You are on page 1of 10

Journal of Environmental Chemical Engineering 8 (2020) 104214

Contents lists available at ScienceDirect

Journal of Environmental Chemical Engineering


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

Low-cost synthesis of Cu/α-Fe2O3 from natural HFeO2: Application in 4- T


nitrophenol reduction
Amal Elfiada,d,*, Federico Gallib,d, Laaldja Moddeur Boukhobzaa, Amar Djadounc,
Daria C. Boffitod
a
Laboratoire des Matériaux Catalytiques et Catalyse en Chimie Organique, Faculté de Chimie, USTHB, BP32 El Alia, Bab Ezzouar, 16111, Algiers, Algeria
b
Dipartimento di Chimica, Universitá degli Studi di Milano, via Golgi 19, 20133, Milano, Italy
c
Laboratoire des Geéophysique, FSTGAT, USTHB, BP32 El Alia, Bab Ezzouar, 16111, Algiers, Algeria
d
Department of Chemical Engineering, Polytechnique de Montréal, C.P. 6079, Succ. CV, Montréal, H3C 3A7, Québec, Canada

A R T I C LE I N FO A B S T R A C T

Editor: Teik Thye Lim Nitro-aromatic compounds pollute the aquatic environment and human being may ingest them through drinking
Keywords: water. Among these, 4-nitrophenol is the most common. In presence of an excess of NaBH4, 4-nitrophenol
Fe- Cu catalyst catalytically reduces to 4-aminophenol, which is not toxic. In this work, the Cu/α-Fe2O3 materials were suc-
Natural hematite cessfully prepared via a simple classical impregnation. The natural hematiteα-Fe2O3 support was obtained from
Goethite treatment natural goethite by a simple thermal treatment and supported on it a mass ratio of 5 % or20 % of Cu at room
4-Nitrophenol reduction temperature. The composition and the structural-morphological properties of natural α-Fe2O3 and Cu/α-Fe2O3
4-Aminophenol were identified by several characterization methods, including thermo gravimetric analysis (TGA), X-ray
Kinetic reaction
fluorescence (XRF), X-ray diffraction (XRD), transmission electron microscopy (TEM), scanning electron mi-
croscopy (SEM), energy dispersive X-ray spectroscopy (EDX), particle size distribution (PSD), and nitrogen
adsorption-desorption at 77 K. The Cu/α-Fe2O3 materials composite presented spherical and uniform particle,
high porosity, good copper dispersion and a mixture structural phase composition, all these characteristics and
the active site nature on her surface bring to approving the excellence catalytic performance of this solid
compared to the α-Fe2O3.The reduction of 4-nitrophenol (4-NP) to 4-aminophenol (4-AP) is completed in 250 s
over natural α-Fe2O3, and in 50 s over Cu/α-Fe2O3 (20 %Wt of Cu), the corresponding rate constant (k) for
reduction reaction was 0.09 s-1 per mg of Cu/α-Fe2O3 catalyst.

1. Introduction system [5]. Nitrophenols reductions to aminophenol in industrial


wastewater are mostly concentrated, but an option to decrease their
Pollutants such as dyes and pigments containing nitro compounds concentration is possible. Aminophenol is a precursor in the manu-
which contaminate agricultural and industrial wastewaters : petro- facture of analgesics, antipyretics and other drugs. It is also a photo-
chemicals, pharmaceutical industries, textiles, printing, …etc. [1–3]. graphic developer, a corrosion inhibitor and a lubricant [6]. There are
These substances are toxic and have a carcinogenic profile. After several methods to remove this pollutant from wastewater like elec-
treatment, industrial wastewaters are discharged as a fresh water and trochemical methods [7], hydrogenation reactions [8], Fenton-like re-
the concentration of nitro-aromatic compounds are detected on the actions [9], photocatalysis [10,11] biological degradation [12] and
order of the ppb. However, long-term exposure and assimilation of adsorption [13]. Most of the techniques mentioned above have lim-
nitro-aromatic compounds through drinking water may damage the itations, such as the production of secondary pollutants, disposal of the
environment and possibly human being health [4]. Nitrophenols have catalyst sludge, formation of hazardous by-products, high energy re-
been listed as priority pollutants by the U.S Environmental Protection quirements and expensive post-processing [14]. Industry favors the
Agency, because of their high solubility and stability in water [4]. 4- hetero-generous catalytic reduction using sodium borohydride NaBH4
nitrophenol (4-NP) is toxic and hardly degradable pollutant which da- as reducing agent [15]. In the presence of catalysts, NaBH4 yields a
mages the central nervous system and the respiratory and digestive stable chemical hydride [16]. 4-NP catalytic reduction to 4-


Corresponding author at: Laboratoire des Matériaux Catalytiques et Catalyse en Chimie Organique, Faculté de Chimie, USTHB, BP32 El Alia, Bab Ezzouar, 16111,
Algiers, Algeria.
E-mail address: aelfiad@usthb.dz (A. Elfiad).

https://doi.org/10.1016/j.jece.2020.104214
Received 9 March 2020; Received in revised form 21 June 2020; Accepted 23 June 2020
Available online 24 June 2020
2213-3437/ © 2020 Published by Elsevier Ltd.
A. Elfiad, et al. Journal of Environmental Chemical Engineering 8 (2020) 104214

aminophenpl (4-AP) has became one of the most studied model reac- Table 1
tions to test the activity of catalysts [17]. Zhao et al. reviewed the main Catalysts and synthesis conditions.
synthesis procedures and tested the activity of diverse Au nanoparticles Catalysts Cu loading Calcination temperature (°C) Abbreviations
for this reaction [18], while Shukla et al. studied the activity of Pt (masse%)
nanoparticles supported on ceria [19]. In this way, Mourya et al. tested
Natural iron – – Fe-NAT
the AgNPs in the 4-NP reduction, the reduction time was estimated
Fe-NAT – 750 α- Fe2O3
around 15 min for 100 % of conversion to 4-AP [20]. Scientific data is 5%Cu/ α- Fe2O3 5 500 5CF500
available on several bimetallic noble metal catalysts for example: 20 %Cu/ α- Fe2O3 20 500 20CF500
AuNPs [21], Au/Pt [22], and Au/Pd [23]. All these catalysts converted 20 %Cu/ α- Fe2O3 20 700 20CF700
more than 99 % of 4-NP, but their precursors are expensive and the
synthesis is complex and lengthy. Transitions metals, such as Co [24],
Ni [25], Fe [26] and Cu [27] may be active as well as in the 4-NP USA X'pertdiffractometer) scanned the powder samples with Cu-
reduction. Among them, Cu is the most active. An approach to increase Kαradiation (λ = 1:5406 Å) and the scanning angle ranged from 10° to
the catalytic performance of Cu nanoparticles is to deposit them on 80°. A scanning electron microscope (FE-SEM-JEOL JSM-7600 F) im-
carriers to establish a synergistic metal-support action [27]. Lei Jin aged the catalysts, and energy dispersive X-ray (EDX) analysis mea-
et al. considered reduced graphene oxide (rGO) as a support, and Cu/ sured their superficial composition. A FEI Tecnaï 20 F transmission
rGO completely reduced 4-NP in 6 min [28]. Darabi et al. [29,30] im- electron microscopy (TEM) imaged the samples. Fe and Cu concentra-
proved the heteropoly compounds performance by adding Cu in the tions were determined by Atomic Adsorption Spectroscopy (AAS)
bulk of the catalyst. Bordbar et al. [31] synthesized CuO/ZnO nano- (Shimadzu PERKIN-ELMER-2380). A Thermo Mattson Fourier
composite and converted 4-NP in 3 min. Elfiad et al. [32] tested Cu/ Transformation Infrared (FTIR) spectroscopy recorded all samples'
NaX for the 4-NP reduction and improved the total reduction to 4-AP in spectra from 400 cm-1 to 4000 cm−1 over a KBr pellet. A laser scat-
60 s with a conversion > 99 %. Wang et al. [33] and Nasrollahzadeh tering analyzer (LA-950 Horiba) quantified the particle size distribution
et al. [34] employed Cu/Fe3O4 and Nandanwar et al. used CuO/γ-Al2O3 (PSD). To determine the textural properties of the prepared samples, N2
catalysts [35]. Our work is original because we support Cu on natural adsorption-desorption measurement was carried out at 77 K using a
hematite (α-Fe2O3). We recently reported the fast and ecofriendly Micrometrics Autosorb-1 (Quantachrome Instruments) after the sample
synthesis of hematite [36]. In this paper we report for the first time an degassed at 200 °C in vacuum for 12 h.
inexpensive and environmentally friendly approach for the preparation
of Cu/α-Fe2O3; the Cu mass fraction was 5 % to 20 %; α-Fe2O3 derived 2.4. Catalytic tests
from natural HFeO2 ore iron via a one-step thermal treatment. We
deposited CuNPs on the surface of natural hematite with different The catalytic activity and stability of α-Fe2O3 and Cu/α-Fe2O3
loadings. Both 5%Cu/α-Fe2O3 and 20 %Cu/α-Fe2O3 reduced 4-NP in materials tested in the reduction of 4-nitrophenol (4-NP) in the pre-
presence of NaBH4at ambient temperature. sence of excess NaBH4 at room temperature. In a typical process,
12.5 mL freshly prepared NaBH4 was added to a batch reactor that
2. Experimental section contains 100 mL of aqueous solution of 4-Nitrophenol, then the mixed
solution was stirred at room temperature (25 ◦C) for 3 min. Here, the
2.1. Reagents final concentrations of 4-NP and NaBH4 were 5 × 10−5 mol.L-1 and
0.5 mol.L-1 respectively, and the pH of the solution was 11.5. After
The catalyst precursor was copper nitrate trihydrate [Cu (NO3)23 wards, 3 ml of the mixed solution were transferred in a quartz cuvette,
H2O, 99 %], it was purchased from Sigma Aldrich. The substrate for the and then 1 mg of as prepared catalyst was added to the cuvette solution.
degradation experiments was 4-nitrophenol (4-NP, 99 %); the reducing The quartz glass was quickly subjected to UV–vis spectroscopy (JASCO
agent was sodium borohydride (NaBH4, 97 %) by Fluka. We did not model V-530 spectrophotometer) and the absorbance of the solution
purify the reagents further. Natural iron rock was the precursor of the was in situ measured every several seconds in the scanning range of
catalyst support. The Geology Laboratory (University of Science and 200–500 nm. The maximum adsorption peak of 4-NP showed at
Technology Houari Boumedien) collected the natural ore iron rock in 400 nm, the conversion of the reaction was calculated according to the
the Melbou-Bejaia region (Algeria). Eq. (1). To optimize the reaction conditions, a series of reactions were
performed using several catalyst amounts and several bases con-
2.2. Catalyst preparation centrations to obtain the best possible combination (Table S1). After the
reaction, a centrifuge separates the catalyst from the solution. Distilled
We treated natural ore iron goethite via a thermal treatment process water and ethanol in sequence rinsed the catalyst removed the 4-AP
to obtain hematite [36]. Detailed synthetic procedures of the natural α- adsorbed on the catalyst surface. The catalysts dried in an overnight for
Fe2O3 were shown in the supporting information. The impregnated Cu/ the next cycle. The catalysts were re-used in five successive tests with
α-Fe2O3 catalysts were prepared by impregnating α-Fe2O3 powder with fresh 4-NP and NaBH4 solution.
an appropriate amounts of copper nitrate (Cu (NO3)2 3 H2O) aqueous X4-NP = (A0-At/A0) *100 (1)
solution, stirred (3500 rpm) at room temperature for 1 h. The obtained
solids were dried at 100 °C overnight. The samples were calcined at A0 is the absorbance at the beginning of the reaction (t = 0) and At is
different temperatures (Table 1) for 5 h with a heating rate of the absorbance at sampling time t.
5 °C.min-1.
3. Results and discussion
2.3. Characterization
3.1. Thermogravimetric analysis (TGA)
A TGA-Q50 (thermo gravimetric analysis Instruments) system
characterized the thermo gravimetric behavior of the precursors with a The TGA curve of natural α-Fe2O3 exhibits three distinct mass loss
temperature interval between 25 ◦C to 800 ◦C and a ramp of 10 ◦C regions (Fig. 1a). The first mass loss of 0.9 % occurs in the temperature
min−1 under an air flow of 10 l min−1. An X-ray fluorescence spec- range of 25 °C–165 °C and belongs to the evaporation of water physi-
trometer PW1400 Philips Instrument evaluated the composition of sorbed on the surface of natural goethite. The second mass loss of 9.1 %
natural iron ore. An X-ray diffractometer (XRD, Philips PANanalytical occurs at 165 ◦C to 288 °C, and is due to the dehydroxylation of

2
A. Elfiad, et al. Journal of Environmental Chemical Engineering 8 (2020) 104214

Fig. 1. Thermal analysis (TGA) for (a) natural α-Fe2O3 (b) 5 %Cu/α-Fe2O3 (c) 20 %Cu/α-Fe2O3.

goethite α-FeO(OH) (identified by XRD) into hematite (α-Fe2O3) The diffraction peaks in the natural ore iron Fe-NAT (Fig. 2.A)
[37,38]. The third mass loss accounts for a mass of 2.2 % and occurs correspond mostly to goethite (α-FeOOH, JCPDS 08-097), and to hex-
between 288 °C–347 °C. It corresponds to the decomposition of the re- agonal α-Fe2O3 hematite (JCPDS 33-664). XRD patterns changed after
maining hydroxyl groups in goethite and partly to non-stoichiometric calcination at 700 °C, where by all goethite's peaks disappear in favor of
hydroxyl units [39]. Above 350 °C there is no further mass loss, in- hematite's, due to the Fe-O restructuring (Fig. 2a). The X-ray dif-
dicating that α-Fe2O3 is stable. The TGA curve of 5%Cu/α-Fe2O3 and 20 fractograms of the Cu/α -Fe2O3 materials with different copper
%Cu/α-Fe2O3 are shown in the (Fig. 1b and c), respectively. The first amounts and calcined at different temperature (500 and 700 °C) are
mass loss of 8.5 % and 9.8 % is between 25 °C–110 °C and 25 °C–174 °C reported in Fig. 2B. Powder X-ray diffraction of 5CF500, 20CF500 and
for the precursor containing a copper mass percentage of 5 % and 20 %, 20CF700 revealed the presence of the rhombo-hexagonal structure of α-
respectively. These losses are attributed to the removal of physically Fe2O3 (space group: R-3C), with lattice parameters of a = b =5.034 Å
adsorbed water from the original precursors [39,40]. The second and c =13.729 Å, which are in good agreement with standard reference
thermal stage occurs in the range of 110 °C–385 °C and 175 °C–490 °C data (PDF-ICCD: 33-0664) [42,43].
with mass percentage losses of 5.2 % and 4.7 % for 5 %Cu/α-Fe2O3 and The diffractograms of 20CF500 and 20CF700 solids exhibit a peak at
20 %Cu/α-Fe2O3 respectively, which are due to the transition phase 2θ = 38:7° corresponding to CuO phase (PDF-ICCD: 01-1117), with
and complete decomposition of copper nitrate into their corresponding lattice parameters a =4.653 Å, b =3.14 Å and c =5.108 Å, the 20CF700
oxide CuO [39,41]. On the other hand, especially for 20 %Cu/α-Fe2O3, material (Fig. 2d) contained another phase refracting at 2θ = 36.9°,
the TGA curve shows a third mass loss (0,6 %) occurring in the range of 43.7°, 58.8°, and 62.2°. These diffraction lines belong to tetragonal type
500 °C–562 °C, attributed to the solid-solid interaction between α-Fe2O3 CuFe2O4 spinel (ICCD: 34-0425), whose presence was confirmed by
and CuO forming CuFe2O4 spinel (Eq. (2)): TGA.
The 5CF500 solid does not contain CuO oxide and it preserved the
α-Fe2O3+CuO → CuFe2O4 (2) rhombohedral hematite structure, suggesting either a homogeneous
The increase in mass loss of Cu supported hematite compared to dispersion of CuO on the hematite matrix, or CuO particle size in the
hematite may be associated to a higher loss of copper precursors and nanometer scale [44], however, increasing the calcination temperature
the effect of the Cu impregnation on α-Fe2O3. The TGA data associated (Fig. 2c-d) and the copper amount increases the crystallinity degree of
with the formation and decomposition of the different phases agree CuO. The diffraction patterns exhibits two peaks belonging to the (1 0
with the XRD analysis. 4) and (1 1 0) plans of α-Fe2O3, in Cu/α-Fe2O3these peaks slightly shift
from 32.9° and 35.6° to 33.2° and 35.8°, respectively (Fig. 2e). This
indicates the change of the α-Fe2O3 lattice parameters after the addition
3.2. Chemical composition and crystal structure of CuO [45].The mean crystallite size was determined from the line
broadening of the attributed XRD lines using the Williamson–Hall
The chemical compositions (wt %) of the natural goethite have been method (Table 2).
examined by the X-ray fluorescence (XRF) analysis, the obtained results
reveal the presence of the iron as majority percentage (96.12 %) with 3.3. Morphology and composition
the minority other elements (SiO2 1.98 %, MnO2 1.60 % and MgO 0.30
%), the results indicated that our ore iron is not pure and contains some We report representative SEM (scanning electron microscopy)
impurities. In the other hand, the EDAX analysis spectrum of the nat- images for the sake of brevity. α-Fe2O3 (Fig. 3a) particles have shuttle-
ural ore iron (Fig. S1) show the presence of the same elements observed like shape with relatively smooth surface. With the increase in the Cu
by XRF, the appearance of copper pick indicated that the holder sample amount the shuttle-like shape disappears and the surface became
used in this measure is made of copper. rougher (Fig. 3b–d). Comparing 20CF500 with 20CF700 (Fig. 3c-d), the

3
A. Elfiad, et al. Journal of Environmental Chemical Engineering 8 (2020) 104214

Fig. 2. (A) XRD patterns of the natural Fe-NAT and α-Fe2O3. (B) XRD patterns of the (a) α-Fe2O3, (b) 5CF500, (c) 20CF500 and (d) 20CF700.

crystallinity increases with the calcined temperature from 500 ◦C to slightly increases the CuO particle size to 40 nm and decrease theα-
700 °C, exhibiting a more ordinate morphology. In conclusion, the mass Fe2O3 particle size to 11 nm (Fig. 4c). The TEM image of the 20CF700
percentage of Cu and the calcination temperature influence the mor- shows uniform and homogenous spherical particles with average dia-
phology of Cu/α-Fe2O3 composite oxide. meter around 15–53 nm (Fig. 4d). The increase in the particle size with
The transmission electron microscopy (TEM) determined the size the calcination temperature is due to the formation of CuFe2O4 phase
and morphology of the nanoparticles (Fig. 4). The original morphology together with sintering. The TEM data agree with the XRD, SEM and
of α-Fe2O3 nanoparticles is non-uniform, spherical with an average surface area analysis (Table 2).
diameter of 50 nm, these agglomerates formed by a superposition of The EDXA analysis is determined the composition of nanoparticles
nonmetric size particles. The TEM image of 5CF500 displays spherical (Table 2, Fig. 3). The EDS spectrum of the Cu-Fe oxide confirmed the
grains (37 nm) more uniform than α-Fe2O3 particles (Fig. 4b). However, presence of Fe and Cu (which is in agreement with the XRD a result
TEM images of the 5CF500 catalyst show that the solid is constituted of discussed earlier (Fig. 2)) and O peaks in the samples; with the Si and
the small particles (3 nm) dispersed on the surface, these particles can Mg impurity, the origin of these impurities was proved previously by
be attributed to the CuO phase. Increasing the Cu content to 20 % XRF analysis. The atomic absorption analysis (AAS) agrees with the

Table 2
Catalysts textural properties and particle size distribution.
Samples Calculated Composition by AAS Composition by S BET pore Average pore PSD Particle size Crystallite size byXRD (nm)
composition % EDX (m2/g) volume diameter (nm) by TEM (nm)
Wt% Wt% (cm3/g)
O Fe Cu Mean size α-Fe2O3 CuO CuFe2O3
Cu Fe Cu Fe (μm)

α-Fe2O3 / 70 / 69.51 /// 38 0.23 19.50 28 50 29.65 – –


5CF500 5 66.5 4.86 67 30 61 6 42 0.24 17.72 25 3−37 27.12 – –
20CF500 20 52 20.12 53.09 31 54 12 29 0.21 20.1 41 11−40 22.25 11.5 –
20CF700 20 52 20.15 52.63 30 55 13 10 0.20 25.9 85 15−53 20.52 50.50 18.33

4
A. Elfiad, et al. Journal of Environmental Chemical Engineering 8 (2020) 104214

Fig. 3. SEM/EDS images of the (a) α-Fe2O3, (b) 5CF500, and (c) 20CF500, (d) 20CF700.

Fig. 4. TEM images of (a) α-Fe2O3, (b) 5CF500, (c) 20CF500, and (d) 20CF700.

5
A. Elfiad, et al. Journal of Environmental Chemical Engineering 8 (2020) 104214

Fig. 5. FT-IR spectra of (a) α-Fe2O3, (b) 5CF500, (c) 20CF500, and (d)
20CF700.
Fig. 7. Particles size distributions of (a) α-Fe2O3, (b) 5CF500, (c) 20CF500, and
expected stoichiometry (Table 2). (d) 20CF700.

3.4. Fourier Transformation Infrared (FT-IR) spectroscopy volume, due to the placement of copper species in the hematite pores
leading to their enlargement. We attribute the higher BET surface area
Two main absorption bands in the range 467 cm−1 to 565 cm−1 of 5CF500 compared to the pure hematite to the creation of new pores,
characterize hematite (Fig. 5a). They both correspond to metal oxygen and to the increase in the total pore volume.
(Fe O) stretching [46]. The Cu/α-Fe2O3 samples exhibit the same peaks The BET surface area decreases from 38 m2 g−1 to 29 m2 g−1 for α-
as hematite, but the bands associated with the Fe-O vibration shift to Fe2O3and 20CF500. This indicates that CuO blocks some of the active
higher wave numbers (Fig. 5b–c). This difference ascribes to the in- sites accessible for nitrogen adsorption when the loading of Cu in-
teraction between Cu and O [47]. The FT-IR results correlate with the creases. A higher calcination temperature (500 ◦C vs. 700 °C) decreased
TGA and XRD analyses, and confirm the formation of CuO and CuFe2O4 the surface area of about 30 % (20CF700 vs. compared to 20CF500). This
for 20CF500 and 20CF700, respectively. decrease correlates with the crystallization of new phases (confirmed by
the appearance of the spinel phase in XRD and TGA).
3.5. Nitrogen adsorption/desorption isotherms
3.6. Particle size distribution (PSD)
All samples exhibit similar shape of isotherms of type-IV according
to the IUPAC classification, typical of mesoporous materials Hematite α-Fe2O3 particle diameter is the smallest (mean size:
(2 nm–50 nm), accompanied by an H2 hysteresis loop (Fig. 6). Ag- 28 μm), whereas 20CF700 has the largest diameter (85 μm) (Table 2).
glomerates of spherical particles generally constitute solids character- After the addition of 5 % of Cu, the particle size distribution does not
ized by this type of hysteresis [48]. From the BET surface area, average change and mean size ∼ 25 (Fig. 7b). When the copper loading in-
pore diameter and corresponding pore volumes (Table 2), we deduce creases to 20 %, the PSD curve becomes bimodal, with a peak at around
that the pore size distribution is in the range of 19 nm–26 nm for all 600 μm (Fig. 7c). This secondary peak is attributed to the agglomera-
samples, which indicates that α-Fe2O3 and Cu-Fe are mesoporous cat- tion of 20CF500 particles in the aqueous solution during the PSD
alysts. The incorporation of 5 % and 20 % of Cu increases the pore measurement.
The average size of 20CF500 was 41 μm. In the sample calcined at
700 °C (20CF500), the two peaks from the bimodal distribution seem to
merge in one larger peak. The main PSD peak shifts at higher diameter
because of particle sintering at higher calcination temperature (700 ◦C
vs. 500 °C). However, the particles do not agglomerate in water during
the analysis because, being bigger, they are more stable than those of
20CF500. The PSD analysis is consistent with the TEM images.

3.7. UV-visible (UV-vis) spectroscopy

4-NP adsorption peak is centered at 317 nm. After adding the so-
dium borohydride (NaBH4) at room temperature, the solution changes
its color from yellow to bright yellow (Fig. 8a–f), and the absorption
peak immediately shifts to 400 nm [49] indicating. the formation of 4-
nitrophenolate ions under alkaline condition [50].The absorption peak
at 400 nm did not shift over 2.5 h, which confirms that the 4-NP cata-
lytic reduction does not occur without a catalyst, i.e. sodium borohy-
dride cannot reduce 4-NP itself [32,51]. Nevertheless, after adding the
nanocatalyst to 4-nitrophenolate ion, the color changes quickly, and the
Fig. 6. N2 adsorption isotherm of (a) α-Fe2O3, (b) 5CF500, (c) 20CF500, and intensity of absorption peak at 400 nm gradually decreases with time: a
(d) 20CF700. new peak at 305 nm appears, corresponding to 4-aminophenol (4-AP)

6
A. Elfiad, et al. Journal of Environmental Chemical Engineering 8 (2020) 104214

Fig. 8. (a) UV–vis absorption spectra of a solution of 4-nitrophenol (4-NP) with and without NaBH4. (b) Time-dependent UV–vis absorption spectra for the reduction
of 4-NP with NaBH4 without catalyst. (c-f) Time-dependent UV–vis absorption spectra for the reduction of 4-NP in the presence of: α-Fe2O3 (c), 5CFT500 (d),
20CFT500 (e), 20CFT700 (f).

Table 3 order is correct. We calculated the rate constants (k) dividing the ap-
Activity tests: reduction of 4-nitrophenol to 4-aminophenol at room tempera- parent rate constant (kapp) by the initial mass of the catalyst (Eq. (4)):
ture.
k=k app/ initial catalysts’ mass (4)
Catalysts Reaction times Conversion (%) kapp(s−1) K (s−1. kg−1)
(s) The rate constant of 20CF700 is the highest (Table 3), this indicates
that increasing of the Cu amount and the calcination temperature in-
α-Fe2O3 250 99.51 0.011 ± 0.0025 11 ± 0.0025
5CF500 115 99.86 0.039 ± 0.0021 39 ± 0.0021 creases the reaction rate. Indeed, increasing the reaction temperature
20CF500 80 99.85 0.056 ± 0.0012 56 ± 0.0012 increases the crystallinity of the sample, i.e. the concentration of the
20CF700 50 99.93 0.090 ± 0.0033 90 ± 0.0033 active phase. The reaction rate constants reported in the present work
for the 4-NP reduction are higher than the other catalysts reported in
literature, whereby expensive metal composites are usually the pre-
(Fig. 8c–f) [52]. UV–vis spectra exhibit an isosbestic point (Fig. 8d) for cursors (Table 4).
the two absorption bands indicating that the 4-NP gradually converts to The Langmuir–Hinshelwood (LH) model was proposed as me-
4-AP with no byproduct [53]. The Cu/ α-Fe2O3 catalysts are more ac- chanism for our reaction. The reduction of 4-NP consists of three me-
tive than α-Fe2O3 (Table 3). The time to complete the reaction de- chanistic steps (Fig. 10): the diffusion and adsorption of 4-NP on to the
creases with increasing of Cu amount. 4-NP completely reduces to 4-AP metal catalysts surface, and the electron transfer from borohydride ion
in 250 s, 115 s, 80 s, and 50 s for α-Fe2O3, 5CF500, 20CF500, and BH4− (donor) to 4-NP (acceptor) by metal catalysts [56]. Here, Fe and
20CF700, respectively. Cu metal particles of catalysts facilitate the electron transfer to the 4-
NP, which eliminates the reaction kinetic barrier [50]. BH4− adsorbs on
3.8. Reaction kinetics the catalyst's surface and liberates an hydrogen as hydride. The hydride
anion transfers to the Fe/NPs or Cu/NPs, yielding a metal-hydride
The concentration of NaBH4 was much higher than that of 4-NP and complex. Simultaneously, the nitro group of 4-nitrophenolate ion re-
thus considered as constant during the course of the reaction. duces by capturing the hydride from the metal-hydride complex
Therefore, the reduction reactions were assumed to be pseudo first [57,58]. Eventually, 4-AP forms from 4-hydroxylaminophenol (stable
order with respect to the concentration of 4-NP [54] (Eq. (3)). More- intermediate) after three hydro-deoxygenation reactions. The last step
over, a NaBH4 excess protects 4-AP from oxidation [55][56]. is the desorption of 4-AP from the surface of catalyst [59].
According to the Table 3, the Cu/α-Fe2O3 catalysts show relatively
ln Ct/C0=-ln At/A0=kapp*t (3)
higher catalytic activity than α-Fe2O3 alone. It is suggested that the α-
kapp is the apparent reaction constant. The slope of the plot of ln Ct/C0 Fe2O3NPs as support play an active part in the catalysis, yielding a
versus t determined the kinetic parameters (Fig. 9a-b). The linearity synergistic effect with copper nanoparticles. In addition, the α-Fe2O3
obtained in Fig. 9a confirms that the assumption of the pseudo first support offered a large number of active sites which were prevented the

7
A. Elfiad, et al. Journal of Environmental Chemical Engineering 8 (2020) 104214

Fig. 9. (a) ln (At/A0) versus time for the reduction of 4-NP (b) At/A0 versus time for the reduction of 4-NP.

Cu nanoparticles from forming agglomerates, leading to the uniform octahedral sites expose metal cations on the surface and act as a
dispersion of the CuNPs which are responsible for the high catalytic medium to transfer electrons from BH4− to 4-NP. In CuFe2O4 both
activity exhibited by Cu/α-Fe2O3 materials. Since, the lowest crystal- Cu2+ and Fe3+ are present, and the transfer of electron between Cu+-
lites size (calculate by XRD) and the smallest particle size (Table 2) of Cu2+ and Fe2+-Fe3+ ion pairs, enhances its catalytic activity [69].Our
the Cu-Fe catalysts facilitate the electron transfer and accelerate the catalytic system has higher activity compared to noble metal catalysts
kinetic reaction. (Table 4). Moreover its synthesis is unexpensive and ecofriendly.
It can be observed that the apparent rate constant (kapp) increases in
the order Kapp 5CF500 < Kapp 20CF500 < Kapp 20CF700, this classification 3.9. Catalysts recycle
reveal that (i) No correlation between the specific surface area and the
catalytic activity. (ii) The highest catalytic activity was related to the All samples completely convert 4-NP to 4-AP and they are stable
highest copper amount (copper %Wt) on hematite surface, which has within the first two runs (Fig. 11). After five cycles, the stability de-
more available actives sites with appropriate surface areas. (iii) The creases and the conversions drop to 86.5 %, 88.7 %, 87.3 %, and 88.1 %
excellence catalytic performance directly correlated the pores radius for α-Fe2O3, 5CF500, 20CF500, and 20CF700, respectively. We ascribe the
values (Table 3). As can be seen from Fig. S2, the rate constants increase decrease to the loss of catalyst during washing and separation and/or
with the pores diameter of catalysts, these results indicated that the the aggregation and destruction of the sample's surface structure by
highest pores values have the ability to accept the 4-nitrophenol mo- external forces, such as centrifugation, washing, and drying of the
lecules leading the high adsorption on catalyst surface. This is im- catalyst during the successive recycling steps.
portant since it is acknowledged that good adsorption is an important
step to accelerate the reaction kinetic and facilitate the 4-NP reduction.
4. Conclusion
As a result, it is wise comprehending that the pores diameter affects the
reduction reaction and the probability of acquiring the reaction not
We synthesized α-Fe2O3 via a simple thermal treatment of natural
only on the surface but also inside the mesopore structure.
goethite. We supported Cu over hematite with impregnation at ambient
The superior activity of 20CF700 compared to 20CF500 ascribes to
temperature. Cu amount influences the textural and structural propri-
the copper ferrite CuFe2O4 (Fig. 2), which is more active in the re-
eties of the catalyst, but all the samples were mesoporous. Cu/α-Fe2O3
duction of 4-NP compared to single metal phase [68] due to the metal
is active in the 4-nitrophenol reduction (initial concentration of 5 mmol
ions distribution among tetrahedral and octahedral sites. The
L−1) to 4-aminophenol in the presence of NaBH4. The reaction

Table 4
Literature data on the reduction of 4-nitrophenol by NaBH4 over different catalysts.
Catalyst 4-NPconcentration Catalysts Times (min) Kapp reference

Ag 4 mM 0.5 g/l 30 0,47.10−3 s-1 [60]


Ag–Au 4 mM 0.5 g/l 4 3,8.10−3 s-1 [60]
RGS@AuNs 0.2 mmol 0.8 mg 5 15,01.10−3 s-1 [61]
Ag/iron oxide 7,95.10−5M 4.10−4g 35 0.305 min−1 [62]
Au/Fe3O4 1,10−3 M 0.49 mg 7 10.10−2 min-1 [63]
Pt/Fe2O3 5 mg 12 0.24mn−1 [64]
Cu 5mol.l−1 1mg 3,5 3,8.10−3mn-1 [65]
CuO 2.5 mM 5 mg 22 / [31]
Cu-Fe 10_4 mol L_1 4 mg L_1 5 0.69 min_1 [66]
Cu/graphene 1 mM 1 mg 6 / [28]
CuO/ɤ-Al2O3 4,3.10−5 mol/L 0.05 g 12 17,37.10−2 min1 [35]
CuO/ZnO 2.5 mM 5 mg 3 / [31]
Cu@Fe3O4 1.5 mM 0.5 mg.mL_1 0.66 0.047s−1 [67]
Cu1.3/Fe3O4 5 mM L_1 1 mg mL_1 3.5 0.71 min−1 [65]
5CF500 1mg 5 ×10−3 1.5 0.039 s−1 This work
20CF500 1mg 5 ×10−3 1.3 0.056 s−1 This work
20CF700 1mg 5 ×10−3 0.8 0.090 s−1 This work

8
A. Elfiad, et al. Journal of Environmental Chemical Engineering 8 (2020) 104214

Fig. 10. Scheme of the 4-NP reduction to 4AP by NaBH4.

Acknowledgements

The authors gratefully acknowledge the support of the Natural


Sciences and Engineering Research Council of Canada (NSERC). This
research was undertaken, in part, thanks to funding from the Canada
Research Chairs program.

Appendix A. Supplementary data

Supplementary material related to this article can be found, in the


online version, at doi:https://doi.org/10.1016/j.jece.2020.104214.

References

[1] S. Ghosh, R. Das, I.H. Chowdhury, P. Bhanja, M.K. Naskar, Rapid template-free
synthesis of an air-stable hierarchical copper nanoassembly and its use as a reusable
catalyst for 4-nitrophenol reduction, RSC Adv. 5 (2015) 101519–101524.
[2] P.R. Prasad, S. Kanchi, E.B. Naidoo, In-vitro evaluation of copper nanoparticles
cytotoxicity on prostate cancer cell lines and their antioxidant, sensing and catalytic
activity: One-pot green approach, J. Photochem. Photobiol. B: Biol. 161 (2016)
375–382, https://doi.org/10.1016/j.jphotobiol.2016.06.008.
[3] A.D. Verma, R.K. Mandal, I. Sinha, Kinetics of p-Nitrophenol reduction catalyzed by
Fig. 11. Reusability of catalysts. Uncertainty is within ± 1 %. PVP stabilized copper nanoparticles, Catal. Lett. 145 (2015) 1885–1892.
[4] J.P. Miller-Schulze, M. Paulsen, T. Kameda, A. Toriba, K. Hayakawa, B. Cassidy,
L. Naeher, M.A. Villalobos, C.D. Simpson, Nitro-PAH exposures of occupationally-
completed (conversion > 99 %) within 250 s, 115 s, 80 s, and 50 s over exposed traffic workers and associated urinary 1-nitropyrene metabolite con-
1 mg of catalyst for α-Fe2O3, 5CF500, 20CF500, 20CF700 respectively. centrations, J. Environ. Sci. 49 (2016) 213–221.
[5] J. Michałowicz, W. Duda, Phenols–sources and toxicity, Polish J. Environ. Stud. 16
The reaction rate constants (k) were between 0.11 s-1 to 0.90 s−1. All (2007).
catalysts exhibited stability, as well as recyclability over 5 runs. [6] S. Saha, A. Pal, S. Kundu, S. Basu, T. Pal, Photochemical green synthesis of calcium-
Natural hematite and Cu supported on natural hematite are eco- alginate-stabilized Ag and Au nanoparticles and their catalytic application to 4-
nitrophenol reduction, Langmuir 26 (2010) 2885–2893.
nomical and eco-friendly catalyst for environmentally friendly appli- [7] F.R. Zaggout, N.A. Ghalwa, Removal of o-nitrophenol from water by electro-
cations as pollutant degradation. Our results can be extrapolated to 1- chemical degradation using a lead oxide/titanium modified electrode, J. Environ.
Manage. 86 (2008) 291–296.
and 2- nitrophenols, which are other pollutants found in wastewaters.
[8] M.J. Vaidya, S.M. Kulkarni, R.V. Chaudhari, Synthesis of p-aminophenol by cata-
lytic hydrogenation of p-nitrophenol, Org. Process Res. Dev. 7 (2003) 202–208.
[9] A. Zhang, N. Wang, J. Zhou, P. Jiang, G. Liu, Heterogeneous Fenton-like catalytic
removal of p-nitrophenol in water using acid-activated fly ash, J. Hazard. Mater.
Declaration of Competing Interest 201 (2012) 68–73.
[10] R. Andreozzi, V. Caprio, A. Insola, G. Longo, V. Tufano, Photocatalytic oxidation of
4‐nitrophenol in aqueous TiO2 slurries: an experimental validation of literature
The authors declare that they have no known competing financial kinetic models, J. Chem. Technol. Biotechnol. Int. Res. Process. Environ. Clean
interests or personal relationships that could have appeared to influ- Technol. 75 (2000) 131–136.
[11] A. Hakki, R. Dillert, D. Bahnemann, Photocatalytic conversion of nitroaromatic
ence the work reported in this paper.
compounds in the presence of TiO2, Catal. Today 144 (2009) 154–159.
[12] Z.I. Bhatti, H. Toda, K. Furukawa, p-Nitrophenol degradation by activated sludge
attached on nonwovens, Water Res. 36 (2002) 1135–1142.
[13] G. Xue, M. Gao, Z. Gu, Z. Luo, Z. Hu, The removal of p-nitrophenol from aqueous

9
A. Elfiad, et al. Journal of Environmental Chemical Engineering 8 (2020) 104214

solutions by adsorption using gemini surfactants modified montmorillonites, Chem. Prot. 91 (2013) 489–494.
Eng. J. 218 (2013) 223–231. [40] A. Biabani-Ravandi, M. Rezaei, Low temperature CO oxidation over Fe–Co mixed
[14] L.G. Devi, S.G. Kumar, K.M. Reddy, C. Munikrishnappa, Photo degradation of oxide nanocatalysts, Chem. Eng. J. 184 (2012) 141–146.
Methyl Orange an azo dye by Advanced Fenton Process using zero valent metallic [41] N.H. Amin, L.I. Ali, S.A. El-Molla, A.A. Ebrahim, H.R. Mahmoud, Effect of Fe2O3
iron: Influence of various reaction parameters and its degradation mechanism, J. precursors on physicochemical and catalytic properties of CuO/Fe2O3 system,
Hazard. Mater. 164 (2009) 459–467. Arab. J. Chem. 9 (2016) S678–S684.
[15] R. Rajesh, E. Sujanthi, S.S. Kumar, R. Venkatesan, Designing versatile hetero- [42] A.S. Teja, P.-Y. Koh, Synthesis, properties, and applications of magnetic iron oxide
geneous catalysts based on Ag and Au nanoparticles decorated on chitosan func- nanoparticles, Prog. Cryst. Growth Charact. Mater. 55 (2009) 22–45.
tionalized graphene oxide, Phys. Chem. Chem. Phys. 17 (2015) 11329–11340. [43] J. Hua, J. Gengsheng, Hydrothermal synthesis and characterization of mono-
[16] Y. Kojima, K. Suzuki, K. Fukumoto, M. Sasaki, T. Yamamoto, Y. Kawai, H. Hayashi, disperse α-Fe2O3 nanoparticles, Mater. Lett. 63 (2009) 2725–2727.
Hydrogen generation using sodium borohydride solution and metal catalyst coated [44] B.M. Reddy, K.N. Rao, Copper promoted ceria–zirconia based bimetallic catalysts
on metal oxide, Int. J. Hydrogen Energy 27 (2002) 1029–1034. for low temperature soot oxidation, Catal. Commun. 10 (2009) 1350–1353.
[17] T.R. Mandlimath, B. Gopal, Catalytic activity of first row transition metal oxides in [45] W. Gao, S. Li, H. Huo, F. Li, Y. Yang, X. Li, X. Wang, Y. Tang, R. Li, Investigation of
the conversion of p-nitrophenol to p-aminophenol, J. Mol. Catal. A Chem. 350 the crystal structure of Cu-Fe bimetal oxide and their catalytic activity for the
(2011) 9–15. Baeyer–Villiger oxidation reaction, Mol. Catal. 439 (2017) 108–117.
[18] P. Zhao, X. Feng, D. Huang, G. Yang, D. Astruc, Basic concepts and recent advances [46] U.C. Rajesh, V.S. Pavan, D.S. Rawat, Copper NPs supported on hematite as mag-
in nitrophenol reduction by gold-and other transition metal nanoparticles, Coord. netically recoverable nanocatalysts for a one-pot synthesis of aminoindolizines and
Chem. Rev. 287 (2015) 114–136. pyrrolo [1, 2-a] quinolines, RSC Adv. 6 (2016) 2935–2943.
[19] A. Shukla, R.K. Singha, T. Sasaki, R. Bal, Nanocrystalline Pt-CeO 2 as an efficient [47] A. Lassoued, M.S. Lassoued, B. Dkhil, A. Gadri, S. Ammar, Structural, optical and
catalyst for a room temperature selective reduction of nitroarenes, Green Chem. 17 morphological characterization of Cu-doped α-Fe2O3 nanoparticles synthesized
(2015) 785–790. through co-precipitation technique, J. Mol. Struct. 1148 (2017) 276–281.
[20] M. Mourya, D. Choudhary, A.K. Basak, C.S.P. Tripathi, D. Guin, [48] A.E.-A.A. Said, M.M.M.A. El-Wahab, S.A. Soliman, M.N. Goda, Synthesis and
Ag‐nanoparticles‐embedded filter paper: an efficient dip catalyst for aromatic ni- characterization of mesoporous Fe–Co mixed oxide nanocatalysts for low tem-
trophenol reduction, intramolecular cascade reaction, and methyl orange de- perature CO oxidation, Process Saf. Environ. Prot. 102 (2016) 370–384.
gradation, ChemistrySelect 3 (2018) 2882–2887. [49] A. Elfiad, F. Galli, A. Djadoun, M. Sennour, S. Chegrouche, L. Meddour-Boukhobza,
[21] G. Zheng, L. Polavarapu, L.M. Liz-Marzán, I. Pastoriza-Santos, J. Pérez-Juste, Gold D.C. Boffito, Natural α-Fe2O3 as an efficient catalyst for the p-nitrophenol reduc-
nanoparticle-loaded filter paper: a recyclable dip-catalyst for real-time reaction tion, Mater. Sci. Eng. B 229 (2018) 126–134.
monitoring by surface enhanced Raman scattering, Chem. Commun. 51 (2015) [50] N. Pradhan, A. Pal, T. Pal, Silver nanoparticle catalyzed reduction of aromatic nitro
4572–4575. compounds, Colloids Surf. A Physicochem. Eng. Asp. 196 (2002) 247–257.
[22] J. Zhang, G. Chen, D. Guay, M. Chaker, D. Ma, Highly active PtAu alloy nano- [51] L. Zhang, Z. Liu, Y. Wang, R. Xie, X.-J. Ju, W. Wang, L.-G. Lin, L.-Y. Chu, Facile
particle catalysts for the reduction of 4-nitrophenol, Nanoscale. 6 (2014) immobilization of Ag nanoparticles on microchannel walls in microreactors for
2125–2130. catalytic applications, Chem. Eng. J. 309 (2017) 691–699.
[23] Z.D. Pozun, S.E. Rodenbusch, E. Keller, K. Tran, W. Tang, K.J. Stevenson, [52] H. Lu, H. Yin, Y. Liu, T. Jiang, L. Yu, Influence of support on catalytic activity of Ni
G. Henkelman, A systematic investigation of p-nitrophenol reduction by bimetallic catalysts in p-nitrophenol hydrogenation to p-aminophenol, Catal. Commun. 10
dendrimer encapsulated nanoparticles, J. Phys. Chem. C. 117 (2013) 7598–7604. (2008) 313–316.
[24] H. Li, J. Liao, Y. Du, T. You, W. Liao, L. Wen, Magnetic-field-induced deposition to [53] J. Lee, J.C. Park, H. Song, A nanoreactor framework of a Au@ SiO2 yolk/shell
fabricate multifunctional nanostructured Co, Ni, and CoNi alloy films as catalysts, structure for catalytic reduction of p‐nitrophenol, Adv. Mater. 20 (2008)
ferromagnetic and superhydrophobic materials, Chem. Commun. 49 (2013) 1523–1528.
1768–1770. [54] Z. Dong, X. Le, X. Li, W. Zhang, C. Dong, J. Ma, Silver nanoparticles immobilized on
[25] S. Li, J. Hao, P. Ning, C. Wang, K. Li, L. Tang, X. Sun, D. Zhang, Y. Mei, Y. Wang, fibrous nano-silica as highly efficient and recyclable heterogeneous catalyst for
Preparation of Cu A Fe nanocomposites loaded diatomite and their excellent per- reduction of 4-nitrophenol and 2-nitroaniline, Appl. Catal. B Environ. 158 (2014)
formance in simultaneous adsorption / oxidation of hydrogen sulfide and phos- 129–135.
phine at low temperature, Sep. Purif. Technol. 180 (2017) 23–35, https://doi.org/ [55] K. Hayakawa, T. Yoshimura, K. Esumi, Preparation of gold− dendrimer nano-
10.1016/j.seppur.2017.02.044. composites by laser irradiation and their catalytic reduction of 4-nitrophenol,
[26] W. Liu, K. Tian, H. Jiang, H. Yu, Harvest of Cu NP anchored magnetic carbon Langmuir 19 (2003) 5517–5521.
materials from Fe / Cu preloaded biomass: their, Green Chem. 16 (2014) [56] K. Esumi, R. Isono, T. Yoshimura, Preparation of PAMAM− and PPI− metal (silver,
4198–4205, https://doi.org/10.1039/c4gc00599f. platinum, and palladium) nanocomposites and their catalytic activities for reduc-
[27] X. Yang, H. Zhong, Y. Zhu, H. Jiang, J. Shen, J. Huang, C. Li, Highly efficient tion of 4-nitrophenol, Langmuir 20 (2004) 237–243.
reusable catalyst based on silicon nanowire arrays decorated with copper nano- [57] W.-J. Liu, K. Tian, H. Jiang, H.-Q. Yu, Harvest of Cu NP anchored magnetic carbon
particles, J. Mater. Chem. A 2 (2014) 9040–9047. materials from Fe/Cu preloaded biomass: their pyrolysis, characterization, and
[28] L. Jin, G. He, J. Xue, T. Xu, H. Chen, Cu/graphene with high catalytic activity catalytic activity on aqueous reduction of 4-nitrophenol, Green Chem. 16 (2014)
prepared by glucose blowing for reduction of p-nitrophenol, J. Clean. Prod. 161 4198–4205.
(2017) 655–662. [58] K. Layek, M.L. Kantam, M. Shirai, D. Nishio-Hamane, T. Sasaki, H. Maheswaran,
[29] M.J.D. Mahboub, S. Lotfi, J.-L. Dubois, G.S. Patience, Gas phase oxidation of 2- Gold nanoparticles stabilized on nanocrystalline magnesium oxide as an active
methyl-1, 3-propanediol to methacrylic acid over heteropolyacid catalysts, Catal. catalyst for reduction of nitroarenes in aqueous medium at room temperature,
Sci. Technol. 6 (2016) 6525–6535. Green Chem. 14 (2012) 3164–3174.
[30] M.J.D. Mahboub, M. Rostamizadeh, J. Dubois, G.S. Patience, Partial oxidation of 2- [59] M. Wang, D. Tian, P. Tian, L. Yuan, Synthesis of micron-SiO2@ nano-Ag particles
methyl-1, 3-propanediol to methacrylic acid: experimental and neural network and their catalytic performance in 4-nitrophenol reduction, Appl. Surf. Sci. 283
modeling, RSC Adv. 6 (2016) 114123–114134. (2013) 389–395.
[31] M. Bordbar, N. Negahdar, M. Nasrollahzadeh, Melissa Officinalis L. leaf extract [60] H. Fu, X. Yang, X. Jiang, A. Yu, Bimetallic Ag–Au nanowires: synthesis, growth
assisted green synthesis of CuO/ZnO nanocomposite for the reduction of 4-ni- mechanism, and catalytic properties, Langmuir 29 (2013) 7134–7142.
trophenol and Rhodamine B, Sep. Purif. Technol. 191 (2018) 295–300. [61] S.K. Maji, A. Jana, Two-dimensional nanohybrid (RGS@AuNPs) as an effective
[32] A. Elfiad, D.C. Boffito, S. Khemassia, F. Galli, S. Chegrouche, L. Meddour- catalyst for the reduction of 4-nitrophenol and photo-degradation of methylene
Boukhobza, Eco-friendly synthesis from industrial wastewater of Fe and Cu nano- blue dye, New J. Chem. 42 (2017,3326-3332), https://doi.org/10.1039/
particles over NaX zeolite and activity in 4-nitrophenol reduction, Can. J. Chem. C6NJ04062D.
Eng.9999 (2018, 1-10), https://doi.org/10.1002/cjce.23083. [62] J.-R. Chiou, B.-H. Lai, K.-C. Hsu, D.-H. Chen, One-pot green synthesis of silver/iron
[33] Z. Wang, S. Zhai, J. Lv, H. Qi, W. Zheng, B. Zhai, Q. An, RSC Advances Versatile oxide composite nanoparticles for 4-nitrophenol reduction, J. Hazard. Mater. 248
hierarchical Cu / Fe 3 O 4 nanocatalysts for e ffi cient degradation of organic dyes (2013) 394–400.
prepared by a facile, controllable hydrothermal method, RSC Adv. 5 (2015) [63] Y.-C. Chang, D.-H. Chen, Catalytic reduction of 4-nitrophenol by magnetically re-
74575–74584, https://doi.org/10.1039/C5RA16027H. coverable Au nanocatalyst, J. Hazard. Mater. 165 (2009) 664–669.
[34] M. Nasrollahzadeh, M. Atarod, S.M. Sajadi, Green synthesis of the Cu/Fe3O4 na- [64] P. Zhang, X. Yang, H. Peng, D. Liu, H. Lu, J. Wei, J. Gui, Magnetically recoverable
noparticles using Morinda morindoides leaf aqueous extract: a highly efficient hierarchical Pt/Fe2O3 microflower: superior catalytic activity and stability for re-
magnetically separable catalyst for the reduction of organic dyes in aqueous duction of 4-nitrophenol, Catal. Commun. 100 (2017) 214–218.
medium at room temperature, Appl. Surf. Sci. 364 (2016) 636–644. [65] Z. Wang, S. Zhai, J. Lv, H. Qi, W. Zheng, B. Zhai, Q. An, Versatile hierarchical Cu/Fe
[35] S.U. Nandanwar, M. Chakraborty, Synthesis of colloidal CuO/γ-Al2O3 by micro- 3 O 4 nanocatalysts for efficient degradation of organic dyes prepared by a facile,
emulsion and its catalytic reduction of aromatic nitro compounds, Chin. J. Catal. 33 controllable hydrothermal method, RSC Adv. 5 (2015) 74575–74584.
(2012) 1532–1541. [66] H. Zhang, Y. Ni, Y. Zhong, H. Wu, M. Zhai, Fast electrodeposition, influencing
[36] A. Elfiad, F. Galli, A. Djadoun, M. Sennour, S. Chegrouche, L. Meddour-Boukhobza, factors and catalytic properties of dendritic Cu–M (M= Ni, Fe, Co) microstructures,
D.C. Boffito, Natural α-Fe2O3 as an efficient catalyst for the p-nitrophenol reduc- RSC Adv. 5 (2015) 96639–96648.
tion, Mater. Sci. Eng. B Solid-State Mater. Adv. Technol. 229 (2018), https://doi. [67] W. Zhao, S. Zhang, J. Ding, Z. Deng, L. Guo, Chemical enhanced catalytic ozonation
org/10.1016/j.mseb.2017.12.009. for NOx removal with CuFe 2 O 4 nanoparticles and mechanism analysis, Journal
[37] N.M. Khalil, M.M.S. Wahsh, E.E. Saad, Hydrothermal extraction of α-Fe2O3 na- Mol. Catal. A Chem. 424 (2016) 153–161, https://doi.org/10.1016/j.molcata.2016.
nocrystallite from hematite ore, J. Ind. Eng. Chem. 21 (2015) 1214–1218. 08.007.
[38] H. Liu, T. Chen, X. Zou, C. Qing, R.L. Frost, Thermal treatment of natural goethite: [68] M. Kooti, M. Afshari, Magnetic cobalt ferrite nanoparticles as an efficient catalyst
thermal transformation and physical properties, Thermochim. Acta 568 (2013) for oxidation of alkenes, Sci. Iran. 19 (2012) 1991–1995.
115–121. [69] E. Casbeer, V.K. Sharma, X.-Z. Li, Synthesis and photocatalytic activity of ferrites
[39] A. Biabani-Ravandi, M. Rezaei, Z. Fattah, Study of Fe–Co mixed metal oxide na- under visible light: a review, Sep. Purif. Technol. 87 (2012) 1–14.
noparticles in the catalytic low-temperature CO oxidation, Process Saf. Environ.

10

You might also like