You are on page 1of 11

Journal of Environmental Management 156 (2015) 225e235

Contents lists available at ScienceDirect

Journal of Environmental Management


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

Hydrotalcite-TiO2 magnetic iron oxide intercalated with the anionic


surfactant dodecylsulfate in the photocatalytic degradation of
methylene blue dye
Liany D.L. Miranda a, Carlos R. Bellato a, *, Jaderson L. Milagres a, Luciano G. Moura b,
Ann H. Mounteer c, Marciano F. de Almeida a
a
b
c

Departamento de Qumica, Universidade Federal de Viosa, Av. PH Holfs, s/n, 36571-000 Viosa, Minas Gerais, Brazil
Departamento de Fsica, Universidade Federal de Viosa, Av. PH Holfs, s/n, 36571-000 Viosa, Minas Gerais, Brazil
Departamento de Engenharia Civil, Universidade Federal de Viosa, Av. PH Holfs, s/n, 36571-000 Viosa, Minas Gerais, Brazil

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 14 November 2014
Received in revised form
15 March 2015
Accepted 29 March 2015
Available online

The new magnetic photocatalysts HT/TiO2/Fe and HT-DS/TiO2/Fe, modied with the anionic surfactant
sodium dodecylsulfate (DS) were successfully synthesized in this work. Titanium dioxide (anatase) followed by iron oxide were deposited on the hydrotalcite support. Several catalyst samples were prepared
with different amounts of titanium and iron. The photocatalysts were characterized by infrared and
Raman spectroscopy, X-ray diffraction, scanning electron microscopy. Photocatalytic performance was
analyzed by UVevisible radiation (lter cutoff, l > 300 nm) of an aqueous solution (24 mg/L) of
methylene blue (MB). The most efcient catalyst was obtained at an iron oxide:TiO2 molar ratio of 2:3.
This catalyst showed high photocatalytic activity, removing 96% of the color and 61% of total organic
carbon from the MB solution after 120 min. It was easily removed from solution after use because of its
magnetic properties. The reuse of the HT-DS/TiO2/Fe23 catalyst was viable and the catalyst was structurally stable for at least four consecutive photocatalytic cycles.
2015 Elsevier Ltd. All rights reserved.

Keywords:
Layered double hydroxides
TiO2
Adsorption
Photocatalysis

1. Introduction
Synthetic dyes are widely used in the textile, food, leather, paper
and plastics industries that often end up in industrial wastewaters.
Dyes are potentially harmful to the environment when discharged
into surface waters without adequate treatment because of their
chemical stability and toxicity (Wang et al., 2014). Therefore,
several methods, including chemical oxidation (Ferrarese et al.,
2008), microbial degradation (Haritash and Kaushik, 2009),
adsorption (Walcarius and Mercier, 2010), occulation (Golob et al.,
2005) and photocatalytic degradation (Tong et al., 2012) have been
developed to eliminate these organic pollutants from wastewaters.
Photocatalytic degradation, in which decomposition or chemical
dissociation is caused by exposure of a substance to visible or ultraviolet light, is considered an especially promising technique
(Chala et al., 2014) (C.-Y. Chen et al., 2012). In general, inorganic
semiconductors are used as photocatalysts, with titanium dioxide

* Corresponding author.
E-mail address: bellato@ufv.br (C.R. Bellato).
http://dx.doi.org/10.1016/j.jenvman.2015.03.051
0301-4797/ 2015 Elsevier Ltd. All rights reserved.

(TiO2) having become the most widely investigated because it is


naturally abundant, inexpensive, non-toxic and stable during
photocatalysis (Ma et al., 2013) (Ma et al., 2014).
When TiO2 is irradiated with photons at an appropriate wavelength, this semiconductor undergoes an electronic excitation
process and electrons in its valence band are promoted to the
conduction band. Through this photoelectronic excitation process,
the photocatalyzer becomes a very efcient oxidizing or reducing
agent. TiO2 has demonstrated exceptional performance in the
photocatalytic degradation of various organic dyes (Chala et al.,
2014) (C.-Y. Chen et al., 2012) (Gao et al., 2013) (Hsiao et al.,
2014). However, nanoparticles of commercial TiO2 are difcult to
recover after use and can accumulate or even plug up reactors, thus
limiting their practical application (Huang et al., 2013). Moreover,
due to their high hydrophilicity, it is very difcult for hydrophobic
organic pollutants to reach the highly hydrophilic nano-TiO2 surface, which is necessary for their photodegradation. To solve these
problems, supports for TiO2 nanoparticles such as montmorillonite
(Ooka et al., 2003) (Miao et al., 2006) (Zhang et al., 2008), zeolites
(Tayade et al., 2007) and carbon nanomaterials (Sellappan et al.,

226

L.D.L. Miranda et al. / Journal of Environmental Management 156 (2015) 225e235

2011) (Sampaio et al., 2011) (Yao et al., 2010) have been effectively
used to improve photocatalytic activity. Takeuchi et al. (2007, 2009)
found that a simple mechanical mixture of zeolites and TiO2
increased the photocatalytic activity of nano-TiO2 signicantly. An
et al. (2008) showed that TiO2 supported on surfactant-modied
montmorillonite completely degraded decabromodiphenyl ether.
 mez-Sols et al. (2012) deposited TiO2 on silicon carbide (SiC)
Go
and found that the SiCeTiO2 catalyst had higher photocatalytic
activity towards organic dyes than unmodied TiO2. Huang et al.
(2013) immobilized nano-TiO2 on hydrophobic layered double hydroxides (LDH) and used it to efciently remove dimethyl phthalate
in water (Huang et al., 2013).
The most important class of LDH anionic clays are the hydrotalcites
(HT)
that
have
the
general
formula
x n
x
[Mg21xAl3
x (OH)2] [Ax/n$mH2O] , where x ranges from 0.17 to
0.33. Hydrotalcites are good anion exchangers since their anions
(An) and interlamellar water molecules can be exchanged for anions in solution (Sahu et al., 2013) (Toledo et al., 2011) (Toledo et al.,
2013) (Miranda et al., 2014). LDHs are porous materials with large
surface areas containing many hydroxyl groups, making them
excellent dispersants for nano-TiO2 particles. Furthermore,
adsorption capacity of hydrophobic organic pollutants on LDHs can
be signicantly improved by modication with surfactants such as
sodium dodecylsulfate (DS) (Miranda et al., 2014) (Bouraada et al.,
2012) (Bruna et al., 2012). The special lamellar structure of LDHs
improves access to adsorption sites where photocatalytic reactions
occur. The immobilization of nano-TiO2 particles on DS-modied
LDH increases the hydrophobicity of the catalyst's surface,
thereby increasing the efciency of photocatalytic degradation of
hydrophobic contaminants (Huang et al., 2013) (Ooka et al., 2003).
The combination of iron oxide with HT to facilitate catalyst
removal from aqueous solution through application of a magnetic
eld reduces catalyst recovery time and costs (Toledo et al., 2011)
(Miranda et al., 2014). Several magnetic catalysts have been
mentioned in the literature for methylene blue (MB) dye photodegradation (Chen et al., 2012) (Sun et al., 2014). In the present
work, we synthesized two catalysts, HT/TiO2/Fe and HT-DS/TiO2/Fe,
which, to the best of our knowledge, have not been reported in the
literature for photodegradation of aqueous organic species. We
evaluated the synergistic effect provided by adsorption of TiO2 on
HT-DS (hydrotalcite intercalated with the dodecylsulfate ion), TiO2
photoactivity and the magnetic properties of iron oxide in the new
catalysts. We also evaluated the quantities of TiO2 and iron oxide
added to the catalysts in MB photocatalytic degradation experiments carried out in a reactor under UVevisible radiation.
2. Material and methods
2.1. Material
Metal nitrates (Mg(NO3)2.6H2O and Al(NO3)3.9H2O), sodium
hydroxide, sodium carbonate, sodium dodecylsulfate, iron III
chloride hexahydrate, ferrous sulfate heptahydrate and titanium
dioxide nanoparticles (particle size 25 nm, 99.7%) in the anatase
form used as photocatalyst were obtained from SigmaeAldrich.
Methylene blue dye was purchased from Vetec (Rio de Janeiro,
Brazil). All solutions were prepared with analytical grade reagents
and high purity deionized water produced by a Milli-Q system
(Millipore, Bedford, MA, USA).
2.1.1. Preparation of HT/TiO2
HT/TiO2 were prepared by xing the Al:(Al Mg) molar ratio at
0.25 and varying the molar amount of TiO2 (0.5, 1, 1.5, 2 and 3 mol of
Ti). A 100 mL solution of Mg(NO3)2$6H2O (0.075 mol and
Al(NO3)3$9H2O (0.025 mol) was added dropwise to a 100 mL

solution of NaOH (0.1805 mol), Na2CO3 (0.084 mol) and TiO2 (0.5, 1,
1.5, 2 or 3 mol) and the mixtures stirred for 24 h. The catalysts were
hydrothermally treated at 80  C for 24 h and the precipitates rinsed
with distilled water and dried at 60  C. The catalysts were labeled as
HT/0.5TiO2, HT/1TiO2; HT/1.5TiO2, HT/2TiO2 and HT/3TiO2.
For comparison purposes, a hydrotalcite was prepared with
interlayered carbonate ions (HT-CO3), [Mg3Al(OH)8]2CO3$nH2O, by
the method of co-precipitation at variable pH, as described in the
literature (Bruna et al., 2012).
2.1.2. Preparation of HT/TiO2/Fe
A 100 mL solution of Mg(NO3)2$6H2O (0.075 mol) and
Al(NO3)3$9H2O (0.025 mol) was added dropwise to a 100 mL solution of NaOH (0.1805 mol), Na2CO3 (0.084 mol) and TiO2 (2 mol in
relation to Ti) and the mixtures stirred for 24 h. The suspensions
were heated to 70  C and 20 mL of different amounts of FeCl3$6H2O
and FeSO4$7H2O (3.7$103 and 8.6$103; 7.4$103 and 1.7$102;
1.1$102 and 2.6$102; 1.5$102 and 3.5$102 mol) solutions were
added, at a xed Fe3: Fe2 molar ratio of 0.5. Iron oxide was
precipitated by addition of NaOH (5 mol/L) to pH10. The catalysts
were hydrothermally treated at 80  C for 24 h and the precipitates
rinsed with distilled water and dried at 60  C. The magnetic catalysts were labeled as HT/TiO2/Fe14, HT/TiO2/Fe24, HT/TiO2/Fe34
and HT/TiO2/Fe44, based on the molar ratios of Fe:Ti used.
2.1.3. Preparation of HT-DS/TiO2/Fe
The photocatalysts were obtained by the co-precipitation
method under an atmosphere of N2. A 150 mL aqueous solution
of Mg(NO3)2$6H2O (0.075 mol) and Al(NO3)3$9H2O (0.025 mol) was
added dropwise to a 750 mL solution of NaOH (0.20 mol) and sodium dodecylsulfate (0.025 mol) and the suspensions stirred for
24 h TiO2 was then added (2 mol relative to Ti) and the suspensions
stirred for another 24 h, after which they were heated to 70  C.
Twenty mL of different amounts of FeCl3.6H2O and FeSO4.7H2O
(9.3$103 and 2.2$103; 7.4$102 and 1.7$102; 1.1$102 and
6.5$102; 3.8$102 and 8.8$102 mol) were added to the suspensions. Iron oxide was precipitated by adding NaOH (5 mol/L) to pH
10. The suspensions were hydrothermally treated at 60  C for 24 h
and the precipitates rinsed with distilled water and dried at 60  C
(Huang et al., 2013) (Toledo et al., 2013) (Bruna et al., 2012). The
magnetic catalysts were labeled as HT-DS/TiO2/Fe23, HT-DS/TiO2/
Fe43, HT-DS/TiO2/Fe63 and HT-DS-TiO2/Fe73, based on the molar
ratios of Fe:Ti used. After the photodegradation experiments, the
catalysts continued to present magnetic properties and all were
removed from solution by a 0.3 T magnet. It was necessary to add
2.5-fold more iron to the HT-DS/TiO2/Fe catalysts than the HT-TiO2/
Fe catalyst to ensure their magnetic properties, which is explained
by the exchange of CO2
3 for the dodecylsulfate ion.
For comparison purposes, prepared hydrotalcite-TiO2 intercalated with anionic surfactant dodecyl, HT-DS/2TiO2.
2.1.4. Characterization of catalysts
Infrared spectroscopy (IR) analysis was carried out directly on
the sample in a VARIAN 660-IR infrared spectrophotometer
equipped with a PIKE GladiATR attenuated reectance accessory
over the region of 400e4000 cm1. X-ray diffraction analysis was
performed on an X'Pert PRO (PANalytical) X-ray diffraction system
using a Ni lter and Cu-ka radiation (l 1.54 ) at an angular
variation of 0e70 (2q). UVevisible spectra were obtained on an
Agilent 8453 spectrophotometer. Diffuse reectance spectra were
acquired on a dual-beam 20 GBC, Cintra model spectrophotometer,
in the region of 350e700 nm. Calcium carbonate was used as a nonabsorbing standard. The spectrum of each sample was obtained by
scanning the established range at a speed of 100 nm/min, for about
4 min. Measurements were acquired at a 0.5 nm resolution and

L.D.L. Miranda et al. / Journal of Environmental Management 156 (2015) 225e235

2 nm slit thickness. Total organic carbon was measured using a


Shimadzu-5000A TOC analyzer. Catalyst particle morphology was
analyzed by scanning electron microscopy (SEM) using a JEOL
model JSM-6010LA electron microscope with a tungsten lament,
operated at an accelerating voltage up to 20 kV. In order to detect
the presence of TiO2 and iron oxide in the catalysts, analyses were
also made by Raman scattering, using a Renishaw Raman Invia
micro spectrometer equipped with an argon laser (514.5 nm) with a
50 objective (NA 0.75, corresponding to a spot of approximately
1 mm in diameter). To avoid heating effects, power used in the
Raman scattering did not exceed 1 mW.
2.2. Adsorption and photocatalytic tests
2.2.1. Adsorption tests
MB adsorption on the catalysts was evaluated using Langmuir
and Freundlich isotherm models. Values of the Langmuir constants
(qmax and KL) were obtained by regression using the linearized form
of the Langmuir equation [Eq. (1)]:

Ce
1
Ce

qe qmax KL qmax

(1)

where qmax (mg/g) and KL (L/mg) are the Langmuir constants


associated with the capacity and the adsorption energy, qe is the
amount of substance adsorbed (mg/g) and Ce is the equilibrium
concentration (mg/L).
Values of the Freundlich constants (n and KF) were obtained by
regression using the linearized form of the Freundlich equation
[Eq. (2)]:

lnqe

1
lnCe lnKF
n

227

bulb, encased in a 70.0 cm tall, 4 cm diameter glass cylinder (lter


cutoff for l > 300 nm). This was placed at the center of another
glass cylinder (60 cm tall, 7 cm diameter, with a total capacity of
1000 mL) containing 300 mL of a 24 mg/L MB solution and 90 mg of
catalyst. The reaction mixture was rst stirred in the dark for
30 min to reach the adsorption equilibrium and then exposed to
UVevisible radiation provided by the mercury lamp.
To avoid overheating caused by the mercury vapor lamp, the
reactor was encased in a water sleeve that maintained the temperature at 30 2  C. At 10 min intervals, 3 mL samples were
removed from the reactor using a syringe. The catalyst was
magnetically separated from the solution before measuring absorption at 665 nm. A blank reaction was carried out under the
same conditions but without catalyst addition.

3. Results and discussion


3.1. Characterization
3.1.1. Infrared spectroscopy
The HT-CO3 IR spectrum (Fig. 2A), shows an absorption band at
3394 cm1 due to stretching of the interlayer water molecule bonds
(n OeH). The strong absorption peak at 1359 cm1 is assigned to the
vibration of carbonate species. The bands in the range of
450e780 cm1 are attributed to stretching of AleO and MgeO
bonds (Toledo et al., 2013). For HT/TiO2/Fe (Fig. 2D), these bands
were superimposed on the FeeO band (Fig. 2B) and the shoulder of
the TieOeO bond in the region of 450e780 cm1 (Fig. 2C).
The modication of HT-CO3 by intercalation of the DS surfactant
is apparent in its IR spectrum (Fig. 2E). The triplet at 2958, 2914 and

(2)

where qe is the amount adsorbed (mg/g), Ce is the equilibrium


concentration of the adsorbate (mg/L); and KF and n are the
Freundlich constants related to the capacity and intensity of
adsorption, respectively.
2.2.2. Photocatalytic tests
The adsorption-photodegradation experiments were performed
in a photo-reactor, schematically represented in Fig. 1 The system
consisted of a 125 W mercury vapor lamp without the protective

Fig. 1. Schematic of the photoreactor employed in the photocatalytic tests.

Fig. 2. IR spectra of: (A) HT-CO3; (B) magnetic iron oxide; (C) nano-TiO2; (D) HT/TiO2/
Fe34 and (E) HT-DS/TiO2/Fe23.

228

L.D.L. Miranda et al. / Journal of Environmental Management 156 (2015) 225e235

2844 cm1, characteristic of stretching of the CeH bonds (n CeH) in


CH3 and CH2 moieties, is evidence of the presence of dodecylsulfate
ions in the interlayer space. Sulfate bond stretching bands at
1210 cm1 (ns S]O) and 1057 cm1 (nass S]O) are also visible
(Miranda et al., 2014) (Bouraada et al., 2008). The absence of a peak
around 1359 cm1 (Fig. 2E) associated with stretching of the CO2
3
group is also evidence of the intercalation of the surfactant in the
interlayer space (Miranda et al., 2014). The characteristic iron oxide
band (Fig. 2B) occurs at around 602 cm1 due to the FeeO bond
(Toledo et al., 2013) (Miranda et al., 2014). The shoulder observed at
650 cm1 (Fig. 2C), is attributed to TieOeTi vibration (Lin et al.,
2011) (Low and Boonamnuayvitaya, 2013). The characteristics
bands of the FeeO and TieO bonds are better visualized using the
Raman scattering technique.
3.1.2. Raman spectroscopy
Fig. 3 shows the Raman spectra between 100 and 1100 cm1 for
catalysts (E and F) and their precursors, HT-CO3 (A), HT-DS (B),
nano-TiO2 (C) and iron oxide (D). The spectra were normalized by
measurement parameters to permit comparison of Raman peak
intensities. Thus, the spectrum in Fig. 3C is about 350-fold and the
spectrum in Fig. 3F about 0.3-fold as intense as the spectra in
Fig. 3A, B, D and E.
The HT-CO3 Raman spectrum (Fig. 3A) presents the strongest
bands at 148, 558 and 1060 cm1, attributed to the interlayer water
molecules, the stretching vibration of the AleOeMg linked to the
3
octahedral layers and CeO stretching in CO2
3 bonded to the Al
1
bound OH groups, respectively. The Raman band at 482 cm are
assigned to AleOeAl bond vibrations (Burrueco et al., 2013).

Fig. 3. Raman spectra of the catalysts and its precursors. (A) HT-CO3; (B) HT-DS; (C)
magnetic iron oxide; (D) TiO2; (E) HT/TiO2/Fe34 and (F) HT-DS/TiO2/Fe23.

After dodecylsulfate anion intercalation the spectrum changed,


as expected (Burrueco et al., 2013), to the prole in Fig. 3B, in which
the 148, 482 and 558 cm1 Raman bands are not present, while the
band at 1060 cm1 shifted to 1074 cm1. The others Raman features
are assigned to the intercalated dodecylsulfate.
The Raman lines at 142, 397, 516 and 636 cm1 in Fig. 3C can be
assigned to the Eg, B1g, A1g B1g, and Eg modes of the TiO2
anatase form (Sellappan et al., 2011). The broad band at 670 cm1 in
Fig. 3D is a characteristic Raman band of magnetic iron oxide and
can be associated to the totally symmetric mode (A1g) of magnetite, Fe3O4(s) (Li et al., 2012).
Fig. 3E and F presents the spectra of HT-CO3 and HT-DS after
incorporation of TiO2 and deposition of magnetic iron oxide. These
spectra are dominated by the TiO2 and iron oxide bands, with the
1060 and 1074 cm1 hydrotalcite bands only weakly visible. These
results indicate a strong incorporation/deposition of TiO2 nanoparticles on the hydrotalcite surface, totally covering the structure.
The stronger intensity of TiO2 Raman peaks in Fig. 3E suggests that
TiO2 nanoparticles are better incorporated onto HT-CO3 than onto
HT-DS.
3.1.3. X-ray diffraction
The X-ray diffraction patterns of the samples are presented in
Fig. 4.
Hydrotalcite (HT-CO3) presented peaks of interplanar distance
(d) and magnitude (hkl) similar to the crystallographic standards
described in the literature (Toledo et al., 2011) (Toledo et al., 2013)
(Bouraada et al., 2008) (Zhao et al., 2008). The 2q (0 0 3) basal
diffraction spacing of HT-CO3 was estimated to be 7.6 , indicative
of highly crystalline hydrotalcite layers. After addition of DS, the
shift in the d(0 0 3) peak to a 2q value of 3.29 and interplanar

Fig. 4. X-ray diffraction spectra. (A) HT-CO3; (B) iron oxide; (C) TiO2; (D) HT/TiO2/Fe34
and (E) HT-DS/TiO2/Fe23.

L.D.L. Miranda et al. / Journal of Environmental Management 156 (2015) 225e235

spacing of 26.8 indicates intercalation of dodecyl ions between


the hydrotalcite lamella (Huang et al., 2013). A proposed structure
for HT-DS/TiO2/Fe and a scheme for hydroxyl radical generation on
its surface are shown in Fig. 5.
The incorporation of nano-TiO2 particles together with the
deposition of iron oxide on the surface of the catalysts led to a
hybrid and disordered compound, leading to broad and weak
diffraction peaks d(0 0 3) for HT/TiO2/Fe34 and HT-DS/TiO2/Fe23
(Miranda et al., 2014) (You et al., 2002) (Barbosa et al., 2005) (Lv
et al., 2008). Magnetite and maghemite (Fig. 4B) can be identied
by the 2q diffraction peaks at 30.1, 35.4, 43, 53.4, 57 and 62.5 ,
indicating their surface deposition which is responsible for the
catalysts' magnetic properties (Quinones et al., 2014) (He et al.,
2012). The magnetic properties are advantageous since they make
it possible to easily remove the catalysts from solution after use by a
simple magnetic process (Toledo et al., 2013) (Miranda et al., 2014).
The characteristic diffraction of the anatase phase d(1 0 1) at a 2q of
about 25 can be seen in the spectra of HT/TiO2/Fe34 (Fig. 4D) and
HT-DS/TiO2/Fe23 (Fig. 4E) (C.-Y. Chen et al., 2012). Fig. 4D and E
shows that deposition of iron oxides and immobilization of TiO2 did
not change the lamellar structure characteristic of HT-CO3 and HT
intercalated with DS since the diffraction characteristics of HT/TiO2/
Fe34 did not change from those of its precursors, HT-CO3, iron
oxides and nano-TiO2. There was only an overlapping of the peaks,
demonstrating that the combination of iron oxide, TiO2 and HT-CO3
is likely to be a physical process. The same was observed for HT-DS/
TiO2/Fe23 (Fig. 4E).

3.1.4. Scanning electron microscopy


HT-CO3 (Fig. 6A) presented a more regular, smoother surface
than the modied hydrotalcites (Fig. 6B and C) (Huang et al., 2013)
(Costa et al., 2008). The deposition of iron oxide and the random

229

distribution of TiO2 particles on the surface of the catalysts formed


agglomerations with a spongy appearance and were responsible for
the increased heterogeneity of the catalysts' surfaces (Fig. 6C)
(Huang et al., 2013) (Toledo et al., 2013).
3.1.5. Diffuse reectance
Measurement of absorption by diffuse reectance in the
UVevisible region is a convenient and effective method to investigate the semiconductor band gap since electrons migrate from the
valence band to the conduction band in photocatalytic semiconductors by absorbing light (Lu et al., 2012). The intensity of the
energy reected by the catalyst can be obtained using the KulbelkaMunk function, F(R) (Eq. (3)).

FR

1  R2
2R

(3)

where R is the absolute diffuse reectance of the beam, that is the


ratio between the reected intensity of the sample and of a nonabsorbent standard, calcium carbonate. The relationship between
diffuse reectance and band gap energy is given in Eq. (4):

FRhl2 Chl  Eg

(4)

where hl is the energy of a photon and C is the constant of proportionality. The band gap energy (Eg) was obtained by plotting
(F(R)hl)2 versus hl, as described in the literature (Lv et al., 2008)
(Song et al., 2012).
The diffuse reectance spectra for the UVevisible absorption of
HT-DS, nano-TiO2, iron oxide and HT-DS/TiO2/Fe23 samples are
shown in Fig. 7. There was no absorption in the UVevisible region
for HT-DS, in agreement with the literature (Huang et al., 2013).
However, nano-TiO2, iron oxides and HT-DS/TiO2/Fe23 presented

Fig. 5. Schematic illustration of the HT-DS/TiO2/Fe catalyst and formation of hydroxyl radicals on its surface.

230

L.D.L. Miranda et al. / Journal of Environmental Management 156 (2015) 225e235

Fig. 6. SEM of the catalysts: (A) HT-CO3; (B) HT-TiO2/Fe34 and (C) HT-DS/TiO2/Fe23.

absorption bands. Absorption shifted towards a longer wavelength,


or lower band gap energy, in HT-DS/TiO2/Fe23 compared to nanoTiO2. The decreased energy suggests that less energy would be
required for photocatalytic activity (Lv et al., 2008) (Song et al.,
2012). The value of the band gap determined for TiO2 (3.3 eV) is
in agreement with the values reported in the literature (between
3.2 and 3.3 eV) (Sellappan et al., 2011) (Rashad et al., 2013) (Song
et al., 2012).
Light absorption in the HT-DS/TiO2/Fe23 catalyst shifted towards longer wavelengths in the visible range. This shift is purportedly due to electronic interaction between the molecular
orbitals of iron oxides (which possess narrow band openings, 0.1 eV
for magnetite and 2.2 eV for maghemite) and TiO2 (3.2e3.3 eV)
forming a new molecular orbital and reducing the band difference.

Fig. 7. The diffuse reectance UVevisible spectra of HT-DS, nano-TiO2, magnetic iron
oxide and HT-DS/TiO2/Fe23.

Similar phenomena were observed between graphene and TiO2


(Liu et al., 2013), carbon nanotubes and TiO2 (Sampaio et al., 2011)
and graphene, TiO2 and magnetite (Tang et al., 2013).
3.2. Adsorption and photocatalytic activity of the synthesized
catalysts
Adsorption isotherms of MB on HT-DS/TiO2/Fe are shown in
Fig. 8 and the values of the adsorption constants (KL, Q max, KF, n and
R2) are shown in Table 1. The Langmuir model (R2 > 0.99) described
the adsorption process better than the Freundlich (R2 > 0.94)
model, as previously found for adsorption of pollutants on hydrophobic organic LDHs (Huang et al., 2013). According to the Langmuir model, when one molecule of MB is adsorbed onto a given
location of the adsorbent, this location becomes unavailable to
other molecules and a single layer of dye molecules is formed
(Miranda et al., 2014). MB adsorption capacities based on Qmax
values decreased in the order: HT-DS/TiO2/Fe23 > HT-DS/TiO2/
Fe43 > HT-DS/TiO2/Fe63 > HT-DS/TiO2/Fe73. The catalyst with an

Fig. 8. Langmuir adsorption isotherm for MB on the HT-DS/TiO2/Fe catalysts.

L.D.L. Miranda et al. / Journal of Environmental Management 156 (2015) 225e235


Table 1
Isotherm model parameters for the adsorption of MB on HT-DS/TiO2/Fe23.
Catalyst

HT-DS/TiO2/Fe23
HT-DS/TiO2/Fe43
HT-DS/TiO2/Fe63
HT-DS/TiO2/Fe73

Langmuir constants

Freundlich constants

KL (L/mmol)

Qmax (mg/g)

R2

KF(min)

R2

0.473
0.532
0.669
0.821

133.37
119.30
98.62
43.41

0.990
0.990
0.996
0.998

71.636
66.007
57.538
26.888

4.926
5.208
5.747
7.092

0.945
0.951
0.947
0.941

iron oxide:TiO2 molar ratio of 2:3 exhibited the highest MB


adsorption (53.4%).
The photocatalytic activity of hydrotalcite can be signicantly
affected by the amount of TiO2 incorporated on its surface, thus the
HT:TiO2 ratio should be optimized. As shown in Fig. 9A, after
120 min reaction, the HT-2TiO2 catalyst exhibited a higher

photocatalytic activity than HT-CO3 and catalysts prepared with


different amounts of TiO2 (HT/0,5TiO2; HT/1TiO2; HT/1,5TiO2 and
HT/3TiO2). Increasing the quantity of TiO2 incorporated into the HTCO3 up to 2 mol Ti led to increased generation of e/H pairs and an
increased photodegradation rate. However, above this optimum
level, TiO2 blocked the passage of light and increased light scattering, resulting in a lower rate of photodegradation for the HT3TiO2 catalyst (Wang et al., 2014) (Huang et al., 2013). Approximately 13% of the MB was degraded by direct photolysis in the
absence of catalyst, indicating that radiation alone had a weak effect on MB degradation (Fig. 9A).
The magnetic characteristic of the catalysts arises from the iron
oxide deposited on their surface, while photocatalytic activity is
mainly located in the TiO2 layer. It is thus necessary to dene a
suitable iron oxide:TiO2 molar ratio to achieve efcient photodegradation while ensuring the magnetic characteristics of the
catalysts (He et al., 2012). The photocatalytic activity of the HT/
2TiO2 catalyst with different amounts of iron oxide was therefore
studied (Fig. 9B) and it was observed that the HT/TiO2/Fe34 catalyst
showed the greatest photo-activity. A small reduction (10%) in MB
degradation was observed using magnetic HT/TiO2/Fe34 compared
to nonmagnetic HT/2TiO2 catalysts (Fig. 9B). However, the magnetic
catalysts have the advantage of being easily removed from solution
by applying a magnetic eld, thus reducing time and costs of their
separation/recovery (Miranda et al., 2014).
Photodegradation of organic pollutants by TiO2 photocatalysis
occurs mainly on or near the surface of the catalyst and thus
adsorption is a critical factor in the efciency of the process (Huang
et al., 2013) (Ooka et al., 2003) (An et al., 2008). Previous investigations have shown that hydrotalcite-iron oxide intercalated
with surfactants are good adsorbents for MB (Miranda et al., 2014).
Therefore, we evaluated the synergistic effects of combining
adsorption on HT-DS, photo-activity of TiO2 and the magnetic
characteristic provided by iron oxide of the new magnetic catalysts
for MB photodegradation.
Photocatalytic activity of the HT-DS/TiO2 catalyst with different
quantities of iron oxide was studied and it was observed that the
HT-DS/TiO2/Fe23 catalyst showed the highest degree of photoactivity (Fig. 9C). Magnetic HT-DS/TiO2/Fe23 photoactivity was
slightly lower than that of nonmagnetic HT-DS/2TiO2, resulting in a
6% reduction in MB degradation, but this decrease is outweighed by
the advantage of ready removal of the magnetic catalyst from solution by applying a magnetic eld (Miranda et al., 2014).
Fig. 10A compares MB degradation by HT-CO3, HT/TiO2/Fe34,
HT-DS/TiO2/Fe23, TiO2 nano-particles and photolysis. The HT-DS/
TiO2/Fe23 catalyst showed the highest rate of MB
photodegradation.
The photodegradation process and the possible MB degradation
pathways are illustrated in Fig. 5. When a photon (hn) with energy
equal to or greater than the band gap energy hits the HT-DS/TiO2/Fe
catalyzer, an electron in the valence band (VB) is promoted to the
conduction band (CB), leading to simultaneous generation of a gap
in the valence band (h) and an excess of electrons in the conduction band (e) (Eq. (5)) (Huang et al., 2013).



HT  DS=TiO2 =Fe hv/HT  DS=TiO2 e h =Fe

Fig. 9. Photocatalytic activity of the HT catalysts: (A) with different amounts of TiO2;
(B) with TiO2 and different amounts of iron oxide and (C) interspersed with DS surfactant, with TiO2 and different amounts of iron oxide in the removal of MB. Initial
concentration of MB 24 mg/L; catalyst dose 0.3 g/L; reaction temperature 30 2  C.

231

(5)

Although various reaction mechanisms for MB degradation have


been reported in the literature, it is accepted that the strongly
oxidizing band gaps can directly attack the MB adsorbed on the
surface (Eqs. (6) and (7)) or can indirectly oxidize MB through the
formation of hydroxyl radicals (OH) (Eq. (8)). The gaps (h)
generated possess sufciently positive potentials to generate hydroxyl radicals from water molecules adsorbed on the surface, and
these radicals can in turn oxidize MB (Eq. (9) and Fig. 5) (Huang

232

L.D.L. Miranda et al. / Journal of Environmental Management 156 (2015) 225e235

et al., 2013) (Konstantinou and Albanis, 2004) (Jing et al., 2011).

h MB/oxidation products


e MB/reduction products

(6)
(7)

h H2 Oads /OH H

(8)

OH MB/CO2 H2 O

(9)

When oxygen is present it can act as a CB electron acceptor,


initiating a series of free radical chain reactions that also result in

OH formation (Eqs. (10)e(14)) (Nogueira and Jardim, 1998) (Ziolli
and Jardim, 1998).

O2 e /O
2

(10)



O
2 H /HO2

(11)

HO2  HO2 /H2 O2 O2

(12)

H2 O2 e /OH OH

(13)




H2 O2 O
2 /OH OH O2

(14)

LDHs are hydroxyl-rich materials and their hydroxyl groups can


react with valence band gaps to produce OH (Eq. (15) and Fig. 5)
(Huang et al., 2013).

h OH /OH

(15)

These OH produced on the surface are considered to be more


effective in attacking the adsorbed MB because of their mutual
proximity (Huang et al., 2013). Therefore, greater adsorption
probably caused the higher percentage of MB degradation observed
for the HT-DS/TiO2/Fe23 catalyst compared to the HT/TiO2/Fe34
catalyst. After 120 min of reaction, MB removal efciency (Fig. 10A)
was 41, 72 and 96% and TOC removal efciency (Fig. 10B) was 15, 42
and 61% for HT-CO3, HT/TiO2Fe34 and HT-DS/TiO2/Fe23, respectively. It should be noted that all tests were performed using the
same amount of catalyst. However, the magnetic HT-DS/TiO2/Fe23
catalyst exhibited photocatalytic performance similar to TiO2 alone,
although a lesser amount of TiO2 was present in the new catalyst.
Typical evolution of MB absorption during radiation is presented in Fig. 10C. A sharp decrease (53.4%) in MB absorption peak
intensity was found after adsorption in the dark for 30 min, because
of the relatively high MB adsorption capacity (133.37 mg/g) of the
HT-DS/TiO2/Fe23 catalyst. Upon radiation, the characteristic MB
absorption peak decreased gradually and almost no color was
observed after 120 min, indicating that MB was degraded by HT-DS/
TiO2/Fe23. Fig. 10D shows the HT-DS/TiO2/Fe23 catalyst attracted to
a 0.3 T magnet after use in MB photodegradation, illustrating its
complete removal from solution.
The similarity in photocatalytic activity of HT-DS/TiO2/Fe23 and
TiO2 can be attributed to synergistic effects of four factors: (1) HTDS/TiO2/Fe23 is an excellent dispersant and transporter for nanoTiO2 particles; (2) HT-DS/TiO2/Fe23 has a high adsorption capacity
for MB (133.37 mg/g), which assists in photocatalytic performance;
(3) the HT surface is rich in hydroxyl groups that favor production of
OH by reaction with gaps; and (4) HT-DS/TiO /Fe23 shifts light
2
absorption to a longer wavelength, and less energy is needed for
photocatalytic activity.
The novel magnetic HT-DS/TiO2/Fe23 catalyst showed satisfactory performance in MB degradation when compared to other TiO2
supported catalysts reported in the literature (Table 2).
3.3. Analysis of the material after reaction
Fig. 10. (A) Comparison of the photocatalytic activity of the catalysts for the removal of
MB; (B) Variation in the total organic carbon content (TOC) of the solution during the
MB photodegradation; (C) Absorption spectra for MB degradation using the HT-DS/
TiO2/Fe23 and (D) Photograph the MB catalyst solution before (left) and after (right)
photodegradation, and HT-DS/TiO2/Fe23 being attracted by a magnet (right).

Four cycles of MB photodegradation were performed with the


same HT-DS/TiO2/Fe23 catalyst in order to evaluate its stability.
There was a small reduction in MB degradation upon reuse of HT-

L.D.L. Miranda et al. / Journal of Environmental Management 156 (2015) 225e235

233

Table 2
MB dye degradation by different TiO2 supported catalysts.
Catalyst

SICeTiO2
TiO2eMn doped
TiO2eCo doped
BiVO4eFe
ZnAlTi-HDL
Fe3O4@HDL@Ag/Ag3PO4
ZnAlTi-HDL
Fe3O4/ZnCr-HDL
HT-DS/TiO2/Fe23

Experimental conditions
Dose (g/L)

% degradation

MB solution (mg/L)

Radiation time (min)

Ref.

1
1
1
1
1
1
2
1
0,3

52
~100
~100
81
~100
100
100
~95
~100

20
5
5
5
10
10
35
5
24

180
60
60
30
120
60
250
180
120

mez-Sols et al., 2012


Go
Rashad et al., 2013
Rashad et al., 2013
Chala et al., 2014
Wang et al., 2014
Sun et al., 2014
Sahu et al., 2013
Chen et al., 2013
This study

DS/TiO2/Fe23 (Fig. 11), with photocatalytic activities of 94, 92, 89


and 83% in the rst to fourth cycles, suggesting a good potential for
catalyst recycling. Wang et al. (2014) reported that MB degradation
was reduced to 76% after four cycles using ZnAlTi-LDH as catalyst.
After use in MB treatment, the HT-DS/TiO2/Fe23 catalyst was
removed from solution using a magnet and then dried at 60  C
without washing. Recovered catalyst was analyzed before and after
both adsorption and irradiation in order to check the stability of HTDS/TiO2/Fe23 after one cycle (2 h) and four cycles (8 h) of MB
degradation using IR spectroscopy (Fig. 12C) and X-ray (Fig. 12D)
analyses. Small changes were observed in the IR absorption bands
after MB adsorption. The 1600 cm1 band became sharper due to
overlapping of the n C]N band and the 1485 and 1339 cm1 band
intensities increased because of overlapping of the MB n C]C and
CAr-N bands, respectively (Yu and Chuang, 2007). After UVevisible
irradiation, these changes disappeared (Fig. 12D), indicating that
MB that had adsorbed on the catalyst was degraded. The similarity
between catalyst spectra before and after photodegradation indicates that the dodecylsulfate ion remained in the interlayer space.
A similar phenomenon was observed for TiO2 supported on LDH
intercalated with DS after photodegradation of dimethyl phthalate
(Huang et al., 2013). The recycled catalyst's IR spectrum (Fig. 12D)
contained characteristic DS bands: a CH2 band between 2844 and
2958 cm1, a symmetrical stretching vibration at ~1210 cm1 and
an asymmetrical stretching vibration at ~1057 cm1 arising from
the sulfonyl (eS]O) vibration (Miranda et al., 2014), evidence that
DS still existed in HT-DS/TiO2/Fe23 after 8 h of irradiation (Fig. 12D).
To investigate further the possible degradation of DS in the

Fig. 11. Effect of HT-DS/TiO2/Fe23 catalyst reuse in MB photodegradation. 1-First cycle,


2-Second cycle, 3-Third cycle and 4-Fourth cycle. Initial concentration of MB 24 mg/L;
catalyst dose 0.3 g/L; reaction temperature 30 2  C; reaction time 2 h.

interlamellar space of HT, the X-ray diffraction patterns of the


samples were obtained (Fig. 13). After repeated use, the catalyst's
crystallinity did not signicantly change, with basal spacing, d(0
0 3), remaining at 25e26 , once again indicating that the DS ions
remained in the HT interlalamellar space, even after 8 h of irradiation. The characteristic TiO2 and the iron oxide peaks were also
preserved (Fig. 13).
4. Conclusions
Novel magnetic catalysts, HT/TiO2/Fe and HT-DS/TiO2/Fe, were
synthesized in this work and their activity evaluated in photodegradation of MB dye. Synergistic effects of adsorption on HT-DS,
photoactivity of TiO2 and magnetic properties provided by iron
oxide were observed. The photocatalytic activity of the catalysts
was signicantly affected by the amount of TiO2 and iron oxide

Fig. 12. IR spectrum of materials: (A) MB; (B) HT-DS/TiO2/Fe23; (C) HT-DS/TiO2/Fe23
after adsorption and (D) HT-DS/TiO2Fe23 after irradiation.

234

L.D.L. Miranda et al. / Journal of Environmental Management 156 (2015) 225e235

Fig. 13. (A) The X-ray of the HT-DS/TiO2/Fe23 catalyst; (B) after 2 h of UVeVisible
radiation and (C) after 8 h of UVeVisible radiation.

incorporated into HT and DS-modied HT. The highest MB (96%)


and TOC (61%) removals after 120 min of reaction were obtained
with the HT-DS/TiO2/Fe catalyst at an iron oxide:TiO2 molar ratio of
2:3. The HT-DS/TiO2/Fe23 catalyst shifted light absorption to a
longer wavelength, and less energy was needed for photocatalytic
activity. Moreover, the catalyst could easily be separated from the
treated solution for reuse by simply applying an external magnetic
eld.
Acknowledgments
~o
The authors acknowledge the nancial support of the Fundaa
 Pesquisa do Estado de Minas Gerais (FAPEMIG, Unide Amparo a
versal Demand, process number: APQ-00416-11) and the Conselho
gico.
Nacional de Desenvolvimento Cientco e Tecnolo
References
An, T., Chen, J., Li, G., Ding, X., Sheng, G., Fu, J., Mai, B., O'Shea, K.E., 2008. Characterization and the photocatalytic activity of TiO2 immobilized hydrophobic
montmorillonite photocatalysts. Catal. Today 139, 69e76.
Barbosa, C.A.S., Dias, P.M., Ferreira, A.M.D.C., Constantino, V.R.L., 2005. MgeAl
hydrotalcite-like compounds containing iron-phthalocyanine complex: effect of
aluminum substitution on the complex adsorption features and catalytic activity. Appl. Clay Sci. 28, 147e158.
Bouraada, M., Lafjah, M., Ouali, M.S., de Menorval, L.C., 2008. Basic dye removal
from aqueous solutions by dodecylsulfate- and dodecyl benzene sulfonateintercalated hydrotalcite. J. Hazard. Mater. 153, 911e918.
norval, L.C., 2012. Dodecylsulfate and dodecyBouraada, M., Ouali, M.S., de Me
benzenesulfonate intercalated hydrotalcites as adsorbent materials for the
removal of BBR acid dye from aqueous solutions. J. Saudi Chem. Soc. 220e225.
Bruna, F., Celis, R., Real, M., Cornejo, J., 2012. Organo/LDH nanocomposite as an
adsorbent of polycyclic aromatic hydrocarbons in water and soil-water systems.
J. Hazard. Mater. 225e226, 74e80.
nez-Sanchidria
n, C., Ruiz, J.R., 2013. Raman
Burrueco, M.I., Mora, M., Jime

microspectroscopy of hydrotalcite-like compounds modied with sulphate and


sulphonate organic anions. J. Mol. Struct. 1034, 38e42.
Chala, S., Wetchakun, K., Phanichphant, S., Inceesungvorn, B., Wetchakun, N., 2014.
Enhanced visible-light-response photocatalytic degradation of methylene blue
on Fe-loaded BiVO4 photocatalyst. J. Alloys Compd. 597, 129e135.
Chen, C.-Y., Cheng, M.-C., Chen, A.-H., 2012. Photocatalytic decolorization of remazol
black 5 and remazol brilliant orange 3R by mesoporous TiO2. J. Environ. Manag.
102, 125e133.
Chen, D., Li, Y., Zhang, J., Zhou, J., Guo, Y., Liu, H., 2012. Magnetic Fe3O4/ZnCr-layered
double hydroxide composite with enhanced adsorption and photocatalytic
activity. Chem. Eng. J. 185e186, 120e126.
Costa, F.R., Leuteritz, A., Wagenknecht, U., Jehnichen, D., H
auler, L., Heinrich, G.,
2008. Intercalation of MgeAl layered double hydroxide by anionic surfactants:
preparation and characterization. Appl. Clay Sci. 38, 153e164.
Ferrarese, E., Andreottola, G., Oprea, I.A., 2008. Remediation of PAH-contaminated
sediments by chemical oxidation. J. Hazard. Mater. 152, 128e139.
Gao, Q., Wu, X., Fan, Y., Zhou, X., 2013. Low temperature fabrication of nanoower
arrays of rutile TiO2 on mica particles with enhanced photocatalytic activity.
J. Alloys Compd. 579, 322e329.
Golob, V., Vinder, A., Simonic, M., 2005. Efciency of the coagulation/occulation
method for the treatment of dyebath efuents. Dyes Pigments 67, 93e97.
mez-Sols, C., Jua
rez-Ramrez, I., Moctezuma, E., Torres-Martnez, L.M., 2012.
Go
Photodegradation of indigo carmine and methylene blue dyes in aqueous solution by SiC-TiO2 catalysts prepared by sol-gel. J. Hazard Mater 30, 217e218.
Haritash, A.K., Kaushik, C.P., 2009. Biodegradation aspects of polycyclic aromatic
hydrocarbons (PAHs): a review. J. Hazard. Mater. 169, 1e15.
He, Z., Hong, T., Chen, J., Song, S., 2012. A magnetic TiO2 photocatalyst doped with
iodine for organic pollutant degradation. Sep. Purif. Technol. 96, 50e57.
Hsiao, Y.-C., Wu, T.-F., Wang, Y.-S., Hu, C.-C., Huang, C., 2014. Evaluating the sensitizing effect on the photocatalytic decoloration of dyes using anatase-TiO2.
Appl. Catal. B Environ. 148e149, 250e257.
Huang, Z., Wu, P., Lu, Y., Wang, X., Zhu, N., Dang, Z., 2013. Enhancement of photocatalytic degradation of dimethyl phthalate with nano-TiO2 immobilized onto
hydrophobic layered double hydroxides: a mechanism study. J. Hazard. Mater.
246e247, 70e78.
Jing, Y., Li, L., Zhang, Q., Lu, P., Liu, P., L, X., 2011. Photocatalytic ozonation of
dimethyl phthalate with TiO2 prepared by a hydrothermal method. J. Hazard.
Mater. 189, 40e47.
Konstantinou, I.K., Albanis, T.A., 2004. TiO2-assisted photocatalytic degradation of
azo dyes in aqueous solution: kinetic and mechanistic investigations. Appl.
Catal. B Environ. 49, 1e14.
Li, Y.-S., Church, J.S., Woodhead, A.L., 2012. Infrared and Raman spectroscopic
studies on iron oxide magnetic nano-particles and their surface modications.
J. Magn. Magn. Mater. 324, 1543e1550.
Lin, S.-H., Chiou, C.-H., Chang, C.-K., Juang, R.-S., 2011. Photocatalytic degradation of
phenol on different phases of TiO2 particles in aqueous suspensions under UV
irradiation. J. Environ. Manag. 92, 3098e3104. http://dx.doi.org/10.1016/
j.jenvman.2011.07.024.
Liu, S., Sun, H., Liu, S., Wang, S., 2013. Graphene facilitated visible light photodegradation of methylene blue over titanium dioxide photocatalysts. Chem.
Eng. J. 214, 298e303.
Low, W., Boonamnuayvitaya, V., 2013. Enhancing the photocatalytic activity of TiO2
co-doping of graphene-Fe3 ions for formaldehyde removal. J. Environ. Manag.
127, 142e149.
Lu, R., Xu, X., Chang, J., Zhu, Y., Xu, S., Zhang, F., 2012. Improvement of photocatalytic
activity of TiO2 nanoparticles on selectively reconstructed layered double hydroxide. Appl. Catal. B Environ. 111e112, 389e396.
Lv, L., Wang, Y., Wei, M., Cheng, J., 2008. Bromide ion removal from contaminated
water by calcined and uncalcined MgAl-CO3 layered double hydroxides.
J. Hazard. Mater. 152, 1130e1137.
Ma, X., Dai, Y., Guo, M., Huang, B., 2013. Insights into the role of surface distortion in
promoting the separation and transfer of photogenerated carriers in anatase
TiO2. J. Phys. Chem. C 117, 24496e24502.
Ma, X., Dai, Y., Huang, B., 2014. Origin of the increased photocatalytic performance
of TiO2 nanocrystal composed of pure core and heavily nitrogen-doped shell: a
theoretical study. ACS Appl. Mater. Interfaces 6, 22815e22822.
Miao, S., Liu, Z., Han, B., Zhang, J., Yu, X., Du, J., Sun, Z., 2006. Synthesis and characterization of TiO2emontmorillonite nanocomposites and their application for
removal of methylene blue. J. Mater. Chem. 16, 579.
Miranda, L.D.L., Bellato, C.R., Fontes, M.P.F., de Almeida, M.F., Milagres, J.L.,
Minim, L.A., 2014. Preparation and evaluation of hydrotalcite-iron oxide magnetic organocomposite intercalated with surfactants for cationic methylene
blue dye removal. Chem. Eng. J. 254, 88e97.
lise heteroge
^nea e sua aplica~
Nogueira, R.F.P., Jardim, W.F., 1998. A fotocata
ao
ambiental. Qum. Nova 2, 69e72.
Ooka, C., Yoshida, H., Horio, M., Suzuki, K., Hattori, T., 2003. Adsorptive and photocatalytic performance of TiO2 pillared montmorillonite in degradation of
endocrine disruptors having different hydrophobicity. Appl. Catal. B Environ. 41,
313e321.

Quinones, D.H., Rey, A., Alvarez,
P.M., Beltr
an, F.J., Plucinski, P.K., 2014. Enhanced
activity and reusability of TiO2 loaded magnetic activated carbon for solar
photocatalytic ozonation. Appl. Catal. B Environ. 144, 96e106.
Rashad, M.M., Elsayed, E.M., Al-kotb, M.S., Shalan, A.E., 2013. The structural, optical,
magnetic and photocatalytic properties of transition metal ions doped TiO2
nanoparticles. J. Alloys Compd. 581, 71e78.

L.D.L. Miranda et al. / Journal of Environmental Management 156 (2015) 225e235


Sahu, R.K., Mohanta, B.S., Das, N.N., 2013. Synthesis, characterization and photocatalytic activity of mixed oxides derived from ZnAlTi ternary layered double
hydroxides. J. Phys. Chem. Solids 74, 1263e1270.
Sampaio, M.J., Silva, C.G., Marques, R.R.N., Silva, A.M.T., Faria, J.L., 2011. Carbon
nanotube e TiO2 thin lms for photocatalytic applications. Catal. Today J. 161,
91e96.
Sellappan, R., Galeckas, A., Venkatachalapathy, V., 2011. Applied catalysis B: environmental on the mechanism of enhanced photocatalytic activity of composite
TiO2/carbon nanolms. Appl. Catal. B Environ. 106, 337e342.
Song, J., Leng, M., Fu, X., Liu, J., 2012. Synthesis and characterization of nanosized
zinc aluminate spinel from a novel ZneAl layered double hydroxide precursor.
J. Alloys Compd. 543, 142e146.
Sun, J., Fan, H., Nan, B., Ai, S., 2014. Fe3O4@LDH@Ag/Ag3PO4 submicrosphere as a
magnetically separable visible-light photocatalyst. Sep. Purif. Technol. 130,
84e90.
Takeuchi, M., Deguchi, J., Hidaka, M., Sakai, S., Woo, K., Choi, P.-P., Park, J.-K.,
Anpo, M., 2009. Enhancement of the photocatalytic reactivity of TiO2 nanoparticles by a simple mechanical blending with hydrophobic mordenite
(MOR) zeolite. Appl. Catal. B Environ. 89, 406e410.
Takeuchi, M., Kimura, T., Hidaka, M., Rakhmawaty, D., Anpo, M., 2007. Photocatalytic
oxidation of acetaldehyde with oxygen on TiO2/ZSM-5 photocatalysts: effect of
hydrophobicity of zeolites. J. Catal. 246, 235e240.
Tang, Y., Zhang, G., Liu, C., Luo, S., Xu, X., Chen, L., Wang, B., 2013. Magnetic TiO2graphene composite as a high-performance and recyclable platform for efcient
photocatalytic removal of herbicides from water. J. Hazard. Mater. 252e253,
115e122.
Tayade, R.J., Kulkarni, R.G., Jasra, R.V., 2007. Enhanced photocatalytic activity of TiO2
-coated NaY and HY zeolites for the degradation of methylene blue in water.
Ind. Eng. Chem. Res. 46, 369e376.

235

Toledo, T.V., Bellato, C.R., Pessoa, D.K., Fontes, M.P.F., 2013. Removal of chromium
(VI) from aqueous solutions using the calcined magnetic composite
hydrotalcite-iron oxide: kinetic and thermodynamic equilibrium studies. Qum.
Nova 36, 419e425.
rio, R.H. do, Neto, J. de O.M., 2011. Adsorption of
Toledo, T.V., Bellato, C.R., Rosa
arsenic(V) by the magnetic hydrotalcite - iron oxide composite. Qum. Nova 34,
561e567.
Tong, D.S., Liu, M., Li, L., Lin, C.X., Yu, W.H., Xu, Z.P., Zhou, C.H., 2012. Transformation
of alunite residuals into layered double hydroxides and oxides for adsorption of
acid red G dye. Appl. Clay Sci. 70, 1e7.
Walcarius, A., Mercier, L., 2010. Mesoporous organosilica adsorbents: nanoengineered materials for removal of organic and inorganic pollutants. J. Mater.
Chem. 20, 4478.
Wang, X., Wu, P., Huang, Z., Zhu, N., Wu, J., Li, P., Dang, Z., 2014. Solar photocatalytic
degradation of methylene blue by mixed metal oxide catalysts derived from
ZnAlTi layered double hydroxides. Appl. Clay Sci. 95, 95e103.
Yao, Y., Xu, F., Chen, M., Xu, Z., Zhu, Z., 2010. Adsorption behavior of methylene blue
on carbon nanotubes. Bioresour. Technol. 101, 3040e3046.
You, Y., Zhao, H., Vance, G.F., 2002. Surfactant-enhanced adsorption of organic
compounds by layered double hydroxides. Colloids Surfaces A Physicochem.
Eng. Asp. 205, 161e172.
Yu, Z., Chuang, S.S.C., 2007. Probing methylene blue photocatalytic degradation by
adsorbed ethanol with in situ IR. J. Phys. Chem. C 111, 13813e13820.
Zhang, W.H., Guo, X.D., He, J., Qian, Z.Y., 2008. Preparation of Ni(II)/Ti(IV) layered
double hydroxide at high supersaturation. J. Eur. Ceram. Soc. 28, 1623e1629.
Zhao, M., Tang, Z., Liu, P., 2008. Removal of methylene blue from aqueous solution
with silica nano-sheets derived from vermiculite. J. Hazard. Mater. 158, 43e51.
~o de compostos
Ziolli, R.L., Jardim, Wilson F., 1998. Mecanismo de fotodegradaa
^nicos catalisada por TiO2. Qum. Nova 21, 319e325.
orga

You might also like