Next Article in Journal
BACE1 Function and Inhibition: Implications of Intervention in the Amyloid Pathway of Alzheimer’s Disease Pathology
Previous Article in Journal
Special Issue “Synthesis and Applications of Oligonucleotide Conjugates”
Previous Article in Special Issue
Solar or UVA-Visible Photocatalytic Ozonation of Water Contaminants
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Participation of the Halogens in Photochemical Reactions in Natural and Treated Waters

Department of Environmental Sciences, The Connecticut Agricultural Experiment Station, 123 Huntington St., P.O. Box 1106, New Haven, CT 06504-1106, USA
*
Author to whom correspondence should be addressed.
Molecules 2017, 22(10), 1684; https://doi.org/10.3390/molecules22101684
Submission received: 18 September 2017 / Revised: 1 October 2017 / Accepted: 4 October 2017 / Published: 13 October 2017
(This article belongs to the Special Issue Photon-involving Purification of Water and Air)

Abstract

:
Halide ions are ubiquitous in natural waters and wastewaters. Halogens play an important and complex role in environmental photochemical processes and in reactions taking place during photochemical water treatment. While inert to solar wavelengths, halides can be converted into radical and non-radical reactive halogen species (RHS) by sensitized photolysis and by reactions with secondary reactive oxygen species (ROS) produced through sunlight-initiated reactions in water and atmospheric aerosols, such as hydroxyl radical, ozone, and nitrate radical. In photochemical advanced oxidation processes for water treatment, RHS can be generated by UV photolysis and by reactions of halides with hydroxyl radicals, sulfate radicals, ozone, and other ROS. RHS are reactive toward organic compounds, and some reactions lead to incorporation of halogen into byproducts. Recent studies indicate that halides, or the RHS derived from them, affect the concentrations of photogenerated reactive oxygen species (ROS) and other reactive species; influence the photobleaching of dissolved natural organic matter (DOM); alter the rates and products of pollutant transformations; lead to covalent incorporation of halogen into small natural molecules, DOM, and pollutants; and give rise to certain halogen oxides of concern as water contaminants. The complex and colorful chemistry of halogen in waters will be summarized in detail and the implications of this chemistry for global biogeochemical cycling of halogen, contaminant fate in natural waters, and water purification technologies will be discussed.

1. Introduction

Halide ions are ubiquitous in natural waters. Ordinary levels of halides in seawater are 540 mM chloride, 0.8 mM bromide, and 100–200 nM iodide [1,2]. Halide levels range downward in estuaries and upward in saltier water bodies relative to typical seawater levels. Surface fresh water and groundwater may contain up to 21 mM chloride and 0.05 mM bromide [1], with higher levels in some places. Even though the halides themselves do not absorb light in the solar region, in nature they provide far more than just background electrolytes—they participate in a rich, aqueous-phase chemistry initiated by sunlight that has many implications for dissolved natural organic matter (DOM) processing, fate and toxicity of organic pollutants, and global biogeochemical cycling of the halogens.
Advanced oxidation processes (AOPs) employing solar, visible, or ultraviolet light have been used or are under study for removal of organic pollutants from reclaimable waters, such as industrial wastewater, petrochemical produced waters, municipal wastewater, and landfill leachates, in order to meet agricultural, residential, business, industrial, or drinking water standards. While generalizations are difficult, such waters often contain moderate-to-very-high halide ion concentrations, as well as high concentrations of other photochemically important solutes like carbonate that can impact halogen chemistry [1].
This review aims to summarize the reactions of halides and their daughter products and offer insight into their effects on photochemical transformations taking place in water. Halides can undergo sensitized photolysis and react with many secondary photoproducts to produce reactive halogen species (RHS) that can participate in a variety of reactions with DOM and anthropogenic compounds, including oxidation and incorporation of halogen. These reactions are described and discussed. Extensive tabulations of rate constants for relevant reactions or RHS generation and decay have been collected for the convenience of the reader in Supplementary Section Table S1. Halides, and the RHS derived from them, affect the concentrations of photogenerated reactive oxygen species (ROS) and other reactive species; influence the photobleaching of DOM; alter the rates and products of pollutant transformations; lead to covalent incorporation of halogen into small natural molecules, dissolved natural organic matter, and pollutants; and give rise to certain halogen oxides of concern as water contaminants. The concentrations of halides is an important consideration in water treatment because halides can scavenge desired reactive oxidants and lead to unwanted halogenated byproducts. The identity of the halogen substituent(s) is critical because toxicity ordinarily increases in the order Cl < Br < I for compounds of similar structure [3,4].
Halogen reactions in the atmosphere have been well studied in relation to ozone chemistry [5]. This article will not discuss gas phase reactions or surface reactions in the atmosphere, a topic recently addressed in a comprehensive review [5]; however, it will cover relevant reactions that occur in the liquid phase or at the air-liquid interface of atmospheric aerosols. A number of important reactions that take place on snow, ice, and solid microparticles actually occur on or within a surface liquid layer that is often rich in salts [6]. Compared to bulk natural waters, aerosol liquid phases can reach lower pH, and the evidence supports altered rates and/or unique chemical reactions close to the air-liquid interface.

2. Sources and Speciation of RHS Produced from Halide Ions

Reactive halogen species are generated by sensitized photochemical reactions or by reaction of halides with other oxidants of a photochemical origin. Halogen interconversion reactions are dealt with in detail. Scheme 1 provides an overview.

2.1. Sensitized Photolysis

Halide ions in aqueous solution have absorption edges below ~260 nm and therefore do not photolyze at solar wavelengths. However, recent studies indicate that photo-sensitization by DOM may be an important source of RHS in natural waters [7,8]. Irradiation of DOM with solar light generates a short-lived excited singlet state (1DOM*) that can relax to the ground state or intersystem crosses (ISC) to a much longer-lived excited triplet state (3DOM*). 3DOM* is a mixture of excited triplet states of diverse structures with energies ranging from 94 kJ·mol−1 to above 250 kJ·mol−1 [9]. While the nature of the chromophoric groups of DOM giving rise to triplet states is not known for certain, it has been said that aromatic ketones and other carbonyl-containing groups (e.g., coumarin and chromone moieties) are candidates for production of the high-energy triplet states of DOM [10]. The steady-state concentration of 3DOM* is estimated to be 10−14 to 10−12 M, depending on light intensities, [DOM] and [O2] [10] and, undoubtedly, the nature of DOM in the water parcel.
3DOM* is a known precursor of photochemically-produced reactive oxygen species (ROS) such as singlet oxygen (1O2) and hydroperoxyl/superoxide (HO2/O2−•, pKa = 4.88), and is a suspected precursor of hydroxyl (HO). In addition, 3DOM* also can engage in triplet energy transfer or oxidation reactions with itself and with other solutes. It has been shown that 3DOM* can oxidize or reduce various organic compounds [11], and that model triplet ketone sensitizers with similar reactivity as 3DOM* can oxidize CO32− to CO3−•, NO2 to NO2 [12], etc.
The question arises whether 3DOM* can oxidize halide ions. The standard reduction potential of 3DOM* obtained in different studies of terrestrial and freshwater NOM reference standards is estimated to be “centered near 1.64 V” [10] and about 1.6–1.8 V [8]. The estimated one-electron reduction potentials of the halogens E ° X · / X are 2.59 V (Cl), 2.04 V (Br), and 1.37 V (I) in water [13]. These values are about 0.4–0.5 V lower in polar organic solvents—an important consideration because DOM exists as supramolecular aggregates and colloids, in which the electric field in the vicinity of the chromophoric site may be somewhere in between water and polar organic solvents. It thus appears that bromide and iodide, and possibly chloride, are potentially susceptible to one-electron oxidation by 3DOM*.
Jammoul et al. [7] found that the triplet excited state of benzophenone, which can be regarded as a surrogate for aromatic carbonyl compounds in seawater DOM, can oxidize halide ions to X2−•, Reaction (1):
[ ( C 6 H 5 ) 2 C = O ] 3 * + 2 X hv ( 355   nm ) [ ( C 6 H 5 ) 2 C O ] + X 2
The rate constant for Reaction (1) follows the order, I (~8 × 109) > Br (~3 × 108) > Cl (<1 × 106 M−1 s−1) which is consistent with the order in their reduction potential. The triplet state of anthraquinone derivatives was observed to oxidize bromide and chloride [12,14].
Building on previous theory [15], Loeff et al. [12] modeled reactions sensitized by simple organic compounds according to Scheme 2.
According to this model, halide ion reacts with the triplet excited state (3M) to form a charge-transfer binary exciplex, 3(M --- X), or, at higher halide concentrations, the ternary exciplex, 3(M---X --- X). Both the binary and ternary exciplexes can decay to the ground state (paths a or c) or dissociate to the radical pair (paths b or d). The ternary exciplex has a lower tendency than the binary exciplex to decay to the ground state because it has weaker spin-orbit coupling of the incipient radical. Therefore, the ternary exciplex more favorably dissociates to the radical products, M−• and X2−•. In seawater, the mixed dihalogen radical anion, BrCl−•, is expected to predominate, since bromide is more readily oxidized [16], while chloride is more abundant.
Comparing artificial seawater with ionic strength controls (NaClO4), Parker and Mitch [8] report that 3DOM* contributes to RHS formation, which, in turn, affects the oxidation of certain added organic compounds. Using a series of radical quenching agents, they found a strong linear correlation between the observed rate constant for degradation of the marine algal toxin domoic acid sensitized by a DOM reference standard, and the same rate constant sensitized by bromoacetophenone which generates Br upon photolysis. In support of Scheme 2 for DOM, the researchers found that chloride enhances bromination in samples containing bromide.
In summary, Scheme 2 has been able to rationalize the behavior of simple sensitizer molecules. Even though the scant data available on DOM is consistent with it, it is far from being “established” for DOM and further studies are called for.

2.2. Oxidation of Halide Ions by Secondary Photo-Products

Sunlight directly or indirectly produces OH, ozone (O3), 1O2, HO2/O2−•, and hydrogen peroxide (H2O2) in natural waters. Such ROS are important in many AOPs, as well. Halide ions are susceptible to oxidation by several of these ROS.
One of the most important is HO. Hydroxyl originates from direct photolysis of H2O2, NO3, NO2, DOM, and dissolved iron species, and can also be produced by (dark) Fenton-type reactions of H2O2 catalyzed by redox-switchable transition metal ions, especially Fe. Which of these sources are most important depends on local conditions and is difficult to ascertain in most situations. The exact mechanism of HO generation from DOM has been the subject of debate for many years, without consensus [17,18,19]. Hydroxyl reacts with halides via the adduct HOX−• to form the corresponding halogen and dihalogen radicals:
X + H O    H O X H + , H 2 O X X X 2
Reaction (2) is fast, reversible, and dependent on [X] and [H+] [20]. Reactions with bromide and iodide lie far to the right at any normal environmental pH, while the oxidation of chloride to Cl and Cl2−• is favorable only under acidic conditions and comparatively high halide concentrations. For example, at pH 3, oxidation of chloride is significant whenever [Cl] is much above a few millimolar [21]. However, oxidation of chloride can be important in aerosols, where the pH can be as low as 2. Bromide and iodide are important OH scavengers in seawater [17]. Scavenging of OH does not necessarily protect other solute molecules from oxidation, as the resulting RHS are themselves strong oxidants, albeit more selective (see Section 4).
Ozone is an important component of the troposphere due to the action of sunlight on nitrogen oxides and organic vapors. Ground-level ozone concentrations can be appreciable especially in urban and industrial areas [22,23,24]. The reaction of ozone with halide initially produces hypohalite or hypohalous acid (XO/XOH; pKa,HOCl = 7.82 (0 °C), 7.54 (25 °C); pKa,HOBr = 8.55; pKa,HOI = 10.5), via a transient halo-ozonide intermediate [25]:
X + O 3 X O O O H 2 O X O H + O 2 + O H
The observed rate constants for overall Reaction (3) differ by more than twelve orders of magnitude among the halogens (kCl− < 3 × 10−3 M−1 s−1; kBr− = 258 M−1 s−1; kI− = 1.2 × 109 M−1 s−1) [25,26]. Given normal seawater halide concentrations, the ratio of rates for ozone oxidation of iodide, bromide, and chloride is thus approximately 2300:130:1.
Reaction with O3 is suggested to be a principal source of bromo- and iodo-RHS in seawater [27]. Since HOI can react with Br and Cl to form molecular bromine and chlorine species and regenerate I (see Section 2.4), iodide has been implicated as a catalyst for volatilization of bromine and chlorine from marine aerosol microdroplets [28].
Halides react only slowly with 1O2; second order rate constants are 1 × 103 for Cl (in D2O); < 1 × 106 for Br (in acetone/bromobenzene solution), and 8.7 × 105 M−1 s−1 for I (in water)—too slow to compete with physical quenching of 1O2 by water (2.5 × 105 s−1 [29]). The halides do not react with HO2/O2−• in water at environmentally significant rates. A few other oxidation reactions of halides are important in aerosol systems (Section 2.3).

2.3. Heterogeneous Reactions Leading to RHS

Halides also participate in both dark and actinic heterogeneous chemistry in or on atmospheric particles [30,31]. Atmospheric aerosols broadly encompass polar stratospheric cloud particles of nitric and sulfuric acid hydrates; cloud particles of water ice; soil dusts; marine boundary layer aerosols consisting of sea salts; secondary organic aerosols resulting from oxidation of biogenic compounds in the troposphere; combustion aerosols of fuels and biomass; and inorganic ammonium salt aerosols. Many of these types of particles are relevant here, either because they are aqueous liquids, or because their surfaces are coated with aqueous films that exist due to the high salt levels which attract water.
Reactions of halides in aerosol liquids can be qualitatively and quantitatively different from reactions in terrestrial waters owing to their small size and the significance of gas-particle interfacial phenomena [30]:
(i)
The pH is often more acidic in the bulk liquid phase of aerosols than in terrestrial water bodies. By contrast, the air-liquid interface can be significantly more basic than the bulk aerosol phase; for example, it is known that the pH is 7 at the surface of bulk water at pH 3 [32].
(ii)
The heavier halide ions (Br, I) concentrate at the air-liquid interface. Evidence exists for unique chemical reactions close to the air-liquid interface [33].
(iii)
Particles may become depleted in bromide and iodide with respect to chloride, so that the chemistry can change over time.
(iv)
Reactions may be sensitive to humidity which governs film thickness.
Halide conversion to RHS on atmospheric aerosols is initiated mainly by reactions with HO, O3, nitrate radical (NO3), and N2O5. Their reactions with HO and O3 are given in Reactions (2) and (3) above. Pratt et al. [6] found that Br2 is generated on arctic fresh snow by oxidation of Br by HO formed by photolysis of NO2 or H2O2 within the quasi-brine layer on the snow surfaces. The volatilized Br2 is postulated to get pumped by the wind into the troposphere where it contributes to the episodic depletion of tropospheric ozone during the Arctic springtime.
Nitrate radical, which originates from oxidation of nitrogen dioxide (NO2) by ozone [34], is an important atmospheric free radical, especially at night. It rapidly oxidizes aerosol halides (Reaction 4) [35,36]:
X + N O 3 X + N O 3 k C l = 3.5 × 1 0 8 M 1 s 1 ; k B r = 4 × 1 0 9 M 1 s 1
The nitrate radical interconverts with dinitrogen pentoxide if a suitable surface is available ( N O 3 + N O 2 N 2 O 5 ) [34]. In water N2O5 dissociates to NO3 and NO2+; the latter pairs with a halide to form XNO2, which reacts with a second halide to give X2 [37,38]:
X + N O 2 + X N O 2 X , H + X 2 + H N O 2
For chloride, Reaction (5) occurs only below pH 2 [38].

2.4. Speciation and Interconversion of RHS in Waters

Radical and non-radical RHS (rRHS and nrRHS) undergo well-known species and interconversion reactions in aqueous solutions. Unfortunately, rate constants are not available for iodine speciation in most cases.
Halogen atoms react rapidly and reversibly with halide ion to form the dihalogen radical anion:
X + X X 2
The equilibrium constants are large (on the order of 105 M−1, Supplementary Table S1) and the equilibria lie far to the right in both seawater and freshwater containing typical levels of halides. When I and Br are generated, the mixed dihalogen radical anion ClX−• can form, as chloride is normally predominant. The reverse of Reaction (6) preferentially gives Cl and the other halogen atom because chlorine is the most electronegative of the pair.
Kinetic modeling for seawater containing phenol in which reactions were initiated with OH indicates that the sum of all X2−• concentrations is more than 1000-times greater than the sum of all X concentrations, and that [Br2−•] is about 2.7 times greater than [BrCl−•] [1].
Interconversion of halogen is possible among the rRHS. Some pertinent reactions and their equilibrium constants are given in Reactions (7) and (8):
H O B r + C l B r C l + O H    K e q = 9.5
H O C l + B r B r C l + O H K e q = 330
B r 2 + C l B r C l + B r    K e q = 5.4 × 10 3
B r C l + C l C l 2 + B r    K e q = 2.75 × 10 8
rRHS dimerize or disproportionate to give the nrRHS:
X + X X 2
2 X 2 X 2 + 2 X
X + X 2 X + X 2
Molecular halogen reacts reversibly with halide to form the trihalide ion Reaction (14). For example,
B r C l + C l B r C l 2 K e q = 5.88 M 1
Pertinent to aerosol chemistry, the reactions of Cl2 and Br2 with bromide and iodide are much faster at the air-microdroplet interface than in bulk aqueous solution presumably due to differences in solvation [39]; the same is likely true for chloride but it was not included in the study.
Molecular halogen and trihalide ions hydrolyze to hypohalous acid or the hypohalite ion [40]. Some relevant reactions are:
X C l + H 2 O H O X + H + + C l
X C l + O H H O C l + X
B r C l 2 + H 2 O H O B r + H + + 2 C l K e q = 3 × 1 0 6 M 3
Reactions (15) and (16) lie far to the right and are complete within seconds.
We may consider speciation of nrRHS in different hypothetical waters (Table 1). One represents seawater (540 mM Cl, 0.8 mM Br, 2.3 mM carbonates, pH 8.1) [1], the other a wastewater (141 mM Cl, 0.05 mM Br, 11.5 mM carbonates, pH 7.0). Modeling was performed with 163 reactions using Kintecus V6.01 [41], with an OH generation rate of 1 × 10−9 M−1 s−1, no organic matter present, and a total simulation time of 5 or 60 min. Iodide was not included because many rate constants are unknown.
It can be seen from Table 1 that the principal X2 species is Br2 and the principal X3 species is Br2Cl. Among all the molecular halogen species, between 87% (seawater) and 92% (wastewater) exist as Br2Cl and the remainder mostly as BrCl2. Nevertheless, the vast majority of the nrRHS are HOX/OX species, with HOBr/OBr dominating over HOCl/OCl by more than a factor of 103 (wastewater) or 104 (seawater). While the concentrations of all X2 and X3 stay constant between 5 and 60 min, the concentrations of HOX/OX continue to increase during this interval because there is no sink for them and the starting concentrations of all reactants and the pH are held constant during the simulations. Interestingly, in seawater where chloride is at much higher concentration than in the wastewater, HOCl/OCl increases at a faster rate than HOBr/OBr between 5 and 60 min. This suggests that Br0 species are partially converted to Cl0 species over time. The most likely explanation is a series of reactions that converts HOBr to HOCl, beginning with (and probably rate-limited by) substitution of Br for Cl in HOBr:
H O B r + C l B r C l + O H    k = 44 M 1 s 1
Following Reaction (18) would be, in sequence: (i) Reaction (14) to give BrCl2; (ii) the reverse of Reaction (14) which gives Cl2 rather than Br2 about 5% of the time; and (iii) hydrolysis of Cl2 to HOCl (via Reactions (15) or (16)).
Both HOCl and HOBr readily oxidize iodide [42,43]:
H O C l + I H O I + C l    k = 4.3 × 10 8   M 1 s 1
H O B r + I k 1 I B r + O H k 2 H O I + B r k 1 = 5 × 10 9   M 1 s 1 k 2 = 6 × 10 9   M 1 s 1
Reactions (19) and (20) will therefore generate a lot of HOI regardless of which RHS is initially formed. In water, HOI is slowly converted to iodate (IO3) [44]. Iodate can be an appreciable fraction of total iodine in the sea [45,46].
Since Reactions (14)–(17) are reversible, and X2 species are volatile, atmospheric aerosols can become depleted in bromide and iodide relative to chloride [30].

3. Reactions of RHS

3.1. Photolysis of nrRHS (X2, X3, HOX)

Molecular halogens, X2 and X3, all absorb at wavelengths in the solar UV and into the visible. Photolysis of X2 yields two X atoms [47,48] ( Φ B r 2 , 500 n m = 0.85 [49]; Φ I B r , 500 n m = 0.73 [49]), while photolysis of X3 yields X2−• and X [50,51,52] ( Φ B r 3 , 260 n m = 0.15 [52]). However, molecular halogens are transient and their concentrations so small that photolysis is not likely an important fate mechanism.
The absorption spectra of the hypohalites partially overlap the UV solar emission spectrum, and the molar absorption of OX is greater than that of HOX. Solar UV cleaves the O-X bond homolytically or heterolytically [53,54,55,56,57] to give halogen atoms, halide ions, and a variety of ROS, including hydroxyl radical OH/O−• (pKa, 11.5 [58]), singlet-state atomic oxygen O(1D), ground-state atomic oxygen O(3P), ozone, and hydrogen peroxide (Scheme 3). For OCl, as wavelength increases the quantum yield of homolytic cleavage decreases while that of heterolytic cleavage increases [53,59].
The absorption spectra of HOBr and HOI are red-shifted in the gas phase compared to the aqueous phase [60,61,62]. Thus, the quantum efficiency of HOBr and HOI reactions may be different in aerosols than in bulk solution due to gas-liquid interfacial effects.

3.2. Reactions of RHS with Inorganic Species

Radical and nrRHS exhibit a complex chemistry with inorganic water constituents. Potentially important scavengers include carbonates, hydrogen peroxide, nitrite, and ozone. Hydrogen peroxide is a common component of natural waters owing to biological and photochemical processes. An overview of the reactions is given in Scheme 4. As strong oxidants, RHS may also oxidize metal ions that are present at low concentrations in natural waters, such as FeII, AsIII, and MnII. Reactions of RHS with metal ions are covered elsewhere [63].

3.2.1. Radical RHS

rRHS can be scavenged by carbonate and bicarbonate ions to form carbonate radicals, which, like rRHS, are strong oxidants of organic compounds:
X ( X 2 ) + C O 3 2 ( H C O 3 ) ( 2 ) X ( + H + ) + C O 3            k C l = ( 0.8 5 ) × 1 0 8 M 1 s 1 ; k B r = ( 0.08 2.0 ) × 1 0 6 M 1 s 1
Carbonates also affect RHS levels indirectly by scavenging OH. Kinetic modeling shows that under OH-generating conditions, addition of 2.3 mM carbonates to a solution containing 0.54 M chloride and 0.8 mM bromide steeply reduces rRHS [64]. Conversely, addition of halide ions to carbonate solutions boosts [CO3−•] [1,64] and increases the contribution of CO3−• to transformation of phenol [1].
rRHS species oxidize H2O2 to HO2/O2−• Reactions (22) and (23). Rate constants are 2 × 109 M−1 s−1 for Cl and 4 × 109 M−1 s−1 for Br, but are much smaller for X2−• ( k Cl 2 · = 1.4 × 10 5 M 1 s 1 ; k Br 2 · = 5.0   ×   10 2 M 1 s 1 . The products HO2/O2−• are not very reactive in water toward most organic compounds:
X + H 2 O 2 H O 2 + X + H +
X 2 + H 2 O 2 H O 2 + 2 X + H +
Nitrite reduces X2−• to the halide and nitrite radical, NO2:
X 2 + N O 2 N O 2 + X   ( k Cl 2 · = 2.5 × 1 0 8 M 1 s 1 ;   k Br 2 · = 2 × 1 0 7 M 1 s 1 )
Ozone reacts rapidly with Br to form XO. Data are unavailable for Cl and I:
B r + O 3 B r O + O 2 k = 1.5 × 1 0 8 M 1 s 1
Ozone also reacts with X2−• ( k Cl 2 · = 9 × 1 0 7 M 1 s 1 ). The ClO radical appears to be much less reactive than X and X2−• toward organic compounds [65].
Since in most waters carbonates will be at millimolar concentrations, whereas H2O2, NO2, and O3 will seldom exceed micromolar concentrations, scavenging of the rRHS by carbonates will usually predominate over the others. For their scavenging rates to be equal, [scavenger] = [carbonate]·kcarbonate/kscavenger. Thus, for example, ozone would have to be >~1 mM for it to out-compete 1 mM CO32− for scavenging of Cl2−•.

3.2.2. Non-Radical RHS

Hypohalites can oxidize H2O2 to give the halide and 1O2 (Reactions (26) and (27)). The highest rate constants are observed when the acidic form of one reactant is paired with the basic form of the other—namely, OX + H2O2 or HOX + HO2. The reaction proceeds by nucleophilic attack of H2O2/HO2 on the electrophilic halogen atom of HOX/OX to give initially H-O-O-X and then H-O-O-O-H [66], which decomposes spontaneously to 1O2 [67]. Singlet oxygen is reactive towards many compounds, but physical quenching by water severely limits this reactivity (Yang et al. manuscript in preparation).
H O X + H O 2 H O O X + O H
H 2 O + H O O X X , H + H O O O H H 2 O 1 O 2
Nitrite attacks hypohalites nucleophilically to generate NO2X Reaction (28) [68,69]. Hypochlorites are more reactive than hypobromites. At typically low NO2 concentrations, the principal decomposition pathway of NO2X is reversible dissociation to X and NO2+, followed by hydrolysis of NO2+ to nitrate (Reactions (29) and (30)):
H O X + N O 2 N O 2 X + O H
N O 2 X N O 2 + + X
N O 2 + + O H N O 3 + H +
In acidic aerosols, it is also possible to regain the nrRHS through Reaction (5). Hypohalites react with O3 giving XO2 and eventually XO3 [25,26,70]. Bromate (BrO3) is of concern in drinking water as a carcinogen [71]:
X O H + O 3 X O 2 + O 2 + H +
X O 2 + O 3 X O 3 + O 2 + H +

4. Involvement of Halogen Species in Organic Matter Processing and Transformations of Organic Compounds

Organic matter entering natural waters is processed in part by its own photo-excitation. Photo-excitation of DOM can lead to bleaching, molecular fragmentation, and mineralization (to CO2). DOM can also sensitize the photolysis of dissolved compounds such as pollutants, either through direct reaction between the solute and the 3DOM* (via either triplet energy transfer or electron transfer [10,11]), or through reactions of the solute with secondary photoproducts of DOM such as 1O2, OH, or HO2/O2−•. Halides can affect photoexcitation and photobleaching of DOM, and give rise to RHS that can react with DOM and other organic compounds.

4.1. Impact of Halide Ions on Photoexcitation and Photobleaching of DOM

DOM-sensitized photolysis is an important mechanism for attenuation of organic contaminants in natural waters [10,11]. Increasing halide concentrations up to seawater levels decreased the DOM-sensitized photolysis rate of the female sex hormone, 17β-estradiol, by 90% [72]. About four fifths of the rate decrease was due to a general ionic strength effect, with the remainder to halide-specific effects, especially for bromide. The halide-specific effect was attributed to halide enhancement of DOM photobleaching, which reduced the concentration of chromophoric groups acting as sensitizers [72]. There have been a few other studies on the effects of halides on DOM-sensitized photodegradation, but they have either neglected to include ionic strength controls, or have attributed the observed effects to unrelated causes (see [72]).
Halide ions have been shown to influence the yield and lifetime of 3DOM*, parameters that can be measured by a sorbic acid isomerization probe method [73]. Two studies independently report substantial increases in the steady-state 3DOM* concentration in photolyzed water as the halide concentration increases to seawater levels [74,75]. One study [74] attributed it to a general ionic strength enhancement of 3DOM* lifetime by slowing intra-organic matter electron transfer, which is known to be an important decay pathway for 3DOM*. In the other [75], it was proposed that halides quench 3DOM*, but at the same time increase the rate of singlet-to-triplet intersystem crossing (1DOM* → 3DOM*). Exactly how halide affects 3DOM* lifetime and intersystem crossing rates remain to be resolved.
Photobleaching has important implications for the depth of the photic zone, the processing of DOM itself, and the ability of DOM to photosensitize transformations of other chemical species. The fundamental mechanisms accounting for photobleaching of DOM are not well understood, but halide ions may have an effect on rate. Using either a terrestrial DOM reference standard or an algal exudate representing seawater DOM, Mitch and co-workers [76] found that seawater levels of Cl and Br enhanced DOM photobleaching rates, independent of ionic strength. About 12% of the rate increase was attributed to the formation of RHS (from reaction with OH) that target electron-rich chromophores more selectively than does OH. The rest was unresolved. Studies on environmental grab samples are mixed; some report no consistent effect, others rate enhancement, and still others rate suppression with increasing salinity (see [76,77]).

4.2. Reactions of RHS with Organic Compounds

It is useful to summarize what is known generally about the reactions of RHS with organic compounds. The reactivities of rRHS (X, X2−•) in aqueous solution have not received a great deal of attention, and rate constants are far more plentiful for X2−• than X. rRHS are commonly generated by pulse radiolysis or flash photolysis, and rate constants are calculated from the decay or growth of the UV/visible signal. As mentioned above, the major nrRHS in aqueous solution are the hypohalites, HOX and OX. A sizable literature on these reactions exists due to their importance in chlorine disinfection chemistry [63,78]. To stay relevant to natural waters we will focus here mainly on initial reactions in dilute solutions.

4.2.1. Radical RHS

Three major pathways for reactions of rRHS with organic compounds have been identified: H-atom abstraction from C-H groups Reaction (33); one electron removal from heteroatoms (Z = N, O, or S; Reaction (34); and addition to unsaturated bonds Reaction (35). Rate constants range from 104 to 109 M−1 s−1 [79,80] (http://kinetics.nist.gov/solution/):
X ( X 2 ) + R H ( 2 ) X + H + + R ( k = 10 7 10 9 M 1 s 1   for   Cl ;   10 3 10 6   M 1 s 1   for   Cl 2 ;   ~ 10 4   M 1 s 1   for   Br )
X ( X 2 ) + R n Z : ( 2 ) X + R n Z + ( k =   10 7 10 9   M 1 s 1   for   Cl ;   10 4 10 9   M 1 s 1   for   Br ;   10 6 10 10   M 1 s 1   for   X 2 )
X ( X 2 ) + C = C X C C ( + X ) ( k = 1 0 6 1 0 9 M 1 s 1 )
As expected from their reduction potentials (Section 2.3), reactivity of rRHS generally follows the order: Cl > Br; Cl2−• > Br2−•; and X >> X2−•. For many organic compounds, the rate constants for reaction with Br and Cl rival those with OH. While rate constants for X may exceed X2−• by several orders of magnitude, the steady-state concentrations of the latter can exceed those of the former by several orders. Thus, both X and X2−• must be considered. The reactivity of BrCl−• with organic compounds is essentially unknown. The reduction potential of BrCl−• ( E ° BrCl · / 2 Cl = 1.85 V ) lies in between that of Cl2−• (2.30 V) and Br2−• (1.66 V) [81], suggesting intermediate reactivity.
H-abstraction [65] seems to occur only for aliphatic C-H groups and the rate constant increases with decreasing C-H bond dissociation energy [79,80]). Molecules containing amino, hydroxyl, ether, keto, and sulfide groups preferentially undergo one-electron oxidation, as in Reaction (34).
rRHS add to the double bond alkenes reversibly Reaction (35). Rate constants for Cl2−• and Br2−• increase with electron-donating ability of the alkene substituents [80]. The resulting β-substituted organoradical can react with oxygen (108–109 M−1 s−1) to give β-halo organoperoxyl radicals (X-C-C-OO) that decompose through various pathways to give such products as halohydrins, haloketones or haloaldehydes, ketones/aldehydes, epoxides, and diols.
Reactions of Br, Cl2−•, and Br2−• with simple aromatic compounds depend on substituents [80]. In general, if the substituent bears an electropositive atom with an electron pair, (e.g., -OH, OR, -NH2), reaction proceeds by electron transfer as in Reaction (34); whereas, if the substituent is H, alkyl, -Cl, -NO2, etc., the radical can add to the aromatic ring.
Mitch and co-workers [1] kinetically modeled the transformation of phenol in artificial saline media employing 180 different elementary reactions with known rate constants. They assumed that the process was initiated by photoproduction of OH, and used short times to minimize involvement of nrRHS. In solutions containing just 540 mM Cl and 0.02 mM Br, the contributions to phenol transformation were 74.4% by OH, 21.9% by BrCl−•, 3.3% by Cl2−•, and 0.4% by Br2−•. In simulated seawater that included 2.3 mM carbonates, they were 52.6% by CO3−•, 6% by OH, 21.5% by ClBr−•, 0.1% by Cl2−•, and 19.7% by Br2−•. The lessons to be learned from this with respect to phenol transformation are that, (a) carbonates divert oxidative power away from OH toward CO3−• and X2−•; (b) X seems to play no significant role; and (c) BrCl−•, is an important rRHS.

4.2.2. Non-Radical RHS

Hypohalites react with organic compounds by electrophilic substitution, electrophilic addition, or oxidation. Known apparent second-order rate constants for HOCl reactions with organic compounds range widely from 10−2 to 107 M−1 s−1 [63]. The most reactive functional groups are amino, keto/aldehyde, phenolic, and low-valent sulfur.
The neutral form of amines reacts rapidly with HOCl (primary > secondary >> tertiary) to form chloramines Reaction (36). α-Amino acid groups undergo further decarboxylation and deamination reactions [63]:
R N H 2 + H O C l R N H C l + H 2 O
α-Amino acid groups undergo further decarboxylation and deamination reactions [63]. Aromatic compounds react with HOX by electrophilic substitution of halogen. HOBr and HOI are more reactive than HOCl [42,82]. Ring substituents increase the rate constant in the approximate order, R– < RO– < HO– < (HO–)n > 1. Phenols give o- and p-X substituted products. The phenoxide ion is ~105-times more reactive than the free phenol, and reactivity correlates with electron donor character of the substituents. Ortho- and para-substituted dihydroxyaromatics undergo oxidation to the corresponding quinone [83].
Above pH ~ 5 ketones and aldehydes are halogenated by electrophilic substitution at the α carbon Reaction (37), an important reaction in disinfection chemistry because it leads to hazardous byproducts:
R C ( = O ) C H 3 O H R C ( O ) = C H 2 H O X R C ( = O ) C H 2 C l
Alkenes are attacked electrophilically by the halogen atom of HOX at the least-substituted end of the double bond to form the halohydrin Reaction (38).
C = C + H O X X C C O H
The halohydrin can undergo internal or solvolytic elimination of halide to form, respectively, the epoxide or the α,β-dihydroxy compound.
Hypohalites can also oxidize some functional groups—for example, primary and secondary alcohols to aldehydes and ketones, respectively; aldehydes to carboxylic acids; and sulfides to sulfones, sulfoxides, or sulfonic acids [78]. Some of those reactions may go through halogenated intermediates via electrophilic pathways.

4.3. Photo-Initiated Incorporation of Halogen into Organic Compounds under Natural Conditions

Given that RHS are photochemically produced in natural waters, the question arises as to whether this could lead to incorporation of halogen into natural compounds and water contaminants. Of the nearly five thousand natural organohalogen compounds identified to date [84], only a few have been assigned an abiotic origin. It is important to understand abiotic halogenation pathways because organohalogen compounds are known to play important roles in climate warming, ozone depletion, and toxicity. In addition, abiotic halogenation may play a role in pollutant fate.

4.3.1. Incorporation of Halogen into Simple, Defined Molecules

A number of volatile gases of importance in stratospheric ozone chemistry and climate warming are thought to originate from abiotic reactions in the oceans. Table 2 lists examples of these compounds and their origins. Moreover, sunlight illumination of natural and artificial saline samples has been observed to halogenate specific organic probe compounds. Table 2 also lists these compounds, which include natural compounds, lignin-like model compounds representing NOM, and pollutants. The mechanism of halogenation is unambiguously established in few of these cases.
A noteworthy example of abiotic halogenation of a natural compound with potentially important ramifications for animal and human exposure was recently reported by Kumar et al. [27]. They found that ozonation of seawater samples in the dark stimulated polyhalogenation (X = Br, Cl) of 1,1-dimethyl-2,2′-bipyrrole and 1′-methyl-1,2′-bipyrrole. The polyhalogenated derivatives of these two bipyrroles are widely distributed among oceanic sealife and found in air samples and human breastmilk, but a satisfactory biological explanation for their existence has been elusive. Under laboratory ozonation conditions, bromination predominated over chlorination and there was no detectable iodine incorporation. Kumar et al. [27] proposed that tropospheric ozone exchanging with seawater generates HOCl and HOBr (see Reaction (3)), which halogenates the bipyrroles. The presence of chlorine RHS is unexpected because, at seawater halide levels, bromide reacts ~130 times faster than chloride with O3 (see Section 2.2). However, since rather high ozone concentrations were employed (0.2–2.2 mmol/L), it is possible that bromide became depleted in solution. Alternative explanations for the chlorinated products are conversion of some bromine RHS to chlorine RHS, or nucleophilic displacement of bromide by chloride on intermediate products. Interestingly, in kinetic modeling of hydroxyl radical-generating systems in water containing 540 mM NaCl and 0.02 mM NaBr, 3.3% of phenol degradation was due to reaction with Cl2·, which could only come from oxidation of chloride [1]. The Cl2−• may originate from chloride displacement of hydroxide from the reversibly-formed species, ClOH−•:
C l + O H k = 6.1 x 10 9 M 1 s 1 k = 4.3 x 10 9 M 1 s 1 C l O H C l ; k = 1 x 1 0 5 M 1 s 1 C l 2 + O H
Or it could come from chloride displacement of bromide in BrCl−• Reaction (10; k = 1.1 × 102 M−1 s−1).
Mitch and co-workers [1] investigated halogen incorporation into phenol both in artificial seawater and wastewater concentrate (141 mM NaCl, 0.05 mM NaBr, and 11.5 mM carbonates at pH 7.0) spiked with H2O2 and irradiated with UV for 35 min. Phenol can be considered a model compound for terrestrial DOM and some pollutants. Both chloro- and bromophenols were produced, with bromophenols constituting the majority of products. However, the total yields based on initial phenol were only 0.52% in seawater and 0.03% in wastewater concentrate. The yields were unaffected by eliminating the carbonate component, despite carbonate’s ability to scavenge rRHS Reaction (19). While not established by these results, it is more likely that phenol was halogenated by nrRHS, given the greater reactivity of nrRHS than rRHS toward phenolic compounds (Section 4.2).

4.3.2. Incorporation of Halogen into Bulk DOM

Recent studies [94,95] show convincingly that bulk natural organic matter is photo-halogenated under natural or simulated natural conditions. In the study by Mitch, Dodd, and co-workers [94], organo-Br and organo-I were quantified by solid-phase extraction and silver-form cation exchange filtration to remove the high background of halide ions, followed by non-specific quantification of Br and I by inductively-coupled plasma mass spectrometry (ICP-MS) (the method was insensitive for Cl). In the study by Hao et al. [95], the organohalogen compounds were identified at the formula level by ultra-high resolution electrospray ionization Fourier transform ion cyclotron resonance mass spectrometry (ESI-FT ICR MS).
Native organobromine and organoiodine compounds were found in a variety of seawater samples at concentrations ranging (3.2–6.4) × 10−4 mol Br/mol C and (1.1–3.8) × 10−4 mol I/mol C (or 19–160 nmol Br L−1 and 6–36 nmol I L−1) [94,95], and diminishing with ocean depth [94]. Simulated and natural solar irradiation of terrestrial NOM spiked to artificial and natural seawaters led to halogenation that increased with light fluence [94]. With added NaI, iodination increased at the expense of bromination [94]. Addition of the probe, 3,5-dimethyl-1H-pyrazole (DMP), to irradiated natural seawater samples generated 4-Br and 4-I DMP. Since rRHS oxidize rather than halogenate DMP, this result verifies production of nrRHS in these systems [94] and points to nrRHS as the most likely source of halogenated DOM.
Control experiments indicated that some of the native and photo-generated organobromine and organoiodine compounds are photolabile [94,95]. This indicates that the prevailing levels of organohalogen found in environmental samples likely reflect a balance between formation and decomposition. Experiments in artificial seawater showed that chloride ion stimulates organobromine production [94,95]. This implies that chloride facilitates oxidation of bromide. If 3DOM* is the active oxidant species, one may postulate a mechanism involving the formation and subsequent decay of a ternary exciplex, as previously discussed (Section 2.1):
D O M * 3 + B r + C l 3 [ D O M B r C l ] D O M 3 + B r C l
The ESI-FT ICR-MS study provided a wealth of information on the types of reactions that occur [95]. Most native and photo-produced organohalogen compounds were mono- or di-Br or I molecules (a few contained Cl) of 250–700 Daltons in size, and there was considerable overlap among the natively-present and photo-produced compounds. Some products could be attributed to simple H-for-X substitution or X-addition reactions, but most were the result of multiple processes, often accompanied by photooxidation. Most brominated compounds fell in regions of the van Krevelin diagram indicating derivation from unsaturated aliphatic compounds and saturated fatty acids and carbohydrates, while smaller numbers were derived from polycyclic aromatic and polyphenol moieties. Most iodo compounds appeared to be derived mainly from lignin- or tannic-like structures.
In summary, the results suggest that sunlight-driven reactions of RHS with DOM play an important role in bromine and iodine geochemical cycling in marine environments. It has been estimated [94] that photochemical halogenation of terrestrial DOM in estuaries could generate 30 Gg of organobromine and 70 Gg of organoiodine annually worldwide. Those values do not even include RHS-driven halogenation of marine DOM in the open ocean.

5. Impacts of Halides on Water Treatment Processes

Photo-driven AOPs using oxygen, ozone, and peroxides as bulk oxidants are frequently used to destroy pollutants in drinking water and wastewaters. Semiconductor materials are often used as photocatalysts. While earlier work employed UV light, recent emphasis has been on reactions that are viable in the visible or solar spectral regions to reduce energy costs. Wastewaters such as landfill leachates, production waters, industrial wastewaters, and reverse osmosis brines intended for reuse or safe disposal often contain high levels of halide ions. Application of AOPs for treating salty waters is challenging due, among other things, to the conversion of ROS to RHS, which can affect the efficiency of organic compound degradation and generate unwanted halogenated byproducts. It should be noted that solutions irradiated with UV wavelengths below the absorption edges of halide ions (~260 nm) may generate rRHS from direct photolysis of halides Reaction (41):
X U V ( X , e ) H + ( H 2 O ) X + H ( + O H )
This reaction proceeds through a reversible halogen atom-electron solvent cage complex that produces free halogen and hydrogen atoms upon reaction with water [96,97]. The hydrogen atoms are normally rapidly scavenged by O2 to produce HO2/O2−•. One study [97] reports that Reaction (41) can be driven in the bay of a diode array spectrophotometer (!), and cautions about the potential for analytical interference.

5.1. Hydroxyl Radical-Based AOPs

Numerous AOPs generate OH, including H2O2/UV, Fenton reactions, and heterogeneous photocatalysis (e.g., TiO2), among others. There are several reports of decreased rates and organohalogen formation when OH-based AOPs were conducted in the presence of elevated halide concentrations. A few examples are given. One study involving a Fenton-based AOP to destroy dyes indicated that dye removal decreased while total halogenated organic compounds (AOX) increased when 57 mM Cl was present at pH 2.8 and 1 [98]. Another observed auto-inhibition of 1,2-dibromoethane oxidation in a (dark) Fenton-based AOP due to the generation of bromide ions during the reaction that scavenged OH [99]. Machulek observed that chloride inhibited mineralization of organic compounds by the photo-Fenton reaction, both in a synthetic phenol wastewater and an extract of gasoline [100]. In Fenton reactions at pH 2.8, the impact of chloride scavenging on organic compound transformation rate was noticeable above 0.01 M Cl [21].
Kinetic modeling of phenol oxidation by an OH-generating reaction (H2O2/UV) in phosphate-buffered water containing 0.8 mM NaBr showed that OH accounted for most (74%) of phenol transformation, Br2−• for 24%, and Br for 0.8% [1]. In a synthetic wastewater (141 mM chloride, 0.05 mM bromide, 11.5 mM carbonates) at pH 7, OH was still the most important radical (67%), followed by CO3−• (31%), BrCl−• (2.1%), and Br2−• (0.3%).
It has been proposed that halide ions can be oxidized on the surfaces of semiconductor photocatalysts such as TiO2, either by surface-associated hydroxyl radicals or valence band holes [101,102]. It is known that chloride forms an inner-sphere complex with Ti at the surface [103].

5.2. The UV/Chlorine AOP

The UV photolysis of HOCl has been proposed as an alternative AOP. The photolysis of HOCl/OCl at 254 nm yields OH and Cl (Section 3.1; Scheme 3). Some OH and Cl are scavenged by HOCl/OCl, however, to give ClO, which is a less-reactive radical towards organics [65,104]:
O H / C l + H O C l / O C l C l O
In bromide-containing waters, HOCl/OCl rapidly converts to HOBr/OBr [105]. Photolysis of HOBr/OBr generates OH and Br (Scheme 3) [106]. HOBr is also known to oxidize HOI to IO3. [107,108] When HOCl is in excess, the oxidation of I to IO3 is catalyzed by Br. Obviously, the use of UV/chlorine AOP has the potential to generate high levels of halogenated byproducts.

5.3. Sulfate Radical-Based AOPs

Sulfate radical (SO4−•)-based AOPs are attractive alternatives to hydroxyl radical-based AOPs. The sulfate radical is nearly as reactive toward organic compounds as hydroxyl. UV/peroxydisulfate (S2O82−) has a higher quantum yield of SO4−• from S2O82− (1.4) at 254 nm [109] than does OH from H2O2 (1.0) [110]. The photolysis of peroxymonosulfate at 254 nm generates OH and SO4−• simultaneously [111]. Sulfate radical can also be generated from peroxysulfates by various non-photolytic means, as well. Sulfate radicals convert to hydroxyl radicals in water above pH 7.
Sulfate radical reacts directly and rapidly with the halide ions:
S O 4 + X S O 4 2 + X
While oxidation of Cl by OH is important only in acidic solution, oxidation of Cl by SO4−• Reaction (43) is pH-independent and therefore impacts water treatment over a much broader pH range. Several studies have shown that halide ions can strongly affect SO4−•-based oxidation rates [64] and lead to halogenated byproducts [112].
Experiment and kinetic modeling show that SO4−•-based processes are more strongly affected by halides than are OH-based processes [64]. In the presence of halides and carbonates, the steady-state concentrations of X2−• and CO3−• are much higher than those of SO4−• and OH [80,113]. This, combined with the fact that X2−• and CO3−• are typically less reactive and more selective toward organic compounds than SO4−• and OH, means that oxidation efficiency can be significantly impacted [80,113]. However, the impact depends on molecular structure. Benzoic acid transformation by UV/S2O82− was strongly suppressed in 0.54 M chloride solution compared to phosphate-buffered water, while cyclohex-3-ene carboxylic acid was hardly affected at all [64]. This is because the major rRHS that formed, Cl2−•, is poorly reactive toward benzoic acid, but highly reactive toward the double bond in cyclohex-3-ene carboxylic acid [64]. A similar reason was offered to explain the effects of halides and carbonates on the UV/S2O82− reactivity of different pharmaceuticals in reverse-osmosis brine compared to water—namely, that X2−• and CO3−• were more reactive toward some than others [114].
Sulfate radical AOPs can yield bromate (BrO3) as a final product under some conditions (Scheme 5) [115,116]. Bromate is a suspected human carcinogen with a drinking water standard of 10 μg/L as set by U.S. EPA and the World Health Organization [117]. Both experiment and modeling indicate that HOBr/OBr is a required intermediate in the production of bromate (Scheme 5) [115,116]. The yield of bromate is pH-dependent, as HOBr is about 2 orders of magnitude less reactive than OBr toward Br [116,118]. Organic solutes, DOM, and generated superoxide can scavenge Br, Br2−•, and HOBr. This has the effect of significantly reducing or eliminating bromate formation, as well as recycling bromine back into the bromide form [116].
Halides can also react directly with peroxymonosulfate. The bimolecular rate constant follows the order I > Br > Cl [119]. The evidence is consistent with nucleophilic attack of halide on the peroxy group. The product HOX (except HOCl) is further oxidized by peroxymonosulfate [120]. Oxidation of HOI to IO2 is strongly pH dependent due to speciation effects [121] (Scheme 6.). The reaction of IO2 to IO3 is very fast. The oxidation of HOBr to BrO3 is much slower [120]:
O 3 S O O H + X H O X + S O 4 2 k C l = 2.1 × 1 0 3 M 1 s 1 ; k B r = 0.7 M 1 s 1 ; k I = 1.1 × 1 0 3 M 1 s 1
In the presence of iodide, peroxymonosulfate reactions can also lead to incorporation of iodine into DOM and form byproducts of concern derived from DOM, namely, iodoform (CHI3) and iodoacetic acid [121].

6. Concluding Remarks

Halogen plays an important and colorful role in environmental photochemical processes in natural waters and in chemical reactions taking place during photochemical water purification. In the environment, halides can be oxidized to rRHS (X, X2−•) principally through DOM-sensitized photolysis and reactions with ROS of photochemical origin, especially hydroxyl radical, but also ozone and nitrate radical in atmospheric aerosols. Much more work needs to be done to establish the mechanism and importance of DOM-sensitized photolysis of halide ions with respect to generation of RHS. The nature of the chromophoric groups and the quantum yields of initial RHS products as a function of DOM type need to be established. It is noteworthy that chloride enhances bromination and further work is needed to establish whether the cause is formation of a ternary exciplex like the one in Scheme 2. DOM photosensitization of RHS formation is diminished with photobleaching of DOM, a process that itself is affected by halide ions.
rRHS dimerize or disproportionate to give nrRHS, chiefly the hypohalites (HOX) interconverting with smaller amounts of molecular halogen species, X2 and X3. nrRHS can photolyze to regenerate halogen atoms and produce hydroxyl radical, ozone, or hydrogen peroxide, depending on pH and wavelength. Hypohalites can react with hydrogen peroxide, nitrite, and ozone to give singlet oxygen, nitrate, and oxyhalide anions, respectively—products that oxidize organic compounds less efficiently than the hypohalites. Rate constants are lacking for many speciation reactions and reactions of RHS with other photo-generated species, especially in the case of I. Usually the most important inorganic scavenger of rRHS will be the carbonates.
Halide ions at relatively high concentrations can apparently increase the steady-state concentration of excited triplet-state dissolved organic matter (3DOM*). Exactly how this happens is not entirely clear and deserves further research; one study attributed it to a general ionic strength effect, while another to an increase in the rate of singlet-to-triplet intersystem crossing. Some studies report that high halide concentration accelerates photobleaching of DOM. The mechanism has not been established with confidence. An increase in the steady-state 3DOM* concentration could lead to a loss of chromophoric groups through intra-DOM reactions, or an increase in the generation of RHS that can oxidize DOM.
Reactions of X, X2−•, and HOX with organic compounds in water have been characterized, although rate constants are sparse for iodine compounds and for X relative to X2−•. Rate constants for the mixed species, BrCl−•, are unavailable. Depending on structure, rRHS can react by H-atom abstraction, one-electron oxidation, and addition to double bonds and aromatic rings. The addition reactions can lead to incorporation of halogen into the products. Hypohalites react principally by non-radical electrophilic reactions, including halogen incorporation into amines, ketones, alkenes, and aromatic rings. Incorporation of halogen seems to be more likely with nrRHS than rRHS. Hypohalites can also oxidize alcohols, aldehydes, and sulfides without halogen incorporation. Limited interconversion of halogen within and between rRHS and nrRHS is possible, and can lead to incorporation of all halogens into organic molecules, regardless of which RHS species are generated initially. Some naturally-occurring halogenated compounds are known to form abiotically, including ozone-depleting gases. Many of these are thought to have a photolytic origin. Evidence has appeared for abiotic incorporation of halogen into water contaminants and model compounds representing natural organic matter initiated by photochemical processes. Evidence has also appeared for the photochemical incorporation of halogen into natural compounds creating products toxic to oceanic sealife. More examples of natural abiotic incorporation of halogen atoms into natural compounds and water contaminants are likely to appear in the future.
Recent studies show that DOM from both oceanic and terrestrial sources is halogenated under simulated or natural conditions of irradiation. The mechanisms of halogen incorporation have not been identified precisely. Likewise, the scope of such reactions and the effects of water chemistry are as yet poorly characterized. Complicating matters is the finding that photo generation and decomposition of halogenated DOM seem to be taking place simultaneously. The results so far indicate that sunlight-driven oxidation by RHS and halogenation reactions may play important roles in halogen geochemical cycling in marine and estuarine environments, especially in regard to bromine and iodine. It is possible that the presence of natural halogenated compounds has contributed to the evolution of enzymatic dehalogenation pathways of halogenated molecules.
Oxidation and halogen incorporation are of demonstrated importance in AOPs for salty water treatment that use light. Rates can be markedly slowed (or sometimes accelerated) and undesirable byproducts can be formed. There is much to be learned about the influence of halide salts. It is noteworthy that halides can be photolyzed by UV below 260 nm. Whether halides are oxidized on photocatalyst surfaces is largely an open question. The rate constants of RHS with many organic compounds are unknown, but important for evaluating the contribution of RHS to organic compound degradation. The yields of halogenated byproducts depend strongly on the parent compound and the solution conditions. An insufficient database exists to predict precisely where rRHS will attack in a complex molecule. The halogenated byproducts may be of greater toxicity than the original contaminants. Although many studies reported the appearance of halogenated products, quantitative yields are often not reported [101,104,122]. For these reasons the impact of halides on toxicity of the treated waters should routinely accompany investigations, and such information should be used to judge the suitability of the AOP.

Supplementary Materials

The following are available online. Table S1: Rate constants for relevant reactions of halides and reactive halogen species.

Acknowledgments

The authors thank the Chinese International Postdoctoral Exchange Fellowship Program (No. 20160074) for support for Y.Y.

Author Contributions

Both Y.Y. and J.J.P. contributed to writing the manuscript.

Conflicts of Interest

The authors declare no conflict of interest. The founding sponsors had no role in the writing of the manuscript, nor in the decision to publish it.

References

  1. Grebel, J.E.; Pignatello, J.J.; Mitch, W.A. Effect of halide ions and carbonates on organic contaminant degradation by hydroxyl radical-based advanced oxidation processes in saline waters. Environ. Sci. Technol. 2010, 44, 6822–6828. [Google Scholar] [CrossRef] [PubMed]
  2. Luther, G.W., III; Swartz, C.B.; Ullman, W.J. Direct determination of iodide in seawater by cathodic stripping square wave voltammetry. Anal. Chem. 1988, 60, 1721–1724. [Google Scholar] [CrossRef]
  3. Plewa, M.J.; Muellner, M.G.; Richardson, S.D.; Fasano, F.; Buettner, K.M.; Woo, Y.-T.; McKague, A.B.; Wagner, E.D. Occurrence, synthesis, and mammalian cell cytotoxicity and genotoxicity of haloacetamides: An emerging class of nitrogenous drinking water disinfection byproducts. Environ. Sci. Technol. 2007, 42, 955–961. [Google Scholar] [CrossRef]
  4. Richardson, S.D.; Plewa, M.J.; Wagner, E.D.; Schoeny, R.; DeMarini, D.M. Occurrence, genotoxicity, and carcinogenicity of regulated and emerging disinfection by-products in drinking water: A review and roadmap for research. Mutat. Res. Rev. Mutat. Res. 2007, 636, 178–242. [Google Scholar] [CrossRef] [PubMed]
  5. Simpson, W.R.; von Glasow, R.; Riedel, K.; Anderson, P.; Ariya, P.; Bottenheim, J.; Burrows, J.; Carpenter, L.J.; Frieß, U.; Goodsite, M.E.; et al. Halogens and their role in polar boundary-layer ozone depletion. Atmos. Chem. Phys. 2007, 7, 4375–4418. [Google Scholar] [CrossRef] [Green Version]
  6. Pratt, K.A.; Custard, K.D.; Shepson, P.B.; Douglas, T.A.; Pohler, D.; General, S.; Zielcke, J.; Simpson, W.R.; Platt, U.; Tanner, D.J.; et al. Photochemical production of molecular bromine in Arctic surface snowpacks. Nat. Geosci. 2013, 6, 351–356. [Google Scholar] [CrossRef]
  7. Jammoul, A.; Dumas, S.; D’Anna, B.; George, C. Photoinduced oxidation of sea salt halides by aromatic ketones: A source of halogenated radicals. Atmos. Chem. Phys. 2009, 9, 4229–4237. [Google Scholar] [CrossRef]
  8. Parker, K.M.; Mitch, W.A. Halogen radicals contribute to photooxidation in coastal and estuarine waters. Proc. Natl. Acad. Sci. USA 2016, 113, 5868–5873. [Google Scholar] [CrossRef] [PubMed]
  9. Zepp, R.G.; Schlotzhauer, P.F.; Sink, R.M. Photosensitized transformations involving electronic energy transfer in natural waters: Role of humic substances. Environ. Sci. Technol. 1985, 19, 74–81. [Google Scholar] [CrossRef]
  10. McNeill, K.; Canonica, S. Triplet state dissolved organic matter in aquatic photochemistry: Reaction mechanisms, substrate scope, and photophysical properties. Environ. Sci. Process Impacts 2016, 18, 1381–1399. [Google Scholar] [CrossRef] [PubMed]
  11. Canonica, S. Oxidation of aquatic organic contaminants induced by excited triplet states. Chim. Int. J. Chem. 2007, 61, 641–644. [Google Scholar] [CrossRef]
  12. Loeff, I.; Rabani, J.; Treinin, A.; Linschitz, H. Charge transfer and reactivity of nπ* and ππ* organic triplets, including anthraquinonesulfonates, in interactions with inorganic anions: A comparative study based on classical Marcus theory. J. Am. Chem. Soc. 1993, 115, 8933–8942. [Google Scholar] [CrossRef]
  13. Isse, A.A.; Lin, C.Y.; Coote, M.L.; Gennaro, A. Estimation of standard reduction potentials of halogen atoms and alkyl halides. J. Phys. Chem. B 2011, 115, 678–684. [Google Scholar] [CrossRef] [PubMed]
  14. De Laurentiis, E.; Minella, M.; Maurino, V.; Minero, C.; Mailhot, G.; Sarakha, M.; Brigante, M.; Vione, D. Assessing the occurrence of the dibromide radical (Br2) in natural waters: Measures of triplet-sensitised formation, reactivity, and modelling. Sci. Total Environ. 2012, 439, 299–306. [Google Scholar] [CrossRef] [PubMed]
  15. Hurley, J.K.; Linschitz, H.; Treinin, A. Interaction of halide and pseudohalide ions with triplet benzophenone-4-carboxylate: Kinetics and radical yields. J. Phys. Chem. 1988, 92, 5151–5159. [Google Scholar] [CrossRef]
  16. Wardman, P. Reduction potentials of one-electron couples involving free radicals in aqueous solution. J. Phys. Chem. Ref. Data 1989, 18, 1637–1755. [Google Scholar] [CrossRef]
  17. Mopper, K.; Zhou, X. Hydroxyl radical photoproduction in the sea and its potential impact on marine processes. Science 1990, 250, 661–664. [Google Scholar] [CrossRef] [PubMed]
  18. Vaughan, P.P.; Blough, N.V. Photochemical formation of hydroxyl radical by constituents of natural waters. Environ. Sci. Technol. 1998, 32, 2947–2953. [Google Scholar] [CrossRef]
  19. Sun, L.; Qian, J.; Blough, N.V.; Mopper, K. Insights into the photoproduction sites of hydroxyl radicals by dissolved organic matter in natural waters. Environ. Sci. Technol. Lett. 2015, 2, 352–356. [Google Scholar] [CrossRef]
  20. Jayson, G.G.; Parsons, B.J.; Swallow, A.J. Some Simple, Highly reactive, inorganic chlorine derivatives in aqueous solution. J. Chem. Soc., Faraday Trans. I 1973, 69, 1597–1607. [Google Scholar] [CrossRef]
  21. Pignatello, J. Dark and photoassisted Fe3+-catalyzed degradation of chlorophenoxy herbicides by hydrogen peroxide. Environ. Sci. Technol. 1992, 26, 944–951. [Google Scholar] [CrossRef]
  22. Ryerson, T.; Trainer, M.; Angevine, W.; Brock, C.; Dissly, R.; Fehsenfeld, F.; Frost, G.; Goldan, P.; Holloway, J.; Hübler, G. Effect of petrochemical industrial emissions of reactive alkenes and NOx on tropospheric ozone formation in Houston, Texas. J. Geophys. Res. Atmos. 2003, 108, 4249. [Google Scholar] [CrossRef]
  23. Zhang, R.; Lei, W.; Tie, X.; Hess, P. Industrial emissions cause extreme urban ozone diurnal variability. Proc. Natl. Acad. Sci. USA 2004, 101, 6346–6350. [Google Scholar] [CrossRef] [PubMed]
  24. Zheng, J.; Shao, M.; Che, W.; Zhang, L.; Zhong, L.; Zhang, Y.; Streets, D. Speciated VOC emission inventory and spatial patterns of ozone formation potential in the Pearl River Delta, China. Environ. Sci. Technol. 2009, 43, 8580–8586. [Google Scholar] [CrossRef] [PubMed]
  25. Liu, Q.; Schurter, L.M.; Muller, C.E.; Aloisio, S.; Francisco, J.S.; Margerum, D.W. Kinetics and mechanisms of aqueous ozone reactions with bromide, sulfite, hydrogen sulfite, iodide, and nitrite ions. Inorg. Chem. 2001, 40, 4436–4442. [Google Scholar] [CrossRef] [PubMed]
  26. Haag, W.R.; Hoigne, J. Ozonation of bromide-containing waters: Kinetics of formation of hypobromous acid and bromate. Environ. Sci. Technol. 1983, 17, 261–267. [Google Scholar] [CrossRef]
  27. Kumar, A.; Borgen, M.; Aluwihare, L.I.; Fenical, W. Ozone-activated halogenation of mono- and dimethylbipyrrole in seawater. Environ. Sci. Technol. 2017, 51, 589–595. [Google Scholar] [CrossRef] [PubMed]
  28. Enami, S.; Vecitis, C.D.; Cheng, J.; Hoffmann, M.R.; Colussi, A.J. Global inorganic source of atmospheric bromine. J. Phys. Chem. A 2007, 111, 8749–8752. [Google Scholar] [CrossRef] [PubMed]
  29. Haag, W.R.; Gassman, E. Singlet oxygen in surface waters-Part I: Furfuryl alcohol as a trapping agent. Chemosphere 1984, 13, 631–640. [Google Scholar] [CrossRef]
  30. Finlayson-Pitts, B.J. The tropospheric chemistry of sea salt: A molecular-level view of the chemistry of NaCl and NaBr. Chem. Rev. 2003, 103, 4801–4822. [Google Scholar] [CrossRef] [PubMed]
  31. Rossi, M.J. Heterogeneous reactions on salts. Chem. Rev. 2003, 103, 4823–4882. [Google Scholar] [CrossRef] [PubMed]
  32. Mishra, H.; Enami, S.; Nielsen, R.J.; Stewart, L.A.; Hoffmann, M.R.; Goddard, W.A.; Colussi, A.J. Brønsted basicity of the air–water interface. Proc. Natl. Acad. Sci. USA 2012, 109, 18679–18683. [Google Scholar] [CrossRef] [PubMed]
  33. Enami, S.; Hoffmann, M.R.; Colussi, A.J. Halogen radical chemistry at aqueous interfaces. J. Phys. Chem. A 2016, 120, 6242–6248. [Google Scholar] [CrossRef] [PubMed]
  34. Wayne, R.P.; Barnes, I.; Biggs, P.; Burrows, J.P.; Canosa-Mas, C.E.; Hjorth, J.; Le Bras, G.; Moortgat, G.K.; Perner, D.; Poulet, G.; et al. The nitrate radical: Physics, chemistry, and the atmosphere. Atmos. Environ. Part A 1991, 25, 1–203. [Google Scholar] [CrossRef]
  35. Poskrebyshev, G.A.; Huie, R.E.; Neta, P. The rate and equilibrium constants for the reaction NO3 + Cl- ⇄ NO3 + Cl in aqueous solutions. J. Phys. Chem. A 2003, 107, 1964–1970. [Google Scholar] [CrossRef]
  36. Neta, P.; Huie, R.E. Rate constants for reactions of NO3 radicals in aqueous solutions. J. Phys. Chem. 1986, 90, 4644–4648. [Google Scholar] [CrossRef]
  37. Schweitzer, F.; Mirabel, P.; George, C. Multiphase chemistry of N2O5, ClNO2, and BrNO2. J. Phys. Chem. A 1998, 102, 3942–3952. [Google Scholar] [CrossRef]
  38. Roberts, J.M.; Osthoff, H.D.; Brown, S.S.; Ravishankara, A.R. N2O5 Oxidizes chloride to Cl2 in acidic atmospheric aerosol. Science 2008, 321, 1059. [Google Scholar] [CrossRef] [PubMed]
  39. Hu, J.H.; Shi, Q.; Davidovits, P.; Worsnop, D.R.; Zahniser, M.S.; Kolb, C.E. Reactive Uptake of Cl2(g) and Br2(g) by aqueous surfaces as a function of Br and I ion concentration: The effect of chemical reaction at the interface. J. Phys. Chem. 1995, 99, 8768–8776. [Google Scholar] [CrossRef]
  40. Wang, T.X.; Kelley, M.D.; Cooper, J.N.; Beckwith, R.C.; Margerum, D.W. Equilibrium, kinetic, and UV-spectral characteristics of aqueous bromine chloride, bromine, and chlorine species. Inorg. Chem. 1994, 33, 5872–5878. [Google Scholar] [CrossRef]
  41. Ianni, J.C. Kintecus V6.01. Available online: www.kintecus.com (accessed on 30 July 2016).
  42. Bichsel, Y.; von Gunten, U. Formation of iodo-trihalomethanes during disinfection and oxidation of iodide containing waters. Environ. Sci. Technol. 2000, 34, 2784–2791. [Google Scholar] [CrossRef]
  43. Troy, R.C.; Margerum, D.W. Non-metal redox kinetics: Hypobromite and hypobromous acid reactions with iodide and with sulfite and the hydrolysis of bromosulfate. Inorg. Chem. 1991, 30, 3538–3543. [Google Scholar] [CrossRef]
  44. Bichsel, Y.; von Gunten, U. Hypoiodous acid: Kinetics of the buffer-catalyzed disproportionation. Water Res. 2000, 34, 3197–3203. [Google Scholar] [CrossRef]
  45. Barkley, R.A.; Thompson, T.G. The total Iodine and Iodate-iodine content of sea-water. Deep Sea Res. 1960, 7, 24–34. [Google Scholar] [CrossRef]
  46. Chen, Z.; Megharaj, M.; Naidu, R. Speciation of iodate and iodide in seawater by non-suppressed ion chromatography with inductively coupled plasma mass spectrometry. Talanta 2007, 72, 1842–1846. [Google Scholar] [CrossRef] [PubMed]
  47. Kurylo, M.J.; Ouellette, P.A.; Laufer, A.H. Measurements of the pressure dependence of the hydroperoxy (HO2) radical self-disproportionation reaction at 298 K. J. Phys. Chem. 1986, 90, 437–440. [Google Scholar] [CrossRef]
  48. Barnes, R.J.; Lock, M.; Coleman, J.; Sinha, A. Observation of a new absorption band of HOBr and its atmospheric implications. J. Phys. Chem. 1996, 100, 453–457. [Google Scholar] [CrossRef]
  49. Haugen, H.K.; Weitz, E.; Leone, S.R. Accurate quantum yields by laser gain vs absorption spectroscopy: Investigation of Br/Br* channels in photofragmentation of Br2 and IBr. J. Phys. Chem. 1985, 83, 3402–3412. [Google Scholar] [CrossRef]
  50. Gershgoren, E.; Banin, U.; Ruhman, S. Caging and geminate recombination following photolysis of triiodide in solution. J. Phys. Chem. A 1998, 102, 9–16. [Google Scholar] [CrossRef]
  51. Callow, A.; Griffith, R.; McKeown, A. The photo-reaction between bromine and hydrogen peroxide in aqueous solution. Trans. Faraday Soc. 1939, 35, 412–420. [Google Scholar] [CrossRef]
  52. Treinin, A.; Hayon, E. Charge transfer spectra of halogen atoms in water. Correlation of the electronic transition energies of iodine, bromine, chlorine, hydroxyl, and hydrogen radicals with their electron affinities. J. Am. Chem. Soc. 1975, 97, 1716–1721. [Google Scholar] [CrossRef]
  53. Forsyth, J.E.; Zhou, P.; Mao, Q.; Asato, S.S.; Meschke, J.S.; Dodd, M.C. Enhanced inactivation of bacillus subtilis spores during solar photolysis of free available chlorine. Environ. Sci. Technol. 2013, 47, 12976–12984. [Google Scholar] [CrossRef] [PubMed]
  54. Jenkin, M.; Cox, R.; Hayman, G. Kinetics of the reaction of IO radicals with HO2 radicals at 298 K. Chem. Phys. Lett. 1991, 177, 272–278. [Google Scholar] [CrossRef]
  55. Francisco, J.S.; Hand, M.R.; Williams, I.H. Ab initio study of the electronic spectrum of HOBr. J. Phys. Chem. 1996, 100, 9250–9253. [Google Scholar] [CrossRef]
  56. Minaev, B.F. The singlet-triplet absorption and photodissociation of the HOCl, HOBr, and HOI molecules calculated by the MCSCF quadratic response method. J. Phys. Chem. A 1999, 103, 7294–7309. [Google Scholar] [CrossRef]
  57. Biedenkapp, D.; Hartshorn, L.G.; Bair, E.J. The O (1D)+ H2O reaction. Chem. Phys. Lett. 1970, 5, 379–381. [Google Scholar] [CrossRef]
  58. Poskrebyshev, G.A.; Neta, P.; Huie, R.E. Temperature dependence of the acid dissociation constant of the hydroxyl radical. J. Phys. Chem. A 2002, 106, 11488–11491. [Google Scholar] [CrossRef]
  59. Buxton, G.; Subhani, M. Radiation chemistry and photochemistry of oxychlorine ions. Part 2.—Photodecomposition of aqueous solutions of hypochlorite ions. J. Chem. Soc. Faraday Trans. 1 1972, 68, 958–969. [Google Scholar] [CrossRef]
  60. Orlando, J.J.; Burkholder, J.B. Gas-phase UV/Visible absorption spectra of HOBr and Br2O. J. Phys. Chem. 1995, 99, 1143–1150. [Google Scholar] [CrossRef]
  61. Rowley, D.M.; Mössinger, J.C.; Cox, R.A.; Jones, R.L. The UV-visible absorption cross-sections and atmospheric photolysis rate of HOI. JAtC 1999, 34, 137–151. [Google Scholar]
  62. Palmer, D.A.; Van Eldik, R. Spectral characterization and kinetics of formation of hypoiodous acid in aqueous solution. Inorg. Chem. 1986, 25, 928–931. [Google Scholar] [CrossRef]
  63. Deborde, M.; von Gunten, U. Reactions of chlorine with inorganic and organic compounds during water treatment—Kinetics and mechanisms: A critical review. Water Res. 2008, 42, 13–51. [Google Scholar] [CrossRef] [PubMed]
  64. Yang, Y.; Pignatello, J.J.; Ma, J.; Mitch, W.A. Comparison of halide impacts on the efficiency of contaminant degradation by sulfate and hydroxyl radical-based advanced oxidation processes (AOPs). Environ. Sci. Technol. 2014, 48, 2344–2351. [Google Scholar] [CrossRef] [PubMed]
  65. Minakata, D.; Kamath, D.; Maetzold, S. Mechanistic insight into the reactivity of chlorine-derived radicals in the aqueous-phase UV–chlorine advanced oxidation process: Quantum mechanical calculations. Environ. Sci. Technol. 2017, 51, 6918–6926. [Google Scholar] [CrossRef] [PubMed]
  66. Von Gunten, U.; Oliveras, Y. Kinetics of the reaction between hydrogen peroxide and hypobromous acid: Implication on water treatment and natural systems. Water Res. 1997, 31, 900–906. [Google Scholar] [CrossRef]
  67. Cerkovnik, J.; Plesničar, B. Recent Advances in the chemistry of hydrogen trioxide (HOOOH). Chem. Rev. 2013, 113, 7930–7951. [Google Scholar] [CrossRef] [PubMed]
  68. Johnson, D.W.; Margerum, D.W. Non-metal redox kinetics: A reexamination of the mechanism of the reaction between hypochlorite and nitrite ions. Inorg. Chem. 1991, 30, 4845–4851. [Google Scholar] [CrossRef]
  69. Lahoutifard, N.; Lagrange, P.; Lagrange, J.; Scott, S.L. Kinetics and mechanism of nitrite oxidation by HOBr/BrO in atmospheric water and comparison with oxidation by HOCl/ClO. J. Phys. Chem. A 2002, 106, 11891–11896. [Google Scholar] [CrossRef]
  70. Von Gunten, U. Ozonation of drinking water: Part II. Disinfection and by-product formation in presence of bromide, iodide or chlorine. Water Res. 2003, 37, 1469–1487. [Google Scholar] [CrossRef]
  71. Environmental Protection Agency, USA. National drinking water regulations: Disinfectants and disinfection byproducts. Fed. Regist. 1998, 63, 69390–69476. [Google Scholar]
  72. Grebel, J.E.; Pignatello, J.J.; Mitch, W.A. Impact of halide ions on natural organic matter-sensitized photolysis of 17-β-estradiol in saline waters. Environ. Sci. Technol. 2012, 46, 7128–7134. [Google Scholar] [CrossRef] [PubMed]
  73. Grebel, J.E.; Pignatello, J.J.; Mitch, W.A. Sorbic acid as a quantitative probe for the formation, scavenging and steady-state concentrations of the triplet-excited state of organic compounds. Water Res. 2011, 45, 6535–6544. [Google Scholar] [CrossRef] [PubMed]
  74. Parker, K.M.; Pignatello, J.J.; Mitch, W.A. Influence of salinity on triplet-state natural organic matter loss by energy transfer and electron transfer pathways. Environ. Sci. Technol. 2013, 47, 10987–10994. [Google Scholar] [CrossRef] [PubMed]
  75. Glover, C.M.; Rosario-Ortiz, F.L. Impact of halides on the photoproduction of reactive intermediates from organic matter. Environ. Sci. Technol. 2013, 47, 13949–13956. [Google Scholar] [CrossRef] [PubMed]
  76. Grebel, J.E.; Pignatello, J.J.; Song, W.; Cooper, W.J.; Mitch, W.A. Impact of halides on the photobleaching of dissolved organic matter. Mar. Chem. 2009, 115, 134–144. [Google Scholar] [CrossRef]
  77. Song, G.; Li, Y.; Hu, S.; Li, G.; Zhao, R.; Sun, X.; Xie, H. Photobleaching of chromophoric dissolved organic matter (CDOM) in the Yangtze River estuary: Kinetics and effects of temperature, pH, and salinity. Environ. Sci. Process Impacts 2017, 19, 861–873. [Google Scholar] [CrossRef] [PubMed]
  78. Larson, R.A.; Weber, E.J. Reaction Mechanisms in Environmental Organic Chemistry; Lewis Publishers: Boca Ratan, FL, USA, 1994. [Google Scholar]
  79. Wicktor, F.; Donati, A.; Herrmann, H.; Zellner, R. Laser based spectroscopic and kinetic investigations of reactions of the Cl atom with oxygenated hydrocarbons in aqueous solution. Phys. Chem. Chem. Phys. 2003, 5, 2562–2572. [Google Scholar] [CrossRef]
  80. Hasegawa, K.; Neta, P. Rate constants and mechanisms of reactions of Cl2 radicals. J. Phys. Chem. 1978, 82, 854–857. [Google Scholar] [CrossRef]
  81. Ershov, B.G.; Kelm, M.; Gordeev, A.V.; Janata, E. A pulse radiolysis study of the oxidation of Br- by dichloro radical anion in aqueous solution: Formation and properties of chlorobromo radical anion. Phys. Chem. Chem. Phys. 2002, 4, 1872–1875. [Google Scholar] [CrossRef]
  82. Lee, Y.; Yoon, J.; von Gunten, U. Kinetics of the oxidation of phenols and phenolic endocrine disruptors during water treatment with ferrate (Fe(VI)). Environ. Sci. Technol. 2005, 39, 8978–8984. [Google Scholar] [CrossRef] [PubMed]
  83. Criquet, J.; Rodriguez, E.M.; Allard, S.; Wellauer, S.; Salhi, E.; Joll, C.A.; von Gunten, U. Reaction of bromine and chlorine with phenolic compounds and natural organic matter extracts—Electrophilic aromatic substitution and oxidation. Water Res. 2015, 85, 476–486. [Google Scholar] [CrossRef] [PubMed]
  84. Gribble, G.W. Naturally Occurring Organohalogen Compopunds--A Comprehensive Update; Springer: Wien, Austria; New York, NY, USA, 2010. [Google Scholar]
  85. Jeffers, P.M.; Wolfe, N.L. On the degradation of methyl bromide in sea water. Geophys. Res. Lett. 1996, 23, 1773–1776. [Google Scholar] [CrossRef]
  86. Moore, R.M. A photochemical source of methyl chloride in saline waters. Environ. Sci. Technol. 2008, 42, 1933–1937. [Google Scholar] [CrossRef] [PubMed]
  87. Moore, R.M.; Zafiriou, O.C. Photochemical production of methyl iodide in seawater. J. Geophys. Res. Atmos. 1994, 99, 16415–16420. [Google Scholar] [CrossRef]
  88. Martino, M.; Mills, G.P.; Woeltjen, J.; Liss, P.S. A new source of volatile organoiodine compounds in surface seawater. Geophys. Res. Lett. 2009, 36. [Google Scholar] [CrossRef]
  89. Jones, C.E.; Carpenter, L.J. Solar Photolysis of CH2I2, CH2ICl, and CH2IBr in water, saltwater, and seawater. Environ. Sci. Technol. 2006, 40, 1372. [Google Scholar] [CrossRef]
  90. Anastasio, C.; Matthew, B.M. A chemical probe technique for the determination of reactive halogen species in aqueous solution: Part 2—Chloride solutions and mixed bromide/chloride solutions. Atmos. Chem. Phys. 2006, 6, 2439–2451. [Google Scholar] [CrossRef]
  91. Liu, H.; Zhao, H.; Quan, X.; Zhang, Y.; Chen, S. Formation of chlorinated intermediate from bisphenol a in surface saline water under simulated solar light irradiation. Environ. Sci. Technol. 2009, 43, 7712–7717. [Google Scholar] [CrossRef] [PubMed]
  92. Tamtam, F.; Chiron, S. New insight into photo-bromination processes in saline surface waters: The case of salicylic acid. Sci. Total Environ. 2012, 435, 345–350. [Google Scholar] [CrossRef] [PubMed]
  93. Narukawa, M.; Kawamura, K.; Hatsushika, H.; Yamazaki, K.; Li, S.-M.; Bottenheim, J.W.; Anlauf, K.G. Measurement of halogenated dicarboxylic acids in the arctic aerosols at polar sunrise. J. Atmos. Chem. 2003, 44, 323–335. [Google Scholar] [CrossRef]
  94. Méndez-Díaz, J.D.; Shimabuku, K.K.; Ma, J.; Enumah, Z.O.; Pignatello, J.J.; Mitch, W.A.; Dodd, M.C. Sunlight-driven photochemical halogenation of dissolved organic matter in seawater: A natural abiotic source of organobromine and organoiodine. Environ. Sci. Technol. 2014, 48, 7418–7427. [Google Scholar] [CrossRef] [PubMed]
  95. Hao, Z.; Yin, Y.; Cao, D.; Liu, J. Probing and comparing the photobromination and photoiodination of dissolved organic matter by using ultra-high-resolution mass spectrometry. Environ. Sci. Technol. 2017, 51, 5464–5472. [Google Scholar] [CrossRef] [PubMed]
  96. Jortner, J.; Ottolenghi, M.; Stein, G. On the photochemistry of aqueous solutions of chloride, bromide, and iodide ions. J. Phys. Chem. 1964, 68, 247–255. [Google Scholar] [CrossRef]
  97. Kalmar, J.; Doka, E.; Lente, G.; Fabian, I. Aqueous photochemical reactions of chloride, bromide, and iodide ions in a diode-array spectrophotometer. Autoinhibition in the photolysis of iodide ions. Dalton Trans. 2014, 43, 4862–4870. [Google Scholar] [CrossRef] [PubMed]
  98. Kiwi, J.; Lopez, A.; Nadtochenko, V. Mechanism and kinetics of the OH-radical intervention during Fenton oxidation in the presence of a significant amount of radical scavenger (Cl). Environ. Sci. Technol. 2000, 34, 2162–2168. [Google Scholar] [CrossRef]
  99. Pignatello, J.J.; Oliveros, E.; MacKay, A. Advanced oxidation processes for organic contaminant destruction based on the Fenton reaction and related chemistry. Crit. Rev. Env. Sci. Technol. 2006, 36, 1–84. [Google Scholar] [CrossRef]
  100. Machulek, A.; Moraes, J.E.F.; Vautier-Giongo, C.; Silverio, C.A.; Friedrich, L.C.; Nascimento, C.A.O.; Gonzalez, M.C.; Quina, F.H. Abatement of the inhibitory effect of chloride anions on the photo-fenton process. Environ. Sci. Technol. 2007, 41, 8459–8463. [Google Scholar] [CrossRef] [PubMed]
  101. Yuan, R.; Ramjaun, S.N.; Wang, Z.; Liu, J. Photocatalytic degradation and chlorination of azo dye in saline wastewater: Kinetics and AOX formation. Chem. Eng. J. 2012, 192, 171–178. [Google Scholar] [CrossRef]
  102. Yamazaki, S.; Tanimura, T.; Yoshida, A.; Hori, K. Reaction mechanism of photocatalytic degradation of chlorinated ethylenes on porous TiO2 pellets: Cl radical-initiated mechanism. J. Phys. Chem. A 2004, 108, 5183–5188. [Google Scholar] [CrossRef]
  103. Kormann, C.; Bahnemann, D.; Hoffmann, M.R. Photolysis of chloroform and other organic molecules in aqueous titanium dioxide suspensions. Environ. Sci. Technol. 1991, 25, 494–500. [Google Scholar] [CrossRef]
  104. Wu, Z.; Fang, J.; Xiang, Y.; Shang, C.; Li, X.; Meng, F.; Yang, X. Roles of reactive chlorine species in trimethoprim degradation in the UV/chlorine process: Kinetics and transformation pathways. Water Res. 2016, 104, 272–282. [Google Scholar] [CrossRef] [PubMed]
  105. Heeb, M.B.; Criquet, J.; Zimmermann-Steffens, S.G.; von Gunten, U. Oxidative treatment of bromide-containing waters: Formation of bromine and its reactions with inorganic and organic compounds—A critical review. Water Res. 2014, 48, 15–42. [Google Scholar] [CrossRef] [PubMed]
  106. Benter, T.; Feldmann, C.R.; Kirchner, U.; Schmidt, M.; Schmidt, S.; Schindler, R. UV/VIS-absorption Spectra of HOBr and CH3OBr; Br(2P3/2) atom yields in the photolysis of HOBr. Ber. Bunsenges. Phys. Chem. 1995, 99, 1144–1147. [Google Scholar] [CrossRef]
  107. Criquet, J.; Allard, S.; Salhi, E.; Joll, C.A.; Heitz, A.; von Gunten, U. Iodate and iodo-trihalomethane formation during chlorination of iodide-containing waters: Role of bromide. Environ. Sci. Technol. 2012, 46, 7350–7357. [Google Scholar] [CrossRef] [PubMed]
  108. Allard, S.; Tan, J.; Joll, C.A.; von Gunten, U. Mechanistic study on the formation of Cl/Br/I trihalomethanes during chlorination/chloramination combined with a theoretical cytotoxicity evaluation. Environ. Sci. Technol. 2015, 49, 11105–11114. [Google Scholar] [CrossRef] [PubMed]
  109. Mark, G.; Schuchmann, M.N.; Schuchmann, H.-P.; von Sonntag, C. The photolysis of potassium peroxodisulphate in aqueous solution in the presence of tert-butanol: A simple actinometer for 254 nm radiation. J. Photochem. Photobiol. A Chem. 1990, 55, 157–168. [Google Scholar] [CrossRef]
  110. Baxendale, J.; Wilson, J. The photolysis of hydrogen peroxide at high light intensities. Trans. Faraday Soc. 1957, 53, 344–356. [Google Scholar] [CrossRef]
  111. Guan, Y.-H.; Ma, J.; Li, X.-C.; Fang, J.-Y.; Chen, L.-W. Influence of pH on the formation of sulfate and hydroxyl radicals in the UV/peroxymonosulfate system. Environ. Sci. Technol. 2011, 45, 9308–9314. [Google Scholar] [CrossRef] [PubMed]
  112. Beitz, T.; Bechmann, W.; Mitzner, R. Investigations of reactions of selected azaarenes with radicals in water. 2. Chlorine and bromine radicals. J. Phys. Chem. A 1998, 102, 6766–6771. [Google Scholar] [CrossRef]
  113. Canonica, S.; Kohn, T.; Mac, M.; Real, F.J.; Wirz, J.; von Gunten, U. Photosensitizer method to determine rate constants for the reaction of carbonate radical with organic compounds. Environ. Sci. Technol. 2005, 39, 9182–9188. [Google Scholar] [CrossRef] [PubMed]
  114. Yang, Y.; Pignatello, J.J.; Ma, J.; Mitch, W.A. Effect of matrix components on UV/H2O2 and UV/S2O82− advanced oxidation processes for trace organic degradation in reverse osmosis brines from municipal wastewater reuse facilities. Water Res. 2016, 89, 192–200. [Google Scholar] [CrossRef] [PubMed]
  115. Fang, J.Y.; Shang, C. Bromate formation from bromide oxidation by the UV/Persulfate process. Environ. Sci. Technol. 2012, 46, 8976–8983. [Google Scholar] [CrossRef] [PubMed]
  116. Lutze, H.V.; Bakkour, R.; Kerlin, N.; von Sonntag, C.; Schmidt, T.C. Formation of bromate in sulfate radical based oxidation: Mechanistic aspects and suppression by dissolved organic matter. Water Res. 2014, 53, 370–377. [Google Scholar] [CrossRef] [PubMed]
  117. Fawell, J.; Walker, M. Approaches to determining regulatory values for carcinogens with particular reference to bromate. Toxicology 2006, 221, 149–153. [Google Scholar] [CrossRef] [PubMed]
  118. Kläning, U.K.; Wolff, T. Laser flash photolysis of HCIO, CIO, HBrO, and BrO in aqueous solution. reactions of Cl-and Br-atoms. Ber. Bunsenges. Phys. Chem. 1985, 89, 243–245. [Google Scholar] [CrossRef]
  119. Fortnum, D.H.; Battaglia, C.J.; Cohen, S.R.; Edwards, J.O. The kinetics of the oxidation of halide ions by monosubstituted peroxides. J. Am. Chem. Soc. 1960, 82, 778–782. [Google Scholar] [CrossRef]
  120. Lente, G.; Kalmár, J.; Baranyai, Z.; Kun, A.; Kék, I.; Bajusz, D.; Takács, M.; Veres, L.; Fábián, I. One-versus two-electron oxidation with peroxomonosulfate ion: Reactions with iron(II), vanadium(IV), halide ions, and photoreaction with cerium(III). Inorg. Chem. 2009, 48, 1763–1773. [Google Scholar] [CrossRef] [PubMed]
  121. Li, J.; Jiang, J.; Zhou, Y.; Pang, S.-Y.; Gao, Y.; Jiang, C.; Ma, J.; Jin, Y.; Yang, Y.; Liu, G.; et al. Kinetics of oxidation of iodide (I) and hypoiodous acid (HOI) by peroxymonosulfate (PMS) and formation of iodinated products in the PMS/I/NOM system. Environ. Sci. Technol. Lett. 2017, 4, 76–82. [Google Scholar] [CrossRef]
  122. Pan, Y.; Cheng, S.; Yang, X.; Ren, J.; Fang, J.; Shang, C.; Song, W.; Lian, L.; Zhang, X. UV/chlorine treatment of carbamazepine: Transformation products and their formation kinetics. Water Res. 2017, 116, 254–265. [Google Scholar] [CrossRef] [PubMed]
Sample Availability: Not available.
Scheme 1. Generation of RHS in waters through the action of sunlight.
Scheme 1. Generation of RHS in waters through the action of sunlight.
Molecules 22 01684 sch001
Scheme 2. Proposed pathways of sensitized oxidation of halide ions in water.
Scheme 2. Proposed pathways of sensitized oxidation of halide ions in water.
Molecules 22 01684 sch002
Scheme 3. Photolysis of hypohalites.
Scheme 3. Photolysis of hypohalites.
Molecules 22 01684 sch003
Scheme 4. Reactions of RHS with inorganic species.
Scheme 4. Reactions of RHS with inorganic species.
Molecules 22 01684 sch004
Scheme 5. The mechanism of BrO3 formation by SO4−•.
Scheme 5. The mechanism of BrO3 formation by SO4−•.
Molecules 22 01684 sch005
Scheme 6. The mechanism of IO3 formation by peroxymonosulfate oxidation of iodide.
Scheme 6. The mechanism of IO3 formation by peroxymonosulfate oxidation of iodide.
Molecules 22 01684 sch006
Table 1. Simulated speciation of nrRHS in different waters. Molar ratio relative to Cl2 after 5 min except where noted.
Table 1. Simulated speciation of nrRHS in different waters. Molar ratio relative to Cl2 after 5 min except where noted.
RHS/Cl2Br2BrClCl3BrCl2Br2ClBr3HOBr/OBrHOCl/OCl
Wastewater4.01 × 1032.270.025733.54173.50.95 × 109
(1.74 × 1010) *
2.57 × 105
(5.92 × 105) *
Seawater1.04 × 10424.70.098253338001456.42 × 109
(7.08 × 109) *
1.79 × 105
(6.2 × 105) *
* After 60 min.
Table 2. Some examples of halogenation reactions of specific organic compounds in illuminated salty water systems.
Table 2. Some examples of halogenation reactions of specific organic compounds in illuminated salty water systems.
CompoundProposed OriginReferences
CH3Cl(a) nucleophilic displacement by chloride on CH3I and/or CH3Br in seawater;
(b) is produced on irradiation of lignin-like DOM model compounds (4-methoxy-1-naphthol; syringic acid; 2-methoxyphenol; 3,4,5-trimethoxy benzoic acid; and2-methoxyhydroquinone) in chloride solution
(a) [85]
(b) [86]
CH3Iformed after simulated solar irradiation of filtered seawater; production was enhanced when samples were degassed or iodide was added; proposed origin is recombination of CH3 and I radicals.[87]
CH2I2, CHI3, and CHI2Clformed by reactions of DOM with HOI generated via oxidation of I by O3[88]
CH2IClphotolysis product of CH2I2 in seawater[89]
Cl-CH2CH(OH)CH2OH and
Br-CH2CH(OH)CH2OH
CH2=CHCH2OH reaction with reactive halogen species[90]
3-Cl and 3,3-diCl bisphenol Asolar irradiation of bisphenol A in coastal seawater and saline solution containing 0.13–0.66 mM Fe(III) and fulvic acid; Cl2−• was detected by its absorption spectrum, and OH as its DMPO adduct by EPR spectroscopy; proposed source of halogen radicals: FeIIICl → FeII + Cl or FeIIIOH → FeII + OH, followed by OH + Cl → Cl.[91]
5-bromo-and 3,5-dibromosalicylic acidssolar irradiation of salicylic acid in artificial seawater and brackish lagoon water[92]
mixed poly-brominated/chlorinated bipyrrolesirradiation of 1,1-dimethyl-2,2′-bipyrrole and 1’-methyl-1,2’bipyrrole in ozonated seawater; proposed oxidation of Br and I by O3 to form HOX/X2.[27]
halogenated dicarboxylic acidsisolated from arctic aerosols; unclear whether transformations occurred in the liquid phase[93]

Share and Cite

MDPI and ACS Style

Yang, Y.; Pignatello, J.J. Participation of the Halogens in Photochemical Reactions in Natural and Treated Waters. Molecules 2017, 22, 1684. https://doi.org/10.3390/molecules22101684

AMA Style

Yang Y, Pignatello JJ. Participation of the Halogens in Photochemical Reactions in Natural and Treated Waters. Molecules. 2017; 22(10):1684. https://doi.org/10.3390/molecules22101684

Chicago/Turabian Style

Yang, Yi, and Joseph J. Pignatello. 2017. "Participation of the Halogens in Photochemical Reactions in Natural and Treated Waters" Molecules 22, no. 10: 1684. https://doi.org/10.3390/molecules22101684

Article Metrics

Back to TopTop