Elsevier

Water Research

Volume 42, Issues 1–2, January 2008, Pages 13-51
Water Research

Review
Reactions of chlorine with inorganic and organic compounds during water treatment—Kinetics and mechanisms: A critical review

https://doi.org/10.1016/j.watres.2007.07.025Get rights and content

Abstract

Numerous inorganic and organic micropollutants can undergo reactions with chlorine. However, for certain compounds, the expected chlorine reactivity is low and only small modifications in the parent compound's structure are expected under typical water treatment conditions. To better understand/predict chlorine reactions with micropollutants, the kinetic and mechanistic information on chlorine reactivity available in literature was critically reviewed. For most micropollutants, HOCl is the major reactive chlorine species during chlorination processes. In the case of inorganic compounds, a fast reaction of ammonia, halides (Br and I), SO32−, CN, NO2, As(III) and Fe(II) with HOCl is reported (103–109 M−1 s−1) whereas low chlorine reaction rates with Mn(II) were shown in homogeneous systems. Chlorine reactivity usually results from an initial electrophilic attack of HOCl on inorganic compounds. In the case of organic compounds, second-order rate constants for chlorination vary over 10 orders of magnitude (i.e. <0.1–109 M−1 s−1). Oxidation, addition and electrophilic substitution reactions with organic compounds are possible pathways. However, from a kinetic point of view, usually only electrophilic attack is significant. Chlorine reactivity limited to particular sites (mainly amines, reduced sulfur moieties or activated aromatic systems) is commonly observed during chlorination processes and small modifications in the parent compound's structure are expected for the primary attack. Linear structure–activity relationships can be used to make predictions/estimates of the reactivity of functional groups based on structural analogy. Furthermore, comparison of chlorine to ozone reactivity towards aromatic compounds (electrophilic attack) shows a good correlation, with chlorine rate constants being about four orders of magnitude smaller than those for ozone.

Introduction

Due to their capability for disinfection (microorganisms) and oxidation (e.g. taste and odor control, elimination of micropollutants, etc.), chemical oxidants (i.e. ozone, chlorine, chlorine dioxide, chloramines, etc.) are commonly used in water treatment processes (Hoff and Geldreich, 1981; Wolfe et al., 1984; Morris, 1986; Burlingame et al., 1992; Hoigné, 1998; Gottschalk et al., 2000; von Gunten, 2003; Legube, 2003; Bruchet and Duguet, 2004). However, under certain circumstances, oxidants can induce formation of potentially harmful by-products or transformation products due to their reactivity with water matrix components or micropollutants (Cancho et al., 2000; Bichsel and von Gunten., 2000; Simmons et al., 2002; Richardson et al., 2003; Plewa et al., 2004; Richardson, 2005; Krasner et al., 2006).

Owing to its low cost, chlorine is globally the most used chemical oxidant for drinking water disinfection. Drinking water disinfection commonly involves the use of chlorine at one or two point(s) in the treatment process, i.e., for pre-treatment (to induce a primary disinfection at the beginning of the treatment process) and/or for post-treatment (to maintain a disinfectant residual in the distribution system). Despite its low activity on microorganisms in biofilms, chlorine can lead to a significant removal of the majority of planktonic bacteria (Le Chevallier et al., 1988; Bois et al., 1997). Added near the end of the treatment process, i.e., before water release in the distribution system (post chlorination), chlorine thus plays an important role to limit the growth of heterotrophic microorganisms. As a chemical oxidant, though less reactive than ozone, chlorine can transform numerous inorganic and organic micropollutants found in water (e.g. Fe(II), As(III), NO2, phenols, pesticides, pharmaceuticals, etc.) (Johnson and Margerum, 1991; Magara et al., 1994; Folkes et al., 1995; Gallard and von Gunten, 2002; Lahoutifard et al., 2003; Diurk and Colette, 2006; Dodd et al., 2006; von Gunten et al., 2006). Chlorination usually represents an efficient process to remove/transform inorganic micropollutants. However, due to the potentially harmful chlorinated transformation products, chlorination is usually not applied for oxidation of organic micropollutants.

Similar to other disinfection processes, chlorination presents certain disadvantages in spite of its broad use and its benefits for the improvement of microbial water quality: (i) Due to its pH-dependent aqueous chemistry, various species of chlorine (HOCl, ClO, Cl2, etc.) may be present in solution (Doré, 1989). These forms of chlorine show significant differences in their reactivity with microorganisms and micropollutants. Therefore, variability in oxidation or disinfection efficiency can be observed depending on the pH of the water. (ii) Chlorine interacts with dissolved natural organic matter (DNOM). Numerous so-called disinfection by-products (DBPs) can result from the reaction of chlorine with DNOM. Among these DBPs, trihalomethanes (THMs) and haloacetic acids (HAAs) were the first chlorine DBPs reported and are currently regulated in the EU (THMs) and the USA (THMs, HAAs) (Richardson, 2005). Currently, about 600 DBPs are identified, among them some highly toxic compounds such as iodo and bromo compounds (Bichsel and von Gunten, 2000; Richardson et al., 2003; Plewa et al., 2004; Krasner et al., 2006), MX (Onstad and Weinberg, 2005; Krasner et al., 2006), halonitromethanes (Krasner et al., 2006) and N-nitrosodimethylamine (Mitch et al., 2003). These individual DBPs or mixtures of DBPs could represent a potential human health risk (Cancho et al., 2000; Simmons et al., 2002; Richardson et al., 2003; Plewa et al., 2004; Richardson, 2005; Krasner et al., 2006). (iii) Because organic micropollutants are typically not mineralized, numerous transformation products can be formed as a result of the oxidation of organic compounds during water chlorination processes (Magara et al., 1994; Gallard and von Gunten, 2002; Hu et al., 2002a, Hu et al., 2002b, Hu et al., 2003; Moriyama et al., 2004; Dodd and Huang, 2004; Dodd et al., 2005; Rule et al., 2005; Diurk and Colette, 2006; von Gunten et al., 2006). Little is known on the stability and the biological effects of these compounds. However, in some cases, certain transformation products are fairly stable against further transformation and could persist for hours to days even in presence of residual chlorine. Moreover, in the case of some endocrine disruptors (i.e. nonylphenol, bisphenol A and hormones), some pesticides (i.e. chlorpyrifos), some pharmaceuticals (i.e. acetaminophen) and some azo-dyes, potentially toxic or biologically active chlorination products were reported (Hu et al., 2002a, Hu et al., 2002b, Hu et al., 2003; Wu and Laird, 2003; Bedner and MacCrehan, 2006; Moriyama et al., 2004; Oliveira et al., 2006). (iv) In bromide-containing waters, chlorination leads to bromine formation. Bromine is usually more reactive than chlorine, especially with phenolic compounds (Gallard et al., 2003; Acero et al., 2005b). Under these conditions, bromination can be highly significant and brominated products can be formed (Gallard et al., 2003; Acero et al., 2005b; Hu et al., 2006).

This study presents an overview on chlorination of drinking water with an emphasis on kinetics and mechanisms of chlorine reactions. Based on literature data, an overview over chlorine reactivity with inorganic and organic compounds is presented. For typical functional groups, chlorination kinetics and mechanisms are described. By structural analogy, linear structure–activity relationships are proposed. Finally, for some organic micropollutants relevant for urban water management, a discussion on expected and observed chlorine reactivities is provided.

Section snippets

Aqueous chlorine chemistry

In water treatment, gaseous chlorine Cl2 or hypochlorite are commonly used for chlorination processes. Chlorine gas (Cl2) hydrolyzes in water according to the following reaction:Cl2+H2Ok-1k1HOCl+Cl+H+, KCl2=k1/k−1,where k1 and k−1 values, calculated at μ=0 M and 25 °C from Wang and Margerum, are 22.3 s−1 and 4.3×104 M−2 s−1, respectively. For temperatures between 0 and 25 °C, KCl2 ranges from 1.3×10−4 to 5.1×10−4 M2 (Wang and Margerum, 1994). Hypochlorous acid resulting from reaction (1), is a weak

Oxidation of inorganic and organic compounds by chlorine

The reactivity of chlorine depends on chlorine speciation as a function of pH. Among the different aqueous chlorine species, hypochlorous acid is the major reactive form during water treatment. The other species are typically present in concentrations that are too low or show insufficient reactivity to be significant (Morris, 1978). For most of the chlorination reactions, the elementary reaction can be formulated asHOCl+B→products,where B is an organic or inorganic compound.

Other elementary

Chlorine reactivity towards organic micropollutants relevant to water treatment

An overview over the chlorine reactivity (kinetic and mechanism) towards the main classes of organic and inorganic compounds has been provided in this study. For organic compounds with complex chemical structures, the main reactive sites can be derived by considering the known chlorine reactivity with the various functional groups. Because numerous kinetic and mechanistic studies on chlorination of pharmaceuticals, endocrine disruptors and cyanotoxins are available, a comparison between

Conclusion

It has been shown that numerous inorganic and organic micropollutants can be transformed by chlorine. However, for certain compounds, the chlorine reactivity is low and only small modifications in the parent compound's structure are expected under water treatment conditions.

During chlorination processes, HOCl is generally the major reactive species for the reaction with micropollutants. A pH dependence of organic and inorganic micropollutant transformation is commonly observed due to chlorine

Acknowledgements

We would like to thank Silvio Canonica, Michael Dodd and William Arnold for fruitful discussions and insightful comments and Claire Wedema for correcting the English. This review was performed within the framework of European Union project TECHNEAU (Contract number 018320). We gratefully acknowledge the financial support.

References (185)

  • J. De Laat et al.

    Chloration de composés organiques: demande en chlore et réactivité vis à vis de la formation des trihalométhanes

    Water Res.

    (1982)
  • M.C. Dodd et al.

    Aqueous chlorination of the antibacterial agent trimethoprim: reaction kinetics and pathways

    Water Res.

    (2007)
  • R. Drozdz et al.

    Oxidation of amino-acids and peptides in reaction with myeloperoxidase, chloride and hydrogen peroxide

    Biochim. Biophys. Acta.

    (1988)
  • L.L. Folkes et al.

    Kinetics and mechanisms of hypochlorous acid reactions

    Arch. Biochem. Biophys.

    (1995)
  • H. Gallard et al.

    Rate constants of reactions of bromine with phenols in aqueous solution

    Water Res.

    (2003)
  • H. Gallard et al.

    Chlorination of bisphenol a: kinetics and byproducts formation

    Chemosphere

    (2004)
  • J.P. Gould et al.

    The kinetics and primary products of uracil chlorination

    Water Res.

    (1984)
  • J.P. Gould et al.

    The formation of stable organic chloramines during the aqueous chlorination of cytosine and 5-methylcytosine

    Water Res.

    (1984)
  • R. Hirsch et al.

    Occurrence of antibiotics in the environment

    Sci. Tot. Environ.

    (1999)
  • L. Ho et al.

    Differences in the chlorine reactivity of four microcystin analogues

    Water Res.

    (2006)
  • H.L. Kopperman et al.

    Aqueous chlorination of α-terpineol

    Tetrahedron

    (1976)
  • N. Lahoutifard et al.

    Kinetics and mechanism of nitrite oxidation by hypochlorous acid in the aqueous phase

    Chemosphere

    (2003)
  • J.M. Antelo et al.

    A kinetic study of the chlorination of tertiary alcoholamines in alkaline media

    Int. J. Chem. Kinet.

    (1985)
  • J.M. Antelo et al.

    Kinetics of the N-chlorination of 2-aminobutyric, 3-aminobutyric, 3-aminoisobutyric and 4-aminobutyric acids in aqueous solution

    Int. J. Kinet.

    (1992)
  • J.M. Antelo et al.

    Kinetic study of the formation of N-chloramines

    Int J. Chem. Kinet.

    (1995)
  • X.L. Armesto et al.

    Chlorination of dipeptides by hypochlorous acid in aqueous solution

    Gazz. Chim. Ital.

    (1994)
  • X.L. Armesto et al.

    N reactivity vs O reactivity in aqueous chlorination

    Int. J. Chem. Kinet.

    (1994)
  • X.L. Armesto et al.

    Concerted grob fragmentation in N-halo-α-amino acid decomposition

    J. Org. Chem.

    (1994)
  • X.L. Armesto et al.

    Intracellular oxidation of dipeptides: very fast halogenation of the amino-terminal residue

    J. Chem. Perkin Trans.

    (2001)
  • J. Arotsky et al.

    Halogen cations

    Q. Rev. Chem. Soc.

    (1962)
  • R. Banker et al.

    Uracil moiety is required for toxicity of the cyanobacterial hepatotoxin cylindrospermopsin

    J. Toxicol. Environ. Heallth, Part A

    (2001)
  • M. Bedner et al.

    Transformation of acetaminophen by chlorination produces the toxicants 1,4-benzoquinone and N-acetyl-p-benzoquinone imine

    Environ. Sci. Technol.

    (2006)
  • Y. Bichsel et al.

    Formation of iodo-trihalomethanes during disinfection and oxidation of iodide-containing waters

    Environ. Sci. Technol.

    (2000)
  • S.D. Boyce et al.

    Reaction pathways of trihalomethane formation from the halogenation of dihydroxyaromatic model compounds for humic acid

    Environ. Sci. Technol.

    (1983)
  • A. Bruchet et al.

    Role of oxidants and disinfectants on the removal masking and generation of tastes and odours

    Water Sci. Technol.

    (2004)
  • G.A. Burlingame et al.

    Cucumber flavour in Philadelphia's drinking water

    J. Am. Water Works Assoc.

    (1992)
  • R.H. Burttschell et al.

    Chlorine derivatives of phenol causing taste and odor

    J. Am. Water Works Assoc.

    (1959)
  • M. Canle

    Mecanismos de Cloracion de α-Aminoacidos y de Fragmentacion de (N-X)-α-Aminoacidos

    (1994)
  • R.M. Carlson et al.

    Organochemical implications of water chlorination

  • R.M. Carlson et al.

    Aqueous chlorination products of polynuclear aromatic hydrocarbons

  • D.P. Cherney et al.

    Monitoring the speciation of aqueous free chlorine from pH 1 to 12 with Raman spectroscopy to determine the identity of potent low-pH oxidant

    Appl. Spectrosc.

    (2006)
  • A.R. Choppin et al.

    The oxidation of aqueous sulfide solutions by hypochlorites

    J. Am. Chem. Soc.

    (1937)
  • B. Conyers et al.

    Chloramines V: products and implications of the chlorination of lysine in municipal wastewaters

    Environ. Sci. Technol.

    (1997)
  • W.R. Cullen et al.

    Arsenic speciation in the environment

    Chem. Rev.

    (1989)
  • M.J. Davies et al.

    Hypochlorite-induces oxidation of thiols: formation of thiyl radicals and the role of sulfenyl chlorides as intermediates

    Free Rad. Res.

    (2000)
  • De Laat, 1981. Contribution à l’étude du mécanisme de formation des trihalométhanes: incidence de l’azote ammoniacal et...
  • M. Deborde et al.

    Aqueous chlorination kinetics of some endocrine disruptors

    Environ. Sci. Technol.

    (2004)
  • M. Deborde et al.

    Kinetics of aqueous ozone-induces oxidation of some endocrine disruptors

    Environ. Sci. Technol.

    (2005)
  • S.E. Diurk et al.

    Degradation of chlorpyrifos in aqueous chlorine solutions: pathways, kinetics and modelling

    Environ. Sci. Technol.

    (2006)
  • M.C. Dodd et al.

    Transformation of the antibacterial agent sulfamethoxazole in reactions with chlorine: kinetics, mechanisms, and pathways

    Environ. Sci. Technol.

    (2004)
  • Cited by (1590)

    • Mercaptan removal with electroscrubbing pilot

      2024, Separation and Purification Technology
    View all citing articles on Scopus
    View full text