl, K. R. Toward elimination of discrepancies between theory and experiment: The rate constant of the atmospheric conversion of SO3 to H2SO4. Proc. Natl. Acad. Sci. U. S. A. 2000, 97, 88748878,  DOI: 10.1073/pnas.97.16.8874
    111. 111
      Seinfeld, J. H.; Pandis, S. N. Atmospheric Chemistry and Physics: From Air Pollution to Climate Change, 2nd ed.; Wiley, 2006.
    112. 112
      Mirabel, P.; Clavelin, J. L. Application of nucleation to the study of the gas-phase photooxidation of sulfur-dioxide. J. Chem. Phys. 1979, 70, 57675772,  DOI: 10.1063/1.437405
    113. 113
      Gurjar, B. R.; Jain, A.; Sharma, A.; Agarwal, A.; Gupta, P.; Nagpure, A. S.; Lelieveld, J. Human health risks in megacities due to air pollution. Atmos. Environ. 2010, 44, 46064613,  DOI: 10.1016/j.atmosenv.2010.08.011
    114. 114
      Donaldson, D. J.; Kroll, J. A.; Vaida, V. Gas-phase hydrolysis of triplet SO2: A possible direct route to atmospheric acid formation. Sci. Rep. 2016, 6, 30000,  DOI: 10.1038/srep30000
    115. 115
      Kroll, J. A.; Frandsen, B. N.; Kjaergaard, H. G.; Vaida, V. Atmospheric Hydroxyl Radical Source: Reaction of Triplet SO2 and Water. J. Phys. Chem. A 2018, 122, 44654469,  DOI: 10.1021/acs.jpca.8b03524
    116. 116
      Anglada, J. M.; Martins-Costa, M. T. C.; Francisco, J. S.; Ruiz-López, M. F. Triplet State Promoted Reaction of SO2 with H2O by Competition Between Proton Coupled Electron Transfer (pcet) and Hydrogen Atom Transfer (hat) Processes. Phys. Chem. Chem. Phys. 2019, 21, 97799784,  DOI: 10.1039/C9CP01105F
    117. 117
      Olivella, S.; Anglada, J. M.; Sole, A.; Bofill, J. M. Mechanism of the hydrogen transfer from the OH group to oxygen-centered radicals: Proton-coupled electron-transfer versus radical hydrogen abstraction. Chem. - Eur. J. 2004, 10, 34043410,  DOI: 10.1002/chem.200305714
    118. 118
      Anglada, J. M. Complex mechanism of the gas phase reaction between formic acid and hydroxyl radical. Proton coupled electron transfer versus radical hydrogen abstraction mechanisms. J. Am. Chem. Soc. 2004, 126, 98099820,  DOI: 10.1021/ja0481169
    119. 119
      Anglada, J. M.; Olivella, S.; Sole, A. Hydrogen transfer between sulfuric acid and hydroxyl radical in the gas phase: Competition among hydrogen atom transfer, proton-coupled electron-transfer, and double proton transfer. J. Phys. Chem. A 2006, 110, 19821990,  DOI: 10.1021/jp056155g
    120. 120
      Gonzalez, J.; Anglada, J. M. Gas Phase Reaction of Nitric Acid with Hydroxyl Radical without and with Water. A Theoretical Investigation. J. Phys. Chem. A 2010, 114, 91519162,  DOI: 10.1021/jp102935d
    121. 121
      Anglada, J. M.; Gonzalez, J. Different Catalytic Effects of a Single Water Molecule: The Gas-Phase Reaction of Formic Acid with Hydroxyl Radical in Water Vapor. ChemPhysChem 2009, 10, 30343045,  DOI: 10.1002/cphc.200900387
    122. 122
      Jorgensen, S.; Jensen, C.; Kjaergaard, H. G.; Anglada, J. M. The gas-phase reaction of methane sulfonic acid with the hydroxyl radical without and with water vapor. Phys. Chem. Chem. Phys. 2013, 15, 51405150,  DOI: 10.1039/c3cp44034f
    123. 123
      Anglada, J. M.; Olivella, S.; Solé, A. Unexpected Reactivity of Amidogen Radical in the Gas Phase Degradation of Nitric Acid. J. Am. Chem. Soc. 2014, 136, 68346837,  DOI: 10.1021/ja501967x
    124. 124
      Anglada, J. M.; Olivella, S.; Sole, A. Atmospheric formation of the NO3 radical from gas-phase reaction of HNO3 acid with the NH2 radical: proton-coupled electron-transfer versus hydrogen atom transfer mechanisms. Phys. Chem. Chem. Phys. 2014, 16, 1943719445,  DOI: 10.1039/C4CP02792B
    125. 125
      Anglada, J. M.; Solé, A. The Atmospheric Oxidation of HONO by OH, Cl, and ClO Radicals. J. Phys. Chem. A 2017, 121, 96989707,  DOI: 10.1021/acs.jpca.7b10715
    126. 126
      Anglada, J. M.; Crehuet, R.; Sole, A. The gas phase oxidation of HCOOH by Cl and NH2 radicals. Proton coupled electron transfer versus hydrogen atom transfer. Mol. Phys. 2019, 117, 14301441,  DOI: 10.1080/00268976.2018.1554829
    127. 127
      Anglada, J. M.; Crehuet, R.; Adhikari, S.; Francisco, J. S.; Xia, Y. Reactivity of hydropersulfides toward the hydroxyl radical unraveled: disulfide bond cleavage, hydrogen atom transfer, and proton-coupled electron transfer. Phys. Chem. Chem. Phys. 2018, 20, 47934804,  DOI: 10.1039/C7CP07570G
    128. 128
      Zhong, J.; Zhu, C.; Li, L.; Richmond, G. L.; Francisco, J. S.; Zeng, X. C. Interaction of SO2 with the Surface of a Water Nanodroplet. J. Am. Chem. Soc. 2017, 139, 1716817174,  DOI: 10.1021/jacs.7b09900
    129. 129
      Jayne, J.; Davidovits, P.; Worsnop, D.; Zahniser, M.; Kolb, C. Uptake of sulfur dioxide (g) by aqueous surfaces as a function of pH: the effect of chemical reaction at the interface. J. Phys. Chem. 1990, 94, 60416048,  DOI: 10.1021/j100378a076
    130. 130
      Donaldson, D.; Guest, J. A.; Goh, M. C. Evidence for adsorbed SO2 at the aqueous-air interface. J. Phys. Chem. 1995, 99, 93139315,  DOI: 10.1021/j100023a002
    131. 131
      Yang, H.; Wright, N. J.; Gagnon, A. M.; Gerber, R. B.; Finlayson-Pitts, B. J. An upper limit to the concentration of an SO2 complex at the air–water interface at 298 K: infrared experiments and ab initio calculations. Phys. Chem. Chem. Phys. 2002, 4, 18321838,  DOI: 10.1039/b108907b
    132. 132
      Shamay, E. S.; Johnson, K. E.; Richmond, G. L. Dancing on Water: The Choreography of Sulfur Dioxide Adsorption to Aqueous Surfaces. J. Phys. Chem. C 2011, 115, 2530425314,  DOI: 10.1021/jp2064326
    133. 133
      Martins-Costa, M. T. C.; Anglada, J. M.; Francisco, J. S.; Ruiz-López, M. F. Photochemistry of SO2 at the Air–Water Interface: A Source of OH and HOSO Radicals. J. Am. Chem. Soc. 2018, 140, 1234112344,  DOI: 10.1021/jacs.8b07845
    134. 134
      Ruiz-Lopez, M. F.; Martins-Costa, M. T. C.; Anglada, J. M.; Francisco, J. S. A New Mechanism of Acid Rain Generation from HOSO at the Air-Water Interface. J. Am. Chem. Soc. 2019, 141, 1656416568,  DOI: 10.1021/jacs.9b07912
    135. 135
      Ciuraru, R.; Fine, L.; van Pinxteren, M.; D’Anna, B.; Herrmann, H.; George, C. Photosensitized production of functionalized and unsaturated organic compounds at the air-sea interface. Sci. Rep. 2015, 5, 12741  DOI: 10.1038/srep12741
    136. 136
      Ciuraru, R.; Fine, L.; van Pinxteren, M.; D’Anna, B.; Herrmann, H.; George, C. Unravelling New Processes at Interfaces: Photochemical Isoprene Production at the Sea Surface. Environ. Sci. Technol. 2015, 49, 1319913205,  DOI: 10.1021/acs.est.5b02388
    137. 137
      Brüggemann, M.; Hayeck, N.; Bonnineau, C.; Pesce, S.; Alpert, P. A.; Perrier, S.; Zuth, C.; Hoffmann, T.; Chen, J.; George, C. Interfacial photochemistry of biogenic surfactants: a major source of abiotic volatile organic compounds. Faraday Discuss. 2017, 200, 5974,  DOI: 10.1039/C7FD00022G
    138. 138
      Shrestha, M.; Luo, M.; Li, Y.; Xiang, B.; Xiong, W.; Grassian, V. H. Let there be light: stability of palmitic acid monolayers at the air/salt water interface in the presence and absence of simulated solar light and a photosensitizer. Chem. Sci. 2018, 9, 57165723,  DOI: 10.1039/C8SC01957F
    139. 139
      Rapf, R. J.; Vaida, V. Sunlight as an energetic driver in the synthesis of molecules necessary for life. Phys. Chem. Chem. Phys. 2016, 18, 2006720084,  DOI: 10.1039/C6CP00980H
    140. 140
      Grosjean, D.; Williams, E. L.; Grosjean, E. Atmospheric chemistry of isoprene and of its carbonyl products. Environ. Sci. Technol. 1993, 27, 830840,  DOI: 10.1021/es00042a004
    141. 141
      Kawamura, K.; Tachibana, E.; Okuzawa, K.; Aggarwal, S. G.; Kanaya, Y.; Wang, Z. F. High abundances of water-soluble dicarboxylic acids, ketocarboxylic acids and alpha-dicarbonyls in the mountaintop aerosols over the North China Plain during wheat burning season. Atmos. Chem. Phys. 2013, 13, 82858302,  DOI: 10.5194/acp-13-8285-2013
    142. 142
      Reed Harris, A. E.; Doussin, J.-F.; Carpenter, B. K.; Vaida, V. Gas-Phase Photolysis of Pyruvic Acid: The Effect of Pressure on Reaction Rates and Products. J. Phys. Chem. A 2016, 120, 1012310133,  DOI: 10.1021/acs.jpca.6b09058
    143. 143
      Chang, X. P.; Fang, Q.; Cui, G. L. Mechanistic photodecarboxylation of pyruvic acid: Excited-state proton transfer and three-state intersection. J. Chem. Phys. 2014, 141, 154311  DOI: 10.1063/1.4898085
    144. 144
      Reed Harris, A. E.; Cazaunau, M.; Gratien, A.; Pangui, E.; Doussin, J.-F.; Vaida, V. Atmospheric Simulation Chamber Studies of the Gas-Phase Photolysis of Pyruvic Acid. J. Phys. Chem. A 2017, 121, 83488358,  DOI: 10.1021/acs.jpca.7b05139
    145. 145
      Chiang, Y.; Kresge, A. J.; Pruszynski, P. Keto-enol equilibria in the pyruvic acid system: determination of the keto-enol equilibrium constants of pyruvic acid and pyruvate anion and the acidity constant of pyruvate enol in aqueous solution. J. Am. Chem. Soc. 1992, 114, 31033107,  DOI: 10.1021/ja00034a053
    146. 146
      Guzman, M. I.; Colussi, A. J.; Hoffmann, M. R. Photoinduced oligomerization of aqueous pyruvic acid. J. Phys. Chem. A 2006, 110, 36193626,  DOI: 10.1021/jp056097z
    147. 147
      Griffith, E. C.; Carpenter, B. K.; Shoemaker, R. K.; Vaida, V. Photochemistry of aqueous pyruvic acid. Proc. Natl. Acad. Sci. U. S. A. 2013, 110, 1171411719,  DOI: 10.1073/pnas.1303206110
    148. 148
      Eugene, A. J.; Xia, S.-S.; Guzman, M. I. Negative production of acetoin in the photochemistry of aqueous pyruvic acid. Proc. Natl. Acad. Sci. U. S. A. 2013, 110, E4274E4275,  DOI: 10.1073/pnas.1313991110
    149. 149
      Griffith, E. C.; Carpenter, B. K.; Shoemaker, R. K.; Vaida, V. Reply to Eugene et al.: Photochemistry of aqueous pyruvic acid. Proc. Natl. Acad. Sci. U. S. A. 2013, 110, E4276E4276,  DOI: 10.1073/pnas.1316367110
    150. 150
      Eugene, A. J.; Guzman, M. I. Reactivity of Ketyl and Acetyl Radicals from Direct Solar Actinic Photolysis of Aqueous Pyruvic Acid. J. Phys. Chem. A 2017, 121, 29242935,  DOI: 10.1021/acs.jpca.6b11916
    151. 151
      Vaida, V.; Reed Harris, A. E.; Rapf, R. J.; Perkins, R. J.; Carpenter, B. K. Comment on “Reactivity of Ketyl and Acetyl Radicals from Direct Solar Actinic Photolysis of Aqueous Pyruvic Acid”. J. Phys. Chem. A 2017, 121, 87388740,  DOI: 10.1021/acs.jpca.7b06018
    152. 152
      Eugene, A. J.; Guzman, M. I. Reply to “Comment on ‘Reactivity of Ketyl and Acetyl Radicals from Direct Solar Actinic Photolysis of Aqueous Pyruvic Acid’”. J. Phys. Chem. A 2017, 121, 87418744,  DOI: 10.1021/acs.jpca.7b08273
    153. 153
      Rapf, R. J.; Perkins, R. J.; Carpenter, B. K.; Vaida, V. Mechanistic Description of Photochemical Oligomer Formation from Aqueous Pyruvic Acid. J. Phys. Chem. A 2017, 121, 42724282,  DOI: 10.1021/acs.jpca.7b03310
    154. 154
      Xia, S. S.; Eugene, A. J.; Guzman, M. I. Cross Photoreaction of Glyoxylic and Pyruvic Acids in Model Aqueous Aerosol. J. Phys. Chem. A 2018, 122, 64576466,  DOI: 10.1021/acs.jpca.8b05724
    155. 155
      Leermakers, P. A.; Vesley, G. F. Photochemistry of alpha-keto acids and alpha-keto esters 0.1. Photolysis of pyruvic acid and benzoylformic acid. J. Am. Chem. Soc. 1963, 85, 37763779,  DOI: 10.1021/ja00906a013
    156. 156
      Closs, G. L.; Miller, R. J. Photo-reduction and photodecarboxylation of pyruvic-acid. Applications of CIDNP to mechanistic photochemistry. J. Am. Chem. Soc. 1978, 100, 34833494,  DOI: 10.1021/ja00479a033
    157. 157
      Eugene, A. J.; Guzman, M. I. Production of Singlet Oxygen (1O2) during the Photochemistry of Aqueous Pyruvic Acid: The Effects of pH and Photon Flux under Steady-State O2(aq) Concentration. Environ. Sci. Technol. 2019, 53, 1242512432,  DOI: 10.1021/acs.est.9b03742
    158. 158
      Rapf, R. J.; Dooley, M. R.; Kappes, K.; Perkins, R. J.; Vaida, V. pH Dependence of the Aqueous Photochemistry of α-Keto Acids. J. Phys. Chem. A 2017, 121, 83688379,  DOI: 10.1021/acs.jpca.7b08192
    159. 159
      Rapf, R. J.; Perkins, R. J.; Dooley, M. R.; Kroll, J. A.; Carpenter, B. K.; Vaida, V. Environmental Processing of Lipids Driven by Aqueous Photochemistry of alpha-Keto Acids. ACS Cent. Sci. 2018, 4, 624630,  DOI: 10.1021/acscentsci.8b00124
    160. 160
      Grgić, I.; Nieto-Gligorovski, L. I.; Net, S.; Temime-Roussel, B.; Gligorovski, S.; Wortham, H. Light induced multiphase chemistry of gas-phase ozone on aqueous pyruvic and oxalic acids. Phys. Chem. Chem. Phys. 2010, 12, 698707,  DOI: 10.1039/B914377G
    161. 161
      Gordon, B. P.; Moore, F. G.; Scatena, L. F.; Richmond, G. L. On the Rise: Experimental and Computational Vibrational Sum Frequency Spectroscopy Studies of Pyruvic Acid and Its Surface Active Oligomer Species at the Air-Water Interface. J. Phys. Chem. A 2019, 123, 1060910619,  DOI: 10.1021/acs.jpca.9b08854
    162. 162
      Eugene, A. J.; Pillar, E. A.; Colussi, A. J.; Guzman, M. I. Enhanced Acidity of Acetic and Pyruvic Acids on the Surface of Water. Langmuir 2018, 34, 93079313,  DOI: 10.1021/acs.langmuir.8b01606
    163. 163
      Fu, Y.; Zhang, Y.; Zhang, F.; Chen, J.; Zhu, Z.; Yu, X.-Y. Does interfacial photochemistry play a role in the photolysis of pyruvic acid in water?. Atmos. Environ. 2018, 191, 3645,  DOI: 10.1016/j.atmosenv.2018.07.061
    164. 164
      Fu, H. B.; Ciuraru, R.; Dupart, Y.; Passananti, M.; Tinel, L.; Rossignol, S.; Perrier, S.; Donaldson, D. J.; Chen, J. M.; George, C. Photosensitized Production of Atmospherically Reactive Organic Compounds at the Air/Aqueous Interface. J. Am. Chem. Soc. 2015, 137, 83488351,  DOI: 10.1021/jacs.5b04051
    165. 165
      Tinel, L.; Rossignol, S.; Bianco, A.; Passananti, M.; Perrier, S.; Wang, X.; Brigante, M.; Donaldson, D. J.; George, C. Mechanistic Insights on the Photosensitized Chemistry of a Fatty Acid at the Air/Water Interface. Environ. Sci. Technol. 2016, 50, 1104111048,  DOI: 10.1021/acs.est.6b03165
    166. 166
      Bernard, F.; Ciuraru, R.; Boreave, A.; George, C. Photosensitized Formation of Secondary Organic Aerosols above the Air/Water Interface. Environ. Sci. Technol. 2016, 50, 86788686,  DOI: 10.1021/acs.est.6b03520
    167. 167
      Rossignol, S.; Tinel, L.; Bianco, A.; Passananti, M.; Brigante, M.; Donaldson, D. J.; George, C. Atmospheric photochemistry at a fatty acid-coated air-water interface. Science 2016, 353, 699702,  DOI: 10.1126/science.aaf3617
    168. 168
      Donaldson, D. J.; Vaida, V. The Influence of Organic Films at the Air-Aqueous Boundary on Atmospheric Processes. Chem. Rev. 2006, 106, 14451461,  DOI: 10.1021/cr040367c
    169. 169
      Andreae, M. O.; Crutzen, P. J. Atmospheric aerosols: Biogeochemical sources and role in atmospheric chemistry. Science 1997, 276, 10521058,  DOI: 10.1126/science.276.5315.1052
    170. 170
      Bruggemann, M.; Hayeck, N.; George, C. Interfacial photochemistry at the ocean surface is a global source of organic vapors and aerosols. Nat. Commun. 2018, 9, 3222  DOI: 10.1038/s41467-018-05687-3
    171. 171
      Martins-Costa, M. T. C.; Anglada, J. M.; Francisco, J. S.; Ruiz-Lopez, M. Reactivity of Atmospherically Relevant Small Radicals at the Air–Water Interface. Angew. Chem., Int. Ed. 2012, 51, 54135417,  DOI: 10.1002/anie.201200656
    172. 172
      Wang, H. F.; Borguet, E.; Eisenthal, K. B. Generalized interface polarity scale based on second harmonic spectroscopy. J. Phys. Chem. B 1998, 102, 49274932,  DOI: 10.1021/jp9806563
    173. 173
      Sen, S.; Yamaguchi, S.; Tahara, T. Different Molecules Experience Different Polarities at the Air/Water Interface. Angew. Chem., Int. Ed. 2009, 48, 64396442,  DOI: 10.1002/anie.200901094
    174. 174
      Martins-Costa, M. C.; Ruiz-Lopez, M. Solvation effects on electronic polarization and reactivity indices at the air–water interface: insights from a theoretical study of cyanophenols. Theor. Chem. Acc. 2015, 134, 17  DOI: 10.1007/s00214-014-1609-z
    175. 175
      Hub, J. S.; Caleman, C.; van der Spoel, D. Organic molecules on the surface of water droplets - an energetic perspective. Phys. Chem. Chem. Phys. 2012, 14, 95379545,  DOI: 10.1039/c2cp40483d
    176. 176
      Caleman, C.; Hub, J. S.; van Maaren, P. J.; van der Spoel, D. Atomistic simulation of ion solvation in water explains surface preference of halides. Proc. Natl. Acad. Sci. U. S. A. 2011, 108, 68386842,  DOI: 10.1073/pnas.1017903108
    177. 177
      Jungwirth, P.; Tobias, D. J. Specific ion effects at the air/water interface. Chem. Rev. 2006, 106, 12591281,  DOI: 10.1021/cr0403741
    178. 178
      Petersen, P. B.; Saykally, R. J. On the nature of ions at the liquid water surface. Annu. Rev. Phys. Chem. 2006, 57, 333364,  DOI: 10.1146/annurev.physchem.57.032905.104609
    179. 179
      Jungwirth, P.; Winter, B. Ions at aqueous interfaces: From water surface to hydrated proteins. Annu. Rev. Phys. Chem. 2008, 59, 343366,  DOI: 10.1146/annurev.physchem.59.032607.093749
    180. 180
      Levin, Y.; dos Santos, A. P.; Diehl, A. Ions at the Air-Water Interface: An End to a Hundred-Year-Old Mystery?. Phys. Rev. Lett. 2009, 103, 257802  DOI: 10.1103/PhysRevLett.103.257802
    181. 181
      Otten, D. E.; Shaffer, P. R.; Geissler, P. L.; Saykally, R. J. Elucidating the mechanism of selective ion adsorption to the liquid water surface. Proc. Natl. Acad. Sci. U. S. A. 2012, 109, 701705,  DOI: 10.1073/pnas.1116169109
    182. 182
      Tobias, D. J.; Stern, A. C.; Baer, M. D.; Levin, Y.; Mundy, C. J. Simulation and Theory of Ions at Atmospherically Relevant Aqueous Liquid-Air Interfaces. Annu. Rev. Phys. Chem. 2013, 64, 339359,  DOI: 10.1146/annurev-physchem-040412-110049
    183. 183
      Levin, Y.; dos Santos, A. P. Ions at hydrophobic interfaces. J. Phys.: Condens. Matter 2014, 26, 203101,  DOI: 10.1088/0953-8984/26/20/203101
    184. 184
      Sun, L.; Li, X.; Tu, Y. Q.; Agren, H. Origin of ion selectivity at the air/water interface. Phys. Chem. Chem. Phys. 2015, 17, 43114318,  DOI: 10.1039/C4CP03338H
    185. 185
      Tong, Y.; Zhang, I. Y.; Campen, R. K. Experimentally quantifying anion polarizability at the air/water interface. Nat. Commun. 2018, 9, 1313,  DOI: 10.1038/s41467-018-03598-x
    186. 186
      Wise, P. K.; Ben-Amotz, D. Interfacial Adsorption of Neutral and Ionic Solutes in a Water Droplet. J. Phys. Chem. B 2018, 122, 34473453,  DOI: 10.1021/acs.jpcb.7b10488
    187. 187
      Agmon, N.; Bakker, H. J.; Campen, R. K.; Henchman, R. H.; Pohl, P.; Roke, S.; Thamer, M.; Hassanali, A. Protons and Hydroxide Ions in Aqueous Systems. Chem. Rev. 2016, 116, 76427672,  DOI: 10.1021/acs.chemrev.5b00736
    188. 188
      Saykally, R. J. Air/water interface: Two sides of the acid–base story. Nat. Chem. 2013, 5, 8284,  DOI: 10.1038/nchem.1556
    189. 189
      Martins-Costa, M. T. C.; Anglada, J. M.; Francisco, J. S.; Ruiz-Lopez, M. F. Reactivity of Volatile Organic Compounds at the Surface of a Water Droplet. J. Am. Chem. Soc. 2012, 134, 1182111827,  DOI: 10.1021/ja304971e
    190. 190
      Sitzmann, E. V.; Langan, J.; Eisenthal, K. B. Intermolecular effects on intersystem crossing studied on the picosecond time scale - the solvent polarity effect on the rate of singlet to triplet intersystem crossing of diphenylcarbene. J. Am. Chem. Soc. 1984, 106, 18681869,  DOI: 10.1021/ja00318a069
    191. 191
      Kellmann, A. Intersystem crossing and internal conversion quantum yields of acridine in polar and nonpolar solvents. J. Phys. Chem. 1977, 81, 11951198,  DOI: 10.1021/j100527a014
    192. 192
      Munoz Losa, A.; Fdez. Galvan, I.; Sanchez, M. L.; Martin, M. E.; Aguilar, M. A. Solvent effects on internal conversions and intersystem crossings: The radiationless de-excitation of acrolein in water. J. Phys. Chem. B 2008, 112, 877884,  DOI: 10.1021/jp075706v
    193. 193
      Sanchez-Rodriguez, J. A.; Mohamadzade, A.; Mai, S.; Ashwood, B.; Pollum, M.; Marquetand, P.; Gonzalez, L.; Crespo-Hernandez, C. E.; Ullrich, S. 2-Thiouracil intersystem crossing photodynamics studied by wavelength-dependent photoelectron and transient absorption spectroscopies. Phys. Chem. Chem. Phys. 2017, 19, 1975619766,  DOI: 10.1039/C7CP02258A
    194. 194
      Toniolo, A.; Olsen, S.; Manohar, L.; Martinez, T. Conical intersection dynamics in solution: the chromophore of green fluorescent protein. Faraday Discuss. 2004, 127, 149163,  DOI: 10.1039/B401167H
    195. 195
      Burghardt, I.; Cederbaum, L. S.; Hynes, J. T. Environmental effects on a conical intersection: A model study. Faraday Discuss. 2004, 127, 395411,  DOI: 10.1039/b315071b
    196. 196
      Yamazaki, S.; Kato, S. Locating the lowest free-energy point on conical intersection in polar solvent: Reference interaction site model self-consistent field study of ethylene and CH2NH2+. J. Chem. Phys. 2005, 123, 114510  DOI: 10.1063/1.2038867
    197. 197
      Spezia, R.; Burghardt, I.; Hynes, J. T. Conical intersections in solution: non-equilibrium versus equilibrium solvation. Mol. Phys. 2006, 104, 903914,  DOI: 10.1080/00268970500417895
    198. 198
      Benjamin, I. Reaction Dynamics at Liquid Interfaces. Annu. Rev. Phys. Chem. 2015, 66, 165188,  DOI: 10.1146/annurev-physchem-040214-121428
    199. 199
      Marcus, R. A. On the theory of electron-transfer reactions. VI. Unified treatment for homogeneous and electrode reactions. J. Chem. Phys. 1965, 43, 679701,  DOI: 10.1063/1.1696792
    200. 200
      Marcus, R. Reorganization free energy for electron transfers at liquid-liquid and dielectric semiconductor-liquid interfaces. J. Phys. Chem. 1990, 94, 10501055,  DOI: 10.1021/j100366a005
    201. 201
      Marcus, R. Theory of electron-transfer rates across liquid-liquid interfaces. J. Phys. Chem. 1990, 94, 41524155,  DOI: 10.1021/j100373a051
    202. 202
      Marcus, R. Theory of electron-transfer rates across liquid-liquid interfaces. 2. Relationships and application. J. Phys. Chem. 1991, 95, 20102013,  DOI: 10.1021/j100158a023
    203. 203
      Eugster, N.; Fermín, D. J.; Girault, H. H. Photoinduced electron transfer at liquid/liquid interfaces. Part VI. On the thermodynamic driving force dependence of the phenomenological electron-transfer rate constant. J. Phys. Chem. B 2002, 106, 34283433,  DOI: 10.1021/jp015533o
    204. 204
      Eugster, N.; Fermín, D. J.; Girault, H. H. Photoinduced electron transfer at liquid| liquid interfaces: Dynamics of the heterogeneous photoreduction of quinones by self-assembled porphyrin ion pairs. J. Am. Chem. Soc. 2003, 125, 48624869,  DOI: 10.1021/ja029589n
    205. 205
      McArthur, E. A.; Eisenthal, K. B. Ultrafast excited-state electron transfer at an organic liquid/aqueous interface. J. Am. Chem. Soc. 2006, 128, 10681069,  DOI: 10.1021/ja056518q
    206. 206
      Rao, Y.; Xu, M.; Jockusch, S.; Turro, N. J.; Eisenthal, K. B. Dynamics of excited state electron transfer at a liquid interface using time-resolved sum frequency generation. Chem. Phys. Lett. 2012, 544, 16,  DOI: 10.1016/j.cplett.2012.05.054
    207. 207
      Cooper, J. K.; Benjamin, I. Photoinduced excited state electron transfer at liquid/liquid interfaces. J. Phys. Chem. B 2014, 118, 77037714,  DOI: 10.1021/jp409541u
    208. 208
      Sagar, D. M.; Bain, C. D.; Verlet, J. R. R. Hydrated electrons at the water/air interface. J. Am. Chem. Soc. 2010, 132, 69176919,  DOI: 10.1021/ja101176r
    209. 209
      Siefermann, K. R.; Liu, Y. X.; Lugovoy, E.; Link, O.; Faubel, M.; Buck, U.; Winter, B.; Abel, B. Binding energies, lifetimes and implications of bulk and interface solvated electrons in water. Nat. Chem. 2010, 2, 274279,  DOI: 10.1038/nchem.580
    210. 210
      Gaiduk, A. P.; Pham, T. A.; Govoni, M.; Paesani, F.; Galli, G. Electron affinity of liquid water. Nat. Commun. 2018, 9, 247  DOI: 10.1038/s41467-017-02673-z