You are on page 1of 8

Journal of Environmental Chemical Engineering 2 (2014) 18811888

Contents lists available at ScienceDirect

Journal of Environmental Chemical Engineering


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

Optimisation of graphene oxideiron oxide nanocomposite in


heterogeneous Fenton-like oxidation of Acid Orange 7
Nor Aida Zubir a,b , Christelle Yacou a , Xiwang Zhang c , Joo C. Diniz da Costa a, *
a
The University of Queensland, FIMLabFilms and Inorganic Membrane Laboratory, School of Chemical Engineering, Brisbane,Queensland 4072, Australia
b
Universiti Teknologi MARA (UiTM), Faculty of Chemical Engineering, Pulau Pinang 13500, Malaysia
c
Department of Chemical Engineering, Monash University, Clayton, Victoria 3800, Australia

A R T I C L E I N F O A B S T R A C T

Article history: Well-dispersed iron oxide nanoparticles supported onto graphene oxide sheets (i.e. GOFe3O4
Received 6 March 2014 nanocomposite) were synthesised and used as heterogeneous Fenton-like catalyst for the degradation
Received in revised form 10 June 2014 of Acid Orange 7 dye (AO7). The reaction was systematically investigated under various experimental
Accepted 1 August 2014
conditions such as nanocomposite dosage, pH, temperature, oxidant and dye concentrations. Best results
showed a fast 80% degradation in 20 min, whilst 98% of AO7 was successfully removed after 180 min of
Keywords: reaction time. The degradation kinetics of AO7 was most inuenced by pH and temperature, and can be
Graphene oxide
described by a pseudo-rst-order reaction following the LangmuirHinshelwood mechanism. Analysis of
Iron oxide
Nanocomposites
the spent nanocomposite suggested that the phase of iron oxide nanoparticles remained unchanged
Heterogeneous Fenton-like reaction whilst minor pore volume losses occurred via carbon deposition and/or re-stacking of GO sheets.
Acid Orange 7 2014 Elsevier Ltd. All rights reserved.

Introduction unattainable regeneration of catalyst which hampers the economic


feasibility of this process [6].
The disposal of organic pollutants into water resources is an To overcome these disadvantages, numerous attempts have
issue of environmental signicance, particularly due to the scarcity been made to develop heterogeneous catalysts including the
of potable water facing our contemporary society. One of the incorporation of iron ions or iron oxides within the structure of
industries under strong scrutiny is the textile industry, as previous catalyst support [710]. The nature and properties of support play a
practices of direct discharge of spent dyes into efuents had led to crucial role in the modulation of the activity in catalytic sites [11].
detrimental effects in both environment and human health due to The decomposition of hydrogen peroxide (H2O2) in heterogeneous
toxicity, biodegradability and aesthetic impacts [13]. Therefore, it Fenton-like reaction was suggested according to the Eqs. (1) and
is imperative that efcient textile wastewater treatment technol- (2) where  represents the iron species bound to the surface of
ogies are put in place to remove these toxic contaminants prior to catalyst support [12]:
discharge into natural water bodies. Oxidation using Fenton or
Fenton-like reaction has been proven as promising and attractive  Fe3 H2 O2 !  Fe2 HOO H (1)
treatment method for effective decolourisation and degradation of
textile dyes. This is attributed to the simplicity of the Fenton  Fe2 H2 O2 !  Fe3 HO OH (2)
reaction operation, together with being an environmentally benign Iron catalyst supports, such as carbon based materials
process which produces highly reactive hydroxyl radicals (HO) (activated carbon; AC) [9,13,14], carbon aerogel [9], ion exchange
that oxidise dyes efciently and non-selectively [4,5]. Nonetheless, resin [15], carbon nanotubes (CNTs) [10,1618], graphene and its
the usage of homogeneous Fenton or Fenton-like reaction derivative graphene oxide (GO) [19,20] have recently been
operation has several signicant disadvantages, such as post- reported. Among these works, carbon nanomaterials (CNTs and
treatment requirements prior to discharge as a result of iron GO) were reported to outperform the macroscale carbon systems
hydroxide sludge formation, narrow pH operational range and (AC) in sustaining stable and higher catalytic activity over a
number of cycles for H2O2 decomposition in aqueous phase [21].
Such improvement is inextricably linked to the morphology,
surface functionalities and chemistry of the carbon nanomaterials.
* Corresponding author. Tel.: +61 7 3365 6960; fax: +61 7 3365 4199. Therefore, immobilisation of iron compounds within the structure
E-mail address: j.dacosta@uq.edu.au (J.C. Diniz da Costa). of these carbon nanomaterials seem to be benecial in improving

http://dx.doi.org/10.1016/j.jece.2014.08.001
2213-3437/ 2014 Elsevier Ltd. All rights reserved.
1882 N.A. Zubir et al. / Journal of Environmental Chemical Engineering 2 (2014) 18811888

the catalytic activity of heterogeneous Fenton-like reaction. resulting black precipitate was magnetically separated and washed
Instead of CNTs, GO is considered to be more economic favourable several times with deionised water and ethanol, and dried at 333 K
as catalyst support since it can be conveniently synthesised by the for 48 h.
oxidation of graphite which is cheap and naturally abundant. Microstructural investigation was carried out by transmission
GO consists of a hexagonal ring-based carbon network with electron microscopy (TEM, JOEL 1010) operated at 100 kV. Samples
both sp2 and sp3 hybridised carbon atoms in two-dimensional were prepared by placing a drop of diluted sample dispersion in
structure that contain copious oxygen functional groups such as ethanol onto a carbon-coated copper grid and dried at room
epoxy, hydroxyl, carbonyl and carboxyl groups on surfaces and temperature. The XRD patterns of nanocomposite were obtained
edges [2225]. These oxygen functional groups serve as activation using X-ray diffraction by a Bruker D8 Advance diffractometer at
sites for nucleation and growth of iron precursors to form GO 40 kV and 40 mA using ltered Cu Ka radiation (l = 1.5418 ). The
supported iron-containing nanocomposites [26,27]. Recently, GO- textural properties of nanocomposite were determined by nitro-
Fe(III) complexes were used for processing Rhodamine B (RhB) dye gen sorption analysis using Tristar II 3020 (Micromeritics).
resulting in a 38% degradation in 120 min (heterogeneous) and
almost 100% within 45 min (photo-assisted) Fenton-like reactions Oxidation degradation and sorption of AO7
[19]. In a similar fashion, GOFe2O3 hybrid exhibited 17% higher
RhB degradation under the visible-light irradiation over the Detailed parametric studies were performed using of GOFe3O4
heterogeneous Fenton-like reactions [20]. These reports have nanocomposite in heterogeneous Fenton-like reaction by changing
highlighted the efcacy of using GO supported iron-containing one variable at a time, whilst keeping the others constant. The
nanocomposites in photo Fenton-like rather than heterogeneous oxidative degradation of AO7 was determined in the following
reaction. initial ranges of parameters: (i) nanocomposite dosage
This work focuses on using GO as a support for magnetite (0.10.3 g L1); (ii) pH (2.56.3); (iii) concentration of H2O2
(Fe3O4) nanoparticles for the heterogeneous Fenton-like reaction, (2.7544 mM); (iv) concentration of dye (1555 mg L1) and
which simplies equipment design by dispensing the need of lamp (v) temperature (298328 K). These testing conditions were
irradiation and extra process controls. The choice of magnetite chosen based on previous works carried out in the heterogeneous
over the other iron oxide phases is based on its highest Fenton Fenton reaction using iron oxide supported based catalysts
activity [28,29]. Exfoliated GO sheets were used to anchor Fe3O4 [9,12,13]. The reaction was initiated by adding H2O2 into the
nanoparticles, thus forming a GOFe3O4 nanocomposite catalyst suspension and stirred at 350 rpm after 30 min of dark adsorption.
[30]. Although GOFe3O4 nanocomposite [31], where the GO was Sampling was carried out periodically. The collected suspension
prepared by the Hummers method [32], were tested for the was ltered through 0.2 mm Millipore syringe lters and the
degradation of Acid Orange 7 (AO7), these works focused mainly on ltrates were then immediately analysed. Concentration of
materials development rather than on catalytic reaction. In other AO7 was determined by measuring the absorbance of the solution
words, there is limited understanding of the catalytic reaction at 484 nm, corresponding to the maximum absorption wavelength
effect of GOFe3O4 nanocomposites on the heterogeneous Fenton- of AO7 in visible region by Evolution 220 UVvis spectrophotom-
like oxidation of AO7, which warrants further research. Therefore, eter (Thermo, Fisher Sci.). The experiments had a reproducible
this work investigates the most inuential operational parameters with an average experimental variation of 7%. The total carbon
(i.e. dosage, pH, concentration of H2O2, concentration of dye and (TC) analysis was carried out using Shimadzu 5000 TOC-VCSH
temperature) on the oxidative degradation of AO7 using GOFe3O4 analyser (Shimadzu), equipped with an automatic sample injector.
nanocomposite, supported by a global parametric optimisation. The TC values represent the average of triplicate measurements.
Further, the reaction kinetics is studied and a model is validated In order to verify if adsorption was concurrently taking place to
following the LangmuirHinshelwood mechanism to explain the the heterogeneous Fenton reaction as reported elsewhere [34,35],
rate of intrinsic chemical reaction on the surface of active sites on adsorption analysis was carried out by mixing an amount of
the nanocomposites. Finally, the characterisation of the spent nanocomposite (0.2 g L1) in a conical ask containing different
catalyst samples was performed by nitrogen sorption and initial concentration of AO7 ranging from 15 to 55 mg L1 and
spectroscopy techniques to verify the stability of the GOFe3O4 magnetically stirred at 350 rpm. For equilibrium adsorption
nanocomposite prior to and after the catalytic degradation of studies, the experiments were carried out at pH 3 and 298 K for
AO7 testing. 24 h to ensure equilibrium was reached. The concentration of
AO7 in the supernatant was analysed consecutively. The AO7
Experimental uptake at equilibrium, qe (mg g1) was calculated according to
Eq. (3).
Materials synthesis and characterisation
C 0  C e V
qe (3)
W
Graphite akes, FeCl36H2O (97%), FeCl24H2O (99%), Na2SO3
(98%), H2O2 (30% w/w) and AO7 (Orange II; 85%) were purchased where C0 and Ce are the solution concentrations at initial and
from SigmaAldrich. Other reagents and solvents used were of equilibrium (mg L1), respectively. V is the volume of solution (L)
analytical grade and without further purication. The GOFe3O4 and W is the mass of the nanocomposite (g).
nanocomposites were synthesised by co-precipitating of pre-
hydrolysed ferric and ferrous salts in the presence of GO as Results and discussion
described elsewhere [31]. Briey, GO was rst prepared via a
modied Hummers method [32,33] and subsequently exfoliated Optimisation analysis
by ultrasonication to attain an aqueous dispersion of GO. Then
NaOH (1 M) was added dropwise into 100 mL of precursor solution GOFe3O4 nanocomposite containing 5 wt% of GO loading was
containing of FeCl36H2O (4 mmol) and FeCl24H2O (2 mmol) until used in this study due to their optimal catalytic activity [31]. Fig. 1
pH 4 under constant stirring. GO solution (50 mL, 0.55 mg mL1) presents the degradation prole of AO7 at different nanocomposite
was then gradually added into a mixture and stirred for another dosage ranging from 0.10.3 g L1. These results demonstrated that
30 min. NaOH (1 M) was continuously added dropwise into the the degradation prole of AO7 increased signicantly as a function
mixture until reaching pH 10 and then aged for 30 min. The of the dosage from 0.1 to 0.2 g L1. Such enhancement in
N.A. Zubir et al. / Journal of Environmental Chemical Engineering 2 (2014) 18811888 1883


H2 O2 H ! H3 O2 (5)

HO H e ! H2 O (6)
Oxidant concentration is also regarded as a key operational
parameter in producing HO radicals in oxidative degradation of
dye. Fig. 3 depicts the relationship between AO7 degradation at
different oxidant (H2O2) concentrations. As the H2O2 concentration
increased from 2.75 to 22 mM, the degradation of AO7 improved
from 75% to 98% over 180 min, respectively. These results suggest
that more HO radicals are formed as H2O2 concentration
increased, which leads to an increased rate in AO7 degradation.
However, when the H2O2 concentration is over 22 mM, for example
44 mM; the degradation rate slowed down and dropped to 90%.
The excess concentration of H2O2 may induce parallel undesirable
reactions between HO radicals with H2O2 and producing hydro-
Fig. 1. Effect of GOFe3O4 nanocomposite dosage on the degradation of AO7. peroxyl radicals (HOO), as expressed in Eq. (7). The oxidation
Experimental conditions: AO7 = 35 mg L1; H2O2 = 22 mM; pH = 3 and 298 K. potential of produced HOO radicals (1.78 V) is much lower so that
they are less reactive to oxidise dyes compared with HO radicals
degradation was ascribed to the increased number of active (2.80 V) [38].
catalytic sites which are responsible for the decomposition of H2O2
into HO radicals. Nevertheless, further increase of dosage H2 O2 HO ! H2 O HOO (7)
(0.3 g L1) inhibited the degradation of AO7. This result is Fig. 4a illustrates the dependence of AO7 degradation on the
attributed to possible scavenging of HO radicals (Eq. (4)) by the initial concentration of dye ranging from 15 to 55 mg L1. The initial
reaction between the excess active sites and the HO radicals, concentration affected the reaction which progressively slowed
therefore induced a decline of AO7 degradation [13,36]. Dosages of down as a function of AO7 concentration. For instance, by
0.20 and 0.25 g L1 delivered comparable degradation prole. increasing the initial AO7 concentration, 80% of
However, lower nanocomposite dosage is always preferred in AO7 degradation was achieved in 30 min (15 mg L1), and doubled
catalysis and thus 0.2 g L1 was chosen for all subsequent to 60 min (25 mg L1) and then increased by more than six fold to
experiments. 180 min (55 mg L1). This is attributed to inductive effects caused
by the high AO7 concentration, as available active sites were
 Fe2 HO !  Fe3 OH (4)
predominantly covered by the dye molecules. As a result, this effect
It is known that pH has a signicant impact in heterogeneous limits the amount of H2O2 freely available to interact with the
Fenton-like reaction, which directly affects the oxidative degrada- active sites of the GOFe3O4 nanocomposite for effective genera-
tion of dyes by HO radicals. This is clearly shown in Fig. 2, as the tion of HO radicals. Similar behaviour was also reported elsewhere
optimal AO7 degradation value of 98% was achieved at pH 3 within for the oxidation of dyes [39,40].
180 min reaction. When the pH was lowered to 2.5, the The prole of normalised concentrations with time (Fig. 4a)
AO7 degradation decreased to nearly 80%. This result suggests follows an exponential pattern, which may describe the
that the protonation of H2O2 in forming an oxonium ion (H3O2+) degradation of AO7 according to the pseudo-rst-order kinetics
had occurred, which then enhances H2O2 stability and reduces its reaction:
reactivity towards Fe2+ (Eq. (5)) [37]. Besides that, the scavenging
dC AO7
effect of HO radicals by the excessive H+ (Eq. (6)) [11] is also kobs C AO7 (8)
dt
enhanced at pH 2.5. The reduction in AO7 degradation was also
1
observed when the initial pH was higher than 3. This can be where the CAO7 represents the concentration of AO7 (mg L ) at
explained by decomposition of more H2O2 into molecular oxygen time t and kobs is the pseudo rst-order rate constant (min1)
and water on the active sites without the formation of appreciable for the degradation of AO7. The integration of Eq. (8) yields
amounts of HO radicals [38]. Eq. (9).

Fig. 2. Effect of initial pH on the degradation of AO7. Experimental conditions: Fig. 3. Effect of H2O2 concentration on the degradation of AO7. Experimental
AO7 = 35 mg L1; Dosage = 0.2 g L1; H2O2 = 22 mM and 298 K. conditions: AO7 = 35 mg L1; Dosage = 0.2 g L1; pH = 3 and 298 K.
1884 N.A. Zubir et al. / Journal of Environmental Chemical Engineering 2 (2014) 18811888

calculated as 0.710 mg L1 min1 and 0.745 L mg1, respectively.


The KAO7 value was found to be lower than the sorption constant
(KL = 1.109 L mg1) in absence of oxidant according to the Langmuir
adsorption model as shown in Fig. S1 and Table S2 (see
Supplementary information), indicating the existence of competi-
tive adsorption between AO7 and H2O2 towards the active sites
during the reaction.
The temperature has a pronounced effect on AO7 degradation
as shown in Fig. 5a. Fast rates are observed within minutes for an
AO7 concentration of 35 mg L1 (expected in textile wastewater),
and 80% degradation was reached less than 20 min at 328 K.
Longer reaction time of 70 min is required as the temperature is
reduced to 298 K. Hence, raising the temperature of the reaction
likewise increases kobs, translating in shorter reaction time. This
phenomenon is attributed to the exponential dependency of kobs
with reaction temperature in accordance with the Arrheniuss law.
High temperature leads to an increased rate between H2O2 and the
active sites of GOFe3O4 nanocomposite in generating more HO
radicals to readily oxidised AO7. These results suggest that the
mobility of AO7 and H2O2 towards the surface of active sites are
promoted by increasing reaction temperature. However, raising
the temperature from ambient conditions at 298328 K requires
energy, which in turn affects the operating costs of wastewater
treatment. Unless a source of waste heat is available, or solar heat
can be used effectively, 298 K was chosen in this work as the
preferred temperature in the heterogeneous Fenton-like reaction
due to the consideration of the potential operating costs.

Fig. 4. (a) Effect of dye initial concentration on the degradation of AO7. (b)
Relationship between the 1/kobs and CAO70 using LangmuirHinshelwood model.
Experimental conditions: Dosage = 0.2 g L1; H2O2 = 22 mM; pH = 3 and 298 K.

C AO70
ln kobs t (9)
C AO7
The slope of the straight line for a plot of ln(CAO70/CAO7) vs. t
determines the kobs value which are summarised in Table S1 (see
Supplementary information). The results in this work shown that
the kobs values decreased as the initial concentration increased.
The heterogeneous reaction for the degradation of organics
encompasses with both reaction and adsorption that occurs
simultaneously [41], and can be described by the Langmuir
Hinshelwood mechanism [39,42], which can be expressed as
follows:
kc K AO7
kobs (10)
1 K AO7 C AO70

1 1  1
C AO70 (11)
kobs kc kc K AO7
where the kc is the intrinsic surface reaction rate constant
(mg L1 min1) and KAO7 is the LangmuirHinshelwood adsorption
equilibrium constant (L mg1). The values of kc and KAO7 can be
calculated by plotting the (1/ kobs) against CAO70. As can be seen in
Fig. 4b, the experimental data ts reasonably well (R2 = 0.983) with
the proposed LangmuirHinshelwood mechanism. In other words,
the catalytic surface reactions between the nanocomposite active Fig. 5. (a) Effect of temperature on AO7 degradation. (b) Arrhenius plot of reaction
sites, AO7 and HO radicals can be used in determining the kinetics rate constant. Experimental conditions: AO7 = 35 mg L1; Dosage = 0.2 g L1; H2O2 =
of the heterogeneous reaction. The kc and KAO7 values were 22 mM and pH 3.
N.A. Zubir et al. / Journal of Environmental Chemical Engineering 2 (2014) 18811888 1885

The kobs parameter for the AO7 degradation for each tempera- (ANOVA) of response surface regression, an empirical relationship
ture was calculated to determine the activation energy of reaction. between the degradation of AO7 and the coded parameters was
It was found that the kobs values at temperatures of 298, 313 and expressed by the following equation:
328 K were 0.022, 0.032 and 0.053 min1, respectively. The
Degradation of AO7% 95:89 2:82A  6:25B 3:66C
correlation between temperature and kobs can be presented in
0:38AB 2:29AC 3:15BC
Fig. 5b using the linear form of Arrhenius equation:
   1:86A2  19:44B2 0:36C 2 (13)
Ea
ln kobs ln A (12) This derived model is graphically represented through the
RT
contour plot (Fig. 6a). This plot illustrates the interactive effects
where A is a constant, R is the gas constant (8.314 J mol1 K1), T is between the pH and temperature at optimal dosage of 0.205 g L1
the solution temperature and Ea is the activation energy. The Ea was on the degradation of AO7. At the same time, determination of the
found to be 24.3 kJ mol1. This value is higher than Ea of optimal condition between those parameters can be facilitated by
the diffusion-controlled reaction that usually ranges within the elliptical nature of the contour plots. Validation of derived
1013 kJ mol1[43], which implies that the kobs for this heteroge- model was conrmed by diagnostic plot (Fig. 6b), which shows
neous reaction is dominated by the rate of intrinsic chemical high correlation between the experimental and predicted values.
reactions on the surface of active sites rather than the rate of mass The model adequacy was corroborated by the high coefcient
transfer as reported elsewhere [39,44]. value of 0.9742, hence given condence to the global optimised
results.
Global optimisation From this analysis, the optimal condition for degradation of
AO7 was obtained at dosage of 0.205 g L1, pH of 3 and temperature
The optimisation carried out so far follows one dimensional of 326.7 K with desirability of 1. In one hand, raising the
pattern for a single condition. Hence, there is a need to understand temperature is benecial in terms of increasing the reaction
the best global optimised condition. To address this point, a kinetics. On the other hand, the increase in temperature must be
23-factorial design was used to elucidate the global interaction traded off against energy input to the reaction which attracts extra
based on a face centred central composite design (FCCD) under undesirable operational cost. Unless free waste heat is available,
response surface methodology (RSM). Three signicant param- we have considered 298 K as the preferred temperature in the
eters (i.e. dosage, pH and temperature) and their ranges were heterogeneous Fenton-like reaction. Based on these ndings, there
chosen based on the ndings of one-parameter-at-a-time method.
The details of the design matrix and the response surface
regression analysis are presented in Tables S3 and S4, respectively
(Supplementary information). From the analysis of variance

Fig. 6. (a) Contour plot of the derived response surface quadratic model.
(b) Correlation between the actual and predicted degradation of AO7. Fig. 7. TEM images of pristine (a) and spent (b) GOFe3O4 nanocomposites.
1886 N.A. Zubir et al. / Journal of Environmental Chemical Engineering 2 (2014) 18811888

Fig. 8. XRD patterns of Fe3O4 nanoparticles, pristine and spent GOFe3O4 Fig. 10. UVvis spectra changes during the oxidative degradation of AO7 by GO
nanocomposites. Fe3O4 nanocomposite in the presence of H2O2. Experimental conditions: AO7 = 35
mg L1; Dosage = 0.2 g L1; H2O2 = 22 mM; pH = 3 and 298 K.

are no signicant differences (except for temperature) on the


which facilitates the nucleation and growth of the Fe3O4 nano-
optimal condition between the one-parameter-at-a-time and the
particles [45]. However, there is a thick area on the spent
global optimisation.
nanocomposite in Fig. 7b (highlighted white circle) possibly
originating from disorder stacking of GOFe3O4 nanocomposite.
Characterisation
XRD patterns in Fig. 8 show that both pristine and spent GOFe3O4
nanocomposites exhibited relatively weak and broad diffraction
The analysis of spent GOFe3O4 nanocomposite was performed
peaks similar to Fe3O4 nanoparticles. Interestingly, the phase of
to compare against its pristine properties. TEM images in Fig. 7
Fe3O4 remained unchanged throughout the AO7 oxidation reaction
indicate that there are no signicant differences in the morphology
because no discernible changes in diffraction peaks of spent
between both samples. The Fe3O4 nanoparticles were well
catalyst were observed.
dispersed throughout the surface of GO sheets in both cases.
Minor changes were observed between the nitrogen isotherms
There is a slight aggregation within the corrugated GO surfaces in
of spent and pristine GOFe3O4 nanocomposites as displayed in
both cases. This is attributed to the surface defects and edges of GO
Fig. 9a. Both isotherms are of type IV with H2 hysteresis loop,
indicative of mesoporous structures. The spent sample resulted in
surface area loss of 19.7% (from 373 to 300 m2 g1) and pore
volume loss of 15.6% (from 0.328 to 0.277 cm3 g1). These results
suggest that the spent nanocomposite underwent structural
densication, though the pore size distribution remained similar
to the pristine catalyst (Fig. 9b). The spent nanocomposite was
washed, dried and degassed prior to the nitrogen adsorption
testing. If the intermediates of the AO7 degradation through
carbon deposition remained adsorbed, then this could explain the
loss in both the surface area and volume of the spent nano-
composite. Alternatively, there may have been re-stacking of the
GO sheets during successive catalytic reaction, washing and drying
processes, thus reecting the observed changes.
Finally, the total carbon (TC) analysis showed that 90.48% of
carbon was successfully being removed and decomposed into
carbon dioxide after undergoing 180 min of reaction. This can be
clearly seen in Fig. 10 as the degradation of AO7 was monitored by
UVvis spectra. There are three absorbance peaks at 484, 310, and
230 nm and one shoulder at 430 nm. The peaks at 484 and 430 nm
in the visible region are ascribed to the hydrazone and azo form,
whilst the other two peaks, at 310 and 230 nm in the ultraviolet
region, are assigned to the naphthalene and benzene rings of AO7,
respectively [6]. The absorbance at 310 and 230 nm increased
slightly in the rst 30 min, attributed to the formation of
naphthalene- and/or benzene-type intermediates during catalysis.
All peaks almost disappeared at 180 min of reaction due to the
destruction of azo bond ( NQN ) in the chromophoric
structure of AO7.

Conclusions

The degradation of AO7 dye by the heterogeneous Fenton-like


Fig. 9. Isotherms (a) and pore size distribution (b) of pristine and spent GOFe3O4 reaction was systematically investigated using GOFe3O4 nano-
nanocomposites. composite catalyst. The optimal operational condition was
N.A. Zubir et al. / Journal of Environmental Chemical Engineering 2 (2014) 18811888 1887

obtained with a catalyst concentration of 0.2 g L1, an initial pH of [11] S. Navalon, A. Dhakshinamoorthy, M. Alvaro, H. Garcia, Heterogeneous Fenton
3 and 22 mM of H2O2 concentration at 298 K. The AO7 degradation catalysts based on activated carbon and related materials, ChemSusChem 4
(2011) 17121730, doi:http://dx.doi.org/10.1002/cssc.201100216. 22162405.
was mainly ascribed to heterogeneous Fenton-like reaction and [12] H. Hassan, B.H. Hameed, Feclay as effective heterogeneous Fenton catalyst for
well described by a pseudo-rst-order kinetic through the the decolorization of reactive Blue 4, Chem. Eng. J. 171 (2011) 912918, doi:
LangmuirHinshelwood mechanism. This heterogeneous reaction http://dx.doi.org/10.1016/j.cej.2011.04.040.
[13] T.D. Nguyen, N.H. Phan, M.H. Do, K.T. Ngo, Magnetic Fe2MO4 (M:Fe, Mn)
is dominated by the rate of intrinsic chemical reactions on the activated carbons: fabrication, characterization and heterogeneous Fenton
surface of active sites rather than the rate of mass transfer. The GO oxidation of methyl orange, J. Hazard. Mater. 185 (2011) 653661, doi:http://
Fe3O4 nanocomposite delivered best AO7 degradation of 98%. The dx.doi.org/10.1016/j.jhazmat.2010.09.068. 20952129.
[14] A. Rodriguez, G. Ovejero, J.L. Sotelo, M. Mestanza, J. Garcia, Heterogeneous
spent catalysts was mainly stable, as the Fe3O4 phase remained Fenton catalyst supports screening for mono azo dye degradation in
similar to a pristine catalyst, though minor changes in surface area contaminated wastewaters, Ind. Eng. Chem. Res. 49 (2010) 498505, doi:
and pore volume were observed possibly attributed to the http://dx.doi.org/10.1021/ie901212m.
[15] Z. Yaping, H. Jiangyong, Photo-Fenton degradation of 17b-estradiol in presence
adsorption of intermediates or re-stacking of the GOFe3O4 sheets
of a-FeOOHR and H2O2, Appl. Catal. B Environ. 78 (2008) 250258, doi:http://
during the reaction. These ndings present new insights into the dx.doi.org/10.1016/j.apcatb.2007.09.026.
inuence of operational parameters in the heterogeneous Fenton- [16] J. Deng, X. Wen, Q. Wang, Solvothermal in situ synthesis of Fe3O4-multi-walled
like oxidation of AO7 using GOFe3O4 nanocomposite. carbon nanotubes with enhanced heterogeneous Fenton-like activity, Mater.
Res. Bull. 47 (2012) 33693376, doi:http://dx.doi.org/10.1016/j.materres-
bull.2012.07.021.
Acknowledgments [17] S. Song, R. Rao, H. Yang, H. Liu, A. Zhang, Facile synthesis of Fe3O4/MWCNTs by
spontaneous redox and their catalytic performance, Nanotechnology 21 (2010)
The authors acknowledge the assistance of Dr Julius Motuzas, 185602185607, doi:http://dx.doi.org/10.1088/0957-4484/21/18/185602.
[18] S. Song, H. Yang, R. Rao, H. Liu, A. Zhang, High catalytic activity and selectivity
and the facilities and the scientic and technical assistance of the for hydroxylation of benzene to phenol over multi-walled carbon nanotubes
Australian Microscopy & Microanalysis Research Facility at the supported Fe3O4 catalyst, Appl. Catal. A Gen. 375 (2010) 265271, doi:http://
Centre for Microscopy and Microanalysis, The University of dx.doi.org/10.1016/j.apcata.2010.01.008.
[19] Y. Dong, J. Li, L. Shi, J. Xu, X. Wang, Z. Guo, W. Liu, Graphene oxide-iron
Queensland. Nor Aida Zubir gratefully acknowledges the generous complex: synthesis, characterization and visible-light-driven photocatalysis, J.
nancial support from Ministry of Higher Education Malaysia Mater. Chem. A 1 (2013) 644650, doi:http://dx.doi.org/10.1039/c2ta00371f.
(MOHE) and Universiti Teknologi MARA (UiTM) for her study leave. [20] S. Guo, G. Zhang, Y. Guo, J.C. Yu, Graphene oxideFe2O3 hybrid material as
highly efcient heterogeneous catalyst for degradation of organic contami-
X. Zhang thanks the Australia Research Council for the support on nants, Carbon 60 (2013) 437444, doi:http://dx.doi.org/10.1016/j.car-
Discovery Project (DP110103533) and his Australian Research bon.2013.04.058.
Fellowship. J. C. Diniz da Costa acknowledges support via the [21] K.V. Voitko, R.L.D. Whitby, V.M. Gunko, O.M. Bakalinska, M.T. Kartel, K. Laszlo,
A.B. Cundy, S.V. Mikhalovsky, Morphological and chemical features of nano
Australian Research Council Future Fellowship Program
and macroscale carbons affecting hydrogen peroxide decomposition in
(FT130100405). aqueous media, J. Coll. Interf. Sci. 361 (2011) 129136, doi:http://dx.doi.org/
10.1016/j.jcis.2011.05.048. 21676406.
Appendix A. Supplementary data [22] D.R. Dreyer, S. Park, C.W. Bielawski, R.S. Ruoff, The chemistry of graphene
oxide, Chem. Soc. Rev. 39 (2010) 228240, doi:http://dx.doi.org/10.1039/
b917103g. 20023850.
Supplementary data associated with this article can be found, in [23] C. Su, K.P. Loh, Carbocatalysts: graphene oxide and its derivatives, Acc. Chem.
the online version, at http://dx.doi.org/10.1016/j. Res. 46 (2013) 22752285.
[24] Y. Matsumoto, M. Koinuma, S. Ida, S. Hayami, T. Taniguchi, K. Hatakeyama, H.
endend.2009.01.011.
Tateishi, Y. Watanabe, S. Amano, Photoreaction of graphene oxide nanosheets
in water, J. Phys. Chem. C 115 (2011) 1928019286, doi:http://dx.doi.org/
References 10.1021/jp206348s.
[25] A. Lerf, H. He, M. Forster, J. Klinowski, Structure of graphite oxide revisited, J.
[1] V.M. Correia, T. Stephenson, S.J. Judd, Characterisation of textile wastewaters Phys. Chem. B102 (1998) 44774482.
a review, Environ. Technol. 15 (1994) 917929, doi:http://dx.doi.org/10.1080/ [26] M.Z. Kassaee, E. Motamedi, M. Majdi, Magnetic Fe3O4-graphene oxide/
09593339409385500. polystyrene: fabrication and characterization of a promising nanocomposite,
[2] M. Aleksi c, H. Kui c, N. Koprivanac, D. Leszczynska, A.L. Boic, Heterogeneous Chem. Eng. J. 172 (2011) 540549, doi:http://dx.doi.org/10.1016/j.
Fenton type processes for the degradation of organic dye pollutant in water cej.2011.05.093.
the application of zeolite assisted AOPs, Desalination 257 (2010) 2229, doi: [27] C. Xu, X. Wang, Graphene oxide-mediated synthesis of stable metal
http://dx.doi.org/10.1016/j.desal.2010.03.016. nanoparticle colloids, Coll. Surf. A Physicochem. Eng. Asp. 404 (2012) 78
[3] J. Labanda, J. Sabat, J. Llorens, Modeling of the dynamic adsorption of an 82, doi:http://dx.doi.org/10.1016/j.colsurfa.2012.04.017.
anionic dye through ion-exchange membrane adsorber, J. Membr. Sci. 340 [28] R. Matta, K. Hanna, S. Chiron, Fenton-like oxidation of 2,4,6-trinitrotoluene
(2009) 234240, doi:http://dx.doi.org/10.1016/j.memsci.2009.05.036. using different iron minerals, Sci. Total Environ. 385 (2007) 242251, doi:
[4] S.P. Sun, C.J. Li, J.H. Sun, S.H. Shi, M.H. Fan, Q. Zhou, Decolorization of an azo dye http://dx.doi.org/10.1016/j.scitotenv.2007.06.030. 17662375.
orange G in aqueous solution by Fenton oxidation process: effect of system [29] R.C.C. Costa, M.F.F. Lelis, L.C.A. Oliveira, J.D. Fabris, J.D. Ardisson, R.R.V.A. Rios, C.
parameters and kinetic study, J. Hazard. Mater. 161 (2009) 10521057, doi: N. Silva, R.M. Lago, Novel active heterogeneous Fenton system based on
http://dx.doi.org/10.1016/j.jhazmat.2008.04.080. 18538927. Fe3xMxO4 (Fe, Co, Mn, Ni): the role of M2+ species on the reactivity towards
[5] J.H. Ramirez, F.M. Duarte, F.G. Martins, C.A. Costa, L.M. Madeira, Modelling of H2O2 reactions, J. Hazard. Mater. 129 (2006) 171178, doi:http://dx.doi.org/
the synthetic dye Orange II degradation using Fentons reagent: from batch to 10.1016/j.jhazmat.2005.08.028. 16298475.
continuous reactor operation, Chem. Eng. J. 148 (2009) 394404, doi:http://dx. [30] N.A. Zubir, X. Zhang, C. Yacou, J.C. Diniz da Costa, Fenton-like degradation of
doi.org/10.1016/j.cej.2008.09.012. acid Orange 7 using graphene oxide-iron oxide nanocomposite, Sci. Adv.
[6] S.H. Tian, Y.T. Tu, D.S. Chen, X. Chen, Y. Xiong, Degradation of Acid Orange II at Mater. 6 (2014) 13821388, doi:http://dx.doi.org/10.1166/sam.2014.1812.
neutral pH using Fe2(MoO4)3 as a heterogeneous Fenton-like catalyst, Chem. [31] N.A. Zubir, C. Yacou, J. Motuzas, X. Zhang, J.C. Diniz da Costa, Structural and
Eng. J. 169 (2011) 3137, doi:http://dx.doi.org/10.1016/j.cej.2011.02.045. functional investigation of graphene oxideFe3O4 nanocomposites for the
[7] N.K. Daud, M.A. Ahmad, B.H. Hameed, Decolorization of acid Red 1 dye solution heterogeneous Fenton-like reaction, Sci. Rep. 4 (2014) 4594, doi:http://dx.doi.
by Fenton-like process using Femontmorillonite K10 catalyst, Chem. Eng. J. org/10.1038/srep04594. 24699690.
165 (2010) 111116, doi:http://dx.doi.org/10.1016/j.cej.2010.08.072. [32] W.S. Hummers, R.E. Offeman, Preparation of graphitic oxide, J. Am. Chem. Soc.
[8] P.P. Gan, S.F.Y. Li, Efcient removal of rhodamine B using a rice hull-based silica 80 (1958) 1339, doi:http://dx.doi.org/10.1021/ja01539a017.
supported iron catalyst by Fenton-like process, Chem. Eng. J. 229 (2013) 351 [33] Z. Xiong, L.L. Zhang, X.S. Zhao, Visible-light-induced dye degradation over
363, doi:http://dx.doi.org/10.1016/j.cej.2013.06.020. copper-modied reduced graphene oxide, Chem. (Weinheim an Der Berg-
[9] J.H. Ramirez, F.J. Maldonado-Hdar, A.F. Prez-Cadenas, C. Moreno-Castilla, C. strasse, Germany) 17 (2011) 24282434, doi:http://dx.doi.org/10.1002/
A. Costa, L.M. Madeira, Azo-dye Orange II degradation by heterogeneous chem.201002906. 21319236.
Fenton-like reaction using carbon-Fe catalysts, Appl. Catal. B Environ. 75 [34] F. Duarte, F.J. Maldonado-Hdar, L.M. Madeira, New insight about orange II
(2007) 312323, doi:http://dx.doi.org/10.1016/j.apcatb.2007.05.003. elimination by characterization of spent activated carbon/Fe Fenton-like
[10] X. Hu, B. Liu, Y. Deng, H. Chen, S. Luo, C. Sun, P. Yang, S. Yang, Adsorption and catalysts, Appl. Catal. B Environ. 129 (2013) 264272, doi:http://dx.doi.org/
heterogeneous Fenton degradation of 17a-methyltestosterone on nano Fe3O4/ 10.1016/j.apcatb.2012.09.037.
MWCNTs in aqueous solution, Appl. Catal. B Environ. 107 (2011) 274283, doi: [35] R. Gonzalez-Olmos, U. Roland, H. Toufar, F.D. Kopinke, A. Georgi, Fe-zeolites as
http://dx.doi.org/10.1016/j.apcatb.2011.07.025. catalysts for chemical oxidation of MTBE in water with H2O2, Appl. Catal. B
1888 N.A. Zubir et al. / Journal of Environmental Chemical Engineering 2 (2014) 18811888

Environ. 89 (2009) 356364, doi:http://dx.doi.org/10.1016/j. peroxide oxidation: kinetic study with the Fermis equation, Appl. Catal. B
apcatb.2008.12.014. Environ. 101 (2011) 197205, doi:http://dx.doi.org/10.1016/j.
[36] J. Herney-Ramirez, M. Lampinen, M.A. Vicente, C.A. Costa, L.M. Madeira, apcatb.2010.09.020.
Experimental design to optimize the oxidation of Orange II dye solution using [41] W. Xu, J.S. Kong, Y.-T.E. Yeh, P. Chen, Single-molecule nanocatalysis reveals
a clay-based Fenton-like catalyst, Ind. Eng. Chem. Res. 47 (2007) 284294. heterogeneous reaction pathways and catalytic dynamics, Nat. Mater. 7 (2008)
[37] N.K. Daud, B.H. Hameed, Decolorization of acid Red 1 by Fenton-like process 992996, doi:http://dx.doi.org/10.1038/nmat2319. 18997774.
using rice husk ash-based catalyst, J. Hazard. Mater. 176 (2010) 938944, doi: [42] N. Daneshvar, M.H. Rasoulifard, A.R. Khataee, F. Hosseinzadeh, Removal of C.I.
http://dx.doi.org/10.1016/j.jhazmat.2009.11.130. 20042285. Acid Orange 7 from aqueous solution by UV irradiation in the presence of ZnO
[38] J.H. Ramirez, C.A. Costa, L.M. Madeira, G. Mata, M.A. Vicente, M.L. Rojas- nanopowder, J. Hazard. Mater. 143 (2007) 95101, doi:http://dx.doi.org/
Cervantes, A.J. Lpez-Peinado, R.M. Martn-Aranda, Fenton-like oxidation of 10.1016/j.jhazmat.2006.08.072. 17030415.
Orange II solutions using heterogeneous catalysts based on saponite clay, Appl. [43] W. Stumm, J.J. Morgan, Aquatic Chemistry: Chemical Equilibria and Rates in
Catal. B Environ. 71 (2007) 4456, doi:http://dx.doi.org/10.1016/j. Natural Waters, third edition, Wiley Interscience, New York, 1996.
apcatb.2006.08.012. [44] S.-S. Lin, M.D. Gurol, Catalytic decomposition of hydrogen peroxide on iron
[39] X. Xue, K. Hanna, M. Abdelmoula, N. Deng, Adsorption and oxidation of PCP on oxide: kinetics, mechanism, and implications, Environ. Sci. Technol. 32 (1998)
the surface of magnetite: kinetic experiments and spectroscopic investiga- 14171423, doi:http://dx.doi.org/10.1021/es970648k.
tions, Appl. Catal. B Environ. 89 (2009) 432440, doi:http://dx.doi.org/10.1016/ [45] Z.G. Geng, Y. Lin, X.X. Yu, Q.H. Shen, L. Ma, Z.Y. Li, N. Pan, X.P. Wang, Highly
j.apcatb.2008.12.024. efcient dye adsorption and removal: a functional hybrid of reduced graphene
[40] J. Herney-Ramirez, A.M.T. Silva, M.A. Vicente, C.A. Costa, L.M. Madeira, oxide-Fe3O4 nanoparticles as an easily regenerative adsorbent, J. Mater. Chem.
Degradation of acid Orange 7 using a saponite-based catalyst in wet hydrogen 22 (2012) 35273535, doi:http://dx.doi.org/10.1039/c2jm15544c.

You might also like