ArticlePDF Available

Decolourization of direct blue 15 by Fenton/ultrasonic process using a zero-valent iron aggregate catalyst

Authors:

Abstract and Figures

Decolourization of direct azo dye, direct blue 15 (DB15), by an advanced Fenton process coupled with ultrasonic irradiation (Fenton/US) was investigated. Zero-valent iron (ZVI) aggregates were used as the catalyst. A positive synergistic effect occurred when Fenton's reagent was combined with ultrasonic irradiation. Experimental results showed that the optimum conditions for decolourization were pH 3.0, Fe(0) 1g/L, H(2)O(2) 5.15×10(-3)mol/L with ultrasound density of 120W/L at 60kHz. These conditions yielded 99% decolouration of 4.7×10(-5) M DB15 (4130 ADMI) solution within 10min. DB15 decolouration follows the first-order decolouration kinetics. Although the solutions containing H(2)CO(3), Cl(-), ClO(4)(-), NO(3)(-) and SO(4)(2-) ions did not have a significant effect on the decolouration, the H(2)PO(4)(-) ion did decrease the decolouration rate. High ultrasonic input power accelerated the reaction and increased decolourization efficiency. The cost effectiveness of this process at high ultrasound density could be controlled despite the high electricity costs incurred by the process. ZVI aggregates were reusable; however, an increase in the number of times ZVI was recycled decreased the decolourization rate. This study demonstrates that a Fenton/US process can effectively decolour the direct azo dye DB15 in wastewater.
Content may be subject to copyright.
Decolourization of direct blue 15 by Fenton/ultrasonic process using
a zero-valent iron aggregate catalyst
Chih-Huang Weng
a,
, Yao-Tung Lin
b
, Cheng-Kuan Chang
a
, Na Liu
c
a
Department of Civil and Ecological Eng., I-Shou University, Kaohsiung 84001, Taiwan
b
Department of Soil and Environmental Sci., National Chung Hsing University, Taichung 402, Taiwan
c
Key Laboratory of Groundwater Resources and Environment, Ministry of Education, Jilin University, Changchun 130021, China
article info
Article history:
Received 24 July 2012
Received in revised form 28 September 2012
Accepted 28 September 2012
Available online 7 November 2012
Keywords:
Direct azo dye
Fenton
Ultrasound
Zero-valent iron aggregate
abstract
Decolourization of direct azo dye, direct blue 15 (DB15), by an advanced Fenton process coupled with
ultrasonic irradiation (Fenton/US) was investigated. Zero-valent iron (ZVI) aggregates were used as the
catalyst. A positive synergistic effect occurred when Fenton’s reagent was combined with ultrasonic irra-
diation. Experimental results showed that the optimum conditions for decolourization were pH 3.0, Fe(0)
1 g/L, H
2
O
2
5.15 10
3
mol/L with ultrasound density of 120 W/L at 60 kHz. These conditions yielded
99% decolouration of 4.7 10
5
M DB15 (4130 ADMI) solution within 10 min. DB15 decolouration fol-
lows the first-order decolouration kinetics. Although the solutions containing H
2
CO
3
,Cl
, ClO
4
,NO
3
and SO
2
4
ions did not have a significant effect on the decolouration, the H
2
PO
4
ion did decrease the decol-
ouration rate. High ultrasonic input power accelerated the reaction and increased decolourization effi-
ciency. The cost effectiveness of this process at high ultrasound density could be controlled despite the
high electricity costs incurred by the process. ZVI aggregates were reusable; however, an increase in
the number of times ZVI was recycled decreased the decolourization rate. This study demonstrates that
a Fenton/US process can effectively decolour the direct azo dye DB15 in wastewater.
Ó2012 Elsevier B.V. All rights reserved.
1. Introduction
Effluents discharged from textile and paper printing industries
normally contain many structural varieties of highly coloured
dye. Azo dyes, which represent approximately 70% of the world’s
dye production, are the major colourants used by these industries.
Azo dyes derived from benzidine have proven to be carcinogenic to
humans and have adverse effects on aquatic life. Therefore, they
need to be properly treated before being discharged. Most azo dyes
are refractory to biodegradation. Moreover, auxiliary textile chem-
icals such as polyvinyl alcohol used in the sizing process cannot be
degraded in an activated sludge process [1]. Thus, conventional
biological treatment may no longer be considered a satisfactory
decolourization process. Physico-chemical decolourization method
such as membrane separation, electrocoagulation and adsorption
are currently available; however, chemical sludge generation,
adsorbent regeneration and maintenance of fouled membranes
may raise serious concerns. A variety of advanced oxidation pro-
cesses (AOPs) have been developed to decolourize textile and
printing industry wastewater [2,3].
Fenton’s reaction and sonolysis are two of the most promising
AOPs for treating organic substances. In general, the Fenton re-
agent can oxidize a variety of organic contaminants. To improve
the effectiveness of Fenton’s reaction, the addition of zero-valent
iron (ZVI) has been explored for the degradation of hazardous or-
ganic compounds [4–6]. Recently, interest in Fenton’s reaction in-
creased when it was determined that combining Fenton’s process
with hydrodynamic cavitation, electrolysis, photocatalysis or son-
olysis resulted in greater versatility for removal of pollutants
[3,7–12]. The Fenton process coupled with ultrasonic irradiation
process (Fenton/US) to destroy recalcitrant organic substances is
of interest because it enables further degradation and is consider-
ably more efficient than Fenton’s reaction alone [13–17]. The
mechanisms of the Fenton/US process are described as follows
[18]:
H
2
O
2
!
ÞÞÞ
2OH
ð1Þ
OH
þOH
!H
2
O
2
ð2Þ
Fe
0
!
ÞÞÞ
Fe
2þ
þ2e
ð3Þ
Fe
2þ
þH
2
O
2
!Fe
3þ
þOH
þOH
ð4Þ
1350-4177/$ - see front matter Ó2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.ultsonch.2012.09.014
Corresponding author. Tel.: +886 929552662; fax: +886 76577461.
E-mail addresses: chweng@isu.edu.tw (C.-H. Weng), yaotung@nchu.edu.tw
(Y.-T. Lin), liuna@jlu.edu.cn (N. Liu).
Ultrasonics Sonochemistry 20 (2013) 970–977
Contents lists available at SciVerse ScienceDirect
Ultrasonics Sonochemistry
journal homepage: www.elsevier.com/locate/ultson
Fe
3þ
þH
2
O
2
!Fe-OOH
2þ
þH
þ
ð5Þ
Fe-OOH
2þ
!Fe
2þ
þOOH
ð6Þ
Fe
3þ
þOOH
!Fe
2þ
þO
2
þH
þ
ð7Þ
Fe
0
þ2Fe
3þ
!3Fe
2þ
ð8Þ
where ‘‘)))’’ represents ultrasound wave.
Ultrasonic irradiation of organics in water results in the gener-
ation of hydroxyl radicals that can react with and destroy a wide
range of organic contaminants. This effect is greatly enhanced by
the addition of hydrogen peroxide in the presence of iron particles
[13]. Although the partial recombination of OH
radicals could self-
produce H
2
O
2
from the sonolysis of water (Eq. (2)), the main source
of H
2
O
2
is external. The use of ZVI as a catalyst in Fenton’s reaction
could be a cost-effective method to treat organic waste [19]. In the
Fenton/US system, ultrasonic irradiation can enhance the corrosion
of ZVI, releasing massive amounts of Fe(II) continuously. Sufficient
amounts of Fe(II) and H
2
O
2
are required to trigger a series of Fen-
ton’s reactions (Eqs. (3)–(8)). Ultrasonic irradiation dramatically
enhances Fenton’s reaction.
In this study, the decolourization of direct blue 15 (DB15) by
Fenton/US was evaluated. Although studies have shown that
DB15 could be decolourized via photocatalysis of Fe(III)–oxalate
complexes/H
2
O
2
[20] and Fenton’s reagent [2], the use of ZVI
aggregates as the catalyst for decolourization of DB15 with the
Fenton/US oxidation is considered to be important and needs to
be assessed further. To evaluate the effectiveness of treating
DB15 with Fenton/US process, the following four objectives were
determined: (1) compare the effectiveness of US/H
2
O
2
, Fenton
and Fenton/US processes; (2) characterise the key parameters
(pH, ZVI dose, H
2
O
2
dose, salts and ultrasonic power) affecting
the Fenton/US process and identify suitable decolourization condi-
tions, (3) evaluate the cost-effectiveness of the Fenton/US process
and (4) test the availability of reusable ZVI in the Fenton/US
process.
2. Materials and methods
2.1. Reagents
DB15 (C
34
H
24
N
6
Na
4
O
16
S
4
, CAS No. 2429–74–5) (Fig. 1) was pur-
chased from Sigma-Aldrich. DB15 concentration was analysed by
measuring the adsorption at 595 nm using a spectrophotometer
(DR2800™, Hach, USA). ZVI aggregates (Fig. 2) with particle size
0.297–2.380 mm and specific weight 2240–2560 kg/m
3
were ob-
tained from Connelly-GPM Inc., USA. H
2
O
2
(analytical grade, 35%
w/w) was purchased from J.T. Baker (USA). Fe(II) and total Fe were
analysed by measuring absorption at 510 nm using Hach FerroVer
Ò
iron reagent [21]. Fe(III) was determined from the difference be-
tween total Fe and Fe(II).
2.2. Fenton/US experiments
Sonication of the dye solution was performed in air atmosphere
with a fixed frequency of 60 kHz generated by an ultrasonic
generator (S-450A, Brason, USA) equipped with a titanium probe
transducer. The tip of the probe was 1.2 cm in diameter; 6 cm of
the tip’s total length was submersed in the dye solution. The
sonication was administered in pulses with a 60% duty cycle. The
ultrasonic generator provides direct sonication, which will not
cause energy loss because the reaction matrix is in direct contact
with the mechanical vibration. Fenton/US experiments were con-
ducted for a 1000 mL DB15 dye placed in a glass beaker. The pH
was adjusted to a predetermined value using a solution of either
HNO
3
or NaOH. The experimental solution was placed in a beaker
and irradiated with an ultrasonic horn (Fig. 3) and appropriate
amounts of ZVI and H
2
O
2
were added. At desired time intervals,
a 5 mL sample was removed from the beaker and immediately fil-
tered using a 0.45
l
m fibre glass filter to collect the supernatant,
which was then analysed to determine the residual concentration.
3. Results and discussion
3.1. Enhancement of sonication in the Fenton process
Fig. 4a shows DB15 depletion as a result of treatment with dif-
ferent processes. Due to a positive synergistic effect, the coupled
Fenton/US process was much more effective than Fenton’s reaction
or sonolysis alone. As shown in the figure, after 20 min treatment,
the decolourization efficiency with either sonolysis or Fenton’s
process was only 15%. However, in the same period, nearly 100%
colour removal was achieved with Fenton/US. The synergistic
mechanism of the Fenton/US process is generally attributed to
the increase in OH
radical concentration, resulting from the influ-
ence of ultrasound [17].
Fig. 1. Chemical structure of DB15.
Fig. 2. ZVI aggregates.
Fig. 3. Fenton/ultrasonic apparatus.
C.-H. Weng et al. / Ultrasonics Sonochemistry 20 (2013) 970–977 971
Fig. 4b shows the effect of the presence of Fe
2+
on decolouriza-
tion of DB15. In Fenton’s process, although Fe
2+
was generated con-
tinuously over a 30 min period, the concentration of Fe
2+
remained
low compared with that in the other two systems, i.e. Fenton/US
and Fe(0)/US. Particularly, in case of the Fenton/US system, the
release of Fe
2+
was significantly enhanced over the entire duration
of the experiment. For Fenton/US, the concentration of Fe
2+
was
even higher than that for Fe(0)/US within 10 min, which could be
attributed to H
2
O
2
as a result of Fe(0) corrosion. In Fenton(0)/US,
the concentration of released Fe(II) decreased after 20 min as a
consequence of Fe
2+
reacting with H
2
O
2
(Eq. (4)). In general, the
quantity of Fe
2+
released from ZVI was largely enhanced by ultra-
sound and influenced by the addition of H
2
O
2
.
3.2. UV–visible spectra changes during the reaction
The graph inset in Fig. 5a shows that the concentration of DB15
decreased dramatically as the Fenton/US reaction time increased.
Complete decolourization of DB15 was achieved within 15 min.
To clarify the changes in molecular and structural characteristics
of DB15 that resulted from oxidation via the Fenton/US process,
representative variations of UV–visible spectra were observed,
and the corresponding spectra are displayed in Fig. 5a. DB15 was
characterised by one main band in the visible region with the
absorbance peak at approximately 600 nm and by another band
in the ultraviolet region with the absorbance peak at 319 nm.
The peak at 600 nm was attributed to the chromophore-containing
azo linkage of the DB15 molecule. The peaks at 319 nm were asso-
ciated with benzene ring structures in the DB15 molecule [15]. The
characteristic peak decreased remarkably within the initial 5 min,
and after 15 min the peaks reduced to the minimum. This implies
that the conjugate azo structure of DB15 was destroyed and the
colour disappeared quickly. In addition to the rapid decolouriza-
tion, the quick decay of the absorbance at 319 nm was the evidence
of degradation of an aromatic fragment in the DB15 molecule and
its intermediates.
The photos in Fig. 5b show that the solution changed from dark
blue to pink after 2 min treatment and then gradually changed to
light brown. The solution became transparent after 15 min. The
American Dye Manufacturers Institute (ADMI) tristimulus filter
method [21] is the standard analytical method for true colour mea-
surements of wastewater samples. In Taiwan, environmental regu-
lations prescribe a discharge limit of 550 ADMI for the textile
industry. As indicated (Fig. 5b), the solution meets the discharge
effluent criteria after 10 min treatment. Overall, the coupled Fen-
ton/US process could decolour DB15 effectively and efficiently.
3.3. Effect of solution pH
The pH of the solution plays a key role in the decolouration effi-
ciency of the Fenton process, because it affects the solubility of
Fe(II)/Fe(III), and ultimately controls the production of OH
radicals.
Fig. 6 shows oxidation of DB15 as a function of pH and time. In gen-
eral, depletion of DB15 increased with decreasing pH. A first-order
equation was used to analyse the decolouration rate:
dC=dt¼kt ð9Þ
C¼C
0
expðktÞð10Þ
where C
0
= initial dye concentration (M); k= rate constant (min
1
);
t= reaction time (min). The obtained kas a function of pH is shown
0 100
1 10-5
2 10-5
3 10-5
4 10-5
5 10-5
0 5 10 15 20 25 30
Fenton/US
Fe(0)/US
H2O2/US
Fenton
US
Direct Blue 15 (mol/L)
Time (min)
C0 4.7x10-5 M
Fe(0) 0.5 g/L
H2O2 5.15x10-3 M
pH 3.0
US 120W/L (60kHz)
(a)
0
1
2
3
4
5
6
7
0 5 10 15 20 25 30
Fenton/US
Fenton
Fe(0)/US
Fe(II) mg/L
Time (min)
Fenton
Fenton/US
Fe(0)/US
(b)
Fig. 4. (a) Depletion of DB15 treated with different processes; (b) concentration of
Fe
2+
during the treatment of different processes.
0
0.5
1
1.5
2
2.5
Abs
0 min
2 min
5 min
10 min
15 min
20 min
25 min
30 min
200 400 600 800 1000
Wave length (nm)
DB 15 4.7x10-5 M
Fe(0) 1.0 g/L
H2O2 1.03x10-2 M
pH 3.0
US 120 W/L
0
1
2
3
4
5
0 5 10 15 20 25 30
Direct Bl ue 15 (x10 -5 mol/L)
Time (min)
(a)
0 min 2 5 10 15 20
(b)
Fig. 5. (a) Changes in the UV–visible spectra of DB15 during Fenton/US treatment.
(b). Colour changes during the treatment. Values of ADMI at 0, 2, 5, 10, 15 and
20 min were 4130, 1640, 495, 144, 48, 34 and 13, respectively.
972 C.-H. Weng et al. / Ultrasonics Sonochemistry 20 (2013) 970–977
in the graph inset in Fig. 6. The high regression coefficient (r
2
> 0.99)
indicates that decolouration of DB15 follows first-order kinetics. As
the pH decreased from 5.0 to 2.0, the oxidation rate increased from
0.095 to 0.264 min
1
. It is evident that pH values <3.0 lead to in-
crease in the extent of DB15 oxidation.
Although the production of the OH
radical could be inhibited by
the formation of the complex species [Fe(II)(H
2
O)
6
]
2+
, which reacts
with H
2
O
2
to form [Fe(II)(OH)(H
2
O)
5
]
+
[22–24], in this study, we
demonstrated that the oxidation was not retarded at pH below
3.0, and in fact, the most effective oxidation occurred in such an
acidic medium. The maintenance of effective oxidation at such
low pH could be explained by the use of ZVI aggregates as the cat-
alyst. ZVI aggregates provide a source of Fe(II) sufficient to ensure a
chain oxidation. As expected, the oxidation decreased substantially
at pH > 4 because of the formation of Fe precipitates (Fe(OH)
2(s)
and Fe(OH)
3(s)
), which suppressed the regeneration of OH
radicals
and lowered catalytic activity during the decomposition of H
2
O
2
.
The decolouration did not improve when pH was lowered from
3.0 to 2.5 or 2.0. Indeed, a sudden drop in the rate of decolouration
was observed at pH > 3.0 (Fig. 6). In terms of cost-effectiveness for
practical applications, the optimum operational condition is pH 3.
The results could be explained by the considerable effect of pH on
sonolysis and the Fenton reaction. Relevant studies of dye decol-
ouration have reported that the optimum pH for Fenton or Fen-
ton/US reaction was 3.0 [16,17,25].
3.4. Effects of ZVI dose
Fig. 7 shows that the depletion of DB15 during the Fenton/US
process was affected by the ZVI dose and the corresponding first-
order rate constants. The decolouration of DB15 appears to be
appreciably influenced by the dosage of ZVI. Within 10 min of
the treatment, when the ZVI dose was increased from 0.25 to
4.0 g/L, the decolouration efficiency increased from 80% to 98%,
respectively. In this study, since ZVI aggregates were used, the cor-
rosion of ZVI surfaces in aqueous solution was the only source of
Fe
2+
. Under the influence of ultrasound, higher doses of ZVI lead
to the release of greater amounts of Fe
2+
(Eq. (3)). In the Fenton
process, Fe
2+
acts as a catalyst, which initiates the decomposition
of H
2
O
2
to produce the reactive OH
radical (Eqs. (4) and (6)). More
OH
radicals were produced from the Fenton’s reaction when the
dosage of ZVI was increased, thereby increasing the decolouration
efficiency. From the results of first-order kinetic analysis, a rela-
tionship of k and the ZVI dose was observed for the ZVI dose be-
tween 0.1 and 4 g/L. Since ZVI aggregates can be reused, 1 g/L of
ZVI was selected as the optimum dosage for the decolouration of
4.7 10
5
M DB15 by the Fenton/US process.
3.5. Effects of H
2
O
2
dose
Fig. 8 shows the influence of H
2
O
2
concentrations on the decol-
ouration of DB15 and the corresponding first-order rate constants.
The findings indicate that the decolouration of DB15 was affected
by the H
2
O
2
dosage. The decolouration efficiency increased from
86% to 98% at 10 min as a consequence of increasing the H
2
O
2
dos-
age from 2.58 10
3
to 5.15 10
3
M, respectively (Fig. 8).
Increasing the H
2
O
2
dosage from 5.15 10
3
to 1.03 10
2
to
2.06 10
2
M had an adverse effect on decolouration.
In the Fenton/US system, the main sources of OH
radicals are
majorly generated via two reactions: (1) H
2
O
2
under the influence
of ultrasound (Eq. (1)) and (2) H
2
O
2
reacting with ferrous irons (Eq.
0 10 0
1 10-5
2 10-5
3 10-5
4 10-5
5 10-5
0 5 10 15 20 25 30 35
pH 2.0
pH 2.5
pH 3.0
pH 3.5
pH 4.0
pH 5.0
Direct Bl ue 15 (mol/L)
Time (min)
0
0.1
0.2
0.3
2345
k (1/min)
pH
Fig. 6. Effect of pH on the decolouration of DB15. Solid lines are the best fit of first-
order kinetic equation. Conditions: ZVI 0.5 g/L, H
2
O
2
5.15 10
3
M, ultrasound
density 120 W/L.
0 100
1 10-5
2 10-5
3 10-5
4 10-5
5 10-5
0 5 10 15 20 25 30 35
0.1 g/L
0.25 g/L
0.5 g/L
1 g/L
2 g/L
4 g/L
Direct Blue 15 (mol/L)
Time (min)
ZVI dose
0
0.2
0.4
0.6
0.8
012345
k (1/min)
Fe(0) dose (g/L)
Fig. 7. Effect of ZVI addition on the decolouration of DB15. Solid lines are the best fit
of first-order kinetic equation. Conditions: pH 3.0, H
2
O
2
5.15 10
3
M, ultrasound
density 120 W/L.
0 100
1 10 -5
2 10 -5
3 10 -5
4 10 -5
5 10 -5
0 5 10 15 20 25 30 35
H2O2 2.58x10-3 M
H2O2 5.15x10-3 M
H2O2 1.03x10-2 M
H2O2 2.06x10-2 M
Direct Blue 15 (mol/L)
Time (min)
0
0.2
0.4
0.6
0.8
0 0.005 0.01 0.015 0.02 0.025
k (1/min)
H2O2 dose (mol/L)
Fig. 8. Effect of H
2
O
2
dose on the decolouration of DB15. Conditions: ZVI 1 g/L, pH
3.0, ultrasound density 120 W/L.
C.-H. Weng et al. / Ultrasonics Sonochemistry 20 (2013) 970–977 973
(4)). Theoretically larger amount of H
2
O
2
should result in the gen-
eration of more active OH
radicals if ferrous irons are not a limiting
factor in the system. However, in this study, such a phenomenon
does not seem to occur. In fact, a high concentration of H
2
O
2
has
an adverse effect on Fenton’s reaction. This result has been re-
ported elsewhere [2,26]. In the Fenton system, when the available
H
2
O
2
exceeds a critical concentration, OH
could be depleted
through the scavenging of OH
by excessive H
2
O
2
and the recombi-
nation of OH
(Eq. (2)). The scavenging effect resulting from an
overdose of H
2
O
2
will decrease the decolouration rate. Compared
with the value of rate constants shown in Fig. 8, the rate constant
for 5.15 10
3
MH
2
O
2
is 0.527 min
1
, which is the highest in the
tested dose range. Therefore, this concentration of H
2
O
2
was
selected as an optimum dose for the decolouration of DB15 by Fen-
ton/US.
3.6. Effects of addition of salts
Textile effluents normally contain a variety of inorganic salts. A
number of studies have shown that the presence of salts could af-
fect the decolouration of dyes in the Fenton process. Malik and
Saha [27] have reported that chloride salts can decrease the rate
of decolouration. However, the effect of SO
2
4
on the decolouration
of some direct dyes is insignificant. In sono-Fenton studies, Li and
Song [28] showed that the presence of SO
2
4
,Cl
and H
2
PO
4
could
retard the oxidation of acid red 97; however, Minero et al. [29] re-
ported that the effects of SO
2
4
and Cl
on the decolouration of acid
blue 40 and methylene blue were insignificant. Apparently, the ef-
fects of various anions on the decolouration in the Fenton or sono-
Fenton processes depend on pollutant characteristics. Thus, it is
necessary to assess the impact of salts on the rate of DB15 decol-
ouration with the Fenton/US process.
A series of experiments were conducted by adding various salts
to the DB15 solution while keeping the remaining variables con-
stant at the optimised decolouration conditions (ZVI 0.5 g/L, pH
3.0, H
2
O
2
5.15 10
3
M). The target salts, NaCl, Na
2
ClO
4
,Na
2
SO
4
,
NaHCO
3
,Na
2
HPO
4
and NaNO
3
, were selected because they are
likely to be present in wastewater from textile and printing indus-
tries. Since the pH of the water has a significant effect on the chem-
ical species in the solution, at pH 3.0 and in an air atmosphere, the
dominant ion species would be Cl
, ClO
4
,NO
3
and SO
2
4
. In the case
of Na
2
HPO
4
and NaHCO
3
salt additions, the dominant species
would be H
2
PO
4
and H
2
CO
3,
respectively. The results (Fig. 9) reveal
that the decolouration of DB15 was not suppressed by the presence
of Cl
, ClO
4
,SO
2
4
,H
2
CO
3
and NO
3
species. However, the extent of
decolouration was not favourable in the presence of H
2
PO
4
. A pos-
sible reason for the retardation of DB15 oxidation is that H
2
PO
4
formed complexes with Fe(III) ion, such as ferric dihydrophosphate
(Fe(H
2
PO
4
)
3
), which are more stable than the complexes formed by
other ions (Cl
, ClO
4
,SO
2
4
and NO
3
). Such complexation formation
greatly affects the distribution and reactivity of Fe species with
H
2
O
2
[30].Fig. 10 shows the effect of the presence of 0.01 M
Na
2
HPO
4
on Fe(II) and Fe(III) concentrations over time. Because
of the precipitation of Fe(H
2
PO
4
)
3(s)
that occurs in the Fenton/US
process with the addition of Na
2
HPO
4
, the concentration of Fe(II)
and Fe(III) is expected to be much lower than the process without
Na
2
HPO
4
. The H
2
PO
4
ion could accept an OH
radical [31] to form
the less reactive (H
2
PO
4
)
radical (Eq. (11)), thereby leading to the
shortage of oxidants in the system; however, H
2
PO
4
is extremely
slow in reacting with OH
, and its scavenging effect can usually
be neglected [32] in the Fenton/US process.
H
2
PO
4
þOH
H
2
PO
4
Þ
þOH
ð11Þ
3.7. Effects of ultrasound density
In a Fenton/US system, ultrasound power can affect the oxida-
tion rate of DB15 to some extent. In a liquid–solid system, ultra-
sonic cavitation produces a series of related phenomena. Cavity
collapse near an extended solid surface becomes non-spherical,
which drives high-speed jets of liquid into the surface and creates
shockwave damage to the surface [33]. Thus, increasing ultrasound
intensity (power) would enlarge the effects of acoustic cavitation,
i.e. increasing the force of bubble collapse. As such, the cavitation
collapse of bubbles on the ZVI surface increases the surface area
and reactivity of ZVI which speedens up the release of Fe(II). Con-
sequently, the production of hydroxyl radicals and the oxidation
rate increase. Fig. 11 shows the effects of ultrasound power on
the decolouration of DB15. The results clearly indicate that oxida-
tion is prominent at a higher acoustic power level. Fitting results of
first-order kinetics show that the oxidation rate accelerated from
7.14 10
2
to 3.90 10
1
min
1
as power increased from 80 to
140 W/L, respectively. If the goal of the treatment is a removal effi-
ciency of 99%, the reaction time for achieving this goal is only
10 min at a power of 140 W/L, while the processing time becomes
three times with a power of 80 W/L. An increase of ultrasound
power leads to an enhancement of the Fenton/US performance
by accelerating the decolouration rate within a short period of
time. Thus, to shorten treatment time and ultimately lower energy
costs, the successful application of this process is achieved by a
higher level of ultrasound power.
3.8. Cost-effectiveness analysis of Fenton/US system
Table 1 provides an economic analysis of the Fenton/US process.
Note that the analysis is preliminary because it only includes the
costs of operating parameters at bench scale, including energy
expenditure, addition of H
2
O
2
and ZVI. When the system is applied
in the field, capital cost, equipment maintenance and depreciation
costs should be included. For a 98% oxidation efficiency in the
treatment of 4.7 10
5
M DB15 (ADMI 4130), energy requirement
ranged from 23.3 to 41.7 kW h/m
3
, which is equivalent to 1.87 to
3.33 USD/m
3
based on the current cost of electricity in Taiwan.
Based on a unit price of 500 USD/tonne, the cost of ZVI aggregates
would be 0.25 USD/m
3
. Assuming that 0.5 g/L ZVI is consumed,
when purchased in large quantities, the cost of ZVI is approxi-
mately 1 USD/kg. As indicated in Table 1, the cost of H
2
O
2
is
0 100
1 10-5
2 10-5
3 10-5
4 10-5
5 10-5
0 10203040506070
No salt
NaCl 1x10-2 M
NaClO4 1x10-2 M
Na2SO4 1x10-2 M
NaHCO3 1x10-2 M
Na2HPO4 1x10-2 M
NaNO3 1x10-2 M
Direct Blue 15 (mol/L)
Time (min)
Fe(0) 0.5 g/L
H2O2 5.15x10 -3 M
pH 3.0
US 120W (60kHz)
Fig. 9. Effect of salt addition on the decolouration of DB15. Solid lines are the best
fit of first-order kinetic equation.
974 C.-H. Weng et al. / Ultrasonics Sonochemistry 20 (2013) 970–977
negligible. The addition of ZVI constitutes approximately 10% of
the total cost, a relatively small expenditure compared to the cost
of power, which constitutes 90% of the total cost. The cost of oper-
ating the process at the higher rate of acoustic power (140 W/L) is
only 2.12 USD/m
3
, which is 60% less than the lower rate (100 W/L).
The results also show that the lowest power density (80 W/L) can
only degrade 80% of 4130 ADMI within 30 min, which does not
meet the effluent discharge limit of 550 ADMI. Although increasing
the treatment time can potentially degrade the dye further, cost of
power will increase proportionally. In general, power consumption
is a major concern. Our results show that, if the acoustic power is
properly controlled, using the Fenton process with ultrasound does
not incur a substantial increase in power consumption.
3.9. Effects of recycled ZVI
Performance using recycled ZVI in the Fenton/US process was
monitored for three consecutive runs. The results for decolouration
and the first-order rate constants are shown in Fig. 12. The exper-
imental data implies that a negative effect results from using recy-
cled ZVI. Unlike the rapid decolouration observed when fresh ZVI
was used, the use of recycled ZVI retarded the decolouration pro-
cess significantly. The value of kis inversely proportional to the
number of times ZVI is recycled. If ZVI is recycled twice over three
runs, the value of kdrops from approximately 0.260 L/min for the
first run to 0.164 and 0.104 L/min for the second and third runs,
respectively. Although the rate of decolouration decreases, a decol-
ouration efficiency of 99% for DB15 could be achieved after 25 min
treatment for the second run using recycled ZVI. The dwindling of
decolouration efficiency evident when recycled ZVI was used could
be partly because of the gradual loss of ZVI activity and Fe content.
In addition, the experimental observations could partly be
explained by the shading effects on the release of Fe(II). An
increase in the number of times ZVI is recycled leads to an increase
0
1
2
3
4
5
0 5 10 15 20 25 30
Fenton/US: No salt
Fenton/US: With 0.01 M Na2HPO4
Fe(II) mg/L
Time (min)
With Na 2HPO4
No salt
(a)
0
1
2
3
4
5
0 5 10 15 20 25 30
Fenton/US
Fenton/US/Na2HPO4
Fe(III) mg/L
Time (min)
With Na2HPO4
No salt
(b)
Fig. 10. Effect of Na
2
HPO
4
on the concentration of (a) Fe(II) and (b) Fe(III) in Fenton/
US process. Condition: Na
2
HPO
4
110
2
M, H
2
O
2
5.15 10
3
M, ZVI 0.5 g/L, pH
3.0, US power density 120 W/L.
0 100
1 10-5
2 10-5
3 10-5
4 10-5
5 10-5
0 5 10 15 20 25 30 35
US 80 W/L
US 100 W/L
US 120 W/L
US 140 W/L
Direct Blue 15(mol/L)
Time (min)
0
0.1
0.2
0.3
0.4
0.5
60 80 100 120 14 0
k (1/min)
US density (W/L)
Fig. 11. Effect of ultrasound density on the decolouration of DB15. Solid lines are
the best fit of first-order kinetic equation. Conditions: ZVI 0.5 g/L, pH 3.0, H
2
O
2
5.15 10
3
M.
Table 1
Cost analysis of Fenton/US process.
Power intensity
(W/L)
Decolourization
efficiency (%)
Reaction time (min) Power consumption
(kW h/m
3
)
Electricity
(USD/m
3
)
Fe(0)
(USD/m
3
)
H
2
O
2
(USD/m
3
)
Total cost
(USD/m
3
)
80 80 30 40.0 3.20 0.25 1 10
2
3.45
100 98 25 41.7 3.33 0.25 1 10
2
3.58
120 98 15 30.0 2.40 0.25 1 10
2
2.65
140 98 10 23.3 1.87 0.25 1 10
2
2.12
Note: (1). The current electricity rate in Taiwan is 0.08 USD/kW h. (2). Power consumption (kW h/m
3
) = Power density (W/L) 1000 W/kW time (min) 60 min/
h1000 L/m
3
; Electricity fee (USD/m
3
) = Power consumption (kW h/m
3
)0.08 USD; Fe(0) cost (USD/m
3
) = 0.5 g/L 0.001 kg/g 1000 L/m
3
500 USD/ton 0.001 ton/kg;
H
2
O
2
cost (USD/m
3
) = 0.0103 mol/L 18 g/mol 0.001 kg/g 1 USD/kg 1000 L/m
3
.
C.-H. Weng et al. / Ultrasonics Sonochemistry 20 (2013) 970–977 975
in the amount of iron hydroxide precipitates that cover the ZVI sur-
face and mask the release of Fe(II), which slows down the forma-
tion of hydroxyl radicals.
4. Conclusions
The Fenton process with ultrasound irradiation using ZVI aggre-
gates as the catalyst is more effective for the decolouration of
direct blue 15 in aqueous solution compared with the use of either
processes alone. The use of ultrasound can enhance the corrosion
of ZVI, resulting in the continuous release of large quantities of
Fe(II), which in turn triggers a more effective Fenton’s reaction.
Compared to the use of the Fenton reagent alone, ultrasonic irradi-
ation enhanced Fenton’s reaction significantly; the production of
OH
radicals was markedly increased. A positive synergistic effect
was observed and the oxidation rate was substantially enhanced
by ultrasonic irradiation.
The decolouration was obviously affected by pH, ZVI dose, H
2
O
2
dose, ultrasound power density and addition of salts. The decolour-
ation kinetics of DB15 can be described by the first-order equation.
Under a constant ultrasonic irradiation frequency (60 kHz) and
power density (120 W/L), the optimum conditions for decolour-
ation were: 4.7 10
5
M DB15, pH 3.0, ZVI 1 g/L and H
2
O
2
5.15 10
3
M. Under these optimum conditions, a decolouration
efficiency of 99% was achieved within 10 min, which complies with
Taiwan’s effluent true colour regulation of 550 ADMI. The decol-
ouration was inhibited by the presence of H
2
PO
4
in the solution,
but it was not affected significantly by the presence of H
2
CO
3
,
Cl
, ClO
4
,NO
3
or SO
2
4
. An increase of ultrasonic power led to
the enhancement of Fenton/US performance and decreased the en-
ergy cost by shortening the treatment time. ZVI can be reused;
however, the rate of decolouration decreased with repeated use
of ZVI.
The operating cost, including chemicals and energy expendi-
ture, for the process with an acoustic power of 140 W/L was esti-
mated to be only 2.12 USD/m
3
for a 99% decolouration of 4130
ADMI within 10 min. This study shows that oxidation power is
remarkably enhanced by ultrasonic irradiation. Overall, it can be
concluded that the Fenton/US process using ZVI aggregates as the
catalyst is a cost-effective technology for treating DB15 provided
the operational conditions are properly controlled.
Acknowledgements
This research was supported by the National Science Council of
R.O.C. (NSC 99-2221-E-214-021-MY3). The authors are very much
thankful to the fellows of Connelly-GPM Inc., USA for providing the
ZVI aggregate samples.
References
[1] R. Ganesh, G.D. Boardman, D. Michelsen, Fate of azo dyes in sludges, Water Res.
28 (1994) 1367–1376.
[2] J.H. Sun, S.H. Shi, Y.F. Lee, S.P. Sun, Fenton oxidative decolorization of the azo
dye Direct Blue 15 in aqueous solution, Chem. Eng. J. 155 (2009) 680–683.
[3] S. Garcia-Segura, F. Centellas, C. Arias, J.A. Garrido, R.M. Rodríguez, P.L. Cabot1,
E. Brillas, Comparative decolorization of monoazo, diazo and triazo dyes by
electro-Fenton process, Electrochim. Acta 58 (2011) 303–311.
[4] S.H. Chang, S.H. Chuang, H.C. Li, H.H. Liang, L.C. Huang, Comparative study on
the degradation of I.C. Remazol Brilliant Blue R and I.C. Acid Black 1 by Fenton
oxidation and Fe
0
/air process and toxicity evaluation, J. Hazard Mater. 166
(2009) 1279–1288.
[5] Y.S. Ma, M. Kumar, J.G. Lin, Degradation of carbofuran-contaminated water by
Fenton process, J. Environ. Sci. Health A: Tox. Hazard. Subst. Environ. Eng. 44
(2009) 914–920.
[6] D.H. Bremner, A.E. Burgess, D. Houllemare, K.C. Namkung, Phenol degradation
using hydroxyl radicals generated from zero-valent iron and hydrogen
peroxide, Appl. Catal. B: Environ. 63 (2006) 15–19.
[7] A.G. Chakinala, P.R. Gogate, A.E. Burgess, D.H. Bremner, Treatment of industrial
wastewater effluents using hydrodynamic cavitation and the advanced Fenton
process, Ultrason. Sonochem. 15 (2008) 49–54.
[8] A.G. Chakinala, P.R. Gogate, A.E. Burgess, D.H. Bremner, Industrial wastewater
treatment using hydrodynamic cavitation and heterogeneous advanced Fenton
processing, Chem. Eng. J. 152 (2009) 498–502.
[9] M. Umar, H.A. Aziz, M.S. Yusoff, Trends in the use of Fenton, electro-Fenton and
photo-Fenton for the treatment of landfill leachate, Waste Manage. 30 (2010)
2113–2121.
[10] A.N. Soon, B.H. Hameed, Heterogeneous catalytic treatment of synthetic dyes
in aqueous media using Fenton and photo-assisted Fenton process,
Desalination 269 (2011) 1–16.
[11] P.V. Nidheesh, R. Gandhimathi, Trends in electro-Fenton process for water and
wastewater treatment: an overview, Desalination 299 (2012) 1–15.
[12] H.-E. Kim, J. Lee, H. Lee, C. Lee, Synergistic effects of TiO
2
photocatalysis in
combination with Fenton-like reactions on oxidation of organic compounds at
circumneutral pH, Appl. Catal. B: Environ. 115 (2012) 219–224.
[13] R. Chand, N.H. Ince, P.R. Gogate, D.H. Bremner, Phenol degradation using 20,
300 and 520 kHz ultrasonic reactors with hydrogen peroxide, ozone and zero
valent metals, Sep. Purif. Technol. 67 (2009) 103–109.
[14] D.H. Bremner, S.D. Carlo, A.G. Chakinala, G. Cravotto, Mineralisation of 2,4-
dichlorophenoxyacetic acid by acoustic or hydrodynamic cavitation in
conjunction with the advanced Fenton process, Ultrason. Sonochem. 15
(2008) 416–419.
[15] H. Zhang, J. Zhang, C. Zhang, F. Liu, D.B. Zhang, Degradation of C.I. acid orange 7
by advanced Fenton process in combination with ultrasonic irradiation,
Ultrason. Sonochem. 16 (2009) 325–330.
[16] C. Özdemir, M.K. Öden, S. Sßahinkaya, E. Kalipçi, Color removal from synthetic
textile wastewater by sono-Fenton process, Clean Soil Air Water 39 (2011) 60–
67.
[17] B. Chen, X. Wang, C. Wang, W. Jiang, S. Li, Degradation of azo dye direct sky
blue 5B by sonication combined with zero-valent iron, Ultrason. Sonochem. 18
(2011) 1091–1096.
[18] J. Liang, S. Komarov, N. Hayashi, E. Kasai, Improvement in sonochemical
degradation of 4-chlorophenol by combined use of Fenton-like reagents,
Ultrason. Sonochem. 14 (2007) 201–207.
[19] T. Zhou, T.T. Lim, X. Lu, Y. Li, F.S. Wong, Simultaneous degradation of 4CP and
EDTA in a heterogeneous ultrasound/Fenton like system at ambient
circumstance, Sep. Purif. Technol. 68 (2009) 367–374.
[20] Y. Dong, J. Chen, C. Li, H. Zhu, Decoloration of three azo dyes in water by
photocatalysis of Fe (III)–oxalate complexes/H
2
O
2
in the presence of inorganic
salts, Dyes Pigm. 73 (2007) 261–268.
[21] American Public Health Association (APHA), Standard Methods for the
Examination of Water and Waste Water, twentieth ed., APHA, Washington,
DC, USA, 1998. Method 2120E, pp. 2-7–2-8.
[22] Y.T. Lin, C.H. Weng, F.Y. Chen, Effective Removal of AB24 Dye by nano/micro-
size zero-valent Iron, Sep. Purif. Technol. 64 (2008) 26–30.
[23] W.Z. Tang, C.P. Huang, 2,4-dichlorophenol oxidation kinetics by Fenton’s
reagent, Environ. Technol. 17 (1996) 1371–1378.
[24] H.J. Fan, S.T. Huang, W.H. Chung, J.L. Jan, W.Y. Lin, C.C. Chen, Degradation
pathways of crystal violet by Fenton and Fenton-like systems: condition
optimization and intermediate separation and identification, J. Hazard Mater.
171 (2009) 1032–1044.
[25] J. Lin, X. Zhao, D. Liu, Z. Yu, Y. Zhang, H. Xu, The decoloration and
mineralization of azo dye C.I. acid red 14 by sonochemical process: rate
improvement via Fenton’s reactions, J. Hazard Mater. 157 (2008) 541–546.
0 100
1 10-5
2 10-5
3 10-5
4 10-5
5 10-5
0 5 10 15 20 25 30 35
1st run
2nd run
3rd run
Direct Blue 15(mol/L)
Time (min)
C04.7x10-5 M
Fe(0) 0.5 g/L
H2O2 5.15x10-3 M
pH 3.0
0
0.1
0.2
0.3
0.4
01234
y = 0.332 - 0 .078x R 2= 0.983
k (1/min)
Regeneration times of Fe(0)
Fig. 12. Effect of recycled ZVI on the decolouration of DB15. Solid lines are the best
fit of first-order kinetic equation.
976 C.-H. Weng et al. / Ultrasonics Sonochemistry 20 (2013) 970–977
[26] S. Wang, A comparative study of Fenton and Fenton-like reaction kinetics in
decolourisation of wastewater, Dyes Pigm. 76 (2008) 714–720.
[27] P.K. Malik, S.K. Saha, Oxidation of direct dyes with hydrogen peroxide using
ferrous ion as catalyst, Sep. Purif. Technol. 31 (2003) 241–250.
[28] J.T. Li, Y.L. Song, Degradation of AR 97 aqueous solution by combination of
ultrasound and Fenton reagent, Environ. Prog. Sustainable Energy 29 (2010)
101–106.
[29] C. Minero, P. Pellizzari, V. Maurino, E. Pelizzetti, D. Vione, Enhancement of dye
sonochemical degradation by some inorganic anions present in natural waters,
Appl. Catal. B: Environ. 77 (2008) 308–316.
[30] E.M. Siedlecka, A. Wie˛ckowska, P. Stepnowski, Influence of inorganic ions on
MTBE degradation by Fenton’s reagent, J. Hazard Mater. 147 (2007) 497–502.
[31] P.A. Morozov, B.G. Ershov, The influence of phosphates on the decomposition
of ozone in water: chain process inhibition, Russ. J. Phys. Chem. A, Focus Chem.
84 (2010) 1136–1140.
[32] J. Hoigne, Chemistry of aqueous ozone and transformation of pollutants by
ozonation and advanced oxidation processes, in: The Handbook of
Environmental Chemistry, in: J. Hrubec (Ed.), Part C, Quality and Treatment
of Drinking Water, vol. 5, Springer-Verlag, Berlin, 1998.
[33] K. Suslick, Sonochemistry, Science 247 (1990) 1439–1445.
C.-H. Weng et al. / Ultrasonics Sonochemistry 20 (2013) 970–977 977
... In Fenton-like (based using nano zero-valent iron) and photo-Fenton-like process (UV/ nZVI/H 2 O 2 ), ferrous and then ferric are formed. During a regular cycle, these ions increase the efficiency of the process [14]. The supported nanoparticle on a various support materials such as zeolite activated carbon etc. has been studied in photodegradation of organic pollutants such as dyes [15]. ...
... AOP is a simple and highly efficient technique for removal of organic pollutant through sonolysis [16][17][18], UV, Fenton processes, and ozonation. Amongst various known methods of AOPs, sonolysis is more advantageous and fascinated the interest of many researchers for elimination of unmanageable contaminants from waste water bodies [19] due to its zero secondary pollutant generation, highly potential and greener technology for water security purpose. In this context, sono-catalysis becomes promising field due to its high efficiency, needs less maintenance, easy to manipulate and low energy input [20]. ...
Article
Full-text available
The industrial revolution has raised major concern of environmental pollution, due to excess release of hazardous chemical, dyes, etc. into water bodies. The present crises of water have attracted considerable attention of scientific drive on development of new technologies on recovery and water purification. Sonochemistry with zero secondary pollutants has been considered as a potential, and greener technologies for water security. In the present work, we synthesized cerium oxide (CeO2) nanoparticles for its dynamic application in opto-electronic field. The work involves the systematic study of sono-catalytic degradation of methylene blue (MB) (40 kHz, 80W) using cerium oxide (CeO2) nanomaterial as a catalyst. Time-dependent synthesis of CeO2 is carried out using co-precipitation method at 120 °C, followed by calcination. The X-ray diffraction (XRD) pattern and Raman studies confirms high crystallinity (~ crystallite size of 10–15 nm) with cubic fluorite phase. The microscopic studies using field emission scanning electron microscope (FESEM), and transmission electron microscope (TEM) confirmed a spindle shaped morphology. Ninety percent sono-cataclytic degradation of methylene blue (MB) is observed using CeO2 synthesized with a reaction time of 2 h (surface area of 62 m²g⁻¹ with pore size of 1.5 nm). The observed sono-catalytic degradation, follow first order reaction kinetics with regression coefficient of 0.92 with rate constant 0.018 min⁻¹. The study reveals that the CeO2 spindles can act as excellent catalysts for sono-catalytic degradation activities and for water purification owing to their excellent surface properties. Graphical Abstract The sono-catalytic degradation of methylene blue (40 kHz, 80W) is systematically studied using synthesized spindle shaped CeO2 catalyst. Ninety percent sono-cataclytic degradation of MB is observed using spindle shaped CeO2. The study reveals that the CeO2 spindles can act as excellent catalysts for sono-catalytic degradation activities owing to their excellent surface properties.
... Based on the presented results, it can be concluded that, as in the case of nZVI concentration, there is a limit value (100 ppm) for H 2 O 2 concentration above which no significantly higher process efficiency is achieved, which is again a consequence of the unwanted removal of ·OH radicals (Eq. 15) [42]. Hydroperoxy radicals (·HO 2 ) formed in this process have significantly less activity compared to ·OH radicals. ...
Article
Full-text available
The application of zero-valent iron nanoparticles (nZVI) is an area of growing interest, especially in the remediation processes of waters of different origin. The standard synthetic procedure for nZVI involves the chemical reduction of Fe²⁺ or Fe³⁺ with the solution of a strong reducing agent. This type of synthetic route leads to the formation of new toxic species. The green synthesis of nZVI using plant extracts is, therefore, an excellent alternative to the above-mentioned conventional methods. Within the scope of this work, the amorphous nZVI was synthesized by green tea mediated single-step method from Fe³⁺ as precursors, denoting an eco-friendly approach. Additionally, the physico–chemical characterization (TEM, SEM/EDS, size of particles/aggregates, XRD, BET) of nZVI was performed. The catalytic activity of the obtained nZVI was tested in a Fenton-like reaction of the removal/degradation of three representative industrial textile dyes (methylene blue, methyl orange and bromothymol blue) from water. In order to optimize the process parameters, the effect of homogenization treatment, nZVI and H2O2 concentration on catalytic performance were also done followed by degradation kinetic study using four different models. The results of this work have shown the exceptionally high efficiency of the green nZVI which is reflected in almost complete degradation of all tested dyes within very short period of time, despite their amorphous nature, aggregates arrangement and very low specific surface area. Furthermore, the examined degradation kinetics has excellently followed the BMG model, indicating that the degradation process of the investigated dyes proceeds via two stages (two-stage pattern)—a fast one and a much slower one, regardless of the used reaction conditions. Based on all gained results, particular degradation mechanisms were proposed. Graphical Abstract
... But it also has some drawbacks, like a narrow pH, high secondary waste mainly Fe +3 , and difficulty restore the catalyst (Fe +2 ) [8]. To solve the challenges, the heterogeneous Fenton is suggested as an effective environmentally friendly, and inexpensive process [10]. ...
Article
This study used a continuous photo-Fenton-like method to remediate textile effluent containing azo dyes especially direct blue 15 dye (DB15). A Eucalyptus leaf extract was used to create iron/copper nanoparticles supported on bentonite for use as catalysts (E@B-Fe/Cu-NPs). Two fixed-bed configurations were studied and compared. The first one involved mixing granular bentonite with E@B-Fe/Cu-NPs (GB- E@B-Fe/Cu-NPs), and the other examined the mixing of E@B-Fe/Cu-NPs with glass beads (glass beads-E@B-Fe/Cu-NPs) and filled to the fixed-bed column. Scanning electron microscopy (SEM), zeta potential, and atomic forces spectroscopy (AFM) techniques were used to characterize the obtained particles (NPs). The effect of flow rate and DB15 concentration on the photo-Fenton-like degradation of DB15 in continuous fixed-bed systems was investigated. To optimize both studied systems, the response surface methodology using the central composite design (CCD) was used. The analysis shows that the removal efficiency for GB-E@B-Fe/Cu-NPs was 81% and for glass beads-E@B-Fe/NPs was 62.6%.
... The values of K L and q max were computed from the slope and intercept of the Langmuir plot of 1/q e versus 1/C e (Langmuir I, 1918). The essential features of the Langmuir isotherm may be expressed in terms of equilibrium parameter R L , which is a dimensionless constant referred to as separation factor or equilibrium parameter (Weng et al., 2013;and Dada et al., 2012). ...
Article
Full-text available
Vo l. 7 , No. 1 F e b r u a r y 2 0 1 8 15 This article can be downloaded from http://www.ijerst.com/currentissue.php Int. Activated carbon produced from Gudali (Bos Inducus) bones which are natural precursors of animal origin and was used in removing of chromium (VI) of wastewater by adsorption.A simple elaboration of an animal activated carbon (AAC) based on chemical activation with phosphoric acid, was used.The tests of adsorption are carried out by mixing by agitation (120 trs/min) 2 g of activated carbon with 50 mL the solution of Cr (VI). The study of the kinetics shows that the maximum of adsorption is obtained at 5 minutes. The adsorbed quantities increase with the pH, the temperature and the concentration, but it decreases with the mass of the activated carbon. The best kinetic model adapted for the adsorption of Cr (VI) is the pseudo-second order. The adsorption isotherms are in conformity with the Langmuir and Freundlich models. The quantities adsorbed on the monolayer from the Langmuir model are of 35,714 mg/g.
Article
Full-text available
NOM ‫محلول‬ ‫در‬) ‫ـدزدايي(‬ ‫گنـ‬ ‫ـانبي‬ ‫یـ‬ ‫ـوزت‬ ‫ـ‬ ‫مح‬ ‫ـاد‬ ‫ايجـ‬ ‫ـبم‬ ‫سـ‬ ‫ـه‬ ‫بـ‬ ‫ـي‬ ‫آبـ‬ ‫ـای‬ ‫هـ‬ DBP ‫ـه‬ ‫یملـ‬ ‫از‬) ‫تری‬ ‫هالومتان‬ ‫ها(‬ THMs ‫اسیدها(‬ ‫استیک‬ ‫هالو‬ ‫و‬) HAAs ‫محلول‬ ‫کلريناسیون‬ ‫نتیجه‬ ‫در‬ ،) ‫فر‬ ‫در‬ ‫آبي‬ ‫های‬ ‫آ‬ ‫متداول‬ ‫فیه‬ ‫ت‬ ‫يند‬ ‫ـه‬ ‫بـ‬ ‫تویه‬ ‫با‬ ‫و‬ ‫خط‬ ‫است.‬ ‫کرده‬ ‫م(طوف‬ ‫خود‬ ‫به‬ ‫را‬ ‫زيادی‬ ‫تویهات‬ ‫عموم،‬ ‫سالمت‬ ‫برای‬ ‫بهداشتي‬ ‫بالقوه‬ ‫رات‬ ‫لذا‬ ‫ـه‬ ‫مطال(ـ‬ ‫اين‬ ‫از‬ ‫هدف‬ ‫و‬ ‫ـین‬ ‫ت(یـ‬ ‫ـي‬ ‫بررسـ‬ ‫فنتون‬ ‫تلفیقي‬ ‫فرآيندهای‬ ‫عملكرد‬ ‫ـ‬ ‫هالومتان‬ ‫تری‬ ‫سازهای‬ ‫پیش‬ ‫حذف‬ ‫در‬ ‫اولتراسونیک‬ ‫محیط‬ ‫از‬ ‫ها‬ ‫ـ‬ ‫صـ‬ ‫ـن‬ ‫آهـ‬ ‫از‬ ‫ـتفاده‬ ‫اسـ‬ ‫ـا‬ ‫بـ‬ ‫آبي‬ ‫های‬ ‫فر‬ ‫بو‬ ‫د.‬ ‫برر‬ ‫روش‬ ‫سي‬ : ‫مطال(ه‬ ‫يک‬ ‫اين‬ ‫مي‬ ‫تجربي‬ ‫باشد‬ ‫سال‬ ‫در‬ ‫که‬ 1400 ‫شد‬ ‫انجام‬ ‫ـت‬ ‫اسـ‬ ‫ه‬ ، ‫روش‬ ‫ـن‬ ‫ايـ‬ ‫در‬ ‫اثر‬ ‫ـا‬ ‫متغیرهـ‬ ‫ـامش‬ ‫شـ‬ ‫ـتقش‬ ‫مسـ‬ ‫؛‬ ‫ـت‬ ‫غلظـ‬ ‫آزينده،‬ ‫اولیه‬ ‫واکنش،‬ ‫زمان‬ ‫استفاده،‬ ‫مورد‬ ‫کاتالیست‬ ‫غلظت‬ ‫و‬ ‫استفاده‬ ‫مورد‬ ‫هیدروژن‬ ‫پروکسید‬ ‫میزان‬ pH ‫ـده‬ ‫شـ‬ ‫مشخص‬ ‫محدوده‬ ‫در‬ ‫یهت‬ ‫پیش‬ ‫حذف‬ ‫تری‬ ‫سازهای‬ ‫هالومتان‬ ‫محی‬ ‫از‬ ‫ها‬ ‫ط‬ ‫صفر‬ ‫آهن‬ ‫از‬ ‫استفاده‬ ‫با‬ ‫آبي‬ ‫های‬ ‫سطح‬ ‫روش‬ ‫با‬ ‫ـ‬ ‫ـي‬ ‫یراحـ‬ ‫ـای‬ ‫مبنـ‬ ‫ـر‬ ‫بـ‬ ‫پاسخ‬ ‫ـم‬ ‫مرکـ‬ ‫مرکزی‬ (CCD) ‫گرفت.‬ ‫قرار‬ ‫ارزيابي‬ ‫مورد‬ ‫آزمايش‬ ‫نرم‬ ‫از‬ ‫استفاده‬ ‫با‬ ‫ها‬ ‫ـزار‬ ‫افـ‬ Design expert ‫روش‬ ‫ـه‬ ‫بـ‬ ‫و‬ CCD-RSM ‫ـد.‬ ‫گرديـ‬ ‫ـي‬ ‫یراحـ‬ ‫مورد‬ ‫اشكال‬ ‫و‬ ‫نمودارها‬ ‫واکنش‬ ‫سنتیک‬ ‫ت(یین‬ ‫یهت‬ ‫نیاز‬ ‫نرم‬ ‫از‬ ‫استفاده‬ ‫با‬ ‫ها‬ ‫اف‬ ‫زا‬ ‫ر‬ Excel ‫ـین‬ ‫ت(یـ‬ ‫ـت‬ ‫یهـ‬ ‫ـاری‬ ‫آمـ‬ ‫ـای‬ ‫پارامترهـ‬ ‫و‬ ‫ـیم‬ ‫ترسـ‬ ‫فر‬ ‫بهینه‬ ‫شرايط‬ ‫آ‬ ‫از‬ ‫استفاده‬ ‫با‬ ‫يند‬ ‫واريانس‬ ‫آنالیز‬ ‫آزمون‬ ، ‫رگرسیون‬ ‫فیشر‬ ‫و‬ ‫شدند‬ ‫تحلیش‬ ‫و‬ ‫تجزيه‬. ‫يافته‬ ‫ها‬ : ‫که‬ ‫است‬ ‫آن‬ ‫از‬ ‫حاکي‬ ‫نتايج‬ ‫نگهداشتن‬ ‫ثابت‬ ‫با‬ ‫ـید‬ ‫اسـ‬ ‫ـک‬ ‫هیومیـ‬ ‫ـت‬ ‫غلظـ‬ ‫ـطح‬ ‫سـ‬ ‫در‬ 75 / 64 ‫میلي‬ ‫ـر‬ ‫لیتـ‬ ‫در‬ ‫ـرم‬ ‫گـ‬ ‫ـت‬ ‫کاتالیسـ‬ ‫ـت‬ ‫غلظـ‬ ‫و‬ 76 / 106 ‫میلي‬ ‫لیتر‬ ‫در‬ ‫گرم‬ ‫و‬ ‫ثابت‬ ‫میزان‬ 39 / 4 pH= ‫و‬ ‫ـید‬ ‫پراکسـ‬ ‫هیدروژن‬ ‫که‬ ‫افتد‬ ‫مي‬ ‫اتفاق‬ ‫زماني‬ ‫اسید‬ ‫هیومیک‬ ‫حذف‬ ‫بازده‬ ‫بیشترين‬ ، ‫خود‬ ‫مقدار‬ ‫بیشینه‬ ‫در‬ ‫واکنش‬ ‫انجام‬ ‫زمان‬ ‫سطح‬ ‫در‬ ‫ترتیم‬ ‫به‬ 35 ‫میلي‬ ‫مول‬ ‫و‬ 75 / 88 ‫دقیقه‬ ‫شرايط‬ ‫اين‬ ‫در‬ ‫که‬ ‫باشند‬ ‫از‬ ‫حذف‬ ‫میزان‬ 80 ‫درصد‬ ‫به‬ 92 ‫يا‬ ‫افزايش‬ ‫درصد‬ ‫است‬ ‫فته‬. ‫هم‬ ‫ـي‬ ‫مـ‬ ‫ـزايش‬ ‫افـ‬ ‫واکنش‬ ‫بازده‬ ‫میزان‬ ‫کاتالیست‬ ‫مقدار‬ ‫افزايش‬ ‫با‬ ‫چنین‬ ‫ـد‬ ‫يابـ‬. ‫ـنش‬ ‫واکـ‬ ‫ـو‬ ‫یبـ‬ ‫ـای‬ ‫هـ‬ ‫فر‬ ‫به‬ ‫مربوط‬ ‫آ‬ ‫راديكال‬ ‫افزايش‬ ‫باعث‬ ‫آهن‬ ‫حضور‬ ‫فنتون،‬ ‫يند‬ ‫مي‬ ‫هیدروکسیش‬ ‫های‬ ‫شود‬ ‫بازده‬ ‫افزايش‬ ‫دزيش‬ ‫از‬ ‫يكي‬ ‫اين‬ ، ‫ـي‬ ‫مـ‬ ‫حذف‬ ‫ـد‬ ‫توانـ
Article
Full-text available
Background & aim: The presence of natural organic substances (NOM) in aqueous solutions due to the creation of disinfection by-products (DBP) including trihalomethanes (THMs) and haloacetic acids (HAAs), as a result of the chlorination of aqueous solutions in the conventional purification process and according to Potential health risks for public health have attracted a lot of attention. Therefore, the aim of the present study was to determine and investigate the performance of combined Fenton-ultrasonic processes in removing the precursors of trihalomethanes from aqueous environments using zero iron. Methods: In the present experimental study conducted in 2021, the effect of independent variables included: the initial concentration of the pollutant, the reaction time, the concentration of the catalyst used, the amount of hydrogen peroxide used and the pH in the specified range to remove the precursors of trihalomethanes from water environments using zero iron with the surface-response method based on composite design. Central (CCD) was evaluated. Experiments were designed using Design expert software and CCD-RSM method. The graphs and figures required to determine the kinetics of the reactions were drawn using Excel software and the statistical parameters were analyzed to determine the optimal conditions of the process using analysis of variance, regression and Fisher tests. Results: The results indicated that by keeping the humic acid concentration at 64.75 mg/liter and the catalyst concentration at 106.76 mg/liter and the constant pH=4.39, the highest humic acid removal efficiency occurs when It happens that the hydrogen peroxide and the reaction time are at their maximum level of 35 mmol and 88.75 minutes, respectively, and in this condition, the removal rate has increased from 80% to 92%, and also with the increase in the amount of the catalyst, the reaction efficiency increases. find According to the reactions related to the Fenton process, the presence of iron causes an increase in hydroxyl radicals, this can be one of the reasons for increasing the removal efficiency. Conclusion: In the mentioned Ultrasonic-Fenton process under laboratory conditions, iron ion as a catalyst reacted with an oxidant such as hydrogen peroxide in an acidic environment and produces hydroxyl radicals. The efficiency of this method was influenced by various factors such as pH, concentration of iron and hydrogen peroxide, initial concentration of the pollutant and reaction time.
Article
Full-text available
An effective hybrid system was applied as the first report for the successful treatment of key pollutants (hydroxypropyl guar gum, HPG) in fracturing flowback fluid, and the synergistic index of the hybrid system was 20.45. In this regard, chemical oxygen demand (COD) removal ratio was evaluated with various influencing operating factors including reaction time, H2O2 concentration, Fe²⁺ concentration, ultrasonic power, initial pH, and temperature. The optimal operating parameters by single-factor analysis method were: the pH of 3.0, the H2O2 concentration of 80 mM, the Fe²⁺ concentration of 5 mM, the ultrasonic power of 180 W, the ultrasonic frequency of 20-25 kHz, the temperature of 39 ℃, the reaction time of 30 min, and the COD removal rate reached 81.15%, which was permissible to discharge surface water sources based on the environmental standards. A possible mechanism for HPG degradation and the generation of reactive species was proposed. Results of quenching tests showed that various impacts of the decomposition rate by addition of scavengers had followed the order of EDTA-2Na < BQ < t-BuOH, therefore ·OH radicals had a dominant role in destructing the HPG. Based on the kinetic study, it was concluded that Chan Kinetic Model was more appropriate to describe the degradation of HPG. Identification of intermediates by GC-MS showed that a wide range of recalcitrant compounds was removed and/or degraded into small molecular compounds effectively after treatment. Under the optimal conditions, the sono-Fenton system was used to treat the fracturing flowback fluid with the initial COD value of 675.21mg/L, and the COD value decreased to 80.83mg/L after 60 min treatment, which was in line with the marine sewage discharge standard. In conclusion, sono-Fenton system can be introduced as a successful advanced treatment process for the efficient remediation of fracture flowback fluid.
Article
Full-text available
This paper present the efficiency of animal activated carbon in removal of methyl orange (MO) from aqueous solutions by adsorption. Batch experiments were conducted to study the effect of reaction parameters on the adsorption of the dye such as: contact time, initial concentrations of MB solution, adsorbent dosage, solution pH, and temperature. The surface properties were measured by some techniques including: Fourier transform infrared spectrometry and point of zero charge. Adsorption capacity for MO is 7.045 mg.g 1 at 298 K under the optimum condition of pH of 3, adsorbent dosage of 10 g.L 1 , particle size of 0.2-0.05 mm, and contact time of 10 min. The kinetic data obtained from different experiments were analyzed usingthree kinetic models namely intra-particulate diffusion, pseudo-first-order and pseudo-second-order kinetics models. Among the kinetics models studied, the pseudo-second-order model was the best applicable model to describe the adsorption of MO onto activated carbon. Freundlich and Langmuir models were used to analyze the obtained experimental data. In comparison, both models were understood to be a better fit for the experimental data, this two models show best fit with R 2 values of 0.999. The values of Gibbs free energy of adsorption ΔG°<0; this negatives values indicated the spontaneity of the adsorption process of MO onto activated carbon. The values of (ΔH°) and (ΔS°) were found to be 57.924 kJ.mol 1 and 198.538 J mol 1 .K 1 , respectively. The thermodynamics parameters indicated that the adsorption of MO onto activated carbon was spontaneous and exothermal.
Chapter
Ozonation is widely and successfully applied for many types of oxidative water treatments. Its chemical effects can be described by considering the sequences of highly selective direct reactions of molecular ozone and the reactions of the more reactive but less selective OH radicals which are always produced from decomposed ozone in aqueous systems. These radicals also control the ozone based AOPs (Advanced Oxidation Processes). In some cases even formation and reactions of additional secondary oxidants, such as carbonate radicals, hypobromite, and hydrogen peroxide have to be accounted for.
Article
The decolorization process of the monoazo Acid Orange 7, diazo Acid Red 151 and triazo Direct Blue 71, with similar aromatics and functional groups, has been comparatively studied by electro-Fenton. Solutions of 100cm3 of each azo dye in 0.05moldm−3 Na2SO4 with 0.5mmoldm−3 Fe2+ of pH 3.0 were electrolyzed in an undivided BDD/air-diffusion cell under galvanostatic conditions. Organics were oxidized with hydroxyl radicals formed at the BDD anode from water oxidation and mainly in the bulk from Fenton's reaction between cathodically generated H2O2 and added Fe2+. A simple parameter so-called initial decolorization rate was used to analyze the behaviour of the azo dyes at the beginning of the process. This parameter decreased with increasing initial azo bonds concentration due to the oxidation of more organic matter with similar amounts of hydroxyl radicals. It rose linearly with current density by the greater production of these oxidants, even when similar color removal was found at long electrolysis times, indicating that the process was mass-transfer controlled. The initial decolorization rate became lower as the number of azo bonds in the molecule increased owing to their smaller reactivity with hydroxyl radicals. Reversed-phase HPLC of electrolyzed solutions revealed the total removal of all azo dyes following a pseudo first-order kinetics with rate constants that showed the same trends as those predicted by initial decolorization rates. The pseudo first-order rate constants for decolorization obtained from absorbance decay also showed similar tendencies, but they were not useful to describe the comparative total decolorization of azo dye solutions because of the slower and parallel destruction of colored conjugated intermediates formed during the electro-Fenton treatment.
Article
Coloration of textiles is improved by inorganic salts, which increase pollution load on dyehouse effluent in general. In this paper, four inorganic salts including sodium chloride, sodium sulphate, sodium nitrate and potassium chloride were used for investigating their effects on photocatalytic decoloration of three azo dyes, Mordant Black PV, Reactive Blue MS and Direct Sky Blue 5B in water based on Fe (III)–oxalate complexes/H2O2 system in a specially designed photoreactor. Also the photocatalytic degradation and mineralization of the dyes were examined by UV–vis spectrum and TOC measurements. The results indicate that photocatalytic decoloration of the dyes is significantly limited in the presence of salts. Increasing concentration of inorganic salts in water can give rise to reduced decoloration efficiency and kinetic reaction constants of the dyes. Sodium chloride and potassium chloride exhibit stronger inhabiting power on photocatalytic decoloration and degradation than sodium sulphate and sodium nitrate under the same reaction condition.
Article
The oxidative degradation of two direct dyes, Blue 2B (B54) and Red 12B (R31) in aqueous solution has been studied using Fenton's reagent (Fe2+ and H2O2). Results show that dyes are decomposed in a two-stage reaction. In the first stage, dyes are decomposed rapidly, and somewhat less rapidly in the second stage. The effects of different system variables like initial pH of the medium, initial concentrations of the dye, Fe2+ and H2O2, reaction temperature, and added Cl− and SO4−2 ions have been investigated. The degradation rate is strongly dependent on the initial concentrations of the dye, Fe2+ and H2O2. The results indicate that the B54 and R31 can be most effectively oxidised in an aqueous medium of pH 3 at an initial [Fe2+]:[H2O2]:[dye] ratio of 1:32.9:2.4 and 1:16.5:1.8 (mM), respectively. At the optimum initial ratio of [Fe2+]:[H2O2]:[dye], 97% degradation can be achieved in 30 min at a temperature of 30°C, and 70% removal of initial chemical oxygen demand (COD) is achieved after 60 min. The degradation of both the dyes obeys first order rate kinetics with respect to the concentration of the dye in the second stage of oxidation. The results will be useful in designing wastewater treatment plants.
Article
The sonochemical degradation rate of the charged substrates Acid Blue 40 (AB40) and methylene blue (MB) is enhanced by scavengers of hydroxyl radicals such as bicarbonate, carbonate, bromide, iodide and (only in the case of AB40) nitrite. No rate variation was observed with chloride, nitrate or sulphate, excluding a mere ionic strength effect. Oxidation of bicarbonate, carbonate, bromide, iodide and nitrite yields the corresponding radicals CO3−, Br2−, I2− and NO2, reactive but less than OH. Degradation enhancement can occur if these radicals are sonochemically formed on the surface of the collapsing cavitation bubbles and undergo there radical–radical recombination at a lesser extent than OH. In this way the radicals would be more available than OH for substrate degradation, both at the bubble surface and in the solution bulk, which could more than compensate for their lower intrinsic reactivity. The varied reactivity toward different substrates of the sonochemically formed radical species could then explain why nitrite inhibits MB degradation while enhancing that of AB40. The sonochemical formation of Br2−, I2− and NO2 can give rise to halogenation and nitration in addition to oxidation processes. In fact bromo-, iodo- and nitrophenols were detected upon sonication of phenol in the presence of the corresponding anions, but only at quite elevated concentration values of nitrite, bromide or iodide (above 10mM). Formation of harmful halogeno- and nitroderivatives could therefore take place on sonication of some wastewater rather than of typical natural waters.
Article
Oxidation kinetics and mechanisms of 2,4-dichlorophenol (2,4-DCP) by Fenton's reagent were studied. The effect of pH, concentration of H2O2, and Fe2+, and 2,4-DCP on both oxidation and dechlorination kinetics was investigated. A mathematical model was developed to describe the kinetics of 2,4-DCP oxidation and chloride ion production at constant concentrations of H2O2 and Fe2+. The optimal ratio of H2O2 to Fe2+ for the oxidation of 2,4-DCP was determined to be 11 which is the same as predicted by the previously developed kinetic model.
Article
Organic compound, especially aromatic compound is the main pollutant in industrial effluent. Conventional wastewater treatment processes are inefficient for the removal of these types of toxic and hazardous pollutants from wastewater. Electro Fenton is one of the powerful and environmentally friendly emerging technologies for the remediation of wastewaters containing organic, especially aromatic compounds. This paper reviews the fundamentals and recent developments in electro Fenton process. Electro Fenton process utilizes different electrolytic reactors such as bubble reactor, filter press reactor, divided double-electrode electrochemical cell, divided three-electrode electrochemical cell and double compartment cell. Different cathodes as working electrode and anodes as counter electrode used in this process are analyzed. The effects of various operating parameters and their optimum ranges for maximum pollutant removal and mineralization are reviewed. Also various pollutants removed by this process are evaluated. Quick removal and mineralization of pollutants and their intermediate reaction products were reported.
Article
An efficient and convenient degradation of AR 97 (C.I. Acid Red 97) in aqueous solution by the combination of ultrasound and Fenton reagent was reported. The various affecting factors were studied. Under ultrasound, the degradation ratio is 99.9% for the initial concentration of 50 mg L−1 using 1.57 mmol L−1 H2O2, 0.054 mmol L−1 Fe2+ at 40°C within 140 min. The combination method was more effective than sonolysis or Fenton reagent process treatment individually. In the combined process, the effect of the kind of anions (Cl−, SO42−, and H2PO4−) and Fenton-like reagents (Cu2+/H2O2, Co2+/H2O2, and Mn2+/H2O2) on degradation was also observed. Except for SO42−, others had a negative effect on degradation. © 2009 American Institute of Chemical Engineers Environ Prog, 2010