Вы находитесь на странице: 1из 10

7414 J. Phys. Chem.

B 2001, 105, 7414-7423

Synthesis and Properties of γ-Fe2O3 Nanoclusters within Mesoporous Aluminosilicate


Matrices

Lei Zhang,† Georgia C. Papaefthymiou,‡,§ and Jackie Y. Ying*,†


Department of Chemical Engineering and Francis Bitter Magnet Laboratory,
Massachusetts Institute of Technology, Cambridge, Massachusetts 02139
ReceiVed: January 17, 2001

Iron oxide nanoclusters were synthesized within mesoporous MCM-41 aluminosilicate matrices via evaporation-
condensation of volatile Fe(CO)5. The well-defined hexagonally packed cylindrical pore structure of MCM-
41 led to the derivation of γ-Fe2O3 particles with spherical and elongated morphologies. Magnetization studies
and Mössbauer spectroscopy indicated that the γ-Fe2O3/MCM-41 nanocomposites exhibited interesting
superparamagnetic behavior. A blue shift in the absorption edge relative to bulk iron oxide was noted in the
UV-Vis spectra. Strain at the particle-support interface and quantum confinement effects played a critical
role in determining the overall magnetic and optical behavior of the γ-Fe2O3/MCM-41 nanocomposites. The
iron oxide nanoclusters within the MCM-41 matrix showed high thermal stability and increased magnetization
when calcined at high temperatures.

Introduction toward filling the mesopores with ultrafine clusters and nano-
wires. Nanoclusters of Fe2O3, polyaniline, Ge, GaAs, InP, Pt,
Nanoclusters and nanoarrays have attracted a great deal of Pd, Ru have been deposited within the MCM-41 inorganic
research attention due to their potential use as structural, matrix.14-21
catalytic, magnetic, electronic, and optical materials.1,2,3,4,5
In this paper, mesoporous MCM-41 matrices were used to
Ferrimagnetic γ-Fe2O3 is of particular interest due to its
host the synthesis of magnetic γ-Fe2O3 nanoclusters. The
dominant role in magnetic storage media.6 Highly dispersed
resulting nanocomposite materials demonstrated unique mag-
γ-Fe2O3 nanoclusters possess unique magnetic and electronic
netic and optical properties associated with quantum confine-
characteristics, such as superparamagnetism and macroscopic
ment effects and particle-support interactions. Various experi-
quantum tunneling associated with size quantization and elec-
mental techniques were employed to characterize the MCM-41
tronic quantum confinement effects.7 By packing magnetic
matrix structure, the iron oxide phase, and the nanocluster
nanoclusters into a host matrix, tailored interfacial confinement
dispersion within the porous support. Size quantization effects
environment, intercluster interaction, and effective charge carrier
on the magnetic and optical properties of the γ-Fe2O3/MCM-
transport may be obtained. Such composite structures have been
41 nanocomposites were investigated. Finally, these new
achieved by using cross-linked ion-exchange polymers, zeolitic
findings were compared to our previously reported studies on
molecular sieves, and sol-gel derived materials.8,9,10
γ-Fe2O3 nanoparticles embedded within amorphous silica
Crystalline zeolitic matrices can host highly ordered super- matrices.11,12
structures of monodispersed magnetic particles but are limited
by their microporous dimensions (typically 0.5 to 1.3 nm). Sol-
gel derived silica materials can support particles of larger sizes Synthesis of Fe2O3 Clusters within Hexagonally-Packed
within their interconnected porous framework but suffer from Mesoporous Hosts
a low degree of particle packing order within the amorphous Synthesis and Characterization of Mesoporous Alumino-
matrices of broad pore size distributions. silicate. Al-doped mesoporous silica was synthesized from an
We have recently reported on two novel approaches for the inorganic siliceous precursor using organic cationic trimethyl-
synthesis of γ-Fe2O3 nanoclusters via (i) controlled chemical ammonium surfactants (C16H33(CH3)3NBr or CTAB) as a
precipitation of iron oxide followed by silica coating11,12 and supramolecular templating agent. A 3.65 g portion of CTAB
(ii) ion-exchange and oxidation of ferrous cations within a surfactants was completely dissolved in H2O. A 22.2 g portion
porous sulfonated silica matrix.11 Here, we describe the use of of 27% sodium silicate solution dispersed in 50 g of water was
MCM-41 with hexagonally-packed mesopores of uniform slowly added to the surfactant solution at room temperature with
diameter13 as a matrix for hosting guest nanoclusters for vigorous stirring. After being stirred for 5 min, the pH of the
fundamental studies and potential nanostructured device ap- mixture was adjusted to 11.5 with diluted H2SO4. A desired
plications. The one-dimensional cylindrical pore channels of amount of aqueous aluminum sulfate solution was then gradually
MCM-41 can be ideally suited for simultaneous control of introduced to the loosely bonded silica gel and stirred for 3 h
cluster size and morphology. Much effort has been devoted at room temperature before the sample was hydrothermally
treated at 100-150 °C for a desired period. The molar
* To whom correspondence should be addressed. composition of the wet gel could be represented as 1 SiO2: 0.1
†Department of Chemical Engineering, M. I. T.
‡ Francis Bitter Magnet Laboratory, M. I. T. CTAB: 120 H2O: x Al2O3 (where x ) 0 to 0.1). The solid
§ Present address: Department of Physics, Villanova University, Vill- powder obtained was washed with ethanol and water, filtered,
anova, PA 19085. and then calcined at 540 °C in air for 6 h to remove the organics.
10.1021/jp010174i CCC: $20.00 © 2001 American Chemical Society
Published on Web 07/18/2001
Synthesis and Properties of γ-Fe2O3 Nanoclusters J. Phys. Chem. B, Vol. 105, No. 31, 2001 7415

Figure 1. XRD patterns of calcined mesoporous AlSi-R samples of Figure 3. BJH desorption pore size distribution of calcined AlSi-25
different Al dopant concentrations aged at 100 °C: (a) R ) 5, (b) R ) calculated from the N2 desorption isotherm.
25, and (c) R ) 50.

Figure 2. N2 adsorption-desorption isotherm of calcined AlSi-25.

The samples were designated as AlSi-R, where R was given


by the precursor molar ratio of Si and Al.
The long-range pore packing order of the mesoporous Al- Figure 4. A schematic of the evaporation-condensation apparatus.
doped silica was characterized by X-ray diffractometry (XRD)
(Siemens D5000 θ-θ diffractometer operated at 45 kV and 40 mesoporous aluminosilicate (AlSi-25) was degassed at 350 °C
mA using Ni-filtered Cu KR, radiation) and transmission electron overnight. Typically, 3 mL of Fe(CO)5 and 0.2 g of mesoporous
microscopy (TEM) (JEOL 200CX, 200 kV). The XRD patterns AlSi-25 were introduced to an evaporation flask and a receiving
of calcined AlSi-R shown in Figure 1 corresponded to a typical tube, respectively. After assembling the two parts, the meso-
hexagonal P6m space group with the intense (100) peak as well porous aluminosilicate was further evacuated for 3 h at 250 °C,
as minor peaks corresponding to (110), (200) and (210) which was maintained by a heating tape around the receiving
diffractions. tube. The receiving tube was then cooled to the desired
N2 adsorption analysis was performed on a Micromeritics temperature. Upon opening valves 1 and 2, Fe(CO)5 was
ASAP 2010 gas sorption instrument. The N2 adsorption- evaporated at room temperature and condensed within the
desorption isotherm of the calcined AlSi-25 sample showed mesopores of the aluminosilicate support. Excess Fe(CO)5 was
negligible hysteresis (Figure 2). The sample has a narrow then pumped out of the receiving tube to the cold trap by
Barrett-Joyner-Halenda (BJH) pore size distribution (Figure opening valve 3. The mesoporous silica containing Fe(CO)5 was
3), with a Brunauer-Emmett-Teller (BET) surface area of 1217 next subjected to a H2 stream at 100-200 °C for several hours
m2/g. to decompose and reduce the iron complex. The as-prepared
Introduction of Fe2O3 Nanoclusters by Evaporation- Fe2O3/AlSi-25 nanocomposite was further calcined in a tube
Condensation. The microstructural characteristics of porous furnace under flowing O2 for 3 h at the desired temperature.
materials, such as surface area, pore diameter, and pore size The iron loading depended on the synthesis parameters, such
distribution, would affect the local environment for hosting as the amount of Fe(CO)5, evaporation rate, saturation time,
quantum-sized metal oxide clusters. MCM-41 materials with reaction temperature, and decomposition conditions. The evapo-
hexagonally packed mesopores of 2-10 nm diameter may ration-condensation procedure could be repeated several times
promote the formation of ordered superstructures of the magnetic to achieve a high iron loading.
phase within their well-defined 1-D cylindrical pore channels.
Both the pore morphology and the dielectric properties of the Results and Discussions
mesoporous aluminosilicate materials are attractive matrix Microstructure of Fe2O3/Mesoporous Aluminosilicate Nano-
characteristics. composites. Three Fe2O3/AlSi-25 nanocomposite samples were
An evaporation-condensation apparatus was used in our synthesized with different Fe loadings. Elemental analysis results
synthesis (Figure 4). Fe(CO)5, a highly volatile precursor, was from direct-current plasma emission spectroscopy (Luvak Inc.)
carefully stored under argon and kept in the dark, while are summarized in Table 1. BET surface area and pore volume
7416 J. Phys. Chem. B, Vol. 105, No. 31, 2001 Zhang et al.

Figure 5. XRD patterns (i: 2θ ) 1.5-20° and ii: 2θ ) 15-75°) of Fe2O3/AlSi-25 nanocomposite #2: (a) as-prepared, (b) calcined at 300 °C,
(c) calcined at 500 °C, and (d) calcined at 800 °C.

TABLE 1: Characterization of Calcined Mesoporous those on the dark field image. Elemental mapping with the point
AlSi-25 and As-Prepared Fe2O3/A1Si-25 Nanocomposites probe (2 nm probe size) was thus employed to verify the Fe
Fe2O3/ Fe2O3/ Fe2O3/ locations. When the X-ray probe was pointed at the bright
sample AlSi-25 AlSi-25 #1 AlSi-25 #2 A1Si-25 #3 spherical spots shown in the dark field image, strong Fe
Fe loading 0 1.27 6.82 57.5 absorption was detected, indicating the presence of Fe. When
(wt %) the probe was moved to the adjacent dark regions, no Fe was
receiving tube - 25 50 80
temperature (°C) detected. In certain areas, a lamellar-like structure was noted
reduction 100 °C, 1 h; 200 °C, 5 h 125 °C, 5hr in the bright field image, corresponding to hexagonally packed
temperature then 150 °C, 5 h pore arrays oriented perpendicular to the electron beam. This
and time
surface area 1217 929 763 362 was viewed as elongated bright lines on the dark field image.
(m2/g) The point mapping along the bright lines corresponded to a
pore volume (cm2/g) 1.10 0.93 0.66 0.29 continuous detection of Fe absorption, indicating the presence
of elongated iron oxide particles. For the sample calcined at
of the Fe2O3/AlSi-25 samples as determined by N2 adsorption
500 °C, a homogeneous Fe mapping with dark field image was
analysis decreased with increasing Fe loading, as shown in Table
obtained, reflecting both spherical and elongated morphologies
1. At a high Fe loading, substantial pore plugging by large Fe2O3
for the iron oxide particles (Figure 8). No particle agglomeration
nanoclusters might have occurred. The particle agglomeration
or sintering was observed at 500 °C, suggesting a high thermal
at the higher loadings led to a significant reduction of the
stability for Fe2O3/AlSi-25 nanocomposite #1.
magnetization due to strong antiferromagnetic coupling among
the closely interacting particles. Magnetic and Optical Properties of γ-Fe2O3/Mesoporous
The XRD pattern of the as-prepared Fe2O3/AlSi-25 nano- Aluminosilicate Nanocomposites. For nanometer-sized single
composite #2 showed that the hexagonally packed mesostructure domain magnetic clusters, magnetic behavior is governed by
of AlSi-25 was preserved (Figure 5i(a)) and no additional peaks the magnetic energy barrier defined as the product of the cluster
associated with Fe2O3 crystalline phase was found (Figure 5ii volume and its magnetic anisotropy density.22 At a temperature
(a)). As the calcination temperature increased, the d(100) spacing T g TB (TB ) critical magnetic blocking temperature), the
of the hexagonal phase decreased slightly due to pore shrinkage thermal energy is comparable to the magnetic energy barrier.
from condensation. A broad band at 2θ ∼ 22° associated with The cluster magnetization vector is thermally excited to fluctuate
amorphous SiO2 was noted in all XRD patterns in Figure 5ii. between the easy directions of magnetization in a manner
No crystalline peaks were observed at high 2θ angles (15°- analogous to that of paramagnetic atoms in an applied field,
75°), indicating the preservation of ultrafine grain sizes for iron except with a much larger magnetic moment. This is the well-
oxide clusters upon calcination. known phenomenon of superparamagnetism. For superpara-
The morphology of the Fe2O3 nanoclusters within the MCM- magnetic particles, the total magnetic moment displays a
41 matrix was studied using both TEM (Akashi 002B, 200 kV), paramagnetic behavior obeying a classical Langevin function
and scanning transmission electron microscopy (STEM) (Vacuum above the TB.
Generators HB603) with elemental mapping. For the as-prepared The macroscopic magnetic properties of our nanocomposites
and 500 °C-calcined Fe2O3/AlSi-25 nanocomposite #1, TEM were investigated by superconducting quantum interference
bright field images showed a hexagonally packed array with device (SQUID). magnetometry (Quantum Design, MPMS,
dark phase contrast inside the pores when viewing normal to characteristic measurement time (τm) ) 102 s) over a wide range
the axis of the hexagonally packed pores (Figure 6). The STEM of temperatures (5 K < T < 300 K) and applied magnetic fields
dark field image showed highly dispersed bright spherical spots (0 T < H < 3 T). Zero field cooling (ZFC) and field cooling
of ∼2-3 nm in diameter (Figure 7). The Fe elemental mapping (FC) magnetization curves were recorded, and magnetic hys-
illustrated that Fe was highly dispersed throughout the image teresis loops at low temperatures were studied.
area, with higher concentrations indicated by the bright spots. The internal magnetization and dynamic spin relaxation
Because of image drifting during scanning, it was very difficult processes of the nanocomposites were also examined with
to precisely match the Fe locations on the elemental map with Mössbauer spectroscopy. The fast 10-8 sec characteristic
Synthesis and Properties of γ-Fe2O3 Nanoclusters J. Phys. Chem. B, Vol. 105, No. 31, 2001 7417

nanocomposites to determine the iron oxide phase as γ-Fe2O3


and to establish its thermal stability against the γ f R phase
transformation.
Mössbauer Studies. In the cluster size range of 2 to 4 nm, a
large fraction of the iron atoms lies on the cluster surface
(assuming a surface layer comprised of two atomic layers), and
the Mössbauer spectra would reflect the electronic and magnetic
properties of surface iron ions to a large extent. Thus, variations
in surface and strain anisotropy would be reflected in (i) the
quadrupole splitting, (ii) the magnetic hyperfine field, (iii) the
magnetic anisotropy energy, blocking temperature, and magnetic
coercivity, and (iv) the collective magnetic excitations detected
below the blocking temperature. A simple idealized core-shell
model is proposed to analyze the electronic and magnetic
contributions from surface and interior iron cations. We assume
that the highly crystalline core is uniformly magnetized and
could undergo a coherent rotation when an external field (H) is
applied. However, the surface spins are canted relative to the
interior lattice magnetization direction due to strong pinning
forces at the surface, which would induce noncoherent spin
reversal modes.22,23
The Mössbauer spectra of the as-prepared Fe2O3/AlSi-25
nanocomposite #1 at various temperatures are shown in Figure
9. The overall temperature profile is consistent with the
observation of magnetically ordered particles of Fe2O3 exhibiting
dynamic superparamagnetic relaxation, due to thermally acti-
vated reversals of the particle magnetization.24 The dependence
of the superparamagnetic relaxation time on temperature (T),
magnetic anisotropy energy density (K), and particle volume
(V), is given by τ ) τo exp (KV/kT), where k is Boltzmann’s
constant and τo is a constant characteristic of the material.24 At
high temperatures, a superparamagnetic relaxation time shorter
than the Mössbauer measuring time (τ < 10-8 s) would obscure
observation of the internal magnetic order of the particles, giving
rise to a quadrupolar spectrum. At lower temperatures, super-
paramagnetic relaxation would be slowed, revealing the internal
magnetic order of the particles in a magnetically split Mössbauer
spectrum.
At T ) 300 K, the broad absorption lines (Figure 9a) called
for spectral fits with two superimposed quadrupole doublets of
similar isomer shift but different quadrupole splitting values,
as tabulated in Table 2. These two subcomponents were
Figure 6. TEM micrographs of Fe2O3/AlSi-25 nanocomposite #1: (a)
associated with interior and surface iron sites according to our
as-prepared and (b) calcined at 500 °C. proposed core-shell model for the analysis of electronic and
magnetic contributions from interior and surface atoms. The
measurement time for Mössbauer spectroscopy yielded informa- larger quadrupole splitting (1.25 mm/s) was associated with
tion on blocking temperature, below which spin relaxation was surface iron atoms due to ligand coordination distortion at the
slowed. This enabled the investigation of dynamic and static surface. A ratio of Asurface:Acore ) 39:61 was obtained, as
magnetic properties of the nanocomposites over a convenient measured by the absorption area under each subspectrum.
temperature range and provided the capability of zero external The spectra at lower temperatures were similarly fitted using
field microscopic magnetic study. Mössbauer spectra for the the core-shell model assuming a distribution of hyperfine fields.
nanocomposites were recorded as a function of temperature for At 4.2 K, where only magnetically split spectra were observed
4.2 K < T < 300 K using a liquid helium cryostat. They were (Figure 9f), two magnetic subsites were poorly resolved. The
computer-fitted to theoretical curves using a nonlinear least- smaller hyperfine field (see Table 2) compared to the corre-
squares curve fitting procedure. sponding values of bulk γ-Fe2O3 (496 kOe at room tempera-
For nanometer-sized particles, the crystallites could be too ture25 and 527 kOe saturation field at 0 K26) was associated
small or poorly developed to be amenable to XRD structural with the surface atoms. It has been known that the sharp
determination, rendering phase identification of the iron oxide discontinuity of the Fe-O-Fe superexchange bonds and canted
difficult. No crystalline peaks associated with the iron oxide spin structure on the surface would result in diminished magnetic
phase were observed in the XRD studies of the nanocomposites exchange. The contribution from surface spins was greatly
in Figure 5, as discussed earlier. We have made use of the suppressed at 4.2 K as Asurface:Acore ) 21:79. We attribute this
Mössbauer electronic parameters, the isomer shift (δ), the to the frozen core spin structure and persistent surface spin
quadrupole splitting (∆Eo), and the internal hyperfine field (Hhf)- fluctuations. At intermediate temperatures, the spectra exhibit
values in conjunction with the magnetization strength of the greater complexity due to the superposition of quadrupolar and
7418 J. Phys. Chem. B, Vol. 105, No. 31, 2001 Zhang et al.

Figure 7. Elemental mappings and STEM images of as-prepared Fe2O3/AlSi-25 nanocomposite #1.

Figure 8. Elemental mappings and STEM images of Fe2O3/AlSi-25 nanocomposite #1 calcined at 500 °C.

magnetic subspectra arising from the presence of a distribution The cluster size distribution can be obtained from the fraction
of particle sizes. f(T) of the clusters with volume greater than the critical volume
The observed temperature dependence of the hyperfine fields (Vc) defined by superparamagnetism.29 The plot of the fraction
(Hhfl and Hhf2) in Table 2 indicated that below the blocking of particles contributing to the magnetic subspectrum as a
temperature, where spin reversals were suppressed, the mag- function of temperature exhibited two inflection points (see
netization vectors of the particles might still undergo collective Figure 10). This suggested that there were two distinct blocking
magnetic excitations around the easy axis of magnetization.27 temperatures as determined by the inflection points, TB1 ) 60
These low-energy excitations could also prevent the observed K and TB2 ) 110 K, corresponding to two particle size ranges
hyperfine fields from reaching their saturation values above 0 in the as-prepared Fe2O3/AlSi-25 nanocomposite #1. This was
K. In addition, the surface fields (Hhf2) exhibited stronger in agreement with our STEM and point X-ray probe mapping
temperature dependence compared to the core fields (Hhfl) results, which indicated the presence of spherical and elongated
presumably due to surface-spin fluctuation modes.28 particles. We could associate the lower blocking temperature
Synthesis and Properties of γ-Fe2O3 Nanoclusters J. Phys. Chem. B, Vol. 105, No. 31, 2001 7419

Figure 9. Mössbauer spectra of as-prepared Fe2O3/AlSi-25 nanocomposite #1 collected at (a) 300 K, (b) 75 K, (c) 60 K, (d) 50 K, (e) 25 K, and
(f) 4.2 K. The solid line is a least-squares fit to the experimental data, assuming a superposition of sites as shown above the experimental points
(see text). The magnetic subcomponent was fitted to a distribution of magnetic hyperfine fields in order to reproduce the observed spectral broadening.

TABLE 2: Mo1 ssbauer Parameters for Fe2O3/AlSi-25 Nanocomposite #1


Fe2O3/AlSi-25 isomer shifta,b δ (mm/sec) quadrupole splittinga ∆Eo (mm/sec) hyperfine field Hhf (kOe)
nanocomposite #1 T (K) surface core surface core Hhf1 Hhf2
300 0.33 0.33 1.25 0.73
75 0.41 0.45 1.30 0.77 428.8 296.4
as-prepared 60 0.54 0.42 1.50 0.77 458.3 375.2
50 0.35 0.39 1.50 0.78 466.0 402.6
25 0.46 0.27 1.51 0.94 466.9 411.3
4.2 -0.026 487.9 458.2
calcination at 300 °C 300 0.36 0.34 1.24 0.74
4.2 -0.056 477.4 447.9
calcination at 500 °C 300 0.39 0.35 1.10 0.81
4.2 -0.132 487.5 442.2
bulk γ-Fe2O3 (A+B sites averaged) 298 0.35c 0 496d
a Uncertainties in δ and ∆E are ( 0.03 mm/s and ( 0.05 mm/s, respectively. b Isomer shift is relative to metallic iron at room temperature; 4.2
o
K values were fixed at those obtained at room temperature + 0.12 mm/s second-order Doppler shift. cReference (27). dReference (25).

identification as γ-Fe2O3. In support of such a conclusion, we


observed (i) the absence or negligible quadrupolar perturbation
on the magnetic spectra and (ii) the reduced values of the
hyperfine fields. Hematite (R-Fe2O3) exhibits a large quadrupolar
perturbation of 0.12 mm/s at 300 K and -0.22 mm/s below the
Morin transition (T ) 260 K for bulk R-Fe2O3), and a saturation
hyperfine field (Hsat) of 544 kOe;30 as opposed to zero
quadrupolar perturbation and a saturation field of 527 kOe for
γ-Fe2O3.26 In addition, hematite has a canted anti-ferromagnetic
structure with very weak magnetization. The SQUID measure-
ments discussed below indicated that these nanocomposites
possessed considerable magnetization, consistent with the fer-
rimagnetic spinnel structure of γ-Fe2O3.
The geometrical confinement of the Fe2O3 clusters by the
Figure 10. The fraction of particles in Fe2O3/AlSi-25 nanocomposite regular pore structure of the inorganic AlSi-25 matrix could
#1 contributing to the Mössbauer magnetic subspectrum as a function potentially stabilize the nanoparticles at high temperatures. The
of temperature. The arrows show the two distinct transitional regions. thermal stability of Fe2O3/AlSi-25 nanocomposite #1 samples
with a distribution of spherical particles, and the higher blocking calcined at 300 and 500 °C was examined by Mössbauer
temperature with a distribution of larger, elongated particles. spectroscopy. The values for the isomer shift and quadrupole
The Mössbauer parameters of the as-prepared Fe2O3/AlSi- splitting of both the surface and core components in the
25 nanocomposite #1 were consistent with the iron oxide phase Mössbauer spectrum at 300 K remained essentially unchanged
7420 J. Phys. Chem. B, Vol. 105, No. 31, 2001 Zhang et al.

Figure 12. Plots of magnetization vs applied field for as-prepared


Fe2O3/AlSi-25 nanocomposite #1 obtained at (a) 300 K and (b) 5 K.

Figure 13. Temperature dependence of the magnetization curves under


(a) field cooling and (b) zero field cooling for as-prepared Fe2O3/A1Si-
25 nanocomposite #1.

Figure 11. Mössbauer spectra collected at 4.2 or 300 K for Fe2O3/


AlSi-25 nanocomposite #1 calcined at (a) 300 °C and (b) 500 °C under due to both spin canting in the ultrafine particle assembly31 and
O2. The solid line is a least-squares fit to the experimental data, distribution in the particle sizes. The lack of magnetic saturation
assuming a superposition of sites as shown above the experimental became more significant at low temperatures. The onset of
points (see text). The magnetic subcomponent was fitted to a distribution hysteresis occurred at ∼50 K. At 5 K, a symmetric hysteresis
of magnetic hyperfine fields in order to reproduce the observed spectral loop was observed with an unsaturated magnetization of 57
broadening. emu/g at 30 kG applied field, a remanence of 1.5 emu/g and a
for the nanocomposite calcined at 300 °C (Figure 11a, Table coercivity of 680 G, demonstrating the ferrimagnetic nature of
2), indicating that no phase transition took place at this the nanocomposite at a low temperature. The spin canting
calcination temperature. Different quadrupole splitting values resulting from the discontinuity of the superexchange bonds
were obtained for both components for the sample calcined at between Fe cations was responsible for the lower magnetization
500 °C, suggesting an improved crystallinity due to atomic (compared to bulk γ-Fe2O3: Ms ) 76 emu/g) and the lack of
restructuring at this calcination temperature (Figure 11b, Table saturation for the nanocomposite.32 Compared to bulk γ-Fe2O3
2). The Mössbauer spectrum at 4.2 K gave a small quadrupole that has a remanence of 63 emu/g33 and a coercivity of 250-
splitting perturbation value (-0.056 mm/s) for the sample 400 G,34 the smaller remanence and larger coercivity of the
calcined at 300 °C, supporting the retention of the γ-Fe2O3 nanocomposite could be attributed to the single magnetic domain
phase. The 4.2 K spectrum of the sample calcined at 500 °C, clusters. For single-domain particles, coercivities are associated
however, showed a large quadrupole splitting perturbation value with the rotation of the magnetization away from the easy axis;
of -0.132 mm/s, which may be interpreted as a partial phase whereas in the bulk system, coercivities reflect magnetic domain
transformation from γ-Fe2O3 to R-Fe2O3 at this calcination wall motion in the presence of a magnetic field.22
temperature (∆Eo of pure R-Fe2O3 ) -0.22 mm/s at T ) 80 The temperature-dependent magnetization in an applied field
K).30 of 200 G was investigated under field cooling (FC) and zero
SQUID Studies. The isothermal magnetization of as-prepared field cooling (ZFC) (Figure 13). The ZFC curve (Figure 13b)
Fe2O3/AlSi-25 nanocomposite #1 was obtained at two different exhibited a maximum at 20 K and a superimposed contribution
temperatures (Figure 12). The hysteresis loop at 300 K gave a to the magnetization with a maximum at 50 K. The sharp Tmax
zero remanence (Mr) and coercivity (Hc), indicating superpara- at 20 K could be related to small spherical single-domain
magnetic behavior. With increasing external magnetic field, the magnetic particles. Because the Tmax is closely related to the
magnetization increased rapidly without saturation up to 30 kG blocking temperature and increases with increasing cluster size
Synthesis and Properties of γ-Fe2O3 Nanoclusters J. Phys. Chem. B, Vol. 105, No. 31, 2001 7421

TABLE 3: Summary of Magnetic Properties of the Fe2O3/


SiO2-based Nanocomposites
M Hc TB TB
as-prepared (emu/g Fe2O3) (G) (K) (K)
sample (at 3 T, 5 K) (at 5 K) (SQUID) (Mössbauer)
Fe2O3/ 30 270 10 25
Sulfonated-SiO2
nanocompositea
Fe2O3/ 20 635 15 23
SiO2-Coated
nanocompositeb
Fe2O3/AlSi-25 57 680 20, 50 60, 110
Figure 14. Temperature dependence of the magnetization curves under
nanocomposite #1
(a) field cooling and (b) zero field cooling for Fe2O3/AlSi-25 nano-
composite #1: as-prepared, or calcined to 300 °C and 500 °C. a Reference (11). b Reference (12).
optical properties of the Fe2O3/AlSi-R nanocomposites reflected
the nanometer-scaled dimensions of the magnetic constituents,
the abundance of interfaces, and particle-support interactions.
These effects may be further brought to focus by a comparison
of the properties exhibited herewith and those for Fe2O3/
sulfonated-SiO2 nanocomposite and Fe2O3/SiO2-coated nano-
particles reported in our previous studies.11,12
The magnetic properties of the three differently prepared
Fe2O3/SiO2-based nanocomposites in our current and previous
studies are summarized in Table 3. A complex magnetic system
was obtained within the cylindrical mesopores of MCM-41
matrix using the evaporation-condensation approach described
in this study. Large magnetization and coercivity values were
obtained using the AlSi-25 supports (bulk γ-Fe2O3: 76 emu/g,
polymer supported γ-Fe2O3 nanoclusters: 46 emu/g,35 block
copolymer supported γ-Fe2O3 nanoclusters: 39 emu/g36). This
Figure 15. UV-vis diffuse reflectance spectra of Fe2O3/AlSi-25
nanocomposite #1: (a) as-prepared, (b) calcined at 300 °C, and (c)
could be explained by the combination of magnetic contributions
calcined at 500 °C. from particles of different morphologies. For single magnetic
domain particles, the magnetization is significantly reduced due
and magnetic anisotropy constant, the broad Tmax at 50 K could to the quantum size effect. The magnetic dipole moment is
be attributed to elongated particles with larger average particle proportional to the particle volume. The combined effective
volume and shape anisotropy. These blocking temperatures are volume of the spherical and elongated Fe2O3 nanoclusters in
lower than those observed with Mössbauer spectroscopy due the cylindrical pores of AlSi-25 could be larger than that of the
to the longer measuring time associated with SQUID measure- spherical SiO2-coated Fe2O3 particles with diameters of ∼4 nm.
ments. Therefore, a larger magnetization was obtained with the Fe2O3/
SQUID magnetometry was also applied to study the magnetic AlSi-25 nanocomposite. The coercivity of spinel-structured
properties of the calcined samples. The magnetization vs nanocrystals greatly depends on morphology and surface
temperature curves measured under FC and ZFC at 200 G anisotropy. Compared to spherical particles, shape anisotropy
showed that the total magnetization increased with calcination in acicular particles can produce a higher coercive field,
temperature (Figure 14). The splitting between FC and ZFC assuming an ellipsoid model with uniform magnetization. The
curves occurred at a lower temperature for samples calcined at higher coercivity of Fe2O3/AlSi-25 nanocomposite #1 compared
a higher temperature. The improved crystallinity of the Fe2O3 with that of Fe2O3/SiO2-coated nanocomposite could be at-
core and the reduced contribution from disordered surface spins tributed to the presence of elongated particles in the former.
gave rise to the increase in magnetization. The partial trans- On the other hand, the low coercivity in the Fe2O3/sulfonated-
formation to R-Fe2O3, observed in the Mössbauer spectrum of SiO2 nanocomposite implied that (i) the internal spin structure
the 500 °C-calcined sample, was not significant enough to lead of the particles in this system significantly deviated from that
to a reduction in the overall magnetization. This study indicated of uniform magnetization assumed in the Stoner and Wohlfarth
that Fe2O3/AlSi-25 nanocomposite #1 would be an excellent model37 and/or (ii) the anisotropy was dominated by strain
candidate for applications requiring high-temperature stability. exerted by the sulfonated SiO2 matrix on the nanoclusters.
Optical Properties: The optical property of the Fe2O3/A1Si- The optical properties of the three nanocomposites are
25 nanocomposites was studied using a Cary II UV-vis compared in Figure 16. The matrix-mediated synthesis of Fe2O3
spectrometer. The spectrum of the as-prepared Fe2O3/AlSi-25 nanoclusters within sulfonated SiO2 gel involved an intercon-
nanocomposite #1 exhibited a blue shift of 70 nm relative to nected porous host with a relatively broad pore size distribution.
bulk γ-Fe2O3 (Figure 15a). The widening of the band gap was This might have resulted in relatively less stress on the magnetic
consistent with quantum confinement. Upon calcination, the particles, and thus, the largest blue shift (110 nm) was obtained
absorption edge shifted to slightly higher wavelengths (Figures with the Fe2O3/sulfonated-SiO2 nanocomposite (Figure 16a). In
15b and c), corresponding to changes in (i) the core structure the case of the Fe2O3/SiO2-coated sample, the presence of high
of the Fe2O3 nanoclusters, and (ii) the interfacial effects from stress exerted by the SiO2 coating might have substantially
the particle-support interaction. counteracted the quantum confinement effects, resulting in a
Comparison of Fe2O3/SiO2-Based Nanocomposites Syn- relatively small blue shift of 20 nm (Figure 16b). The strain
thesized via Different Routes. The electronic, magnetic, and induced by particle-support interactions could significantly
7422 J. Phys. Chem. B, Vol. 105, No. 31, 2001 Zhang et al.

References and Notes


(1) (a) Gletier, H. Prog. Mater. Sci. 1989, 33, 233. (b) Tronc, E.; Jolivet,
J. P. In Nanophase Materials: Synthesis-Properties-Applications; Hadji-
panayis, G. C., Siegel, R. W., Eds.; Kluwer: Boston, 1994; p 21. (c) Gleiter,
H. Mater. Sci. Forum 1995, 189-190, 67. (d) Shi, J.; Gider, S.; Bakcock,
K.; Awschalom, D. D. Science 1996, 271, 937. (e) Greer, A. L. Mater. Sci.
Forum 1998, 269-272, 3.
(2) (a) Thomas, J. M. Pure Appl. Chem. 1988, 60, 1517. (b) Ying, J.
Y. AIChE J. 2000, 46, 902. (c) Gates, B. C. Chem. ReV. 1995, 95, 511.
(3) (a) Simon, A. Angew. Chem., Int. Ed. Engl. 1988, 27, 159. (b)
Chien, C. L. Annu. ReV. Mater. Sci. 1995, 25, 129. (c) Leslie-Pelecky, D.
L.; Rieke, R. D. Chem. Mater. 1996, 8, 1770. (d) Schmid, G.; Hornyak, G.
L. Curr. Opin. Solid State Mater. Sci. 1997, 2, 204.
(4) (a) Valokitin, Y.; Sinzig, J.; DeJongh, L. J.; Schmid, G.; Moiseev,
I. I. Nature 1997, 384, 621. (b) Luth, H. Appl. Surf. Sci. 1998, 130-132,
855. (c) Chen, W.; Ahmed, H. AdV. Imaging Electron Phys. 1998, 102, 87.
(5) (a) Simon, U.; Schön, G. Angew. Chem., Int. Ed. Engl. 1993, 32,
250. (b) Gehr, R. J.; Boyd, R. W. Chem. Mater. 1996, 8, 1807.
(6) Berkowitz, A. E. In Nanomaterials: Synthesis, Properties and
Figure 16. UV-vis diffuse reflectance spectra of (a) Fe2O3/sulfonated- Applications; Edelstein, A. S., Cammarata, R. C., Eds.; Institute of Physics
SiO2 nanocomposite, (b) Fe2O3/SiO2-coated nanocomposite, and (c) Publishing: Bristol, 1996; p 569.
Fe2O3/AlSi-25 nanocomposite #1. (7) (a) Morrish, A. H. The Principles of Magnetism; Wiley: New York,
1996; Chapter 7. (b) Stucky, G. D.; MacDougall, J. Science 1990, 247,
669. (c) Gunther, L. In Magnetic Properties of Fine Particles; Dormann, J.
distort the particle’s crystal lattice as to produce red-shifted L.; Fiorani, D., Eds.; Elsevier: Amsterdam, 1992; p 213. (d) Dormann, J.
optical absorption as reported for a γ-Fe2O3/polymer nanocom- L.; Fiorani, D.; Tronc, E. NATO ASI Series E 1994, 260, 635.
posite system.35 The hexagonally packed array of cylindrical (8) (a) Golden, J. H.; Deng, H.; DiSalvo, F. J.; Fréchet, J. M. J.;
Thompson, P. M. Science 1995, 268, 1463. (b) Ziolo, R. F.; Giannelis, E.
mesopores (2.5 nm pore diameter) of the AlSi-25 support could P.; Shull, R. D. Nanostr. Mater. 1993, 3, 85. (c) Nguyen, M. T.; Diaz, A.
induce more stress on the surface of Fe2O3 clusters than the F. AdV. Mater. 1994, 6, 858.
irregular pore structure of the sulfonated SiO2 gel (4 nm average (9) (a) Shull, R. D.; Ritter, J. J.; Shapiro, A. J.; Swartzendruber, L. J.;
pore size) but less stress than that imposed by the silica coating. Bennett, L. H. J. Appl. Phys. 1990, 67, 4490. (b) Shull, R. D.; Kerch, H.
M.; Ritter, J. J. J. Appl. Phys. 1994, 75, 6840. (c) Wang, J. P.; Luo, H. L.
The combined quantum confinement and stress effects led to J. Appl. Phys. 1994, 75, 7425.
the intermediate level of optical blue shift (70 nm) in Fe2O3/ (10) (a) Bein, T.; Schmiester, G.; Jacobs, P. A. J. Phys. Chem. 1986,
AlSi-25 nanocomposite #1 (Figure 16c). 90, 4851. (b) Herron, N.; Wang, Y.; Eddy, M.; Stucky, G. D.; Cox, D. E.;
Moller, K.; Bein, T. J. Am. Chem. Soc. 1989, 111, 530. (c) Lazaro, F. J.;
Lopez, A.; Larrea, A.; Pankhurst, Q. A.; Nieto, J. M.; Corma, A. IEEE
Conclusions Trans. Magn. 1998, 34, 1030.
(11) Zhang, L.; Papaefthymiou, G. C.; Ziolo, R. F.; Ying, J. Y. Nanostr.
Evaporation-condensation of volatile Fe(CO)5 was used to Mater. 1997, 9, 185.
produce Fe2O3 nanoclusters within a hexagonally packed (12) Zhang, L.; Papaefthymiou, G. C.; Ying, J. Y. J. Appl. Phys. 1997,
81, 6892.
mesoporous aluminosilicate matrix. Various Fe loadings could (13) (a) Kresge, C. T.; Leonowicz, M. E.; Roth, W. J.; Vartuli, J. C.;
be obtained through control of cluster synthesis parameters. Beck, J. S. Nature 1992, 359, 710. (b) Huo, Q.; Margolese, D. I.; Ciesla,
Superparamagnetism was demonstrated for the Fe2O3/AlSi-25 U.; Feng, P.; Gier, T. E.; Sieger, P.; Leon, R.; Petroff, P. M.; Schüth, F.;
nanocomposites at low Fe loadings. TEM and STEM with Fe Stucky, G. D. Nature 1994, 368, 317. (c) Tanev, P. T.; Chibwe, M.;
Pinnavaia, T. J. Nature 1994, 368, 321. (d) Yang, H.; Kuperman, A.;
elemental mapping indicated the coexistence of spherical and Coombs, N.; Mamiche-Afara, S.; Ozin, G. A. Science 1996, 379, 703. (e)
elongated iron oxide particles within the cylindrical pores of Schulz-Ekloff, G.; Rthaisky, J.; Zukol, A. J. Inorg. Mater. 1999, 1, 97. (f)
AlSi-25 supports. The SQUID and Mössbauer properties were Ying, J. Y.; Mehnert, C. P.; Wong, M. S. Angew. Chem., Int. Ed. 1999, 38,
56.
consistent with γ-Fe2O3. Two distinct relaxation phenomena (14) MacLachlan, M. J.; Aroca, P.; Coombs, N.; Manners, I.; Ozin. G.
were observed in both SQUID and Mössbauer measurements, A. AdV. Mater. 1998, 10, 144.
associated with the presence of two sets of particle size (15) Wu, C.-G.; Bein, T. Chem. Mater. 1994, 6, 1109.
distributions. The large magnetization and coercivity of Fe2O3/ (16) Leon, R.; Margolese, D.; Stucky, G. D.; Petroff, P. M. Phys. ReV.
B 1995, 52, 2285.
AlSi-25 nanocomposite can be attributed to the presence of (17) Srdanov, V. I.; Alxneit, I.; Stucky, G. D.; Reaves, C. M.; DenBaars,
spherical and elongated magnetic particles. The Fe2O3 nano- S. P. J. Phys. Chem. B 1998, 102, 3341.
clusters in this nanocomposite exhibited high stability against (18) Agger, J. R.; Anderson, M. W.; Pemble, M. E.; Terasaki, O.; Nozue,
Y. J. Phys. Chem. B 1998, 102, 3345.
γ f R phase transition. Only a partial phase transition to (19) (a) Junges, U.; Jacobs, W.; Voiget-Martin, I.; Krutzsch, B.; Schüth,
R-Fe2O3 was detected by Mössbauer spectroscopy for the sample F. J. Chem. Soc., Chem. Commun. 1995, 2283. (b) Corma, A.; Martı́nez,
calcined at 500 °C. The high thermal stability and increased A.; Martı́nez-Soria, V. J. Catal. 1997, 169, 480.
magnetization from calcination make this material an excellent (20) Mehnert, C. P.; Ying, J. Y. Chem. Commun. 1997, 2215. (b)
Mehnert, C. P.; Weaver, D. W.; Ying, J. Y. J. Am. Chem. Soc. 1998, 120,
candidate for high-temperature applications. A blue shift in the 12 289. (c) Koh, C. A.; Nooney, R.; Tahir, S. Catal. Lett. 1996, 163, 148.
absorption edge was noted in the optical spectrum of this (21) Mumukutla, R. S.; Asakura, K.; Namba, S.; Iwasawa, Y. Chem.
nanocomposite system. Commun. 1998, 1425.
(22) Cullity, B. D. Introduction to Magnetic Materials; Addison-
Wesley: Reading, 1972; Chapters 9 and 11.
Acknowledgment. This work was supported by the Camille (23) Pankhurst, Q. A.; Pollard, R. J. Phys. ReV. Lett. 1991, 67, 248.
and Henry Dreyfus Foundation. The authors thank M. Frongillo (24) Aharoni, A. In Magnetic Properties of Fine Particles; Dormann,
and A. G. Garratt-Reed (MIT CMSE) for their assistance in J. L., Fiorani, D., Eds.; Elsevier: North-Holland, 1992; p 3.
(25) Kelly, H. V.; Folen, V. J.; Hass, M.; Schreiner, W. N.; Beard, G.
the TEM and STEM studies. G.C.P. thanks the Centro Brasileiro- B. Phys. ReV. 1961, 122, 1447.
de Pesquisas Fı́sicas for support during her Visiting Scientist (26) Murad, E. In Magnetic Properties of Fine Particles; Dormann, J.
appointment in Brazil. The authors would also like to acknowl- L.; Fiorani, D., Eds.; Elsevier: North-Holland, 1992; p 339.
(27) Mørup, S.; Topsøe, H. J. Appl. Phys. 1976, 11, 63.
edge Drs. R. Scorzelli and A. Bustamante Dominguez of the (28) de Jongh, L. J.; Miedema, A. R. AdV. Phys. 1974, 23, 1.
Centro Brasileiro de Pesquisas Fı́sicas for the low-temperature (29) Kaufman, K. S.; Papaefthymiou, G. C.; Frankel, R. B.; Rosenthal,
Mössbauer measurements. A. Biochim. Biophys. Acta 1980, 522.
Synthesis and Properties of γ-Fe2O3 Nanoclusters J. Phys. Chem. B, Vol. 105, No. 31, 2001 7423

(30) Greenwood, N. N.; Gibb, T. C. Mössbauer Spectroscopy; Chapman (35) (a) Ziolo, R. F.; Giannelis, E. P.; Weinstein, B. A.; O’Horo, M. P.;
Hall: London, 1971; p 241. Ganguly, B. N.; Mehrotra, V.; Russell, M. W.; Huffman, D. R. Science
(31) Coey, J. M. D. Phys. ReV. Lett. 1971, 27, 1140. 1992, 257, 219. (b) Vassiliou, J. K.; Mehrotra, V.; Russell, M. W.; Giannelis,
(32) Teillet, J.; Bouree, F.; Krishnan, R. J. Magn. Magn. Mater. 1993, E. P.; McMichael, R. D.; Shull, R. D.; Ziolo, R. F. J. Appl. Phys. 1993, 73,
123, 93. 5109.
(33) Imaoka, Y. J. Electrochem. Jpn. 1968, 36, 15. (36) Sohn, B. H.; Cohen, R. E.; Papaefthymiou, G. C. J. Magn. Magn.
(34) Morrish, H. In Crystals: Growth, Properties, and Applications; Mater. 1998, 182, 216;
Freyhardt, H. C., Mgr Ed.; Springer-Verlag: New York, Vol. 2, 1980; p (37) Stoner, E. C.; Wohlfarth, E. P. Philos. Trans. R. Soc. A 1948, 240,
171. 599.

Вам также может понравиться