You are on page 1of 7

Science of the Total Environment 506507 (2015) 252258

Contents lists available at ScienceDirect

Science of the Total Environment


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

Oxidation of diclofenac by potassium ferrate (VI): Reaction kinetics and


toxicity evaluation
Yingling Wang a,c, Haijin Liu a, Guoguang Liu a,b,, Youhai Xie a, Shuyan Gao a
a
School of Environment, Henan Normal University, Henan Key Laboratory for Environmental Pollution Control, Key Laboratory for Yellow River and Huaihe River Water Environment and Pollution
Control, Ministry of Education, Xinxiang 453007, PR China
b
Faculty of Environmental Science and Engineering, Guangdong University of Technology, Guangzhou 510006, PR China
c
School of Basic Medicine, Xinxiang Medical University, Xinxiang 453003, PR China

H I G H L I G H T S

The reaction kinetics of diclofenac oxidation by ferrate(VI) was investigated.


The inuences of pH and temperature on reactivities were elucidated.
Microtox bioassay was employed to evaluate acute toxicity of reaction solutions.
DCF intermediates were observed to be more toxic than the parent compound.

a r t i c l e i n f o a b s t r a c t

Article history: The reaction kinetics and toxicity of diclofenac (DCF) oxidation by ferrate (VI) under simulated water disinfection
Received 12 September 2014 conditions were investigated. Experimental results indicated that the reaction between DCF and Fe(VI) followed
Received in revised form 31 October 2014 rst-order kinetics with respect to each reactant. Furthermore, the effects of pH and temperature on DCF
Accepted 31 October 2014
oxidation by Fe(VI) were elucidated using a systematic examination. The apparent second-order rate
Available online xxxx
constants (kapp) increased signicantly from 2.54 to 11.6 M 1 s 1, as the pH of the solution decreased
Editor: D. Barcelo from 11.0 to 7.0, and the acidbase equilibriums of Fe(VI) and DCF were proposed to explain the pH dependence
of kapp. The acute toxicity of DCF solution during Fe(VI) oxidation was evaluated using a Microtox bioassay.
Keywords: Overall, the DCF degradation process resulted in a rapid increase of the inhibition rate of luminescent bacteria.
Diclofenac These toxicity tests suggest that the formation of enhanced toxic intermediates during the Fe(VI) disinfection
Ferrate (VI) process may pose potential health risk to consumers.
Reaction kinetics 2014 Elsevier B.V. All rights reserved.
Toxicity
Water treatment

1. Introduction subsequent to enterohepatic circulation (Forrez et al., 2010; Zhang


et al., 2008). Typically, only about 30% of DCF is removed in conven-
Diclofenac (DCF), a synthetic non-steroidal anti-inammatory drug, tional sewage treatment plants (STPs) (Surez et al., 2008), due in
is one of the most commonly used pain killers, which is primarily used part to its low biodegradability (Joss et al., 2006) and its limited
clinically as the sodium salt. The global consumption of DCF is estimated sorption properties onto activated sludge (Ternes et al., 2004). DCF
to be approximately 940 tons per year, with a dened daily dose of has been detected in STP efuent and surface water at levels of up
100 mg (Michael et al., 2014; Vieno et al., 2007; Zhang et al., 2008). to 4.7 g L 1 and 1.2 g L 1, respectively, and even in groundwater
Approximately 65% of the dosage is usually excreted through urine, and tap water at concentrations of up to 380 ng L 1 and below
mainly as hydroxylated metabolites that are conjugated with glucuronides 10 ng L 1, respectively (Aguinaco et al., 2012; Heberer, 2002). DCF
is considered to be one of the most relevant compounds in terms of
ecotoxicity and persistence in the environment. The contamination
of primary drinking water sources with DCF would constitute a serious
Corresponding author at: School of Environment, Henan Normal University, Henan threat to public health.
Key Laboratory for Environmental Pollution Control, Key Laboratory for Yellow River and
Huaihe River Water Environment and Pollution Control, Ministry of Education, Xinxiang
Therefore, new and more stringent standards in the EU water
453007, PR China. Tel.: +86 373 3325971; fax: +86 373 3326335. framework directive necessitate the advanced treatment of secondary
E-mail address: liugg@gdut.edu.cn (G. Liu). efuents toward the reduction of trace organic compounds such as

http://dx.doi.org/10.1016/j.scitotenv.2014.10.114
0048-9697/ 2014 Elsevier B.V. All rights reserved.
Y. Wang et al. / Science of the Total Environment 506507 (2015) 252258 253

DCF, prior to efuent discharge into ambient waterways (Huebner and Shandong (China). Fe(VI) stock solutions were prepared in a 5 mM
Jekel, 2013), utilizing technologies such as photocatalysis with TiO2 Na2HPO4/1 mM Na2B4O7 10H2O buffer (pH 9.0 0.2), the pH at
(Miranda-Garca et al., 2011) or Fenton reagent (Bae et al., 2013) or which Fe(VI) is the most stable (Lee et al., 2005), and quickly ltrated
H2O2 (Shu et al., 2013), chlorination (Soufan et al., 2012), ozonation with a 0.22 m hydrophilic nylon syringe lter (Shanghai, China).
(Sein et al., 2008) and sonolysis procedures (Nie et al., 2014). During Concentrations of Fe(VI) in solutions were determined spectroscop-
the phototransformation process, the major byproducts of DCF are ically at a maximum absorbance of 510 nm using an UVVis spectro-
composed of substituted diphenylamines and carbazoles, which are of photometer (722 grating spectrophotometer, Shanghai, China).
critical importance for further studies on the ecotoxicological effects A molar absorption coefcient of 510nm = 1150 M cm 1 was
of the drug in the ambient environment (Svanfelt and Kronberg, used to determine the Fe(VI) concentration at pH 9.0 0.2
2011). Although DCF was rapidly degraded during drinking water disin- (Lee et al., 2009; Luo et al., 2011). Stock solutions were utilized with-
fection using free chlorine, it may generate chlorinated disinfection in 3 h of preparation to minimize the autodecomposition of Fe(VI).
byproducts (Soufan et al., 2012). Ozonation was highly effective for A stock solution of DCF was prepared in ultrapure Milli-Q water at
the treatment of DCF in aqueous solutions with second-order rate con- concentrations of 3.0 mM. HPLC-grade solvents of acetonitrile and
stants of ~ 106 M 1 s1 (Huber et al., 2003) and 1.8 104 M 1 s 1 methanol were obtained from Suqian Guoda Chemical Reagents
(Vogna et al., 2004). However, bromated byproducts that are generated Co. Ltd. (Jiangsu, China). Additional analytical grade reagents (Na2S2O3,
through ozonation should be considered, as well as hydroxyl radicals, NaCl, phosphate, etc.) were employed without further purication.
which have no selectivity for target contaminants (Garoma and Freeze-dried Vibrio (V.) scheri bacteria were obtained from Nanjing
Matsumoto, 2009). Ultrasound irradiation techniques usually suffer CAS Kuake Technology Co. Ltd. (Nanjing, China). Ultrapure water
from high operational costs and relatively slow removal kinetics from a Milli-Q water purication system (Millipore, USA) was used
(Hartmann et al., 2008). Since there is, as yet, no efcacious process in the preparation of all aqueous solutions. The pH of test solutions
for the oxidative removal of DCF from water, the development of was adjusted to the desired level through the addition of either
more promising technologies is warranted (Jiang, 2013). NaOH or phosphoric acid.
Ferrate has been widely used in potable water treatment for the
enhancement of coagulation, removal of micropollutants, and toward 2.2. DCF oxidation experiments
the remediation of contaminated groundwater (Jiang et al., 2012;
Sharma, 2013; Zimmermann et al., 2012). Although ferrate exhibits a Kinetic experiments of DCF oxidation by K2FeO4 were conducted in
strong oxidizing potential (E0H;K 2 FeO4 2:20V at acidic conditions) 60 mL brown glass bottles with stopples on a collector-type magnetic
(Jiang, 2014), it is considered to be a selective oxidizing agent due to stirrer at pH 9.0 0.2 and 23 2C unless otherwise specied. The
its high reactivity with electron-rich organic moieties of target com- reactions were initiated by spiking an aliquot of the Fe(VI) stock solu-
pounds, such as phenol, olen, amine, and aniline (Anquandah et al., tion into pre-equilibrated solutions containing DCF under rapid stirring.
2011; Jiang, 2014; Sharma, 2013). The instability of Fe(VI) salts has lim- In 50.00 mL reaction mixture solutions, the initial concentration of DCF
ited their use in eld treatment applications; however, recent efforts was 0.030 mM while the initial concentration of Fe(VI) was in the range
have focused on the development of technologies for onsite Fe(VI) from 0.15 mM to 0.90 mM. At preselected time intervals, 2.00 mL of the
generation (Licht and Yu, 2005). Fe(VI) is also attracting growing atten- reaction liquid was transferred using a pipette, from the brown glass
tion because it is non-halogenating and generates insoluble and envi- bottle to a small beaker, which contained 50 L of preloaded Na2S2O3
ronmentally benign reduction products (e.g., Fe(OH)3(s)) that may (0.10 M), to immediately terminate the reaction. It was then ltrated
be readily removed through sedimentation/ltration. Furthermore, with a 0.22 m nylon lter into high performance liquid chromatogra-
Fe(OH)3(s) products can act as coagulants and adsorbents to promote phy (HPLC) vial. The concentrations of residual DCF in the resulting
the physical removal of pharmaceuticals and other contaminants of reaction mixtures were analyzed directly using HPLC. Changes in
concern (e.g., disinfection byproducts precursors) (Jiang, 2014; Licht the pH values were negligible throughout the entire process, and all
and Yu, 2005). Although ferrate is currently utilized as an oxidant for of the concentrations of Fe(VI) and pH reported were initial values.
the removal of micropollutants from aqueous media, the degradation All samples were prepared in triplicate.
of DCF by Fe(VI) has rarely been investigated. It is also important to
determine whether ferrate (VI) treated water that contains toxic or 2.3. Analytical methods
mutagenic substances, such as DCF, will allay public health concerns
when this novel chemical is employed for water treatment. The concentrations of DCF solutions were determined using a
This study aims to determine the rate constants and inuential reversed-phase HPLC system, which consisted of two Waters 1525
factors that are involved with the reaction of Fe(VI) with DCF and Binary HPLC pumps and a Waters 2998 Photodiode Array detector
evaluate the changes in toxicity during the Fe(VI) oxidation of DCF (Waters, Massachusetts, USA). Analytical column temperatures were
using a Microtox bioassay testing system. The reaction kinetics and ef- controlled with a Model 1500 Column Heater (Waters, and Product
fects of pH and temperature on DCF oxidation by Fe(VI) are discussed of Singapore). The analytical column was a 150 mm 4.6 mm Waters
using a systematic examination under simulated water disinfection con- C18 column, (particle size 5 m). A Waters guard column (C18,
ditions. Toxicity studies of a DCF laced solution during Fe(VI) oxidation 4.6 mm 20 mm, particle size 5 m) was employed to protect the an-
were carried out using acute toxicity luminescent bacteria tests. alytical column (both purchased from Waters), and the injection vol-
These results will assist with elucidating the fate and behavior of ume was 20 L. The mobile phase consisted of a mixture of 75% HPLC-
DCF to determine whether ferrate (VI) reagents might produce grade methanol and 25% Milli-Q water (containing 1% acetic acid) at
mutagenic byproducts, thus contributing to the establishment of a constant ow rate of 1.0 mL min 1, with the detection wavelength
criteria for the safe discharge of DCF into ambient aquatic environments. set at 276 nm. The pH of the solution was measured using a Mettler
Toledo Delta 320 pH meter, whereas the solution temperatures
2. Materials and methods were simultaneously controlled by an HX-08 Cryostat (Shanghai
Bilon Instruments Co. Ltd., China).
2.1. Standards and reagents
2.4. Toxicity measurements
DCF, 2-[(2,6-dichlorophenyl) amino] benzeneacetic acid, sodium
salt (98% purity), was purchased from J&K Chemical Co. Ltd. (Beijing, The samples (10.00 mL) that were collected after different oxidation
China). Solid potassium ferrate (K2FeO4, 86%) was purchased from times were pretreated using the following procedures: (1) immediate
254 Y. Wang et al. / Science of the Total Environment 506507 (2015) 252258

quenching with 150 L 0.10 M Na2S2O3; (2) addition of 0.30 g NaCl to


each sample in order to attain a concentration of 3.0%; (3) ltration
0.0 a
using 0.22 m nylon lters to remove Fe(OH)3 ; (4) adjustment
-0.5
of the sample (2.00 mL) pH to ~ 7.0 with 350 L M/15 H3PO4. The tox-
icity of the resulting samples was subsequently examined using a

ln[DCF]/[DCF]0
Microtox Model DXY-2 toxicity analyzer, which measures the -1.0
capacity of the byproducts to inhibit the bioluminescence of the
bacterium V. scheri (Calza et al., 2006; Marco-Urrea et al., 2010). -1.5
Freeze-dried bacteria pellets were reconstituted by adding 1.00 mL 0.15 mM
of diluent (2.5% NaCl). Briey, the samples were tested in a medium -2.0 0.30 mM
containing 3.0% NaCl; toxicity data were recorded on 15 min expo- 0.45 mM
0.60 mM
sure of 10.00 L reconstituted bacteria solution to each sample at -2.5 0.75 mM
23 2 C. The inhibition of the luminescence, in contrast to blank 0.90 mM
controls, to quantify the percentage of inhibition (I%), was calculated -3.0
based on Eq. (1), where Ix and I0 are the luminosity of sample solution 0 1 2 3 4 5 6 7 8 9 10 11
and blank solution without DCF and K2FeO4, respectively. The initial Time (min)
DCF concentration was 0.030 mM.
5.0

I0 Ix 4.5 b
I%  100% 1
I0 4.0
y = - 0.06825 + 5.21891x
3.5 2
In addition, the total organic carbon (TOC) was determined by a R = 0.9991

1000kobs (s-1)
multi N/C 2100S/1 TOC Analyzer (Analytik Jena AG, Germany). The 3.0
removal rate (%) of TOC was calculated by Eq. (2):
2.5

2.0
TOC0 TOCx
 100% 2
TOC0 1.5

1.0
where [TOC]0 is the initial TOC concentration of 0.030 mM DCF, and 0.5
[TOC]x is the TOC concentration at certain reaction time.
0.0
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
3. Results and discussion
[Fe(VI)] (mM)
3.1. Reaction orders and constants of ferrate with DCF
Fig. 1. Determination of reaction order for DCF oxidation by ferrate: (a) plot of ln
[DCF]/[DCF]0 vs. reaction time at different initial Fe(VI) concentrations; and (b) linear
Initially, the DCF oxidation experiments at different initial Fe(VI)
relationship between measured pseudo-rst-order kinetic constants (kobs, s1) and
concentrations were carried out at pH 9.0 0.2 under 23 2 C. The initial Fe(VI) concentrations. Reaction conditions: [DCF]0 = 0.030 mM, pH = 9.0 0.2,
rate expression for the reaction of Fe(VI) with DCF may be expressed T = 23 2 C.
as the following Eq.:
with respect to each reactant. Accordingly, the rate expression for the
dDC F m n oxidation of DCF by Fe(VI) could be described as:
kapp DC F FeVI 3
dt
dDCF
where [DCF] and [Fe(VI)] are the DCF and Fe(VI) concentrations, m and kobs DCF kapp FeVIDCF 5
dt
n are the orders of the reaction with respect to the concentrations of
DCF and Fe(VI), respectively, and kapp is the apparent kinetic constant. For each experiment performed with various concentrations of Fe(VI),
The kinetic studies were investigated under pseudo-rst-order condi- kapp values were determined by considering Eq. (5). To the best of our
tions with a molar ratio of Fe(VI) to DCF 5:1 in a batch reactor. knowledge the current study is the rst to report the rate constants of
Eq. (3) can be rewritten as: the oxidation of DCF by commercial grade ferrate. Nevertheless, the rate
constants reported here agree well with previous studies, which investi-
dDC F m gated the chlorination and ozonation of DCF, and the oxidation of other
kobs DC F 4
dt pharmaceuticals with Fe(VI). The value of kapp was calculated to be
5.17 M1 s1 at a molar ratio of Fe(VI) to DCF of 15:1, which is very
where kobs = kapp[Fe(VI)]n, kobs (s1) is the pseudo-rst-order kinetic close to that of the chlorination of DCF (kapp 3.89 1.17 M1 s1 at pH
constant. Although the Fe(VI) was unstable, it was always in excess 7.0) reported by Soufan et al. (Soufan et al., 2012); however, it is several
within the studied reaction time. orders of magnitude lower than that of O3 (kapp 6.8 105 M1 s1)
The loss of DCF followed pseudo-rst-order kinetics within the (Sein et al., 2008). Yang et al. reported the pH-dependent kapp of the
time scales investigated at a given Fe(VI) concentration, i.e., m = 1, oxidation of triclson using Fe(VI) to be 754.7 M1 s1 at 7.0 (Yang
as illustrated in Fig. 1a. For a constant initial concentration of DCF et al., 2011), which is markedly higher than the rate constant of Fe(VI) re-
(0.030 mM), different values of kobs were determined from the slopes ported in this study. Other pharmaceuticals containing amine moieties
of linear plots at various concentrations of Fe(VI) (Fig. 1a). The value (e.g., tramadol, enrooxacin, and ciprooxacin) have kapp values of
of kobs increased linearly (R2 = 0.9991) as Fe(VI) concentrations were 7.4, 24, and 1.7 102 M 1 s1, respectively, at pH 8.0 (Anquandah
increased from 0.15 to 0.90 mM (Fig. 1b), demonstrating a rst-order et al., 2011; Lee et al., 2009; Zimmermann et al., 2012). Comparatively,
dependence of the reaction on Fe(VI), i.e., n = 1. Thus, the reaction sulfonamides had generally higher rates than those of amines that
between DCF and ferrate was of second-order in total, and rst-order contained pharmaceuticals (Sharma et al., 2006).
Y. Wang et al. / Science of the Total Environment 506507 (2015) 252258 255

a 2 represent the molar fraction of non-ionizable DCF and ionized


0.0
DCF , respectively; 1 , 2 and 3 represent the molar fraction of
H2FeO4, HFeO 2
4 and FeO4 , respectively.
-0.5 Chemically, due to the pKa value of DCF and Fe(VI), DCF is the
major form in solution, whereas two major species of Fe(VI)
ln[DCF]/[DCF]0

(i.e., HFeO 2
4 and FeO4 ) may be observed under experimental pH con-
-1.0
ditions (7.0 pH 11.0). Considering these conditions, although there
were six possible reactions from Eq. (9), only two elementary reactions
-1.5 (10)(11) might be contributing to the observed kinetics constants as
pH=11.0
pH=10.0 relates to the oxidation of DCF by Fe(VI) (Anquandah et al., 2013;
pH=9.0 Yang et al., 2011):
-2.0
pH=8.0
pH=7.0
HFeO4 DCF Products FeOH3 k10 10
-2.5
0 1 2 3 4 5 6 7 8 9 10 11 2
FeO4 DCF Products FeOH3 k11 11
Time (min)
As a result, Eq. (9) can be rewritten as:
12 b 1.0    
ka k10 ka1 H k11 ka1 ka2
kapp k10 2 2 k11 2 3   12
10 Meas. kapp
H  ka H 2 ka1 H  ka1 ka2
0.8
Model kapp
8 Fig. 2b represents the experimental and modeled pH proles
kapp (M-1 s-1)

Mole fraction
0.6 that were obtained from these constants and Eq. (12). The species
6 specic second-order kinetics constants, k10 (16.8( 0.55) M1 s1)
0.4
and k11 (3.61( 0.25) M 1 s 1) were calculated from least-squares
4 nonlinear regressions of the experimental kapp data by using
SigmaPlot 10.0 software (Systat Software Inc.). A fairly good correla-
0.2 tion (R2 = 0.994) between measured (symbols, Fig. 2b) and modeled
2
kapp (thick solid line, Fig. 2b) was obtained, whatever the pH level
0 0.0 might be. The tting results indicated that HFeO 4 was much more
reactive than FeO24 , in good agreement with a number of previous
7.0 7.5 8.0 8.5 9.0 9.5 10.0 10.5 11.0 reports that described Fe(VI) reactivity (Anquandah et al., 2013; Yang
pH et al., 2011).
As can be seen from the speciation of Fe(VI) (thin solid line in
Fig. 2. Effect of pH on DCF oxidation by ferrate: (a) plot of ln[DCF]/[DCF]0 vs. reaction time Fig. 2b), the molar fraction of HFeO 4 decreased, whereas that of
under different pH conditions; and (b) plot of apparent second-order kinetic constant (kapp, FeO24 increased accordingly, as the solution pH was increased. The
M1 s1) vs. pH. Reaction conditions: [DCF]0 = 0.030 mM, [Fe(VI)]0 = 0.45 mM, T =
contribution of the individual reactions to the kinetics constant is also
23 2 C. Symbols represent measured data and thick line represents the model calculations.
shown (dashed lines, Fig. 2b). At pH 7.09.0, the reaction (10) between
HFeO
4 and DCF (k10) was the key contributor to the rate constants,
3.2. Effect of pH on DCF oxidation while the reaction (11) between FeO24 with DCF (k11) was the
major reaction at pH 9.011.0. In general, the kapp value decreased as
Similar experiments were also conducted to determine the values of the solution pH was increased. Moreover, the pH dependence of kapp
kapp at different pH levels. Fig. 2 indicates that the kapp of DCF oxidation could also be explained by the redox potential of FeO2
4 /Fe(OH)3 couple
by ferrate was heavily dependent on pH. Kapp increased signicantly (Jiang and Zhou, 2013). Based on the following electrode reaction (12)
from 2.54 to 11.6 M1 s 1, as the pH was decreased from 11.0 to and Nernst Eq. (13) in the alkaline medium:
7.0 at 23 2C, which should be associated with acid catalysis under
an increased [H+] condition. A similar trend of pH dependence was FeO4
2
l 4H2 O 3e FeOH3 s 5OH l E FeO 2 = FeOH 0:72V 12
0
4 3
also observed with the oxidation of carbamazepine, nonylphenol, by
Fe(VI) (Hu et al., 2009; Sharma et al., 2009). The pH dependence of
0:0592 c5 OH
the rate constants might be attributed to the speciations of both DCF E FeO 2
= FeOH3 0:72 lg 13
and Fe(VI). The acidbase equilibrium of DCF and Fe(VI) may be
4 3 c FeO 2
4

expressed in Eqs. (6), (7) and (8):


Eq. (12) calculations show that the electrode potential of

DC FDC F H pKa 4:2 (Huber et al., 2003) 6 FeO24 /Fe(OH)3 couple increased with pH decreased from 11.0 to
7.0 at 23 2 C, which indicated that the oxidation ability of Fe(VI)

H2 FeO4 H HFeO4 pKa1 3:5 (Sharma et al., 2001) 7 had been increased as well. Thus the increase of k app was likely
attributed to the promotion of H+ at a lower pH, which facilitated
2
H FeO4 H FeO4 pKa2 7:23 8 an easier reaction between DCF and Fe(VI). Given that Fe(VI) suf-
fered variable autodecomposition at below pH 8.0, the optimal initial
The second-order kinetics could therefore be modeled by the pH was determined to be 9.0.
following Eq.:
3.3. Effect of temperature on DCF oxidation
dDC F X
k FeVIDCF
i1;2; j1;2;3 i j i j
9
dt
The oxidation of DCF by ferrate was dependent on temperature since
where kij is species-specic second-order kinetics constant for the the kapp of DCF destruction by ferrate was increased gradually from 2.64
reaction between the DCF species i and the Fe(VI) species j; 1 and to 7.20 M 1 s 1 with elevated temperatures from 5 C to 35 C, as
256 Y. Wang et al. / Science of the Total Environment 506507 (2015) 252258

depicted in Fig. 3. Considering the needs of relatively high reaction rates From the previous literature, the activation parameters for the
and low energy cost, the optimal temperature was close to ambient reaction of Mn(VII) with carbamazepine (an anticonvulsant and
at 25 C. mood-stabilizing drug), similar to DCF oxidation by Fe(VI) in this
According to the Arrhenius Eq.: study, were reported to be 22( 2) kJ mol 1, 19( 1) kJ mol 1 and
140( 5) J mol 1 K 1 (Hu et al., 2009). The small difference in
Ea Ea and H values indicated a subtle energy difference between the tran-
lnkapp lnA 14
RT sition state and the reactants, resulting in the relative lower effect of
temperature on the reaction between Fe(VI) and DCF (Sharma et al.,
where Ea is the apparent activation energy, R is the universal gas 2006).
constant, and A is the Arrhenius pre-exponential constant. Plotting In contrast to previously reported literature, the properties of
ln(kapp) vs. 1/T would yield a linear relationship between measured contaminants, as well as the initial concentrations of reactants and pH
kinetic constants and the reciprocal of absolute temperature (T), as is in different solutions, exhibited a signicant inuence on the rate con-
shown in Fig. 3b. Ea was determined to be 23.3( 0.5) kJ mol 1 by stants. The ndings presented above suggest that the control of reaction
tting Eq. (14) to the experimental data. The t-derived value of conditions should be considered in the determination of degradation
Ea fell at the lower end of the range expected for reactions under rates and efcacy toward the practical application of Fe(VI) for water
chemical control, and indicated a rather small temperature depen- treatment.
dence (Hu et al., 2009). The enthalpic (H) and entropic (S) con-
tributions to the activation energy were also determined to be 3.4. Toxicity assessment
20.9( 0.5) kJ mol 1 and 161.0( 1.6) J mol 1 K 1, respectively,
by tting the temperature-dependent kinetic data with the Eyring Toxicity is a critical factor in the assessment of drinking water safety.
Eq. (Hu et al., 2009): Since DCF byproducts may be more toxic than the original DCF com-
pound, it is prudent to assess the toxicity of DCF contaminated water
kapp H 1 k S that is treated with ferrate. The toxicity of water samples was evaluated
ln  ln B 15
T R T h R by monitoring changes in the natural luminescence of the bacteria
V. scheri when challenged with toxic compounds. Because Fe(VI) resi-
where kB is Boltzmann's constant and h is Planck's constant. dues in the samples had already been quenched with Na2S2O3 prior to
the Microtox test, the presence of Fe(VI) itself in the reaction did not in-
hibit the growth of the luminescent bacteria. Sixteen samples subjected
a to oxidation by Fe(VI) (0.30 mM) at different reaction times were
0.0
analyzed to estimate their TOC removal rates (left bars) and inhibition
-0.3 percentages (right bars), as shown in Fig. 4. The initial toxicity of DCF
solution (0.030 mM DCF, 0 min of oxidation-distilled water) exhibited
-0.6 an inhibition rate of 19.1% (15 min of incubation) in Fig. 4a, according
ln[DCF]/[DCF]0

to Eq. (1). There was a slight decrease observed, between 8.1% and
-0.9 o
5 C 16.9% in the inhibition rate, whereas the TOC conversion was signi-
o
10 C cantly increased, from 3.6% to 50.9% when the reaction time was
-1.2 o
15 C increased from 2 min (Fig. 4b) to 20 min (Fig. 4g), which should primar-
-1.5
o
20 C ily be associated with the reduction of the DCF, coupled with a minor
o
25 C toxicity fraction derived from the (P1, P2, P3) byproducts that are
-1.8 o
30 C formed. Subsequently, the overall toxicity of the solution increased rap-
o
35 C idly from 30.2% to 71.0%, reaching a maximum inhibition rate at 80 min
-2.1 (Fig. 4l). By this time, the parent DCF compound had already completely
0 1 2 3 4 5 6 7 8 9 10 11 disappeared. Regarding the TOC removal within the 3080 min of
Time (min) reaction time (Fig. 4g l), there was a slower increase, from 61.7% to
72.2%, which clearly demonstrated that compounds having a higher
2.2 toxicity than DCF were generated (Mendez-Arriaga et al., 2008). From
80120 min of oxidation (Fig. 4l n), the inhibition rate gradually
b decreased; however, it remained at higher than 28.7%. Additionally,
2.0
during this period, P3 and other unidentied intermediates were still
1.8 present in the solution, while synergistic effects among them were
also considered in terms of the corresponding mineralization rate
ln(kapp M-1 s-1)

1.6 (Scheurell et al., 2009). Following 120 min (Fig. 4n q), the toxicity
of the solution was reduced steadily, from 28.7% to 12.9%, since some
1.4 higher mineralization was observed up to 84.6%, illustrating the
detoxication efcacy of the disinfection process. As treated under
1.2 y = 11.07477 - 2.80268x the added volume of Fe(VI), the formed byproducts were further oxi-
2 dized to other nontoxic compounds (Martinez et al., 2011; Pera-Titus,
R = 0.9975
1.0 2004), and thus, resulted in a continued slowdown in toxicity. Although
the changes in toxicity were primarily attributed to the concentrations
0.8 of the intermediates, the highest toxicity, observed in Fig. 4l, was
3.20 3.25 3.30 3.35 3.40 3.45 3.50 3.55 3.60 3.65 associated with the maximum load of P3 that was generated in lower
concentrations (Mendez-Arriaga et al., 2008). These results suggested
1000/T (K-1)
that higher concentrations of P1 and P2 did not yield an appreciable
Fig. 3. Effect of temperature on DCF oxidation by ferrate: (a) plot of ln[DCF]/[DCF]0 vs.
effect on the overall toxicity, relative to P3, etc. Therefore, synergistic
reaction time under different temperature; and (b) Arrhenius plot of lnkapp vs. 1/T. effects were considered to be the major factor in determining
Reaction conditions: [DCF]0 = 0.030 mM, [Fe(VI)]0 = 0.45 mM, pH = 9.0 0.2. the toxicity of the treated samples. This signicant increase in the
Y. Wang et al. / Science of the Total Environment 506507 (2015) 252258 257

Fig. 4. Variation of removal rate of TOC (left bars), inhibition rate of V. scher bioluminescence (right bars) and HPLC spectra (in integrogram) during DCF oxidation by Fe(VI) at different
reaction times. Reaction conditions: [DCF]0 = 0.030 mM, [Fe(VI)]0 = 0.30 mM, pH = 9.0 0.2, T = 23 2 C.

toxicity of intermediates raises a serious concern when DCF contam- 2.54 to 11.6 M1 s1, as pH was decreased from 11.0 to 7.0. This pH
inated source water is treated with ferrate for disinfection in water dependence was due to the changing acidbase speciation of Fe(VI).
treatment plants. Moreover, the slight variation of kapp in the studied temperature range
of 535 C exhibited relatively lower dependent on temperature and
4. Conclusions E a , H and S were determined to be 23.3( 0.5) kJ mol 1 ,
20.9( 0.5) kJ mol 1 and 161.0( 1.6) J mol1 K 1, respectively.
This study demonstrated that ferrate could be applied to effectively Toxicity tests suggest that more toxic intermediates were formed. Fur-
remove DCF in the disinfection of potable water. The reaction between ther efforts will be required to methodically investigate individual
DCF and Fe(VI) followed rst-order kinetics with respect to each byproducts in order to conrm their toxicity, thus providing useful
reactant. The value of kapp was observed to increase signicantly from information regarding the safety of drinking water.
258 Y. Wang et al. / Science of the Total Environment 506507 (2015) 252258

Acknowledgments Marco-Urrea E, Perez-Trujillo M, Cruz-Morato C, Caminal G, Vicent T. Degradation of the


drug sodium diclofenac by Trametes versicolor pellets and identication of some
intermediates by NMR. J Hazard Mater 2010;176:83642.
This work is supported by the National Natural Science Foundation Martinez C, Canle LM, Fernandez MI, Santaballa JA, Faria J. Aqueous degradation of
of China (Nos. 21377031, 21071047), the Scientic Research Project diclofenac by heterogeneous photocatalysis using nanostructured materials. Appl
Catal Environ 2011;107:1108.
of Guangdong Province (No. 2013B020800009), the Natural Science Mendez-Arriaga F, Esplugas S, Gimenez J. Photocatalytic degradation of non-steroidal
Foundation of Henan Province (Nos. 132300410158, 112300410104), anti-inammatory drugs with TiO2 and simulated solar irradiation. Water Res
and the Science and Technology Research Key Project of Education 2008;42:58594.
Michael I, Achilleos A, Lambropoulou D, Osorio Torrens V, Perez S, Petrovic M, et al.
Department of Henan Province (Nos. 13A610516, 13A610853). Proposed transformation pathway and evolution prole of diclofenac and ibuprofen
transformation products during (sono) photocatalysis. Appl Catal Environ 2014;
References 147:101527.
Miranda-Garca N, Surez S, Snchez B, Coronado JM, Malato S, Maldonado MI. Photocat-
Aguinaco A, Beltran FJ, Garcia-Araya JF, Oropesa A. Photocatalytic ozonation to remove alytic degradation of emerging contaminants in municipal wastewater treatment
the pharmaceutical diclofenac from water: inuence of variables. Chem Eng J 2012; plant efuents using immobilized TiO2 in a solar pilot plant. Appl Catal Environ
189:27582. 2011;103:294301.
Anquandah GAK, Sharma VK, Knight DA, Batchu SR, Gardinali PR. Oxidation of Nie E, Yang M, Wang D, Yang X, Luo X, Zheng Z. Degradation of diclofenac by ultrasonic
trimethoprim by ferrate(VI): kinetics, products, and antibacterial activity. Environ irradiation: kinetic studies and degradation pathways. Chemosphere 2014;113:
Sci Technol 2011;45:1057581. 16570.
Anquandah GAK, Sharma VK, Panditi VR, Gardinali PR, Kim H, Oturan MA. Ferrate(VI) Pera-Titus M. Garca-Molina V, Baos MA, Gimnez J, Esplugas S. Degradation of
oxidation of propranolol: kinetics and products. Chemosphere 2013;91:1059. chlorophenols by means of advanced oxidation processes: a general review. Appl
Bae S, Kim D, Lee W. Degradation of diclofenac by pyrite catalyzed Fenton oxidation. Appl Catal Environ 2004;47:21956.
Catal Environ 2013;134135:93102. Scheurell M, Franke S, Shah RM, Huehnerfuss H. Occurrence of diclofenac and its
Calza P, Sakkas VA, Medana C, Baiocchi C, Dimou A, Pelizzetti E, et al. Photocatalytic metabolites in surface water and efuent samples from Karachi, Pakistan. Chemosphere
degradation study of diclofenac over aqueous TiO2 suspensions. Appl Catal Environ 2009;77:8706.
2006;67:197205. Sein MM, Zedda M, Tuerk J, Schmidt TC, Golloch A, Von Sonntao C. Oxidation of diclofenac
Forrez I, Carballa M, Verbeken K, Vanhaecke L, Schlusener M, Ternes T, et al. Diclofenac with ozone in aqueous solution. Environ Sci Technol 2008;42:665662.
oxidation by biogenic manganese oxides. Environ Sci Technol 2010;44:344954. Sharma VK. Ferrate(VI) and ferrate(V) oxidation of organic compounds: kinetics and
Garoma T, Matsumoto S. Ozonation of aqueous solution containing bisphenol A: effect of mechanism. Coord Chem Rev 2013;257:495510.
operational parameters. J Hazard Mater 2009;167:118591. Sharma VK, Anquandah GAK, Nesnas N. Kinetics of the oxidation of endocrine disruptor
Hartmann J, Bartels P, Mau U, Witter M, Von Tumpling W, Hofmann J, et al. Degradation of nonylphenol by ferrate(VI). Environ Chem Lett 2009;7:1159.
the drug diclofenac in water by sonolysis in presence of catalysts. Chemosphere Sharma VK, Burnett CR, Millero FJ. Dissociation constants of the monoprotic ferrate(VI)
2008;70:45361. ion in NaCl media. Phys Chem Chem Phys 2001;3:205962.
Heberer T. Tracking persistent pharmaceutical residues from municipal sewage to Sharma VK, Mishra SK, Ray AK. Kinetic assessment of the potassium ferrate(VI) oxidation
drinking water. J Hydrol 2002;266:17589. of antibacterial drug sulfamethoxazole. Chemosphere 2006;62:12834.
Hu L, Martin HM, Arcs-Bulted O, Sugihara MN, Keatlng KA, Strathmann TJ. Oxidation of Shu Z, Bolton JR, Belosevic M. Gamal El Din M. Photodegradation of emerging
carbamazepine by Mn(VII) and Fe(VI): reaction kinetics and mechanism. Environ micropollutants using the medium-pressure UV/H 2 O 2 Advanced Oxidation
Sci Technol 2009;43:50915. Process. Water Res 2013;47:28819.
Huber MM, Canonica S, Park GY, Von Gunten U. Oxidation of pharmaceuticals during Soufan M, Deborde M, Legube B. Aqueous chlorination of diclofenac: kinetic study and
ozonation and advanced oxidation processes. Environ Sci Technol 2003;37:101624. transformation products identication. Water Res 2012;46:337786.
Huebner U, Jekel M. Tertiary treatment of Berlin WWTP efuents with ferrate (Fe(VI)). Surez S, Carballa M, Omil F, Lema J. How are pharmaceutical and personal care
Water Sci Technol 2013;68:166571. products (PPCPs) removed from urban wastewaters? Rev Environ Sci Biotechnol
Jiang JQ. The role of ferrate(VI) in the remediation of emerging micro pollutants. Procedia 2008;7:12538.
Environ Sci 2013;18:41826. Svanfelt J, Kronberg L. Synthesis of substituted diphenylamines and carbazoles:
Jiang JQ. Advances in the development and application of ferrate(VI) for water and phototransformation products of diclofenac. Environ Chem Lett 2011;9:1414.
wastewater treatment. J Chem Technol Biotechnol 2014;89:16577. Ternes TA, Herrmann N, Bonerz M, Knacker T, Siegrist H, Joss A. A rapid method to
Jiang JQ, Zhou Z, Pahl O. Preliminary study of ciprooxacin (cip) removal by potassium measure the solidwater distribution coefcient (Kd) for pharmaceuticals and musk
ferrate(VI). Sep Purif Technol 2012;88:958. fragrances in sewage sludge. Water Res 2004;38:407584.
Jiang J, Zhou Z. Removal of pharmaceutical residues by ferrate(VI). PLos One 2013;8. Vieno NM, Harkki H, Tuhkanen T, Kronberg L. Occurrence of pharmaceuticals in river
Joss A, Zabczynski S, Gobel A, Hoffmann B, Lofer D, McArdell CS, et al. Biological water and their elimination a pilot-scale drinking water treatment plant. Environ
degradation of pharmaceuticals in municipal wastewater treatment: proposing a Sci Technol 2007;41:507784.
classication scheme. Water Res 2006;40:168696. Vogna D, Marotta R, Napolitano A, Andreozzi R. d'Ischia M. Advanced oxidation of
Lee Y, Yoon J, Von Gunten U. Kinetics of the oxidation of phenols and phenolic endocrine the pharmaceutical drug diclofenac with UV/H2O2 and ozone. Water Res 2004;
disruptors during water treatment with ferrate (Fe(VI)). Environ Sci Technol 2005; 38:41422.
39:897884. Yang B, Ying GG, Zhao JL, Zhang LJ, Fang YX, Nghiem LD. Oxidation of triclosan by ferrate:
Lee Y, Zimmermann SG, Kieu AT, Von Gunten U. Ferrate (Fe(VI)) application for municipal reaction kinetics, products identication and toxicity evaluation. J Hazard Mater
wastewater treatment: a novel process for simultaneous micropollutant oxidation 2011;186:22735.
and phosphate removal. Environ Sci Technol 2009;43:38318. Zhang Y, Geissen SU, Gal C. Carbamazepine and diclofenac: removal in wastewater
Licht S, Yu XW. Electrochemical alkaline Fe(VI) water purication and remediation. treatment plants and occurrence in water bodies. Chemosphere 2008;73:
Environ Sci Technol 2005;39:80716. 115161.
Luo Z, Strouse M, Jiang JQ, Sharma VK. Methodologies for the analytical determination Zimmermann SG, Schmukat A, Schulz M, Benner J, Von Gunten U, Ternes TA. Kinetic and
of ferrate(VI): a review. J Environ Sci Health A Tox Hazard Subst Environ Eng mechanistic investigations of the oxidation of tramadol by ferrate and ozone. Environ
2011;46:45360. Sci Technol 2012;46:87684.

You might also like