ISSN: 2456–5474 RNI No.  UPBIL/2016/68367 VOL.- VIII , ISSUE- I February  - 2023
Innovation The Research Concept
Marine Iodine Chemistry Affects the Climate and Air Quality
Paper Id :  17404   Submission Date :  01/02/2023   Acceptance Date :  19/02/2023   Publication Date :  23/02/2023
This is an open-access research paper/article distributed under the terms of the Creative Commons Attribution 4.0 International, which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited.
For verification of this paper, please visit on http://www.socialresearchfoundation.com/innovation.php#8
Shailly Yadav
Student
Chemistry
M.M. College
Meerut ,Uttar Pradesh, India
Abstract Iodine is a crucial trace element that is engaged in a wide range of significant processes that take place within the Earth system. Iodine has been known to be important for human health for more than a century, and low iodine diets have been related to goitre, cretinism, and newborn deaths. Iodine has been found to have considerable effects on tropospheric photochemistry, which in turn affects climate by lowering ozone's radiative forcing and air quality by lowering extreme O3 concentrations in polluted areas. The majority of the iodine that naturally occurs in the water is in the form of aqueous iodide and iodate. The largest single supply of gaseous iodine to the atmosphere is thought to be the quick reaction of sea-surface iodide with O3. This release of iodine is thought to have increased significantly over the twentieth century, maybe by a factor of three, as a result of rising anthropogenic O3. However, there are significant obstacles to making accurate predictions about how iodine emissions and its biogeochemical cycle may change in the future or have changed in the past due to uncertainties in the marine iodine distribution and world cycle. Here, we summarize recent findings from our team and others that offer a novel viewpoint on the biogeochemical cycle of iodine on a global scale. We specifically contend that aqueous iodide concentrations in the surface ocean may shift significantly as a result of future climate-induced oceanographic changes, with implications for climate and air quality.
Keywords Iodine, Photochemistry, Climate, Earth.
Introduction
Iodine found on Earth's surface is mostly found in the oceans, which hold 70% of the planet's total natural iodine supply. Both biological and non-biological mechanisms are involved in the volatilization of marine iodine and, to a lesser extent, terrestrial iodine, to the atmosphere. Iodine is distinct from the other halogens due to its ease of volatilization, which is significantly larger than that of chlorine and bromine in both inorganic and organic forms. Chameides and Davis, who utilized a photochemical model to infer large affects on tropospheric photochemistry mostly produced by oceanic emissions of methyl iodide, conducted the first investigation into the role of iodine on the atmosphere (CH3I). Nearly three decades later, observational evidence of reactive iodine's extensive effects emerged, confirming that iodine may play a large role in affecting atmospheric photochemistry over the oceans. Since then, a number of observational investigations have established that iodine oxide radicals (IO) are consistently present in the marine troposphere. O3 in the atmosphere reacts with gaseous iodine compounds released from the ocean to form the IO radical after being photolyzed in the atmosphere quickly (between minutes to days). IO is a "smoking gun" that indicates the presence of active iodine chemistry. Important parts of the chemistry of the atmosphere are disturbed by its additional reactions with nitrogen and hydrogen oxides. Atmospheric iodine cycling is now taken into account in chemical transport and air quality models due to its substantial significance in a number of atmospheric processes. These models demonstrate that iodine has a significant effect on the photochemistry of the troposphere, resulting in a 15% reduction in tropospheric O3 (a crucial gas for climate and air quality) globally, a 15% reduction in summertime O3 exposure over Europe, a significant negative feedback mechanism on O3, and a source of aerosols. Additionally, it has recently been demonstrated that iodine is pumped into the stratosphere, where it may contribute in a negligible but important way to the depletion of O3. Determining the effects of iodine on the environment and how they might alter in the future, however, is still fraught with serious uncertainty.
Aim of study In spite of certain advanced discoveries related to iodine chemistry, still many unanswered questions remain there. Certain models have been proposed by different researchers as discussed previously. Certain sources of iodine and its impacts on atmosphere as well as human have been discussed but there may be more of different sources and widespread impacts. There are still significant uncertainties in the magnitude and impacts of the ocean- atmosphere cycling of iodine. However, it would appear from these calculations that iodine aerosol may play an important regional role in determining the aerosol load of the remote tropical ocean both in the present and in the preindustrial. Further observations in these regions would help us to constrain the magnitude of this role. The main objective of this work is to develop a global model to define the iodine chemistry in the atmosphere completely using certain programming with JAVA, C++ etc. and the results may be compared with the experimental data measurements done in the atmosphere.
Review of Literature

Atmospheric Iodine

The chemistry of bromine and chlorine has received a lot of attention in the work on the stratosphere. Iodine, however, is thought to play a major role in the troposphere (Simpson et al., 2015). Iodinated organic molecules from the ocean were once assumed to be the main source of iodine in the atmosphere (Carpenter, 2003). Initially, methyl iodide (CH3I) was thought to be the main source, but it was discovered that other organo-halogens (CH2I2, CH2IBr, and CH2ICl) with higher photolysis rates would result in these compounds having a bigger flux despite their low concentrations (Chuck et al., 2005; Jones et al., 2010; Law and Sturges, 2006). Iodine has recently been discovered to be released into the atmosphere via inorganic halogen compounds (I2 and HOI) (Carpenter et al., 2013).

Iodine species measurement is difficult because of their brief lives and low atmospheric concentrations, and as a result, there is a dearth of observational data. Measurements have been primarily on CH3I and organic chemicals for decades (Saiz-Lopez and von Glasow, 2012). Over the past ten years, there has been a rise in the amount of data available for both organic (such as CH3I and CH2IX, where X=Cl, Br, I) and inorganic (such as IO, OIO, I2) species (Saiz-Lopez et al., 2012b). A global coverage attempt has also been made by satellites, however retrieval at tropospheric concentrations is challenging (Schonhardt et al., 2008).

Main Text

Chemistry of Iodine and Its Destruction of Ozone (O3)

Even if there are some doubts, the primary iodine cycles that result in the degradation of O3 were initially emphasized by Chameides and Davis (1980), and these are reviewed (Atkinson et al., 2000, 2007; Sander et al., 2011).

The sections that follow now discuss iodine reactions. Starting with interactions with the HOx and NOx cycles. Iodine self reactions (IOx) are then taken into account The heterogeneous responses of iodine are then taken into consideration.

Figure 1– Simplified Inorganic Iodine Gas-Phase Chemistry Schematic.

Iodine in the Ocean

The two predominant iodine species in the ocean are aqueous iodide (I, reduced form) and iodate (IO3, oxidized form), with a combined concentration of typically 400–500 nM. Iodate is the preferred form of iodine from a thermodynamic standpoint (apart from in extremely oxygen-depleted waters), and it predominates above the marine mixed layer in oxygenated saltwater. Iodide's presence and distribution in the surface ocean are primarily controlled by processes of physical mixing and advection as well as its physiologically mediated interconversion with iodate. Up to 50% of iodine may be detected as iodide in the euphotic zone due to the reduction of iodate to iodide, which has been associated with primary productivity. Days to weeks have been reported for the iodate to iodide transition to occur in natural saltwater, but the mechanism is poorly understood, and it is unknown if the transformation occurs as an extracellular, cell surface, or dissimilatory reaction. Although iodate interactions with reduced sulfur species generated from cells during senescence and reduction by nitrate reductase enzymes have both been hypothesized, neither method has been proven to be a substantial pathway for conversion.

Once created, kinetically stable iodide progressively oxidizes back to iodate; however, the duration of this process is unknown, with estimates varying from six months to 40 years. In the past, mass balance methods have often been used to determine the rate of iodide oxidation. These estimations' oxidation rates for natural seawater were obtained from recent studies utilizing a radiotracer technique. Rates of chemical oxidation of iodide to iodate in seawater are too slow to account for the observed distribution of iodine species, and the process has been considered to be biologically mediated aside from processes unique to the sea-surface microlayer, such as oxidation by O3. The creation of ocean models of iodine transformations has been argued to be severely constrained by the ambiguity surrounding the rates and mechanisms involved. It has been suggested that bacterial nitrification may be related to the oxidation of iodide to iodate based on studies of the water column. Our latest research for laboratory cultures of ammonium-oxidizing bacteria has validated this. In the ammonium-oxidizing bacteria Nitrosomonas sp. (Nm51) and Nitrosoccocus oceani (Nc10) supplied with I and NH4+, we found a significant increase in iodate concentrations compared to media-only controls, showing that iodide oxidation to IO3 is linked to nitrification, and particularly ammonium oxidation. Nitrosomonas sp. had cell-normalized production rates of 15.69 (4.71) fmol IO3 cell1 d1 while Nitrosococcus oceani had 14.35 (8.55) fmol IO3 cell1 d1. Iodide oxidation to iodate may be common throughout the waters of the planet because nitrification is known to occur throughout the oceanic water column. As hypothesized by others, this mechanism offers a different or additional connectivity between the iodine and nitrogen cycles in the ocean.

Figure-2: Simplified Schematic of Iodine Cycling in the Surface Ocean

HOx Interactions

As atomic oxygen is regenerated after atomic iodine is photolysed, a "null cycle" is completed. Atomic iodine directly combines with O3 to make IO, but this reaction does not directly deplete O3.

I + O3 → IO+O2    (1.1)

IO hν  I+O             (1.2)

O2 + O + M → O3 + M       (1.3)

The transfer of this unusual oxygen is made possible by secondary interactions involving IO and other trace gases, even if the majority of IO is photolysed to regenerate atomic iodine and oxygen (Ox). A catalytic cycle is created if processes can replenish atomic iodine without releasing oxygen, which results in the loss of oxygen and a consequent net drop in O3. The IO reaction with the hydroperoxy radical making HOI, which is photolyzed to replenish atomic iodine without generating an Ox molecule, is a good illustration of this. Additionally, this alters the HO2:OH ratio, which reduces the formation of O3 and alters the lives of species that depend on OH (such as CH4) (Saiz-Lopez et al., 2012b). When IO concentrations are lower ( 1 pmol mol-1), this IO reaction route through HO2 is the main loss pathway (Saiz-Lopez et al., 2012a).

I + O3 → IO + O2  (1.31)

IO + HO2 → HOI + O2        (1.32)

HOI hν  I + OH                (1.33)

net : O3 + HO2 → OH + 2 O2

Iodine Self Reactions and Higher Oxide Chemistry

Higher IO concentrations (>2 pmol mol-1, Saiz-Lopez et al. 2012b) can also cause IO to interact with itself, resulting in the creation of an additional efficient catalytic Ox loss cycle through the I+O2 product channel (Reaction 1.46). The alternative channel that produces I2O2 has two possible outcomes: it can either form a "null cycle" by photolyzing back to OIO+I or thermally decomposing to IO+IO, or it can be physically lost (for example, by deposition or aerosol uptake) leading to a termination step.

I + O3 → IO + O2  (1.43)

IO + IO −M→ OIO + I         (1.44)

OIO hν  I + O2     (1.45)

net : 2 O3 → 3 O2

Iodine has been demonstrated to produce iodine oxide particles, principally through the condensation of higher iodine oxides (Burkholder et al., 2004; G'omez Mart'n et al., 2007; McFiggans et al., 2004). (IxOy). However, observations for specific higher iodine oxide reactions rate constants remain poorly characterized, and discussion regarding potential production pathways as outlined in recent studies continues (Saiz-Lopez et al., 2012b; Sommariva et al., 2012). The proposed pathways essentially include sequential development through addition of OIO, as in the continuation of Reactions 1.53, or sequential addition of O by reaction of I2OX with O3 to generate I2OX+1 (Sommariva et al., 2012). The validity of these approaches has recently been challenged by new research, which contends that the synthesis of I2O4 is the crucial stage in the formation of iodine particles (Gomez Mart'n et al., 2013a).

Because of their potential to produce particles in the marine boundary layer, greater iodine oxide pathways have drawn a lot of attention in the literature. The marine atmosphere is the planet's most pristine environment, and it is extremely susceptible to changes in the types and quantities of these aerosol precursors (Quinn and Bates, 2011). The marine boundary layer may provide a new pathway for the formation of particles, such as a source of cloud condensation nuclei (CCN), which could have important climatic implications.

I + O3 IO + O2

(1.49)

IO + IO → OIO + I

(1.50)

I2O2

(1.51)

IO + OIO I2O3

(1.52)

OIO + OIO I2O4

(1.53)

IxOy aerosol

(1.54)

Studies have different "cut-off" points for iodine's oxidation to higher oxides. Some studies (Breider, 2010) or all processes past I2O2 (Ord'on ez et al., 2012) do not consider O3 loss via I2O2 + O3 oxidation. Due to a lack of knowledge regarding the reversibility of the greater oxide loss route, the choice of which reactions to add is quite unpredictable. One study in particular draws attention to this, showing that the included I2O2 + O3 oxidation route accounts for the majority of the iodine-driven O3 loss (Breider, 2010), thereby treating all higher iodine oxides as being totally stable. Literature pays a lot of attention to the stability of these compounds (Saiz-Lopez et al., 2012b), and some recent work has even explored the effects of iodine photolysis without higher oxides on the world as a whole (Saiz-Lopez et al., 2014).

Uncertainties in Iodine Chemistry

The chemical kinetics of iodine is subject to substantial uncertainty. Iodine's reaction mechanisms have been thoroughly discussed in a number of papers (Carpenter, 2003; Saiz-Lopez et al., 2012b,a; Simpson et al., 2015), and more recent work has used box-modeling to examine uncertainty within known reactions (Sommariva et al., 2012).

A substantial corpus of experimental and theoretical work on iodine has been evaluated by JPL (Sander et al., 2011) and IUPAC (Atkinson et al., 2000, 2006, 2007, 2008) compilations. However, for other processes, the experimental rate data are scant or nonexistent. For instance, it is difficult to quantify the ultimate chemical destiny of higher iodine oxides (I2OX, where X 2), despite the fact that these events have a considerable impact on the lifetime of IOx (Sommariva et al., 2012). I2OX is created through combination processes, however as further addressed by Saiz-Lopez et al., there are still uncertainties regarding their polymerization, photolytic characteristics, and ultimate fate (2012b). Recent research shows the sensitivity of iodine concentrations and, consequently, the atmospheric implications of this higher oxide chemistry while demonstrating that some of the major iodine species are effectively buffered to mechanism uncertainties (Sommariva et al., 2012).

Atmospheric Processing

Dry and Wet Deposition

Henry's law values and formation enthalpies for a certain species must be understood in order to use dry and wet deposition calculation methods. These values for iodine are known for the majority of species (Sander, 1999, 2015) or have been estimated by analogy with related most previously investigated bromine species (Saiz-Lopez et al., 2007b). There are theoretical calculations for enthalpies of production for less well characterized species, including I2OX (Kaltsoyannis and Plane, 2008).

Photolysis

Iodine forms labile species and has a low ionization energy. Short lifespan species result from this, including those with seconds (I2/HOI), minutes (CH2I2), hours (CH2ICl), and days (I2/HOI) (CH3I).

According to the most recent JPL collection, absorption cross-sections are advised for all major iodine species (I2, HOI, IO, OIO, INO, INO2, INO3, CH3I, CH2I2, CH2IBr, and CH2ICl) (Sander et al., 2011). With the exception of CH3I, temperature dependencies are not stated, and quantum yields are still subject to uncertainty. In the most recent JPL/IUPAC compilations, higher iodine oxide cross-sections (for IxOy where x and y are greater than 1) are not reported (Sander et al., 2011). I2O2 and I2O3 "likely" values have been published by a single study using information taken from a study of the IO spectrum (Spietz et al., 2005). According to Bloss et al. (2010), several modeling studies made the assumption that IxOy had cross sections analogous to INO3 based on estimations from Bloss et al. (2001) and G'omez Mart'n et al. (2005) as well as ab initio results (Kaltsoyannis and Plane, 2008).

OIO's cross-section size had been the subject of debate in the past because the species was only observed during the day (Carpenter, 2003). (Stutz et al., 2007). However, thanks to the work of G'omez Mart'n et al., issue is now resolved (2009). Photolysis has been demonstrated to occur between 500 and 650 nm, and JPL provides the suggested cross-section (Sander et al., 2011). The photolysis routes of INO3 were the subject of a second historical dispute that limited knowledge of the efficacy of iodine-initiated O3 destruction (Carpenter, 2003), but this issue has also been settled (Saiz-Lopez et al., 2012b).

Observational Constraints

Organic Iodine Species

Iodine species are difficult to observe because they are quickly destroyed by photolysis or oxidation and have a very short life span. Unsurprisingly, CH3I, the iodine species with the longest lifetime (6 days; Bell et al. 2002), is the most measured. Having said that, the expansion of model comparisons and the creation of worldwide organic halogen emission schemes have been made possible by the improved availability of observations (Ordonez et al., 2012). Using satellite chlorophyll-a as a proxy for concentration, Ordonez and colleagues (2012) used the association between observations of species and their sources (such as biotic, abiotic), and they employed scaling factors to account for greater emission in coastal regions versus open waters. Iodocarbon species (CH3I, CH2IBr, CH2ICl, and CH2I2) and bromocarbon species (CHBr3, CH2Br2, CH2BrCl, CHBrCl2, and CHBr2Cl) were both present in this work. Ziska et al. (2013) recently estimated fluxes of halogen species using the expanding dataset of ocean surface concentrations, although only CH3I was calculated for iodine.


Conclusion Iodine's biogeochemical cycling in its marine and atmospheric compartments has traditionally been investigated separately and by various scientific communities. Because its biogeochemical cycles take place over a wide range of durations, from seconds for some atmospheric processes to up to millennia in the ocean, studying iodine on an Earth system scale is incredibly difficult. Global ocean iodine modeling is in its infancy, despite significant work being made in recent decades on building atmospheric models of iodine cycling. We still have several fundamental questions that severely restrict the representation of iodine biogeochemistry. One of these is how iodide present at the ocean's surface is quantitatively converted into iodine emissions to the atmosphere. Another one of these is the rates and controls of iodine cycling in the ocean, including in oxygen-depleted waters. Weak laboratory and field data on atmospheric HOI, which is thought to be the main carrier of iodine from the ocean to the atmosphere, as well as a dearth of observations of ocean iodine speciation in some regions, particularly in the subtropical gyres and in the Arctic, are major observational gaps that limit our fundamental understanding. At any particular site, the range of seasonal fluctuation in sea-surface iodide concentrations is likewise very weakly limited. We suggest that iodine should be taken into account from this perspective even though the tools to study iodine's role in the Earth system are not yet fully developed. This is because understanding the interactions and feedbacks between biogeochemical and physical processes in the ocean and ozone (and other oxidants) in the atmosphere, which have policy-relevant effects ranging from emissions of ozone precursors to climate change, ocean acidification, and stratospheric ozone, are crucial. The possible effect of ocean acidification on atmospheric iodine emissions over the next two to three decades has been noted here. Continuing ocean acidification, deoxygenation, and decreased productivity, as well as modifications to ocean circulation and vertical mixing, could result in additional substantial alterations over the ensuing decades. The management of tropospheric ozone levels via precursor (nitrogen oxides and hydrocarbons) emission control is affected by these fluctuations in iodine.
References
1. Fuge R, Johnson CC. 2015 Iodine and human health, the role of environmental geochemistry and diet, a review. Appl. Geochem. 63, 282–302. (doi:10.1016/j.apgeochem.2015.09.013) 2. Carpenter LJ. 2003 Iodine in the marine boundary layer. Chem. Rev. 103, 4953–4962. (doi:10.1021/Cr0206465) 3. Redeker KR, Cicerone RJ. 2004 Environmental controls over methyl halide emissions from rice paddies. Glob. Biochem. Cycles 18, 1–26. (doi:10.1029/2003GB002092) 4. Amachi S. 2008 Microbial contribution to global iodine cycling: volatilization, accumulation, reduction, oxidation, and sorption of iodine. Microbes Environ. 23, 269–276. (doi:10.1264/jsme2.ME08548) 5. Saiz-Lopez A, Plane JMC, Baker AR, Carpenter LJ, Von Glasow R, Gómez Martín JC, McFiggans G, Saunders RW. 2012 Atmospheric chemistry of iodine. Chem. Rev. 112, 1773– 1804. (doi:10.1021/cr200029u) 6. Chameides WL, Davis DD. 1980 Iodine: its possible role in tropospheric photochemistry. J. Geophys. Res. Ocean 85, 7383–7398. (doi:10.1029/JC085iC12p07383) 7. Read KA et al. 2008 Extensive halogen-mediated ozone destruction over the tropical Atlantic Ocean. Nature 453, 1232–1235. (doi:10.1038/nature07035) 8. Mahajan AS et al. 2010 Evidence of reactive iodine chemistry in the Arctic boundary layer. J. Geophys. Res. 115, D20303. (doi:10.1029/2009JD013665) 9. Puentedura O, Gil M, Saiz-Lopez A, Hay T, Navarro-Comas M, Gómez-Pelaez A, Cuevas E, Iglesias J, Gomez L. 2012 Iodine monoxide in the north subtropical free troposphere. Atmos. Chem. Phys. 12, 4909–4921. (doi:10.5194/acp-12-4909-2012) 10. Mahajan AS et al. 2012 Latitudinal distribution of reactive iodine in the Eastern Pacific and its link to open ocean sources. Atmos. Chem. Phys. 12, 11 609–11 617. (doi:10.5194/acp-12- 11609-2012) 11. Großmann K et al. 2013 Iodine monoxide in the Western Pacific marine boundary layer. Atmos. Chem. Phys. 13, 3363–3378. (doi:10.5194/acp-13-3363-2013) 12. Dix B, Baidar S, Bresch JF, Hall SR, Schmidt KS, Wang S, Volkamer R. 2013 Detection of iodine monoxide in the tropical free troposphere. Proc. Natl Acad. Sci. USA 110, 2035–2040. (doi:10.1073/pnas.1212386110) 13. Prados-Roman C et al. 2015 Iodine oxide in the global marine boundary layer. Atmos. Chem. Phys. 15, 583–593. (doi:10.5194/acp-15-583-2015) 14. Volkamer R et al. 2015 Aircraft measurements of BrO, IO, glyoxal, NO2, H2O, O2-O2 and aerosol extinction profiles in the tropics: comparison with aircraft-/ship-based in situ and lidar measurements. Atmos. Meas. Tech. 8, 2121–2148. (doi:10.5194/amt-8-2121-2015) 15. Inamdar S et al. 2020 Estimation of reactive inorganic iodine fluxes in the Indian and Southern Ocean marine boundary layer. Atmos. Chem. Phys. 2020, 1–57. (doi:10.5194/acp-2019-1052) 16. Sherwen T et al. 2016 Iodine’s impact on tropospheric oxidants: a global model study in GEOS- Chem. Atmos. Chem. Phys. 16, 1161–1186. (doi:10.5194/acp-16-1161-2016) 17. MacDonald SM, Gómez Martín JC, Chance R, Warriner S, Saiz-Lopez A, Carpenter LJ, Plane JMC. 2014 A laboratory characterisation of inorganic iodine emissions from the sea surface: dependence on oceanic variables and parameterisation for global modelling. Atmos. Chem. Phys. 14, 5841–5852. (doi:10.5194/acp-14-5841-2014) 18. Sherwen T et al. 2016 Global impacts of tropospheric halogens (Cl, Br, I) on oxidants and composition in GEOS-Chem. Atmos. Chem. Phys. 16, 12 239–12 271. (doi:10.5194/ acp-16-12239-2016) 19. Sherwen TM, Evans MJ, Spracklen DV, Carpenter LJ, Chance R, Baker AR, Schmidt JA, Breider TJ. 2016 Global modeling of tropospheric iodine aerosol. Geophys. Res. Lett. 43, 10 012–10 019. (doi:10.1002/2016GL070062) 20. Sherwen T et al. 2017 Effects of halogens on European air-quality. Faraday Discuss. 200, 75–100. (doi:10.1039/C7FD00026J) 21. Carpenter LJ, MacDonald SM, Shaw MD, Kumar R, Saunders RW, Parthipan R, Wilson J, Plane JMC. 2013 Atmospheric iodine levels influenced by sea surface emissions of inorganic iodine. Nat. Geosci. 6, 108–111. (doi:10.1038/ngeo1687) 22. Prados-Roman C, Cuevas CA, Fernandez RP, Kinnison DE, Lamarque J-F, Saiz-Lopez A. 2015 A negative feedback between anthropogenic ozone pollution and enhanced ocean emissions of iodine. Atmos. Chem. Phys. 15, 2215–2224. (doi:10.5194/acp-15-2215-2015) 23. Sherwen T, Evans MJ, Carpenter LJ, Schmidt JA, Mickley LJ. 2017 Halogen chemistry reduces tropospheric O3 radiative forcing. Atmos. Chem. Phys. 17, 1557–1569. (doi:10.5194/acp- 17-1557-2017) 24. Sipilä M et al. 2016 Molecular-scale evidence of aerosol particle formation via sequential addition of HIO3. Nature 537, 532–534. (doi:10.1038/nature19314) 25. Gómez Martín JC, Lewis TR, Blitz MA, Plane JMC, Kumar M, Francisco JS, Saiz-Lopez A. 2020 A gas-to-particle conversion mechanism helps to explain atmospheric particle formation through clustering of iodine oxides. Nat. Commun. 11, 1–14. (doi:10.1038/s41467-020- 18252-8) 26. Koenig TK et al. 2020 Quantitative detection of iodine in the stratosphere. Proc. Natl Acad. Sci. USA 117, 1860–1866. (doi:10.1073/pnas.1916828117) 27. (WMO) WMO. 2018 Scientific assessment of ozone depletion: 2018. Global Ozone Research and Monitoring Project-Report No 58. Geneva, Switzerland: WMO. 28. Jones CE, Hornsby KE, Sommariva R, Dunk RM, von Glasow R, McFiggans G, Carpenter LJ. 2010 Quantifying the contribution of marine organic gases to atmospheric iodine. Geophys. Res. Lett. 37, 1–6. (doi:10.1029/2010GL043990) 29. Tinel L, Adams TJ, Hollis LDJ, Bridger AJM, Chance RJ, Ward MW, Ball SM, Carpenter LJ. 2020 Influence of the sea surface microlayer on oceanic iodine emissions. Environ. Sci. Technol. 54, 13 228–13 237. (doi:10.1021/acs.est.0c02736) 30. Bell N, Hsu L, Jacob DJ, Schultz MG, Blake DR, Butler JH, King DB, Lobert JM, Maier-Reimer E. 2002 Methyl iodide: atmospheric budget and use as a tracer of marine convection in global models. J. Geophys. Res. Atmos. 107, ACH 8-1–ACH 8-12. (doi:10.1029/2001JD001151) 31. Ordóñez C, Lamarque J-F, Tilmes S, Kinnison DE, Atlas EL, Blake DR, Sousa Santos G, Brasseur G, Saiz-Lopez A. 2012 Bromine and iodine chemistry in a global chemistry-climate model: description and evaluation of very short-lived oceanic sources. Atmos. Chem. Phys. 12, 1423–1447. (doi:10.5194/acp-12-1423-2012) 32. Hansell DA, Carlson CA. 2014 Biogeochemistry of marine dissolved organic matter. New York, NY: Academic Press. 33. Hayase S, Yabushita A, Kawasaki M, Enami S, Hoffmann MR, Colussi AJ. 2010 Heterogeneous reaction of gaseous ozone with aqueous iodide in the presence of aqueous organic species. J. Phys. Chem. A 114, 6016–6021. (doi:10.1021/jp101985f) 34. Reeser DI, Donaldson DJ. 2011 Influence of water surface properties on the heterogeneous reaction between I(aq)−1 and O3(g). Atmos. Environ. 45, 6116–6120. (doi:10.1016/J. ATMOSENV.2011.08.042) 35. Hayase S, Yabushita A, Kawasaki M, Enami S, Hoffmann MR, Colussi AJ. 2011 Weak acids enhance halogen activation on atmospheric water’s surfaces. J. Phys. Chem. A 115, 4935–4940. (doi:10.1021/jp2021775) 36. Shaw MD, Carpenter LJ. 2013 Modification of ozone deposition and I2 emissions at the air aqueous interface by dissolved organic carbon of marine origin. Environ. Sci. Technol. 47, 10 947–10 954. 37. Saiz-Lopez A, Fernandez RP, Ordóñez C, Kinnison DE, Martín JCG, Lamarque JF, Tilmes S. 2014 Iodine chemistry in the troposphere and its effect on ozone. Atmos. Chem. Phys. 14, 13 119–13 143. (doi:10.5194/acp-14-13119-2014) 38. Mahajan AS, Plane JMC, Oetjen H, Mendes L, Saunders RW, Saiz-Lopez A, Jones CE, Carpenter LJ, Mcfiggans GB. 2010 Measurement and modelling of tropospheric reactive halogen species over the tropical Atlantic Ocean. Atmos. Chem. Phys. 10, 4611–4624. (doi:10.5194/acp-10-4611-2010) 39. Saiz-Lopez A, Mahajan AS, Salmon RA, Bauguitte SJ-B, Jones AE, Roscoe HK, Plane JMC. 2007 Boundary layer halogens in coastal Antarctica. Science 317, 348–351. (doi:10.1126/ science.1141408) 40. Allan JD et al. 2015 Iodine observed in new particle formation events in the Arctic atmosphere during ACCACIA. Atmos. Chem. Phys. 15, 5599–5609. (doi:10.5194/acp-15-5599-2015) 41. Fernandez RP et al. 2019 Modeling the sources and chemistry of polar tropospheric halogens (Cl, Br, and I) using the CAM-Chem global chemistry-climate model. J. Adv. Model Earth Syst. 11, 2259–2289. (doi:10.1029/2019MS001655) 42. Chang W, Heikes BG, Lee M. 2004 Ozone deposition to the sea surface: chemical enhancement and wind speed dependence. Atmos. Environ. 38, 1053–1059. (doi:10.1016/j.atmosenv. 2003.10.050) 43. Ganzeveld L, Helmig D, Fairall CW, Hare J, Pozzer A. 2009 Atmosphere-ocean ozone exchange: a global modeling study of biogeochemical, atmospheric, and waterside turbulence dependencies. Global Biogeochem. Cycles 23, 1–16. (doi:10.1029/2008GB003301) 44. Hardacre C, Wild O, Emberson L. 2015 An evaluation of ozone dry deposition in global scale chemistry climate models. Atmos. Chem. Phys. 15, 6419–6436. (doi:10.5194/acp-15-6419- 2015) 45. Young PJ et al. 2018 Tropospheric ozone assessment report: assessment of global-scale model performance for global and regional ozone distributions, variability, and trends. Elem. Sci. Anthr. 6, 1–6. (doi:10.1525/elementa.265) 46. Sarwar G, Kang D, Foley K, Schwede D, Gantt B, Mathur R. 2016 Technical note: examining ozone deposition over seawater. Atmos. Environ. 141, 255–262. (doi:10.1016/j. atmosenv.2016.06.072) 47. Luhar AK, Woodhouse MT, Galbally IE. 2018 A revised global ozone dry deposition estimate based on a new two-layer parameterisation for air–sea exchange and the multi-year MACC composition reanalysis. Atmos. Chem. Phys. 18, 4329–4348. (doi:10.5194/acp-18-4329- 2018) 48. Pound RJ, Sherwen T, Helmig D, Carpenter LJ, Evans MJ. 2020 Influences of oceanic ozone deposition on tropospheric photochemistry. Atmos. Chem. Phys. 20, 4227–4239. (doi:10.5194/acp-20-4227-2020) 49. Chance R, Baker AR, Carpenter L, Jickells TD. 2014 The distribution of iodide at the sea surface. Environ. Sci. Process Impacts 16, 1841–1859. (doi:10.1039/c4em00139g) 50. Sherwen T, Chance RJ, Tinel L, Ellis D, Evans MJ, Carpenter LJ. 2019 A machine- learning-based global sea-surface iodide distribution. Earth Syst. Sci. Data 11, 1239–1262. (doi:10.5194/essd-11-1239-2019) 51. Wadley MR, Stevens DP, Jickells TD, Hughes C, Chance R, Hepach H, Tinel L, Carpenter LJ. 2020 A global model for iodine speciation in the upper ocean. Global Biogeochem. Cycles 34, e2019GB006467. (doi:10.1029/2019GB006467) 52. Mahajan AS, Tinel L, Sarkar A, Chance R, Carpenter LJ, Hulswar S, Mali P, Prakash S, Vinayachandran PN. 2019 Understanding iodine chemistry over the northern and equatorial Indian Ocean. J. Geophys. Res. Atmos. 124, 8104–8118. (doi:10.1029/2018JD029063)