Вы находитесь на странице: 1из 9

Journal of Magnetism and Magnetic Materials 387 (2015) 37–45

Contents lists available at ScienceDirect

Journal of Magnetism and Magnetic Materials


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

Role of mechanochemical milling in FeVO4 synthesis


Monica Sorescu a,n, Tianhong Xu a, Johanna D. Burnett b, Jennifer A. Aitken b
a
Duquesne University, Department of Physics, 309 B Fisher Hall, Pittsburgh, PA 15282-0321, USA
b
Duquesne University, Department of Chemistry and Biochemistry, Mellon Hall, Pittsburgh, PA 15282-1503, USA

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

Article history: Single phase, a FeVO4 triclinic crystalline structure was successfully synthesized by annealing the me-
Received 9 March 2012 chanochemically milled xV2O5  (1  x)α-Fe2O3 composites (x ¼0.5) at 550 °C for 1 h. X-ray powder dif-
Received in revised form fraction (XRD), simultaneous differential scanning calorimetry and thermogravimetric analysis (DSC–
21 October 2014
TGA), Mössbauer spectroscopy, scanning electron microscopy (SEM), and optical diffuse reflectance
Accepted 24 March 2015
spectroscopy were combined for a detailed study of the assisting role of the mechanochemical milling
Available online 26 March 2015
process. Mechanochemical milling homogeneously mixed the starting materials of α-Fe2O3 and V2O5 and
Keywords: substantially decreased their average grain sizes. The Mössbauer spectroscopy studies showed that the
FeVO4 spectrum of the mechanochemically milled composites consisted of three sextets and one doublet, in-
Mechanochemical milling
dicating the occurrence of V5 þ –Fe3 þ ion substitutions in the corresponding α-Fe2O3 and V2O5 lattices,
X-ray diffraction
respectively. The partially V5 þ -substituted α-Fe2O3 phase and Fe3 þ -substituted V2O5 could be the im-
Mössbauer spectroscopy
Simultaneous DSC–TGA portant intermediate phases in the production of FeVO4 single phase. The synthesized FeVO4 phase had a
slightly distorted nature with an unequal ratio in Fe3 þ population in three inequivalent sites. Simulta-
neous DSC–TGA studies indicated that the synthesized FeVO4 is thermally stable up to 600 °C. SEM
images of the formed FeVO4 confirmed the wide particles size distribution range composed of nano-
grains. Optical diffuse reflectance spectroscopy studies showed that the synthesized FeVO4 phase had
semiconductor properties, with the band gap energy of  2.44 eV.
& 2015 Elsevier B.V. All rights reserved.

1. Introduction atmospheric environment [10].


Recently, more research interests are focused on the water
Iron (III) vanadate, FeVO4, appears in four polymorphs: FeVO4-I splitting capability of photocatalytic materials due to the huge
[1], -II [2], -III [3], and -IV [4]. It has attracted attention because of global energy consumption. Solar hydrogen production, through
its interesting physical and chemical properties with wide appli- the use of a semiconductor photocatalyst employed to split water
cations [5–10]. The FeVO4 ceramics have a high dielectric constant, into oxygen and hydrogen, is a pollution-free energy source [11].
low dissipation factors and high quality factors. These attributes Fe–V–O based photocatalysts may be a suitable choice due to the
have lead to the wide spread use of these ceramics within many narrower band gaps [12]. It is well known that the photocatalytic
fields [5]. Crystals of the incongruently melting compound FeVO4, activity is closely related to the diameter size, morphology and
with shapes of both rods and platelets, have been grown by slowly surface area of the photocatalysts [13,14]. Thus, the synthesis of a
cooling melts of the composition 34 mol% Fe2O3, 66 mol% V2O5. nano-sized FeVO4 photocatalyst with high surface area is a subject
Mössbauer and susceptibility measurements showed a transition of considerable research interest, in regards to improving the
photocatalytic efficiency.
from paramagnetic to an ordered antiferromagnetic state at
FeVO4 materials have been synthesized via a surfactant-as-
TN ¼22 71 K [6]. FeVO4 was also used as a two-way Fenton-like
sisted sol–gel method [15,16], hydrothermal method, and a con-
catalyst in the degradation of Orange II contaminant, which is a
ventional solid state reaction. However, these methods either re-
primary contaminant in the textile industry [7]. FeVO4 also dis-
quire expensive chemical reagents or sintering processes at high
plays unusual electrochemical properties with respect to Li ac-
temperature. The high temperature synthesis leads to aggregation
ceptance/removal at low average voltages. This is an important
and sintering of the reaction products, which is expected to lower
factor in the performance of a lithium battery [8,9]. FeVO4 can be
the catalytic activities of FeVO4 due to the reduction of surface
used as a gas sensor material for detecting H2S traces in the area. High energy ball-milling is a well established method for
preparing extended solid solutions, composite and nanostructure
n
Corresponding author. Fax: þ 1 412 396 4829. systems using commercially obtained oxides as starting materials.
E-mail address: sorescu@duq.edu (M. Sorescu). This preparation method is promising for production scale use due

http://dx.doi.org/10.1016/j.jmmm.2015.03.074
0304-8853/& 2015 Elsevier B.V. All rights reserved.
38 M. Sorescu et al. / Journal of Magnetism and Magnetic Materials 387 (2015) 37–45

to its relatively low cost and simple operation [17]. Although top of double side carbon tape which was attached onto a standard
FeOOH and V2O5 are structurally active substances for stimulating aluminum stub and examined under high vacuum conditions, re-
mechanochemical reaction, it is not a sufficient condition, as no spectively. A 1.0 kV accelerating voltage and a 5.0 mm working
reaction occurs between the two active substances to form FeVO4 distance were employed.
[18]. According to the literature survey, few reports are available An optical diffuse reflectance spectrum was obtained using a
on the mechanochemical milling assisted synthesis of FeVO4 Varian Cary 5000 UV/vis/NIR spectrophotometer. The sample was
[19,20]. The final products in these studies contain  7% un-re- loaded into a Harrick Praying Mantis diffuse reflectance accessory
acted α-Fe2O3 phase and the assisting role of mechanochemical that uses elliptical mirrors. BaSO4 was used as a 100% reflectance
milling in FeVO4 synthesis was not discussed. standard. Scans were performed from 2500 to 200 nm at a rate of
In this work, we report the successful synthesis of single-phase 600 nm/min, wavelength data were converted to electron volts,
FeVO4 by a mechanochemical milling assisted method through the and the percent reflectance data were converted to absorbance
ball-milling of V2O5–α-Fe2O3 mixtures. The mixture consists of a units using the Kubelka–Munk equation [21].
stoichiometric ratio of starting materials, with the reaction being
carried out at room temperature, and then followed by a heat
treatment at 550 °C for 1 h. X-ray powder diffraction, simulta-
3. Results and discussion
neous DSC–TGA, Mössbauer spectroscopy, scanning electron mi-
croscopy, and UV–vis spectroscopy have been employed to in-
3.1. XRD
vestigate the phase evolution, thermal behavior, magnetic prop-
erties and morphology of the as-synthesized FeVO4 as well as the
V2O5-doped hematite, xV2O5  (1  x)α-Fe2O3 in which x¼0.5,
ball-milled oxides at various ball-milling times. The assisting role
was milled from 2 to 12 h (Fig. 1). The starting materials were pure
of mechanochemical milling in connection with the synthesis of
α-Fe2O3 and V2O5, no diffraction peaks from other phases were
FeVO4 is discussed.
detected after the physical mixing process (Fig. 1a). For the ball-
milled composites (Fig. 1b–e), the patterns show progressive peak
broadening with increased milling time. This peak broadening is
2. Experimental
associated with the decrease in the grain size of both the hematite
The sources of vanadium (V) and iron (III) oxides were com-
mercially purchased from Alfa Aesar: vanadium (V) pentoxide
(99.2% metals basis, average particle size is about 93.6 nm), and
hematite (α-Fe2O3, 99% metal basis, average particle size is about
49.2 nm). Powders of hematite and vanadium oxides were milled
at 1:1 M ratio in a hardened steel vial with 12 stainless-steel balls
(type 440; eight of 0.25 in diameter and four of 0.5 in diameter) in
the SPEX 8000 mixer mill for time periods ranging from 2 to 12 h.
The ball/powder mass ratio was 5:1. Prior to their introduction in
the ball milling device, the powders were manually ground in air
to obtain a homogeneous mixture.
The X-ray powder diffraction patterns of samples were ob-
tained using a PANalytical X'Pert Pro MPO powder diffractometer
with CuKα radiation (45 kV/40 mA, λ ¼1.54187 Å) with a nickel
filter on the diffracted side. A silicon-strip detector called X'cel-
lerator was used. The scanning range was 10–80° (2θ) with a step
size of 0.02°. The average grain size was determined by the
Scherrer method. The lattice parameters were extracted from
Rietveld structural refinement of the XRD patterns using GSAS
software to perform least-square fitting.
Simultaneous DSC–TGA experiments were performed using a
Netzsch Model STA 449F3 Jupiter instrument with a Silicon Car-
bide (SiC) furnace. Samples were contained in an alumina crucible
fitted with an alumina lid. A series of experiments were performed
using a 20 72 mg sample size. The atmosphere consisted of
flowing protective argon gas at a rate of 50 ml/min. DSC and TGA
curves were obtained by heating samples from room temperature
to 800 °C or 600 °C with a ramp rate of 10 °C/min. Both DSC and
TGA curves were corrected by subtraction of a baseline which was
run under identical conditions as DSC–TGA measurement with
residue of samples in the crucible. The Netzsch Proteus Thermal
Analysis software was used for DSC and TGA data analysis.
Room temperature transmission Mössbauer spectra were re-
corded using an MS-1200 constant acceleration spectrometer with
a 10 mCi 57Co source diffused in Rh matrix. Least-squares fittings
of the Mössbauer spectra were performed with the NORMOS
program.
Scanning electron microscopy was performed using a Hitachi
S-3400N scanning electron microscope. The powders of both 12 h Fig. 1. XRD patterns of mechanochemically milled xV2O5  (1  x)α-Fe2O3 (x ¼ 0.5)
ball-milled sample and the final FeVO4 product were adhered on composites at ball-milling time of: (a) 0 h; (b) 2 h; (c) 4 h; (d) 8 h; and (e) 12 h.
M. Sorescu et al. / Journal of Magnetism and Magnetic Materials 387 (2015) 37–45 39

Table 1
Rietveld refinement parameters from XRD patterns of xV2O5  (1  x)α-Fe2O3 (x¼ 0.5) composites at different ball-milling times and the as-synthesized FeVO4 sample.

BMT (h) Phase Lattice parameters Grain size ( 7 2 nm)

a ( 70.001 Å) b ( 7 0.001 Å) c ( 7 0.001 Å)

0 α-Fe2O3 5.0311 – 13.7386 49.2


V2O5 11.5133 3.5643 4.3730 93.6
2 α-Fe2O3 5.0369 – 13.7548 36.6
V2O5 11.5172 3.5666 4.3742 19.4
4 α-Fe2O3 5.0360 – 13.7615 33.6
V2O5 11.5178 3.5693 4.3866 18.5
8 α-Fe2O3 5.0388 – 13.7668 30.0
V2O5 11.5312 3.5678 4.3781 18.5
12 α-Fe2O3 5.0344 – 13.7726 30.1
V2O5 11.6414 3.5652 4.4124 14.9
12 h þ500 °C FeVO4 6.7114 8.0561 9.3545 84.0

and V2O5 samples. It can also be seen that the diffraction peak phases can still be seen and no other phase was observed. From
intensities of α-Fe2O3 and V2O5 decrease with the increase of ball- the variations in lattice parameters and average grain sizes of
milling time, indicating the possible ion substitutions between α-Fe2O3 and V2O5, it can be inferred that the mechanochemical
V5 þ and Fe3 þ in the corresponding hematite and V2O5 lattices. milling of the α-Fe2O3–V2O5 mixtures only reduces the average
Only α-Fe2O3 and V2O5 phases are present in the XRD patterns up grain size and introduces the ion substitutions between V5 þ and
to 12 h of milling time. This indicates that there is no solid state Fe3 þ in α-Fe2O3/V2O5 lattices, respectively. However, the me-
reaction between α-Fe2O3 and V2O5 under mechanochemical chanochemical milling process is not energetically enough to in-
milling up to 12 h. Table 1 presents the lattice parameters of the itiate the solid state reaction between α-Fe2O3 and V2O5, which is
Rietveld structural refinement. It can be seen that the average in good agreement with reported results [18,20].
grain size of α-Fe2O3 and V2O5 decreases with the ball-milling For the 12 h ball-milled sample after being annealed in air at
time. The original α-Fe2O3 has an average grain size of 49.2 nm 550 °C for 1 h, none of the peaks corresponding to the starting
and decreases slightly to 36.6 nm after 2 h of ball milling. It de- materials of V2O5 or α-Fe2O3 were observed, indicating the com-
creases continuously at increased ball milling times, from 33.6 nm pletion of the solid state reaction between V2O5 and α-Fe2O3.
for 4 h milling down to 30.1 nm for 12 h milling. The original V2O5 X-ray analysis, alongside the Rietveld refinement of the mechan-
has an average grain size of 93.6 nm, and it drops dramatically to
ochemically milled composites after annealing process is shown in
19.4 nm during the first 2 h of ball-milling time. It only decreases
Fig. 2, indicating that FeVO4 was successfully obtained after a very
slightly at long ball milling times, from 18.5 nm for 4 h milling
short time annealing process for the 12 h ball-milled composites. A
down to 14.9 nm for 12 h milling. The difference in the decrease in
model with a single phase of FeVO4 (triclinic P-1 space group) was
the average grain sizes of α-Fe2O3 and V2O5 may arise from the
employed to perform the refinement. The refined lattice para-
different micro-strains which were applied to the two different
meters of the synthesized FeVO4 have values of a¼ 6.7114 Å,
materials during milling process.
b¼8.0561 Å, c¼ 9.3545 Å, α ¼ 96.730°, β ¼106.672°, γ ¼ 101.565°,
The variations in lattice parameters of hematite (a and c) and
and unit cell V ¼466.837 Å3, respectively, which are similar to the
vanadium pentoxide (a, b, and c) of xV2O5  (1  x)α-Fe2O3 (x ¼0.5)
previously reported results [1,24]. The average grain size of the
are due to the high energy ball milling which causes the decrease
in grain size, the gradual V5 þ substitution of Fe3 þ in the hematite
lattice, and Fe3 þ substitution of V5 þ in V2O5 lattice. For V2O5
phase, lattice parameter a, b, and c increase from 11.5133 Å,
3.5643 Å, and 4.3730 Å in 0 h ball-milled sample to 11.6414,
3.5652, and 4.4124 Å in 12 h ball-milled sample, respectively. The
increase in the lattice parameters a, b, and c of V2O5 with the in-
crease in ball-milling time is consistent with the radius difference
between Fe3 þ and V5 þ , with Fe3 þ (0.63 Å) bigger than V5 þ
(  0.54 Å). Interestingly, lattice parameters a and c of α-Fe2O3
phase also increase with the increase in the ball-milling time, a
increases from 5.0311 Å for 0 h ball-milled sample to 5.0344 Å for
12 h milled sample, while c correspondingly increases from
13.7386 Å to 13.7726 Å. The change in the lattice parameters of
α-Fe2O3 is due to the high energy ball-milling effects, which de-
crease the grain size during the ball-milling process as well as V5 þ
substitution of Fe3 þ in α-Fe2O3 lattice. During the ball-milling
process, a microstrain concentrates in the lattice and increases the
lattice distortion and strain energy. The increase in the lattice
distortion, decrease in grain size and ion substitutions result in the
variation of lattice parameters of both α-Fe2O3 and V2O5. In fact, it
is well documented in the literature that lattice parameter changes Fig. 2. XRD patterns of the synthesized FeVO4 material. Individual data points are
under ball-milling process, either contraction or expansion, are shown as discrete open circles (o), the Rietveld fitted pattern is shown as the
continuous red line, pink pattern is the tick mark for the FeVO4 phase, and blue
expected when the grain sizes decrease as compared to the values pattern is the difference between experimental and fitted patterns. (For inter-
of bulk materials [22,23]. pretation of the references to color in this figure legend, the reader is referred to
After continuous ball-milling up to 12 h, α-Fe2O3 and V2O5 the web version of this article.)
40 M. Sorescu et al. / Journal of Magnetism and Magnetic Materials 387 (2015) 37–45

obtained FeVO4 is estimated from Scherrer formula to be


84.0 nm.
To obtain single phase of FeVO4, the conventional solid state
reactions between V2O5 and α-Fe2O3 are normally performed in
the temperature range of 800–1000 °C for different durations of
time [1,25], which is far above the eutectic point (  615 °C at
5 mol% Fe2O3). On the other hand, the conventional solid state
reaction requires a long heat treatment time (normally 410 h) to
ensure the V2O5 and α-Fe2O3 melt completely and homo-
geneously. In our case, after 12 h of mechanochemical milling
process, the temperature required to complete the solid state re-
action is not only much lower than that of the conventional solid
state reaction method, but is comparative to that of a wet chem-
istry method (  600 °C) [26]. The heat treatment time can be
dramatically dropped to 1 h to achieve the single phase FeVO4. The
lower synthesis temperature and shorter heat treatment time to
achieve single phase FeVO4 for this mechanochemical milling
method is due to the Fe3 þ and V5 þ substitution into V2O5/α-Fe2O3
lattices, which may lower the activation energy for the formation
of FeVO4 phase. On the other hand, homogeneous mixture of V2O5
and α-Fe2O3, with much smaller particles after long time ball-
milling process, may be the result to obtain the single phase of
FeVO4 requiring shorter heat treatment time as short as only 1 h.

3.2. Simultaneous DSC–TGA

The DSC–TGA curves of the original α-Fe2O3 and V2O5 samples


are shown in Fig. 3. One sharp endothermic peak can be been seen
for the original V2O5 sample (Fig. 3b). The endothermic peak is at
672 °C. The TGA curve of the V2O5 sample shows that there is no
distinguished weight loss corresponding to the endothermic peak.
This sharp endothermic peak is due to melting of V2O5, the peak
temperature corresponds to the V2O5 melting point, which is re-
ported as 670 °C [27]. The detailed explanation of DSC–TGA curves
of original α-Fe2O3 samples have been discussed in a previous
study [28], with the first exothermic peak and weight loss corre-
sponding to the decomposition of α-Fe2O3 under argon atmo- Fig. 3. DSC–TGA curves of the starting materials (a) α-Fe2O3 and (b) V2O5.
sphere. In order to avoid the melting of V2O5, the DSC–TGA mea-
surement for the ball-milled composites was performed only up to
600 °C. suggests that the manually grinding process does not dramatically
The small amount of weight loss up to temperature of about change the phase transition of α-Fe2O3 decomposition. The de-
150 °C for all the samples can be attributed to the physically ad- crease in the enthalpy for this sample is most likely due to the
sorbed water on the surface. The larger water weight losses from solid–solid interactions.
all of the ball-milled samples also indicate indirectly that the After ball milling for 2, 4, 8 and 12 h, the DSC curves of
mechanochemically milled composites possess smaller average xV2O5  (1  x)α-Fe2O3 (x ¼0.5) nanostructure system (Fig. 4b–e)
grain sizes because the smaller grains sizes provide a larger sur- change dramatically compared to the sample for 0 h of ball milling,
face area that can absorb more water. indicating the strong effect of ball-milling on the thermal behavior
Fig. 4 shows the DSC and TGA curves of xV2O5  (1  x)α-Fe2O3 of the xV2O5  (1  x)α-Fe2O3 system. Only one broad exothermic
composites (x¼ 0.5) after ball milling for 0, 2, 4, 8, 12 h, and the as- peak is observed for all of the ball-milled samples. The integration
synthesized FeVO4 sample. At 0 h of milling time, the DSC curve of this broad exothermic peak on the DSC curves gives much larger
showed one small exothermic peak (Fig. 4a) with the peak tem- enthalpy value (as indicated on the graphs) compared to the 0 h
perature at 344 °C. The characteristics of this DSC curve are not ball-milled sample. However, the enthalpy values of the exother-
just the simple sum of the DSC curves of the original α-Fe2O3 and mic peaks do not change dramatically with the increase of the ball
V2O5 samples. The integration of this exothermic peak gives an milling time. The peak temperatures of the broad exothermic peak
enthalpy value of 4.74 J/g, which is much smaller than that of the vary in the range of 460–474 °C. From the TGA curves of
original α-Fe2O3 sample ( 1.2 kJ/g). This means that the manually xV2O5  (1  x)α-Fe2O3 (x¼ 0.5) nanostructure system after different
ground mixture of xV2O5  (1  x)α-Fe2O3 (x ¼0.5) shows a different ball-milling times (Fig. 4h and l), it was found that the amount of
thermal behavior compared to samples with α-Fe2O3 or V2O5 weight loss for each sample varies with the ball-milling time. The
alone. The difference in DSC curves may characterize strong solid– weight loss percentage ranges from 0.84% to 0.93%, indicating that
solid interactions between α-Fe2O3 and V2O5, which affect the there is a solid–solid interaction between α-Fe2O3 and V2O5 after
crystallization of α-Fe2O3 or V2O5 fine grains. ball-milling. The increase in the enthalpy value after ball-milling is
Fig. 4g shows the TGA curve of xV2O5  (1  x)α-Fe2O3 composite possibly due to the decrease in average grain size of α-Fe2O3 and
(x ¼0.5) for 0 h of ball milling. A weight loss value of 0.66% can be V2O5. The crystallization of α-Fe2O3 and V2O5 finer grains releases
observed for temperature up to 600 °C, which is smaller than that more energy under heat treatment. The similarity in enthalpy
of the pure original α-Fe2O3 (0.93%) and is reasonable considering values and weight loss as a function of ball-milling time for the
that there is 50% of V2O5 in molar fraction in the mixtures. This ball-milled samples also confirms that. The dramatic changes in
M. Sorescu et al. / Journal of Magnetism and Magnetic Materials 387 (2015) 37–45 41

Fig. 4. DSC curves of mechanochemically milled xV2O5  (1  x)α-Fe2O3 (x¼ 0.5) composites at ball milling time of: (a) 0 h; (b) 2 h; (c) 4 h; (d) 8 h; (e) 12 h; and (f) as-
synthesized FeVO4 sample, and TGA curves of mechanochemically milled xV2O5  (1  x)α-Fe2O3 (x ¼0.5) composites at ball milling time of: (g) 0 h; (h) 2 h; (i) 4 h; (j) 8 h;
(k) 12 h; and (l) as-synthesized FeVO4 sample.

enthalpy and weight loss occurred only after 2 h of ball-milling time, the Mössbauer spectrum was fitted with 3 sextets and
time, which is consistent with the dramatic change in the average 1 doublet. The three sextets can be attributed to the α-Fe2O3 phase
grain sizes of α-Fe2O3 and V2O5 after 2 h of ball-milling time. The and the vanadium-substituted α-Fe2O3 phase with V5 þ substitu-
longer ball-milling time does slightly decrease the average grain tion of Fe3 þ in the α-Fe2O3 lattice, respectively, and the doublet
size of α-Fe2O3 and V2O5, but the induced change is not as large as can be assigned to Fe3 þ ions which substitute V5 þ in the V2O5
that of the 2 h ball-milled sample, which is also reflected in the lattice. When part of V5 þ substitutes Fe3 þ in α-Fe2O3 lattice, an-
similarities of the DSC and TGA behavior of the ball-milled tiferromagnetic ordering among Fe3 þ can still be induced by an
samples. exchange interaction, which gives sextets in Mössbauer spectrum
Fig. 4f and l presents the DSC and TGA curves of the as-obtained with different strengths in hyperfine magnetic fields. However,
FeVO4 sample. It can be seen that synthesized FeVO4 is thermally when a small amount of Fe3 þ substitution of V5 þ in the V2O5
stable in the studied temperature range. No obvious weight loss lattice occurs, no antiferromagnetic ordering among Fe3 þ in the
and crystallization behavior is observed, which is consistent with V2O5 lattice can be induced by an exchange interaction, therefore,
the reported results that FeVO4 is stable up to 700 °C [24,29]. only a quadrupole doublet appears. The isomer shift of Fe3 þ ex-
tracted from the doublet is higher compared to the value of Fe3 þ
3.3. Mössbauer spectroscopy extracted from the sextet pattern of Fe2O3; this may be attributed
to the change of neighbors in these two different cases. The re-
The room temperature transmission Mössbauer spectra of the lative abundance of a magnetic phase with highest hyperfine field
xV2O5  (1  x)α-Fe2O3 composites (x¼ 0.5) after ball milling for 0, (48.23 T) is 67.97%, and the relative population of a magnetic
2, 4, 8 and 12 h, respectively, are represented in Fig. 5a–e. The phase with lower hyperfine fields of 47.79 and 45.17 T is 10.63%
hyperfine parameters corresponding to these spectra are given in and 17.27%, respectively.
Table 2. At 0 h of milling time, the spectrum was fitted with After 12 h of milling time, the spectrum can still be fitted with
1 sextet (Fig. 5a), corresponding to α-Fe2O3. After 2 h of milling 3 sextets and 1 doublet. Similar to the 4 h ball-milled sample, the
42 M. Sorescu et al. / Journal of Magnetism and Magnetic Materials 387 (2015) 37–45

Table 2
Mössbauer parameters for xV2O5  (1  x)α-Fe2O3 (x ¼0.5) composites at different
ball-milling times and the as-synthesized FeVO4 sample.

BMT (h) I.S. (mm/s) Q.S. (mm/s) B (T) Abundance (%)

0 0.568  0.290 47.68 100


2 0.517  0.192 48.23 67.97
0.565  0.290 47.79 10.63
0.571  0.336 45.17 17.27
0.531 0.574 – 4.13
4 0.550  0.267 48.27 62.40
0.580  0.349 47.15 23.14
0.529  0.258 45.27 7.66
0.559 0.618 – 6.80
8 0.591  0.348 48.33 39.54
0.548  0.248 47.89 36.51
0.557  0.228 45.13 14.59
0.606 0.511 – 9.36
12 0.581  0.318 48.23 50.08
0.548  0.274 47.69 21.80
0.557  0.307 44.49 12.77
0.615 0.549 – 15.36
FeVO4 sample 0.758 0.559 – 37.45
0.785 0.136 – 32.12
0.145 0.300 – 30.43
Error 7 0.01 70.02 71.0 7 0.1

Notes: BMT: ball-milling time; I.S.: isomer shift (relative to α-Fe); Q.S.: quadrupole
splitting/shift; and B: hyperfine magnetic field.

proportions [1], with two in octahedral site and one in trigonal


bipyramidal environment. The three independent iron atoms
create a doubly bent chain of six edge-sharing polyhedral, the
chains are joined by VO34  tetrahedra. The observed isomer shift
values of the three doublets for the as-synthesized FeVO4 sample
correspond to high spin Fe3 þ ions (Table 2). The attribution of the
Fe3 þ to three doublets originates from the significant differences
between the isomer shift values. The doublet, with the isomer
shift and quadrupole splitting values of 0.145 mm/s and
0.300 mm/s, respectively, can be assigned to a five-folded co-
ordinated Fe3 þ (trigonal bipyramidal sites); each of the other two
doublets can be assigned to octahedrally coordinated Fe3 þ . The
populations of these three different sites are 37.45%, 32.10%, and
30.43%, respectively, which is not in an exact 1:1:1 ratio, indicating
the slight distortion of the iron sites. The similar distortion, with
unequal Fe3 þ populations occupying the three inequivalent sites,
Fig. 5. Mössbauer spectra of mechanochemically milled xV2O5  (1  x)α-Fe2O3 was also found in the Mössbauer spectrum of amorphous FeVO4
(x¼ 0.5) composites at ball milling time of: (a) 0 h; (b) 2 h; (c) 4 h; (d) 8 h; (e) 12 h; phase [30]. On the other hand, the difference in the isomer shift,
and (f) as-synthesized FeVO4 sample. quadrupole splitting values, and relative populations of Fe3 þ sites
depends significantly on the preparation methods which produce
three sextets can be attributed to the α-Fe2O3 phase and the va- the FeVO4, with considerable difference in the degree of crystal-
nadium-substituted α-Fe2O3 phase with V5 þ substitution of Fe3 þ linity and structural imperfections [2,31].
in the α-Fe2O3 lattice, respectively, and the doublet can be as-
signed to Fe3 þ ions which substitute V5 þ in the V2O5 lattice. In 3.4. Scanning electron microscopy
addition, the population of this quadrupole splitting doublet
component increases from 4.13% for 2 h ball-milled sample to SEM was employed to study the morphology and particle size
15.36% for 12 h ball-milled sample, indicating that long ball-mil- of the as-obtained products. Fig. 6 shows secondary electron SEM
ling time introduces not only smaller grain sizes but also more ion images of the xV2O5  (1  x)α-Fe2O3 (x ¼0.5) composite after 12 h
substitutions between Fe3 þ and V5 þ , which are believed to be the of ball-milling time and the as-synthesized FeVO4 sample. For the
precursors to produce the single phase of FeVO4 material, as the 12 h ball-milled composites, a mixture of large crystal agglomer-
solid state reaction occurs at a lower temperature. ates in the size range 1–30 mm coated with much finer grains with
Three doublets are observed in the Mössbauer spectrum of as- nanometer dimensions is observed. Several agglomerates with 10–
synthesized FeVO4 (Fig. 5f). The absence of a sextet corresponding 20 μm in size can be seen (Fig. 6a). However, most particles are
to α-Fe2O3 indicates the complete consumption of α-Fe2O3 com- still less than 10 μm at the magnification scale of  250. Fig. 6b
ponents, which is in good agreement with the XRD results, as no shows more details of the as-obtained composite at a magnifica-
diffraction peaks corresponding to the starting materials are ob- tion of  3.0K. It can be seen clearly that the large agglomerates
served. The existence of three doublets for the as-synthesized consist of fine grains with redundant sizes in nanometer range.
FeVO4 sample is in good agreement with the three inequivalent Grains in the nanometer size range are distributed randomly on
iron positions in the triclinic crystalline structure of FeVO4. Fe3 þ the edge and top of the big agglomerates, indicating ball milling is
ions occupy the three inequivalent sites in slightly different a useful method to prepare composites at the nanometer scale.
M. Sorescu et al. / Journal of Magnetism and Magnetic Materials 387 (2015) 37–45 43

Fig. 6. SEM images of mechanochemically milled xV2O5  (1  x)α-Fe2O3 (x¼ 0.5) composites at ball-milling time of 12 h (a, b) and as-synthesized FeVO4 sample (c, d).

Fig. 6c and d shows the SEM images of the as-synthesized FeVO4 occupying the Pmnm space group, and its unit cell contains two
material. Similar to the ball-milled composites, the as-synthesized formula units. The estimated band gap energy of the original V2O5
triclinic FeVO4 has a wide-range distribution in particle size dia- has a value of  2.33 eV (Fig. 7b), which is in very good agreement
meters, consisting of micrometer-sized agglomerates and nan- with the reported value of  2.3 eV [37–39]. For the sample
ometer-sized grains. The average grain sizes of FeVO4 are much xV2O5  (1  x)α-Fe2O3 (x¼ 0.5) with 0 h of milling time, in other
larger than those of the ball-milled xV2O5  (1 x)α-Fe2O3 (x ¼0.5) words, just after physically mixing of V2O5 and α-Fe2O3, the band
composites, due to the crystallization of finer grains during the gap energy of this oxide mixture changes slightly, with an average
thermal annealing process. However, the as-synthesized FeVO4 value of  2.23 eV (Fig. 7c). Due to the extremely close band gap
through the mechanochemical milling assisted method can energies of V2O5 and α-Fe2O3, the individual band gap energies for
maintain an average grain size in a nanometer range, which is the mixed oxides can hardly be determined. The optical band gap
useful in the catalytic application of FeVO4. Agglomerates were measurement only shows an overall band gap value, which lies in
also found in the ball-milling prepared photocatalyst the range of band gaps of α-Fe2O3 and V2O5.
p-CaFe2O4/n-Ag3VO4 [32], and ball-milling synthesized LaFeO3 After different hours of milling time, band gap energies of the
perovskite materials [33], indicating it is a general phenomenon to ball-milled samples do not change dramatically (Fig. 7d–g). Similar
produce some agglomerates with micrometer size using the high to that of 0 h ball-milled samples, the band gap energies of all of
energy ball milling method. the ball-milled samples are 2.23 eV, suggesting that the me-
chanochemical milling process does not alter the band gap en-
3.5. Optical diffuse reflectance spectroscopy ergies of the ball-milled oxide mixtures, though it does change its
microstructures, average grain sizes, as well as the magnetic
The optical diffuse reflectance spectrum for hematite, V2O5, and properties. The similar optical properties of the ball-milled
xV2O5  (1  x)α-Fe2O3 (x¼ 0.5) with different ball milling times, xV2O5  (1  x)α-Fe2O3 (x ¼0.5) are due to the very close band gap
and the as-synthesized FeVO4 are shown in Fig. 7. For hematite, energies between V2O5 and α-Fe2O3.
the valence and conduction bands arise from crystal field splitting For the as-synthesized FeVO4 sample, the UV–vis spectrum
of the Fe 3d levels due to the octahedral coordination of oxygen changes dramatically (Fig. 7h). The band gap energy is  2.44 eV,
around Fe [34]. As shown in Fig. 7a, diffuse reflectance spectra of which is higher than that of both original α-Fe2O3 and V2O5,
the original hematite exhibited a band gap energy value of suggesting the formation of a new phase after the annealing
2.19 eV, which is consistent with reported value 2.14–2.2 eV for process. This is in good agreement with XRD and Mössbauer re-
bulk hematite [35,36]. V2O5 forms an orthorhombic crystal sults which confirm that single phase FeVO4 formed. The plot of
44 M. Sorescu et al. / Journal of Magnetism and Magnetic Materials 387 (2015) 37–45

sextets and one doublet, indicating the occurrence of V5 þ -Fe3 þ


ion substitutions in the corresponding α-Fe2O3 and V2O5 lattices,
respectively. The partially V5 þ -substituted α-Fe2O3 phase and
Fe3 þ -substituted V2O5 phase could be important intermediate
phases, which required a heat treatment in air at a lower tem-
perature with a shorter time duration than that of the conven-
tional solid state reaction to produce FeVO4 single phase. Further
analysis of Mössbauer spectrum revealed that the synthesized
FeVO4 material had a distorted nature with an unequal ratio in
Fe3 þ population in the three inequivalent sites. Simultaneous
DSC–TGA studies indicated that the synthesized FeVO4 is ther-
mally stable up to 600 °C. SEM images of the as-synthesized FeVO4
confirmed the wide range distribution of particles sizes, from
nanometer-sized particles to micrometer-sized agglomerates, but
the grains remained at nanometer scale. Optical diffuse reflectance
spectroscopy studies showed that the synthesized FeVO4 phase
had semiconductor properties, with the band gap energy of
 2.44 eV.

Acknowledgments

This work was supported by the National Science Foundation,


USA under Grant number DMR-0854794 and the National Science
Foundation (NSF) Major Research Instrumentation (MRI) program
under Grant number CHE-0923183.

References

[1] B. Robertson, E. Kostiner, Crystal structure and Mössbauer effect investigation


of FeVO4, J. Solid State Chem. 4 (1972) 29–37.
[2] Y. Oka, T. Yao, N. Yamamoto, Y. Ueda, S. Kawasaki, M. Azuma, M. Takano, Hy-
drothermal synthesis, crystal structure, and magnetic properties of FeVO4-II, J.
Solid State Chem. 123 (1996) 54–59.
[3] J. Muller, J.C. Jaubert, Synthese sous haute pression d'oxygene d'une forme
dense ordonne´e de FeVO4 et mise en evidence d'une varie´te´allotropique de
structure CrVO4, J. Solid State Chem. 14 (1975) 8–13.
[4] F. Laves, A.P. Young, C.M. Schwartz, On the high-pressure form of FeVO4, Acta
Crystallogr. 17 (1964) 1476–1477.
[5] S. Gupta, Y.P. Yadava, R.A. Singh, Electrical transport properties of iron vana-
Fig. 7. The optical diffuse reflectance spectrum converted to absorption for sam- date, J. Mater. Sci. Lett. 5 (1986) 736–738.
ples: (a) original hematite, (b) original V2O5, and mechanochemically milled [6] L.M. Levison, B.M. Wanklyn, Crystal growth and magnetic behavior of FeVO4, J.
xV2O5  (1  x)α-Fe2O3 (x¼ 0.5) composites at ball milling time of: (c) 0 h; (d) 2 h; Solid State Chem. 3 (1971) 131–133.
(e) 4 h; (f) 8 h; (g) 12 h, and (h) as-synthesized FeVO4 sample, respectively. On the y [7] J.H. Deng, J.Y. Jiang, Y.Y. Zhang, X.P. Lin, C.M. Du, Y. Xiong, FeVO4 as a highly
axis, α is the absorbance and s is the scattering coefficient constant. active heterogeneous Fenton-like catalyst towards the degradation of Orange
II, Appl. Catal. B 84 (2008) 468–473.
[8] P. Poizot, E. Baudrin, S. Laruelle, L. Dupont, M. Touboul, J.M. Tarascon, Low
absorption vs energy is similar to that of a FeVO4 thin film elec- temperature synthesis and electrochemical performance of crystallized
trode synthesized using a low temperature and aqueous pre- FeVO4  1.1H2O, Solid State Ion. 138 (2000) 31–40.
cipitation reaction, and the band gap of FeVO4 is believed to cor- [9] M. Hayashibara, M. Eguchi, T. Miura, T. Kishi, Lithiation characteristics of FeVO4,
Solid State Ion. 98 (1997) 119–125.
respond to an indirect optical transition [40]. The as-synthesized [10] G. Mangamma, E. Prabhu, T. Gnanasekaran, Investigations on FeVO4 as a gas
FeVO4 sample has a band gap energy of  2.44 eV, which indicates sensor material, Bull. Electrochem. 12 (1996) 696–699.
that it is a semiconductor. It has been previously confirmed that [11] A. Fujishima, K. Honda, Electrochemical photolysis of water at a semi-
conductor electrode, Nature 238 (1972) 37–38.
the FeVO4 is an n-type semiconductor [41], and the mechanism of
[12] T. Arai, Y. Konishi, Y. Iwasaki, H. Sugihara, K. Sayama, High-throughput
electronic conduction in FeVO4 is due to the thermally activated screening using porous photoelectrode for the development of visible-light-
hopping of a charge carrier on equivalent iron lattice sites [42]. responsive semiconductors, J. Comb. Chem. 9 (2007) 574–581.
[13] D. Chen, Design, Synthesis and properties of highly functional nanostructured
photocatalysts, Recent Pat. Nanotechnol. 2 (2008) 183–189.
[14] D.F. Wang, Z.G. Zou, J.H. Ye, A novel Zn-doped Lu2O3/Ga2O3 composite pho-
4. Conclusions tocatalyst for stoichiometric water splitting under UV light irradiation, Chem.
Phys. Lett. 384 (2004) 139–143.
[15] A.S. Vuk, B. Orel, G. Dražič, F. Decker, P. Colomban, I.R. UV-Visible, spectro-
After mechanochemical milling of α-Fe2O3 and V2O5 mixtures electrochemical studies of FeVO4 sol-gel films for electrochromic applications,
with a 1:1 molar ratio for 12 h, the single phase FeVO4 triclinic J. Sol-Gel Sci. Technol. 23 (2002) 165–181.
crystalline structure was successfully synthesized by annealing the [16] H.Y. Zhou, T.L. Gu, D.G. Yang, Z.Y. Jiang, J.M. Zeng, Characterization and pho-
tocatalytic activity of FeVO4 photocatalysts synthesized via a surfactant-as-
mechanochemically milled composite at 550 °C for 1 h. The me-
sisted sol–gel method, Adv. Mater. Res. 197–198 (2011) 926–930.
chanochemical milling process played an important role in the [17] T. Tojo, Q.W. Zhang, F. Saito, Mechanochemical synthesis of indium complex
synthesis of FeVO4. It homogeneously mixed the starting materials oxides (InAO4; A ¼P, V, Nb, Ta, Sb) and their solid solution, J. Mater. Sci. 43
of α-Fe2O3 and V2O5 and decreased their average grain sizes. The (2008) 2962–2966.
[18] Q.W. Zhang, T. Tojo, W. Tongamp, F. Saito, Correlation between mechan-
Mössbauer spectroscopy studies indicated that the spectrum of ochemical reactivity formation ABO4-type complex oxides and the structures
the mechanochemically milled composites consisted of three of product materials, J. Mater. Sci. 195 (2009) 40–43.
M. Sorescu et al. / Journal of Magnetism and Magnetic Materials 387 (2015) 37–45 45

[19] D. Klissurski, D. Radev, R. Iordanova, M. Milanova, Synthesis of iron (III) va- [32] S.F. Chen, W. Zhao, W. Liu, H.Y. Zhang, X.L. Yu, Y.H. Chen, Preparation, char-
nadate from mechanically activated precursors, Chim. Inorg. 56 (2003) 27–30. acterization and activity evaluation of p–n junction photocatalyst
[20] D. Klissurski, R. Iordanova, D. Radev, S.T. Kassabov, M. Milanova, K. Chakarova, p-CaFe2O4/n-Ag3VO4 under visible light irradiation, J. Hazard. Mater. 172
Mechanochemically assisted synthesis of FeVO4 catalysts, J. Mater. Sci. 39 (2009) 1415–1423.
(2004) 5375–5377. [33] M. Sorescu, T.H. Xu, J.D. Burnett, J.A. Aitken, Investigation of LaFeO3 perovskite
[21] P. Kubelka, F. Munk, Ein Beitrag zur Optik der Farbanstriche, Z. Tech. Phys. 12 growth mechanism through mechanical ball milling of lanthanum and iron
(1931) 593–601 (English translated by Westin, S.). oxides, J. Mater. Sci. 46 (2011) 6709–6717.
[22] M. Sorescu, L. Diamandescu, Mechanochemical and magnetomechanical [34] F.J. Morin, Electrical properties of α-Fe2O3, Phys. Rev. 93 (1954) 1195–1199.
synthesis of hematite nanoparticles, Hyperfine Interact. 196 (2010) 349–358. [35] L.A. Marusak, R. Messier, W.B. White, Optical absorption spectrum of hematite,
[23] H. Dutta, S.K. Pradhan, Microstructure characterization of high energy ball- α-Fe2O3 near IR to UV, J. Phys. Chem. Solids 41 (1980) 981–984.
milled nanocrystalline V2O5 by Rietveld analysis, Mater. Chem. Phys. 77 (2002) [36] M.P. Dare-Edwards, J.B. Goodenough, A. Hamnett, P.R. Trevellick, Electro-
868–877. chemistry and photoelectrochemistry of iron (III) oxide, J. Chem. Soc. Faraday
[24] P.I. Cowin, R. Lan, L. Zhang, C.T.G. Petit, A. Kraft, S.W. Tao, Studies on con- Trans. 79 (1983) 2027–2041.
ductivity and redox stability of iron orthovanadate, Mater. Chem. Phys. 126 [37] D.C. Conlon, W.P. Doyle, Absorption spectra of vanadium, niobium, and tan-
(2011) 614–618. talum pentoxides, J. Chem. Phys. 35 (1961) 752–753.
[25] Z.Z. He, J.I. Yamaura, Y. Ueada, Flux growth and magnetic properties of FeVO4 [38] N. Kenny, C.R. Kannewurf, D.H. Whitmore, Optical absorption coefficients of
single crystal, J. Solid State Chem. 181 (2008) 2346–2349. vanadium pentoxide single crystals, J. Phys. Chem. Solids 27 (1966)
[26] V.D. Nithya, R.K. Selvan, Synthesis, electrical and dielectric properties of FeVO4 1237–1246.
nanoparticles, Physica B 406 (2011) 24–29. [39] V.G. Mokerov, V.L. Makarov, V.B. Tulvinskii, A.R. Begishev, Optical properties of
[27] F.D. Rossini, D.D. Wagman, Selected Values of Chemical Thermodynamic vanadium pentoxide in the region of photon energies from 2 eV to 14 eV, Opt.
Properties, National Bureau of Standards (US), Washington, D.C., 1952. Spectrosc. 40 (1976) 58–61.
[28] M. Sorescu, T.H. Xu, L. Diamandescu, Synthesis and characterization of [40] C.D. Morton, I.J. Slipper, M.J.K. Thomas, B.D. Alexander, Synthesis and char-
xTiO2  (1  x)α-Fe2O3 magnetic ceramic nanostructure system, Mater. Charact. acterization of Fe–V–O thin film photonodes, J. Photochem. Photobiol. A 216
61 (2010) 1103–1118. (2010) 209–214.
[29] T. Yang T, D.G. Xia, Z.L. Wang, Y. Chen, A novel anode material of Fe2VO4 for [41] T. Groń, J. Krok-Kowalowski, M. Kurzawa, J. Walczak, Electrical conductivity in
high power Lithium ion battery, Mater. Lett. 63 (2009) 5–7. the antiferromagnetic compounds FeVO4, FeVMoO7 and Fe4V2Mo3O20, J.
[30] H. Mehner, W. Meisel, A. Brückner, A. York, Investigation of alkali doped Magn. Magn. Mater. 101 (1991) 148–150.
Fe2O3–V2O5 catalysts by transmission and conversion electron Mössbauer [42] N.S. Rao NS, O.G. Palanna, Electrical and magnetic studies of iron (III) vana-
spectroscopy, Hyperfine Interact. 111 (1998) 51–56. date, Bull. Mater. Sci. 18 (1995) 229–236.
[31] S. Varma, B.N. Wani, A. Sathyamoorthy, N.M. Gupta, On the role of lattice
distortion in the catalytic properties of substituted orthovanadates
La1  xFexVO4, J. Phys. Chem. Solids 65 (2004) 1291–1296.

Вам также может понравиться