You are on page 1of 9

Journal of Magnetism and Magnetic Materials 342 (2013) 38–46

Contents lists available at SciVerse ScienceDirect

Journal of Magnetism and Magnetic Materials


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

Synthesis, characterization and magnetic behavior of Mg–Fe–Al mixed


oxides based on layered double hydroxide
Angélica C. Heredia a,n, Marcos I. Oliva b,c, Ulises Agú a,c, Carlos I. Zandalazini c,d, Sergio
G. Marchetti e, Eduardo R. Herrero a, Mónica E. Crivello a
a
Universidad Tecnológica Nacional, Facultad Regional Córdoba-CITeQ, Maestro López esq. Cruz Roja Argentina, Ciudad Universitaria, 5016 Córdoba,
Argentina
b
IFEG, Universidad Nacional de Córdoba, Córdoba, Argentina
c
CONICET, Argentina
d
INFIQC, FCQ Universidad Nacional de Córdoba, Córdoba, Argentina
e
CINDECA, UNLP, Buenos Aires, Argentina

art ic l e i nf o a b s t r a c t

Article history: In the present work, Mg–Al–Fe layered double hydroxides were prepared by coprecipitation reaction
Received 22 November 2012 with hydrothermal treatment. The characterization of precursors and their corresponding calcinated
Received in revised form products (mixed oxides) were carried out by X ray diffraction, X-ray photoelectron spectroscopy (XPS),
12 April 2013
termogravimetric analysis and differential scanning calorimetry, diffuse reflectance UV–vis spectroscopy,
Available online 25 April 2013
specific surface area, Mössbauaer and magnetic properties. The Fe3+ species were observed in
Keywords: tetrahedrally and octahedrally coordination in brucite layered. The XPS analysis shows that the Fe3+
Layered doubled hydroxides ions can be found in two coordination environments (tetrahedral and octahedral) as mixed oxides, and as
Coprecipitation spinel-structure. Oxides show a decrease in the specific surface areas when the iron loading is increased.
Mixed oxides
The magnetic and Mössbauaer response show that MgAlFe mixed oxides are different behaviours such as
Magnetic properties
different population ratios of ferromagnetic, weak-ferromagnetic, paramagnetic and superparamagnetic
phases. The better crystallization of spinel structure with increased temperature, is correlated with the
improved magnetic properties.
& 2013 Published by Elsevier B.V.

1. Introduction large variety of anions that can be incorporated between the


brucite-like layers, the high anionic exchange capacity of these
Mixed oxides with iron obtained by thermal decomposition of materials, and the large specific surface area values, LDH can be
layered double hydroxides (LDH) have been attracting interest successfully used as matrices for tailoring specific organic–clay
because of their exclusive properties and potential applications in hybrid structures with new potential applications in pharmaceu-
numerous technological fields. In particular, they are mainly ticals or as new biocompatible materials [8–10]. It is possible to
important in medicine, as adsorbents, anion exchangers, and most synthesize LDH containing more of two cations in the layers.
importantly as basic catalysts. Besides, if iron is used, the final To use LDH precursors produces an excellent dispersion of metals
compounds show interesting magnetic properties [1–5]. LDH are compounds on a matrix of magnesium and aluminum. Thermal
represented by the general formula [M2+ (1−x) Mx
3+
(OH)2] (An−x/n)  decomposition of LDH containing iron start at 300 1C and leads to
mH2O], where the divalent ion may be Mg , Co2+, Zn2+, Ni2+,
2+
mixed oxides. The final decomposition product is a mixture of
Mn2+, and the trivalent ion may be Al3+, Fe3+, Cr3+. Their structure spinel and oxides depending of M2+/M3+ ratio and synthesis
consists of brucite-type layers, where the substitution of M2+ with method. This allows obtaining a wide range of magnetic proper-
M3+ cations results into a net positive charge, compensated by ties, according to the variable cation distribution in the crystalline
interlayer anions (A). There is also water crystallization into the structure because magnetic properties are strongly dependent of
interlayer region [6,7]. The interlayer distance depends on both, spin structure. Thus, magnesium-ferrite aluminate is the first
the nature of anions and the state of hydration. Because of the system where canting of Fe3+ spins on A-site has been observed
when Fe3+ spins on B-site has a collinear spin structure [11].
Recent studies have indicated interesting magnetic properties for
n novel drugs–anionic clay structures, strongly dependent on synth-
Corresponding author. Tel.: +543514690585.
E-mail addresses: angelicacheredia@gmail.com, esis conditions, internal structure and the drug nature stability
aheredia@scdt.frc.utn.edu.ar (A.C. Heredia). [12,13]. Qi et al. [14] reported the synthesis of Co0.9Mn0.1Fe2O4

0304-8853/$ - see front matter & 2013 Published by Elsevier B.V.


http://dx.doi.org/10.1016/j.jmmm.2013.04.057
A.C. Heredia et al. / Journal of Magnetism and Magnetic Materials 342 (2013) 38–46 39

using the LDH precursors, which displayed magnetostrictive scan speed of 1/41 min in 2θ and interfaced to a DACO-MP data
properties that were similar to the traditional ceramic method. acquisition microprocessor provided with Diffract/AT software.
Zhang et al. [15] investigated the effects of the high magnetic field The diffraction pattern was identified by comparing with those
on products obtained by calcinations of Co–Fe LDH precursors included in the JCPDS (Joint Committee of Powder Diffraction
under the applied field (10 T) at different temperatures. Standards) data base.
In the last years, some oxides mixed have been successfully Inductively coupled plasma (ICP) optical emission spectroscopy
prepared using LDH as precursors [16,17], but they were synthe- was used for the determination of the metal content in the oxides.
sized varying the temperature, cations ratio or pH value. The measurements were performed with a Varian Spectra AA.
In this work, we report that mixed oxides can be prepared from Diffuse reflectance UV–vis (DRUV-vis) spectroscopic measure-
Mg2+/Fe3+/Al3+–LDH precursors. They were synthesized by copre- ments of the LDH precursors were recorded using an Optronics OL
cipitation at 60 1C under ambient atmosphere, the pH value was 750-427 spectrometer in the wavelength range 200–600 nm.
9 and hydrothermal method was used. The content of iron was Termogravimetric analysis (TGA) and differential scanning
changing in order to assess this effect on the structural properties calorimetry (DSC) were carried out in a TA instrument SDT Q600
of the LDH and magnetic behaviors of the calcinations materials. (0–1500 1C) instrumentation apparatus in a flowing air atmo-
Due to the properties of these materials, they can be used as sphere. Approximately 30 mg of sample were loaded and heated
catalysts in reactions of ethylbenzene dehydrogenation [18], at a rate of 10 1C/min up to 550 1C.
photocatalysis [19], catalysis for remediating the contaminated X-ray photoelectron spectroscopy (XPS) analyses were carried
waters with organic compounds in solution [20,21]. out using an ESCA (VG microtech) spectrometer with a non-
monochromatic Mg Kα radiation (υ¼ 1253.6 eV) as the excitation
source. High-resolution spectra were recorded in the constant pass
2. Experimental energy mode at 20 eV, using a 720 mm diameter analysis area.
Under these conditions, the Au 4f7/2 line was recorded with 1.16 eV
2.1. Sample preparation FWHM (full width at half maximum) at a binding energy (BE) of
84.0 eV. The spectrometer energy scale was calibrated using Cu
The Mg, Al, Fe LDH precursors were prepared by coprecipita- 2p3/2, Ag 3d5/2 and Au 4f7/2 photoelectron lines at 932.7, 368.3 and
tion at low supersaturation method at constant pH (9 70.5), with 84.0 eV, respectively. Charge referencing was done against adven-
M2+/M3+ ¼3 M ratio constant [22] while the molar ratios Al3+/Fe3+ titious carbon (C 1s, 284.8 eV). The pressure in the analysis
was changed in order to assess this effect on the structural and chamber was maintained lower than 10–9 Torr. PHI ACCESS
magnetic properties of the LDH and their calcinations products. ESCA-V6.0 F software package was used for acquisition and data
Two solutions, A and B, were prepared. In order to obtain A analysis. A Shirley-type background was subtracted from the
solution, Mg(NO3)2  6H2O, Al(NO3)3  9H2O and Fe(NO3)3  9H2O were signals.
dissolved together in distilled water. The amounts of nitrates were Specific surface area was determined by the BET method, which
selected to obtain the total cation concentration of 0.7 M. The B was recorded on a Micromeritics ASAP 2000 instrument. The
solution contains 0.085 M of Na2CO3 (according to the relation precursors were degassed at 200 1C and the calcined materials at
of [CO32−]¼0.5 [M3+]); both solutions were added simultaneously 390 1C, both for 60 min.
to 30 mL of distilled water at drip rate of 50 mL/h, the solution was The magnetic measurements were performed in a commercial
continuously magnetically stirred, while the pH was maintained vibrating sample magnetometer (Lakeshore 7300) at room tem-
constant by adding NaOH 2 M. The coprecipitation was carried out at perature varying the applied field between 0 and 15000 Oe. The
60 1C. The gel was transferred into Teflon-lined stainless-steel powder samples were compacted applying 5 Tn to make a disc
autoclave and kept in an oven at 200 1C for 18 h, then was separated shaped sample of 5 mm diameter and 2 mm height to measure
by centrifugation at 2800 rpm, and washed with distilled water until magnetization as function of applied field (M vs H). Hysteresis
a sodium content lower than 0.13 wt%. Finally, the solid was dried loops were well fitted using the function:
overnight at 90 1C in air. All samples were calcined in atmosphere of  h i  
BH
air at 550 1C for 9 h. The HT100 sample was calcined at three different MðHÞ ¼ M s  ð1−αÞ  1−eð−H=Hs Þ þ α  L
T
temperatures to analyze the different structures formed. The LDH
precursors will be called HT100, HT75, HT50, HT25 and HT0, according where LððB  HÞ=TÞ is the Langevin's function [23], T is the system
to the iron loading (keeping in mind only cations M3+), and the temperature, α is the paramagnetic and/or superparamagnetic
calcined solids will be named CHT100, CHT75, CHT50, CHT25 and CHT0. fraction, (1−α) is the ferromagnetic fraction, Hs is the ferromag-
Compositions of the samples are presented in Table 1. netic saturation magnetic field, Ms is the total magnetization of
saturation. The B factor is an adjustable parameter, B ¼μ/k where μ
2.2. Sample characterization is the dipolar magnetic moment per paramagnetic molecule [11]
and k is the Boltzmann's constant. The parameters α, Ms y Hs are
The XRD powder patterns were collected on a Rigaku diffract- also obtained from the fitting.
ometer, using monochromatized Cu Kα radiation (λ ¼1.54 Ǻ) at a The SEM images were obtained using JEOL JSM-6380 LV
(accelerating voltage: 20 kV) and a Zeiss Supra 40.
Table 1 The Mössbauer spectra were obtained in transmission geome-
Chemical composition of the studied LDH. try with a 512-channel constant acceleration spectrometer.
A source of 57Co in Rh matrix of nominally 50 mCi was used.
LDH sample M3+ M2þ
%Fe3þ ¼ Fe3þ
 100
M3þ Al

þFe3þ Velocity calibration was performed against a 12-μm-thick α-Fe foil.
All isomer shifts (δ) mentioned in this paper are referred to this
Theoretical ICP Theoretical ICP
standard at 298 K. The Mössbauer spectra were evaluated using a
HT0 Al 3 2.82 0 0 commercial program with constraints named Recoil [24].
HT 25 Al–Fe 3 3.03 25 23.51 Lorentzian lines with equal widths were considered for each
HT 50 Al–Fe 3 3.22 50 45.02 spectrum component. The spectra were folded to minimize geo-
HT 75 Al–Fe 3 3.09 75 70.69
metric effects. In order to consider the existence of different iron
HT 100 Fe 3 2.42 100 100
environments, hyperfine magnetic field distributions were used.
40 A.C. Heredia et al. / Journal of Magnetism and Magnetic Materials 342 (2013) 38–46

3. Results and discussion After calcinations at 550 1C, the (0 0 3) and (0 0 6) reflections
disappeared (Fig. 2). The MgO phase (periclase PCPDFWIN
3.1. X-ray diffraction 78-0430) was detected in all samples. The XRD patterns showed
that Fe+3 species were crystallized in two structures: MgFe2O4
By X-ray diffraction, the hydrotalcite phase was detected in all (Magnesium Iron oxide PCPDFWIN 71-1232) and α-Fe2O3 (Hema-
precursor samples. Most of the peak positions were matched with tite PCPDFWIN 79-1741). The transformation of the phases pro-
the ICDD (International Centre for Diffraction Data) PCPDFWIN duced after calcinations are also demonstrated by the analysis of
data. the correspondent XRD patterns. Fig. 3 displays the XRD patterns
Powder X-ray diffraction patterns of precursors with different of HT100 sample heated in air at different temperatures. From this
(Al3+/Fe3+) molar ratio are shown in Fig. 1. The hydrotalcite phase XRD patterns, three peaks are observed at 35.41, 431, 62.21 for
(PCPDFWIN 70-2151) was observed in all samples. Two peaks samples calcined corresponding to characteristic diffractions of
located at 2θ of 11.61 and 23.41 are associated to diffraction by MgFe2O4, MgO and Fe2O3. As the calcination temperature further
(0 0 3) and (0 0 6) planes characteristics of hydrotalcite phase. The increases, the intensities of these peaks are increased together
secondary phase MgCO3 (magnesium carbonate PCPDFWIN 83- with the peaks at 301 and 571 of MgFe2O4 spinel structure,
1761) was observed only in HT75. indicating that this phase is enhanced at high temperature [25].

3.2. DRUV–vis spectroscopy

The DRUV–vis spectra of the LDH precursors are shown in


(003)

(006)

Fig. 4. All precursor exhibited band at ∼210 and ∼275 nm. They
HT0 could be assigned to Fe3+ in tetrahedrally and octahedrally [26,27]
(009)

(015)

(018)

(110)
(113)

coordination in brucite layered. The band between∼350 and


550 nm is assigned to octahedral Fe3+ present in small clusters,
isolated species and large particles [27]. The temperature during of
HT25
the coprecipitation (60 1C) and the aging by hydrothermal method
in the autoclave, can promote the clusters production and particles
Intensity (a.u)

of great size outside the laminar structure [7].


HT50 Fig. 5 show the diffuse reflectance UV–vis spectra of mixed oxides.
The two characteristic bands at ∼210 nm and ∼275 nm present in LDH
lose their definition and intensity, when the laminar structure
disappears in the calcined samples. In all samples the band at
HT75
∼290 nm was observed, indicating the presence of Fe3+ species. The
peak around 350 nm is assigned to isolated Fe3+ in either periclase Mg
(Fe, Al)O or MgFe2O4 spinel in mixed oxide [27]. Between450
HT100
and 550 nm is observed a band, which can probably be assigned to
the aggregated Fe3+ oxide clusters.
10 20 30 40 50 60 70
3.3. Measurement of specific surface areas

Fig. 1. X-ray diffraction patterns of the LDH precursors with different iron The specific surface areas of precursor and calcined samples
percentage. (▲) Magnesium Carbonate MgCO3. were determined by BET method. The values obtained are

CHT0 850 °C

CHT25
Intensity (a.u)

Intensity (a.u)

700 °C
CHT50

CHT75
550 °C

CHT100

10 20 30 40 50 60 70 80 30 40 50 60 70 80
2θ 2θ
Fig. 2. X-ray diffraction patterns of calcined samples. (●) spinel MgFe2O4, Fig. 3. X-ray diffraction patterns of HT100 at different temperatures. (□) periclase
(□) periclase MgO, and (+) hematite Fe2O3. MgO, (●) spinel MgFe2O4, (+) α-Fe2O3.
A.C. Heredia et al. / Journal of Magnetism and Magnetic Materials 342 (2013) 38–46 41

Fe3+ content is increased. This result is attributed to the decrease of


Octahedral
aluminum content, since the Al2O3 amorphous phase produces an
increase of the specific surface area. In agreement with this
Fe3+
Tetrahedral concept, Table 2 shows that the largest area was obtained in the
CHT0, while in the sample without iron the area was only 48 m2/g.

3.4. Thermal analyses (TGA and DSC)


Absorbance (a.u)

HT100

HT75 Thermal properties of the samples have been assessed by TGA


and DSC. These studies were carried out in air. To understand the
decomposition procedure of the Mg–Al–Fe LDH the TGA–DSC profiles
HT50
were analyzed in detail. Two weight loss stages can be observed on
the TGA curve, corresponding to the two endothermic peaks on DSC
HT25
profile, demonstrating that the decomposition proceeded in two
steps. In the TGA curves (Fig. 6), all samples show a weight loss below
∼100 1C, due to the water physically adsorbed, which corresponds to
a shoulder in the DSC profiles (Fig. 7), after this step starts the
200 250 300 350 400 450 500 550 600
decomposition proceeds. In the region between 100 and ∼300 1C the
TGA profiles show a weight loss which corresponds to the loss of
Wavelength (nm)
interlayer water, with a first maximum endothermic peak in the DSC
Fig. 4. UV–vis diffuse reflectance spectra of precursors with different iron profiles centered between 119 and 217 1C (Table 3). The dehydrox-
percentages. ylation of the brucite-like sheets and the loss of carbonates take place
between 300 and 410 1C. This weight loss is accompanied by a
maximum endothermic peak in the DSC profile centered between
345 and 396 1C approximately (Fig. 7). TGA curves showed a decrease
in the total weight loss when the iron content is increased due the
lower expulsion of CO2 and H2O from the LDH. This is consistent with
a decrease of LDH phase production observed by DRX and the
decrease of the specific surface area. Besides, the shift to lower
Absorbance (a.u)

CHT100
temperature of dehydroxylation of the brucite-like sheets with an
increase in the Fe content was observed, indicating a weakening of
CHT75 the brucite-like structure.
Final thermal decomposition products have shown to be metal
CHT50 oxides as well as mixed oxides and spinel-like species.

CHT25 3.5. XPS analysis

The XPS analysis has been done in order to obtain information


about the surface composition of the Mg–Al–Fe mixed oxides
derived from LDH precursors. Fig. 8 shows the Fe 2p spectra for
200 250 300 350 400 450 500 550 600 the oxides. All spectra show the main Fe 2p3/2 peak at BE of around
Wavelength (nm) 710.15 70.27 eV, accompanied by a satellite line visible at BE of

Fig. 5. UV–vis diffuse reflectance spectra of the oxides with different iron
percentages.
100

Table 2
Specific surface area of the samples. 90

% Fe (M+3) Specific surface area (m2/g)

Precursor Oxide 80
Weight (%)

0 48 179
25 20 150
50 19 119 70
75 18 96
100 17 48 HT100
HT75
60 HT50
summarized in Table 2. The specific surface area is inversely HT25
related with the iron content. HT0
Upon calcination at 550 1C for 9 h the surface areas of the
oxides are higher than their corresponding precursors. This 50
35 100 165 230 295 360 425 490 555 620 685
increase can be attributed to the formation of micro and meso-
Temperature (°C)
pores due to expulsion of CO2 and H2O from the LDH [28]. The
samples calcined showed a decrease in the areas when the iron Fig. 6. Precursor TGA curves with different iron percentages.
42 A.C. Heredia et al. / Journal of Magnetism and Magnetic Materials 342 (2013) 38–46

Table 3
Results of the TGA–DSC for different samples. Fe2p3/2
Sample temperature range Observed weight loss DSC maximum (1C) Fe3+ Satellite
(1C) (%) 7 0.01 1C
Fe2p1/2
Peak II Fe3+
HT0
Until 100 1.92 45.22
CHT100
100–300 15.6 261.9
300–550 25.19 395.03
Total weight loss 42.71
HT25

Intensity (a.u)
Until 100 1.57 64.41
100–300 13.84 206.62
CHT75
300–550 25.14 378.73
Total weight loss 40.55
HT50
Until 100 1.19 64.33
100–300 8.99 151.97
300–550 26.62 373.43 CHT50
Total weight loss 36.8
HT75
Until 100 1.87 102.59
100–300 5.56 179.34
300–550 22.94 363.37
Total weight loss 30.37
CHT25
HT100
Until 100 1.53 119.36
100–300 4.08 120
300–550 15.86 345
Total weight loss 21.47 700 705 710 715 720 725 730
BE (eV)
Fig. 8. XPS of the Fe 2p regions in the oxides.

-1
O1s OI
-3
OII
CHT100
Heat Flow (W/g)

-5

-7
Intensity (a.u)

CHT75
HT100 -9
HT75
HT50
HT25 -11
HT0
CHT50
-13
35 100 165 230 295 360 425 490 555 620 685
Temperature (°C)
Fig. 7. Precursor DSC curves with different iron percentages.

CHT25

around 718.87 70.40 eV, only indicating the presence of Fe3+


cations [29–31]. Whereas the contribution at 723.9 eV is assigned 526 528 530 532 534 536
to Fe 2p1/2. [32]. BE (eV)
The signal at 710.1570.27 eV can be decomposed in two
Fig. 9. XPS of the O1s regions in the oxides.
contributions; which indicate that the Fe3+ species exist in more
than one chemical state. Most probably, the two chemical states
may be related to different coordination environments of the Fe3+:
tetrahedral (A sites) at higher binding energy and octahedral OII, representing two different kinds of surface oxygen species.
(B sites) at lower binding energy [31,33]. The peak at 714.027 There is general agreement between the literature and the present
0.28 eV (peak II) could be related to the coordination environment results such that the OI with the lower oxygen bound to metal
of the Fe3+ cations in spinels phase (tetrahedral sites) [29]. cations of the structure, while OII with the higher BE al ca. 531.1 eV
Fig. 9 shows the O1s spectra for the surface oxides. The signal belongs most likely to surface oxygen, including mainly oxygen
to 529.5 70.2 eV can be deconvoluted in two contributions, OI and species of hydroxyl group [34].
A.C. Heredia et al. / Journal of Magnetism and Magnetic Materials 342 (2013) 38–46 43

3.6. Magnetism 15

CHT100
The room temperature hysteresis loops (only the first quadrant
is shown in order to get a better comprehension) for all samples 10
are displayed in Fig. 10. The magnetization curves of these mixed
oxides are well fitted by equation (1), supposing a paramagnetic
and/or superparamagnetic contribution and a ferromagnetic one. 5
From these fittings the saturation magnetizations Ms, saturation

M [emu/g]
fields Hs, and the B parameter, were calculated and they are shown
0
in Table 4. 550°C
The saturation magnetization of sample with 100% Fe content,
is lower than reported values (10.6–24.1 emu/g) for MgFe2O4 -5 700°C
produced by calcination of the intercalated LDH at 750–1100 1C
for 2 h [35]. This result would indicate that the precursor is not
fully transformed in MgFe2O4 after calcination at 550 1C for 9 h. -10
It is in total agreement with both XRD analysis and calorimetric
850°C
studies. As expected, the substitution of Fe+3 cations by Al+3 affect
these magnetic parameters. There are different behaviors with the -15
incorporation of aluminum, given that MgFe2O4 is almost an -15 -10 -5 0 5 10 15
inverse spinel, and MgAl2O4 is a normal spinel, then when we H [KOe]
replace Fe by Al, to obtain compounds of the type MgAlxFe2−x O4, Fig. 11. Hysteresis curves and their corresponding fittings for CHT100 at different
the Fe3+ ions, that occupy A sites, could be replaced by Al3+ ions or calcinations temperatures.
by the migration of Mg2+ ions from site B to site A [11]. This phase
was not detected by XRD in the samples; this could be due to their
small size. In any case, if a magnetic ion is substituted in any of the
sub-lattices of the spinel by a diamagnetic one, the magnetic
interactions between networks of the spinel will be changed.
These phenomena can be explained by ramdon canting model
[36,37] and for a chemically disordered system such as MgAlxFe2−x
O4, Modi et al. [38] explained that the canting is no uniform shown
a locally dependent upon non-magnetic neighboring ions

CHT25
Fit Ec (1) CHT75
3.5

3.0
CHT100
2.5
M [emu/g]

2.0
Transmission (a.u)

CHT50

1.5 CHT50

1.0
CHT25
CHT75
0.5

0.0
0 2500 5000 7500 10000 12500 15000
H [Oe]
Fig. 10. First quadrant of the room temperature hysteresis loops whit fitting curve
using equation (1).

CHT100
Table 4
Saturation magnetization, saturation field, B parameter and paramagnetic percen-
tage fraction (α) resulting from fitting Eq. (1) to the curves in Fig. 10.

% Fe Ms [emu/g] Hs [Oe] B¼μ/k %α

100 2.93 7900 7.7 26.6


75 3.88 7100 5.6 43.4
-12 -10 -8 -6 -4 -2 0 2 4 6 8 10 12
50 1.33 7700 4.1 57.9 Velocity (mm/s)
25 1.13 6800 4.1 65
Fig. 12. Mössbauer spectra of the mixes oxides samples.
44 A.C. Heredia et al. / Journal of Magnetism and Magnetic Materials 342 (2013) 38–46

distribution. Also the magnetic behavior of this compound and 9% for B site). The isomorphic Fe3+ substitution by diamag-
depends on procedures and synthesis conditions [11]. However, netic Al3+ ions would explain this result. This would produce
a correlation between the magnetic properties and aluminum different environments for Fe3+ ions depending on the nature and
content was not observed. These results would indicate that the number of nearest neighbors which is reflected in a broadening of
samples are complex, with different population ratios of ferrimag- the hyperfine magnetic field distributions of both sites. However, it
netic (MgFe2O4), weak-ferromagnetic (α-Fe2O3), paramagnetic and can not completely ruled out the existence of collective magnetic
superparamagnetic phases depending on the aluminum content. excitations.
Fig. 11 shows the curves of hysteresis and their corresponding Finally, the doublet present in all samples could be attributed to
fittings for CHT100 for different calcination temperatures. The several species:
magnetization of saturation increases with the calcination tem-
perature, corresponding to the observed increase of the ferrimag- – isolated paramagnetic Fe3+ions dissolved in MgO matrix.
netic phase and the crystallinity appreciated in X-ray diagrams. – superparamagnetic α-Fe2O3 (as very small crystals)
– susperparamagnétic MgFe2O4 with crystallite sizes below
3.7. Mössbauer 12 nm [43]
– MgFexAl2−xO4 with x o1, since, the presence of a high concen-
Fig. 12 shows the Mössbauer spectra of mixed oxides at room tration of diamagnetic cations (Al3+) in the oxide, causes the
temperature and their hyperfine parameters are listed in Table 5. formation of small magnetic “clusters” inside the particles (for
The spectra were fitted with one doublet and three sextets, this reason the MgFexAl2−xO4 phase was not detected by XRD),
except the CHT100 spectrum in which only two sextets were used. which show superparamagnetism. If the diamagnetic ions
The sextet of higher hyperfine magnetic field (H≅51.5 T) present in concentration is very high, these “clusters” can be completely
all samples corresponds to α-Fe2O3 with all hyperfine parameters isolated, and the spectrum shows a doublet even at low
consistent with this species, with a hyperfine field value slightly temperatures [44].
decreased in comparison with the “bulk” value [39]. This decrease
in H can be attributed to crystal size effects due to the process of
collective magnetic excitations [40], and/or to isomorphic replace- 3.8. Scanning electron micrograph (SEM)
ment of Fe3+ ions by Al3+[39].
The second sextet of the sample CHT100 corresponds to the sum Fig. 13(a)–(d) shows the micrographs of MgAlFe LDH with
of the two sites, tetrahedral (A sites) and octahedral (B sites) different iron content. The images were taken with secondary
occupied by Fe3+ions in MgFe2O4 spinel [41]. The presence of a electrons; all images show the presence of lamellar phase. When
large distribution of neighbors would produce a broadening of the the iron content is increased, smaller and exfoliated sheets are
signals. Therefore, is not possible to distinguish between both produced. The layered double hydroxide structure is distinguished
sites. in all samples except in the HT100.
The other two sextuplets present in the samples with iron Fig. 14(a) shows the image of the CHT75 sample. It was formed
content between 25 and 75% have hyperfine parameters assigned by secondary electrons indicating a rough surface, and the
to Fe3+ ions located in tetrahedral sites (sites A) and octahedral presence of small clusters generated by the calcination. When this
(B sites) of MgFe2O4 [42]. micrograph is compared with the corresponding LDH (Fig. 13(c)) it
The presence of A and B sites occupied by Fe3+ ions in the is possible to see that has a greater porosity and dispersion of the
spinel structure is consistent with that observed by XPS surface particles as well as the breakdown of the laminar structure. This
analysis. observation is consistent with the increased specific surface area,
Another aspect is that the magnetic hyperfine field values of when precursors are calcined.
the two sites are significantly diminished compared to a pure The image of the CHT100 sample (Fig. 14(b)) was formed with
spinel MgFe2O4 (between 14 and 15% for A site and between 11 secondary electron and backscattered. Note the presence of a

Table 5
Oxides Mössbauer parameters at 298 K.

Species Parameters HT25 HT50 HT75 HT100

α-Fe2O3 H (T) 51.487 0.08 51.46 70.06 51.5 70.3 51.3 7 0.1
δ (mm/s) 0.37 70.01 0.38 70.01 0.39 70.05 0.38 70.02
2ε (mm/s) −0.21 70.02 −0.217 0.02 −0.20 7 0.09 −0.20 7 0.03
% 297 6 24 72 57 1 317 2
Fe3+ ions in sites (A+B) of MgFe2O4 H (T) — — — 42.8 70.7
δ (mm/s) — — — 0.32 70.06
2ε (mm/s) — — — 0.17 0.1
% — — — 157 2
Fe3+ ions in A sites of MgFe2O4 H (T) 397 1 39.8 70.8 39.5 70.4 —
δ (mm/s) 0.27n 0.27n 0.27n —
2ε (mm/s) −0.017 0.06 0.017 0.04 0n —
% 257 11 217 4 14 72 —
Fe3+ ions in B sites of MgFe2O4 H (T) 447 1 45.2 70.2 45.3 70.2 —
δ (mm/s) 0.35n 0.35n 0.35n —
2ε (mm/s) 0.047 0.06 0.02 70.02 0n —
% 247 11 297 3 257 2 —
Fe3+ paramagnetic and/or superparamagnetic Species Δ 0.85 7 0.03 0.777 0.01 0.707 0.01 0.687 0.01
δ 0.32 7 0.02 0.337 0.01 0.357 0.01 0.31 70.01
% 227 4 267 2 567 2 54 7 2

H: hyperfine magnetic field in Tesla; δ: isomer shift (all the isomer shifts are referred to α-Fe at 298 K); 2ε: quadrupole shift; Δ: quadrupole splitting.
n
Parameters held fixed in fitting.
A.C. Heredia et al. / Journal of Magnetism and Magnetic Materials 342 (2013) 38–46 45

2 µm 2 µm

5 µm 2 µm

Fig. 13. Micrographs of MgAlFe LDH with different iron content. (a) HT0, (b) HT50, (c) HT75, (d) HT100.

1 µm 1 µm

Fig. 14. Micrographs of mixed oxides. (a) CHT75, (b) CHT100.

smooth surface roughness and small clusters. This can be attrib- the layer structure not was modified. On the other hand, when the
uted to sintering of the oxides, in absence of the amorphous samples were calcined at 550 1C, MgO, MgFe2O4 and α-Fe2O3 were
phase (Al2O3). This phase and the MgO are the cause of the iron detected by DRX; but by Mössbauer and magnetism the MgFexAl2−xO4
oxides dispersion [45]. Comparing the precursor area (17 m2/g) phase could be assigned.
with its oxide (48 m2/g) is not observed a significant increase after The formation of MgFe2O4 spinel structure is enhanced at high
calcination. temperature. By UV–vis-DRS were observed Fe3+ in tetrahedrally
and octahedrally coordination in brucite layered. In the calcined
samples was observed the presence of isolated Fe3+ in either
4. Conclusions periclase Mg(Fe, Al)O or MgFe2O4 spinel in mixed oxide and Fe3+
oxide clusters.
A series layered double hydroxides (LDH) as potential mixed A lower expulsion of CO2 and H2O of the interlayer when the
oxides precursors containing Mg2+, Al3+ and Fe3+ cations in the iron content is increased could be observed by a decrease in the
brucite-like layers were obtained. The precursors have been total weight loss from the LDH and in the specific surface areas in
prepared by coprecipitation method at 60 1C with hydrothermal the oxides.
treatment and a M2+/M3+ ¼3 constant molar ratio. The XPS analysis shows that the Fe3+ ions can be found in two
Powder X-ray diffraction shows the presence of the hydrotal- coordination environments (tetrahedral and octahedral) as mixed
cite phase. It was observed that the change the aluminum by iron, oxides, and as spinel-structure.
46 A.C. Heredia et al. / Journal of Magnetism and Magnetic Materials 342 (2013) 38–46

The combined magnetic and Mössbauer results analysis of the [16] J.M. Fernandez, M.A. Ulibarri, F.M. Labajos, V. Rives, Journal of Materials
complete series show that the samples are complex, with different Chemistry 8 (1998) 2507.
[17] Qinghong Xu, Yabo Wei, Yao Liu, Xuemei Ji, Lan Yang, Mingguang Gu, Solid
population ratios of ferrimagnetic (MgFe2O4), weak-ferromagnetic State Sciences 11 (2009) 472.
(α-Fe2O3), paramagnetic and superparamagnetic phases depending [18] X. Li, Q. Wei, W. Li, Advances in Materials Research (2012) 431347–353 (2012) 431.
on the aluminum content. Magnetic properties were significantly [19] A. Mantilla, F. Tzompantzi, J.L. Fernández, J.A.I. Díaz Góngora, R. Gómez,
Catalysis Today 150 (2010) 353.
improved when the calcination temperature was increased as a [20] Regina C.C. Costa, M.F.F Lelis, L.C.A. Oliveira, J.D. Fabris, J.D. Ardisson, R.R.V.
consequence of a better crystallization of spinel structure. A. Rios, C.N Silva, R.M. Lago, Journal of Hazardous materials 129 (2006) 171.
Due to the magnetic phase (spinel) in the materials calcined at [21] Qiujing Yang, Hyeok Choi, Souhail R. Al-AbedDionysios D. Dionysiou, Applied
Catalysis B: Environmental 88 (2009) 462.
temperatures above 550 1C, they could be used as catalysts in [22] A. Heredia, M. Oliva, C. Zandalazini, U. Agú, G. Eimer, S. Casuscelli, E. Herrero,
reactions of photocatalysis and dehydrogenation of ethylbenzene. C. Pérez, M. Crivello, I Industrial and Engineering Chemistry Research 50
(2011) 6695.
[23] P. Langevin, Annals of Chemistry and Physics 5 (1905) 70.
[24] K. Lagarec, D.G. Rancourt, Mossbauer Spectral Analysis Software, Department
Acknowledgements
of Physics, University of Otawa, 1998, Version 1.0.
[25] X. Ye, N. Ma, W.H.Y. Yue, C. Miao, Z. Xie, Z. Gao, Journal of Molecular Catalysis
This work was supported by the UTN-FRC of Argentina. A: Chemical 217 (2004) 103.
[26] T. Kawabata, N. Fujisaki, T. Shishido, K. Nomura, T. Sano, K. Takehira, Journal of
We thank UTN for doctoral fellowship for Angélica C. Heredia.
Molecular Catalysis A: Chemical 253 (2006) 279.
We thank geol. Julio D. Fernández (UTN-FRC, Córdoba, Argentina) [27] Y. Ohishi, T. Kawabata, T. Shishido, K. Takaki, Q. Zhang, Y. Wang, K. Nomura,
for help in recording specific surface area data and to Dra. K. Takehira, Applied Catalysis A: General 288 (2005) 220.
[28] J.L. Shumaker, C. Crofcheck, S.A. Tackett, E. Santillan-Jimenez, T. Morgan, Y. Ji,
Valentinuzi of LAMARX for SEM photografies.
M. Crocker, T.J. Toops, Applied Catalysis B: Environmental 82 (2008) 120.
[29] L.H. Zhang, X. Xiang, L. Zhang, F. Li, J. Zhu, D.G. Evans, X. Duan, Journal of
References Physics and Chemistry of Solids 69 (2008) 1098.
[30] F. Tihay, G. Pourroy, M. Richard-Plouet, A.C. Roger, A. Kiennemann, Applied
Catalysis A: General 206 (2001) 29.
[1] J. Roelofs, A. van Dillen, Y.K. de Jong, Catalysis Today 60 (2000) 297. [31] Xiaofeng Feng Li, Qiaozhen Liu, Junjie Yang, David G. Liu, Xue Duan. Evans,
[2] K. Simeonidis, S. Mourdikoudis, M. Moulla, I. Tsiaoussis, C. Martinez-Boubeta, Materials Research Bulletin 40 (2005) 1244.
M. Angelakeris, C. Dendrinou-Samara, O. Kalogirou, Journal of Magnetism and [32] Akihiro Miyakoshi, Akifumi Ueno, Masaru Ichikawa, Applied Catalysis A:
Magnetic Materials 316 (2007) e1–e4. General 219 (2001) 249.
[3] V. Rives, O. Prieto, A. Dubey, S. Kannan, Journal of Catalysis 220 (2003) 161. [33] H. Pines, The Chemistry of Catalytic Hydrocarbon Conversions, Academic
[4] F. Prinetto, D. Tichit, R. Teissier, B. Coq, Catalysis Today 55 (2000) 103. Press, New York, 1981, p. 208.
[5] S. Murcia-Mascarós, R. Navarro, L. Gómez-Sainero, U. Costantino, M. Nocchetti, [34] L. Zhang, J. Zhu, X. Jian, D.G. Evans, F. Li, Journal of Physics and Chemistry of
J.L.G. Fierro, Journal of Catalysis 198 (2001) 338. Solids 67 (2006) 1678.
[6] M. Crivello, C. Pérez, E. Herrero, G. Ghione, S. Casuscelli, E. Rodríguez- [35] W. Meng, F. Li, D.G. Evans, Xue Duan, Materials Chemistry and Physics 86
Castellón, Catalysis Today 107–108 (2005) 215. (2004) 1.
[7] F. Cavani, F. Trifiro, A. Vaccari, Catalysis Today 11 (1991) 173. [36] A. Rosencwaig, Canadian Journal of Physics 48 (1970) 2857.
[8] S. Kwak, W. Kriven, M. Wallig, Biomaterials 25 (2004) 5995. [37] J.M.D. Coey, Canadian Journal of Physics 65 (1989) 1210.
[9] H. Nakayama, N. Wada, M. Tsuhako, International Journal of Pharmaceutics [38] K.B. Modi, H.H. Joshi, R.G. Kulkarni, Journal of Materials Science 31 (1996)
269 (2004) 469. 1311.
[10] G. Carja, H. Chiriacb, N. Lupu, Journal of Magnetism and Magnetic Materials [39] R.E. Vandenberghe, E. De Grave, C. Landuydt, L.H. Bowen, Hyperfine Interac-
311 (2007) 26. tions 53 (1990) 175.
[11] M.D. Sundararajan, A Narayanasamy, T. Nagarajan, L. Haggstrom, C.S. Swamy, [40] S. Mørup, H. Topsøe, Applied Physics 11 (1976) 63.
K.V. Ramanujachary, Journal of Physics C: Solid State Physics 17 (1984) 2953. [41] H.H. Hamdeh, Z. Xia, R. Foehrweiser, B.J. McCormick, R.J. Willey, G. Busca,
[12] H. Zhang, K. Zou, H. Sun, X. Duan, Journal of Solid State Chemistry 178 (2005) Journal of Applied Physics 76 (2) (1994) 1135.
3485. [42] E. De Grave, A. Govaert, D. Chambaere, G. Robbrecht, Physica B: Condensed
[13] P.M. Forster, M.M. Tafoya, A.K. Cheetham, Journal of Physics and Chemistry of Matter 96 (1979) 103.
Solids 65 (2004) 11. [43] Q. Chen, A.J. Rondinone, B.C. Chakoumakos, Z.J. Zhang, Journal of Magnetism
[14] X. Qi, D. Wu, Journal of Magnetism and Magnetic Materials 320 (2008) 666. and Magnetic Materials 194 (1999) 1.
[15] X. Zhang, D. Wang, S. Zhang, Y. Ma, W. Yang, Y. Wang, S. Awaji, K. Watanabe, [44] S.C. Bhargava, Physical Review B: Condensed Matter 58 (6) (1998) 3240.
Journal of Magnetism and Magnetic Materials 312 (2010) 3023. [45] F. Cavani, F. Trifiro, A. Vaccari, Catalysis Today 11 (1991) 173.

You might also like