Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Fe3O4 Magnetic Nanoparticles: Characterization and Performance

Exemplified by the Degradation of Methylene Blue in the Presence of


Persulfate
Chang-Mao Hung, Chiu-Wen Chen, Yu-Zhe Jhuang, Cheng-Di Dong*
Department of Marine Environmental Engineering, National Kaohsiung Marine University, 142 Haijhuan Road,
Nanzih District, Kaohsiung City 81157, Taiwan

Abstract:
In this study, the oxidation of methylene blue (MB) over iron oxide magnetic nanoparticles (Fe3O4), which
effectively activates persulfate anions (S2O82−) to form sulfate free radicals (SO4−•), was explored. In addition,
the effect of the initial pH, sodium persulfate (Na2S2O8, PS) concentration, and Fe3O4 content on the
decolorization of MB was investigated. The results revealed that the decolorization rate increased when the
persulfate concentration increased from 0.03 to 0.12 g/L and the Fe3O4 content from 0.1 to 0.8 g/L. Therefore,
the Fe3O4 nanoparticles enhanced the decolorization of MB. The catalyst was analyzed using cyclic voltammetry
(CV), three-dimensional excitation-emission fluorescence matrix (EEFM) spectroscopy, and zeta potential
measurements. The CV spectra indicated that a reversible redox reaction may explain the high catalytic activity
of the catalyst. EEFM was used to evaluate the yield of a fresh Fe3O4 catalyst, and two peaks were observed at
EX/EM wavelengths of 230/300 nm and 270/300 nm. Furthermore, the structure and surface morphology of the
catalyst were characterized using X-ray diffraction (XRD) and environmental scanning electron microscopy
(ESEM)–energy dispersive spectroscopy (EDS), respectively. The XRD result confirmed the existence of Fe3O4
in the catalyst. ESEM was used to determine the Fe3O4 particle size, indicating a high degree of nanoparticle
dispersion.

Keywords: methylene blue (MB); sodium persulfate (Na2S2O8, PS); Fe3O4 nanoparticle materials; three-
dimensional excitation-emission fluorescence matrix (EEFM) spectroscopy.

Introduction or yield contaminated sludge or adsorbents that


In recent years, synthetic dye emission has require further treatment to satisfy stringent discharge
become a severe environmental problem worldwide, regulations.
involving the exposure of animals and human to To overcome these drawbacks, the most
aqueous toxic chemical and hazardous materials (1-3). promising and widely used advanced water treatment
Dyes with a complex aromatic molecular structure are techniques for controlling dye effluents are sulfate-
emitted through various processes, including those radical-based (SO4–•) advanced oxidation processes
used in leather and textile manufacturing and (AOPs) for oxidizing persulfate that produce a
pharmaceutical and food production (4). Synthetic powerful oxidizing species known as SO4–• (7,8). The
dyes can be converted to highly mutagenic and persulfate anion (S2O82−) is an oxidant that is highly
carcinogenic compounds under ambient conditions soluble in aqueous solutions and thermodynamically
and are a potential public health and environmental more stable (E0 = 2.01 V) than H2O2 (E0 = 1.76 V);
hazard (5). Hence, removing industrial dyes from waste however, direct reactions between persulfate and most
streams and controlling dye emissions are of global reductants are slow. When appropriately activated,
concern. In recent years, several approaches have been persulfate decomposes into a highly reactive species,
developed to biologically, physically, and chemically SO4–• (9), that SO4–• has a high redox potential (E0 =
treat dyes, and methods such as bio-degradation, 2.6 V) near that of hydroxyl radicals (2.8 V), which
filtration, coagulation, ultraviolet-light irradiation, are strong one-electron oxidants that can degrade
ozonation, adsorption, and catalytic oxidation have contaminants. Generally, persulfate can be chemically
been used for removing dyes from water (6). However, activated by transition metals (Mn+):
2  2
these approaches either result in phase transformation S2O8  M n  SO4  SO4  M ( n1)
(1)
This critical heterogeneous catalytic process has
*Corresponding author; E-mail: cddong@mail.nkmu.edu.tw attracted considerable attention in materials chemistry,

ISSN 1203-8407 © 2016 Science & Technology Network, Inc. J. Adv. Oxid. Technol. Vol. 19, No. 1, 2016 43
C-M Hung et al.

and the effectiveness of the active persulfate process Magnetite (Fe3O4) is a promising heterogeneous
has been improved by using high-performance ferrous catalyst because of its natural abundance, low-cost,
ions, which potentially shorten the reaction time under and environmentally friendly properties and can be
mild operating conditions. used as an activator of persulfate (15, 16). Recently,
2  2
applying magnetite to oxidize organic compounds in
S2O8  Fe 2  SO4  SO4  Fe 3 (2) dye-contaminated wastewater has attracted considerable
Typically, sulfate-radical-based AOPs selectively interest (7). Fe3O4 improves the persulfate oxidation
produce carbon dioxide and involve the oxidation of properties by accelerating the degradation rate through
the formation of SO4–•. Moreover, to the best of our
small molecules through aqueous phase oxidation.
Various types of catalysts that oxidize dye- knowledge, few studies have investigated using Fe3O4
contaminated wastewater have been studied. For for the catalytic liquid-phase oxidation of MB.
example, Yan et al. (10) reported that the complete Accordingly, in the current study, a Fe3O4 catalyst was
removal of sulfamonomethoxine (SMM) through used to oxidize MB, a brightly colored cationic
degradation was achieved within 15 min by using 2.4 thiazine dye, in a sulfate-radical-based AOP process
mmol/L Fe3O4 magnetic nanoparticles as the under various conditions, and the effectiveness of the
heterogeneous activator of 1.2 mmol/L persulfate. catalyst was examined.
Nguyen et al. (11) used Fe2MnO4/AC-H catalysts to Excitation-emission fluorescence matrix (EEFM)
oxidize methyl orange at pH 3.0 and found that spectroscopy was used across a range of excitation
activated carbon can be treated with HNO3 to and emission wavelengths to effectively understand
functionalize the carbon surface through the formation catalyst characteristics during the catalytic process.
of carboxyl groups. Additionally, a recent study (12) Thus, this study investigated the effect of various
described the efficient catalytic oxidation of the azo parameters, such as Fe3O4 content, PS concentration
dye orange G (OG) in an aqueous stream achieved by and pH, on a Fe3O4 nanoparticle catalyst, applied to
using a Fe2+/persulfate reagent in the temperature range activate persulfate and oxidize MB, to establish the
of 293–313 K at a pH of 3.5. The OG degradation optimal conditions for the catalytic system. Moreover,
increased when the amount of Fe2+ and persulfate the catalysts were characterized using cyclic
increased. Yang and co-workers (13) reported the voltammetry (CV), zeta potential measurements, X-
persulfate oxidative degradation of the azo dye acid ray diffraction (XRD), and environmental scanning
orange 7 (AO7) in the presence of suspended granular electron microscopy (ESEM)–energy dispersive
activated carbon (GAC) with a remarkable synergistic spectroscopy (EDS).
effect in the GAC/sodium persulfate (PS) combined
system. Ghauch et al. (14) studied the thermal Materials and Methods
discoloration of methylene blue (MB) by using an Materials and Chemicals
efficient and affordable process involving activating MB was purchased from Kojima Chemicals Co.
persulfate at 303–343 K to form powerful sulfate Ltd (Japan) and used without further purification.
radicals in an aqueous solution. They separated and Fe3O4 powders with particle sizes of 50–100 nm and a
identified the main intermediates by using liquid Brunauer–Emmett–Teller surface area of 20–50 m2/g
chromatography–mass spectrometry and proposed a were obtained from Alfa Aesar Co. (USA). PS
plausible degradation pathway for MB. Zhu et al. (9) (Na2S2O8, 98%) was purchased from Showa Chemical
examined the oxidation of methyl orange in an aqueous Industry Co. (Japan). All other reagents were of
solution over core-shell Fe-Fe2O3 nanostructures analytical grade. Deionized water was used for solution
(FNs) and found that the conversion of methyl orange preparation in all experiments. Before each run, a
reached 90% in 10 min in an FN/Na2S2O8 process. Lin stock solution of MB was prepared with deionized
et al. (5) converted AO7 to sodium sulfanilamide and water and the initial concentration (C0) was maintained
1-amino-2-naphthol by using Fe3O4–catalyst–activated at 30 mg/L (0.08 mM). Furthermore, H2SO4 or NaOH
peroxydisulfate in an electro/Fe3O4/PDS process. was used to adjust the initial pH (pH0) of an MB
They found that the Fe3O4 particles were stable and solution.
reusable. Wang et al. (4) investigated the degradation
conversion of AO7 by using persulfate activated over Experimental Methods
zero-valent iron catalysts in the presence of ultrasonic The experimental setup included a 500 mL
irradiation. The results of the aforementioned studies cylindrical reactor containing 250 mL of a model dye
revealed the potential of persulfate reactivity to solution. All experiments were conducted at room
degrade organic contaminants. temperature (298 K). A mechanical stirrer, used to

44 J. Adv. Oxid. Technol. Vol. 19, No. 1, 2016


C-M Hung et al.

maintain uniform suspension, was switched on when


persulfate and Fe3O4 powders were added to the 1.0

solution. At various time intervals, 2 mL of the liquid Fe3O4


was sampled and the absorbance was measured at 664 0.8 PS
Fe3O4/PS
nm by using an ultraviolet spectrophotometer (Merck,
Spectroquant Vega 400, Germany). The decolorization 0.6

C/C0
efficiency (η%) was calculated according to the
following equation: (%)  ( A0  A) / A0  100 , 0.4

where A0 and A are the absorbance of the sample at


time instants 0 and t, respectively. 0.2

CV measurements were conducted at room


temperature by using an electrochemical analyzer (CH 0.0
0 10 20 30 40 50
Instrument, CHI 6081D, USA) and a three-electrode
Time (min)
electrochemical cell. The working electrode was a
glassy carbon electrode, and the samples were Figure 1. Plots of the decrease in the MB concentration (C/C0)
versus time for various processes. Experimental conditions:
scanned at a rate of 50 mV/s with potential cycling reaction volume = 250 mL, T = 298 K, pH0 = 6.0, [MB] = 30
between -0.2 and 1.2 V for each analysis. The counter mg/L, [Na2S2O8] = 0.03 g/L, and [Fe3O4] = 0.1 g/L.
electrode was a platinum wire, Ag/AgCl (in saturated
KCl solution) was employed as the reference electrode, Fe3O4/PS. MB could not be effectively degraded when
and H2SO4 (0.5 M) was used as the electrolyte solution. only Fe3O4 was used, indicating the in effect
In this study, fluorescence excitation-emission adsorption of MB under the investigated conditions.
matrix (FEEM) spectroscopy was used to obtain Figure 1 shows that the conversion rate of MB
complete spectral characterization of the catalyst. increased moderately (41%) after 9 min of treatment,
FEEM spectra were obtained using a luminescence during which 0.03 g/L PS was added. The experimental
spectrophotometer (F-4500, Hitachi, Japan) with a results revealed that the MB conversion rate increased
xenon lamp as the excitation source. The emission to 73% at a PS concentration of 0.03 g/L, a Fe3O4
spectra were plotted on the x-axis, and the excitation concentration of 0.1 g/L, and an initial pH value of 6.0
spectra were plotted parallel to the y-axis. The widths at 298 K in the sulfate-radical-based AOP when Fe3O4
of all slits in both the excitation and emission was used as the activator. This observation can be
monochromators were 10 nm. The FEEM comprised explained as follows: in the Fe3O4/PS process, the
60 excitation and 60 emission spectra from 200 to 800 Fe3O4 may influenced the electron mass transfer of the
nm and provided discrete values for the fluorescence reacting species (MB and PS), thereby producing
intensity at 3600 excitation/emission wavelength SO4−• (5,7,14) through the chemical reaction expressed
pairs. Spectral subtraction was performed to remove in Eqs. (3) and (4). These SO4−• enhanced the
the blank spectra of pure water. A zeta potential degradation of MB.
2  2
analyzer (Zetasizer 2000HAS, Malvern, UK) was used S2O 8 e   SO4  SO4 (3)
for determining the particle zeta potential. XRD  2
analysis was performed using a Diano-8536 MB  SO4  [ MB]  SO4 (4)
2−
diffractometer equipped with a CuKα radiation S2O8 can be activated at the Fe3O4 surface through
source. During analyses, samples were scanned from electron transference to produce sulfate radicals that
10° to 80° at a rate of 0.4°/min. We used an ESEM increase the conversion yield. Thus, the combination
system equipped with an energy-dispersive X-ray of S2O82− and Fe3O4 exerts a synergistic effect on the
spectrum analyzer (Quanta 200 FEG, FEI Company, decolorization of MB through oxidation. Recently,
Czech Republic) to determine the morphology of the Ghauch and his collaborators (17) used Fe0 as a PS
catalyst. These aforementioned techniques yielded activator. The use of Fe0 caused iron corrosion
information on the distribution of Fe3O4 nanoparticles products to form in solution, thereby enhancing the
on the catalyst surface. activation process. Yan et al. (10) observed iron ions
(Fe2++ Fe3+) and highly dispersed clusters in iron
Results and Discussions oxide magnetic nanoparticles that were coupled to
Effect of Different Parameters on persulfate. Therefore, using an appropriate metal
Decolorization of Methylene Blue oxide in the persulfate system may effectively provide
Figure 1 presents the results of experiments reactive species. Furthermore, in the present study, as
conducted using Fe3O4 alone, PS alone, and both the reaction proceeded, the degradation kinetics of

J. Adv. Oxid. Technol. Vol. 19, No. 1, 2016 45


C-M Hung et al.

1.0
1.0 (a)
pH 3.0
pH 6.0
pH 9.0
0.01 g/L
0.8
0.8 0.03 g/L
0.09 g/L
0.12 g/L
0.6 0.6
C/C0

0.4 0.4

0.2
0.2

C/C0
0.0
0 10 20 30 40 50 0.0
1.0
Time (min)
(b)
Figure 2. Effect of the initial pH on the MB degradation.
Experimental conditions: reaction volume = 250 mL, T = 298 K, 0.8 0.01 g/L
[MB] = 30 mg/L, [Na2S2O8] = 0.03 g/L, and [Fe3O4] = 0.1 g/L. 0.03 g/L
0.09 g/L
0.6 0.12 g/L
MB during persulfate oxidation over a Fe3O4 catalyst
could be explained by mass-transfer-controlled first-
0.4
order kinetics, which can be determined from a linear
plot of ln(C/C0) versus the reaction time t; the rate
constants (k) were 0.014 and 0.029 min-1 for PS and 0.2

the Fe3O4/PS process, respectively. Hence, activating


persulfate by using Fe3O4 is a more rapid method of 0.0
0 10 20 30 40 50
degrading MB.
Figure 2 presents plots showing the effect of the Time (min)

initial pH on the decolorization of MB over the Fe3O4 Figure 3. Effect of the persulfate concentration on the MB
catalyst in the Fe3O4/PS process. The experiments degradation of (a) Na2S2O8 and (b) Na2S2O8 + Fe3O4.
demonstrated that this effect was appreciable at pH0 Experimental conditions: reaction volume = 250 mL, T = 298 K,
[MB] = 30 mg/L, pH0 = 6.0, and [Fe3O4] = 0.1 g/L.
values near 6.0. Moreover, the removal of MB was
suppressed at pH0 9.0 and 3.0. The difference in color
removal at different pH0 values suggests that the rate radical and the hydroxyl radical is enhanced at a high
of the MB reaction in the persulfate catalytic pH (5). In other words,
 2
oxidation system depended on the pH value. This SO4  OH  SO4  H   0.5O2 (7)
dependence was confined to the duration during which
sulfate radicals generated in the Fe3O4/PS process Accordingly, the Fe3O4/PS process can apparently
reacted with H2O or OH– to produce hydroxyl radicals be performed using a wide range of initial pH values,
that reduced the solution pH. Acidic conditions from 3.0 to 9.0. Because pH0 6.0 was the natural pH
enhance the dissolution of iron powder by producing of the MB solution, it was used in the subsequent
more Fe(II). Thus, more sulfate radicals can be experiments.
formed by Fe(II) through the catalytic activation of Figure 3 shows the PS concentration effect for the
persulfate. However, excessive Fe(II) can act as using PS alone and both Fe3O4/PS on MB
scavengers of sulfate radical. In other words, decomposition. The decolorization rate of MB
 2 increased slowly when only the PS concentration
Fe 2  SO4  SO4  Fe 3 (5)
increased from 0.01 to 0.12 g/L (Fig. 3a). The total
Under alkaline conditions, more hydroxyl radicals conversion efficiency of MB was 84% at a PS
are generated from sulfate radicals according to the concentration of 0.12 g/L (2.0 mM) and a Fe3O4 dose
following reaction: of 0.1 g/L (0.4 mM) (Fig. 3b). The [MB]0/[PS]0 molar
 2 ratio was 0.04. A higher PS concentration causes an
SO4  OH   SO4  OH (6) increased degradation of MB. In addition, a sufficient
However, the oxidation potential of the hydroxyl amount of persulfate produces adequate SO4−• to
radical may decrease with an increase in the pH. In enhance the activation activity of Fe3O4, thus leading
addition, the scavenging reaction between the sulfate to the enhanced degradation of MB (9). The

46 J. Adv. Oxid. Technol. Vol. 19, No. 1, 2016


C-M Hung et al.

decomposition of MB over the Fe3O4/PS process 1.0


increased with the amount of sulfate radicals because 0.1 g/L
of an increased direct electron transfer between the 0.8
0.8 g/L
1.6 g/L
components of the Fe(III)/Fe(II) redox pair (7,16). 2.4 g/L
Consequently, the increased PS concentration increased
the production of SO4−• and provided a strong driving
0.6

C/C0
force for oxidation. The oxidation reaction can also be
0.4
explained by the synergistic effect of Fe3O4 materials
and SO4−•. However, after 18 min of reaction, the
decolorization efficiency was almost the same as that 0.2

for PS concentrations from 0.03 to 0.12 g/L,


indicating that the amount of SO4−• produced was 0.0
0 10 20 30 40 50
sufficient for achieving decolorization. Compared
Time (min)
with the situation in Fig.1, in which the rate of MB
conversion increased moderately (41%) after 9 min of Figure 4. Effect of iron oxide addition on the MB degradation.
Experimental conditions: reaction volume = 250 mL, T = 298 K,
treatment (during which 0.03 g/L of PS was added), [MB] = 30 mg/L, pH0 = 6.0, and [Na2S2O8] = 0.03 g/L.
the conversion after 18 min of reaction in the
Fe3O4/PS process was nearly 60%. A higher PS decreasing the reaction stoichiometric efficiency after
concentration (>0.12 g/L) may not enhance MB the formation of S2O82− and Fe(III). This observation
oxidation because of chain termination between is consistent with previously reported results (10).
sulfate radicals. In this decolorization system, MB
constituents exhibited a considerable influence and led Characterization of Methylene Blue for Types of
to the gradual reduction of PS of experimental run. Fe3O4
The treated sample had over 70% of the PS still Figure 5 shows the electrochemical behavior of
remaining at an initial PS dose of 0.03 g/L in treated freshly prepared and used Fe3O4 catalysts. The CV
sample during 45 min of experimental run. The results plot reveals that the freshly prepared Fe3O4 catalyst
indicated that adequate PS must be available because had a higher reversible redox capacity and oxidation
it is the source of the SO4−• responsible for the current density than did the used Fe3O4 catalyst; the
degradation of MB. Therefore, low PS consumption is well-marked state was at 0.55 and 0.61 V for the
required for achieving an acceptable reaction freshly prepared and used Fe3O4 catalysts, respectively.
stoichiometric efficiency, which is necessary for This reversible redox ability of the fresh catalyst could
driving the reaction until MB removal is complete explain its high activity. By contrast, the used Fe3O4
(18). This finding was similar to that of Wang et al. catalyst produced negligible oxidation–reduction
(4), who experimented on the persulfate oxidation of currents in the potential window. Lambrou et al. (20)
AO7 in wastewater. showed that Fe3O4 had the most active phase in the
Figure 4 shows the effect of Fe3O4 on the MB catalytic reaction because Fe3O4 strongly promotes
removal efficiency as a function of time in the oxygen storage. This finding suggests that Fe3O4 can
Fe3O4/PS process. Generally, the degradation of MB promote PS activation during MB oxidation.
increased sharply when the Fe3O4 increased from 0.1 Furthermore, the generation of Fe3O4 species was
to 1.6 g/L. Specifically, Figure 4 shows that using 1.6 evidenced by the reduction peaks near 0.6 V (21,22).
g/L of Fe3O4 as an activator for the effective Therefore, MB is believed to have been adsorbed onto
degradation of MB yielded a maximum conversion the surface of the catalyst before the oxidation–
rate of 70%. This rate can be attributed to the presence reduction cycle reaction commences at the iron oxide
of a sufficient quantity of Fe(II), which serves as an active sites.
electron donor, for catalytically decomposing persulfate To further elucidate the reactive properties of the
anions (19). However, the MB degradation rate catalyst, EX/EM plot positions obtained from the
increased slightly when the Fe3O4 dose was increased EEFM of the fluorescence spectrometer were used to
to 2.4 g/L, and a further increase in the reaction time generate effective data for the preliminary estimation
little affected the decolorization rate. This is because of the catalyst properties. Figure 6a shows the
the excess Fe3O4 may have reduced PS consumption, fingerprint of the three-dimensional EEFM for a fresh
and the produced Fe(II) could have scavenged SO4−• Fe3O4 nanoparticle; two significant excitation/emission
through the reaction expressed in Eq. (3). Moreover, a plots were generated at 230/300 and 270/300 nm.
high Fe(II) dose could cause SO4−• quenching, thereby Figure 6b shows the fingerprint of the EEFM for the

J. Adv. Oxid. Technol. Vol. 19, No. 1, 2016 47


C-M Hung et al.

(a) Fresh
2 (b) Used
Current (mA)

-2

-4

-0.4 -0.2 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4

Potential (V)

Figure 5. CV profiles in a 0.5 M H2SO4 solution for the Fe3O4


nanoparticle used for degrading MB at a scan rate of 50 mV/s (a)
before and (b) after the catalytic test.

Fe3O4 catalyst after the activity test; two substantial


excitation/emission plots at 290/300 and 270/384 nm
were observed. The emission wavelength of the used
Fe3O4 nanoparticle increased remarkably to 384 nm
because of effective electron transfer from the excited
state to the Fe3O4 nanoparticle conduction band. This
result is similar to that reported by Alveroğlu et al.
(23), who determined that magnetite nanoparticle
wavelengths were 380 and 405 nm in an emission
spectrum. These excitation/emission plots can be
described by considering the metal-enhanced
fluorescence effect, in which fluorescence is
associated with iron and ionic iron species of Fe3O4
Figure 6. Contour plots of the excitation–emission fluorescence
during the catalytic reaction. Moreover, the observed matrix for the Fe3O4 nanoparticle used for degrading MB (a)
changes in the excitation/emission plots may be before and (b) after the catalytic test.
attributable to the presence of surface functional
groups and surface energy traps that become emissive analysis confirmed the existence of the Fe3O4 state in
upon stabilization as a result of surface passivation the catalyst. The XRD patterns of the freshly prepared
after the activity test (24). Fe3O4 and used Fe3O4 after the Fe3O4/PS process were
The particle zeta potential is a crucial physical somewhat similar. The dominant Fe3O4 diffraction
property of a catalyst. Figure 7 displays the variation peaks detected at 2θ values of approximately 18.4o,
of the zeta potential for freshly prepared and used 30.3 o, 35.6 o, 37.3 o, 43.3 o, 53.6 o, 57.2 o, 62.8 o, 71.1 o,
catalysts. The variation reveals that the surface charge and 74.2o were related to the (111), (220), (311),
of the catalyst changed during exposure to the (222), (400), (422), (511), (440), (620), and (533)
catalytic oxidation environment. The mean particle planes, respectively. Thus, the diffraction peaks of
zeta potential changed from -35.6 mV for the freshly nanoparticles corresponded with those of Fe3O4
prepared Fe3O4 catalyst to -20.9 mV for the reacted confirming the high crystallinity of the samples. This
Fe3O4 catalyst. The zeta potential changed inappreci- result is consistent with those of Yan et al. and Dong
ably, even after decolorization through persulfate et al. (10,25). The results of the XRD analysis
addition. The weak electrostatic repulsion between the indicated that Fe3O4 active sites were formed on the
surface charges of a solid particle and the tendency of Fe3O4 catalyst. Therefore, MB was adsorbed on the
the particles to flocculate after a catalytic reaction surface of the catalyst before undergoing a catalytic
may explain the experimental results. reaction at the Fe3O4 active sites. On the basis of XRD
Figure 8 shows the XRD pattern of the fresh and patterns, the crystal size of the particles can be
used catalysts; the catalyst changed during exposure to evaluated by using the following Debye–Scherrer
the catalytic environment. Furthermore, the XRD equation:

48 J. Adv. Oxid. Technol. Vol. 19, No. 1, 2016


C-M Hung et al.

8e+5 ●
● Fe3O4
(a) Fresh
(b) Used

6e+5

● ●

Total Counts

(a) Fresh

Intensity
● ● ●
4e+5 ● ●

2e+5

● ● (b) Used


● ● ● ● ●
0
-200 -100 0 100 200

Zeta Potential (mV) 10 20 30 40 50 60 70 80

Figure 7. Zeta potential of the Fe3O4 nanoparticle on the MB 2θ


degradation (a) before and (b) after the catalytic test. Figure 8. XRD pattern of the Fe3O4 nanoparticle used for
degrading MB (a) before and (b) after the catalytic test.

D  K /(  cos  ) (8)
The performance of the Fe3O4 catalyst was poor
where K is the Scherrer factor (0.89), λ is the X-ray mainly because of iron corrosion, which caused
wavelength (0.15418 nm), β is the peak full width at nanoparticles to agglomerate on the catalyst surface.
half-maximum (FWHM), and θ is the Bragg In future studies, recycling experiments using Fe3O4
diffraction angle. According to the most intense peak catalysts should be systematically conducted.
(311), the average size of the particles was calculated
to be 26.9 nm. Conclusions
Figure 9 presents ESEM-EDS images showing the The results of this study indicated that in a
surface morphology changes of the Fe3O4 nanoparticle solution with an MB concentration of 30 mg/L, a PS
catalyst. These data provide crucial information concentration of 0.03 g/L, a Fe3O4 dosage of 0.1 g/L,
regarding the surface structure of the catalyst during and an optimal initial pH of 6.0, MB can be
the MB oxidation reaction. ESEM observations of the decolorized in 45 min. The decolorization rate can be
nanoparticles indicated that the Fe3O4 catalyst formed substantially enhanced by using a combination of
small spherical particles. Furthermore, these experi- Fe3O4 and sulfate radicals. Overall, the Fe3O4 nano-
mental results confirm that the dispersion of the particle catalyst was crucial in the MB catalytic
catalyst increased the efficiency of the MB catalysis. process. CV curves revealed the reversibility of the
Figure 9a presents a more crystalline catalyst surface redox, which may explain the high activity of the
compared with that in Figure 9b. The small crystal catalyst. According to a fluorescence spectrometry
phases may explain the high activity of the catalysts; evaluation, a fresh Fe3O4 nanoparticle catalyst yielded
the catalysts were nearly nanoscale in size (approxi- fluorescence plots at 230/300 nm and 270/300 nm,
mately 80 nm) and exhibited high dispersion. which were identified at room temperature. Moreover,
Specifically, disaggregated phases formed when the the loss of activity of the Fe3O4 nanoparticle catalyst
catalyst surface aged or was poisoned because of was attributed to either the over-oxidation of the
plugging, implying that the porosity of the particles catalyst surface or the inhibition of the surface by
changed (Figure 9b). EDS provided evidences of the reaction intermediates. Specifically, while past studies
presence of iron species in the catalyst. Moreover, the have suggested that the role of Fe3O4 is confirmed to
elemental-composition analysis of the test catalyst providing active sites for the reaction during a catalytic
surfaces revealed that the leaching of metal ions from run, in the current study, the Fe3O4 nanoparticle
the Fe3O4 catalyst varied slightly. For the Fe3O4 catalyst was involved in the persulfate oxidation of
catalyst, the elemental percentage of Fe atoms was MB. Furthermore, according the experimental data,
70%–67% and that of O atoms was 30%–33%. Less the Fe3O4 nanoparticles performed favorably in
iron and more oxygen were observed after the PS environmental treatment applications involving the
chemical activation test, possibly because the MB removal of MB, and are therefore expected to be useful
oxidation was caused by PS on the surface of Fe3O4 for improving industrial plant effluents to satisfy
(14). Fe3O4 emerged from the catalyst, which likely increasingly stringent regulations on azo dyes discharges
exposed the active site on the nanoparticle surface. and for achieving environmental sustainability. In the

J. Adv. Oxid. Technol. Vol. 19, No. 1, 2016 49


C-M Hung et al.

Figure 9. ESEM–EDS photographs of the Fe3O4 nanoparticle used for degrading MB (a) before and (b) after the catalytic test.

future, we intend to use a mass spectrometer analyzer (2) Hsieh, W.P.; Pan, J.R.; Huang, C.; Su, Y.C.; Juang,
to identify intermediate transformation products and Y.J. Sci. Total Environ. 2010, 408, 672-679.
final products in the oxidation reaction of MB over (3) Hou, Y.; Qu, J.; Zhao, X.; Lei, P.; Wan, D.; Huang,
Fe3O4/PS systems. C. Sci. Total Environ. 2009, 407, 2431-2439.
(4) Wang, X.; Wang, L.; Li, J.; Qiu, J.; Cai, C.; Zhang,
Acknowledgments H. Sep. Purif. Technol. 2014, 122, 41-46.
The authors would like to thank the National (5) Lin, H.; Zhang, H.; Hou, L. J. Hazard. Mater. 2014,
Science Council of the Republic of China, Taiwan, for 276, 182-191.
financially supporting this study under Contract Nos. (6) Yang, A.; Huang, C.; Wei, B.; Zhang, Z. Water Sci.
MOST 103-2221-E-022-001-MY3 and 103-2221-E- Technol. 2014, 695, 1094-1100.
022-002-MY3. The author is also grateful to Prof. W. L. (7) Xu, X.R.; Li, S.; Hao, Q.; Liu, J.L.; Yu, Y.Y.; Li,
Lai of the Department of Environmental Resource H.B. Int. J. Environ. Bioener. 2012, 1, 60-81.
Management, Tajen University of Science and (8) Tsitonaki, A.; Petri, B.; Crimi, M.; Mosbaek, H.;
Technology, for his support and discussions, and to Siegrist, R.L.; Bjerg, P.L. Critical Reviews in
Prof. C. P. Huang of the University of Delaware for Environ. Sci. Tech. 2010, 40, 55-91.
his valuable comments on the initial version of this (9) Zhu, L.; Ai, Z.; Ho, W.; Zhang, L. Sep. Purif. Tech.
manuscript. 2013, 108, 159-165.
(10) Yan, J.; Lei, M.; Zhu, L.; Anjum, M.N.; Zou, J.;
References Tang, H. J. Hazard. Mater. 2011, 186, 1398-1404.
(1) Juang,Y.; Nurhayati, E.; Huang, C.; Pan, J.R.; (11) Nguyen, T.D.; Phan, N.H.; Do, M.H.; Ngo, K.T. J.
Huang, S. Sep. Purif. Technol. 2013, 120, 289-295. Hazard. Mater. 2011, 185, 653-661.

50 J. Adv. Oxid. Technol. Vol. 19, No. 1, 2016


C-M Hung et al.

(12) Xu, X.R.; Li, X.Z. Sep. Purif. Tech. 2010, 72, 105- (20) Lambrou, P.S.; Efstathiou, A.M. J. Catal. 2006,
111. 240, 182-193.
(13) Yang, S.Y.; Yang, X.; Shao, X.T.; Niu, R.; Wang, (21) Wu, F.; Huang, R.; Mu, D.; Wu, B.; Chen, S. ACS
L.L. J. Hazard. Mater. 2011, 186, 659-666. Appl. Mater. Interfaces 2014, 6, 19254-19264.
(14) Ghauch, A.; Tuqan, A.M.; Kibbi, N.; Geryes, S. (22) Karunakaran, C.; Jayabharathi, J.; Sathishkumar, R.
Chem. Eng. J. 2012, 213, 259-271. Spectrochim. Acta Part A. 2013, 110, 151-156.
(15) Si, S.; Kotal, A.; Mandal, T.K.; Giri, S.; Nakamura, (23) Alveroğlu, E.; Sözeri, H.; Baykal, A.; Kurtan, U.;
H.; Kohara, T. Chem. Mater. 2004, 16, 3489-3496. Şenel, M. J. Mater. Sci. 2013, 1037, 361-366.
(16) Tang, S.C.N.; Lo, I.M.C. Water Res. 2013, 4, 2613- (24) He, X.; Liu, Y.; Li, H.; Huang, H.; Liu, J.; Kang, Z.;
2632. Lee, S.T. J. Colloid Inter. Sci. 2011, 356, 107-110.
(17) Ghauch, A.; Ayoub, G.; Naim, S. Chem. Eng. J. (25) Dong, Y.C.; Ma, R.G.; Hu, M.J.; Cheng, H.;
2013, 228, 1168-1181. Tsang, C.K.; Yang, Q.D.; Li, Y.Y.; Zapien, J.A. J.
(18) Ayoub, G.; Ghauch, A. Chem. Eng. J. 2014, 256, Solid State Chem. 2013, 201, 330-337.
280-292.
(19) Anotai, J.; Masomboon, N.; Chuang, C.L.; Lu, M. Received for review February 16, 2015. Revised manuscript
C. Water Sci. Tech. 2011, 63, 1434-1440. received July 7, 2015. Accepted July 14, 2015.

J. Adv. Oxid. Technol. Vol. 19, No. 1, 2016 51

You might also like