Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Available online at www.sciencedirect.

com

Ultrasonics Sonochemistry 15 (2008) 257–264


www.elsevier.com/locate/ultsonch

Sonochemical synthesis of amorphous nanoscopic iron(III) oxide


from Fe(acac)3
a,*
Jiri Pinkas , Vendula Reichlova a, Radek Zboril b, Zdenek Moravec a,
Petr Bezdicka c, Jirina Matejkova d
a
Department of Inorganic Chemistry, Masaryk University, Kotlarska 2, CZ-61137 Brno, Czech Republic
b
Department of Physical Chemistry, Palacky University, Svobody 26, CZ-77146 Olomouc, Czech Republic
c
Institute of Inorganic Chemistry ASCR, CZ-25068 Rez, Czech Republic
d
Institute of Scientific Instruments ASCR, Kralovopolska 147, CZ-61264 Brno, Czech Republic

Received 14 October 2006; accepted 23 March 2007


Available online 3 April 2007

Abstract

Amorphous nanoscopic iron(III) oxide with interesting magnetic properties was prepared by sonolysis of Fe(acac)3 under Ar in
tetraglyme with a small amount of added water. The organics content and the surface area of the Fe2O3 nanoparticles can be controlled
with an amount of water in the reaction mixture and it increases from 48 m2 g 1 for dry solvent up to 260 m2 g 1 when wet Ar is
employed. For further monitoring of the particle size and morphology and for the study of the surface, magnetic and thermal properties,
the sample with 2 vol.% of H2O was chosen. SEM showed nanoscopic composite particles of a uniform size distribution and nearly
spherical shapes with an estimated diameter of 20 nm. Such composites are built from amorphous iron(III) oxide nanoparticles
(3 nm) embedded in an acetate matrix as proved by TEM and IR spectroscopy. Temperature-dependent Mössbauer spectra demonstrate
a very narrow magnetic transition with an unusually low transition temperature around 25 K reflecting the system of magnetically non-
interacting ultrasmall particles with a narrow size distribution. The in-field (5 T) Mössbauer spectrum recorded at 5 K shows a minimum
change compared to the zero-field spectrum indicating an absence of the long-range magnetic ordering. The composite particles are ther-
mally stable up to 150 C, which is confirmed by DSC, TG, and by the constant surface area. At higher temperatures, acetate groups are
removed from the particle surface, which is documented by the increased surface area and disappearance of their IR bands.
 2007 Elsevier B.V. All rights reserved.

Keywords: Iron; Amorphous oxides; Ultrasound; Sonochemistry; Solvents; Nanoparticles

1. Introduction able method for the preparation of amorphous phases, new


mixed-element compounds, and nanoscopic structures. The
The ultrasound-activated reactions were in recent times study of ultrasound-activated routes to nanomaterials is a
employed in many areas of chemistry, one of them being rapidly growing research area [3,4]. Judicious choice of pre-
the field of inorganic materials synthesis [1,2]. The sono- cursors, solvents, additives and other variations in the reac-
chemical process relies on the cavitation effects produced tion parameters and conditions allow control of the
by collapsing bubbles in liquids. Factors such as rapid chemical nature of the product, its composition, external
cooling rates, atomic level mixing and the nonequilibrium form and morphology, and crystallinity.
nature of the reactions render sonochemistry the most suit- Nanoparticles of various polymorphic forms of iron oxi-
des [5] found their applications in such diverse areas as
magnetic liquids, photocatalysis, catalytic destruction of
*
Corresponding author. Tel.: +420 54946493; fax: +420 549492443. chlorinated organics, diagnostic imaging, and drug deliv-
E-mail address: jpinkas@chemi.muni.cz (J. Pinkas). ery. A multitude of preparative methods for iron oxide

1350-4177/$ - see front matter  2007 Elsevier B.V. All rights reserved.
doi:10.1016/j.ultsonch.2007.03.009
258 J. Pinkas et al. / Ultrasonics Sonochemistry 15 (2008) 257–264

nanoparticles were developed, such as MW irradiation [6], centers and solvate large cations. Therefore, it should sta-
forced hydrolysis and precipitation [7], laser pyrolysis [8], bilize colloidal particles similarly to PEG in the MW irra-
microemulsion [9], sol–gel, electrochemical [10], solid-state diation assisted synthesis of Fe2O3 from FeCl3 and urea
thermal decomposition [11], oxidation of metallic iron [6]. Use of tetraglyme in sonochemical systems has not
nanoparticles [12,13], solvothermal decomposition [14], been reported so far.
combustion method [15], and others. Considering the favorable properties of Fe(acac)3 as a
Thermally induced processes provide crystalline nano- precursor and tetraglyme as a solvent, we set out to inves-
particles in many cases. In contrast, ultrasound-driven tigate this system. Herein we report on the sonochemical
reactions yield amorphous materials that are produced in synthesis of amorphous Fe2O3 nanoparticles with an
collapsing cavitation bubbles as enormous cooling rates organic matrix and uniform size distribution which exhibit
(1010 K s 1) prevent their crystallization during quenching. unusual magnetism. Decomposition under power ultra-
Sonochemical synthetic routes leading to iron oxides rely sound activation of nontoxic precursor Fe(acac)3 in high-
most frequently on Fe(CO)5 [16,17] as a precursor, how- boiling coordinating solvent tetraglyme provided nano-
ever, FeCl3 [18], Fe(NO3)3 [19], Fe(OAc)2 [20], and powders with large surface areas of up to 260 m2 g 1.
Fe(OEt)3 [21] were also employed. Nanoscopic amorphous The effect of water added to the reaction mixture was
Fe2O3 occupies a special position among various forms of examined with the aim to facilitate the ligand removal as
iron oxides because of several factors [11]. A large surface Hacac. The addition of water decreased organic content
area predestines it to be a good candidate for gas sorption, in the powders and increased surface area of the products.
sensor, and electrode materials. Its disturbed surface struc-
ture with a large number of unsaturated bonds endows it 2. Experimental
with a high catalytic activity and superparamagnetic
behavior of Fe2O3 nanoparticles in colloidal solutions led 2.1. Chemicals
to their use as magnetic fluids. Moreover, amorphous
Fe2O3 nanopowder was found to be a suitable precursor Fe(acac)3 was prepared from acetylacetone and
for solid-state systems of Fe2O3 or a-Fe magnetic nanopar- FeCl3 Æ 6H2O. Tetraethyleneglycol dimethylether (tetrag-
ticles [22]. The ultrasound-activated route led to the synthe- lyme) from Aldrich was vacuum distilled (93–105 C/
sis of amorphous Fe2O3 nanopowders of variable particle 3 Pa) and stored under dry N2.
size [16,17] and parameters influencing the yield in the
sonoreaction of Fe(CO)5 were examined in detail [23]. Fur- 2.2. Sonochemical experiments
thermore, nanocomposites of amorphous Fe2O3 with SiO2
[24] and TiO2 [25], mesoporous forms [21], amorphous Sonication was carried out on a Sonics and Materials
Fe2O3 coating layers on alumina [26] and surface-function- VXC system with 500 W input power and working fre-
alized amorphous Fe2O3 particles can be prepared by soni- quency of 20 kHz. In a typical experiment, iron(III) acetyl-
cations of Fe(CO)5 solutions [27]. acetonate precursor (0.50 g, 1.4 mmol) was dissolved in
Fe(acac)3 is an easily available, organics-soluble, solid tetraglyme (50 cm3) to a clear red solution (0.028 M) which
precursor that is, in contrast to Fe(CO)5, nontoxic and thus was then purged with argon for 10 min before irradiation
safe to handle on a large scale. Thermal decomposition of and also during the reaction (flow rate 60 cm3 min 1).
this precursor to Fe2O3 in air at normal pressure starts at The high-intensity Ti horn with a diameter of 13 mm and
170 C and is completed below 400 C [28]. It was used in tip surface area of 3.5 cm2 was inserted into a standard
the preparation of iron oxide nanoparticles by solution conical 250 cm3 sonochemical vessel (Ace Glass) which
thermolysis in various organic solvents [14,29], sol–gel was filled with 50 cm3 of the reaction mixture to ensure a
methods [30], and precursor thermolysis in an MCM41 constant immersion depth of 4 cm. The power transmitted
matrix [31]. Sonochemical decomposition of Fe(acac)3 into the solution was determined calorimetrically as
was studied only sparsely. For example, sonolysis in hex- 13 W cm 2 at 75% amplitude. The gas-tight connection
adecane was found to follow a first order kinetic law with of the horn to the vessel was ensured by a Teflon adapter
respect to Fe(acac)3 concentration and from the estimated attached at the nodal point and excess gas was released
sublimation temperature of 220 C, authors concluded that through an oil bubbler. The irradiation was set to an inter-
the reaction takes place inside the cavitation bubbles [32]. val of 2 s with a delay of 2 s. Before starting any experi-
The solid oxide product was X-ray amorphous, contained ment, the content of the reaction vessel was precooled
an admixture of ligand degradation products, and featured with a Julabo F 25-MP thermostat to 0 C. Sonication
characteristic acetate bands in the IR spectra. By heating to was run for 4, 8, 12, 16, 18, and 38 h. The solution temper-
700 C, it was converted to a-Fe2O3 with 20–40 nm particle ature rose to about 40 C during irradiation. After the reac-
size. tion, the resulting colloidal solution was left to stand for a
Tetraglyme is an ether type of solvent that has physical week and then filtered to remove a small amount of fine Ti
and chemical properties suitable for sonochemical experi- powder sputtered from the horn. The product was precip-
ments. It dissolves well diketonate precursors and is also itated by the addition of hexane (20 cm3) and the solid
known to coordinate through its oxygen atoms to metal was separated from the solvents by centrifugation for
J. Pinkas et al. / Ultrasonics Sonochemistry 15 (2008) 257–264 259

20 min on a Heraeus Labofuge 400 at 3000 rpm. The pow- ous temperatures between 300 and 5 K using a cryomag-
ders were washed to remove organic-soluble residues by netic system of Oxford Instruments. In-field spectra were
three cycles of resuspending them in isopropanol or light recorded at 5 K in an external magnetic field of 5 T applied
petroleum (15 cm3) and subsequent centrifugations. They parallel with the gamma ray propagation. Isomer shift
were left to dry in an open air. parameters are related to metallic a-Fe as a reference mate-
rial. IR spectra (4000–400 cm 1) were recorded on a Bru-
2.3. Characterization techniques ker EQUINOX 55/S/NIR or IFS 28 instrument. Samples
of solids were prepared as KBr pellets while neat liquids
Thermal analyses (TG/DTG/DTA) were carried out on were spread between KBr disks. Microanalyses were per-
a MOM Derivatograph-C instrument under static air of formed on a ThermoFinnigan Co. CHN Analyzer Flash
flowing Ar (7 dm3 min 1) with 5 C min 1 ramp from EA 1112.
room temperature to 1000 C. Samples (20–50 mg) were
contained in corundum crucibles. The a-Al2O3 powder 3. Results and discussion
was used as a reference material. DSC measurements
(DSC XP-10, THASS GmbH) were performed under static 3.1. Synthesis
air within the range of 25–400 C with the heating rate of
2.5 C min 1. A Stoe-Cie transmission diffractometer Tetraglyme was selected as an inert solvent for the sono-
STADI P operating with a Ge monochromatized Co chemical experiments because its physical and chemical
(k = 0.1788965 nm) radiation (40 kV, 30 mA) and parameters are favorable for the ultrasound processing:
equipped with a PSD detector was used for the XRD data its high normal boiling point (276 C) imparts a low vapor
acquisition at room temperature. ICP spectrometer Jobin pressure that is necessary for effective cavitation. This boil-
Yvon 170 Ultrace (lateral observation, generator ing point is higher than the one of a frequently used sono-
40 MHz, output 1.0 kW, plasma gas flow 12 l min 1, chemical solvent decalin (cis-decalin 195.8 C). Tetraglyme
monochromator 1 m) was used for the determination of density (1005.68 kg m 3 [33]) and ultrasound speed
the Fe metal content in the synthesized samples. The mea- (1381 m s 1 [34]) also allow to attain a sufficient acoustic
surements were performed at the 238.204 nm line. Scanning pressure amplitude.
electron microscopy (SEM) micrographs were taken on a We studied the influence of water added to the reaction
Jeol 6700 field emission microscope. Dried Fe2O3 powder solution with the aim of facilitating the protonation and
samples were resuspended in isopropanol and a drop of removal of the diketonate ligand from the iron center as
colloidal solution was placed on a cleaned silicon (1 0 0) Hacac and the formation of an oxide. Beneficial effects of
slide. After solvent evaporation, the sample was briefly added H2O were observed in the CVD fabrication of cop-
sputtered with Au/Pd alloy to avoid charging. Transmis- per oxide films from Cu(hfacac)2, where 18O labelling study
sion electron microscopy (TEM) was carried out on JEOL evidenced the retention of water oxygen in the solid oxide
JEM2010 and 3010 microscopes operated at 200 and product and decomposition of the hfacac ligands to ace-
300 kV, respectively. Instruments were equipped with a tone and acetate [35]. Sonolytical reactions were carried
LaB6 cathode, an EDX (energy dispersive X-ray) detector, out for comparison in dried tetraglyme under dry Ar gas,
and characterized by a point-to-point resolution of 1.7 Å. with water added to the reaction mixture (2% or 6%) or
Images were recorded by a CCD camera with resolution under Ar gas wetted by passing through a water bubbler
of 1024 · 1024 pixels using the Digital Micrograph soft- at room temperature. The amount of water had an impact
ware package. Powder samples were dispersed in ethanol on the surface area and residual organics content in the
and the suspension was treated in ultrasound for 10 min. prepared samples as it will be discussed below (Table 1).
A drop of very dilute suspension was placed on a car- Sonolysis was typically carried out for 8 h. The variation
bon-coated grid and allowed to dry by evaporation at of the reaction time between 4 and 38 h had little effect
ambient temperature. Surface areas and pore volumes were on the yield.
determined by N2 adsorption at 77.4 K by volumetric tech- The ultrasound treatment of the Fe(acac)3 solution pro-
nique on a Quantachrome Autosorb-1MP instrument. duced ruby red colloidal solutions that were stable for
Prior to the measurements, the samples were degassed at weeks. We observed a small amount of rutile (JCPDS card
25 C for at least 24 h until the outgas rate was less than 21-1276) and pseudobrookite (JCPDS card 41-1432) in the
0.4 Pa min 1. The adsorption–desorption isotherm was calcined oxide product as a result of sputtering of metallic
measured for each sample at least three times. The specific Ti particles from the ultrasound horn. The synthetic proce-
surface area was determined by the multipoint BET dure was therefore modified and settling of Ti powder and
method with at least five data points with relative pressures filtration through a fine frit was performed to completely
between 0.05 and 0.23. Analysis was performed with the eliminate these admixtures. The addition of hexane to the
instrument software package. Transmission 57Fe Möss- colloidal solution induced the precipitation of a red-brown
bauer spectra of 512 channels were collected using a Möss- solid on standing overnight. Centrifugation, washing with
bauer spectrometer in a constant acceleration mode with a 3–4 portions of hexane and drying in air yielded 100–
57
Co(Rh) source. Measurements were carried out at vari- 190 mg of the product. Expected yield of pure Fe2O3 from
260 J. Pinkas et al. / Ultrasonics Sonochemistry 15 (2008) 257–264

Table 1
Dependence of the surface area, pore volume, and Fe and organics
contents on the amount of water in the reaction solution
Water SAa Hb Pore Fe C H TG mass
content (m2 g 1) volume (%) (%) (%) loss (%) at
(vol.%) (cm3 g 1) 500 C
Dry 48 H3 0.304 22.0 30.5 4.52 63.1
2 105 H3 0.611 35.6 22.6 3.48 46.0
6 147 H3 0.455 43.6 15.3 2.75 35.2
2 + wet 260 H2 0.291 42.0 15.5 2.73 37.3
Arc
a
SA = surface area.
b
H = hysteresis type.
c
Ar gas wetted by passing through a water bubbler.

0.50 g of staring iron complex is 113 mg. A higher product


weight resulted from residual organic groups that are
retained in the final solid. The higher retention of organics Fig. 1. SEM of sonochemically prepared Fe2O3 powder.
was particularly prominent in the reaction carried out
under dry conditions. Actual isolated yields recalculated
from elemental analyses to pure Fe2O3 were between 60%
and 75%. For elemental compositions see Table 1. The particle surface. Typically, C constants were in the range
solid product could be redispersed in isopropanol by shak- 35–80 attesting to the validity of BET approximation.
ing overnight. The total pore volumes (cm3 g 1) were estimated from
the amounts adsorbed at a relative pressure of about 0.97
(Table 1). The N2 isotherm of the as-prepared nanoparti-
3.2. Characterization cles is of IIb type with the hysteresis loop of H3 type.
The surface area of iron(III) oxides displayed relation to
SEM was used to establish the morphology of materials the amount of water in the reaction mixture and increased
isolated from the colloidal reaction solution. SEM images with its larger quantity (Table 1). For dry reaction condi-
of as-prepared iron(III) oxide were taken on samples pre- tions, a value of 48 m2 g 1 was obtained. Addition of
pared by resuspending the isolated powders in isopropanol 2 vol.% of H2O to the starting solution led to 105–
and these colloidal solutions were deposited on Si wafers. 110 m2 g 1, 6 vol.% to 147 m2 g 1 and wetting of the Ar
They were observed both as native and Au sputtered. There stream in a bubbler filled with water maximized the surface
was no noticeable difference in the particle size or shape of area to 260 m2 g 1. The difference between adding of excess
Au-coated and uncoated Fe2O3. The identical appearance water directly to the reaction mixture and delivering of
of the particles was also observed in the samples prepared moistened Ar gas was reflected in the larger surface area
from the tetraglyme colloidal solutions obtained directly for the later, but the organic group content reflected in ele-
after sonolysis by the solvent evaporation at 260 C under mental analysis data and weight loss at 500 C were
vacuum. A micrograph of a powder obtained from the approximately the same for both reaction setups.
reaction with 2 vol.% of water is shown in Fig. 1. It We tried to establish the chemical and phase composi-
revealed uniform spherical nanoparticles with a diameter tion of the prepared particles by the X-ray diffraction
of about 20 nm. Increasing the amount of water in the reac- experiment. However, the sonochemically prepared iron
tion mixture to 6 vol.% led to a smaller particle size (about oxide powders appeared amorphous in XRD. The crystal-
5–10 nm). However, the smaller particles are more highly lization behavior was examined in a separate study by the
agglomerated. On the other hand, the use of rigorously authors. The variable temperature XRD under isothermal-
dried tetraglyme solvent and desiccated Ar gas resulted in dynamic heating program and also under linear heating
larger particles (about 70 nm) that were fused to large ramp provided above 300 C different modifications of iron
irregular chunks. This shows that the product particle size oxide, namely hematite and maghemite with a controllable
can be controlled by adding water to the reaction. Addi- particle size. Amorphous state of the as-prepared oxide
tional complementary information was obtained form the below 300 C cannot be proved unambiguously from the
surface area measurements. The BET method applied to featureless XRD pattern alone. Therefore, the results of
the nitrogen isotherms provided the surface area of pow- the TEM and Mössbauer spectroscopy experiments have
ders. The specific surface area data of as-prepared samples to be also considered. As it was discussed above, the as-pre-
of Fe2O3 were obtained by N2 adsorption–desorption mea- pared iron oxide powders contain organic residues whose
surements at 77.4 K. Prior to analyses, samples were amount can be controlled by added water (see the results
degassed at 25 C for 24 h. This low temperature was of CHN elemental analyses in Table 1). We examined the
employed to avoid any thermally induced changes in the sonochemically prepared samples of iron(III) oxide by
J. Pinkas et al. / Ultrasonics Sonochemistry 15 (2008) 257–264 261

thermal analysis methods in the 25–1000 C range under is also of interest; therefore, we studied the product after
air and Ar to establish the amount of removable organic sonication by FT-IR spectroscopy in KBr pellets to gain
groups, dynamics of decomposition and transformation information about the characteristic organic groups. A rep-
temperatures and crystallization mechanism. Except a resentative IR spectrum is shown in Fig. 4. The striking
slightly different shape of exothermic peaks, no remarkable feature is the absence of the bands characteristic for the
changes in the heat effects, mass loss and transformation residual acac groups. Strong vibrations ms(C@O)ring and
temperatures were observed in argon atmosphere com- mas(CACAC)ring, that were observed in the product of
pared to air. A typical thermogram (simultaneous TG/ interaction of Fe(acac)3 with Fe2O3 at 1590 and
DTA) and a DSC curve recorded in air with heating rates 1530 cm 1, respectively [36], are clearly missing. The
of 5 and 2.5 C min 1 are depicted in Figs. 2 and 3, methine stretch at 3090 cm 1 is also absent. The spectrum
respectively. is dominated by the bands at 1566 and 1432 cm 1 that
Generally, decomposition of organics takes place in belong to the mas(COO) and ms(COO) vibrations of chelat-
three partially overlapping steps and the separation of ing acetate, respectively. The carboxylate deformation
the mass loss events is better at faster ramping. The end bands are observed at 660 and 615 cm 1. It is reasonable
temperatures of the complete organic group removal shift to suppose that under high-power ultrasonic conditions,
progressively to higher values at faster heating rates. The the conversion of acac to acetate takes place. Thermal
value of the total mass loss at 500 C depends on the decomposition of Fe(acac)3 under N2 was studied by ther-
amount of water added to the reaction mixture. Powders moanalytic and IR spectroscopic methods [28] which
prepared in dry tetraglyme under dried Ar featured the revealed the presence of acetates as intermediate species
largest weight drop of about 63%, samples synthesized in in the course of the reaction. Similarly, the conversion of
the presence of 2 vol.% of water released about 46% of surface bound acac groups to acetates was observed by
weight, while products of the reaction with 6 vol.% of IR on heating the films of alumina (200 C) [37] and zirco-
added water showed only 35% mass decrease on heating. nia (110 C) [38]. The binding mode of acetates on the par-
This demonstrates the beneficial influence of water on ticle surface can be deduced from the Deacon–Phillips rules
ligand release and the lower content of organic matrix. [39]. The difference between the asymmetric and symmetric
The nature of the organic species entrained in the particles COO stretch is 134 cm 1 which is comparable to 164 cm 1
for ionic acetate and points to the bridging binding mode.
Amorphous Fe2O3 samples were treated with water in
attempts to remove residual acetate. IR spectra after wash-
ing still displayed acetate bands attesting to stability of ace-
tate binding. Calcination at 5 C min 1 to 300 C decreased
the acetate band intensities and heating to 500 C resulted
in their near disappearance (see Fig. 4).
The results of all methods used to characterize the sono-
chemically prepared iron oxide powders show that particles
are composed of organic material and metal oxide. The

Fig. 2. TG and DTA traces of amorphous Fe2O3, heating rate 5 C min 1,


in air. 1.0

0.9

0.8

0.7
Absorbance

0.6

0.5

0.4

0.3

0.2

0.1

0.0
4000 3000 2000 1000
Wavenumber(cm-1)

Fig. 4. IR spectrum of the as-synthesized Fe2O3 powder in a KBr pellet


(red) and after calcination to 500 C (blue). (For interpretation of the
Fig. 3. DSC curve of amorphous Fe2O3 sample. Heating rate of references to color in this figure legend, the reader is referred to the web
2.5 C min 1. version of this article.)
262 J. Pinkas et al. / Ultrasonics Sonochemistry 15 (2008) 257–264

transmission electron microscopy provided several pieces techniques hinted at the composite character. Therefore,
of information on crystallinity and constitution of these we heated the samples dynamically up to 250 C to par-
composite particles. The amorphous nature of the sono- tially remove the acetate groups (see TGA and DSC results
chemical iron oxide product was revealed in TEM micro- above) and to expose the iron oxide primary particles. At
graphs shown in Fig. 5. The apparent size and shape also 250 C the crystallization exothermic effect is not yet
agree well with the data obtained from SEM. However, apparent in DSC curve and also XRD pattern reflects a
the particles appear to contain smaller oxide islands (1– high content of amorphous phase. TEM micrographs of
2 nm) dispersed in organic matrix Fig. 6. We tried to eluci- the heated sample showed clearly small oxide particles
date the true nature of the primary oxide particles as other (around 3 nm) embedded in a matrix. A conspicuous
change from featureless amorphous particles to crystallites
was observed in samples heated to 300 C with 5 C min 1
heating rate; the size was around 5 nm, in accordance with
XRD data. Furthermore, the analysis of electron diffrac-
tion pattern revealed the presence of maghemite.
It is worth to mention the gradual crystallization of
amorphous nanoparticles under the electron beam during
the TEM experiments. It was demonstrated by recording
the electron diffraction pattern during the exposure. At
the beginning of experiments, the sample is completely
amorphous. After several minutes, Bragg diffractions
appeared, that were characteristic for c-Fe2O3 (or Fe3O4)
produced by the crystallization (reduction of iron) under
the electron beam.
Magnetic characterization of the materials was obtained
from Mössbauer spectroscopy. Room temperature (RT)
Mössbauer spectra of sonochemically prepared iron oxides
display the broadened doublet (C = 0.53–0.56 mm s 1)
with the isomer shift of 0.33–0.34 mm s 1 and quadrupole
splitting of 0.81–0.85 mm s 1 (see Table 2). Such doublet
with a not-fully Lorentzian shape and similar isomer shift
values, corresponding to the high-spin Fe3+ ions in octahe-
dral environment, was previously observed in amorphous
Fig. 5. TEM of amorphous Fe2O3. Fe2O3 samples prepared by various other routes [11,40–
44]. As a unique phenomenon, only quadrupole interaction
is observed even at very low temperatures up to 28 K as
clearly seen from the thermal evolution of Mössbauer spec-
tra (see Fig. 7). The sextet component appears firstly at
20 K and the hyperfine magnetic field increases with the
temperature lowering down to 5 K. Let us emphasize that
such low magnetic transition temperature has not been
described for any of amorphous or nanocrystalline iro-
n(III) oxides yet [5,11,45]. For comparison, amorphous
Fe2O3 nanopowder prepared by the solid-state reaction
and containing ultrasmall particles with similar size (1–
3 nm) exhibited Mössbauer transition temperature of
about 60 K [11]. It seems that the unrecorded low magnetic
transition yielded by sonochemically prepared amorphous
particles is related not only to the restricted particle dimen-
sions but also to the suppressed interparticle magnetic
interactions due the acetate matrix.
In order to understand the magnetic regime and local
ordering in the iron environment, Mössbauer spectrum
was recorded at 5 K in an external magnetic field of 5 T
and compared with that collected in a zero-field at the same
temperature. The zero-field spectrum can be fitted by two
Fig. 6. TEM of amorphous Fe2O3 showing 2–3 nm oxide particles sextets differing slightly in isomer shift and mainly in
embedded in a matrix. hyperfine magnetic fields (see Table 2). Their origin is prob-
J. Pinkas et al. / Ultrasonics Sonochemistry 15 (2008) 257–264 263

Table 2
Hyperfine parameters and relative spectral areas of Mössbauer spectra of as-prepared amorphous and thermally treated Fe2O3 samples measured at
various temperatures and applied fields
Sample Bexta Tb Spectrum dc ±0.01 DEQ(eQ)d ±0.01 Bhyp(Beff)e ±0.5 RAf ±2 Assignment
(T) (K) (mm s 1) (mm s 1) (T) (%)
As-prepared am-Fe2O3g 0 300 Doublet 0.34 0.81 – 100 Octahedral Fe(III)
As-prepared am-Fe2O3 0 5 Sextet 0.46 0.02 43.4 68 Surface Fe(III)
Sextet 0.47 0.01 47.3 32 Bulk Fe(III)
As-prepared am-Fe2O3 5 5 Sextet 0.46 0.02 42.1 67 Surface Fe(III)
Sextet 0.48 0.01 46.7 33 Bulk Fe(III)
am-Fe2O3 heated up to 0 300 Doublet 0.34 0.84 – 92 am-Fe2O3 + SPh
250 C c-Fe2O3
Doublet 1.02 2.34 – 8 Fe(II)
am-Fe2O3 heated up to 0 300 Doublet 0.34 0.86 – 96 am-Fe2O3 + SP
300 C c-Fe2O3
Doublet 1.05 2.30 – 4 Fe(II)
am-Fe2O3 heated up to 6 20 Sextet 0.44 0.01 43.4 21 am-Fe2O3*
300 C Sextet 0.36 0.02 53.2 31 c-Fe2O3 – tetrahedral
Sextet 0.47 0 47.2 48 c-Fe2O3 – octrahedral
am-Fe2O3 heated up to 0 300 Sextet 0.34 0 41.8 39 Ferrimagnetic c-Fe2O3
360 C Doublet 0.34 0.90 – 58 am-Fe2O3 + SP c-Fe2O3
Doublet 1.04 2.32 – 3 Fe(II)
a
Bext = external magnetic field.
b
T = temperature of measurement.
c
d = isomer shift related to metallic iron.
d
DEQ(eQ) = quadrupole shift (splitting).
e
Bhyp(Beff) = hyperfine (effective) magnetic field; Beff = Bhyp+Bext.
f
RA = relative spectrum area.
g
am = amorphous.
h
SP = superparamagnetic fraction.
*
Amorphous phase fitted by one broadened sextet including contributions of both surface and bulk Fe(III) atoms.

also used for the in-field spectrum, where the effective mag-
netic fields are reduced partially compared to the values of
Bhyp in zero-field spectrum as an expected response to the
applied field. As a key conclusion, the second and the fifth
spectral lines are neither suppressed nor intensified as
should be expected in the external magnetic field applied
parallel with the gamma ray propagation if the ferrimag-
netic (maghemite) or antiferromagnetic (hematite) compo-
nents are present in the sample, respectively [5,45]. This is
another indirect marker of amorphicity of sonochemically
prepared iron(III) oxide. No or negligible change of inten-
sities of the second and the fifth spectral lines evidence for
an absence of the long range magnetic ordering as in spero-
magnetic systems. Such speromagnetic behavior has been
previously observed in Mössbauer spectra of ultrasmall
amorphous nanoparticles of various ferrites such as
CdFe2O4, ZnFe2O4 and NiFe2O4 [46,47].

Fig. 7. Zero-field Mössbauer spectra of amorphous Fe2O3 recorded


4. Conclusions
between 300 and 5 K and in-field spectrum measured at 5 K in external
field of 5 T applied parallel with gamma ray. A novel ultrasound-driven synthesis of the amorphous
Fe2O3 nanoparticles starting from nontoxic Fe(acac)3 as
a precursor and tetraglyme as a new sonochemical solvent
ably in non-equivalent surrounding of the surface Fe3+ was developed. As-prepared iron(III) oxide nanoparticles
ions being in the bonding interaction with the acetate are embedded in biocompatible acetate matrix resulting
groups compared to the bulk iron. The similar fit can be in magnetic composites with large surface area applicable
264 J. Pinkas et al. / Ultrasonics Sonochemistry 15 (2008) 257–264

mainly in biomagnetic separation processes. The amount of [18] W. Huang, X. Tang, I. Felner, Y. Koltypin, A. Gedanken, Mater.
water in the reaction mixture represents a key variable Res. Bull. 37 (2002) 1721.
[19] H. Schmidt, Appl. Organometall. Chem. 15 (2001) 331.
allowing control the organics content as well as the size [20] R. Vijayakumar, Y. Koltypin, X.N. Xu, Y. Yeshurun, A. Gedanken,
and surface area of the composite particles. The maximum I. Felner, J. Appl. Phys. 89 (2001) 6324.
surface area of 260 m2 g 1 can be reached if moistened Ar [21] D.N. Srivastava, N. Perkas, A. Zaban, A. Gedanken, Pure Appl.
gas is delivered to the reaction mixture. The presence of Chem. 74 (2002) 1509.
organic matrix significantly affects the magnetism and ther- [22] O. Schneeweiss, R. Zboril, N. Pizurova, M. Mashlan, E. Petrovsky, J.
Tucek, Nanotechnology 17 (2006) 607.
mal behavior of amorphous Fe2O3 nanophase. Such sur- [23] M. Sivakumar, A. Gedanken, Ultrason. Sonochem. 11 (2004)
face modified ultrasmall (1–2 nm) nanoparticles exhibit 373.
an unrecorded low magnetic transition temperature of [24] K.S. Suslick, T. Hyeon, M. Fang, A.A. Cichowlas, Mater. Sci. Eng. A
about 25 K, below which they behave as speromagnetic 204 (1995) 186.
highly magnetically disordered systems with a high contri- [25] N. Perkas, O. Palchik, I. Brukental, I. Nowik, Y. Gofer, Y. Koltypin,
A. Gedanken, J. Phys. Chem. B 107 (2003) 8772.
bution of the surface anisotropy. [26] Z. Zhong, Y. Zhao, Y. Koltypin, A. Gedanken, J. Mater. Chem. 8
(1998) 2167.
Acknowledgements [27] K.V.P.M. Shafi, A. Ulman, X. Yan, N.-L. Yang, C. Estournes, H.
White, M. Rafailovich, Langmuir 17 (2001) 5093.
This work was supported by the Grant Agency of the [28] H.M. Ismail, J. Anal. Appl. Pyrolysis 21 (1991) 315.
[29] N. Pinna, G. Garnweitner, M. Antonietti, M. Niederberger, J. Am.
Czech Republic (203/04/0296) and the Ministry of Educa- Chem. Soc. 127 (2005) 5608.
tion, Youth and Sports (MSM0021622410, 1M6198959201 [30] V. Matsura, Y. Guari, J. Larionova, C. Guerin, A. Caneshi, C.
and MSM6198959218). Authors thank Prof. Z. Travnicek Sangregorio, E. Lancelle-Beltran, A. Mehdi, R.J.P. Corriu, J. Mater.
for CHN elemental analyses, Dr. K. Novotny for ICP mea- Chem. 14 (2004) 3026.
surements, and Dr. M. Klementova for TEM micrographs. [31] S. Liu, P. Cool, L. Lu, E. Beyers, P. Van der Voort, E.F. Vansant, M.
Jiang, Micropor. Mesopor. Mater. 79 (2005) 299.
[32] S.I. Nikitenko, P. Moisy, A.F. Seliverstov, P. Blanc, C. Madic,
References Ultrason. Sonochem. 10 (2003) 95.
[33] C.A. Tovar, E. Carballo, C.A. Cerdeirina, M.I.P. Andrade, L.
[1] K.S. Suslick, J.P. Price, Annu. Rev. Mater. Sci. 29 (1999) 295. Romani, Fluid Phase Equilb. 136 (1997) 223.
[2] T.J. Mason, J.P. Lorimer, Applied Sonochemistry, Wiley–VCH, [34] J.N. Real, T.P. Iglesias, S.M. Pereira, M.A. Rivas, J. Chem.
Weinheim, 2002. Thermodyn. 34 (2002) 1029.
[3] A. Gedanken, Ultrason. Sonochem. 11 (2004) 47. [35] J. Pinkas, J.C. Huffman, D.V. Baxter, M.H. Chisholm, K.G. Caulton,
[4] Y. Mastai, A. Gedanken, in: C.N.R. Rao, A. Mueller, A.K. Chem. Mater. 7 (1995) 1589.
Cheetham (Eds.), The Chemistry of Nanomaterials, Wiley–VCH, [36] M. de Ridder, P.C. van de Ven, R.G. van Welzenis, H.H.
NY, 2004, pp. 113–169. Brongersma, S. Helfensteyn, C. Creemers, P. Van Der Voort, M.
[5] R. Zboril, M. Mashlan, D. Petridis, Chem. Mater. 14 (2002) 969. Baltes, M. Mathieu, E.F. Vansant, J. Phys. Chem. B 106 (2002)
[6] X. Liao, J. Zhu, W. Zhong, H.-Y. Chen, Mater. Lett. 50 (2001) 341. 13146.
[7] A.A. Khaleel, Chem. Eur. J. 10 (2004) 925. [37] J. Guzman, B.C. Mates, Langmuir 19 (2003) 3897.
[8] S. Veintemillas-Verdaguer, M.P. Morales, C.J. Serna, Appl. Orga- [38] P. Van Der Voort, R. van Welzenis, M. de Ridder, H.H. Brongersma,
nometall. Chem. 15 (2001) 365. M. Baltes, M. Mathieu, P.C. van de Ven, E.F. Vansant, Langmuir 18
[9] V. Chhabra, P. Ayyub, S. Chattopadhyay, A.N. Maitra, Mater. Lett. (2002) 4420.
26 (1996) 21. [39] G.B. Deacon, R.J. Phillips, Coord. Chem. Rev. 3 (1980) 227.
[10] C. Pascal, J.L. Pascal, F. Favier, M.L. Elidrissi Moubtassim, C. [40] P. Ayyub, M. Multani, M. Barma, V.R. Palkar, R. Vijayaraghavan, J.
Payen, Chem. Mater. 11 (1999) 141. Phys. C: Solid State Phys. 21 (1988) 2229.
[11] R. Zboril, L. Machala, M. Mashlan, V. Sharma, Cryst. Growth Des. [41] G. Schimanke, M. Martin, Solid State Ionics 136–137 (2000)
4 (2004) 1317. 1235.
[12] T. Hyeon, S.S. Lee, J. Park, Y. Chung, H.B. Na, J. Am. Chem. Soc. [42] C. Cannas, G. Concas, A. Musinu, G. Piccaluga, G. Spano, Z.
123 (2001) 12798. Naturforsch. 54 (1999) 513.
[13] M.F. Casula, Y.-W. Jun, D.J. Zaziski, E.M. Chan, A. Corrias, A.P. [43] M.D. Mukadam, S.M. Yusuf, P. Sharma, S.K. Kulshreshtha, J.
Alivisatos, J. Am. Chem. Soc. 128 (2006) 1675. Magn. Magn. Mater. 269 (2004) 317.
[14] F.X. Redl, C.T. Black, G.C. Papaefthymiou, R.L. Sandstrom, M. [44] C. Cannas, G. Concas, A. Falqui, A. Musinu, G. Spano, G.
Yin, H. Zeng, C.B. Murray, S.P. O’Brien, J. Am. Chem. Soc. 126 Piccaluga, J. Non-Cryst. Solid 286 (2001) 64.
(2004) 14583. [45] J. Tucek, R. Zboril, D. Petridis, J. Nanosci. Nanotechnol. 6 (2006)
[15] N.N. Mallikarjuna, B. Govindaraj, A. Lagashetty, A. Venkataraman, 926.
J. Therm. Anal. Calorim. 71 (2003) 915. [46] H. Guérault, J.-M. Grenèche, J. Phys.: Condens. Matter 12 (2000)
[16] X. Cao, R. Prozorov, Y. Koltypin, G. Kataby, I. Felner, A. 4791.
Gedanken, J. Mater. Res. 12 (1997) 402. [47] C.N. Chinnasamy, A. Narayanasamy, N. Ponpandian, K. Chatto-
[17] X. Cao, Y. Koltypin, R. Prozorov, G. Kataby, A. Gedanken, J. padhyay, H. Guéeault, J.-M. Grenèche, J. Phys.: Condens. Matter 12
Mater. Chem. 7 (1997) 2447. (2000) 7795.

You might also like