Vous êtes sur la page 1sur 8

Formation Pathways of Magnetite Nanoparticles by Coprecipitation

Method
Taebin Ahn,

Jong Hun Kim,

Hee-Man Yang,

Jeong Woo Lee,

and Jong-Duk Kim*


,,

Department of Chemical and Biomolecular Engineering (BK21 Graduate Program), Korea Advanced Institute of Science and
Technology, Daejeon 305-701, Republic of Korea

Center for Energy and Environment Engineering, Korea Advanced Institute of Science and Technology, Daejeon 305-701, Republic
of Korea
*S Supporting Information
ABSTRACT: Magnetite nanoparticles for biomedical appli-
cations are typically prepared using the coprecipitation
technique, which is the most convenient method. However,
the reaction pathways leading to the production of the
magnetite phase in the coprecipitation reaction are not fully
understood, despite the fact that the reaction path may be of
significant importance in controlling the crystal structure,
morphology, and particle size of the magnetite nanoparticles.
In the present study, we identified the reaction pathways in the
coprecipitation of magnetite; when base was slowly added to
an iron chloride solution, akaganeite nucleated and transformed through goethite to magnetite. At high addition rates, an
additional pathway in which ferrous hydroxide nucleated and transformed through lepidocrocite to magnetite competed with the
former pathway. This difference was due to the pH inhomogeneity in the reaction medium that was present before homogeneous
mixing. In most coprecipitation reactions, these magnetite formation pathways coexist, but the dominant process is the topotactic
transformation of goethite to magnetite, mediated by arrow-shaped nanoparticles. The morphology of the arrow-shaped
nanoparticles was explained on the basis of specific crystallographic relationships among the iron oxide phases. The proposed
reaction scheme for magnetite coprecipitation could assist in devising a more detailed study of the reaction mechanism.
1. INTRODUCTION
Processes for synthesizing magnetite (Fe
3
O
4
) nanoparticles
have been extensively investigated over the past several decades,
owing to the biocompatibility and high saturation magnet-
ization of such particles, which make them suitable for diverse
biomedical applications.
15
The synthesis of iron oxides can be
complicated, as there are as many as 16 distinct species in the
form of oxides, hydroxides, and oxyhydroxides. However,
magnetite, with an inverse spinel structure, can be synthesized
by well-defined processes, such as the thermal decomposition,
6
microemulsion,
7
hydrothermal,
8
and coprecipitation
912
meth-
ods.
Among the various techniques for magnetite synthesis, the
coprecipitation method is a convenient way to synthesize
magnetite nanoparticles from an aqueous iron salt (Fe
2+
+
Fe
3+
) solution; a base is simply added under an inert
atmosphere at room temperature. The coprecipitation process
does not produce or use any toxic intermediates or solvents,
does not require precursor complexes, and proceeds at
temperatures under 100 C. This process has been recognized
for its industrial importance because of its ability to be scaled
up, its reproducibility, and its eco-friendly reaction con-
ditions.
4,5
However, it yields particles with a broad size
distribution,
1,4,5
probably because of the complicated set of
pathways that lead to the formation of magnetite.
In general, the magnetite particles produced in the
coprecipitation process are crystallized in a quasi-immediate
process at room temperature, via a rather complex mecha-
nism.
13
It is, therefore, of significant interest to increase our
understanding of the underlying mechanism, and to improve
the size distribution and crystallinity of the magnetite
nanoparticles produced. After early pioneering work by
Massart,
9
many research groups reported various phase
transformations between iron oxides. Cornell and Giovanoli
investigated the phase transformation of akaganeite into
goethite and/or hematite in alkaline media.
14
Similar results
were also observed by other research groups.
1518
Gualtieri and
Venturelli used in situ synchrotron X-ray powder diffraction to
study the transformation of goethite to hematite.
19
Hematite
magnetite
2024
and lepidocrocitemaghemite
25,26
transforma-
tions were also reported. Abou-Hassan et al. recently studied
the kinetics of the transformation using a coaxial flow
microreactor and demonstrated the importance of pH gradients
for superparamagnetic nanoparticle synthesis.
27
They also
separated the nucleation and growth process of ferrihydrite
nanoparticles.
28
However, all these studies covered only a small
Received: December 8, 2011
Revised: February 23, 2012
Published: February 23, 2012
Article
pubs.acs.org/JPCC
2012 American Chemical Society 6069 dx.doi.org/10.1021/jp211843g | J. Phys. Chem. C 2012, 116, 60696076
part of the entire transformation involving the coprecipitation
reaction of magnetite. Here, we report the phase transformation
of the intermediates of goethite (-FeOOH), akaganeite (-
FeOOH), and lepidocrocite (-FeOOH). A reaction scheme is
proposed in which Fe
2+
and Fe
3+
have separate, but interrelated,
pathways toward magnetite nanoparticles. We also report the
topotactic transformation of goethite to magnetite mediated by
arrow-shaped nanoparticles.
2. EXPERIMENTAL SECTION
2.1. Materials. Ferric chloride hexahydrate (FeCl
3
6H
2
O),
ferrous chloride tetrahydrate (FeCl
2
4H
2
O), and ammonia
solution (NH
3
(aq)) were purchased from Sigma-Aldrich. All
reagents were used as received without further purification.
Deionized water was used throughout the experiments.
2.2. Synthesis. Experiments were carried out at 25 C, in a
1 L jacketed glass reactor equipped with a mechanical stirrer, a
temperature sensor, a gas inflow port (for N
2
(g)), an exit gas
tube with a water-cooled condenser, and a port for the addition
of base solution.
29
The iron salt solution (500 g, containing 60
mmol of FeCl
3
6H
2
O and 30 mmol of FeCl
2
4H
2
O) was
transferred to an oxygen-free reactor, and a 0.8 M NH
3
solution
was added to the iron salt solution. This addition was
performed either continuously, using a peristaltic pump at a
constant speed (1.88 mL/min), or abruptly, where the entire
NH
3
solution volume was added at once. Hence, the molar
ratio of ammonia to iron ions (R = [NH
3
]/[Fe
2+
+ Fe
3+
]) was
varied with sampling intervals of 0.5 or 0.1 in R. In the abrupt
addition case, samples were removed 60 min after the addition
of the base solution (preliminary experimental results indicated
60 min was enough to complete the reaction). Samples were
centrifuged, repeatedly washed with deionized water, and then
dried at room temperature under vacuum.
2.3. Characterization. Transmission electron microscopy
(TEM) images and fast Fourier transform (FFT) patterns were
recorded using a Tecnai G2 F30 (FEI) at 300 kV, or a Tecnai
F20 (FEI) at 200 kV. A drop of the colloidal solution in
methanol was deposited onto a thin carbon film supported by a
copper TEM grid, and the solvent was then allowed to
evaporate. Field emission scanning electron microscopy (FE-
SEM) images were obtained using an S-4800 instrument
(Hitachi). Attenuated total reflectance Fourier transform
infrared (ATR-FTIR) spectra were measured using an FTS
3000 instrument (Bio-Rad). Powder X-ray diffraction (XRD)
patterns were recorded on a Rigaku D/MAX-RB diffractometer
with Cu K radiation ( = 1.5406 ), at 40 kV and 100 mA
(see Table S1, Supporting Information, for the crystal
structures referred to in this study).
3. RESULTS AND DISCUSSION
3.1. Continuous Addition of Base. An NH
3
solution was
added continuously to an iron salt solution (1:2 mol ratio
mixture of Fe
2+
and Fe
3+
) over the range of R = 04.0; the pH
of the solution varied from 1.5 to 9.0. The color of the iron
solution changed slowly from light brown, through dark brown,
and finally to black, indicating the formation of magnetite
nanoparticles (Figure S1, Supporting Information). Two pH
plateaus (which indicated that the added base was consumed by
the precipitation reaction) were observed near pH 1.5 and 5.0,
corresponding to the hydroxylations of Fe
3+
and Fe
2+
,
respectively (Figure S2, Supporting Information).
30,31
Near
pH 5, the mixed iron salt solutions crystallized the magnetite
nanoparticles without changing the pH, even though ammonia
was added such that the R value changed from 2.0 to 2.5.
3.1.1. Akaganeite Dissolves near R = 2. The evolution of
the morphology over the course of the synthetic process is
shown in Figure 1. The TEM images in Figure 1, XRD patterns
in Figure 2, and FTIR spectra in Figure S3 (Supporting
Information) indicated that akaganeite and goethite were
intermediates for magnetite. At R = 1.5, rod-shaped akaganeite
particles were precipitated with low crystallinity, as indicated by
the broad XRD peaks. With an increase in R to 2.0, the
akaganeite particles dissolved and became rounded and smaller
(23 nm), consistent with the broader XRD peaks. At R =
2.5, these small akaganeite nanoparticles evolved into a mixture
of rod-shaped goethite nanoparticles (minor component) and
Figure 1. TEM images of samples at R values of (a) 1.5, (b) 2.0, (c)
2.5, and (d) 3.0 (continuous addition of base). The black arrow in (c)
indicates a detached arrowhead part.
Figure 2. XRD patterns of samples at R values of (a) 1.5, (b) 2.0, (c)
2.5, and (d) 3.0 (continuous addition of base): (pink cross) akaganeite
(JCPDS 34-1266), (blue square) goethite (JCPDS 29-0713), (black
diamond) magnetite (JCPDS 19-0629). Subscript: G, goethite.
The Journal of Physical Chemistry C Article
dx.doi.org/10.1021/jp211843g | J. Phys. Chem. C 2012, 116, 60696076 6070
rounded magnetite nanoparticles (main component). At R >
3.0, pure rounded magnetite nanoparticles were obtained. Near
pH 5 (2.0 R 2.5), our results indicated that akaganeite
dissolved, followed by the crystallization of goethite and
magnetite.
Because such a dramatic and complex transition occurred
between R = 2.0 and 2.5, we took samples at narrow intervals
(intervals of 0.1 R) in this range and characterized the crystal
structures (Figures 3, 4, and S3, spectrum b (Supporting
Information)). At R = 2.1, nearly pure goethite nanorods were
observed as an intermediate phase. As R was increased to 2.5,
these goethite nanoparticles gradually disappeared, and the
proportion of magnetite nanoparticles simultaneously in-
creased. These results indicated that the overall variation of
the crystal structures during the continuous addition of the base
solution could be expressed as akaganeite goethite
magnetite.
3.1.2. Oriented Aggregation to Goethite. Rod-shaped
akaganeite nanoparticles grown along the [001] direction are
produced by the hydrolysis of FeCl
3
solution at pH 12
([OH]/[Fe
3+
] < 2.7).
32,33
The akaganeite nanoparticles take up
protons in acidic solution. The akaganeite structure is stabilized
by the incorporation of a stoichiometric amount of Cl

ions,
but these are easily replaced by OH

ions at higher pH. Such


ion replacement may induce the phase transformation of
akaganeite to goethite, which can be described as a
dissolutionrecrystallization process.
1416
Recent studies indicate that, when the structure of akaganeite
nanoparticles collapses and their particle size becomes
comparable to a few unit cells, akaganeite fragments may
have a goethite-like crystalline structure.
17,18
Our results also
demonstrate that, as chloride ions were replaced by hydroxide
ions and the akaganeite dissolved, the crystalline structure of
the small (23 nm) fragments of akaganeite became nearly
indistinguishable from that of goethite (Figure 5a,b, Table S2
(Supporting Information)). The HRTEM image and the
corresponding FFT pattern (Figure 5b) of the fragment show
the akaganeite structure, confirming the presence of (211) and
(330) planes, which are very similar to the (021) and (040)
planes of goethite, respectively (Table S2, Supporting
Information). Neighboring akaganeite fragments may have
experienced an oriented aggregation due to a thermodynamic
driving force working to reduce the surface energy (Figure
5c,d).
3436
Akaganeite fragments with high OH

replacement
could, therefore, be used as building blocks for goethite
structures with only slight rearrangements; thus, the species
dissolved at high pH may also have recrystallized on the blocks,
making the oriented aggregates more perfect. This crystal
growth by oriented aggregation often produced small
misorientations at the interface between the blocks in the
goethite nanorods (Figure 5c,d).
37
Therefore, the evolution of
akaganeite to goethite proceeded via a fragmentation by
Figure 3. TEM images of samples at R values of (a) 2.1, (b) 2.2, (c)
2.3, and (d) 2.4 (continuous addition of base). The black arrow in (d)
indicates a detached arrowhead part.
Figure 4. XRD patterns of samples at R values of (a) 2.1, (b) 2.2, (c)
2.3, and (d) 2.4 (continuous addition of base): (blue square) goethite,
(black diamond) magnetite.
Figure 5. TEM and HRTEM images of samples at R values of (a, b)
2.0 and (c, d) 2.1 (continuous addition of base). (b) [111]
A
zone axis
HRTEM image obtained from the rectangular area in (a). (c) The
lattice fringe spacings of 0.31 nm corresponded to 2 times the (160)
G
and (160)
G
planes. (d) The lattice fringe spacings of 0.27 nm
corresponded to the (130)
G
plane (black arrows indicate junctions).
Subscripts: A, akaganeite; G, goethite.
The Journal of Physical Chemistry C Article
dx.doi.org/10.1021/jp211843g | J. Phys. Chem. C 2012, 116, 60696076 6071
dissolution of akaganeiteoriented aggregation to goethite
recrystallization process.
3.1.3. Topotactic Transformation. TEM and SEM images
taken during the transformation from goethite to magnetite
(2.1 R 2.5) show the diverse crystal habits (Figures 6 and
7, and S4 and S5 (Supporting Information)). Of particular note
were the arrow shapes, which formed parallelograms with
identical angles. For these arrow-shaped nanoparticles, a
specific crystallographic relationship would exist between
goethite (orthorhombic) and hematite (hexagonal), or between
hematite and magnetite (cubic).
The formation of arrow-shaped nanoparticles can be
explained by the following possible mechanisms. The structure
of goethite and hematite can be described as consisting of
slightly distorted hcp arrays of anions (O
2
, OH

) stacked
along the [100]
G
direction and the [001]
H
direction, for
goethite and hematite, respectively.
38
In addition, the b axis of
goethite is 2 times the a axis of hematite and the c axis of
hematite is 3 times the a axis of goethite (Table S1, Supporting
Information).
16,19
Since the structure of hematite can also be
expressed using rhombohedral axes,
39
the crystallographic
relationships among the three kinds of crystal axes can be
depicted schematically, as shown in Figure 6a (see Figure S6,
Supporting Information, for detailed crystal models). On the
basis of these relationships, the formation of the arrow-shaped
nanoparticles shown in Figure 6b would be possible if the
goethite nanorods grew along the [100]
G
direction (Figure
5c,d) and hematite grew on the surface of the tips of the
goethite nanorods.
Magnetite has ccp arrays of oxygen anions along the [111]
Mt
,
which is parallel to [001]
H
, and that [110]
Mt
is parallel to
[110]
H
in the case of the topotactic transformation of hematite
to magnetite (Figure S6, Supporting Information).
20,21
In that
case, the hcp oxygen packing of hematite can play a role as a
template for the ccp oxygen packing of magnetite.
22,23,40
Thus,
magnetite can also develop arrowheads similar to the hematite
shape from two imaginary octahedra of magnetite (Figure 6c).
TEM and HRTEM images show the arrow-shaped nano-
particles, which have magnetite lattice planes in the arrowhead
region and goethite lattice planes in the arrow shaft region
(Figures 6 and 7). Therefore, the transformations of goethite
through hematite to magnetite proceeded via a topotactic
process based on the above structural relationships.
The topotactic transformation of goethite to hematite
involves dehydration and local rearrangement processes,
19,41,42
which are accelerated by the presence of a small amount of
Fe
2+
, which enables electron hopping between Fe
2+
and
Fe
3+
.
4346
As the transformation proceeds, the destruction of
the crystal structure of goethite nanorods progresses gradually
due to the dehydration and rearrangement processes.
Conversely, the crystal structure of magnetite forms in the
arrowhead region. As a result, the arrow-shaped nanoparticles,
which show magnetite lattice fringes, have an obscure goethite
crystalline structure, while the arrow-shaped nanoparticles,
which display obvious goethite lattice fringes, possess
ambiguous arrowhead shapes and an unclear magnetite
crystalline structure (Figures 6 and 7). This destruction of
the goethite structure explains the absence of double-arrow-
shaped nanoparticles.
3.1.4. Adsorption of Fe
2+
Ions on Maghemite. Although
hematite is more stable than maghemite in the bulk state, the
thermodynamic stability of maghemite becomes comparable to
Figure 6. (a) Crystallographic relationship between goethite and
hematite. (b) Schematic drawing of an arrow-shaped nanoparticle
consisting of a nanorod of goethite and a rhombohedron of hematite
with a [110]
H
zone axis. (c) Schematic drawing of an arrow-shaped
nanoparticle consisting of a nanorod of goethite and two halves of two
imaginary octahedra of magnetite with a [110]
Mt
zone axis. (d) TEM
image of an arrow-shaped nanoparticle at R = 2.2. (e) [110]
Mt
zone
axis HRTEM image obtained from the rectangular area in (d). The
lattice fringe spacings of 0.42 and 0.18 nm corresponded to 2 times the
(004)
Mt
and (664)
Mt
planes, respectively. Blue, goethite; red, hematite
(rhombohedral); black, hematite (hexagonal); and green, magnetite.
Subscripts: G, goethite; H, hematite; Mt, magnetite.
Figure 7. HRTEM images obtained from the rectangular area in each
inset. (a) The lattice fringe spacing of 0.42 nm corresponded to 2
times the (400)
Mt
plane (R = 2.2). (b) The lattice fringe spacings of
0.225 and 0.97 nm corresponded to the (121)
G
and 2 times the
(111)
Mt
planes, respectively (R = 2.2). (c) The lattice fringe spacing of
0.225 nm corresponded to the (121)
G
plane (R = 2.1). (d) The lattice
fringe spacings of 0.495 and 0.254 nm corresponded to the (020)
G
and
(101)
G
planes, respectively (R = 2.3).
The Journal of Physical Chemistry C Article
dx.doi.org/10.1021/jp211843g | J. Phys. Chem. C 2012, 116, 60696076 6072
that of hematite at the nanoscale, due to the high surface energy
of hematite. Consequently, hematite nanoparticles may alter
their structure and may possess a maghemite-like structure
(inverse spinel structure) near the surface; this tendency is
amplified as the particle size decreases.
18,40,47
The structure of
magnetite resembles that of maghemite closely enough that
they can form a continuous solid solution with each other,
18
and the adsorption of Fe
2+
ions on the maghemite surface
induces the conversion of maghemite to magnetite by raising
R.
13,24
In fact, several driving forces may have combined to cause
phase transformations at this stage (pH 5). Increases in R,
the thermodynamic stability of hematite and maghemite, and
the adsorption of Fe
2+
gave rise to the transformations from
goethite to hematite,
15,48
hematite to maghemite, and
maghemite to magnetite, respectively. Moreover, the presence
of Fe
2+
in the reaction medium catalyzed all these phase
transformations.
49
The conversion of goethite to magnetite
should pass through the hematite and maghemite phases from a
crystallographic viewpoint. Therefore, hematite and maghemite
may be transient phases in the transformation of goethite into
magnetite. The angles of the arrowheads shown in Figures S4
and S5 (Supporting Information) were more similar to those of
hematite than magnetite (Figure 6b,c), but the lattice planes
were similar to those of magnetite rather than hematite (albeit
imperfectly similar) (Figures 6e and 7). This suggests that the
arrowhead structure was an intermediate phase between
goethite and magnetite that was not observed as distinct
peaks of hematite or maghemite in the XRD patterns (Figure
4). The transformation to magnetite occurred almost
simultaneously with the generation of hematite on the surface
of the goethite nanoparticles. We used ex situ characterization
methods in this study, which may have hindered the ability to
detect the intermediate phases. Further studies of phase
transformations during coprecipitation reactions by in situ
analyses will be implemented for stronger support of the
proposed formation pathways.
3.1.5. Magnetite Seed. The intermediate character of the
arrowhead regions implies that the arrowhead sections
detached from the goethite before, or soon after, it formed
the magnetite crystal structures. Other work has suggested that
dehydration and iron migration may induce void structures in
goethite, which subsequently collapse at the end of the reaction,
during the topotactic transformation to hematite.
19,50
The
topotactic transformations between hematite and magnetite
may trigger dissolution and fracturing along the interphase
boundaries, due to the induced internal stress resulting from
the crystal structural difference between two phases, which
finally leads to separation.
22,23
These voids and fractures also
occurred in our experiments, and the magnetite-like structures
eventually detached from the goethite nanoparticles as nuclei
(Figures 1c, 3d, and 8). Once detached, the separated
arrowheads acted as seeds for magnetite growth and underwent
oriented aggregation, in a similar fashion as in the phase
transformation of akaganeite to goethite. Both a slight
misorientation between the primary particles and an overall
parallelism of the lattice fringes (which are usually found in
crystals grown via the oriented aggregation mechanism)
37
were
observed in the magnetite nanoparticles at R = 3.0 (Figure 9).
The goethite nanorods of the arrow-shaped nanoparticles
finally collapsed, as explained above, and then provided the Fe
3+
ions for the growth reaction of the oriented aggregates with the
magnetite-like structure. Meanwhile, the Fe
2+
ions, which were
initially added to the reaction medium but had not precipitated,
were used for the growth reaction. Therefore, the trans-
formation of goethite to magnetite occurred through a
topotactic transformationdetachment produced by internal
stressoriented aggregationgrowth of the oriented aggre-
gates process.
3.2. Abrupt Addition of Base (R < 2.67). It was
determined that the transformation (taking place with changes
from low pH to high pH) occurred as akaganeite goethite
(hematite maghemite) magnetite when the base was
added continuously. In conventional batch-operation copreci-
pitation reactions, however, large amounts of base solution are
abruptly (all at once) added to the iron salt solution. The
intermediates produced in the course of the reaction could be
different when using the two different methods, because the
abrupt addition of base solution can intensify the inhomoge-
neity of the pH in the reaction medium before complete mixing
and can alter the reaction rate and reaction pathway. It is,
therefore, important to control the process of mixing the iron
salt solution and the base solution.
3.2.1. Metastable Lepidocrocite from Ferrous Hydroxide.
In the range of R 2.02.2, we observed lepidocrocite,
consisting of layered iron(III) oxide octahedra bonded by
hydrogen bonding via hydroxide layers. Interestingly, no
lepidocrocite was observed in the continuous addition
experiments. A large proportion of the particles were tabular
lepidocrocite, with small amounts of goethite at R = 2.0. As R
increased to 2.2, the proportion of lepidocrocite decreased
continuously, and the proportion of goethite and magnetite
increased. Finally, at R = 2.6, lepidocrocite totally disappeared
and magnetite became the main phase, accompanied by
goethite as a minor phase (Figures 10, 11, and S3, spectrum
c (Supporting Information)).
Lepidocrocite is produced by the oxidation of ferrous
hydroxide (Fe(OH)
2
) under slightly acidic conditions (pH
57), and ferrous hydroxide is formed at pH > 67.
13,51
The
abrupt addition of base, as indicated previously, may have
Figure 8. (a, b) TEM images of goethite showing the void
morphology. (c) TEM image showing the detached arrowhead parts.
Figure 9. [111]
Mt
zone axis HRTEM images of oriented aggregates of
magnetite nanoparticles. The lattice fringe spacings of 0.3 nm
corresponded to {220}
Mt
planes.
The Journal of Physical Chemistry C Article
dx.doi.org/10.1021/jp211843g | J. Phys. Chem. C 2012, 116, 60696076 6073
created incomplete mixing locally, leading to a local pH
gradient in the reaction medium before it could homogeneously
mix and reach pH 5. It is likely that, in the local high-pH region,
ferrous hydroxide was formed first and then (as the mixing
proceeded) underwent a transformation to lepidocrocite via the
substitution of Fe
2+
(in the structure) with Fe
3+
(in solution),
or via the oxidation of Fe
2+
ions.
The fact that goethite was formed as a main phase at R =
2.2even under the abrupt addition conditions (Figure 11,
pattern c)indicated that the akaganeite goethite
(hematite maghemite) magnetite route observed in the
continuous addition method was still being followed. The
overall direction of the pH change in the abrupt addition
method was also from pH 1.5 to pH 5. Consequently, as R
increased, the formation of goethite increased, and the
formation and size of lepidocrocite, therefore, decreased, due
to the lack of available Fe
3+
(which was consumed for goethite).
3.2.2. Topotactic Transformation. Magnetite can be formed
not only from goethite but also from lepidocrocite. The
structure of lepidocrocite consists of ccp arrays of anions
stacked along the [150] direction, which corresponds to the
[111] direction of inverse spinel structures (maghemite and
magnetite).
52
There were large quantities (half the amount of
Fe
3+
) of Fe
2+
in the reaction solution, enough for lepidocrocite
to transform into magnetite through a topotactic process.
11
Magnetite with a sheet structure was observed at R = 2.1, which
may have originated from lepidocrocite (Figure 10d).
We now consider the possible reasons for the observation
that lepidocrocite was present and maghemite was absent at R
= 2.0. The presence of lepidocrocite implied that some regions
of the reaction medium had a pH high enough to form ferrous
hydroxide, where Fe
3+
naturally precipitates. Since we dissolved
Fe
2+
and Fe
3+
in a 1:2 molar ratio, magnetite would form in that
high-pH region if the direct reaction between Fe
2+
and Fe
3+
occurred in the aqueous phase. The magnetite phase should
then have transformed into maghemite as the pH decreased
with further solution mixing.
13
At R = 2.0, however, large
amounts of lepidocrocite and an absence of maghemite were
observed. Therefore, magnetite did not form by the direct
reaction of Fe
2+
and Fe
3+
in the aqueous phase, but via the
phase transformation of iron oxyhydroxides (goethite and
lepidocrocite).
3.3. Conventional Coprecipitation Reaction (Abrupt
Addition of Base, R 2.67). From the results described
above, we propose a complete reaction scheme for the
coprecipitation of magnetite (see Figure 12). In the conven-
tional coprecipitation process, the iron salt solution and the
base solution are mixed abruptly and quickly form a contact
interface between the iron-rich solution (pH 1.5) and the
base-rich solution (pH 11). In the iron-rich solution, the base
diffuses across the contact interface, while iron ions diffuse in
the opposite direction. Finally, the contact interface will be
dispersed in a precipitation event. Although the interface will be
maintained for a short period, the low-pH iron-rich side will
accept the base, which will increase the pH and react with the
less-stable Fe
3+
ions to form akaganeite, while the high-pH
base-rich side will incorporate not only Fe
3+
but also Fe
2+
ionic
diffusion to form the ferrous hydroxide. Therefore, two
different nucleation processes will be initiated as the solutions
begin to mix, and both synthetic routes (the akaganeite
goethite (hematite maghemite) magnetite route and
the ferrous hydroxide lepidocrocite (maghemite)
magnetite route) will be followed to create the magnetite
structure.
Figure 10. TEM images of samples at R values of (a) 2.0, (b, d) 2.1,
and (c) 2.2 (abrupt addition of base). (d) [112]
Mt
zone axis HRTEM
image obtained from the rectangular area in (b).
Figure 11. XRD patterns of samples at R values of (a) 2.0, (b) 2.1, (c)
2.2, and (d) 2.6 (abrupt addition of base): (red circle) lepidocrocite
(JCPDS 08-0098), (blue square) goethite, (black diamond) magnetite.
Figure 12. Formation pathways of magnetite nanoparticles by
coprecipitation method. Main intermediate phases are shown in
yellow areas.
The Journal of Physical Chemistry C Article
dx.doi.org/10.1021/jp211843g | J. Phys. Chem. C 2012, 116, 60696076 6074
When a large quantity (R 2.67) of base solution is mixed
abruptly with the iron solution, the pH may increase at a more
rapid rate compared with continuous addition of base. The
growth of akaganeite may then be suppressed, the formation of
goethite may be strengthened, and the formation of
lepidocrocite may be weakened, due to the lack of Fe
3+
. In
addition, the hematite and maghemite may be part of a
transient phase. Therefore, goethite and ferrous hydroxide
would be the major intermediates controlling the phase
transformation and particle growth.
4. CONCLUSIONS
The present results showed that magnetite nanoparticles were
formed in the coprecipitation process by the phase trans-
formation of iron oxyhydroxides, rather than the direct reaction
of Fe
2+
and Fe
3+
in the aqueous phase. This was detected using
X-ray diffraction, electron microscopy, and FTIR spectroscopy.
Akaganeite nucleated and transformed to goethite, which
underwent a topotactic transformation to magnetite as the
pH of the iron salt solution increased slowly and continuously.
Simultaneously with the above reaction route, an additional
reaction route was followed: ferrous hydroxide nucleated and
transformed to lepidocrocite, which underwent a topotactic
transformation to magnetite as the iron salt solution and the
base solution mixed quickly and abruptly. Especially, the
topotactic transformation of goethite to magnetite was
mediated by the unprecedented arrow-shaped nanoparticles.
These phase transformations were consistent with the specific
crystallographic relationships among the iron oxide phases. The
growth of goethite and magnetite proceeded through an
oriented aggregation mechanism.
In coprecipitation reactions, the formation of magnetite
proceeds via the reaction pathways described above, which are
complex and interrelated. This may explain the wide particle
size distribution and low crystallinity of magnetite nanoparticles
prepared using the coprecipitation process. The findings of this
study indicate the importance of the mixing process (i.e., the
mixing of the iron salt solution and the base solution) in
determining the coprecipitation reaction route and provide the
foundations for a more detailed study of the reaction
mechanism.

ASSOCIATED CONTENT
*S Supporting Information
Crystal structures, color photograph of reaction solution, pH
curve, FTIR spectra, further details in Figure 5b, TEM and
SEM images of arrow-shaped nanoparticles, crystal models, and
magnetic attraction. This material is available free of charge via
the Internet at http://pubs.acs.org.

AUTHOR INFORMATION
Corresponding Author
*E-mail: kjd@kaist.ac.kr.
Notes
The authors declare no competing financial interest.

ACKNOWLEDGMENTS
This research was supported by the Basic Science Research
Program, through the National Research Foundation of Korea
(NRF), funded by the Ministry of Education, Science and
Technology (2009-0076882). We thank Hyung Bin Bae in the
KAIST Research Supporting Team for TEM measurements.

REFERENCES
(1) Qiao, R.; Yang, C.; Gao, M. J. Mater. Chem. 2009, 19, 6274.
(2) Yang, H.-M.; Lee, H. J.; Jang, K.-S.; Park, C. W.; Yang, H. W.;
Heo, W. D.; Kim, J.-D. J. Mater. Chem. 2009, 19, 4566.
(3) Yang, H.-M.; Lee, H. J.; Park, C. W.; Yoon, S. R.; Lim, S.; Jung, B.
H.; Kim, J.-D. Chem. Commun. 2011, 47, 5322.
(4) Lu, A.-H.; Salabas, E. L.; Schuth, F. Angew. Chem., Int. Ed. 2007,
46, 1222.
(5) Laurent, S.; Forge, D.; Port, M.; Roch, A.; Robic, C.; Elst, L. V.;
Muller, R. N. Chem. Rev. 2008, 108, 2064.
(6) Park, J.; An, K.; Hwang, Y.; Park, J.-G.; Noh, H.-J.; Kim, J.-Y.;
Park, J.-H.; Hwang, N.-M.; Hyeon, T. Nat. Mater. 2004, 3, 891.
(7) Baier, J.; Koetz, J.; Kosmella, S.; Tiersch, B.; Rehage, H. J. Phys.
Chem. B 2007, 111, 8612.
(8) Taniguchi, T.; Nakagawa, K.; Watanabe, T.; Matsushita, N.;
Yoshimura, M. J. Phys. Chem. C 2009, 113, 839.
(9) Massart, R. IEEE Trans. Magn. 1981, 17, 1247.
(10) Tronc, E.; Jolivet, J. P.; Massart, R. Mater. Res. Bull. 1982, 17,
1365.
(11) Neveu-prin, S.; Cabuil, V.; Massart, R.; Escaffre, P.; Dussaud, J.
J. Magn. Magn. Mater. 1993, 122, 42.
(12) Massart, R.; Dubois, E.; Cabuil, V.; Hasmonay, E. J. Magn. Magn.
Mater. 1995, 149, 1.
(13) Jolivet, J.-P.; Chaneac, C.; Tronc, E. Chem. Commun. 2004, 481.
(14) Cornell, R. M.; Giovanoli, R. Clays Clay Miner. 1990, 38, 469.
(15) Cai, J.; Liu, J.; Gao, Z.; Navrotsky, A.; Suib, S. L. Chem. Mater.
2001, 13, 4595.
(16) Cornell, R. M.; Schwertmann, U. The Iron Oxides: Structure,
Properties, Reactions, Occurrences and Uses, 1st ed.; VCH: Weinheim,
Germany, 1996.
(17) Deore, S. W. A. Thermodynamic and Structural Study of
Atomistic, Nano and Bulk Systems. Ph.D. Thesis, University of
California, Davis, CA, 2007.
(18) Navrotsky, A.; Mazeina, L.; Majzlan, J. Science 2008, 319, 1635.
(19) Gualtieri, A. F.; Venturelli, P. Am. Mineral. 1999, 84, 895.
(20) Chueh, Y.-L.; Lai, M.-W.; Liang, J.-Q.; Chou, L.-J.; Wang, Z. L.
Adv. Funct. Mater. 2006, 16, 2243.
(21) Watanabe, Y.; Takemura, S.; Kashiwaya, Y.; Ishii, K. J. Phys. D:
Appl. Phys. 1996, 29, 8.
(22) Barbosa, P. F.; Lagoeiro, L. Am. Mineral. 2010, 95, 118.
(23) Lagoeiro, L. E. J. Metamorph. Geol. 1998, 16, 415.
(24) Otake, T.; Wesolowski, D. J.; Anovitz, L. M.; Allard, L. F.;
Ohmoto, H. Earth Planet. Sci. Lett. 2007, 257, 60.
(25) Fang, J.; Kumbhar, A.; Zhou, W. L.; Stokes, K. L. Mater. Res.
Bull. 2003, 38, 461.
(26) Sudakar, C.; Subbanna, G. N.; Kutty, T. R. N. J. Phys. Chem.
Solids 2003, 64, 2337.
(27) Abou-Hassan, A.; Dufreche, J.-F.; Sandre, O.; Meriguet, G.;
Bernard, O.; Cabuil, V. J. Phys. Chem. C 2009, 113, 18097.
(28) Abou-Hassan, A.; Sandre, O.; Neveu, S.; Cabuil, V. Angew.
Chem., Int. Ed. 2009, 48, 2342.
(29) Ahmed, I. A. M.; Benning, L. G.; Kakonyi, G.; Sumoondur, A.
D.; Terrill, N. J.; Shaw, S. Langmuir 2010, 26, 6593.
(30) Domingo, C.; Rodriguez-clemente, R.; Blesa, M. A. Solid State
Ionics 1993, 59, 187.
(31) Salazar, J. S.; Perez, L.; de Abril, O.; Phuoc, L. T.; Ihiawakrim,
D.; Vazquez, M.; Greneche, J.-M.; Begin-Colin, S.; Pourroy, G. Chem.
Mater. 2011, 23, 1379.
(32) Bottero, J.-Y.; Manceau, A.; Villieras, F.; Tchoubar, D. Langmuir
1994, 10, 316.
(33) Atkinson, R. J.; Posner, A. M.; Quirk, J. P. Clays Clay Miner.
1977, 25, 49.
(34) Penn, R. L.; Banfield, J. F. Science 1998, 281, 969.
(35) Banfield, J. F.; Welch, S. A.; Zhang, H. Z.; Ebert, T. T.; Penn, R.
L. Science 2000, 289, 751.
(36) Yuwono, V. M.; Burrows, N. D.; Soltis, J. A.; Penn, R. L. J. Am.
Chem. Soc. 2010, 132, 2163.
(37) Chan, C. S.; De Stasio, G.; Welch, S. A.; Girasole, M.; Frazer, B.
H.; Nesterova, M. V.; Fakra, S.; Banfield, J. F. Science 2004, 303, 1656.
The Journal of Physical Chemistry C Article
dx.doi.org/10.1021/jp211843g | J. Phys. Chem. C 2012, 116, 60696076 6075
(38) Yang, H.; Lu, R.; Downs, R. T.; Costin, G. Acta Crystallogr., Sect.
E: Struct. Rep. Online 2006, 62, i250.
(39) Gonzalez, G.; Sagarzazu, A.; Villalba, R. Mater. Res. Bull. 2000,
35, 2295.
(40) Chernyshova, I. V.; Hochella, M. F.; Madden, A. S. Phys. Chem.
Chem. Phys. 2007, 9, 1736.
(41) Ruan, H. D.; Frost, R. L.; Kloprogge, J. T.; Duong, L.
Spectrochim. Acta, Part A 2002, 58, 967.
(42) Rebolledo, A. F.; Laurent, S.; Calero, M.; Villanueva, A.; Knobel,
M.; Marco, J. F.; Tartaj, P. ACS Nano 2010, 4, 2095.
(43) Pedersen, H. D.; Postma, D.; Jakobsen, R.; Larsen, O. Geochim.
Cosmochim. Acta 2005, 69, 3967.
(44) Williams, A. G. B.; Scherer, M. M. Environ. Sci. Technol. 2004,
38, 4782.
(45) Schaefer, M. V.; Gorski, C. A.; Scherer, M. M. Environ. Sci.
Technol. 2011, 45, 540.
(46) Yang, L.; Steefel, C. I.; Marcus, M. A.; Bargar, J. R. Environ. Sci.
Technol. 2010, 44, 5469.
(47) Randrianantoandro, N.; Mercier, A. M.; Hervieu, M.; Greneche,
J. M. Mater. Lett. 2001, 47, 150.
(48) Das, S.; Hendry, M. J.; Essilfie-Dughan, J. Environ. Sci. Technol.
2011, 45, 268.
(49) Rosso, K. M.; Yanina, S. V.; Gorski, C. A.; Larese-Casanova, P.;
Scherer, M. M. Environ. Sci. Technol. 2010, 44, 61.
(50) Watari, F.; Delavignette, P.; Van Landuyt, J.; Amelinckx, S. J.
Solid State Chem. 1983, 48, 49.
(51) Cornell, R. M.; Giovanoli, R. Clays Clay Miner. 1988, 36, 385.
(52) Fasiska, E. J. Corros. Sci. 1967, 7, 833.
The Journal of Physical Chemistry C Article
dx.doi.org/10.1021/jp211843g | J. Phys. Chem. C 2012, 116, 60696076 6076

Vous aimerez peut-être aussi