You are on page 1of 12

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/323387116

Photocatalytic degradation of methylene blue dye by iron oxide (α-Fe2O3)


nanoparticles under visible irradiation

Article  in  Journal of Materials Science Materials in Electronics · February 2018


DOI: 10.1007/s10854-018-8819-4

CITATIONS READS

2 1,156

5 authors, including:

Abdelmajid Lassoued Mohamed Saber Lassoued


University of Gabès University of Gabès
22 PUBLICATIONS   116 CITATIONS    25 PUBLICATIONS   92 CITATIONS   

SEE PROFILE SEE PROFILE

Brahim Dkhil Salah Ammar


CentraleSupélec The Faculty of Science of Gabes
300 PUBLICATIONS   7,494 CITATIONS    98 PUBLICATIONS   1,094 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Polymeric Materials View project

photocatalysis View project

All content following this page was uploaded by Abdelmajid Lassoued on 25 February 2018.

The user has requested enhancement of the downloaded file.


Journal of Materials Science: Materials in Electronics
https://doi.org/10.1007/s10854-018-8819-4

Photocatalytic degradation of methylene blue dye by iron oxide


(α-Fe2O3) nanoparticles under visible irradiation
Abdelmajid Lassoued1,2 · Mohamed Saber Lassoued1 · Brahim Dkhil2 · Salah Ammar1 · Abdellatif Gadri1

Received: 16 December 2017 / Accepted: 23 February 2018


© Springer Science+Business Media, LLC, part of Springer Nature 2018

Abstract
Iron oxide (α-Fe2O3) nanoparticles were precipitated from iron (III) chloride hexahydrate (­ FeCl3·6H2O) at the reaction
temperature 80 °C for 3 h and were further calcinated at 600, 700, 750, 800 and 850 °C for 4 h. The effect of calcination
temperature on the structural, morphological, optical, magnetic and photocatalytic activity was investigated. XRD data
revealed a rhombohedral (hexagonal) structure with the space group R-3c in all samples. The calcination temperature was
significantly influenced the particle size and morphological properties of hematite nanoparticles and the particle size was
increased from 18 to 37 nm with the increased of the calcination temperature. The synthesized nanoparticles were roughly in
spherical morphology was confirmed by TEM and SEM. FT-IR confirms the phase purity of the nanoparticles synthesized.
Raman spectroscopy was used not only to prove that we have synthesized pure α-Fe2O3 but also to identify their phonon
modes. The thermal behavior of compound was studied by using TGA/DTA results: the TGA showed three mass losses,
whereas DTA resulted in three endothermic peaks. Besides, the optical investigation revealed that samples have an optical
gap of about 2.1 eV and that this value varies as a function of the calcination temperature. The products exhibited the attrac-
tive magnetic properties with high saturation magnetization, which were examined by (VSM). The photocatalytic activities
of the samples were studied based on the degradation of methylene blue as a model compound, where the results showed
that hematite (α-Fe2O3) a good photocatalytic activity.

1 Introduction (hematite), β-Fe2O3, γ-Fe2O3 (maghemite), ε-Fe2O3 [4–7].


Hematite (α-Fe2O3) has a corundum structure showing weak
The three main types of iron oxide nanoparticles (FeO, ferromagnetism below 950 K. β-Fe2O3 has a bixbyite struc-
­Fe2O3 and ­Fe3O4) are considered to be important materi- ture showing antiferromagnetism below 119 K. Maghemite
als due to their biocompatibility, non-toxicity, catalytic (γ-Fe2O3) has a spinel structure showing ferrimagnetism
activity, low-cost and environmentally friendly nature. below 928 K. ε-Fe2O3 has an orthorhombic structure show-
Nanostructured iron oxides have attracted the attention of ing ferrimagnetism below 495 K. Hematite and maghemite
several researchers because of their technological utiliza- phases have been well-studied and extensively used whereas
tions such as media, catalysis, biomedicine, biotechnol- β- and ε-Fe2O3 are considered to be rare phases [8].
ogy, magnetic storage, biosensors, etc due to their large α-Fe2O3 is the most stable ferric oxide under ambient con-
surface area, robustness and availability [1–3]. ­Fe2O3, the ditions and has n-type semiconducting properties. It in the bulk
ferric oxide has four crystallographic phases, viz α-Fe2O3 form undergoes two types of magnetic transition: the Neel
temperature ­(TN) at about 950 K and the Morin temperature
­(TM) at about 260 K. The nanosized hematite can display three
* Abdelmajid Lassoued transition (critical) temperatures: the Neel ­(TN), Morin ­(TM)
abdelmajidlassoued15@gmail.com and blocking ­(TB) temperatures [9]. Hematite has antiferro-
1
Unité de Recherche Electrochimie, Matériaux et magnetism with canting ferromagnetic responses at room tem-
Environnement UREME (UR17ES45), Faculté des Sciences perature. Also, it has a complex defect structure having defect
de Gabès, Université de Gabès, Cité Erriadh, 6072 Gabès, species such as oxygen vacancies and F ­ e2+ interstitials [10]. In
Tunisia addition, it has attracted the attention of researchers in recent
2
Laboratoire Structures, Propriétés et Modélisation des years due to its application in catalysis, magnetic recording
Solides CentraleSupélec, CNRS-UMR8580, Université (storage) devices, sorbents, rechargeable lithium batteries,,
Paris‑Saclay, 91190 Gif‑sur‑Yvette, France

13
Vol.:(0123456789)
Journal of Materials Science: Materials in Electronics

pigments, gas sensors, biomedical materials, photocatalyst in Double distilled water was used for all experiments. Raw
­N2 fixation, optical devices, solar cells and electromagnetic materials are: iron (III) chloride hexahydrate (­ FeCl3·6H2O)
devices [11–17]. To date, a variety of techniques have been (98%, Sigma-Aldrich), ammonia hydroxide (­ NH4OH) (98%,
used to prepare hematite in powder or thin film form: polyol Sigma-Aldrich) and ethanol ­(C2H6O) (99%, Sigma-Aldrich).
method [18], sol–gel method [19–21], spray pyrolysis [22–24],
hydrothermal technique [25], chemical vapor deposition [26], 2.2 Synthesis of hematite nanoparticles
pulsed laser deposition [27], co-precipitation [28, 29] and
high vacuum evaporation [28]. Though mono-disperse hema- Hematite nanoparticles were synthesized with the chemical
tite particles with controlled shapes and sizes are obtainable, precipitation method .In this procedure, aqueous solution
it’s not easy enough to scale up the synthesis in view of the was prepared by dissolving 4.052 g of iron (III) chloride
complex reaction processes with time consuming preparation hexahydrate ­(FeCl3·6H2O) in 100 mL of double distilled
conditions, high-energy consumption, and low product yields. water. 50  mL aqueous solution of 2  M of N ­ H4OH, was
As a matter of fact, finding a simple synthetic pathway for used as the precipitating agent. Base solution (­NH4OH)
the production of fairly mono-disperse α-Fe2O3 nanoparticles, was added gradually drop wise to maintain a pH value of
remains a true challenge. The chemical precipitation method 11. The reaction vessel was heated up to 80 °C temperature
proves to be the best owing to its simplicity, effectiveness and under magnetic stirring for 3 h. The resulting precipitations
its low cost as it requires no special equipment or additives. were collected and centrifuged at 6000 rpm and then washed
Liu et al. [30, 31] successfully obtained nearly spherical nano- with distilled water and ethanol for several times and finally
particles of α-Fe2O3 with a diameter of 60–80 nm. dried in air at 80 °C and calcined at (600, 700, 750, 800 and
Controlling the particle size and morphology is extremely 850 °C) for 4 h.
significant in the preparation of hematite nanoparticles, since
their properties of the ultrafine particles depend largely on 2.3 Characterization
the morphology, particle size, and dispersion of the prod-
ucts [32]. So far, several researchers have focused on the The X-ray diffraction patterns of the samples were identi-
investigation of the factors influencing the particle shape and fied using German Bruker D2 PHASER X-ray diffractometer
size of the synthesized hematite nanoparticles. These factors with ­CuKα radiation (1.5418 Å) as a source. The intensity
have been fully-detailed in literature, such as the reactant data were collected over the range of 20°–80° using a step
concentration, time and temperature reaction, pH solution, scan mode (0.001°/s). The TEM micrographs were obtained
calcination temperature, ionic strength, the anions, the sur- on a JEOL 2011 transmission electron microscope with an
factant, and the nature of iron salts [33–38]. accelerating voltage of 200 kV. Micro structural charac-
In the present study, we aimed at obtaining iron oxide teristics such as morphology and particle size of samples
(α-Fe2O3) nanoparticles with varying of calcination tem- were analyzed through SEM with KYKY-EM3200, 25KV
perature (600, 700, 750, 800 and 850 °C) using the precipita- type. FT-IR spectra of samples were explored by a NICO-
tion technique. The main advantage of precipitation method LET IR200 FT-IR spectrometer with transmission from
is the good productivity and the relatively low temperature 4000 to 400 cm−1. The Raman spectra were collected using
required which limits the formation of large grains. The a Renshaw in Via Raman microscope with a 50× objec-
nanoparticles are well characterized for their structural, mor- tive coupled to a 632.8 nm He–Ne laser excitation source
phological and optical properties by various characterization (Renshaw RL633). The UV–Vis absorption of the samples
techniques such as the X-ray diffraction (XRD), transmission was recorded on SHIMADZU (UV-3101 PC) UV–Vis spec-
electron microscopy (TEM), scanning electron microscopy trophotometer. Solid photoluminescence (PL) spectra were
(SEM), Fourier transform infra-red spectroscopy (FT-IR), taken using a time-resolved Edimbugh Instrument FLSP920
Raman spectroscopy, thermo gravimetric analysis (TGA), spectrofluorimeter with a Red-PMT detector and a Xenon
differential thermal analysis (DTA), ultraviolet–visible bulb as an excitation source. The thermal behavior was
(UV–Vis) analysis, photoluminescence (PL) and vibrating evaluated by TGA and DTA in the air using TGA Q500 TA
sample magnetometer (VSM). instrument. Magnetization measurements were performed
using the VSM.

2 Experiments 2.4 Photocatalytic experiments

2.1 Materials Photocatalytic efficiency was estimated by the decompo-


sition of methylene blue (MB) dye by using a previously
Chemicals with high purity were bought from shopping established methodology. MB was reagent grade and used
centers (Aldrich) and used without additional purification. as supplied. All the experiments were conducted at room

13
Journal of Materials Science: Materials in Electronics

temperature in air. In each experiment, 0.1 g of photocatalyst narrow sharp peaks indicate that the hematite products are
was sprinkled in 100 mL of MB solution (10 mg/L) to obtain highly crystalline, implying that the high purity of synthe-
the concentration of the catalyst at 1.0 g/L. The experiments sized hematite particles is obtained by using this synthesis
were carried out in a Pyrex beaker illuminated with medium method. Moreover, Fig. 1 indicate that when increasing the
pressure xenon lamp (150 W) was used as the light source. calcination temperature, the intensity of the (104) diffrac-
The visible lamp was localized above the beaker containing tion peak decreases gradually and that its half height width
the solution. The elder was stirred for 30 min prior to irra- is decreased and subsequently the particle size increases.
diation to reach the adsorption/desorption equilibrium. Then Given that line broadening of diffraction peaks is affected
it was continuously stirred during the experiments. At given by the crystallite size and internal strains, the approximate
irradiation time intervals, 4 mL of the suspension were col- particle size “DXRD” is evaluated using the following Scher-
lected, and then centrifuged (6000 rpm, 10 min) to segregate er’s equation [42, 43].
the photocatalyst particles. The MB concentration was eval-
K.𝜆
uated by UV–Vis spectrophotometer (Shimadzu, UV-1650 DXRD = (1)
𝛽 cos θ
pc) monitoring the absorption maximum at λ = 664 nm. A
calibration plot based on Beer–Lambert’s law was settled by where k is the so-called shape factor (0.9), λ is the wave-
relating the absorbance to the concentration. length (0.15418 nm, ­CuKα), β is the full width at half maxi-
mum (FWHM), and θ is the diffraction angle. The (104)
plane was selected to calculate the particle size α-Fe2O3.
3 Results and discussion The average crystallite sizes associated with XRD data
ranged between 18 and 37 nm. It is thus observed that the
3.1 X‑ray diffraction analysis size of pure α-Fe2O3 nanoparticles synthesized through the
precipitation method varied (Table 1) when the calcination
X-ray diffraction patterns of precipitated and calcinated temperature is changed.
α-Fe2O3 nanoparticles are shown in Fig. 1. The XRD pat-
terns are well matched with the standard diffraction data 3.2 Transmission electron microscopy (TEM)
JCPDS No: 33-0664 [39, 40]. All the observed peaks can
be indexed in agreement with the expected rhombohedral Size and morphology of the nanoparticles are determined by
(hexagonal) structure of α-Fe2O3 (space group: R-3c), with analyzing the recorded TEM images. The TEM images of
lattice constants of a = 0.5034 nm and c = 1.375 nm [41]. α-Fe2O3 nanoparticles obtained from precipitation method
The peaks appearing at 2θ range of 24.16°, 33.12°, 35.63°, and calcined at different calcination temperature for 4 h are
40.64°, 49.47°, 54.08° and 57.42° can be attributed to the shown in Fig. 2a–e. After the heat treatment for 4 h, the
(012), (104), (110), (113), (024), (116) and (018) crystalline α-Fe2O3 particles were found in the range of 19–37 nm. It
structures corresponding to pure α-Fe2O3 nanoparticles. The is clear that hematite nanoparticles are mainly present as
granules with small and big spherical shaped particles and
are well crystallize in nature. The nanoparticles size depends
largely on the calcination temperature used in the synthesis
of hematite. In fact the former varies as we vary the lat-
ter. The sizes of nanoparticles were listed in Table 2, which
proved that the particle size increased with the raise of the
calcination temperature (Fig. 2a–e). This enhanced particle
size is due to the coalescence of smaller particle and the

Table 1  Particle size of different synthesized samples


Samples Particle sizes
‘DXRD’ (nm)

α-Fe2O3 calcined at 600 °C 18


α-Fe2O3 calcined at 700 °C 21
α-Fe2O3 calcined at 750 °C 28
α-Fe2O3 calcined at 800 °C 33
Fig. 1  XRD patterns of iron oxide (α-Fe2O3) obtained with chemical α-Fe2O3 calcined at 850 °C 37
precipitation method using different calcination temperature

13
Journal of Materials Science: Materials in Electronics

Fig. 2  TEM observation of α-Fe2O3 prepared with different calcination temperature: a 600 °C, b 700 °C, c 750 °C, d 800 °C and e 850 °C

13
Journal of Materials Science: Materials in Electronics

Table 2  The average particle sizes of the samples measured by TEM activation energy required for nucleation is lowered at higher
Samples Particle temperatures.
sizes (nm)
TEM 3.3 Scanning electron microscopy (SEM)
α-Fe2O3 calcined at 600 °C 19
The surface morphologies of the prepared samples were
α-Fe2O3 calcined at 700 °C 22
studied using SEM. The average sizes of the synthesized
α-Fe2O3 calcined at 750 °C 26
iron oxide nanoparticles under the effects of various prepa-
α-Fe2O3 calcined at 800 °C 34
ration conditions are shown in Fig. 3a–f. All the nanoparti-
α-Fe2O3 calcined at 850 °C 37
cles are spherical in shape. For the effect of the calcination

Fig. 3  SEM observation of α-Fe2O3 prepared with different calcination temperature: a 600 °C, b 700 °C, c 750 °C, d 800 °C, e 850 °C and f
EDX spectrum of α-Fe2O3 nanoparticles

13
Journal of Materials Science: Materials in Electronics

temperature (Fig. 3a–f), the average particle diameter of


hematite increases from 18 to 37 nm with increasing calci-
nation temperature from 600 to 850 °C.
The EDX spectrum of hematite (α-Fe2O3) revealed the
presence of iron (Fe) and oxygen (O) in Fig. 3 f.

3.4 Fourier transform infra‑red (FT‑IR) spectroscopy

The formation of nanoparticles was further confirmed by


FT-IR spectroscopy. Figure 4a shows FT-IR spectra of hem-
atite uncalcined and calcined at 700 °C synthesized through
precipitation method.
For the uncalcined hematite product, the strong absorp-
tion peaks at 556 and 443 cm−1 can be attributed to the
Fe–O band vibrations [44, 45]. The very broad absorption
band centered at 3331 cm−1 and reaching peak at 1625 cm−1
is assigned to the stretching and bending vibrations of the
hydroxyl groups and/or water molecules [46], respectively.
The findings indicate the presence of not only a small
amount of absorbed water on the surface of the product but
also structural hydroxyl groups on the same surface as the
products were prepared in the aqueous solution. In addition,
there is a peak at 1437 cm−1 that is assigned to the deforma-
tion of ­CH3.
FT-IR spectrum investigations of pure hematite cal-
cined at 700 °C prepared by precipitation method shows the
absence of all bands related to the hydroxyl group OH. Yet,
all organic species were removed after calcination. Then we
noticed the appearance of two bands at 527 and 434 cm−1
that can be attributed to the Fe–O vibration in the rhombo-
hedral lattice of hematite [47].

3.5 Raman spectroscopy

Figure 4b portrays the Raman spectra of pure α-Fe2O3 syn-


thesized with different calcination temperature (600, 700,
750, 800 and 850 °C) recorded at room temperature using
632.80 nm excitation wavelengths. Pure hematite belongs
to the R-3c crystal space group and seven phonon lines are
expected to appear in the Raman spectrum, namely two A ­ 1g
phonon modes and five E ­ g phonon modes [48]. The peaks at
227.12 and 494.14 cm−1 are associated with the ­A1g phonon
mode while the peaks at 245.73, 292.74, 299.00, 409.90 and
609.40 cm−1 are related to E ­ g phonon modes. These results
prove that the calcined product is α-Fe2O3. No other iron
oxide, such as magnetite or maghemite, was detected, which Fig. 4  a Spectra FT-IR of pure α-Fe2O3 uncalcined and calcined at
indicates the high purity of the product. Similar results have 700  °C synthesized by precipitation method. b Raman spectra for
hematite synthesized with different calcination temperature. c Ther-
been reported with α-Fe2O3 nanoparticles [49]. Indeed, Xu
mal analyses (TGA/DTA) for pure hematite
et al. [50] used Raman spectrum to explore the nanoparticles
of α-Fe2O3 synthesized by oxygenating pure iron and iden-
tified ­A1g (225, 498 cm−1), ­Eg (252, 293, 411, 612 cm−1).

13
Journal of Materials Science: Materials in Electronics

The Raman spectrum of hematite calcined at 600 °C where α is the absorption coefficient, A is constant, hν is the
is identical to the Raman spectra of α-Fe2O3 synthesized energy of light and n is a constant depending on the nature
with the other calcination temperature (700, 750, 800 and of the electron transition [55]. Hematite has a direct band
850 °C). gap (n = 2) [56]. Figure 5b shows the plot of (αhν)2 versus
hν. The energy gap can be obtained from the intercept of
3.6 Thermal study the linear absorption edge part with the energy axis. When
(αhν)2 is zero, the photon energy is e.g. The optical band gap
The findings associated with formation and decomposition of samples showed a Table 3. The gap energy of the syn-
phase occurring during heat treatment of synthesized sam- thesized samples decreases as the calcination temperature
ples are in good agreement with thermo gravimetric analysis increases. This result is in agreement with the bibliographic
and differential thermal analysis (TGA and DTA) results. data found by [57, 58].
Thermal analyses were carried out from room temperature
to 700 °C. TGA of the prepared compound was performed 3.8 Photoluminescence
by heating the samples in air atmosphere at 10°C/min in
alumina crucible. Figure 4c depicts the thermal analyses Taking into account the excellent luminescent properties of
(TGA, DTA) for pure hematite synthesized by precipitation ­d6 metal complexes and to explore the potential application
technique. as luminescent materials, solid-state fluorescent properties
As for α-Fe 2 O 3 prepared by precipitation method of the title compound was studied. The PL spectrum of the
(Fig. 4c), there are three distinct mass loss steps in the tem- sample at room temperature excited at 250 nm is depicted
perature ranges. The first weight loss step occurred gradu- in Fig. 5c. This material shows one band of luminescence
ally between 44 and 182 °C. The mass loss was of 5.56%, located at 370  nm (3.35  eV), this luminescence can be
and this loss of weight is attributed to the removal of water observed even with the naked eye at room temperature and
existing on the surface of α-Fe2O3. Using DTA enabled us to is due to exciton emission [59]. The luminescence originates
find one endothermic peak at 99.82 °C, The second step cor- from electronic transition within the iron oxide (α-Fe2O3).
responds to a mass loss of 3.04% occurring at 182–540 °C, In this compound, the lowest exciton state arises from exci-
which is due to the combustible organic products present in tation between the valence band (VB), which consists of a
our prepared sample. Through the use of DTA we recorded mixture of Fe (3d) and O (2p) states, and the conduction
one endothermic peak at 186.28 °C. The third step stands band (CB) which derives primarily from Fe (4 s) states.
for a minor weight loss (1.38%) occurring in the range of Under excitation of 363 nm irradiation, an electron (−) is
540–660 °C, which is due to the transition phase of syn- excited from the VB to the CB, leaving a hole (+) in the
thesized compounds. Finally using DTA helped us find one VB. The exciton (−) and the hole (+) move freely in the
endothermic peak at 656.02 °C. After 660 °C, the curve CB and VB, forming an exciton, the recombination of the
becomes parallel to the temperature axis, which emphasizes electron and hole in the exciton yields a red emission at
high stability of α-Fe2O3 nanoparticles. There is no associ- 370 nm (Fig. 5d).
ated signal with the thermal processes of α-Fe2O3 nanopar-
ticles in the TGA curve confirming the crystallization and 3.9 Magnetic properties of the α‑Fe2O3
phase transition of α-Fe2O3 nanoparticles associated with nanoparticles
them.
Magnetic properties of the prepared samples were investi-
3.7 UV–Vis analysis gated by VSM at 300 K and it exhibits the ferromagnetic
behavior. The various parameters like saturation magnetiza-
The absorption spectra in the UV–Vis range of hematite tion ­(Ms), remanent ­(Mr) and coercivity ­(Hc) were tabulated
synthesized with different calcination temperature show that in Table 4 and magnetic hysteresis loop is shown in Fig. 6.
all absorption curves exhibit an intense absorption in the Iron oxide (α-Fe2O3) is an important magnetic material and
range of 500–700 nm wavelength (Fig. 5a).This result is shows an unusual magnetic behavior, due to the particle size
consistent with data from other studies [50–52]. The optical and morphology of the samples [60]. The observed results
band gap ­(Eg) for hematite nanoparticles can be determined indicated a ferromagnetic behavior. The saturation magneti-
by extrapolation from the absorption edge which is given by zation ­(Ms), which was obtained from the extrapolation of
the following equation [53, 54]: the M vs. 1/H curves, is in the order of 27.32 (emu/g). The
­Hc value was obtained from the α-Fe2O3 calcined at 700 °C
(αhν)n = A hν − Eg is equal to 138 (Oe).
(2)
( )

13
Journal of Materials Science: Materials in Electronics

Fig. 5  a UV–Vis DRS spectra of α-Fe2O3 obtained through the chem- temperature. c Excitation fluorescence profile of pure hematite cal-
ical precipitation method with different calcination temperature. b cined at 700 °C. d Simple model for the formation and recombination
Tauc plot obtained with UV–Vis DRS spectra of α-Fe2O3 resulting of the exciton in the title compound
from the chemical precipitation method using different calcination

Table 3  The optical band gap of the samples 3.10 Photocatalytic activity


Samples Optical band
gap ‘Eg’ (eV) The capability of hematite (α-Fe2O3) calcined at 700 °C to
degrade dye was evaluated through its application for deg-
α-Fe2O3 calcined at 700 °C 2.14 radation of MB under visible light irradiation are shown in
α-Fe2O3 calcined at 750 °C 2.10 Fig. 7. The characteristic absorption peak of MB at 664 nm
α-Fe2O3 calcined at 800 °C 2.05 decreased in intensity over time and it disappeared com-
pletely after about 140 min.
The percentage of degradation of MB dye was calculated
Table 4  Magnetic parameters of hematite calcined at 700 °C from the following equation:
Samples Ms (emu/g) Mr (emu/g) Hc (Oe)
(3)
( )
Percentage of degradation = 1 − Ct ∕C0 × 100
α-Fe2O3 calcined at 700 °C 27.32 17.79 138

13
Journal of Materials Science: Materials in Electronics

Fig. 6  Hysteresis loop of the α-Fe2O3 calcined at 700 °C Fig. 8  Photodegradation of MB in the presence of α-Fe2O3 nanoparti-
cles calcined at 700 °C

leading to the formation of an electron–hole pair. Further


charge separation and migration of the generated charge
carriers towards the surface of catalyst can then lead to
redox reactions with organics. The high oxidative potential
of the hole ­(h+) (VB) in the catalyst allows direct oxidation
of MB to reactive intermediates. Very reactive hydroxyl
radicals can also be formed by the decomposition of water.
On the other hand, the reaction of electrons with dis-
solved oxygen molecules to give superoxide radical anions,
­O2·−, yielding hydroperoxyl radicals ­HO2· on protonation
and finally OH· radicals will be more efficient.
α - Fe2 O3 + hν → α - Fe2 O3 + e− + h+

h+ + MB → MB⋅+
O2 + e− → O2 ⋅−

O2 ⋅− + H2 O → HO2 ⋅ + OH⋅−
Fig. 7  Time-dependent UV–Vis absorption spectra for MB dye in the HO2 ⋅ + H2 O → OH⋅ + H2 O2
presence of hematite calcined at 700 °C
As can be seen from Fig. 10, the photodegradation of MB
by α-Fe2O3 calcined at 700 °C followed a first-order rate law,
where ­C0 and C­ t in Fig. 8 represent the initial concentration ln ­(Ct/C0), where C­ 0 is initial concentration of methylene
after the adsorption–desorption equilibrium for 30 min and blue (mg/l). ­Ct is the concentration of the dye at various
the real-time concentration of MB, respectively. Figure 8 interval times (mg/l), t is the illumination time (min) and k
showed that the decolorization rate of MB using α-Fe2O3 is the reaction rate constant.
nanoparticles calcined at 700 °C reached 89%.
The photocatalytic mechanism proposed for the iron
oxide (α-Fe2O3) nanoparticles under visible light is shown
in Fig. 9. Upon visible illumination, electrons in the VB
could be excited to the CB of the oxide, with the concomi-
tant formation of the same quantity of holes in the VB,

13
Journal of Materials Science: Materials in Electronics

Fig. 9  Schematic illustration of
the photocatalytic mechanism
using α-Fe2O3 nanoparticles
under visible light

4 Conclusion increase of the calcination temperature leads to an increase


of size of nanoparticles. In conclusion, the lowest calcination
In this study, hematite nanoparticles were prepared by pre- temperature (700 °C) resulted in the smallest size (21 nm)
cipitation method with different calcination temperature and best crystallinity of hematite nanoparticles. Moreover,
(600, 700, 750, 800 and 850 °C). The effects of changes hematite calcined at 700 °C yielded the biggest band gap
in the calcination temperature used in the synthesis of ­(Eg = 2.14 eV). From the hysteresis loop, it is clear that the
α-Fe2O3 nanoparticles at the level of size, morphology and synthesized sample can be control, for α-Fe2O3 nanoparti-
optical band gap were explored. XRD, TEM, SEM, FT-IR, cles. The photodegradation analysis showed that the α-Fe2O3
Raman spectroscopy, DTA, TGA, UV–Vis, VSM and PL nanoparticles performed the best as a photocatalyst during
techniques were used to characterize the synthesized sam- the degradation of MB under visible light irradiation.
ples. The results of the different techniques show that the
Acknowledgements  The present work was supported by the Research
Funds of Electrochemistry, Materials and Environment Research Unit
UREME (UR17ES45), Faculty of Sciences Gabes University, Tunisia
and Structures, Properties and Modeling of Solids (SPMS) Laboratory,
University Paris-Saclay, France.

References
1. L. Machala, R. Zboril, A. Gedanken, J. Phys. Chem. B 111, 4003–
4018 (2007)
2. S. Sakurai, A. Namai, K. Hashimoto, S.J. Ohkoshi, J. Am. Chem.
Soc. 131, 18299–18303 (2009)
3. N.D. Phu, D.T. Ngo, L.H. Hoang, N.H. Luong, N. Chau, N.H. Hai,
J. Phys. D 44, 345002–345008 (2011)
4. R.M. Cornell, U. Schwertmann, Reactions and Uses (Wiley-VCH,
Weinhein, 2003)
5. R. Zboril, M. Mashlan, D. Petridis, Chem. Mater. 14, 969–982
(2002)
6. J. Tucek, R. Zboril, A. Namai, S. Okoshi, Chem. Mater. 22, 6483–
6505 (2010)
Fig. 10  The pseudo first order kinetics of degradation of MB dye

13
Journal of Materials Science: Materials in Electronics

7. L. Machala, J. Tucek, R. Zboril, Chem. Mater. 23, 3255–3272 35. K. Kandori, S. Ohnishi, M. Fukusumi, Y. Morisada, Colloid Surf.
(2011) 331, 232–238 (2008)
8. S.I. Srikrishna Ramya, C.K. Mahadevan, Mater. Lett. 89, 111–114 36. N.K. Ilona, R. Aleksander, P. Mihaly, J. Nanopart. Res. 14, 1150–
(2012) 1159 (2012)
9. A.S. Teja, P.Y. Koh, Prog. Cryst. Growth Charact. Mater. 55, 37. M. Ming, Z. Yu, G. Zhirui, G. Ning, Facile synthesis of ultrathin
22–45 (2009) magnetic iron oxide nanoplates by Schikorr reaction. Nanoscale
10. R. Suresh, R. Prabu, A. Vijayaraj, K. Giribabu, A. Stephen, V. Res. Lett. 8, 16–22 (2013)
Narayanan, Mater. Chem. Phys. 134, 590–597 (2012) 38. A. Ibrahim, A.B. Abubakar, Afr. J. Pure Appl. Chem. 7, 114–121
11. M.M. Khadar, N.N. Lichtin, G.H. Vurnes, M. Salmenon, G.A. (2013)
Somoraj, Langmuir 3, 303–304 (1987) 39. J. Hua, J. Gengsheng, Mater. Lett. 63, 2725–2727 (2009)
12. J. Shakhapure, H. Vijayanand, S. Basavaraja, V. Hiremath, A. 40. T. Almeida, M. Fay, Y.Q. Zhu, P.D. Brown, J. Phys. Chem. C 113,
Venkataraman, Bull. Mater. Sci. 28, 713–718 (2005) 18689–18698 (2009)
13. H. Wang, W. Geng, Y. Wang, Res. Chem. Intermed. 37, 389–395 41. A.S. Teja, P.Y. Koh, Prog. Cryst. Growth 55, 22–45 (2009)
(2011) 42. A. Lassoued, M.S. Lassoued, B. Dkhil, A. Gadri, S. Ammar, J.
14. M. Tadic, N. Citakovic, M. Panjan, Z. Stojanovic, D. Markovic, Mol. Struct. 1141, 99–106 (2017)
V. Spasojevic, J. Alloys Compd. 509, 7639–7644 (2011) 43. W. Ben Soltan, M.S. Lassoued, S. Ammar, T. Toupance, J. Mater.
15. M. Tadic, N. Citakovic, M. Panjan, Z. Stojanovic, K. Markovic, D. Sci.: Mater. Electron. 28, 15826–15834 (2017)
Jovanovic, V. Spasojevic, J. Alloys Compd. 543, 118–124 (2012) 44. H. Liu, P. Li, B. Lu, Y. Wei, Y. Sun, J. Solid State Chem. 182,
16. D.A. Wheeler, G. Wang, Y. Ling, Y. Li, J.Z. Zhang, Energy Envi- 1767–1771 (2009)
ron. Sci. 5, 6682–6702 (2012) 45. Z. Jing, S. Wu, Mater. Lett. 58, 3637–3640 (2004)
17. M. Zhu, Y. Wang, D. Meng, X. Qin, G. Diao, J.Phsy. Chem. C 46. E. Darezereshki, Mater. Lett. 65, 642–645 (2011)
116, 16276–16285 (2012) 47. K.H. Wu, T.H. Ting, G.P. Wang, C.C. Yang, B.R. Mc, Garvey,
18. D.W. Jung, D.W. Park, Appl. Surf. Sci. 255, 5409–5413 (2009) Mater. Res. Bull. 40, 2080–2088 (2005)
19. M.I.B. Bernardi, S. Cava, C.O. Paiva-Santos, E.R. Leite, C.A. 48. A. Lassoued, M.S. Lassoued, B. Dkhil, A. Gadri, S. Ammar, J.
Paskocimas, E. Longo, J. Eur. Ceram. Soc. 22, 2911–2919 (2002) Mol. Struct. 1148, 276–281 (2017)
20. F. Lan, X. Wang, X. Xu, React. Kinet. Mech. Catal. 106, 113–125 49. D. Bersani, P. Lottici, A. Montenero, J. Raman Spectrosc. 30,
(2012) 355–360 (1999)
21. F. Morazzoni, C. Canevali, N. Chiodini, C. Mari, R. Ruffo, R. 50. Y.Y. Xu, D. Zhao, X.J. Zhang, W.T. Jin, P. Kashkarov, H. Zhang,
Scotti, L. Armelao, E. Tondello, L. Depero, E. Bontempi, Mater. Physica E 41, 806–811 (2009)
Sci. Eng. C 15, 167–169 (2001) 51. S. Sivakumar, D. Anusuya, C.P. Khatiwada, J. Sivasubramanian,
22. J. Zhang, K. Colbow, J. Appl. Phys. 71, 2238–2242 (1992) A. Venkatesan, P. Soundhirarajan, Spectrochim. Acta 128, 69–75
23. A. Messad, J. Bruneaux, H. Cachet, M. Froment, J. Mater. Sci. 29, (2014)
5095–5103 (1994) 52. A. Lassoued, B. Dkhil, A. Gadri, S. Ammar, J. Res. Phys. 7,
24. H. Kim, A. Pique, Appl. Phys. Lett. 84, 218 (2004) 3007–3015 (2017)
25. J.R. Zhang, L. Gao, Mater. Chem. Phys. 87, 10–13 (2004) 53. A. Lassoued, M.S. Lassoued, B. Dkhil, S. Ammar, A. Gadri, J.
26. K.S. Kim, S.Y. Yoon, W.J. Lee, K.H. Kim, Surf. Coat. Technol. Physica E 97, 328–334 (2018)
138, 229–236 (2001) 54. W. Ben Soltan, S. Nasri, M.S. Lassoued, S. Ammar, J. Mater. Sci.:
27. T.M.N. Thai, S.R. Kim, H.J. Kim, New Phys. 64, 252–255 (2014) Mater. Electron. 28, 6649–6656 (2017)
28. A. Lassoued, M. Ben hassine, F. Karolak, B. Dkhil, S. Ammar, A. 55. R. Branek, H. Kisch, Photochem. Photobiol. Sci. 7, 40–48 (2008)
Gadri, J. Mater. Sci.: Mater. Electron. 28, 18857–18864 (2017) 56. J.I. Pankove, Optical Processes in Semiconductors (Prentice-Hall
29. A. Lassoued, M.S. Lassoued, F. Karolak, S. García-Granda, B. Inc., Englewood Cliff, 1971), pp 34–86
Dkhil, S. Ammar, A. Gadri, J. Mater. Sci.: Mater. Electron. 28, 57. K. Varunkumara, R. Hussaina, G. Hegdeb, A.S. Ethiraj, Mater.
18480–18488 (2017) Sci. Semicond. Process. 66, 149–156 (2017)
30. H. Liu, H. Guo, P. Li, Y. Wei, J. Solid State Chem. 181, 2666– 58. A. Amirsalari, S. Farjami, Shayesteh, Superlattices Microstruct.
2671 (2008) 82, 507–524 (2015)
31. H. Liu, Y. Wei, Y. Sun, J. Mol. Catal. A 226, 135–140 (2005) 59. Y. Ling, D.A. Wheeler, J.Z. Zhang, Y. Li, One-Dimens. Nano-
32. L. Diamandescu, D.M. Tarabasanu, N.P. Pogrion, A. Totovina, I. struct. 167–184 (2013)
Bibicu, Ceram. Int. 25, 689–692 (1999) 60. H. Ni, Y.Y. Zhou, J.M. Hong, Mater. Lett. 73, 206–208 (2012)
33. R.M. Cornell, R. Giovanoli, Clay Clay Miner. 33, 424–432 (1985)
34. U. Schwertmann, J. Friedl, H. Stanjek, J. Colloid Interface Sci.
209, 215–223 (1999)

13

View publication stats

You might also like