You are on page 1of 28

w a t e r r e s e a r c h 4 8 ( 2 0 1 4 ) 1 5 e4 2

Available online at www.sciencedirect.com

journal homepage: www.elsevier.com/locate/watres

Oxidative treatment of bromide-containing waters:


Formation of bromine and its reactions with
inorganic and organic compounds d A critical
review
Miche`le B. Heeb a, Justine Criquet b,c,d, Saskia G. Zimmermann-Steffens a,
Urs von Gunten a,b,e,*
a

School of Architecture, Civil and Environmental Engineering (ENAC), Ecole Polytechnique Federale de Lausanne
(EPFL), CH-1015 Lausanne, Switzerland
b
Eawag, Swiss Federal Institute of Aquatic Science and Technology, CH-8600 Dubendorf, Switzerland
c
Universite Lille 1, Laboratoire Geosyste`mes, UMR CNRS 8217, 59655 Villeneuve dAscq, France
d
Curtin Water Quality Research Centre, Curtin University, GPO Box U1987, Perth, WA 6845, Australia
e
Institute of Biogeochemistry and Pollutant Dynamics, ETH Zurich, CH-8092 Zurich, Switzerland

article info

abstract

Article history:

Bromide (Br) is present in all water sources at concentrations ranging from w10 to

Received 25 June 2013

>1000 mg L1 in fresh waters and about 67 mg L1 in seawater. During oxidative water

Received in revised form

treatment bromide is oxidized to hypobromous acid/hypobromite (HOBr/OBr) and other

23 August 2013

bromine species. A systematic and critical literature review has been conducted on the

Accepted 25 August 2013

reactivity of HOBr/OBr and other bromine species with inorganic and organic compounds,

Available online 4 September 2013

including micropollutants.

Keywords:

been calculated and it could be shown that HOBr/OBr are the dominant species in fresh

Bromine

waters. In ocean waters, other bromine species such as Br2, BrCl, and Br2O gain importance

HOBr

and may have to be considered under certain conditions.

The speciation of bromine in the absence and presence of chloride and chlorine has

Oxidation kinetics

HOBr reacts fast with many inorganic compounds such as ammonia, iodide, sulfite,

Water treatment

nitrite, cyanide and thiocyanide with apparent second-order rate constants in the order of

Inorganic compounds

104e109 M1 s1 at pH 7. No rate constants for the reactions with Fe(II) and As(III) are

Organic compounds

available. Mn(II) oxidation by bromine is controlled by a Mn(III,IV) oxide-catalyzed process


involving Br2O and BrCl.
Bromine shows a very high reactivity toward phenolic groups (apparent second-order
rate constants kapp z 103e105 M1 s1 at pH 7), amines and sulfamides (kapp z 105
e106 M1 s1 at pH 7) and S-containing compounds (kapp z 105e107 M1 s1 at pH 7). For
phenolic moieties, it is possible to derive second-order rate constants with a Hammett-sbased QSAR approach with logkHOBr=PhO 7:8  3:5Ss. A negative slope is typical for
electrophilic substitution reactions.
In general, kapp of bromine reactions at pH 7 are up to three orders of magnitude greater
than for chlorine. In the case of amines, these rate constants are even higher than for

* Corresponding author. Eawag, Swiss Federal Institute of Aquatic Science and Technology, CH-8600 Dubendorf, Switzerland. Tel.: 41 58
765 5270; fax: 41 58 765 5210.
E-mail addresses: urs.vongunten@eawag.ch, vongunten@eawag.ch (U. von Gunten).
0043-1354/$ e see front matter 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.watres.2013.08.030

16

w a t e r r e s e a r c h 4 8 ( 2 0 1 4 ) 1 5 e4 2

ozone. Model calculations show that depending on the bromide concentration and the pH,
the high reactivity of bromine may outweigh the reactions of chlorine during chlorination
of bromide-containing waters.
2013 Elsevier Ltd. All rights reserved.

1.

Introduction

Oxidative water treatment such as chlorination or ozonation


has been applied in the first half of the 20th century primarily
for disinfection purposes. Consecutively, other treatment
goals such as the removal of taste and odor compounds and
the elimination of (anthropogenic) micropollutants were
tackled as well (Sedlak and von Gunten, 2011). One major
drawback of oxidative water treatment is the formation of
disinfection by-products (DBPs) and transformation products
resulting from the reaction of oxidants with moieties of the
water matrix and micropollutants, respectively, that might be
of human or environmental health concern (Krasner et al.,
2006; von Sonntag and von Gunten, 2012).
Besides natural organic matter (NOM), which reacts with
all oxidants applied in water treatment, bromide (Br) is one of
the key components of the water matrix relevant in oxidation
processes; it can be oxidized to hypobromous acid (HOBr, Eqs.
(1) and (7)) and other bromine species. Various studies have
illustrated that the presence of bromide during oxidation
processes can have a significant impact: the formation of
bromine can accelerate the transformation of undesired
compounds (Haag et al., 1984; Lee and von Gunten, 2009),
prevent the formation of undesired by-products (Criquet et al.,
2012) and affect the extent of formation of halogenated DBPs
(e.g., Amy et al., 1984; Bulloch et al., 2012; Chowdhury et al.,
2010; Cowman and Singer, 1996; Gallard et al., 2003; Hu
et al., 2006; Hua et al., 2006; Inaba et al., 2006; Jones et al.,
2012; Pan and Zhang, 2013; Richardson et al., 2003; Rodil
et al., 2012; Rook et al., 1978; Singer, 1994; Symons et al.,
1993; Zhang et al., 2005; Zhao et al., 2012) and N-nitrosodimethylamine (NDMA) (Le Roux et al., 2012; Luh and Marinas,
2012; von Gunten et al., 2010). Bromide often acts as a catalyst
in these reactions (Criquet et al., 2012; Duirk et al., 2008; Haag
et al., 1984; von Gunten et al., 2010). In Fig. 1, a schematic
overview of the role of bromide in oxidative water treatment
is given.
This review provides an assessment of the role of bromine
(the term bromine is used for all active bromine species,
mainly HOBr, OBr, and Br2, but also BrCl, Br2O, and BrOCl)
during oxidation processes with an emphasis on kinetics and
mechanisms of bromine reactions. This information may also
help to better understand the formation of (brominated)
disinfection by-products such as trihalomethanes (THMs) and
haloacetic acids (HAAs). However, their formation is not the
main focus of this review. Based on literature data, an overview of bromine speciation and bromine reactivity with
inorganic and organic compounds is presented. For relevant
inorganic substances (NH3, NO2 , SO3 2, I, CN, SCN) and
typical functional groups, such as activated aromatic systems,
amines, sulfur groups, and olefines, bromination kinetics and
mechanisms are described. Furthermore, for some organic

micropollutants relevant for urban water management, a


discussion on expected and observed bromine reactivities is
provided. Finally, a kinetic assessment is made on the
contribution of bromine to the overall reaction during chlorination of bromide-containing waters.

2.

Aqueous bromine chemistry

2.1.

Bromide levels in natural waters

Bromide is present in all fresh waters in concentrations in the


range of a few mg L1 to several mg L1 (Magazinovic et al.,
2004; von Gunten and Hoigne, 1992) and is non-toxic at
these concentrations (Flury and Papritz, 1993). The concentration of bromide in seawater is about 67 mg L1 (Millero,
1974). Local geological situations and salt-water intrusion in
coastal areas are responsible for the observed bromide levels
in natural waters (Agus et al., 2009). However, also anthropogenic activities such as potassium and coal mining or chemical industries may result in elevated bromide levels (Valero
and Arbos, 2010; von Gunten, 2003).

2.2.

Oxidation of bromide

2.2.1.

Water treatment

If bromide-containing water is treated with chlorine, bromide


is oxidized to hypobromous acid and hypobromite (HOBr/

HOBr

AO - + H +
ii)

H2O

i)

Oxidant

Br -

iii)

ABr

Oxidant

Fig. 1 e Role of bromide (BrL) in oxidative water treatment:


BrL is oxidized to bromine (HOBr for simplicity), which
reacts with organic or inorganic reaction partners (A),
leading to (i) brominated compounds (ABr), (ii) AOL via
hydrolysis of ABr or (iii) to other oxidized species (e.g.,
DBPs, A0 ) after further oxidation of ABr. In reactions (ii) and
(iii) bromide acts as a catalyst.

w a t e r r e s e a r c h 4 8 ( 2 0 1 4 ) 1 5 e4 2

17

OBr, Eqs. (1) and (2)) (Bousher et al., 1986; Farkas et al., 1949;
Kumar and Margerum, 1987):
HOCl Br /HOBr Cl


ClO Br /BrO Cl

k1 1:55  6:84  103 M1 s1


k2 9  104 M1 s1

(1)
(2)

In chloramination processes, bromide can also be oxidized


(Eqs. (3)e(6)). The reaction is first-order in monochloramine,
bromide, and the proton concentration and its main product is
assumed to be a mixed haloamine, NHBrCl (Bousher et al.,
1989; Gazda et al., 1993; Trofe et al., 1980).

NH2 Cl H $NH3 Cl

K3 2:8  101 M1

NH3 Cl Br /NH3 Br Cl

k4 5  6  104 M1 s1

NH3 Br NH2 Cl/NHBrCl NH4

fast

(3)
(4)
(5)

This leads to the following overall reaction (Eq. (6), the rate
constant is based on the decrease of NH2Cl, for the decrease of
Br/increase of NHBrCl the rate constant is half the value):


2NH2 Cl H Br /NHBrCl Cl NH4


k6 2:8  0:3  106 M2 s1

(6)

Bromine in bromochloramine is highly reactive (Valentine,


1986). Further species can be formed depending on pH
and the Br/Cl ratio (Bousher et al., 1989). The reaction of bromochloramine with monochloramine leads to the decomposition of both haloamines into N2, Cl, Br and H. Bromide
thus catalyzes the autodecomposition of monochloramine
(Vikesland et al., 2001). However, in the presence of NOM, the
accelerated autodecomposition is reduced, as bromamine
reacts mainly with NOM (Duirk and Valentine, 2007).
The oxidation of bromide by chlorine dioxide (ClO2) is very
slow (<0.05 M1 s1 at pH 8) and can be neglected under
typical water treatment conditions (Hoigne and Bader, 1994). It
has been shown that peracetic acid is also able to oxidize
bromide to HOBr (Booth and Lester, 1995). However, this is as
well a very slow process and is probably not relevant for
brominated disinfection by-product formation (DellErba
et al., 2007).
The formation of HOBr from Br by ozone has been studied
in the context of bromate (BrO3 ) formation (Haag and Hoigne,
1983; von Gunten, 2003; von Gunten and Hoigne, 1994).
Second-order rate constants for the reaction between bromide
and ozone have been determined (Eq. (7)) (Haag and Hoigne,
1983; Haruta and Takeyama, 1981; Liu et al., 2001).
O3 Br /BrO O2

k7 1:60  2:58  102 M1 s1

Fig. 2 e Bromate formation during ozonation of bromidecontaining waters (from von Sonntag and von Gunten
2012, with permission).

(7)

As shown in Fig. 2, HOBr plays an important role as an


intermediate in the bromate formation mechanism during
ozonation. Bromate has been classified as a potential carcinogen and a drinking water standard/guideline of 10 mg L1 has
been established worldwide (European Union, 1998; United
States Environmental Protection Agency, 2009; World Health
Organization, 2011). Bromate is formed by a complex multistep mechanism including ozone and hydroxyl radicals (OH,
Fig. 2) (von Gunten and Hoigne, 1994; von Sonntag and von
Gunten, 2012).

HOBr can also be formed in advanced oxidation processes


(AOPs) such as UV/H2O2 or O3/H2O2, where hydroxyl radicals
are the only or the main oxidants, via the formation of the
bromine atom (Br) (Pinkernell and von Gunten, 2001; von
Gunten and Oliveras, 1998). Hydrogen peroxide (H2O2), which
is present in such processes reduces HOBr to bromide with a
high second-order rate constant (Table 3, Section 3.2) and
therefore fully (UV/H2O2) or partially (O3/H2O2) inhibits further
oxidation of HOBr to bromate or its reaction with other compounds (Symons and Zheng, 1997; von Gunten and Oliveras,
1997).
In oxidative water treatment, bromide is thus oxidized by
the three commonly used oxidants chlorine, chloramine and
ozone. The oxidation of bromide by chlorine and ozone is
relatively fast, with hypochlorous acid reacting about ten
times faster with bromide than ozone, while the oxidation of
bromide by chloramine is a significantly slower process.

2.2.2.

Biological systems

Hypohalous acids are also formed in biological systems such


as the mammalian host defense system (Thomas et al., 1995;
Weiss et al., 1986). White blood cells are able to oxidize bromide (and chloride) with H2O2. These reactions are catalyzed
by myeloperoxidase (MPO) and eosinophil peroxidase (EPO),
two heme enzymes (Thomas et al., 1995). The hypohalous
acids react with different biological molecules of pathogens
but can e if produced in excess or unintentionally e also harm
the tissue of the respective organism itself, leading to inflammatory diseases (Pattison and Davies, 2006).

2.2.3.

Atmospheric waters

Bromine plays an important role in the troposphere at the


marine boundary layer in a number of processes such as
ozone depletion, reactions with organic compounds, HOx and
NOx, and oxidation of Hg(0) (Finlayson-Pitts, 2003; Saiz-Lopez
and von Glasow, 2012; Sommariva and von Glasow, 2012).
The formation of bromine in the marine boundary is complex,
involving reactions in the gas phase, in aerosols and on salt
surfaces such as the oxidation of bromide by ozone and HOCl
(formed by the reaction of chlorine monoxide with hydroperoxyl radical) (Finlayson-Pitts, 2003; Vogt et al., 1996).

18

w a t e r r e s e a r c h 4 8 ( 2 0 1 4 ) 1 5 e4 2

Table 1 e Reactions of bromine and chlorine species: Equilibrium and rate constants.a
Equilibrium constanta

No.

Reaction

(1)
(8)
(9)

HOCl Br $HOBr Cl
HOBr$OBr H
Br2 H2 O$HOBr Br H

(10)

Br2 Br $Br3

(11)
(12)
(13)

2HOBr$Br2 O H2 O
HOCl$OCl H
Cl2 H2 O$HOCl Cl H

6.31 M1h
pKa 7.47
(1.04  0.07)  103 M2i

(14)
(15)
(16)
(17)
(18)
(19)
(20)
(21)
(22)
(23)
(24)

BrCl H2 O$HOBr Cl H


BrCl H2 O$HOCl Br H
2BrCl$Cl2 Br2
BrCl Cl $Cl2 Br
Cl2 Cl $Cl3
Cl2 Br $BrCl2
BrCl Cl $BrCl2
Br2 Cl $Br2 Cl
BrCl Br $Br2 Cl
HOBr HOCl$BrOCl H2 O
2HOCl$Cl2 O H2 O

(1.3  0.1)  104 M2


8.7  1010 M2
(7.6  1.7)  103
9.1  107
(1.8  0.2)  101 M1e
4.2  106 M1
3.8  0.3 M1j
1.3  0.3 M1e
(1.8  0.2)  104 M1e
3.47  101 M1h
1.51  102 M1h,k

Rate constanta
k1 (1.55e6.84)  103 M1 s1b

1.5  10
See Table 2
(6.1  0.1)  109 M2c

k9 9.7  101 s1d


k9 (1.6  0.2)  1010 M2 s1d
k10 9.6  108 M1 s1g
k10 5.5  107 s1g

(1.61  0.03)  101 M1e,f

k13 (2.23  0.06)  101 s1i


k13 (2.14  0.08)  104 M2 s1i
k14 (3.0  0.4)  106 s1
k15 1.32  106 M2 s1

k19 7.7  109 M1 s1e

k22 > 108 M1 s1e

a Liu and Margerum (2001) and references within unless otherwise indicated, m 1 M, 25  C.
b Bousher et al. (1986), Farkas et al. (1949), Kumar and Margerum (1987).
c Beckwith et al. (1996), m 0.5 M, 25  C; Toth and Fabian (2004): (7.17  0.04)  109 M2, m 1 M, 25  C; for an overview of K at various ionic
strengths and reported values, see Beckwith et al. (1996).
d Beckwith et al. (1996), m 0.5 M, 25  C; Eigen and Kustin (1962): k9 110 s1, k9 1.6  1010 M2 s1, m 0.1 M, 20  C.
e Wang et al. (1994), m 1 M, 25  C.
f Toth and Fabian (2000): (1.93  0.12)  101 M1, m 1 M, 25  C; Ershov (2004): 1.75  101 M1.
g Ershov (2004); other values reported: k10 (1.5  0.4)  109 M1 s1, k10 (5  1.3)  107 s1 (Ruasse et al., 1986).
h Sivey et al. (2013).
i Wang and Margerum (1994), m 0.5 M, 25  C; an overview of reported values is also given there.
j Wang et al. (1994): 6.0  0.3 M1, m 1 M, 25  C.
k Roth (1929): 8.71  103 M1 at 19  C.

Furthermore, similar to the discussion in Section 2.3, species


such as Br2 and BrCl are also formed in aerosols and in exchange with the gas phase.

disproportionation (Beckwith and Margerum, 1997). The


overall disproportionation can be formulated as
3HOBr/BrO3  2Br 3H

(27)

2.3.
Speciation of bromine in aqueous solution in the
absence and presence of chlorine
The reactions occurring in the CleHOCleBr system are
compiled in Table 1. Figs. 3 and 4 show the reactions/equilibria
of bromine species in the absence and presence of chlorine,
respectively.
According to Fig. 3, HOBr undergoes acidebase speciation
(Eq. (8), Table 1). Table 2 lists reported pKa values, with 8.8  0.1
being selected for low ionic strength.
HOBr disproportionates leading to the formation of bromide and bromate (Fig. 3). The reaction occurs in two steps
according to Eqs. (25) and (26) (Br(-I): Br, Br(I): HOBr/BrO,
Br(III): HBrO2/BrO2 , Br(V): HBrO3/BrO3 ) (Beckwith and
Margerum, 1997).
BrI BrI/BrIII BrI
BrI BrIII/BrV BrI

k25 2  103 M1 s1


faster than 25

(25)
(26)

The reaction of HOBr with bromite (Eq. (26)) is faster


than reaction (25), the latter thus dominating the

Br2O
HOBr
HOBr/OBr OBr - + H +

HOBr
(CuO)

Br -

HOBr/OBr BrO2 -

(CuO)

Br BrO3 -

Br -, H +
Br2
Br Br3-

Fig. 3 e Bromine species and their formation pathways in


absence of ClL and HOCl: HOBr dissociates with a pKa of 8.8
(see Table 2). At low pH, Br2 is formed, which can react
further to Br3 L. HOBr may form Br2O in a self-reaction.
Disproportionation of HOBr leads to Br L, BrO2 L and finally
BrO3 L. This is a slow process unless catalyzed by CuO or
other corrosion products.

19

w a t e r r e s e a r c h 4 8 ( 2 0 1 4 ) 1 5 e4 2

BrOCl

Br2O

HOCl

HOBr

Br -

OBr - + H +

HOBr

Cl2O
HOCl

OCl - + H +

BrCl

HOCl
Cl -, H +

Cl3-

Cl -

Cl -

Br -, H +
Br -

Cl Br2Cl -

BrCl2-

Cl -

Br Cl2

Br2

Br -

Br3-

Fig. 4 e Overview of the formation pathways and the speciation of all chlorine, bromine and brominechlorine species in the
ClLeHOCleBrL system. Equilibria and rate constants are compiled in Table 1. HOBr and HOCl disproportionations have
been neglected since they are only relevant at high concentrations or for very long reaction times.

Disproportionation is fastest in the pH range 3e8, with a


maximum at pH 7.3 (Beckwith and Margerum, 1997; Chapin,
1934). Above pH 8, disproportionation is significantly slower
due to the deprotonation of HOBr, while initial steps of the
disproportionation are thought to be reversible below pH 3,
thus decreasing the rate of reaction (Beckwith and Margerum,
1997). Beckwith and Margerum (1997) gave an overview of
reported rate constants at various pH values. It is suggested
that the disproportionation is accelerated by bases, metals
and light (Beckwith and Margerum, 1997). Under acidic conditions (pH < 4), molecular bromine is formed in addition to
bromate in the disproportionation process (Eq. (28)) (Beckwith
and Margerum, 1997).
5HOBr$2Br2 BrO3  2H2 O H

K28 1  1010 M1

(28)

Due to the low bromine concentration and the low rate


constant, disproportionation is a very slow process under
typical water treatment conditions and bromate formed by
this pathway is negligible. However, in case of concentrated
hypochlorite solutions containing bromide, disproportionation becomes important and can lead to a dosing of bromate
during chlorination of drinking water (Weinberg et al., 2003).

Table 2 e Dissociation constants for HOBr.


Reported
Ionic
Temperature
pKa
strength
[ C]
[M]
8.76
8.8 0.1*
8.59
7.87
8.7
8.7

0.02e0.15
0.5
1
1
n.a.
n.a.

n.a. not available.


* Selected value.

20
25
25
25
20
n.a.

Reference

(Haag and Hoigne, 1983)


(Troy and Margerum, 1991)
(Gerritsen et al., 1993)
(Lahoutifard et al., 2002)
(Shilov, 1938)
(Farkas and Lewin, 1950)

Furthermore, it has been observed that cupric oxide (CuO), a


corrosion product of copper pipes, catalyzes the disproportionation of hypobromous acid (Liu et al., 2012). This may lead
to elevated bromate concentrations during chlorination of
bromide-containing waters if copper pipes are used in drinking water distribution systems (Liu et al., 2012).
With the equilibria listed in Table 1 and illustrated in Fig. 4,
the chlorine and bromine equilibrium-speciation as a function
of pH was calculated using PHREEQC (Parkhurst and Appelo,
1999) and based on a model by Korshin (2011). Three
different conditions have been considered: water in absence
of chlorine with 99% of bromide oxidized to bromine (Fig. 5a),
fresh water after chlorination with 2 mg L1 total active
chlorine (Fig. 5b) and seawater after chlorination with
2 mg L1 total active chlorine (Fig. 5c). The fresh water system
contained 100 mg L1 bromide and 10 mg L1 chloride, which
represents average fresh water concentrations for both ions.
In seawater, concentrations of 67 mg L1 and 20 g L1 were
assumed for bromide and chloride, respectively. Seawater is
oxidized/disinfected, e.g., for treatment of ballast water, aqua
cultures or seawater aquaria (Goncalves and Gagnon, 2011;
Grguric and Coston, 1998; Jorquera et al., 2002; Werschkun
et al., 2012). In absence of chlorine (Fig. 5a), HOBr and OBr
dominate the system. Br2 and (to a much lesser extend) Br3 
gain importance under acidic conditions with their concentrations depending on the residual bromide concentration.
The concentration of Br2O is proportional to [HOBr]2. In chlorinated fresh water systems (Fig. 5b), HOBr and OBr remain
the dominant bromine species. In the pH range relevant for
water treatment (pH 6e8), the concentrations of other species
such as BrCl, BrOCl, Br2O and Br2 are five to eight orders of
magnitude lower than the HOBr concentration. Due to their
low concentrations, their reactivity with compounds that
react quickly with HOBr can normally be neglected under
water treatment conditions, even though these species can
also be very reactive (Odeh et al., 2004; Sivey et al., 2013).
BrCl2 , Br3  and Br2Cl are not relevant under the conditions
considered here. In the case of chlorinated seawater (Fig. 5c),

20

w a t e r r e s e a r c h 4 8 ( 2 0 1 4 ) 1 5 e4 2

a)

-3
-5
-7
-9
-11
-13
-15

b)

log (concentration, M)

-3
-5
-7
-9
-11
-13
-15

c)

-3
-5
-7
-9
-11
-13
-15
4

10

11

12

pH
Fig. 5 e Concentrations of halogen species as a function of pH in (a) fresh water containing 100 mg LL1 bromide
([1.25 3 10L6 M), of which 99% is oxidized to bromine, without active chlorine, (b) fresh water containing 100 mg LL1
bromide ([1.25 3 10L6 M) and 10 mg LL1 chloride ([2.82 3 10L4 M), which is chlorinated with 2 mg LL1 ([2.82 3 10L5 M)
active chlorine, (c) seawater containing 67 mg LL1 bromide ([8.38 3 10L4 M) and 20 g LL1 chloride ([0.56 M), which is
chlorinated with 2 mg LL1 ([2.82 3 10L5 M) active chlorine. Calculations were done using PHREEQC (Parkhurst and Appelo,
1999), based on the equilibrium constants in Table 1. The shaded area represents the typical pH range for water treatment.

the concentrations of most bromine species are higher than in


fresh water due the higher natural abundance of chloride and
bromide. HOBr and OBr are still the most important bromine
species in the pH range 6e8. Yet, other bromine species
(i.e. Br2, Br2Cl, BrCl, Br3 , BrCl2  and Br2O) gain importance
and their concentrations are only about two to four orders of
magnitude lower than the HOBr concentration at pH 7. At
pH < 5, Br2 and Br2Cl become the dominant species.

3.
Reactions of HOBr/OBrL with inorganic
and organic compounds
3.1.

Kinetic considerations

In most cases, the oxidation of 1 mol of a compound C by h


moles of oxidant Ox (e.g., bromine) resulting in 1 mol of a

21

w a t e r r e s e a r c h 4 8 ( 2 0 1 4 ) 1 5 e4 2

product P can be described by second-order kinetics, firstorder with respect to each reaction partner (von Sonntag
and von Gunten, 2012):
1 dOx
dC dP


kox;C  Ox  C

h dt
dt
dt

kapp kHOBr;C  aHOBr  1  aCH kOBr ;CH  1  aHOBr  aCH


kHOBr;CH  aHOBr  aCH kOBr ;C  1  aHOBr

(29)

where kox,C is the second-order rate constant [M1 s1].


In case compound C and/or the oxidant undergo acidebase
equilibria, the speciation has to be taken into account. If
hypobromous acid reacts with a compound that does not
speciate, the oxidation of the compound by the protonated
form (HOBr) and the anion (OBr) has to be considered in the
rate law, which leads to


dC
kHOBr;C  HOBr  C kOBr ;C  OBr  C

dt

(30)

with kHOBr,C and kOBr ;C being the species-specific rate constants. Based on the pH and the dissociation constant Ka, the
degree of dissociation aHOBr (i.e. the protonated fraction) can
be defined (Eq. (31)).
aHOBr

accordingly (Eq. (34), (Criquet et al., 2012; von Gunten and


Oliveras, 1997)).

H 

Ka  H 

(31)

 1  aCH
(34)
In the following sections, rate constants for the reaction of
bromine with various classes of compounds are discussed.
Besides the kinetics for the reactions with HOBr and OBr, rate
constants at lower pH, indicated by HOBr H, have often
been proposed in literature. It has to be noted that these reactions may also reflect the reactivity of Br2, which is formed
under these conditions (Eq. (9), Fig. 5a). Furthermore, in
certain cases rate constants for Br2 and Br3  are also given.

3.2.

Reactions of bromine with inorganic compounds

Table 3 provides a compilation of rate constants for the reaction of HOBr and OBr with various inorganic compounds.

3.2.1.

Eq. (30) then becomes:



dC
kHOBr;C  aHOBr kOBr ;C  1  aHOBr  HOBrtot  C

dt
(32)
The expression in brackets represents the apparent secondorder rate constant kapp, which is pH-dependent:
kapp kHOBr;C  aHOBr kOBr ;C  1  aHOBr

(33)

If compound C speciates as well (as CH/C), the rate law


and the apparent rate constant have to be extended

Ammonia

The reaction of HOBr with ammonia to NH2Br is very fast (Eq.


(35)).
HOBr NH3 $NH2 Br H2 O k35 4  7:5  107 M1 s1

(35)

Judging from two conflicting equilibrium constants that have


been reported (Soulard et al., 1981; Tremblay-Goutaudier et al.,
1994), the rate constant for the back reaction should be in the
order of 1e103 s1. In a model, a value of 1.5  103 s1 was
estimated (Haag and Lietzke, 1980), while a fitted hydrolysis rate
constant of 7.5  106 M1 s1 which is first-order in OH was
used in another model by Pinkernell and von Gunten (2001).

Table 3 e Rate constants for the reaction of inorganic compounds with bromine.
Compound pKa

Species-specific rate constants


kHOBrH
 2 1 
M s

Br
BrO
ClO2
CN
HO2
I
OI
HOI
IO2
NH3
NH2Cl
NO2
O2
SCN
SO32

k(HOBr)
[M1 s1]

kOBr
 1

M s1

Apparent rate Temperature


constant at pH 7
kapp
[M1 s1]

Reference

[ C]

See Table 1
8.8
2.0
9.2
11.6
10
10.4
9.3
1.4
3.4
4.8
1.1
7.2

Disproportionation:
See Section 2.3
9.7  101
(4.2  0.9)  109
(7.6  1.3)  108
(5.0  0.3)  109
1.9  106
Not significant
High
(7.5  0.4)  107
(4  1)  107
(2.9  0.1)  105
(1.4  1)  104
3.5  109
2.3  109
(5  1)  109

(5.7  0.4)  107


(6.8  0.4)  105
1.8  103
High
(7.6  0.4)  104
(2.2  0.1)  104

3.8  104
(1.0  0.1)  108

9.6  101
2.6  107
1.9  104
4.9  109
7.3  102

4.1  105
2.2  105
2.8  105
1.4  104
3.4  109
2.3  109
1.9  109

25
25
25
25
24  1
24  1
24  1
20
25
25

18
25

(Furman and Margerum, 1998)


(Gerritsen et al., 1993)
(von Gunten and Oliveras, 1997)
(Troy and Margerum, 1991)
(Criquet et al., 2012)
(Criquet et al., 2012)
(Criquet et al., 2012)
(Wajon and Morris, 1982)
(Inman and Johnson, 1984)
(Gazda and Margerum, 1994)
(Lahoutifard et al., 2002)
(Schwarz and Bielski, 1986)
(Nagy et al., 2006)
(Troy and Margerum, 1991)

22

w a t e r r e s e a r c h 4 8 ( 2 0 1 4 ) 1 5 e4 2

The second-order rate constant for the reaction of HOBr


with NH3 is approximately three orders of magnitude higher
than the corresponding rate constant for OBr (Table 3), which
is due to the higher electrophilicity of HOBr compared to OBr.
The activation energies for these reactions are 15.7 (HOBr) and
48.4 kJ mol1 (OBr) (Wajon and Morris, 1982). The mechanism
of the reaction between HOBr and ammonia is an electrophilic
substitution of H(I) in ammonia by Br(I), forming monobromamine (Wajon and Morris, 1982). For comparison, the speciesspecific rate constant for the reaction of HOCl with ammonia
is about one order of magnitude smaller in the order of
106 M1 s1 (Deborde and von Gunten, 2008).
Fig. 6 shows the apparent second-order rate constants for
the reaction of chlorine and bromine with ammonia as a
function of pH. Since the main reacting species are NH3 and
HOBr or HOCl, respectively, the pH of the maximum apparent
second-order rate constants equals the average of the pKa
values of NH4 and the hypohalous acid (pH 9.1 for HOBr/NH3
and 8.4 for HOCl/NH3). At pH 7, the difference between the
apparent second-order rate constants of the two reactions
(HOBr NH3 vs. HOCl NH3) is a factor of around 30, at pH 8 a
factor of around 80. The fraction of ammonia fHOBr;NH3
reacting with HOBr at a certain pH can be calculated by Eq.
(36):
fHOBr;NH3

kapp;HOBr  HOBrtot
kapp;HOBr  HOBrtot kapp;HOCl  HOCltot

(36)

with kapp,HOBr and kapp,HOCl being the apparent second-order


rate constants for the reaction of HOBr and HOCl with
ammonia. Under the conditions in Fig. 5b (water containing
1.25  106 M bromide, chlorinated with 2.8  105 M active
chlorine) and assuming that bromide was oxidized to HOBr
prior to ammonia addition, at pH 7 around half of the
ammonia present reacts with HOBr, while at pH 8, the fraction
increases to around 80%. However, in the more realistic case
of a water containing both bromide and ammonia at the same
time when chlorinated, the majority of chlorine will react
immediately with ammonia (fraction of ammonia reacting
with chlorine >99%, assuming the same concentrations as

1.0

NH3

log (kapp, M-1 s-1)

6
5

ine

om

br

0.5

k ap

ine

lor

ch

k ap

HOCl

HOBr

Relative species concentration

0.0
4

10

12

pH

Fig. 6 e Reaction of HOX with ammonia: pH dependence of


reactive species concentration and apparent second-order
rate constants for the reaction of HOBr and HOCl with
ammonia.

before and with [NH3]tot [HOCl]tot at neutral pH), suppressing


the oxidation of bromide in the first place. Only if the
ammonia concentration is lower than the active chlorine
concentration, a relevant bromine formation will be observed.
The reactions of the formed chloramine with bromide are
discussed above (Section 2.2.1).
Due to its high reactivity with HOBr, ammonia addition has
been suggested as a measure for bromate mitigation during
ozonation of bromide-containing waters, because ammonia
masks HOBr, a decisive intermediate species in the bromate
formation pathway (Haag et al., 1984; Pinkernell and von
Gunten, 2001). Based on the masking of HOBr as NH2Br, the
chlorineeammonia process has been developed (Neemann
et al., 2004). In this process, a pre-chlorination to oxidize
bromide to HOBr is carried out, followed by the addition of
ammonia to mask HOBr as NH2Br before ozone is added.
Compared to conventional ozonation, bromate formation can
be reduced by up to a factor of ten by this process (Buffle et al.,
2004). This process can be applied if high ozone exposures are
required (e.g., for Cryptosporidium parvum oocysts inactivation)
in waters with elevated bromide levels (Buffle et al., 2004).
Bromine can react further with NH2Br to form dibromamine (NHBr2) and eventually tribromamine (NBr3) (GalalGorchev and Morris, 1965). The relative concentrations of
the three bromamine species after reaction depend on the pH
and the N/Br ratio (Galal-Gorchev and Morris, 1965; Johnson
and Overby, 1971). Dibromamine is also formed through the
disproportionation of monobromamine to dibromamine and
ammonia (Eq. (37)) and analogously tribromamine can be
formed (Inman and Johnson, 1984; Lei et al., 2004; Soulard
et al., 1981).
2NH2 Br H $NHBr2 NH4 K37 8  107  5  109 M1

(37)

It was reported that the disproportionation of monobromamine/formation of dibromamine is catalyzed by phosphate


buffer and the following rate law for the pH range 7e8.5 was
proposed (Inman and Johnson, 1984) (Eq. (38)):
 
 
dNHBr2 =dt k38a H PO4 Br2tot k38b H Br2tot

(38)

[Br]tot: sum of HOBr and OBr, [PO4]: total phosphate concentration; with k38a 9.9  1011 M3 s1 and k38b 2.4  108 M2
s1 (Inman and Johnson, 1984). The catalytic effect of acids on
the disproportionation of monobromamine was confirmed,
however with different rate constants in another study (Lei
et al., 2004). The measured rate constants for the forward reaction ranged from 0.5 M1 s1 (with H2O as a catalyst) to
5  108 M2 s1 (H catalysis) and for the backward reaction
rate constants varied between 1 (with H2O as a catalyst) and
1  109 M2 s1 (H catalysis).
While the disproportionation of monobromamine/formation of dibromamine is a reversible process (Eq. (37)), further
decomposition reactions of monobromamine and dibromamine (Eqs. (39) and (40)) were found to be irreversible and
base-catalyzed (Lei et al., 2004).
NH2 Br NHBr2 /products

(39)

2NHBr2 /products

(40)

23

w a t e r r e s e a r c h 4 8 ( 2 0 1 4 ) 1 5 e4 2

The corresponding rate constants were determined in H2O


and for catalysis by OH, PO4 3, CO3 2, NH3 and HPO4 2. The
rate constants without catalysis are 8.9 M1 s1 (Eq. (39)) and
6.2 M1 s1 (Eq. (40)); the values for OH catalysis are
4.1  107 M2 s1 (Eq. (39)) and 8.3  104 M2 s1 (Eq. (40)), the
rate constants for catalysis with other bases lie between the
two values for H2O and OH (Lei et al., 2004).
Little is known about the rate and mechanism of tribromamine formation (Hofmann and Andrews, 2001); its decomposition is discussed by Inman et al. (1976) and Johnson and
Overby (1971).

3.2.2.

Reactivity of bromamines

Not much information is available on the reactivity of bromamines with inorganic and organic compounds. The reactivity
of bromine (Br(I)) with NOM in a mixed bromideechloramine
system was investigated by Duirk and Valentine (2007).
Another study discusses the kinetics of the reaction between
mono- and dibromamine and CN (Eqs. (41) and (42)). The
product of these reactions is in both cases cyanogen bromide
(BrCN) (Lei et al., 2006).

>105 M1 s1 (pH 7.2) for glutathione, 3.6  103 M1 s1 for
ascorbate (pH 6.9) and 1.02  102 M1 s1 for Fe(III)cytochrome
c (at pH 7.2) (Prutz et al., 2001). Under acidic conditions, the
reaction of monobromamine with iodide forms IBr and
ammonia (Prutz et al., 2001). For the reactivity of organic
bromamines, see Section 3.3.2.

3.2.3.

Halides and other inorganic anions

Based on the equilibrium constant and the rate constant for


the reaction of HOCl with Br, it can be assumed that the back
reaction, i.e. the reaction of HOBr with Cl, is quite slow with a
second-order rate constant in the order of 102 M1 s1 (see
Table 1). As far as the reaction of HOBr with the other halides
is concerned, HOBr reacts with Br to form Br2 in acidic solution (Eq. (9), Table 1) and oxidizes I in a very fast process
(Eq. (44), Table 3) leading to the formation of the intermediate
IBr. IBr hydrolyzes quickly to hypoiodous acid (HOI/OI) with
hydrolysis constants of 8  105 s1 with H2O and
6  109 M1 s1 with OH (Troy et al., 1991).
HOBr I /HOI Br

k44 5:0  0:3  109 M1 s1

NH2 Br CN H2 O/NH3 BrCN OH


k41 2:63  104 M1 s1
NHBr2 CN H2 O/NH2 Br BrCN OH
1

k42 1:31  10 M
8

1

(41)

(42)

The rate constants of Eqs. (41) and (42) are five to six orders of
magnitude higher than the analogous rate constants of chloramines (Lei et al., 2006). Compared to HOBr, the second-order
rate constant for the reaction of NHBr2 with CN is only a
factor of about 30 lower (Table 3). The reaction of monobromamine with cyanide is general-acid-catalyzed, while no
catalytic effect was found for the reaction with dibromamine
(Lei et al., 2006). Third-order catalysis rate constants for H,
H2PO4 , HPO4 2, H3BO3 and NH4 have also been determined
(Lei et al., 2006). For the hydrolysis of BrCN, see Section 3.2.3.
The reactions of bromamines with acetic acid forming
haloacetic acids were studied by Pope and Speitel (2008). It
was found that bromamines are significantly more reactive
with acetic acid than their chlorine analogs (Pope and Speitel,
2008).
The reaction of monobromamines with ozone yields bromide and nitrate (Eq. (43)) (Haag et al., 1984).
NH2 Br 3O3 /Br NO3  3O2 2H
k43 4:0  1:0  101 M1 s1

(43)

For the analogous reaction of dibromamine with ozone, the


second-order rate constant is a factor of four smaller (Haag
et al., 1984). No rate constant has been determined for tribromamine. Because bromamines react faster with ozone
than ammonia/ammonium at circumneutral pH, an enhanced
ammonia oxidation during ozonation of bromide-containing
waters has been proposed (Haag et al., 1984).
The reaction of monobromamine with biological substrates has been studied in the context of inflammatory diseases. The apparent second-order rate constant of NH2Br was
found to be 1.8  103 M1 s1 (pH 7.3) for methionine,

(44)

HOI/OI can be further oxidized to iodite (IO2 , Eq. (45)), which


is then again quickly oxidized to iodate (IO3 , Eq. (46)) by HOBr
(Criquet et al., 2012).
HOBr OI /IO2  Br H

k45 1:9  106 M1 s1

(45)


HOBr IO
2 /IO3 Br H

faster than 45

(46)

The oxidation of I to iodate by HOBr is preferable in drinking


water treatment because iodate is non-toxic (Burgi et al., 2001)
and concurrently the formation of iodo-organic DBPs is
minimized (Criquet et al., 2012). In the sequence of iodide
oxidation to iodate (Eqs. (44)e(46)), the oxidation of hypoiodite
(Eq. (45)) is the rate-determining step (Criquet et al., 2012). The
rate constant for the reaction of HOBr with HOI is considerably
lower than the one with OI, because of the higher nucleophilicity of OI compared to HOI. Analogous reactions also
occur with OBr, however, the corresponding second-order
rate constants are about three to four orders of magnitude
lower (Table 3).
Similar to the reaction of HOBr with iodide, reactions of
HOBr with sulfite (SO3 2), cyanide (CN) and thiocyanate
(SCN) are also all very rapid and close to the diffusion limit
(Gerritsen et al., 1993; Nagy et al., 2006; Troy and Margerum,
1991). Typically, the reaction of HOBr is one to nearly five orders of magnitude faster than OBr (Table 3). The proposed
pathway for the reactions with I, SO3 2and CN is a Br
transfer resulting in IBr, SO3Br and CNBr as intermediates
(Fig. 1, (i)), which subsequently hydrolyze to HOI, SO4 2 and
OCN, respectively (Eq. (47), Fig. 1, (ii)) (Gerritsen et al., 1993;
Troy and Margerum, 1991).
HOBr A $ABr OH  $

Br
AOH
e
AO H

(47)

SO3Br hydrolyzes to SO4 2 with a first-order rate constant


of 230 s1 (0  C) (Troy and Margerum, 1991) and for CNBr hydrolysis rate constants of 0.53 M1 s1 with OH and
7.5  103 M1 s1 with CO3 2 were reported, resulting in the

24

w a t e r r e s e a r c h 4 8 ( 2 0 1 4 ) 1 5 e4 2

formation of OCN and Br (Gerritsen et al., 1993). For the fast
reaction of SCN with HOBr two parallel pathways of generalacid-catalyzed Br transfer and direct reaction of HOBr were
proposed. The product of these reactions is OSCN (Nagy
et al., 2006).
Species-specific second-order rate constants for the reactions of bromine with the anions discussed above (I, SO3 2,
CN, SCN) are 2e100 times higher than those of chlorine
kHOCl;I 1:4  108 ; kHOCl;SO 2 7:6  108 ; kHOCl;CN
3
1:2  109 and kHOCl;SCN 2:3  107 M1 s1 (Ashby et al.,
2004; Deborde and von Gunten, 2008). The differences are
more pronounced (up to a factor of around 350) when
comparing the respective apparent second-order rate constants at pH 7 and 8 (Table 4).
For nitrite, a rate constant of 1.4  104 M1 s1 in 1 M
Na2SO4 was reported by Lahoutifard et al. (2002). This rate
constant was about a factor of 3 smaller when the ionic
strength was adjusted with NaCl instead of Na2SO4
(Lahoutifard et al. 2002). However, the bromine speciation
changes significantly under these conditions (see Fig. 4,
Table 1). The reaction pathway proposed by Lahoutifard et al.
(2002) consists of a fast formation of BrNO2 (nucleophilic
attack of Br), which then e in the rate-determining step e
reacts with NO2  to N2O4. N2O4 hydrolyzes to NO2  and NO3 
(Lahoutifard et al., 2002). This partly agrees with a mechanism
suggested for the oxidation of NO2  by HOCl (Johnson and
Margerum, 1991). After the formation of ClNO2 two pathways were observed: either N2O4 formation and subsequent
hydrolysis or decay of ClNO2 yielding Cl and NO2 , where
NO2 hydrolyzes to NO3  (Johnson and Margerum, 1991).
The rate constants for the reaction of HOBr with HO2  (von
Gunten and Oliveras, 1997) and O2  (Schwarz and Bielski,
1986) are very high (Table 3). The reaction with HO2  is of
particular interest in H2O2-based AOPs (i.e. O3/H2O2, UV/H2O2)
as it minimizes the formation of bromate and brominated
DBPs as discussed in Section 2.2.1.

3.2.4.

Iron, manganese and arsenic

The oxidation of Fe(II) and Mn(II) by bromine has not received


much attention in literature except in the context of chemical
oscillation reactions (e.g., Chou et al., 1993; Hegedus et al.,
2006; Melichova et al., 1995, 2001; Sasaki, 1990). Also for the

Table 4 e Comparison of apparent second-order rate


constants for the reaction of chlorine and bromine with
selected anions at pH 7 and pH 8 (for references of
bromine rate constants see Table 3; chlorine rate
constants from Ashby et al., 2004 and Deborde and von
Gunten, 2008).
kapp [M1 s1]

Compound pKa

Apparent rate
constant at pH 7

Apparent rate
constant at pH 8

Chlorine Bromine Chlorine Bromine




CN
I
SCN
SO3 2

9.2
10
1.1
7.2

5.8  106
1.1  108
1.8  107
2.2  108

2.6  107
4.9  109
2.3  109
1.9  109

1.7  107
3.4  107
5.5  106
1.6  108

2.2  108
4.3  109
2.0  109
3.7  109

oxidation of arsenite (As(III)) by bromine, no rate constant is


available in literature.
In general, it can be assumed that an increase in pH will
result in a faster reaction of Fe(II) due to the increasing concentration of Fe(II) hydroxy complexes, which are more susceptible to oxidation (Deborde and von Gunten, 2008; King,
1998).
In the case of manganese it is known that, besides the slow
direct oxidation of Mn(II) by bromine, mostly heterogeneous
Mn(III,IV) oxide-catalyzed reactions occur: Mn(II) adsorbs to
Mn(III,IV) oxide, which results in a faster oxidation of Mn(II)
(Allard et al., 2013; Melichova et al., 2001). This autocatalytic
mechanism has also been observed for the analogous Mn(II)
oxidation by chlorine (Hao et al., 1991 in Deborde and von
Gunten, 2008). It has been suggested that the oxidation of
Mn(II) by bromine occurs to a significant extent through BrCl
and Br2O (Allard et al., 2013). This is thus a case where
brominated electrophiles other than HOBr need to be taken
into account (see Section 2.3).
Because no rate constants for the reaction of Fe(II) and
As(III) with bromine are available, the corresponding rate
constants for chlorine may give an indication (Fe(II):
1.7  104 M1 s1 at acidic pH (Folkes et al., 1995); As(III):
4.3  103 M1 s1 for As(OH)3, 5.8  107 M1 s1 for As(OH)2O
and 1.4  109 M1 s1 for As(OH)O2 2 (Dodd et al., 2006)). Based
on these rate constants, it can be expected that the corresponding rate constants for the reaction of bromine with Fe(II)
and As(III) are very high.

3.3.

Reactions of bromine with organic compounds

3.3.1.

Aromatic compounds

Table 5 summarizes published rate constants for the reaction


of bromine with various aromatic compounds. Most rate
constants are available for the reaction of bromine with phenols, since an important class of taste and odor compounds is
attributed to the products of these reactions (bromophenols,
medicinal taste and odor) (Piriou et al., 2007). The odor
threshold concentrations are very low with 30 ng L1 and
0.5 ng L1 for 2-bromophenol and 2,6-bromophenol, respectively (Acero et al., 2005).
The rate constants for the reaction of bromine with phenol
and phenolic compounds are very high (Table 5). Yet, the
species-specific rate constants for the reaction with phenol
determined by Gallard et al. (2003) and the phenolic compounds determined by Guo and Lin (2009) seem to be slightly
overestimated due to the use of quench-flow systems, which
do not allow proper mixing for such fast reactions. In the
compilation in Table 5, a value based on new measurements
with an advanced quench-flow system has been added
(Criquet et al., in preparation).
Due to the acid/base speciation of both HOBr and phenol,
the apparent reaction rate constant varies with pH, similar to
that observed for the reaction of HOBr with NH3 shown above
(Eq. (35), Table 3, Fig. 6). A comparison of the species-specific
rate constants clearly shows that the protonated form,
i.e. HOBr has the highest reactivity with the phenolate ion
(Table 5). This can be explained by the higher electrophilicity
of HOBr compared to OBr and the increased electron density

Table 5 e Rate constants for the reaction of aromatic compounds with bromine.
Compound

Species-specific rate constants

pKa

k

2 =PhOH
2 =PhO 
 Br1
 kBr1
M s1
M s1

4-Acetylphenol
N-Acetyl-tyrosine
p-Aminophenol
Anisole
Benzene

8.6

10.3

k(HOBr/PhOH)
[M1 s1]

kHOBr=PhO
M1 s1

kOBr =PhO
M1 s1

(1.0  0.4)  104 (4.1  0.6)  106

<2.0  103

8.9  104

4.9  103

5.4  108

Apparent rate constant Temperature


at pH 7 or at given pH
kapp
[ C]
[M1 s1]
1.1  105
(2.6  0.2)  105 (pH 7.2e7.5)
(5.8  0.2)  105 (pH 7.2e7.5)
3.5  105
5.2  101
<1  102 (pH 4)

1.1  105
2-Bromo-4-methylphenol
2-Bromophenol

1.0  10

3.9  10

6.2  10

9a

5.5  10

9.2

6.4  106
4.8  106

3.2  104

2-Chlorophenol
8.6
3-Chlorophenol
7.9b
4-Chlorophenol
9.4
m-Cresol (3-methylphenol) 10.1

7.2  106
6.5  104
7.9  106
(6.0  3.0)  103 (7.0  0.8)  106
(3.5  0.3)  108

o-Cresol (2-methylphenol) 10.3


p-Cresol (4-methylphenol) 10.3

<1.0  104

(9.6  0.4)  107


(2.1  0.5)  108

<3.0  105

1.2  104

8.9  105

4.2  104

4.5  104
1.2  105
(2.6  0.1)  105 (pH 7.2e7.5)
w1.2  105 (pH 7.8)
1.3  105

1.7  104

4.8  105

8.2  104

3.3  105

(3.0  2.0)  104


1.6  104

1.3  105
2.5  105
(2.4  0.2)  105 (pH 7.4)
5.2  104
(1.6  0.3)  105 (pH 7.2e7.5)

7.6  104

2.1  106
3.1  104

Cyclo(Serine-Tyrosine)
2,4-Dibromophenol
2,6-Dibromophenol
2,4-Dichlorophenol
2,6-Dichlorophenol
Hydroquinone
4-Hydroxybenzoic acid
3-(4-Hydroxyphenyl)
propionic acid (HPPA)
3-Methoxyphenol
4-Methoxyphenol
4-Nitro-3-methylphenol
4-Nitrophenol

7.8
w3.0  102

3.7  109

w5.0  102

2.7  109

6.7
7.9
7.0

(3.0  1.0)  104 (8.8  0.9)  105


3.5  104
4.5  105

9.5

(1.4  0.1)  107

6.8  105

9.7
10.2

7.5  105
9.2  103

7.2
<6.0  101

1.2  109a

6.5  108
(4.9  0.6)  107
3.2  108
8.8  106

7.5  102
(5.5  4.0)  104

5.6  102
6.4  103

1.9  105
1.0  106
3.2  104
2.7  105

3.6  106

20
23  2
22
20  2
25  2
25
25  2
25
23  2
25  2
22
20
22

(Gallard et al., 2003)


(Pattison and Davies, 2004)
(Pattison and Davies, 2004)
(Guo and Lin, 2009)
(Echigo and Minear, 2006)
(Pinkernell and von
Gunten, 2001)
(Tee et al., 1989)
(Pattison and Davies, 2004)
(Echigo, 2002)
(Tee et al., 1989)
(Echigo, 2002)
(Tee et al., 1989)
(Echigo, 2002)
(Guo and Lin, 2009)
(Gallard et al., 2003)
(Echigo and Minear, 2006;
Echigo, 2002)
(Echigo, 2002)
(Gallard et al., 2003)
(Pattison and Davies, 2004)
(Prutz et al., 2001)
(Acero et al., 2005)
(Tee et al., 1989)
(Acero et al., 2005)
(Tee et al., 1989)
(Gallard et al., 2003)
(Acero et al., 2005)
(Skaff et al., 2007)
(Echigo, 2002)
(Pattison and Davies, 2004)

25
20
25
25
25

(Guo and Lin, 2009)


(Echigo, 2002)
(Guo and Lin, 2009)
(Guo and Lin, 2009)
(Tee et al., 1989)

25
22
25  2
25
25  2
25
25  2
25
23  2
20

w a t e r r e s e a r c h 4 8 ( 2 0 1 4 ) 1 5 e4 2

4-Bromophenol

8.5

(6.4  0.3)  105 (pH 7.2e7.5)


2.2  105

23  3
22
37
25
20
20

Reference

(continued on next page)

25

26

Table 5 e (continued )
Compound

pKa

Species-specific rate constants



k

2 =PhOH
2 =PhO 
 Br1
 kBr1
M s1
M s1

Phenol

10.0

4.3  10
3-Phenylpropionic acid
Pyrene
2,4,6-Tribromophenol

p-Xylene

7.4

w1.2  10

kOBr =PhO
M1 s1

kHOBr=PhO
M1 s1

<5.0  102

(1.8  0.2)  108


(4.1  0.1)  107

<1.0  105

1.8  105
4.1  104

23  2
20

8.0  103

2.3  108

5.3  102

<1

6.6  107

3.5  104

2.3  105
5.0  102 (pH 4)
6.5  104

3.3  103

<1

Very slow (pH 7.2e7.5)


1.1  101
2.1  103

(1.4  0.1)  103

1

1.2  103
(6.4  0.6)  104 (pH 7.4)
(2.5  0.3)  106 (pH 7.4)
6.0  105

25
20
25
25
22
20
25  2
25
23  2
22
22
20

2  101 (pH 4)
1.1  104

20
20

9c

2.0  102
w1.7  103a

(1.3  0.1)  106

Reference

(Gallard et al., 2003)


(Echigo and Minear, 2006),
(Echigo, 2002)
(Guo and Lin, 2009)
(Pinkernell and von Gunten, 2001)
(Criquet et al., in preparation)
(Tee et al., 1989)
(Pattison and Davies, 2004)
(Hu et al., 2006)
(Acero et al., 2005)
(Tee et al., 1989)
(Gallard et al., 2003)
(Skaff et al., 2007)
(Skaff et al., 2007)
(Echigo and Minear, 2006;
Echigo, 2002)
(Pinkernell and von Gunten, 2001)
(Voudrias and Reinhard, 1988)

a These values may be overestimated due to polybromination (Tee et al., 1989).


b pKa used by Guo and Lin (2009) for the fitting of the rate constants. The CRC Handbook of Chemistry and Physics gives a value of 9.1 (Haynes, 2013).
c A second-order rate constant of 8.5  108 M1 s1 has been determined for the reaction of Br3  with PhO (Tee et al., 1989). However, this is not relevant under typical water treatment conditions (see
text).

w a t e r r e s e a r c h 4 8 ( 2 0 1 4 ) 1 5 e4 2

2,4,6-Trichlorphenol
Trolox
Ubiquinol-0
Vanillin

6.8
5.9
6.2

k(HOBr/PhOH)
[M1 s1]

Apparent rate constant Temperature


at pH 7 or at given pH
kapp
[ C]
[M1 s1]

w a t e r r e s e a r c h 4 8 ( 2 0 1 4 ) 1 5 e4 2

at the ortho and para position of the phenolate compared to the


phenol.
The mechanism of the reaction of bromine with phenols is
mainly an electrophilic substitution caused by the positive
partial charge (d) of the Br in HOBr (Acero et al., 2005). The
bromination occurs in ortho or para position. Statistically, the
ortho position is favored which leads to approximately 2/3
ortho substitution and 1/3 para substitution (Acero et al., 2005).
When all the ortho and para positions have been substituted by
a halogen, further halogenation leads to ring opening and
formation of THMs (Acero et al., 2005; Arnold et al., 2008;
Gallard and von Gunten, 2002).
While electrophilic substitution is the main reaction
mechanism of HOBr with phenolic compounds, for catechol
and hydroquinone electron transfer has been observed, which
leads to the formation of 1,2-benzoquinone and 1,4benzoquinone, respectively (Criquet et al., in preparation).
In addition to the rate constants of HOBr and OBr, secondorder rate constants for the reaction of Br2 with phenols and
phenolates (and some other compounds, not listed) have been
measured (Tee and Bennett, 1988; Tee and Iyengar, 1985; Tee
et al., 1985, 1989) (Table 5). The second-order rate constants
for the reaction of Br2 with phenol and phenolates are
extremely high, close to diffusion limitation. Furthermore, a
second-order rate constant has been determined for the reaction of Br3  with phenolate. This is, however, irrelevant
under any natural water conditions, as very high bromide
concentrations would be necessary for Br3  formation at a pH
for which there is sufficient overlap with phenolate (4 mM of
bromide at pH 6 for Br3  to contribute to 1% of the total reaction). In contrast, the reaction of Br2 with phenol can play a
role at low pH: In the case of a water initially containing
1.25 mM bromide (100 mg L1) of which 90% has been oxidized
(i.e. 1.25  107 M bromide remaining) at pH 6, around 1& of
the total phenol reacts with Br2, while at pH 5 it is already
around 10%.
The analogous reactions of HOCl with phenol and halophenols follow the same reaction pathway as observed for
HOBr (Deborde and von Gunten, 2008). Fig. 7 shows the correlation between second-order rate constants for the reaction
of HOCl kHOCl;PhO and HOBr kHOBr;PhO with
various phenolates (R2 0.998). The slope of the linear correlation is 3000, which represents the ratio between
kHOBr;PhO and kHOCl;PhO . This means that the speciesspecific rate constants of the reactions of HOBr with phenolate and halophenolates are on average around 3000 times
greater than the species-specific rate constants of the analog
reactions of HOCl.
Generally, a higher degree of halogenation results in lower
species-specific rate constants due to the electronwithdrawing effect of the halogens. However, with respect
to the apparent rate constant at near neutral pH, this effect is
in some cases compensated due to the lower pKa of the
halogenated phenols (Fig. 8). For example, the species-specific
rate constant for the reaction of HOBr with 2-bromophenol
(6.4  106 M1 s1) is more than one order of magnitude
higher than the corresponding rate constant for its reaction
with 2,6-dibromophenol (4.8  105 M1 s1). In contrast, the
apparent second-order rate constant for the reaction of HOBr
with 2,6-dibromophenol at pH 7 (3.3  105 M1 s1) is higher

27

than the apparent rate constant for the reaction with 2bromophenol at the same pH (2.2  105 M1 s1). This contraintuitive effect is explained by the considerably lower pKa of
2,6-dibromophenol compared to 2-bromophenol (6.7 vs. 8.5). A
comparison of the apparent second-order rate constants for
the reaction of bromine with various substituted and unsubstituted phenols as a function of pH is shown in Fig. 8.
Quantitative StructureeActivity Relationships (QSARs) between species-specific rate constants for oxidation reactions
of closely related compounds and substituent descriptor variables such as Hammett substituent constants (s) have been
successfully applied to predict and quantify rate constants for
oxidation of organic compounds (Canonica and Tratnyek,
2003; Lee and von Gunten, 2012). This empirical relationship
implies a linear correlation between the logarithm of the
species-specific rate constant (log(k)) and the Hammett substituent constant (s) calculated as the sum of the effects of the
different ring substituents. The parameter s reflects the effect
of the substitution on the ring on electron density by inductive
and resonance effects compared to the phenolate for which s
equals zero. From the compiled data of rate constants for the
reaction of HOBr with aromatic compounds (Tables 5 and 10),
a Hammett-type correlation was established (Eq. (48), Fig. 9).

log kHOBr=PhO 7:8  3:5Ss

(48)

Hammett constants from Hansch et al. (1991) and Lee and von
Gunten (2012) have been used. The negative slope is typical for
electrophilic reactions and its magnitude shows the sensitivity of the reaction to substituent effects (Hansch et al.,
1991). The correlation includes the rate constants of the reactions of HOBr with the dissociated species of phenol, ethinylestradiol (EE2, Table 10), and their chlorinated and
brominated derivatives and agrees well with correlations
found in other studies (Acero et al., 2005; Gallard et al., 2003).
All rate constants obtained by Guo and Lin (2009) as well as the
rate constants for phenol and p-cresol from Gallard et al.
(2003) have been excluded due to the relatively high uncertainty of these measurements (see above). Despite arriving at

Fig. 7 e Correlation of the species-specific rate constants


for the reaction of HOBr with various phenolates with the
corresponding species-specific rate constants of HOCl.

28

w a t e r r e s e a r c h 4 8 ( 2 0 1 4 ) 1 5 e4 2

Apparent second-order rate constant [M-1s-1], log scale

_________________________________________________

3
6

Phenol
(unsubstituted)
2-Bromophenol

2,6-Dibromophenol

2,4-Dibromophenol

4-Bromophenol

2,4,6-Tribromophenol

10

11

12

pH

Fig. 8 e Apparent second-order rate constants for the


reaction of bromine with phenol, mono-, di-, and
tribromophenols. At pH 7, the apparent second-order rate
constants of some of the brominated phenols are in the
same range or even higher than the corresponding rate
constant of (non-substituted) phenol, even though their
species-specific rate constants are more than one order of
magnitude lower. This effect is due to the lower pKa of the
Br-substituted compounds.

approximately the same correlation coefficients, also trichlorinated and tribrominated phenols were excluded from
the correlation since the preferred sites of substitution (ortho
and para position relative to the hydroxyl group) are not
available and the Hammett calculation does not take this
particular case into account.
Fig. 10 shows a comparative Hammett plot for the reaction
of various oxidants including HOBr with phenolates. It shows
that only ozone reacts faster than HOBr with rate constants
that are around one order of magnitude higher. The reactivity
of HOBr is similar to the one of chlorine dioxide. While HOBr
exhibits a much higher reactivity toward phenolic compounds
than HOCl (ca. three orders of magnitude), the two oxidants
show approximately the same dependence on the phenolic
substituents moieties (slope 3.0 and 3.5 for HOCl and HOBr,
respectively). HOI and to a lesser extend ferrate are more
sensitive to substitution effects, which is reflected in the
steeper slopes compared to the other oxidants (Lee et al.,
2005).
Compared to phenolic compounds, aromatic compounds
without activating or only slightly activating moieties such as
methoxy or methyl substituents (benzene, anisole and xylene)
have very low rate constants for the reaction with bromine
(Table 5).
Aromatic functional groups can make up a significant part
of NOM. The reaction of bromine with NOM typically shows
biphasic reaction kinetics: a fast initial phase followed by a
slower one (Echigo and Minear, 2006; Westerhoff et al., 2004).
The first phase has been attributed to the reaction with aromatic functional groups by means of specific UV-absorption

Fig. 9 e Correlation between the second-order rate


constants (log(k)) for the reaction of HOBr with dissociated
phenols and the Hammett constant (sigma, s). Correlation:
logkHOBr=PhOL [7:8L3:5 Ss, R2 [ 0.8. 2,4,6-tribromophenol
and 2,4,6-trichlorophenol were excluded from the
correlation (see text).

with apparent rate constants ranging in the order of


102e104 M1 s1 (rate constants based on C ) around neutral
pH, while the rate constants for the slower phase are around
100 times lower (Echigo and Minear, 2006; Westerhoff et al.,
1998, 2004). However, rate constants for the reaction with
NOM are difficult to compare due to the great variability of its
nature. It is unclear whether the dominant reaction mechanism is bromination (bromine atom transfer) or electron

Fig. 10 e Comparison of the correlations between the


second-order rate constants of the reaction of dissociated
phenols with HFeOL
4 , O3, ClO2, HOI, HOCl and HOBr as a
function of the Hammett constant (sigma, s). Reprinted
(adapted) with permission from Lee et al. (2005). Copyright
(2005) American Chemical Society. The bold line for HOBr is
taken from Fig. 9.

w a t e r r e s e a r c h 4 8 ( 2 0 1 4 ) 1 5 e4 2

transfer. During the initial phase of the reaction, bromination


of NOM dominated with more than 75% (Echigo and Minear,
2006). In the case of preozonated NOM, however, bromination occurred to 15e25% and electron transfer to 75e85% for
reaction times longer than the initial phase (Westerhoff et al.,
1998).

3.3.2.

Nitrogen-containing compounds

Generally, hypohalous acids (i.e. HOCl and HOBr) react


quickly with nitrogen-containing compounds, forming
varying N-halo-derivatives, which depend on the N/hypohalous acid ratio and pH. At a ratio of 1:1 or higher, the
product is a monohalogenated N-compound; at lower ratios,
dihalogenated compounds are formed in a second reaction,
which is slower than the first halogenation (Armesto et al.,
1998).

3.3.2.1. Amines.

Most of the kinetic information for


HOBreamine reactions deals with primary amines; little or no
information is available on secondary and tertiary amines
(Table 6). The reaction of HOBr with primary and secondary
amines is very fast with apparent second-order rate constants
in the order of 105e106 M1 s1 at near neutral pH (Pattison
and Davies, 2004; Skaff et al., 2007; Wajon and Morris, 1982).
A rate constant for the reaction of HOBr with glycine has been
determined in two studies with apparent second-order rate
constants of 2.6  106 M1 s1 for a pH between 7.2 and 7.5
(Pattison and Davies, 2004) and 1.4  106 M1 s1 for pH 7.5
(Wajon and Morris, 1982) (Table 6). A difference of about a
factor of two is quite common for experimentally determined
rate constants. Based on the limited data in Table 6, it seems
that the nature of the substituent R on the primary amine
(NH2R) has no significant effect on the magnitude of the
second-order rate constant.
The products of the reactions of bromine with amines are
bromamines formed by an electrophilic substitution of a
proton by Br (Wajon and Morris, 1982). It has been found that
the resultant organic bromamines can react further and
brominate phenolic structures (Wu et al., 1999). Also the reaction of some organic bromamines with nicotinamide
adenine dinucleotide (NADH) and ethylene glycol vinyl ether
has been reported: the rate constant for the reaction of
brominated taurine with NADH at pH 7.3 was found to be
2.7  103 M1 s1 and the rate constants for the reaction
of brominated phosphoryl-ethanolamine, brominated phosphoryl-serine, brominated N-a-Lysine and brominated
taurine with ethylene glycol vinyl ether ranged between
1.3  103 and w2.7  104 M1 s1 at pH 7.4 (Prutz et al., 2000;
Skaff et al., 2008).
Due to the higher electrophilicity of HOBr compared to
HOCl, the species-specific rate constants of the analogous
reactions for HOBr and primary and secondary amines are
around one to two orders of magnitude higher than for HOCl
(Deborde and von Gunten, 2008). Similar to the case of
ammonia (Section 3.2.1, Fig. 6), the reaction of amines (e.g.,
glycine) with bromine is not only favored due to the higher
species-specific rate constants but also due to the difference
in pKa values of HOBr (8.8) and HOCl (7.5). This difference in
pKa values means that the pH range over which HOBr and the
neutral form of the amine (i.e. the two reacting species) are

29

simultaneously present is larger than for HOCl. At the same


time, over this pH range the relative HOBr concentration is
always higher than the relative HOCl concentration.
The rate constants of the reaction of bromine with primary
and secondary amines are higher not only than the corresponding rate constants of chlorine, but also than the corresponding rate constants of ozone reactions (von Sonntag and
von Gunten, 2012). The second-order rate constants for the reaction of ozone with glycine and dimethylamine are in the order
of 105 M1 s1 and 107 M1 s1, respectively (Hoigne and Bader,
1983b), whereas the corresponding species-specific rate constants are in the order of 108 M1 s1 and 109 M1 s1 for HOBr
(Table 6). This may affect oxidation product formation from
amines or other nitrogen-containing compounds during ozonation of bromide-containing waters (von Gunten et al., 2010).

3.3.2.2. Other nitrogen-containing organic compounds. Similar


to the corresponding reactions of HOCl (Deborde and von
Gunten, 2008), HOBr reacts only slowly with amides (Table
6). This can be generally expected for oxidation of amides
since the nitrogen is deactivated by the adjacent carbonyl
group (von Sonntag and von Gunten, 2012). In the case of urea,
it was found that the chlorination was initiated by Cl2, rather
than HOCl (Blatchley and Cheng, 2010). This might also be the
case for the reaction of bromine with amides. In contrast to
linear amides, second-order rate constants for the reaction of
HOBr with cyclic amides such as Cyclo(Gly)2 or Cyclo(Ala)2
have been reported to be around two orders of magnitude
greater (Pattison and Davies, 2004; Prutz, 1999). The brominated amide Cyclo(Gly)2 reacts further with NADH with a rate
constant of at least 4  105 M1 s1 (Prutz, 1999).
Furthermore, apparent second-order rate constants for the
reaction of guanidine derivatives with bromine at pH 7.2e7.5
are in the order of 103 M1 s1. For the reaction of bromine
with imidazole, indole, and pyridine moieties and urate
apparent rate constants in the order of 106 M1 s1 at pH
7.2e7.5 have been reported (Table 6).
3.3.3.

Sulfur-containing compounds

Apparent second-order rate constants for the reaction of HOBr


with sulfur-containing compounds at near neutral pH are
listed in Table 7. In the case of the moieties of the amino acids
cysteine and methionine, rate constants in the order of
107 M1 s1 (sulfhydryl group) and 106 M1 s1 (thioether
group) have been reported (Pattison and Davies, 2004; Prutz
et al., 2000). These rate constants are high, but lower than
the corresponding apparent rate constants of the reaction
with HOCl, which are 3.0  107 M1 s1 for N-acetyl-cysteine
and 3.8  107 M1 s1 for N-acetyl-methionine (Pattison and
Davies, 2001, 2004). For the reaction of HOBr with the disulfide bond in 3,30 -dithio-dipropionic acid (DTDPA) and in
cystine (disulfide bond), apparent second-order rate constants
in the order of 106 M1 s1 and 105 M1 s1, respectively, were
reported at pH 7.2e7.5 (Pattison and Davies, 2004). The lower
rate constant for cystine compared to DTDPA might be due to
sterical hindrance (Pattison and Davies, 2004). The secondorder rate constant for the reaction of DTDPA with HOBr is
around one order of magnitude higher than for HOCl, while
the rate constants for cystine differ by a factor of two (Pattison
and Davies, 2001, 2004).

30

Table 6 e Rate constants for the reaction of nitrogen-containing compounds with bromine.
Compound

Reactive Moiety

pKa

Species-specific
rate constants
k(HOBr)
[M1 s1]

Primary
Primary
Primary
Primary
Primary

amine
amine
amine
amine
amine

Phosphoryl-ethanolamine

Primary amine

Phosphoryl-serine

Primary amine

Valine
Dimethylamine
N-Acetyl-alanine
N-Acetyl-alanine-OMe
Cyclo(Alanine)2
Cyclo(Aspartic acid)2
Cyclo(Glycine)2

Primary amine
Secondary amine
Amide
Amide
Amide
Amide
Amide

Cyclo(Serine)2
2-Methylpropionamide
Propionamide
Trimethylacetamide
N,N-Dimethylsulfamide
Urate
N-Acetyl-arginine-OMe
Ethylguanidine
N-Acetyl-tryptophan
4-Imidazoleacetic acid
NMNH (Nicotinamide mononucleotide)
NMNH, brominated

Amide
Amide
Amide
Amide
Sulfamide
Guanine
Guanidine
Guanidine
Indole
Imidazole
Pyridine
Br-Pyridine

kapp
[M1 s1]

OBr 
 k1
M s1


10.5
9.9
10
9.8

3.5  108
(3.8  0.3)  108

9.7
10.7

3.0  109

10.5

8.1  108

5.0  104
(2.1  0.2)  105

(3.6  0.3)  10 (pH 7.2e7.5)


(1.6  0.1)  106 (pH 7.2e7.5)
(2.6  0.2)  105 (pH 7.2e7.5)
3.0  105
4.6  105
(2.6  0.2)  106 (pH 7.2e7.5)
(8.8  0.4)  105 (pH 7.4)
(2.6  0.4)  105 (pH 7.4)
(9.3  0.8)  105 (pH 7.4)
(3.1  0.4)  105 (pH 7.4)
(1.7  0.1)  106 (pH 7.2e7.5)
6.0  105
(7  2)  102 (pH 7.2e7.5)
2.1  0.2 (pH 7.2e7.5)
(2.5  0.3)  102a (pH 7.2e7.5)
(5.0  2.0)  101a (pH 7.2e7.5)
(9.0  1.0)  102a (pH 7.2e7.5)
(2.9  0.3)  103a (pH 7.2e7.5)
6  102a (pH 7.2)
(5.5  0.9)  102a (pH 7.2e7.5)
1.5  0.3 (pH 7.2e7.5)
3.3  0.4 (pH 7.2e7.5)
(9  1)  101 (pH 7.2e7.5)
2.7  105
(3.4  0.2)  106 (pH 7.4)
(2.2  0.4)  103 (pH 7.2e7.5)
(1.3  0.2)  103 (pH 7.2e7.5)
(3.7  0.3)  106 (pH 7.2e7.5)
w3.0  106 (pH 7.2e7.5)
w4  106 (pH 7.2)
w9  105 (pH 7.5)

a Rate constant given per amide group, i.e. the rate constant for the molecule would double (Pattison and Davies, 2004).

Temperature

Reference

[ C]
22
22
22
20
20
22
22
10
22
10
22
20
22
22
22
22
22
37
20  2
22
22
22
22
25  2
22
22
22
22
22
20  2
20  2

(Pattison and Davies, 2004)


(Pattison and Davies, 2004)
(Pattison and Davies, 2004)
(Wajon and Morris, 1982)
(Wajon and Morris, 1982)
(Pattison and Davies, 2004)
(Skaff et al., 2007)
(Skaff et al., 2007)
(Skaff et al., 2007)
(Skaff et al., 2007)
(Pattison and Davies, 2004)
(Wajon and Morris, 1982)
(Pattison and Davies, 2004)
(Pattison and Davies, 2004)
(Pattison and Davies, 2004)
(Pattison and Davies, 2004)
(Pattison and Davies, 2004)
(Pattison and Davies, 2004)
(Prutz, 1999)
(Pattison and Davies, 2004)
(Pattison and Davies, 2004)
(Pattison and Davies, 2004)
(Pattison and Davies, 2004)
(Heeb et al., in preparation)
(Skaff et al., 2007)
(Pattison and Davies, 2004)
(Pattison and Davies, 2004)
(Pattison and Davies, 2004)
(Pattison and Davies, 2004)
(Prutz et al., 2000)
(Prutz et al., 2000)

w a t e r r e s e a r c h 4 8 ( 2 0 1 4 ) 1 5 e4 2

N-a-Acetyl-lysine
Alanine
-Aminocaproic acid
Glutamate
Glycine

Apparent rate constant


at pH 7 or at given pH

31

w a t e r r e s e a r c h 4 8 ( 2 0 1 4 ) 1 5 e4 2

Table 7 e Rate constants for the reaction of sulfur-containing compounds with bromine.
Compound

Apparent rate constant


at pH 7.2e7.5

pKa

Temperature [ C]

Reference

kapp [M1 s1]


N-Acetyl-cysteine
N-Acetyl-cystine (N-Acetyl-Cysteine)2
N-Acetyl-methionine-OMe

(1.2  0.2)  107


(3.4  0.8)  105
(3.6  0.3)  106
(9.6  1.3)  106
(1.1  0.2)  106
(2.3  0.2)  106
w4  106 (pH 7.2)

8.1 (eSH)

3,30 -Dithio-dipropionic acid (DTDPA)


Methionine

22
22
22
37
22
37
20  2

(Pattison and Davies,


(Pattison and Davies,
(Pattison and Davies,
(Pattison and Davies,
(Pattison and Davies,
(Pattison and Davies,
(Prutz et al., 2000)

2004)
2004)
2004)
2004)
2004)
2004)

Table 8 e Rate constants for the reaction of olefines with bromine.


Compound

pKa

Temperature [ C]

Reference

(1.7  0.2)  106

22

(Skaff et al., 2007)

(3.5  0.1)  106

22

(Skaff et al., 2008)

(1.1  0.2)  104


(1.3  0.2)  103

22
22

(Skaff et al., 2007)


(Skaff et al., 2007)

Apparent rate constant at pH 7.4


1

kapp [M
Ascorbate

4.1
11.8

Ethylene glycol
vinyl ether
3-pentenoic acid
Sorbate

3.3.4.

4.5
4.8

1

s ]

Olefines

Table 8 provides the available kinetic information for the reaction of olefines with HOBr. At pH 7.4, ascorbate has a high
apparent second-order rate constant because one hydroxylic
group is deprotonated, which significantly increases the
electron density in the double bond. The double bond in
ethylene glycol vinyl ether reacts very rapidly with HOBr,
while the corresponding rate constants for the reaction with
3-pentenoic acid and sorbate are about two and three orders
of magnitude lower (Table 8). The lower rate constant for the
reaction with sorbate compared to 3-pentenoic acid can be
explained by a conjugation of the electron-withdrawing
carboxyl group in the case of sorbate.

3.3.5.

Carboxylic acids, aldehydes, alcohols, and ketones

The reported apparent second-order rate constants for the


reaction of HOBr with carboxylic acids are low. Formic and
acetic acid are practically unreactive, while the dicarboxylic

acids oxalic and malonic acid show a low reactivity (Table 9).
Even lower rate constants have been reported for aldehyde
(formaldehyde, pH 4) and alcohol (2-propanol, pH 6.7)
(Table 9). Rate constants for the reaction of Br2 with acetaldehyde and the alcohols 2-propanol and ethanol near neutral
pH have been reported to be in the order of 101 M1 s1
and 104 M1 s1, respectively (Perlmutter-Hayman and
Weissmann, 1962, 1969). Species-specific rate constants for
the reaction of OBr for various alcohols range between
4.1  107 and 3.4  104 M1 s1 (Negi et al., 1987).
Ketones only react with HOBr in their enol form (Fig. 11).
For the reaction of HOBr with ketones, their enolization is the
rate-determining step, because the reaction of the enol (olefine, see above) with bromine is fast (Pinkernell and von
Gunten, 2001). The ketone-enol equilibrium is generally
heavily on the ketone side and enolization rate constants are
low; in the order of 106 s1 for aliphatic ketones in acidic
media (Dubois and Toullec, 1969). Similar to the bromination

Table 9 e Rate constants for the reaction of carboxylic acids, aldehydes, and alcohols with bromine.
Compound

pKa

Apparent rate constant


1

kapp [M
Acetic acid
Formic acid
Malonic acid
Oxalic acid
Formaldehyde
2-Propanol

4.8
3.8
2.8
5.7
1.3
4.1

Temperature [ C]

Reference

1

s ]

<1 (pH 6)
<1 (pH 6)
3.0  101 (pH 4)

20
20
20

(Pinkernell and von Gunten, 2001)


(Pinkernell and von Gunten, 2001)
(Pinkernell and von Gunten, 2001)

4.0  101 (pH 6)

20

(Pinkernell and von Gunten, 2001)

1.8  103 (pH 4)


3.9  104 (pH 6.7)

20
25

(Pinkernell and von Gunten, 2001)


(Perlmutter-Hayman and Weissmann, 1969)

32

w a t e r r e s e a r c h 4 8 ( 2 0 1 4 ) 1 5 e4 2

naphthalene (3  103 M1 s1) is 1500 times greater than the


rate constant for its reaction with benzene (2 M1 s1) (Hoigne
and Bader, 1983a; von Sonntag and von Gunten, 2012).

HOBr

R2

R1

OH

R1
R3

R3

R2
H
Ketone

Enol

Fig. 11 e Enolization of ketones followed by reaction with


bromine. Bromine attack occurs at the double bond of the
enol (arrow). The rate-determining step of this reaction is
the enolization.

of aromatic compounds, bromination of enols (olefines) is an


important source of THMs.

3.3.6.

Micropollutants

The reactivity of HOBr with several micropollutants has been


investigated and depends on the functional groups present in
the micropollutant (Table 10). Fig. 12 shows the chemical
structures of micropollutants for which data on their reactivity with bromine is available, including the likely site of
reaction with bromine.

3.3.6.1. Aromatic compounds. 17a-Ethinylestradiol (EE2) and


its halogenated transformation products have high secondorder rate constants for their reactions with HOBr, due to
their phenolic moieties (Table 10) (Lee and von Gunten, 2009).
The species-specific rate constants decrease with increasing
degree of halogenation of the phenolic ring (for discussion see
Section 3.3.1). Since the reactivity of HOBr is about three orders of magnitude higher than the one of HOCl (see Fig. 7), the
transformation of EE2 is accelerated during chlorination of
bromide-containing waters (Lee and von Gunten, 2009). It has
been shown that bromination of EE2 is an effective means to
destroy the estrogenic properties of EE2 (Lee et al., 2008).
Endpoints other than estrogenicity have not yet been investigated for brominated EE2. Chlorophene is as well attacked at
its phenolic moiety. Its apparent rate constant at pH 7
(1.9  102 M1 s1) is a factor 100 lower than what could be
expected by comparison to the rate constant of 4chlorophenol (3.2  104 M1 s1, Table 5).
In contrast, the phenylurea herbicides diuron and isoproturon react only slowly with bromine (Acero et al., 2007)
(Table 10), because the aromatic ring is deactivated (diuron) or
only slightly activated (isoproturon) and nitrogen is present as
an amide, which is characterized by a low reactivity with
HOBr (see Section 3.3.2). DEET shows an apparent rate constant comparable to diuron, which is explained by its deactivated aromatic ring. Phenacetin is an aromatic compound
with only a slight activation by an ethoxy group, which results
in an apparent second-order rate constant of 7.3 M1 s1 for its
reaction with bromine at pH 7 (Table 10). Naproxen has a
higher reactivity due to the presence of a naphthalene moiety
(k 25 M1 s1, Table 10). This is comparable to the case of
ozone, whose second-order rate constant for the reaction with

3.3.6.2. Nitrogen-containing compounds. The reaction sites for


potential bromine attack on metoprolol are the aromatic ring
or the secondary amine group. Based on the reactivity of
dimethylamine (apparent second-order rate constant of
around 105 M1 s1 at pH 7, Table 6) it would be expected that
the secondary amine moiety dominates the kinetics. However, there is a discrepancy between the low rate constant for
the reaction with metoprolol at pH 7 (Table 10) and the reported rate constant for the reaction with dimethylamine
(Table 6). The same is true for nortriptyline and hydrochlorothiazide (its aromatic ring is unreactive to electrophilic
attack due to the electron-withdrawing chloro- and
sulfoxide-moieties), where the attack is also expected at the
secondary amine group. More data is needed to interpret
these findings.
As far as attacks on heterocyclic nitrogen are concerned,
bromine shows a very high reactivity with 3-methylindole
(1.1  108 M1 s1 at pH 7). A high rate constant was also
found for the reaction of the indole moiety in the amino acid
tryptophan (3.7  106 M1 s1 at pH 7.2e7.5, Table 6). The
reactivity of bromine with benzotriazole is rather low
(8.5 M1 s1 at pH 7).
3.3.6.3. Sulfur-containing compounds. The pesticide chlorpyrifos shows a high second-order rate constant for the reaction with bromine (Table 10), which is three orders of
magnitude greater than the corresponding rate constant of
the reaction with chlorine (4.8  102 M1 s1) (Duirk et al.,
2008). Because the aromatic ring is highly deactivated by
chlorine substitution, the thiophosphate group is oxidized,
forming chlorpyrifos oxon, which is more toxic than the
parent compound (Wu and Laird, 2003). It was found that OBr
and OCl accelerate the hydrolysis of both the parent compound and the oxon to form 3,5,6-trichloro-2-pyridinol (Duirk
and Collette, 2006; Duirk et al., 2008). The rate constants for
the hydrolysis by OBr and OCl range between 0.27 and
0.39 M1 s1 (Duirk and Collette, 2006; Duirk et al., 2008).
The species-specific rate constants for the reaction of
bromine and chlorine with ametryn were reported to be
higher for the positively charged, protonated species than for
the neutral species of ametryn (Xu et al., 2009). However, the
reason for an increasing apparent rate constant with
decreasing pH is probably not the speciation of ametryn and
thus not a higher species-specific rate constant for the reaction with protonated ametryn compared to the neutral form.
This effect is more likely due to the contribution of Br2 at low
pH. As observed for the other compounds as well, the secondorder rate constant of the reaction of bromine with ametryn
was significantly higher than that for chlorine (Xu et al., 2009).
An attack on the sulfur, resulting in a sulfoxide was proposed
for the reaction of chlorine (Xu et al., 2009); the same mechanism can be assumed for bromine, because the rest of the
molecule (triazine ring and deactivated secondary amines) is
not susceptible to oxidative attack (von Sonntag and von
Gunten, 2012). The apparent second-order rate constant for
the reaction of bromine with ametryn at pH 7 is more than two

Table 10 e Rate constants for the reaction of micropollutants with bromine.


Compound

Species-specific rate constants

pKa
kHOBrH a
[M2 s1]

Ametryn
Amoxicillin

Benzotriazole

<1  107

k(HOBr/P)b
[M1 s1]

kHOBr=P b
[M1 s1]

3.5  106
9.1  103
(2.4  0.4)  104 (2.4  0.4)  104 (9.9  1.5)  106

kHOBr=P2 b
[M1 s1]

kOBr =P b
[M1 s1]

(3.8  1.2)  109

(3.2  0.6)  105


0.67
1.2  105
10.4
8.7
8.6
8.9
9.0
7.1
7.1
7.1
7.1
7.9 (4.2  0.8)  103 3.3  0.4
3.1  107
9.7 (2.8  0.2)  103 3.9  0.2

1.2  101
(7.7  2.2)  103 (5.3  0.7)  108
(7.2  1.3)  107
(2.8  1.4)  107
(1.3  0.4)  108
(7.4  2.6)  107
(4.9  1.2)  105
7.3  105
7.3  105
(9.7  2.4)  105
(5.8  1.0)  101
1.8  101
(1.0  0.1)  102

4.2 <1  108


10.2
1.4 <1  104
14

(2.1  0.4)  104 (2.7  0.3)  102


(5.4  0.6)  102 1.7  0.2

(7.4  0.5)  107

8  102
(1.2  0.8)  106

2  102

Reference

1. 3  104
6.7  106

25  2
20  0.2

(Xu et al., 2009)


(Benitez et al., 2011)

8.5  0.3

20

(Acero et al., 2013)

(1.9  0.2)  102


3.1  105
(9  2)  102
1.3  101
2.1  105
1.5  106
6.6  105
1.7  106
7.0  105
2.1  105
3.2  105
3.2  105
4.1  105
8.1  1.2
2.5  101
5.2  0.8
(1.1  0.1)  108
(2.5  0.1)  101
7.4  0.3
7.3  1.0

20
25  1
20
20  0.2
23  2
23  2
23  2
23  2
23  2
23  2
23  2
23  2
23  2
20  0.2
20  0.2
20  0.2
20
20  0.2
20
20  0.2

(Acero et al., 2013)


(Duirk et al., 2008)
(Acero et al., 2013)
(Acero et al., 2007)
(Lee and von Gunten,
(Lee and von Gunten,
(Lee and von Gunten,
(Lee and von Gunten,
(Lee and von Gunten,
(Lee and von Gunten,
(Lee and von Gunten,
(Lee and von Gunten,
(Lee and von Gunten,
(Benitez et al., 2011)
(Acero et al., 2007)
(Benitez et al., 2011)
(Acero et al., 2013)
(Benitez et al., 2011)
(Acero et al., 2013)
(Benitez et al., 2011)

2009)
2009)
2009)
2009)
2009)
2009)
2009)
2009)
2009)

w a t e r r e s e a r c h 4 8 ( 2 0 1 4 ) 1 5 e4 2

Chlorophene
Chlorpyrifos
DEET
Diuron
EE2
2-Cl EE2
4-Cl EE2
2-Br EE2
4-Br EE2
2,4-DiCl EE2
2,4-Br,Cl EE2
2,4-Cl,Br EE2
2,4-DiBr EE2
Hydrochlorothiazide
Isoproturon
Metoprolol
3-Methylindole
Naproxen
Nortriptyline
Phenacetin

4.1
2.6
7.3
9.7
0.4
8.2
9.8

kHOBr=P b
[M1 s1]

Apparent rate Temperature


constant at pH 7
kapp
[M1 s1]
[ C]

a It is not clear, whether these rate constants truly characterize an acid-catalysis or rather the effect of other bromine species formed at low pH (see Section 3.1).
b P, P, P and P2 represent the charge of the molecule and not in each case its actual protonation.

33

34

w a t e r r e s e a r c h 4 8 ( 2 0 1 4 ) 1 5 e4 2

Fig. 12 e Structures of micropollutants for which second-order rate constants for the reaction with bromine are available (in
alphabetical order). Arrows indicate the possible reaction sites for bromine.

w a t e r r e s e a r c h 4 8 ( 2 0 1 4 ) 1 5 e4 2

orders of magnitude smaller than the corresponding rate


constant for N-acetyl-methionine-OMe with a similar reactive
moiety (thioether, Table 7). The lower reactivity of ametryn
can be explained by the electron-withdrawing effect of the
triazine moiety.
For amoxicillin, an attack of bromine is possible at the
phenolic, the amine or the sulfur moiety. Based on the reactivity of the different functional groups, the phenolic moiety
can be ruled out. It also seems that the second-order rate
constants for the reaction of HOBr with the primary amine
(see Table 6) are too low to explain the apparent rate constant
of the reaction of bromine with amoxicillin at pH 7. Therefore,
it is hypothesized that the primary attack of bromine occurs at
the thioether group.

4.
Modeling of chlorination processes in the
presence of bromide
To assess the effect of bromide on oxidative transformation
reactions during chlorination, a kinetic model was set up,
which is illustrated in Fig. 13. It combines the formation of
HOBr through the oxidation of bromide by HOCl (Eq. (1),
Fig. 13, (ii)) with the reaction of HOCl (Fig. 13, (i)) and HOBr
(Fig. 13, (iii and v)) with an organic compound (e.g., NOM, a
micropollutant, a taste and odor compound, etc.). The oxidation of bromide by HOCl (Fig. 13, (ii)) competes with the reaction of HOCl with the organic compound (Fig. 13, (i)). For the
reaction of HOBr with the organic compound, two possibilities
were considered: electron transfer (Fig. 13, (iii), model results
shown in Fig. 14a and b) and substitution reaction (Fig. 13, (v),
model results shown in Fig. 14c and d). The electron transfer
reaction results in an oxidized organic compound and bromide, which can be re-oxidized by HOCl to HOBr (Fig. 13, (iv)),
resulting in a catalytic reaction. In the case of the substitution

HO

Cl

(iv)

Br - + Org. Cox
(iii)

Br

HOCl
(i)

HOBr

Org. C

(ii)

(v)

Org

Br-Org. C

.C
Org. Cox / Cl-Org. C

Fig. 13 e Illustration of the reactions considered in the


kinetic model. Chlorine reacts (i) with an organic
compound (Org. C) or (ii) oxidizes bromide to bromine.
Bromine then either (iii) oxidizes the organic compound by
electron transfer (forming Org. Cox) during which bromide
is released and can be re-oxidized to bromine (iv), or (v)
undergoes a substitution reaction forming a brominated
organic compound (Br-Org. C). All the compounds undergo
acidebase speciation. For better readability, only the
relevant species are included in this graph. The
corresponding rate and equilibrium constants are given in
Table 11.

35

reaction, the product is a brominated organic compound, a


sink for Br.
Half-lives of the organic compound (t1/2) and the fraction
of the organic compound reacting with bromine (a in the
case of electron transfer and b in the case of substitution,
each for a 90% degradation of the organic compound) were
calculated as a function of pH and the initial bromide concentration. For the organic compound, similar kinetic characteristics as reported for phenol (rate constants see Table
11) were chosen for both pathways (electron transfer and
substitution, Fig. 14aed). Bromide concentrations of up to
250 mg L1 (3 mM) and a solution pH of 7e9 were used for the
calculations. An initial concentration of 10 mM organic
compound (equivalent to 0.72 mg C L1) and a HOCl dose of
15 mM were used (1 mg Cl2 L1). In the model system, either
electron transfer or substitution is considered; no calculations were done for a system where both reactions occur
simultaneously.
In general, the model confirms that the overall rate of reaction is enhanced in the presence of bromide. The half-life of
the target compound decreases with increasing bromide
concentrations (up to 250 mg L1) over the whole pH range
from 3500 s to 100 s for the electron transfer and from 3500 s to
500 s for the substitution reaction (Fig. 14a and c). However,
the extent of the bromide contribution depends on the pH of
the solution and the type of reaction involved (electron
transfer vs. substitution). Globally, the reaction is faster in the
electron transfer process, because of the overall higher HOBr
production due to the catalytic effect of bromide. Both the
reactions of HOCl and HOBr depend on pH, the maximum
reactivity occurs at a pH equal to the mean of the pKa values
(i.e. 8.75 and 9.40, respectively, for HOCl and HOBr in a system
containing an organic compound with a pKa of 10.0, see Section 3.3.1). The system also depends on the pH dependence of
the HOClebromide reaction (Eq. (1)), which is why the HOBr
formation can be the limiting factor for the degradation of
organic compounds at high pH. This is illustrated by the
increasing half-life toward higher pH for a bromide concentration of 250 mg L1 (Fig. 14a and c). The effect of the initial
bromide concentration on the half-life time is much more
important at pH 7 than at higher pH values.
The type of reactivity (electron transfer vs. substitution) of
the oxidants is not only important in terms of kinetics but also
for the formation of halogenated by-products. In natural waters, a mix of the two types of reactions can be assumed. For
electron transfer, the fraction of the HOBr reaction varies
dramatically with the bromide concentration and the pH of
the solution. Under the evaluated conditions, the HOBr fraction is at its maximum (reaching almost 100%) at neutral pH
and in the presence of high levels of bromide (Fig. 14b). In
contrast, when substitution is considered, the bromide concentration is the key factor for the incorporation of bromine,
while the pH affects bromine binding only at a pH above 8.5,
due to the slower oxidation of bromide by HOCl at higher pH
(Fig. 14d). In fact, in the pH range 7e8.5, the b-values correspond to an incorporation of all of the initial bromide, at pH 9
this decreases to 70% of the initial bromide. This is due to
incomplete oxidation of bromide by HOCl (reaction (ii) in
Fig. 13), which results in lower HOBr concentrations and thus
a higher fraction of chlorine substitution.

36

w a t e r r e s e a r c h 4 8 ( 2 0 1 4 ) 1 5 e4 2

Fig. 14 e Results of the model illustrated in Fig. 13. (a)e(d): Half-lives of the organic compound (t1/2) and fractions of it
reacting with bromine (a/b) as a function of pH and initial bromide concentration for catalytic electron transfer (a, b) and
substitution (c, d). (e, f): Fractions of the organic compound reacting with bromine (a/b) as a function of initial bromide
concentration and the ratio of the apparent rate constants (kHOBr/kHOCl) at pH 8 for electron transfer (e) and substitution (f)
(kHOBr constant according to Table 11). Calculations were done for a 90% transformation of the organic compound with initial
concentrations of 10 mM organic compound (equivalent to 0.72 mg C/L) and 15 mM HOCl (1 mg Cl2/L).

37

w a t e r r e s e a r c h 4 8 ( 2 0 1 4 ) 1 5 e4 2

Table 11 e Rate and equilibrium constants used in the model illustrated in Fig. 13.
Reaction
HOCl Br /HOBr Cl
ClO Br /BrO Cl
HOBr$OBr H
HOCl$OCl H
Org:Cprot $Org:C H
HOCl Org:C$Cl  Org:C=Org:Cox Cl
HOBr Org:C/Br  Org:C=Org:Cox Br
OBr Org:C/Br  Org:C=Org:Cox Br

Rate/Equilibrium constant

Corresponding pathway in Fig. 13

k 1.55  103 M1 s1


k 9  104 M1 s1
pKa 8.8
pKa 7.5
pKa 10.0
k 2.2  104 M1 s1
k 6.6  107 M1 s1
k 3.5  104 M1 s1

(ii) and (iv)


(ii) and (iv)

To show the impact of the reactivities of HOCl and HOBr


with organic compounds, model calculations with varying
ratios of the corresponding rate constants (kHOBr/kHOCl) were
performed at pH 8 (Fig. 14e and f for electron transfer and
substitution, respectively). The ratios were varied from 102 to
104, keeping the apparent second-order rate constant for the
reaction of HOBr with the organic compound constant
(k 6.6  107 M1 s1, Table 11) and varying the rate constant
of HOCl. When bromide acts as a catalyst, the fraction of reaction occurring through HOBr progressively increases with
the bromide concentration and the ratio kHOBr/kHOCl (Fig. 14e,
parameter a). In contrast, for substitution the bromination
reaches a maximum for a kHOBr/kHOCl ratio of 3  103 (Fig. 14f,
parameter b). There is no increase of bromine substitution at
higher ratios, because no free bromide is available anymore in
solution. Thus, the effect of bromide during chlorination depends on both the relative reactivity of HOBr and HOCl and on
the mechanism involved (electron transfer vs. substitution).

5.

Conclusion

Reactive bromine species can be formed during oxidative


water treatment of bromide-containing waters with chlorine,
ozone or chloramines and react with dissolved inorganic and
organic compounds. A critical evaluation of kinetic and
mechanistic information allows the following conclusions:
 In the pH range relevant for water treatment, the speciation
of bromine in the absence and presence of chlorine in fresh
waters is governed by hypobromous acid/hypobromite
(HOBr/OBr), which are present in concentrations at least
five orders of magnitude greater than other reactive
bromine species such as Br2, Br2O, BrOCl or BrCl.
 Reactions with many inorganic anions (A) are fast and
often proceed through Br transfer to ABr, which hydrolyzes
to AO, an oxygen addition product.
 The reaction of bromine with ammonia is fast and leads to
the formation of bromamines.
 Based on the corresponding chlorine reactions, the oxidation of Fe(II) and As(III) by bromine at circumneutral pH is
expected to be very fast. The oxidation of Mn(II) by bromine
is much slower and proceeds through a heterogeneous
process catalyzed by Mn(III,IV) oxides.
 HOBr has a high reactivity with many organic compounds
containing electron-rich moieties such as activated




(i)
(iii) and (v)
(iii) and (v)

aromatic systems, deprotonated amines, olefins, and


reduced sulfur species with apparent second-order rate
constants (kapp) in the range of 103e107 M1 s1 at pH 7.
Overall, kapp for the reaction of bromine with organic compounds at pH 7 vary over more than seven orders of
magnitude from <1 to 107 M1 s1.
Hammett-type QSARs and cross correlations with the kinetics of the corresponding chlorine reactions can be used
to predict the kinetics of bromine reactions.
Bromine has typically higher second-order rate constants
than chlorine (e.g., factor of w3000 for substituted phenols).
Kinetic data on oxidative transformation of organic micropollutants show similar trends to those of organic moieties
in model compounds.
Kinetic calculations of chlorination of bromide-containing
waters show that HOBr can account for a large fraction of
the transformation of organic compounds. The extent of
this fraction depends on the bromide concentration, pH,
type of bromine reaction (electron transfer vs. electrophilic
substitution) and the ratio of the second-order rate constants for the reactions of HOBr and HOCl, respectively.

references

Acero, J.L., Piriou, P., von Gunten, U., 2005. Kinetics and
mechanisms of formation of bromophenols during drinking
water chlorination: assessment of taste and odor
development. Water Research 39 (13), 2979e2993.
Acero, J.L., Real, F.J., Benitez, F.J., Gonzalez, M., 2007. Kinetics of
reactions between chlorine or bromine and the herbicides
diuron and isoproturon. Journal of Chemical Technology and
Biotechnology 82 (2), 214e222.
Acero, J.L., Benitez, F.J., Real, F.J., Roldan, G., Rodriguez, E., 2013.
Chlorination and bromination kinetics of emerging
contaminants in aqueous systems. Chemical Engineering
Journal 219, 43e50.
Agus, E., Voutchkov, N., Sedlak, D.L., 2009. Disinfection byproducts and their potential impact on the quality of water
produced by desalination systems: a literature review.
Desalination 237 (1e3), 214e237.
Allard, S., Fouche, L., Heitz, A., von Gunten, U., 2013. Oxidation of
manganese (II) during chlorination: role of bromide.
Environmental Science and Technology 47 (15), 8716e8723.
Amy, G.L., Chadik, P.A., King, P.H., Cooper, W.J., 1984. Chlorine
utilization during trihalomethane formation in the presence
of ammonia and bromide. Environmental Science and
Technology 18 (10), 781e786.

38

w a t e r r e s e a r c h 4 8 ( 2 0 1 4 ) 1 5 e4 2

Armesto, X.L., Canle, L.M., Garc`a, M.V., Santaballa, J.A., 1998.


Aqueous chemistry of N-halo-compounds. Chemical Society
Reviews 27 (6), 453e460.
Arnold, W.A., Bolotin, J., von Gunten, U., Hofstetter, T.B., 2008.
Evaluation of functional groups responsible for chloroform
formation during water chlorination using compound specific
isotope analysis. Environmental Science and Technology 42
(21), 7778e7785.
Ashby, M.T., Carlson, A.C., Scott, M.J., 2004. Redox buffering of
hypochlorous acid by thiocyanate in physiologic fluids. Journal
of the American Chemical Society 126 (49), 15976e15977.
Beckwith, R.C., Margerum, D.W., 1997. Kinetics of hypobromous acid
disproportionation. Inorganic Chemistry 36 (17), 3754e3760.
Beckwith, R.C., Wang, T.X., Margerum, D.W., 1996. Equilibrium
and kinetics of bromine hydrolysis. Inorganic Chemistry 35 (4),
995e1000.
Benitez, F.J., Acero, J.L., Real, F.J., Roldan, G., Casas, F., 2011.
Bromination of selected pharmaceuticals in water matrices.
Chemosphere 85 (9), 1430e1437.
Blatchley, E.R., Cheng, M., 2010. Reaction mechanism for
chlorination of urea. Environmental Science and Technology
44 (22), 8529e8534.
Booth, R.A., Lester, J.N., 1995. The potential formation of
halogenated by-products during peracetic acid treatment of
final sewage effluent. Water Research 29 (7), 1793e1801.
Bousher, A., Brimblecombe, P., Midgley, D., 1986. Rate of
hypobromite formation in chlorinated seawater. Water
Research 20 (7), 865e870.
Bousher, A., Brimblecombe, P., Midgley, D., 1989. Kinetics of
reactions in solutions containing monochloramine and
bromide. Water Research 23 (8), 1049e1058.
Buffle, M.O., Galli, S., von Gunten, U., 2004. Enhanced bromate
control during ozonation: the chlorine-ammonia process.
Environmental Science and Technology 38 (19), 5187e5195.
Bulloch, D.N., Lavado, R., Forsgren, K.L., Beni, S., Schlenk, D.,
Larive, C.K., 2012. Analytical and biological characterization of
halogenated gemfibrozil produced through chlorination of
wastewater. Environmental Science and Technology 46 (10),
5583e5589.
Burgi, H., Schaffner, T., Seiler, J.P., 2001. The toxicology of iodate:
a review of the literature. Thyroid 11 (5), 449e456.
Canonica, S., Tratnyek, P.G., 2003. Quantitative structure-activity
relationships for oxidation reactions of organic chemicals in
water. Environmental Toxicology and Chemistry 22 (8),
1743e1754.
Chapin, R.M., 1934. The effect of hydrogen-ion concentration on
the decomposition of hypohalites. Journal of the American
Chemical Society 56 (11), 2211e2215.
Chou, Y.C., Lin, H.P., Sun, S.S., Jwo, J.J., 1993. Kinetic study of the
ferriin oxidation of malonic acid and its derivatives:
implication in the Belousov-Zhabotinskii reaction. The Journal
of Physical Chemistry 97 (32), 8450e8457.
Chowdhury, S., Champagne, P., James McLellan, P., 2010.
Investigating effects of bromide ions on trihalomethanes
and developing model for predicting bromodichloromethane
in drinking water. Water Research 44 (7), 2349e2359.
Cowman, G.A., Singer, P.C., 1996. Effect of bromide ion on
haloacetic acid speciation resulting from chlorination and
chloramination of aquatic humic substances. Environmental
Science and Technology 30 (1), 16e24.
Criquet, J., Allard, S., Salhi, E., Joll, C.A., Heitz, A., von Gunten, U.,
2012. Iodate and iodo-trihalomethane formation during
chlorination of iodide-containing waters: role of bromide.
Environmental Science and Technology 46 (13), 7350e7357.
Criquet, J., Rodriguez, E., Wellauer, S., Salhi, E., Joll, C.A., von
Gunten, U. Reactivity of bromine and chlorine with phenolic
compounds and natural organic matter extracts, in
preparation.

Deborde, M., von Gunten, U., 2008. Reactions of chlorine with


inorganic and organic compounds during water treatment e
kinetics and mechanisms: a critical review. Water Research 42
(1-2), 13e51.
DellErba, A., Falsanisi, D., Liberti, L., Notarnicola, M., Santoro, D.,
2007. Disinfection by-products formation during wastewater
disinfection with peracetic acid. Desalination 215 (1e3),
177e186.
Dodd, M.C., Vu, N.D., Ammann, A., Le, V.C., Kissner, R.,
Pham, H.V., Cao, T.H., Berg, M., von Gunten, U., 2006. Kinetics
and mechanistic aspects of As(III) oxidation by aqueous
chlorine, chloramines, and ozone: relevance to drinking water
treatment. Environmental Science and Technology 40 (10),
3285e3292.
Dubois, J.E., Toullec, J., 1969. Bromination of saturated aliphatic
ketones in an acidic medium e effects of structure on
apparent rate constants for enolization and for bromination of
enol. Journal of the Chemical Society D: Chemical
Communications (6), 292e293.
Duirk, S.E., Collette, T.W., 2006. Degradation of chlorpyrifos in
aqueous chlorine solutions: pathways, kinetics, and modeling.
Environmental Science and Technology 40 (2), 546e551.
Duirk, S.E., Valentine, R.L., 2007. Bromide oxidation and
formation of dihaloacetic acids in chloraminated water.
Environmental Science and Technology 41 (20), 7047e7053.
Duirk, S.E., Tarr, J.C., Collette, T.W., 2008. Chlorpyrifos
transformation by aqueous chlorine in the presence of
bromide and natural organic matter. Journal of Agricultural
and Food Chemistry 56 (4), 1328e1335.
Echigo, S., Minear, R.A., 2006. Kinetics of the reaction of
hypobromous acid and organic matters in water treatment
processes. Water Science and Technology 53 (11), 235e243.
Echigo, S., 2002. Kinetics and Speciation of Brominated
Disinfection By-products During Ozonation. PhD Thesis.
University of Illinois, Urbana-Champaign, Illinois.
Eigen, M., Kustin, K., 1962. The kinetics of halogen hydrolysis.
Journal of the American Chemical Society 84 (8), 1355e1361.
Ershov, B.G., 2004. Kinetics, mechanism and intermediates of
some radiation-induced reactions in aqueous solutions.
Russian Chemical Reviews 73 (1), 101e113.
European Union, 1998. Council Directive 98/83/EC of 3 November
1998 on the Quality of Water Intended for Human
Consumption. European Union, Brussels.
Farkas, L., Lewin, M., 1950. The dissociation constant of
hypobromous acid. Journal of the American Chemical Society
72 (12), 5766e5767.
Farkas, L., Lewin, M., Bloch, R., 1949. The reaction between
hypochlorite and bromides. Journal of the American Chemical
Society 71 (6), 1988e1991.
Finlayson-Pitts, B.J., 2003. The tropospheric chemistry of sea salt:
a molecular-level view of the chemistry of NaCl and NaBr.
Chemical Reviews 103 (12), 4801e4822.
Flury, M., Papritz, A., 1993. Bromide in the natural environment:
occurrence and toxicity. Journal of Environmental Quality 22
(4), 747e758.
Folkes, L.K., Candeias, L.P., Wardman, P., 1995. Kinetics and
mechanisms of hypochlorous acid reactions. Archives of
Biochemistry and Biophysics 323 (1), 120e126.
Furman, C.S., Margerum, D.W., 1998. Mechanism of chlorine
dioxide and chlorate ion formation from the reaction of
hypobromous acid and chlorite ion. Inorganic Chemistry 37
(17), 4321e4327.
Galal-Gorchev, H., Morris, J.C., 1965. Formation and stability of
bromamide bromimide and nitrogen tribromide in aqueous
solution. Inorganic Chemistry 4 (6), 899e905.
Gallard, H., von Gunten, U., 2002. Chlorination of natural organic
matter: kinetics of chlorination and of THM formation. Water
Research 36 (1), 65e74.

w a t e r r e s e a r c h 4 8 ( 2 0 1 4 ) 1 5 e4 2

Gallard, H., Pellizzari, F., Croue, J.P., Legube, B., 2003. Rate
constants of reactions of bromine with phenols in aqueous
solution. Water Research 37 (12), 2883e2892.
Gazda, M., Margerum, D.W., 1994. Reactions of monochloramine

with Br2, Br
3 , HOBr, and OBr e formation of
bromochloramines. Inorganic Chemistry 33 (1), 118e123.
Gazda, M., Dejarme, L.E., Choudhury, T.K., Cooks, R.G.,
Margerum, D.W., 1993. Mass-spectrometric evidence for the
formation of bromochloramine and N-bromo-Nchloromethylamine in aqueous-solution. Environmental
Science and Technology 27 (3), 557e561.
Gerritsen, C.M., Gazda, M., Margerum, D.W., 1993. Nonmetal
redox kinetics e hypobromite and hypoiodite reactions with
cyanide and the hydrolysis of cyanogen halides. Inorganic
Chemistry 32 (25), 5739e5748.
Goncalves, A.A., Gagnon, G.A., 2011. Ozone application in
recirculating aquaculture system: an overview. Ozone-Science
& Engineering 33 (5), 345e367.
Grguric, G., Coston, C.J., 1998. Modeling of nitrate and bromate in
a seawater aquarium. Water Research 32 (6), 1759e1768.
Guo, G., Lin, F., 2009. The bromination kinetics of phenolic
compounds in aqueous solution. Journal of Hazardous
Materials 170 (2-3), 645e651.
Haag, W., Lietzke, M., 1980. A kinetic model for predicting the
concentrations of active halogen species in chlorinated saline
cooling waters. Water Chlorination: Environmental Impact
and Health Effect 3, 415e426.
Haag, W.R., Hoigne, J., 1983. Ozonation of bromide-containing
waters: kinetics of formation of hypobromous acid and
bromate. Environmental Science and Technology 17 (5),
261e267.
Haag, W.R., Hoigne, J., Bader, H., 1984. Improved ammonia
oxidation by ozone in the presence of bromide ion during
water treatment. Water Research 18 (9), 1125e1128.
Hansch, C., Leo, A., Taft, R.W., 1991. A survey of Hammett
substituent constants and resonance and field parameters.
Chemical Reviews 91 (2), 165e195.
Hao, O.J., Davis, A.P., Chang, P.H., 1991. Kinetics of manganese(II)
oxidation with chlorine. Journal of Environmental Engineering
117 (3), 359e374.
Haruta, K., Takeyama, T., 1981. Kinetics of oxidation of aqueous
bromide ion by ozone. The Journal of Physical Chemistry 85
(16), 2383e2388.
Haynes, W.M. (Ed.), 2013. CRC Handbook of Chemistry and
Physics, 93rd ed. CRC Press/Taylor and Francis, Boca Raton, FL.
Internet Version.
Heeb, M.B., Zimmermann-Steffens, S.G., von Gunten, U. The
reactivity of bromine with N-containing compounds, in
preparation.
Hegedus, L., Forsterling, H.-D., Onel, L., Wittmann, M.,
Noszticzius, Z., 2006. Contribution to the chemistry of
the Belousov-Zhabotinsky reaction. Products of the
Ferriin-Bromomalonic acid and the Ferriin-Malonic acid
reactions. The Journal of Physical Chemistry A 110 (47),
12839e12844.
Hofmann, R., Andrews, R.C., 2001. Ammoniacal bromamines: a
review of their influence on bromate formation during
ozonation. Water Research 35 (3), 599e604.
Hoigne, J., Bader, H., 1983a. Rate constants of reactions of ozone with
organic and inorganic compounds in water. I: Non-dissociating
organic compounds. Water Research 17 (2), 173e183.
Hoigne, J., Bader, H., 1983b. Rate constants of reactions of ozone
with organic and inorganic compounds in water. II:
Dissociating organic compounds. Water Research 17 (2),
185e194.
Hoigne, J., Bader, H., 1994. Kinetics of reactions of chlorine dioxide
(OClO) in water. 1. Rate constants for inorganic and organiccompounds. Water Research 28 (1), 45e55.

39

Hu, J., Jin, X., Kunikane, S., Terao, Y., Aizawa, T., 2006.
Transformation of pyrene in aqueous chlorination in the
presence and absence of bromide ion: kinetics, products, and
their aryl hydrocarbon receptor-mediated activities.
Environmental Science and Technology 40 (2), 487e493.
Hua, G., Reckhow, D.A., Kim, J., 2006. Effect of bromide and iodide
ions on the formation and speciation of disinfection
byproducts during chlorination. Environmental Science and
Technology 40 (9), 3050e3056.
Inaba, K., Doi, T., Isobe, N., Yamamoto, T., 2006. Formation of
bromo-substituted triclosan during chlorination by chlorine in
the presence of trace levels of bromide. Water Research 40
(15), 2931e2937.
Inman, G.W., Johnson, J.D., 1984. Kinetics of monobromamine
disproportionation-dibromamine formation in aqueous
ammonia solutions. Environmental Science and Technology
18 (4), 219e224.
Inman, G.W., LaPointe, T.F., Johnson, J.D., 1976. Kinetics of
nitrogen tribromide decomposition in aqueous solution.
Inorganic Chemistry 15 (12), 3037e3042.
Johnson, D.W., Margerum, D.W., 1991. Nonmetal redox kinetics e
a reexamination of the mechanism of the reaction between
hypochlorite and nitrite ions. Inorganic Chemistry 30 (25),
4845e4851.
Johnson, J.D., Overby, R., 1971. Bromine and bromamine
disinfection chemistry. ASCE Journal of the Sanitary
Engineering Division 97 (5), 617e628.
Jones, D.B., Saglam, A., Song, H., Karanfil, T., 2012. The impact of
bromide/iodide concentration and ratio on iodinated
trihalomethane formation and speciation. Water Research 46
(1), 11e20.
Jorquera, M.A., Valencia, G., Eguchi, M., Katayose, M.,
Riquelme, C., 2002. Disinfection of seawater for hatchery
aquaculture systems using electrolytic water treatment.
Aquaculture 207 (3-4), 213e224.
King, D.W., 1998. Role of carbonate speciation on the oxidation
rate of Fe(II) in aquatic systems. Environmental Science and
Technology 32 (19), 2997e3003.
Korshin, G.V., 2011. In: Tratnyek, P.G., Grundl, T., Haderlein, S.B.
(Eds.), Aquatic Redox Chemistry. American Chemical Society,
Washington, D.C, pp. 223e245.
Krasner, S.W., Weinberg, H.S., Richardson, S.D., Pastor, S.J.,
Chinn, R., Sclimenti, M.J., Onstad, G.D., Thruston, A.D., 2006.
Occurrence of a new generation of disinfection byproducts.
Environmental Science and Technology 40 (23), 7175e7185.
Kumar, K., Margerum, D.W., 1987. Kinetics and mechanism of
general-acid-assisted oxidation of bromide by hypochlorite
and hypochlorous acid. Inorganic Chemistry 26 (16),
2706e2711.
Lahoutifard, N., Lagrange, P., Lagrange, J., Scott, S.L., 2002.
Kinetics and mechanism of nitrite oxidation by HOBr/BrO in
atmospheric water and comparison with oxidation by HOCl/
ClO. The Journal of Physical Chemistry A 106 (49),
11891e11896.
Le Roux, J., Gallard, H., Croue, J.P., 2012. Formation of NDMA and
halogenated DBPs by chloramination of tertiary amines: the
influence of bromide ion. Environmental Science and
Technology 46 (3), 1581e1589.
Lee, Y., von Gunten, U., 2009. Transformation of 17 alphaethinylestradiol during water chlorination: effects of
bromide on kinetics, products, and transformation pathways.
Environmental Science and Technology 43 (2), 480e487.
Lee, Y., von Gunten, U., 2012. Quantitative structure-activity
relationships (QSARs) for transformation of organic
micropollutants during oxidative water treatment. Water
Research 46 (19), 6177e6195.
Lee, Y., Yoon, J., von Gunten, U., 2005. Kinetics of the oxidation of
phenols and phenolic endocrine disruptors during water

40

w a t e r r e s e a r c h 4 8 ( 2 0 1 4 ) 1 5 e4 2

treatment with ferrate (Fe(VI)). Environmental Science and


Technology 39 (22), 8978e8984.
Lee, Y., Escher, B.I., von Gunten, U., 2008. Efficient removal of
estrogenic activity during oxidative treatment of waters
containing steroid estrogens. Environmental Science and
Technology 42 (17), 6333e6339.
Lei, H., Marinas, B.J., Minear, R.A., 2004. Bromamine
decomposition kinetics in aqueous solutions. Environmental
Science and Technology 38 (7), 2111e2119.
Lei, H., Minear, R.A., Marinas, B.J., 2006. Cyanogen bromide
formation from the reactions of monobromamine and
dibromamine with cyanide ion. Environmental Science and
Technology 40 (8), 2559e2564.
Liu, Q., Margerum, D.W., 2001. Equilibrium and kinetics of
bromine chloride hydrolysis. Environmental Science and
Technology 35 (6), 1127e1133.
Liu, Q., Schurter, L.M., Muller, C.E., Aloisio, S., Francisco, J.S.,
Margerum, D.W., 2001. Kinetics and mechanisms of aqueous
ozone reactions with bromide, sulfite, hydrogen sulfite, iodide,
and nitrite ions. Inorganic Chemistry 40 (17), 4436e4442.
Liu, C., von Gunten, U., Croue, J.P., 2012. Enhanced bromate
formation during chlorination of bromide-containing waters
in the presence of CuO: catalytic disproportionation of
hypobromous acid. Environmental Science and Technology 46
(20), 11054e11061.
Luh, J., Marinas, B.J., 2012. Bromide Ion effect on Nnitrosodimethylamine formation by monochloramine.
Environmental Science and Technology 46 (9), 5085e5092.
Magazinovic, R.S., Nicholson, B.C., Mulcahy, D.E., Davey, D.E.,
2004. Bromide levels in natural waters: its relationship to
levels of both chloride and total dissolved solids and the
implications for water treatment. Chemosphere 57 (4),
329e335.
Melichova, Z., Melichercik, M., Olexova, A., Treindl, L., 1995.
Chemical oscillators based on the oxidation of Mn(II) by some
halogens. Reaction Kinetics and Catalysis Letters 55 (1), 51e57.
Melichova, Z., Treindl, L., Valent, I., 2001. Kinetics and
mechanism of the autocatalytic oxidation of Mn(II) by
bromine. Reaction Kinetics and Catalysis Letters 74 (1), 79e86.
Millero, F.J., 1974. Physical-chemistry of seawater. Annual Review
of Earth and Planetary Sciences 2, 101e150.
Nagy, P., Beal, J.L., Ashby, M.T., 2006. Thiocyanate is an efficient
endogenous scavenger of the phagocytic killing agent
hypobromous acid. Chemical Research in Toxicology 19 (4),
587e593.
Neemann, J., Hulsey, R., Rexing, D., Wert, E., 2004. Controlling
bromate formation during ozonation with chlorine and
ammonia. Journal American Water Works Association 96 (2),
26e29.
Negi, S.C., Shah, B., Banerji, K.K., 1987. Kinetics and mechanism
of the oxidation of aliphatic alcohols by sodium hypobromite.
Oxidation Communications 10 (1-2), 85e94.
Odeh, I.N., Nicoson, J.S., Hartz, K.E.H., Margerum, D.W., 2004.
Kinetics and mechanisms of bromine chloride reactions with
bromite and chlorite ions. Inorganic Chemistry 43 (23),
7412e7420.
Pan, Y., Zhang, X., 2013. Four groups of new aromatic halogenated
disinfection byproducts: effect of bromide concentration on
their formation and speciation in chlorinated drinking water.
Environmental Science and Technology 47 (3), 1265e1273.
Parkhurst, D.L., Appelo, C.A.J., 1999. Users Guide to PHREEQC
(Version 2) e A Computer Program for Speciation, Batch
Reaction, One-dimensional Transport, and Inverse
Geochemical Calculations. Water-Resources Investigations
Report 99-4259. Water-Resources Investigations, Denver,
Colorado.
Pattison, D.I., Davies, M.J., 2001. Absolute rate constants for the
reaction of hypochlorous acid with protein side chains and

peptide bonds. Chemical Research in Toxicology 14 (10),


1453e1464.
Pattison, D.I., Davies, M.J., 2004. Kinetic analysis of the reactions
of hypobromous acid with protein components: implications
for cellular damage and use of 3-bromotyrosine as a marker of
oxidative stress. Biochemistry 43 (16), 4799e4809.
Pattison, D.I., Davies, M.J., 2006. Reactions of myeloperoxidasederived oxidants with biological substrates: gaining chemical
insight into human inflammatory diseases. Current Medicinal
Chemistry 13 (27), 3271e3290.
Perlmutter-Hayman, B., Weissmann, Y., 1962. The kinetics of the
oxidation of ethanol and of acetaldehyde by bromine in
aqueous solution. Influence of pH. Journal of the American
Chemical Society 84 (12), 2323e2326.
Perlmutter-Hayman, B., Weissmann, Y., 1969. Oxidation of 2propanol by bromine and by hypobromous acid in aqueous
solution. Journal of the American Chemical Society 91 (3),
668e672.
Pinkernell, U., von Gunten, U., 2001. Bromate minimization
during ozonation: mechanistic considerations. Environmental
Science and Technology 35 (12), 2525e2531.
Piriou, P., Soulet, C., Acero, J.L., Bruchet, A., von Gunten, U.,
Suffet, I.H., 2007. Understanding medicinal taste and odour
formation in drinking waters. Water Science and Technology
55 (5), 85e94.
Pope, P.G., Speitel, G.E., 2008. In: Karanfil, T., Krasner, S.W.,
Westerhoff, P., Yuefeng, X. (Eds.), Disinfection By-products in
Drinking Water. American Chemical Society, Washington, DC,
pp. 182e197.
Prutz, W.A., 1999. Consecutive halogen transfer between various
functional groups induced by reaction of hypohalous acids:
NADH oxidation by halogenated amide groups. Archives of
Biochemistry and Biophysics 371 (1), 107e114.
Prutz, W.A., Kissner, R., Koppenol, W.H., Ruegger, H., 2000. On the
irreversible destruction of reduced nicotinamide nucleotides
by hypohalous acids. Archives of Biochemistry and Biophysics
380 (1), 181e191.
Prutz, W.A., Kissner, R., Nauser, T., Koppenol, W.H., 2001. On the
oxidation of cytochrome c by hypohalous acids. Archives of
Biochemistry and Biophysics 389 (1), 110e122.
Richardson, S.D., Thruston, A.D., Rav-Acha, C., Groisman, L.,
Popilevsky, I., Juraev, O., Glezer, V., McKague, A.B., Plewa, M.J.,
Wagner, E.D., 2003. Tribromopyrrole, brominated acids, and
other disinfection byproducts produced by disinfection of
drinking water rich in bromide. Environmental Science and
Technology 37 (17), 3782e3793.
Rodil, R., Benito Quintana, J., Cela, R., 2012. Transformation of
phenazone-type drugs during chlorination. Water Research 46
(7), 2457e2468.
Rook, J.J., Gras, A.A., Van der Heijden, B.G., De Wee, J., 1978.
Bromide oxidation and organic substitution in water
treatment. Journal of Environmental Science and Health, Part
A: Environmental Science and Engineering 13 (2), 91e116.
Roth, W.A., 1929. On the thermochemistry of chlorine and
hypochlorous acid ((Zur Thermochemie des Chlors und der
unterchlorigen Saure)). Zeitschrift fur physikalische Chemie
Abteilung A 145, 289e297.
Ruasse, M.F., Aubard, J., Galland, B., Adenier, A., 1986. Kinetic
study of the fast halogen-trihalide ion equilibria in protic
media by the Raman-laser temperature-jump technique. A
non-diffusion-controlled ion-molecule reaction. The Journal
of Physical Chemistry 90 (18), 4382e4388.
Saiz-Lopez, A., von Glasow, R., 2012. Reactive halogen chemistry
in the troposphere. Chemical Society Reviews 41 (19),
6448e6472.
Sasaki, Y., 1990. Tentative reaction network for bromate bromide
ferroin system and simulation of behavior of the system
under batch reactor and continuous-flow stirred tank reactor

w a t e r r e s e a r c h 4 8 ( 2 0 1 4 ) 1 5 e4 2

conditions. Bulletin of the Chemical Society of Japan 63 (12),


3521e3527.
Schwarz, H.A., Bielski, B.H.J., 1986. Reactions of HO2 And O
2 with
iodine and bromine and the I
2 and I atom reduction
potentials. The Journal of Physical Chemistry 90 (7),
1445e1448.
Sedlak, D.L., von Gunten, U., 2011. The chlorine dilemma. Science
331 (6013), 42e43.
Shilov, E.A., 1938. On the calculation of the dissociation constants
of hypohalogenous acids from kinetic data. Journal of the
American Chemical Society 60 (2), 490e491.
Singer, P.C., 1994. Control of disinfection by-products in
drinking-water. Journal of Environmental Engineering 120 (4),
727e744.
Sivey, J.D., Arey, J.S., Tentscher, P.R., Roberts, A.L., 2013. Reactivity
of BrCl, Br2, BrOCl, Br2O, and HOBr toward dimethenamid in
solutions of bromide aqueous free chlorine. Environmental
Science and Technology 47 (3), 1330e1338.
Skaff, O., Pattison, D.I., Davies, M.J., 2007. Kinetics of
hypobromous acid-mediated oxidation of lipid components
and antioxidants. Chemical Research in Toxicology 20 (12),
1980e1988.
Skaff, O., Pattison, D.I., Davies, M.J., 2008. The vinyl ether linkages
of plasmalogens are favored targets for myeloperoxidasederived oxidants: a kinetic study. Biochemistry 47 (31),
8237e8245.
Sommariva, R., von Glasow, R., 2012. Multiphase halogen
chemistry in the tropical Atlantic Ocean. Environmental
Science and Technology 46 (19), 10429e10437.
Soulard, M., Bloc, F., Hatterer, A., 1981. Diagrams of existence of
chloramines and bromamines in aqueous solution. Journal of
the Chemical Society, Dalton Transactions (12), 2300e2310.
Symons, J.M., Krasner, S.W., Simms, L.A., Sclimenti, M., 1993.
Measurement of THM and precursor concentrations revisited:
the effect of bromide ion. Journal of the American Water
Works Association 85 (1), 51e62.
Symons, J.M., Zheng, M.C.H., 1997. Technical note: Does hydroxyl
radical oxidize bromide to bromate? Journal American Water
Works Association 89 (6), 106e109.
Tee, O.S., Iyengar, N.R., 1985. The bromination of salicylate
anions e evidence for the participation of the orthocarboxylate group. Journal of Organic Chemistry 50 (23),
4468e4473.
Tee, O.S., Bennett, J.M., 1988. Catalysis of the bromination of
phenols and phenoxide ions in aqueous solution by alphacyclodextrin. Journal of the American Chemical Society 110
(1), 269e274.
Tee, O.S., Iyengar, N.R., Kraus, B., 1985. The bromination of parahydroxybenzoic acid in aqueous solution e reaction via the
minor para-carboxyphenoxide anion tautomer. Journal of
Organic Chemistry 50 (7), 973e976.
Tee, O.S., Paventi, M., Bennett, J.M., 1989. Kinetics and
mechanisms of the bromination of phenols and phenoxide
ions in aqueous solution. Diffusion-controlled rates. Journal of
the American Chemical Society 111 (6), 2233e2240.
Thomas, E.L., Bozeman, P.M., Jefferson, M.M., King, C.C., 1995.
Oxidation of bromide by the human-leukocyte enzymes
myeloperoxidase and eosinophil peroxidase e formation of
bromamines. Journal of Biological Chemistry 270 (7),
2906e2913.
Toth, Z., Fabian, I., 2000. Kinetics and mechanism of the initial
phase of the bromine-chlorite ion reaction in aqueous
solution. Inorganic Chemistry 39 (20), 4608e4614.
Toth, Z., Fabian, I., 2004. Oxidation of chlorine(III) by
hypobromous acid: kinetics and mechanism. Inorganic
Chemistry 43 (8), 2717e2723.
Tremblay-Goutaudier, C., Rizk-Ouaini, R., Cohen-Adad, M.T.,
Jenin, P., 1994. Monobromamine in aqueous solution. II:

41

Stability and formation via electrochemical pathways


((Monobromamine en milieu aqueux. II: Stabilite et
elaboration par voie electrochimique)). Journal de Chimie
Physique et de Physico-Chimie Biologique 91 (5), 535e546.
Trofe, T.W., Inman, G.W., Johnson, J.D., 1980. Kinetics of
monochloramine decomposition in the presence of bromide.
Environmental Science and Technology 14 (5), 544e549.
Troy, R.C., Margerum, D.W., 1991. Nonmetal redox kinetics e
hypobromite and hypobromous acid reactions with iodide and
with sulfite and the hydrolysis of bromosulfate. Inorganic
Chemistry 30 (18), 3538e3543.
Troy, R.C., Kelley, M.D., Nagy, J.C., Margerum, D.W., 1991.
Nonmetal redox kinetics e iodine monobromide reaction with
iodide-ion and the hydrolysis of IBr. Inorganic Chemistry 30
(25), 4838e4845.
United States Environmental Protection Agency, 2009. National
Primary Drinking Water Regulations. United States
Environmental Protection Agency.
Valentine, R.L., 1986. Bromochloramine oxidation of N,N-diethylp-phenylenediamine in the presence of monochloramine.
Environmental Science and Technology 20 (2), 166e170.
Valero, F., Arbos, R., 2010. Desalination of brackish river water
using Electrodialysis Reversal (EDR): control of the THMs
formation in the Barcelona (NE Spain) area. Desalination 253
(1-3), 170e174.
Vikesland, P.J., Ozekin, K., Valentine, R.L., 2001. Monochloramine
decay in model and distribution system waters. Water
Research 35 (7), 1766e1776.
Vogt, R., Crutzen, P.J., Sander, R., 1996. A mechanism for halogen
release from sea-salt aerosol in the remote marine boundary
layer. Nature 383 (6598), 327e330.
von Gunten, U., 2003. Ozonation of drinking water: Part II.
Disinfection and by-product formation in presence of
bromide, iodide or chlorine. Water Research 37 (7), 1469e1487.
von Gunten, U., Hoigne, J., 1992. Factors controlling the formation
of bromate during ozonation of bromide-containing waters.
Aqua 41 (5), 299e304.
von Gunten, U., Hoigne, J., 1994. Bromate formation during
ozonization of bromide-containing waters: interaction of
ozone and hydroxyl radical reactions. Environmental Science
and Technology 28 (7), 1234e1242.
von Gunten, U., Oliveras, Y., 1997. Kinetics of the reaction
between hydrogen peroxide and hypobromous acid:
implication on water treatment and natural systems. Water
Research 31 (4), 900e906.
von Gunten, U., Oliveras, Y., 1998. Advanced oxidation of
bromide-containing waters: bromate formation mechanisms.
Environmental Science and Technology 32 (1), 63e70.
von Gunten, U., Salhi, E., Schmidt, C.K., Arnold, W.A., 2010.
Kinetics and mechanisms of N-nitrosodimethylamine
formation upon ozonation of N,N-dimethylsulfamidecontaining waters: bromide catalysis. Environmental Science
and Technology 44 (15), 5762e5768.
von Sonntag, C., von Gunten, U., 2012. Chemistry of Ozone in
Water and Wastewater Treatment: From Basic Principles to
Applications. IWA Publishing, London.
Voudrias, E.A., Reinhard, M., 1988. Reactivities of hypochlorous
and hypobromous acid, chlorine monoxide, hypobromous
acidium ion, chlorine, bromine, and bromine chloride in
electrophilic aromatic-substitution reactions with p-xylene in
water. Environmental Science and Technology 22 (9),
1049e1056.
Wajon, J.E., Morris, J.C., 1982. Rates of formation of N-bromo
amines in aqueous solution. Inorganic Chemistry 21 (12),
4258e4263.
Wang, T.X., Margerum, D.W., 1994. Kinetics of reversible chlorine
hydrolysis e temperature dependence and general acid baseassisted mechanisms. Inorganic Chemistry 33 (6), 1050e1055.

42

w a t e r r e s e a r c h 4 8 ( 2 0 1 4 ) 1 5 e4 2

Wang, T.X., Kelley, M.D., Cooper, J.N., Beckwith, R.C.,


Margerum, D.W., 1994. Equilibrium, kinetic, and UVSpectral characteristics of aqueous bromine chloride,
bromine, and chlorine species. Inorganic Chemistry 33 (25),
5872e5878.
Weinberg, H.S., Delcomyn, C.A., Unnam, V., 2003. Bromate in
chlorinated drinking waters: occurrence and implications for
future regulation. Environmental Science and Technology 37
(14), 3104e3110.
Weiss, S.J., Test, S.T., Eckmann, C.M., Roos, D., Regiani, S., 1986.
Brominating oxidants generated by human eosinophils.
Science 234 (4773), 200e203.
Werschkun, B., Sommer, Y., Banerji, S., 2012. Disinfection byproducts in ballast water treatment: an evaluation of
regulatory data. Water Research 46 (16), 4884e4901.
Westerhoff, P., Song, R., Amy, G., Minear, R., 1998. NOMs role
in bromine and bromate formation during ozonation.
Journal of the American Water Works Association 90 (2),
82e94.
Westerhoff, P., Chao, P., Mash, H., 2004. Reactivity of natural
organic matter with aqueous chlorine and bromine. Water
Research 38 (6), 1502e1513.

World Health Organization, 2011. Guidelines for Drinking-water


Quality, fourth ed. World Health Organization.
Wu, J.G., Laird, D.A., 2003. Abiotic transformation of chlorpyrifos
to chlorpyrifos oxon in chlorinated water. Environmental
Toxicology and Chemistry 22 (2), 261e264.
Wu, W.J., Chen, Y.H., dAvignon, A., Hazen, S.L., 1999. 3bromotyrosine and 3,5-dibromotyrosine are major products of
protein oxidation by eosinophil peroxidase: potential markers
for eosinophil-dependent tissue injury in vivo. Biochemistry
38 (12), 3538e3548.
Xu, B., Gao, N.Y., Cheng, H.F., Hu, C.Y., Xia, S.J., Sun, X.F.,
Wang, X.J., Yang, S.G., 2009. Ametryn degradation by aqueous
chlorine: kinetics and reaction influences. Journal of
Hazardous Materials 169 (1e3), 586e592.
Zhang, X., Echigo, S., Lei, H., Smith, M.E., Minear, R.A., Talley, J.W.,
2005. Effects of temperature and chemical addition on the
formation of bromoorganic DBPs during ozonation. Water
Research 39 (2-3), 423e435.
Zhao, Y., Anichina, J., Lu, X., Bull, R.J., Krasner, S.W., Hrudey, S.E.,
Li, X.-F., 2012. Occurrence and formation of chloro- and
bromo-benzoquinones during drinking water disinfection.
Water Research 46 (14), 4351e4360.

You might also like