2015 JMagMagMater 387 37
2015 JMagMagMater 387 37
2015 JMagMagMater 387 37
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.
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.
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
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.
Notes: BMT: ball-milling time; I.S.: isomer shift (relative to α-Fe); Q.S.: quadrupole
splitting/shift; and B: hyperfine magnetic field.
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
Acknowledgments
References
[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.