Crystengcomm: Paper

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

CrystEngComm

This article is licensed under a Creative Commons Attribution-NonCommercial 3.0 Unported Licence.

View Article Online


PAPER View Journal | View Issue

Oxidation of magnetite nanoparticles: impact on


surface and crystal properties†
Open Access Article. Published on 02 December 2016. Downloaded on 7/24/2019 12:21:42 PM.

Cite this: CrystEngComm, 2017, 19,


246 S. P. Schwaminger,a D. Bauer,a P. Fraga-García,a F. E. Wagnerb and S. Berensmeier*a

Iron oxide nanoparticles are of great scientific interest due to their huge versatility of applications. The oxi-
dation process of magnetite to maghemite is difficult to monitor as both iron oxide polymorphs possess
connatural chemical properties. Especially the surface composition and reactivity of these nanosystems,
which are most relevant for interactions with their environment, are not completely understood. Here, the
oxidation of magnetite is investigated under mild and harsh conditions in order to understand the oxidation
behaviour and the chemical stability of transition forms. Therefore, the oxidation process, is investigated
with Raman, Mössbauer and X-ray photoelectron spectroscopy as well as X-ray diffraction and magnetom-
etry. The multi-analytical approach allows new insights into surface composition and rearrangement
according to respective different depth profiles. For both conditions investigated, the ferrous iron compo-
Received 21st November 2016, nents are oxidised prior to structural changes in the Fe–O vibrations and crystal structure. The process
Accepted 2nd December 2016
starts from the outer layers and is acid catalysed. Oxidation leads to a decrease of magnetisation which still
remains higher than 54 emu g−1. The charge and surface reactivity can be affected by the different oxida-
DOI: 10.1039/c6ce02421a
tion methods and the irreversible adsorption of acid molecules. Biocompatibility and catalytic properties of
www.rsc.org/crystengcomm iron oxide nanoparticles open doors to future applications.

Introduction imaging5,18–20 are possible for MNP as they have been ap-
proved by the US Food and Drug Administration (FDA).21 On
Magnetic nanoparticles (MNP) have been intensively studied the other hand, for catalytic applications, such as Fenton
and optimised for many different scopes of application over chemistry, a high reactivity of MNP is desired.22 While most
the past decade. Most MNPs applied industrially consist of investigations focus on the modification and
the superparamagnetic iron oxide nanoparticles magnetite functionalisation of MNP,23–27 in depth fundamental under-
(Fe3O4) and maghemite (γ-Fe2O3). The versatility of possible standing of the formation, phase transition and factors
applications reaches from adsorbents in separation pro- influencing the surface chemistry is still required. Many inter-
cesses,1,2 enzyme immobilisation,3 biomedicine4,5 and cataly- esting studies focusing on the formation pathways of iron ox-
sis6 to energy storage.7–9 In mineral processing and wastewa- ide nanoparticles exist for different routes such as polyol,28,29
ter treatment10 MNP are used as low-cost adsorbents which hydrothermal,30 electrochemical,31 pyrolytic32 and co-
can be magnetically separated and aid in removing toxic11 or precipitative synthesis.33,34 The influence of different synthe-
radioactive ions.12 For biological separation processes such as sis parameters, such as temperature, pH and iron ion concen-
protein or DNA/RNA purification, cell sorting, algae tration, on the size, morphology and composition of the iron
harvesting and virus removal MNP should be chemically and oxide nanoparticles are widely discussed in literature.35–39
mechanically inert toward the target products.13 Additionally, However, as numerous applications require the interaction
great separation capabilities and affinities to biomolecules with different ions, biomolecules or polymers, the control of
are requested for bioseparation engineering.14,15 However, the surface reactivity is an important topic.27 Especially fer-
even biomedical in vivo applications such as hypothermia,16 rous ions occurring in magnetite are known to have an im-
drug delivery17 and as contrast agents for magnetic resonance pact on cytotoxicity40,41 for microorganisms and on catalytic
activity towards organic substrates.22 Aruoja et al. investigated
the cytotoxicity of different metal oxide nanoparticles on
a
Bioseparation Engineering Group, Technical University of Munich, bacteriae, algae and protozoa and observed a cell growth inhi-
Boltzmannstraße 15, Garching, 85748, Germany. E-mail: [email protected]
b
bition for magnetite nanoparticles.42 While magnetite seems
Physik-Department E15, Technical University of Munich, James-Franck-Straße 1,
Garching, 85748, Germany
to demonstrate toxic behaviour towards biological organisms,
† Electronic supplementary information (ESI) available. See DOI: 10.1039/ no cell damaging effect could be evidenced for
c6ce02421a maghemite.40,41 This cell-damaging effect is often ascribed to

246 | CrystEngComm, 2017, 19, 246–255 This journal is © The Royal Society of Chemistry 2017
View Article Online

CrystEngComm Paper

the generation of reactive oxygen species. Such radicals can with approximately same size. The structure is examined by
be formed under acidic conditions, especially if H2O2 is pres- Mössbauer and Raman spectroscopy as well as X-ray diffrac-
ent and the Fe2+ ions of the magnetite particles can be used tion. While distinguishing between magnetite and
as Fenton catalysts (Scheme 1).22,43 maghemite from XRD is quite difficult,46 in Mössbauer
This article is licensed under a Creative Commons Attribution-NonCommercial 3.0 Unported Licence.

Iron ions are known to crystallise in more than 20 oxide/ spectroscopy, divalent and trivalent iron ions can easily be
(oxy)hydroxide forms. Different reaction conditions such as distinguished due to different isomer shifts.47 Bulk materials
pH, temperature, precursor and atmosphere can influence are often measured at room temperature where the electron
the entire crystal structure beside particle size and hopping between octahedral positions leads to an effective
magnetisation. Moreover, the different materials are also Fe2.5+ component, which is characteristic for magnetite48 and
does not appear in maghemite.47 Since thermal effects such
Open Access Article. Published on 02 December 2016. Downloaded on 7/24/2019 12:21:42 PM.

known to be transformed into each other, which occurs faster


in nanoparticles than in the bulk.44 Therefore, many transi- as Brown and Néel relaxation disturb the magnetic splitting
tion states and especially the understanding of the transition at room temperature for nanoparticles,49 a differentiation be-
on the nanoparticle level remain unknown. Particularly, the tween maghemite and magnetite nanoparticles with diame-
divalent iron ions on tetrahedral sites in the surface of mag- ters below 20 nm by Mössbauer spectroscopy is only possible
netite nanoparticles tend to oxidise very fast under ambient for measurements in strong magnetic fields (≥5 T) or low
conditions changing the composition and properties. This temperatures (<120 K).50 Below the Verwey transition at
sensitivity to oxygen often limits the application as distinct ∼120 K, a distinction between tetrahedral and octahedral po-
magnetic properties and particle size can be altered.28 Hence, sitions in Mössbauer spectra becomes difficult while the diva-
it is not only the synthesis conditions that restrict the appli- lent iron components show significantly smaller magnetic hy-
cations of particles but also the environment where the parti- perfine fields and larger isomer shifts than trivalent iron
cles are used. Furthermore, with regard to certain applica- ions. Besides Mössbauer spectroscopy, this investigation in-
tions, the surface of particles can be tuned with mild cludes Raman spectroscopy results to further demonstrate
oxidation of the particle surface leading to thermodynami- the distinction between magnetite and maghemite.51 With
cally more stable particles and different structure–property Raman spectroscopy, magnetite and maghemite nano-
relationships.45 The aim of our investigation is to monitor particles can even be accurately differentiated at room tem-
the oxidation of magnetite nanoparticles under mild (ele- perature. Different phonon modes exist for the inverse spinel
vated temperature and ambient atmosphere) and harsh (0.07 configuration of magnetite and the defect spinel structure of
mol L−1 HNO3) oxidation conditions. Therefore, we directly maghemite which correspond to different Raman shifts.52,53
investigate the structure–property relationships of the oxida- However, many reports indicate a possible oxidation of iron
tion process concerning magnetic properties, size and zeta oxides exposed to laser beams under ambient atmosphere
potential. Furthermore, we analyse the changes in catalytic which has to be considered.54 Another spectroscopic ap-
and cytotoxic behaviour of magnetite and oxidised particles proach to distinguish between magnetite and maghemite is
X-ray photoelectron spectroscopy (XPS). The shifts in core
electron orbitals can indicate a difference in the oxidation
state of iron ions.55 Although, this technique is not the best
indicator for the oxidation state in nanoparticles as the
electron mean free path is limited and not the whole particle
is demonstrated equally in the spectrum compared to a pene-
trating technique such as Mössbauer spectroscopy.

Experimental
Magnetite was synthesised as described by Roth et al.39 The
synthesis was carried out by the co-precipitation of Fe2+ and
Fe3+ aqueous salt solutions in an alkaline environment as pre-
viously reported. Ferric chloride (FeCl3·6H2O) and sodium hy-
droxide (NaOH) were purchased from AppliChem GmbH, Ger-
many. Ferrous chloride (FeCl2·4H2O) was purchased from
Bernd Kraft GmbH, Germany. 200 mL of ferrous chloride (100
mmol), ferric chloride (200 mmol) and 500 mL sodium hy-
droxide (1 mol) were prepared with degassed and deionised
water. The co-precipitation of magnetite nanoparticles was
performed in a stirred tank reactor under nitrogen atmo-
Scheme 1 Schematic illustration of different applications, structural
sphere to prevent oxidation of the precursors and the product.
transformations and surface interactions of nanoscale iron oxide The suspension was washed several times with degassed and
nanoparticles. deionised water in order to remove extraneous ions.

This journal is © The Royal Society of Chemistry 2017 CrystEngComm, 2017, 19, 246–255 | 247
View Article Online

Paper CrystEngComm

The mild oxidation of magnetite was performed in a ser was used and the laser power was reduced by optical filters
double-walled glass reactor. Here, 1 L of a suspension of to 0.1 mW for each measurement. In order to quantify the
freshly prepared magnetite nanoparticles with a concentra- progress of oxidation, the A1g band between 600 and 750 cm−1
tion of 7 g L−1 was stirred at 400 rpm and heated to 60 °C. Af- was fitted with Voigt functions in Origin 2015 Pro.
This article is licensed under a Creative Commons Attribution-NonCommercial 3.0 Unported Licence.

ter having maintained the temperature constant for 10 min, X-ray Photoelectron Spectroscopy was accomplished with a
compressed air was bubbled through the reaction mixture. Leybold–Heraeus LHS 10 XPS system in ultrahigh vacuum
Aliquots of the reaction mixture were taken after 0, 15, 45 (UHV) hosting a nonmonochromatised Al Kα source (1486.7
min and 2, 4 and 24 h and were frozen at −80 °C prior to eV). The powder samples were fixed on a vacuum compatible
Mössbauer spectroscopy or stored at −20 °C under nitrogen copper foil adhesive tape. The spectra were recorded at a con-
stant pass energy mode set to 100 eV and a FWHM of ∼1.1
Open Access Article. Published on 02 December 2016. Downloaded on 7/24/2019 12:21:42 PM.

for other analyses.


Nitric acid (HNO3) purchased from Merck KGaA, Germany, eV. The C 1s (284.5 eV) peak corresponding to adventitious
was utilised as oxidising agent for the harsh oxidation condi- carbon was used as energy reference to compensate energy
tions. 700 mL of the magnetite (7 g) nitric acid (490 mmol) shifts due to charging. Detail spectra of the C 1s, O 1s N 1s
suspension was stirred at 250 rpm and held at 60 °C for 24 and Fe 2p regions were acquired by repeatedly scanning the
hours. Aliquots were retained from the reaction mixture after same region 30 times in order to reduce statistical noise. All
0, 15, 45 min, 2, 4 and 24 h and neutralised with NaOH. Each spectra were recorded in a UHV at a pressure below 5 × 10–
sample was washed with deionised and degassed water be- 8 mbar. The core level spectra were fitted by a mix of Gauss-
fore being frozen and stored under nitrogen atmosphere at ian and Lorentzian functions (Gaussian line width (0.7 eV)
−80 °C or at −20 °C. and Lorentzian line width (0.3 eV)). Shirley backgrounds were
Particle size and shape were investigated by transmission subtracted from Fe 2p3/2 and O 1s spectra.
electron microscopy (TEM) with a JEM 100-CX from JEOL
GmbH, Germany. For the TEM measurements the colloidal Results and discussion
samples were diluted in degassed and deionised water, ultra-
sonicated and placed on carbon coated copper grids. In order Changes in size of nanoparticles upon oxidation were studied
to receive statistical particle distributions at least 100 parti- by TEM and XRD. While the spherical particle shape
cles were counted per picture. remained constant for both oxidation routes, a slight particle
For further analytical methods, suspensions were growth can be observed (Fig. 1). Besides counting of particle
lyophilised in an ALPHA 1-2LD plus from Martin Christ diameters from TEM micrographs, the crystallite size can be
Gefriertrocknungsanlagen GmbH, Germany. estimated from XRD with the Scherrer equation. We used a
X-ray diffraction (XRD) was performed in transmission ge- spherical shape factor (0.89) and obtained smaller particle di-
ometry with a Stadi-P from STOE & Cie GmbH, Germany ameters than those measured manually on TEM images
equipped with a MoKα source (λ = 0.7093 Å). The full width at which is consistent with literature.56 The synthesised magne-
half maximum (FWHM) of the (311) reflection was used for tite particles exhibit a median particle diameter of 9 and 7.2
particle size determination with the Scherrer equation (ESI†). nm for TEM and XRD analysis, respectively. At this size, mag-
In order to quantify oxidation progress the (440) reflection was netic iron oxide nanoparticles are known to be super-
fitted with Voigt functions in Origin 2015 Pro (S2 and S3†). paramagnetic with a high saturation magnetisation and a
Magnetic properties of the precipitates were characterised high specific surface area.25,39 The diameter for the mildly
with a superconducting quantum interference device (SQUID) oxidised particles increases to 9.4 and 7.3 nm while the
magnetometer MPMS (Quantum Design Inc., USA) at a tem- harshly oxidised particles demonstrate a diameter gain to 10
perature of 300 K. The magnetic field was varied from −50 and 8.0 nm for TEM and XRD, respectively. The particle
kOe to +50 kOe. The 57Fe Mössbauer measurements were growth can be explained by the phase distortion of the origi-
performed in transmission geometry with a source of 57Co in nal inverse spinel and migration of iron ions to the surface
a rhodium matrix at 4.2 K in a liquid helium bath cryostat. and subsequent oxidation.7,44,57,58 Indeed, the XRD patterns
The Mössbauer spectra were fitted with Lorentzian lines confirm preservation of the spinel structure during both oxi-
grouped into appropriate patterns. In addition to magnetic dation processes. Structural reorientation of maghemite par-
sextets and electric quadruple doublets, patterns correspond- ticles smaller than 100 nm in diameter is known to require
ing to Gaussian distributions of magnetic hyperfine fields temperatures over 500 °C.59 Thus, the lattice constant calcu-
were used. Isomer shifts are given as measured with respect lated from XRD decreases with oxidation of nanoparticles
to the source having the same temperature as the absorber. from 8.40 to 8.36 and 8.35 for mild and harsh oxidation con-
In order to convert them to isomer shifts, with respect to me- ditions, respectively (S1†). Nevertheless, approaches from
tallic iron at room temperature, 0.245 mm s−1 was added to Kim et al. exist, which try to differentiate between the
each spectrum obtained. distorted spinel of maghemite and the inverse spinel of mag-
Raman spectroscopy was performed with a Raman Senterra netite in a powder XRD.46 Accordingly, the (440) reflection of
spectrometer from Bruker Optics, Germany. Suspension drops the investigated particle demonstrates a decrease of the mag-
had been deposited on microscope slides and dried under ni- netite share to 20% and 60% under harsh and mild condi-
trogen atmosphere before spectra were recorded. A 488 nm la- tions, respectively (S2 and S3†).

248 | CrystEngComm, 2017, 19, 246–255 This journal is © The Royal Society of Chemistry 2017
View Article Online

CrystEngComm Paper
This article is licensed under a Creative Commons Attribution-NonCommercial 3.0 Unported Licence.
Open Access Article. Published on 02 December 2016. Downloaded on 7/24/2019 12:21:42 PM.

Fig. 2 XRD of synthesised nanoparticles oxidised for different times at


Fig. 1 TEM images of as-synthesised magnetite nanoparticles (a), par- 60 °C in air atmosphere (a) and in nitric acid (b).
ticles oxidised for 24 h at 60 °C in air atmosphere (b) and particles
oxidised for 24 h in nitric acid (c).

effect can be explained by form, size and shape anisotropy of


Maybe the most crucial property of magnetite and the synthesised particles which was verified by evaluating the
maghemite is their ferrimagnetism. The saturation TEM images (Fig. 1).25 The remanence as well as coercivity
magnetisation of as-synthesised 9 nm magnetite nano- are marginally increasing with oxidation time and reaching
particles is determined to be around 75 emu g−1 at ±50 kOe. the highest level after 24 h (S5†). The negative coercivity as
The reduced saturation magnetisation compared to bulk well as negative remanence magnetisation can be explained
magnetite can be explained by a disordered structure near by site-specific surface anisotropy, noncollinear spin struc-
the surface leading to canted spins in the surface layer.20,60–63 ture or antiferromagnetic coupling of particle core and
Our mild oxidation experiment reveals no changes in satura- shell.20,49,61,64,65 The saturation magnetisation of nano-
tion magnetisation as well as in the shape of the particles oxidised in nitric acid decreases much faster
magnetisation curve in the first two hours (Fig. 2). After (Fig. 3). After 15 minutes in the nitric acid, the nanoparticles'
seven hours of oxidation in air the saturation magnetisation magnetisation reaches 67 and after 2 hours 65 emu g−1. The
decreases to 68 emu g−1 and finally reaches a value of 66 emu lowest value observed in our experiments is 54 emu g−1
g−1 after 24 h. On the other hand, the susceptibility of nano- obtained after 24 h in oxidising conditions in a nitric acid so-
particles slightly increases with oxidation resulting in parti- lution. In Fig. 3 the magnetization behaviour of the particles
cles to reach their saturation magnetisation earlier (S4†).49 oxidised by nitric acid is shown as well as the inset for the al-
The initial magnetic susceptibility is a measure for the effec- iquot after 24 hours compared to magnetite at small values
tive anisotropy barrier of nanoparticles. The observed in- for the magnetic field. For this experiment, the
crease in susceptibility for the particles oxidised in air can be magnetisation decreased to 67 emu g−1 already after 15 mi-
explained with increased interparticle interactions upon oxi- nutes and to 65 emu g−1 after 45 minutes. Even though the
dation. The increase in initial susceptibility is lower for change of the saturation magnetization after 2 hours is mar-
harshly oxidised particles as the adsorption of nitrate ions ginal, after 24 hours it decreased to 54 emu g−1. The shape of
might lead to lesser interparticle interaction.64 In the absence the hysteresis curve with a coercive field around −15 Oe as
of a magnetic field, all samples show a similar low rema- well as the remanence do not change significantly upon oxi-
nence around ±1 emu g−1 which is generally discussed as dation which is in good agreement with values generally
magnetic viscosity for superparamagnetic materials.27 This stated in the literature.28,49 The first drop in coercivity and

This journal is © The Royal Society of Chemistry 2017 CrystEngComm, 2017, 19, 246–255 | 249
View Article Online

Paper CrystEngComm

15 minutes only half of the divalent ions can be observed.


With the end of the experiments after 24 h, no divalent ions
can be evidenced in the spectra.66 We were only able to con-
sistently differentiate between divalent and trivalent iron spe-
This article is licensed under a Creative Commons Attribution-NonCommercial 3.0 Unported Licence.

cies as the arrangement of iron ions in magnetite and


maghemite and transition states is very similar (S6–17†).
Besides Mössbauer spectroscopy, Raman spectroscopy, of-
fers a great alternative to investigate iron oxide nanoparticles.
While Mössbauer spectroscopy tends to be an analysis tech-
Open Access Article. Published on 02 December 2016. Downloaded on 7/24/2019 12:21:42 PM.

nique for bulk materials, Raman spectroscopy with a limited


penetration depth, especially for a 488 nm laser, renders in-
formation on surface/nanoscale properties of materials. The
oxidation of the divalent iron ions can be verified by the Ra-
man mode (A1g) near 660 cm−1 (S18†). For maghemite this
mode is split into two equal components around 660 and 710
cm−1 (S19†).64,67 According to literature, the second mode re-
fers to the oxidation of divalent iron ions on octahedral
sites.68 Under mild oxidation conditions after 24 h the ratio
between the band at 660 and 710 cm−1 is still 1.3 : 1.
Mössbauer spectroscopy indicates significantly decreasing
amounts of divalent iron ions even after 15 min of mild oxi-
dation. Such slow oxidation disagrees with our Mössbauer
data. Therefore, we suggest a beginning lattice distortion
which occurs more slowly than the oxidation of divalent ions
and explains the increasing peak around 710 cm−1.51,67,68 The
distortion is further emphasised by the 2 bands around 300
and 530 cm−1 corresponding to T2g modes of magnetite
Fig. 3 SQUID magnetisation measurements of synthesised changing with oxidation. Here, an increase in intensity as
nanoparticles oxidised at different times at 60 °C in air atmosphere (a) well as a redshift of the band at 300 and a blueshift of the
and in nitric acid (b).
band around 530 cm−1 which overlaps with the Eg band of
maghemite can be observed (S20†). The change of phonon
modes is faster for the oxidation in nitric acid compared to
remanence magnetisation can be explained by the the oxidation in air. After 24 h under harsh oxidation condi-
stabilisation of the surface layer by nitrate ions which is tions almost a ratio of around 1 : 1 for the component around
reported in literature for other stabilisers.25,28,63 The further 660 cm−1 to the component around 710 cm−1 is reached. We
increase of the coercivity can be related to an increasing were not able to observe a hydroxide surface layer with Ra-
anisotropy by phase transformation to maghemite.28 man spectroscopy but the broad bands can be related to
Mössbauer spectroscopy is a powerful tool for the identifi- quantum effects of nanomaterials or the existence of multi-
cation of the chemical state and crystal environment of iron ple iron oxide or (oxy)hydroxide species which are close to
species. Hence, the distinction between the relatively similar bands corresponding to maghemite (Fig. 5).51
iron oxides maghemite and magnetite is possible. At 4.2 K XPS enables analysing the outmost surface layers of mate-
the contributions of the divalent and trivalent iron ions can rials. While a differentiation between trivalent iron ions in
be separated and quantified. At room temperature electron iron oxides and (oxy)hydroxides is quite difficult by analyzing
hopping occurs and only the tetrahedral Fe3+ and a combined the Fe 2p bands of nanoparticles, a discrimination between
Fe2.5+ species on octahedral sites can be differentiated.47,48 divalent and trivalent iron ions occuring in magnetite is in-
The main problem for nanoparticles are the Néel relaxation deed possible. The Fe 2p3/2 spectra were fitted according to
and the Brownian motion, which make it practically impossi- the model suggested by Grosvenor et al. and are shown in
ble to obtain resolved magnetic hyperfine splitting at ambi- Fig. 6.55 We were able to verify Fe2+ ion in the as-synthesised
ent temperature. However, at 4.2 K the localized Fe2+ ions magnetite nanoparticles while the particles oxidised for 24
show a significantly smaller magnetic field and a shift com- hours only demonstrate trivalent iron ions. Spectra in the
pared to the trivalent ions on octahedral sites. The trivalent whole Fe 2p range shown in the ESI† demonstrate broader
ions on tetrahedral sites show a similar behaviour as Fe3+ on peaks corresponding to Fe 2p3/2 and Fe 2p1/2 for magnetite
octahedral sites. Thus, we can follow the oxidation of iron than for the oxidised nanoparticles (S21†). Moreover, the
ions and the change of the crystal structure in the Mössbauer shake-up satellites of the oxidised species are more pro-
spectra (Fig. 4). For mild and for harsh oxidation conditions, nounced and at higher binding energies which is in good
the amount of divalent ions decreases rapidly with time. After agreement with literature spectra of magnetite and

250 | CrystEngComm, 2017, 19, 246–255 This journal is © The Royal Society of Chemistry 2017
View Article Online

CrystEngComm Paper
This article is licensed under a Creative Commons Attribution-NonCommercial 3.0 Unported Licence.
Open Access Article. Published on 02 December 2016. Downloaded on 7/24/2019 12:21:42 PM.

Fig. 5 Raman spectra of synthesised nanoparticles oxidised at


different times at 60 °C in air atmosphere (a) and in nitric acid (b).

The multi-analytical characterisation of magnetite oxida-


tion under harsh and mild conditions, presented in this
study illustrates two different oxidation routes and final
nanoparticles. Whereas magnetite nanoparticles oxidise al-
most completely to maghemite under harsh conditions, the
Fig. 4 Mössbauer spectra of synthesised nanoparticles oxidised at mild oxidation leads to particles in a transition state between
different times at 60 °C in air atmosphere (a) and in nitric acid (b). magnetite and maghemite. The harshly oxidised particles
demonstrate a lower saturation magnetisation, the lattice
constant is similar to bulk maghemite, only 3% divalent iron
ions remain in the particle core and the Raman phonon
maghemite, respectively.55 The ratio of divalent to trivalent modes are also almost similar to bulk maghemite.52,69 Simi-
iron ions is 1 : 2 which is quite close to the value reported in lar Fe2+ concentrations have also been observed for
literature. The analysis of the oxygen core spectra yields the maghemite in literature.69 The oxidation is much faster un-
existence of iron oxide species (530 eV) as well as hydroxyl der harsh conditions and the particles derived after 2 h of ox-
groups (532 eV) which were assigned to a surface hydroxide idation show the same characteristics as the particles
in all samples.15 The nanoparticles oxidised in acid show an oxidised for 24 h but the saturation magnetisation (Fig. 7a).
additional chemical oxygen shift around 533.5 eV which can For the mild oxidation conditions, especially the transition of
be assigned to nitric acid. As O 1s fit corresponding to the hy- the magnetite crystal phase to maghemite is incomplete after
droxyl group is below 10% of the whole oxygen content we as- 24 h (Fig. 7b). The oxidation of divalent iron ions as well as
sume that only the outermost layer is hydroxylised while the the increase of the phonon mode at 710 cm−1 is significantly
residual part of the particles contains iron oxide. Fig. 6c dem- slower than under the harsh oxidation conditions. Hence, we
onstrates nitric acid bound to the particles while surface hy- assume an oxidation of nanoparticles from the surface to the
droxide can still be observed. A nitrogen band around 408 eV core where the divalent irons are oxidised prior to phase tran-
which can be assigned to a core shift of nitric acid can be sition. After an oxidation of ferrous to ferric ions which starts
evidenced (S22†). While acidic groups are often reported to on the surface, the inverse spinel structure is degraded before
replace surface hydroxides we cannot confirm this behaviour being reoriented as defect spinel. The ferrous iron ions are
for the maghemite nanoparticles.15,23,63 migrating from core to surface. This process is faster in nitric

This journal is © The Royal Society of Chemistry 2017 CrystEngComm, 2017, 19, 246–255 | 251
View Article Online

Paper CrystEngComm
This article is licensed under a Creative Commons Attribution-NonCommercial 3.0 Unported Licence.
Open Access Article. Published on 02 December 2016. Downloaded on 7/24/2019 12:21:42 PM.

Fig. 6 XP spectra of synthesised magnetite nanoparticles (a) and oxidised particles after 24 at 60 °C under air atmosphere (b) and in nitric acid (c).

acid than in air and the ions are able to precipitate on the for discussion on the possibility of an oxyhydroxide surface.
particle surface leading to slightly larger particles as observed However, from our Mössbauer data we were able to exclude
in the TEM and XRD size analysis. The surface of magnetite loosely bound iron ions adsorbed on the surface. From X-ray
nanoparticles seems to be disordered even before oxidation photoelectron spectroscopy we evidence that the magnetite
as the saturation magnetization is lower than bulk material nanoparticle surface contains hydroxyl groups. These groups
and the remanence magnetisation as well as the coercivity play a key role for the interactions with molecules.15
are inverse to the magnetisation in the respective magnetic In order to further analyse the surface of the magnetic
field prior to the shutdown of the magnetic field. The surface nanoparticles for application scopes, besides zeta potential
layer of magnetic nanoparticles is widely discussed in litera- measurements, the cytotoxicology and the capability of acting
ture.60,68,70 Most groups assume that the nanoparticle surface as a Fenton catalyst were investigated. The isoelectric point
contains an non-stoichiometric oxyhydroxide but were not (IEP) for magnetite is determined at a pH value of 7.0, for the
able to prove this with Raman or Mössbauer spectroscopy as particles oxidised by air at 8.4 and for the ones treated with
the vibrations of different iron oxides and (oxy)hydroxides nitric acid at 6.3. For magnetite a value of 6.5 was reported
are quite similar. Our Raman results still leave some space by Plaza et al. and others; they also reported a change to 5.5

252 | CrystEngComm, 2017, 19, 246–255 This journal is © The Royal Society of Chemistry 2017
View Article Online

CrystEngComm Paper

dation of rhodamine 6G than magnetite nanoparticles, which


can be ascribed to the lack of ferrous ions on the surface
(S24†). However, the cytotoxicity of the synthesised magnetite
nanoparticles seems to be negligible as Escherichia coli
This article is licensed under a Creative Commons Attribution-NonCommercial 3.0 Unported Licence.

HMS174IJDE3) pET28a cells were able to grow though being


incubated with these particles. Magnetite is usually known to
have an impact on the growth of microorganisms, while
maghemite nanoparticles demonstrate no cytotoxicity; we
were not able to detect any difference between the influences
of these particles on the bacterial growth (S25–27†).41,42
Open Access Article. Published on 02 December 2016. Downloaded on 7/24/2019 12:21:42 PM.

Conclusions
In summary, our investigations on the oxidation of magnetite
nanoparticles reflect the behaviour complexity of nanoscale
iron oxides. We monitored the oxidation process of magne-
tite nanoparticles with an average diameter, obtained from
TEM values, of 9 nm in harsh and mild oxidative conditions.
The harsh oxidation conditions, where magnetite particles
were incubated with 0.07 mol L−1 HNO3, lead to a complete
phase transformation to maghemite after 24 h. This transfor-
mation could be evidenced by the amount of ferrous iron
ions, obtained by Mössbauer spectroscopic measurements
and by the change of the lattice constant as well as by the
appearing of a maghemite-specific phonon species. Further-
more, a decrease of the saturation magnetisation was ob-
served for the oxidation with nitric acid. The nitric acid catal-
Fig. 7 Change of magnetite content calculated from Raman,
Mössbauer and XRD data of nanoparticles oxidised for different times yses the oxidation of magnetite to maghemite and is
at 60 °C in air atmosphere (a) and in nitric acid (b). adsorbed to the surface of the particles. The mild oxidation,
where the particles were heated to 60 °C and stirred vigor-
ously in air atmosphere, lead to different particles. Here, a
higher saturation magnetisation was conserved over 24 h of
after 7 days.66,71 They suggested this shift of the IEP for oxidation and the crystal phase was not completely trans-
maghemite due to oxidation of the particles in the suspen- formed to maghemite as observed with XRD and Raman
sion and verified this assumption by measuring a commer- spectroscopy. On the other hand, a complete oxidation of fer-
cial maghemite sample.71 In our case, we observed this be- rous iron ions was also observed for mild oxidation condi-
haviour only for the particles resulting from the oxidation in tions. The different development of the surface charge, indi-
nitric acid solution. For the maghemite oxidised in air the cated by the zeta potentials, upon both synthesis routes is
IEP shifts, contrary to the findings of Plaza et al., to a higher noteworthy. While the incubation with nitric acid lead to a
value (S23†).71 Even after several months the zeta potentials lower IEP of nanoparticles, the mild oxidation lead to a
of both oxidised iron oxide nanoparticle species stored in higher IEP. This enables easy tuning of the surface charge by
deionised water showed no changes. We assume the integra- change of oxidation parameters for particles still possessing
tion or adsorption of negatively charged nitrate into the sur- a high saturation magnetisation (>54 emu g−1) and super-
face layer of the particles oxidised in nitric acid which is also paramagnetic relaxation. Furthermore, the surface spin cant-
confirmed with XP spectra in the N 1s region (S22†). Similar ing was observed in all particles as the coercivity and rema-
behaviour was observed by Daou et al., where phosphate was nence saturation are negative for all samples. This means the
strongly adsorbed on magnetic nanoparticles.72 As we as- surface layer is similar for wet chemically synthesised
sume a transport of ferrous iron ions to the surface in our maghemite and magnetite nanoparticles. We were able to
magnetite nanoparticles, the Fenton reaction which is depen- verify the Fenton activity of our magnetite particles due to
dent on the existence of these ions was investigated. Thus, their ferrous iron ion content while no cytotoxicity towards E.
we were able to determine a catalytic activity of the as coli was detected.
synthesised magnetite nanoparticles towards the degradation
of rhodamine 6G in the presence of H2O2. The activity is sig- Acknowledgements
nificantly higher than of commercially available magnetite
and represents a promising catalyst for dye removal.73 How- The authors would like to thank Dr. Marianne Hanzlik and
ever, the oxidised nanoparticles demonstrate a lower degra- Dr. Carsten Peters for support with TEM imaging, Prof.

This journal is © The Royal Society of Chemistry 2017 CrystEngComm, 2017, 19, 246–255 | 253
View Article Online

Paper CrystEngComm

Sebastian Günther for support with XPS and Prof. Dr. Tom 20 C. Graf, C. Goroncy, P. Stumpf, E. Weschke, C. Boeglin, H.
Nilges for the provision of the X-ray diffractometer. We thank Ronneburg and E. Rühl, J. Phys. Chem. C, 2015, 119,
the students Matthias Alt and Teng Daquan for toxicity and 19404–19414.
catalytic activity tests, respectively. Furthermore, we are partic- 21 E. Duguet, M.-H. Delville and S. Mornet, in Magnetic
This article is licensed under a Creative Commons Attribution-NonCommercial 3.0 Unported Licence.

ularly appreciative for the financial support of this work by Nanoparticles. From Fabrication to Clinical Applications, ed.
the Federal Ministry of Education and Research (Grant num- N. T. Thanh, CRC Press, Hoboken, 2012, pp. 47–72.
ber 031A173A). 22 P. V. Nidheesh, RSC Adv., 2015, 5, 40552–40577.
23 H.-C. Roth, S. Schwaminger, P. Fraga García, J. Ritscher and
References S. Berensmeier, J. Nanopart. Res., 2016, 18, 99.
Open Access Article. Published on 02 December 2016. Downloaded on 7/24/2019 12:21:42 PM.

24 T. Marin, P. Montoya, O. Arnache and J. Calderon, J. Phys.


1 P. Fraga García, M. Brammen, M. Wolf, S. Reinlein, M. Chem. B, 2016, 120, 6634–6645.
Freiherr von Roman and S. Berensmeier, Sep. Purif. Technol., 25 A. G. Kolhatkar, A. C. Jamison, D. Litvinov, R. C. Willson
2015, 150, 29–36. and T. R. Lee, Int. J. Mol. Sci., 2013, 14, 15977–16009.
2 M. Franzreb, M. Siemann-Herzberg, T. J. Hobley and O. R. T. 26 A.-H. Lu, E. L. Salabas and F. Schuth, Angew. Chem., Int. Ed.,
Thomas, Appl. Microbiol. Biotechnol., 2006, 70, 505–516. 2007, 46, 1222–1244.
3 H.-C. Roth, S. P. Schwaminger, F. Peng and S. Berensmeier, 27 S. Laurent, D. Forge, M. Port, A. Roch, C. Robic, L. Vander
ChemistryOpen, 2016, 5, 183–187. Elst and R. N. Muller, Chem. Rev., 2008, 108, 2064–2110.
4 G. Liu, J. Gao, H. Ai and X. Chen, Small, 2013, 9, 1533–1545. 28 W. Baaziz, B. P. Pichon, S. Fleutot, Y. Liu, C. Lefevre, J.-M.
5 L. M. Bauer, S. F. Situ, M. A. Griswold and A. C. S. Samia, Greneche, M. Toumi, T. Mhiri and S. Begin-Colin, J. Phys.
Nanoscale, 2016, 8, 12162–12169. Chem. C, 2014, 118, 3795–3810.
6 L. M. Rossi, N. J. S. Costa, F. P. Silva and R. Wojcieszak, 29 T. Fan, D. Pan and H. Zhang, Ind. Eng. Chem. Res., 2011, 50,
Green Chem., 2014, 16, 2906. 9009–9018.
7 J. Liu, Z. Wu, Q. Tian, W. Wu and X. Xiao, CrystEngComm, 30 T. J. Daou, G. Pourroy, S. Bégin-Colin, J. M. Grenèche, C.
2016, 18, 6303–6326. Ulhaq-Bouillet, P. Legaré, P. Bernhardt, C. Leuvrey and G.
8 L.-L. Tian, M.-J. Zhang, C. Wu, Y. Wei, J.-X. Zheng, L.-P. Lin, Rogez, Chem. Mater., 2006, 18, 4399–4404.
J. Lu, K. Amine, Q.-C. Zhuang and F. Pan, ACS Appl. Mater. 31 D. Ramimoghadam, S. Bagheri and S. B. A. Hamid, J. Magn.
Interfaces, 2015, 7, 26284–26290. Magn. Mater., 2014, 368, 207–229.
9 A. M. Bruck, C. A. Cama, C. N. Gannett, A. C. Marschilok, 32 S. Kluge, L. Deng, O. Feroughi, F. Schneider, M. Poliak, A.
E. S. Takeuchi and K. J. Takeuchi, Inorg. Chem. Front., Fomin, V. Tsionsky, S. Cheskis, I. Wlokas, I. Rahinov, T.
2016, 3, 26–40. Dreier, A. Kempf, H. Wiggers and C. Schulz, CrystEngComm,
10 P. Xu, G. M. Zeng, D. L. Huang, C. L. Feng, S. Hu, M. H. 2015, 17, 6930–6939.
Zhao, C. Lai, Z. Wei, C. Huang, G. X. Xie and Z. F. Liu, Sci. 33 J. Baumgartner, A. Dey, P. H. H. Bomans, C. Le Coadou, P.
Total Environ., 2012, 424, 1–10. Fratzl, N. A. J. M. Sommerdijk and D. Faivre, Nat. Mater.,
11 S. R. Chowdhury, E. K. Yanful and A. R. Pratt, J. Hazard. 2013, 12, 310–314.
Mater., 2012, 235–236, 246–256. 34 T. Ahn, J. H. Kim, H.-M. Yang, J. W. Lee and J.-D. Kim,
12 R. A. Crane, M. Dickinson, I. C. Popescu and T. B. Scott, J. Phys. Chem. C, 2012, 116, 6069–6076.
Water Res., 2011, 45, 2931–2942. 35 B. Bateer, C. Tian, Y. Qu, S. Du, T. Tan, R. Wang, G. Tian
13 R. Turcu, V. Socoliuc, I. Craciunescu, A. Petran, A. Paulus, and H. Fu, CrystEngComm, 2013, 15, 3366.
M. Franzreb, E. Vasile and L. Vekas, Soft Matter, 2015, 11, 36 J. Lim, K. Sim and J.-K. Lee, CrystEngComm, 2016, 18,
1008–1018. 2155–2162.
14 K. Pušnik, M. Peterlin, I. K. Cigić, G. Marolt, K. Kogej, A. 37 D. Forge, A. Roch, S. Laurent, H. Tellez, Y. Gossuin, F.
Mertelj, S. Gyergyek and D. Makovec, J. Phys. Chem. C, Renaux, L. Vander Elst and R. N. Muller, J. Phys. Chem. C,
2016, 120, 14372–14381. 2008, 112, 19178–19185.
15 S. P. Schwaminger, P. Fraga García, G. K. Merck, F. A. 38 R. M. Fratila, S. Rivera-Fernandez and J. M. de La Fuente,
Bodensteiner, S. Heissler, S. Günther and S. Berensmeier, Nanoscale, 2015, 7, 8233–8260.
J. Phys. Chem. C, 2015, 119, 23032–23041. 39 H.-C. Roth, S. P. Schwaminger, M. Schindler, F. E. Wagner
16 C. A. Quinto, P. Mohindra, S. Tong and G. Bao, Nanoscale, and S. Berensmeier, J. Magn. Magn. Mater., 2015, 377,
2015, 7, 12728–12736. 81–89.
17 U. Ikoba, H. Peng, H. Li, C. Miller, C. Yu and Q. Wang, 40 M. Auffan, J. Rose, J.-Y. Bottero, G. V. Lowry, J.-P. Jolivet and
Nanoscale, 2015, 7, 4291–4305. M. R. Wiesner, Nat. Nanotechnol., 2009, 4, 634–641.
18 H. Liu, J. Zhang, X. Chen, X.-S. Du, J.-L. Zhang, G. Liu and 41 M. Auffan, W. Achouak, J. Rose, M.-A. Roncato, C. Chanéac,
W.-G. Zhang, Nanoscale, 2016, 8, 7808–7826. D. T. Waite, A. Masion, J. C. Woicik, M. R. Wiesner and J.-Y.
19 B. H. Kim, N. Lee, H. Kim, K. An, Y. I. Park, Y. Choi, K. Shin, Bottero, Environ. Sci. Technol., 2008, 42, 6730–6735.
Y. Lee, S. G. Kwon, H. B. Na, J.-G. Park, T.-Y. Ahn, Y.-W. 42 V. Aruoja, S. Pokhrel, M. Sihtmäe, M. Mortimer, L.
Kim, W. K. Moon, S. H. Choi and T. Hyeon, J. Am. Chem. Mädler and A. Kahru, Environ. Sci.: Nano, 2015, 2,
Soc., 2011, 133, 12624–12631. 630–644.

254 | CrystEngComm, 2017, 19, 246–255 This journal is © The Royal Society of Chemistry 2017
View Article Online

CrystEngComm Paper

43 P. Avetta, A. Pensato, M. Minella, M. Malandrino, V. 60 R. Frison, G. Cernuto, A. Cervellino, O. Zaharko, G. M.


Maurino, C. Minero, K. Hanna and D. Vione, Environ. Sci. Colonna, A. Guagliardi and N. Masciocchi, Chem. Mater.,
Technol., 2015, 49, 1043–1050. 2013, 25, 4820–4827.
44 R. M. Cornell and U. Schwertmann, The Iron Oxides, Wiley- 61 E. Tronc, A. Ezzir, R. Cherkaoui, C. Chanéac, M. Noguès, H.
This article is licensed under a Creative Commons Attribution-NonCommercial 3.0 Unported Licence.

VCH Verlag GmbH & Co. KGaA, Weinheim, FRG, 2003. Kachkachi, D. Fiorani, A. Testa, J. Grenèche and J. Jolivet,
45 P. Dhakal, C. J. Matocha, F. E. Huggins and M. M. J. Magn. Magn. Mater., 2000, 221, 63–79.
Vandiviere, Environ. Sci. Technol., 2013, 47, 6206–6213. 62 M. P. Morales, S. Veintemillas-Verdaguer, M. I. Montero,
46 W. Kim, C.-Y. Suh, S.-W. Cho, K.-M. Roh, H. Kwon, K. Song C. J. Serna, A. Roig, L. Casas, B. Martínez and F.
and I.-J. Shon, Talanta, 2012, 94, 348–352. Sandiumenge, Chem. Mater., 1999, 11, 3058–3064.
Open Access Article. Published on 02 December 2016. Downloaded on 7/24/2019 12:21:42 PM.

47 F. J. Berry, S. Skinner and M. F. Thomas, J. Phys.: Condens. 63 J. Salafranca, J. Gazquez, N. Perez, A. Labarta, S. T.
Matter, 1998, 10, 215–220. Pantelides, S. J. Pennycook, X. Batlle and M. Varela, Nano
48 G. A. Sawatzky, J. Appl. Phys., 1969, 40, 1402. Lett., 2012, 12, 2499–2503.
49 J. Santoyo Salazar, L. Perez, O. de Abril, L. Truong Phuoc, D. 64 R. L. Rebodos and P. J. Vikesland, Langmuir, 2010, 26,
Ihiawakrim, M. Vazquez, J.-M. Greneche, S. Begin-Colin and 16745–16753.
G. Pourroy, Chem. Mater., 2011, 23, 1379–1386. 65 S. Gu, W. He, M. Zhang, T. Zhuang, Y. Jin, H. ElBidweihy, Y.
50 I. Dézsi, C. Fetzer, Á. Gombkötő, I. Szűcs, J. Gubicza and T. Mao, J. H. Dickerson, M. J. Wagner, E. Della Torre and L. H.
Ungár, J. Appl. Phys., 2008, 103, 104312. Bennett, Sci. Rep., 2014, 4, 6267.
51 D. L. A. de Faria, S. Venâncio Silva and M. T. de Oliveira, 66 S. J. Iyengar, M. Joy, C. K. Ghosh, S. Dey, R. K. Kotnala and
J. Raman Spectrosc., 1997, 28, 873–878. S. Ghosh, RSC Adv., 2014, 4, 64919–64929.
52 A. M. Jubb and H. C. Allen, ACS Appl. Mater. Interfaces, 67 Y. El Mendili, F. Grasset, N. Randrianantoandro, N.
2010, 2, 2804–2812. Nerambourg, J.-M. Greneche and J.-F. Bardeau, J. Phys.
53 I. Chamritski and G. Burns, J. Phys. Chem. B, 2005, 109, Chem. C, 2015, 119, 10662–10668.
4965–4968. 68 G. V. Jacintho, P. Corio and J. C. Rubim, J. Electroanal.
54 O. N. Shebanova and P. Lazor, J. Raman Spectrosc., 2003, 34, Chem., 2007, 603, 27–34.
845–852. 69 P. Belleville, J.-P. Jolivet, E. Tronc and J. Livage, J. Colloid
55 A. P. Grosvenor, B. A. Kobe, M. C. Biesinger and N. S. Interface Sci., 1992, 150, 453–460.
McIntyre, Surf. Interface Anal., 2004, 36, 1564–1574. 70 I. Chourpa, L. Douziech-Eyrolles, L. Ngaboni-Okassa, J.-F.
56 Y.-k. Sun, M. Ma, Y. Zhang and N. Gu, Colloids Surf., A, Fouquenet, S. Cohen-Jonathan, M. Souce, H. Marchais and
2004, 245, 15–19. P. Dubois, Analyst, 2005, 130, 1395–1403.
57 T. W. Swaddle and P. Oltmann, Can. J. Chem., 1980, 58, 71 R. C. Plaza, J. L. Arias, M. Espin, M. L. Jimenez and A. V.
1763–1772. Delgado, J. Colloid Interface Sci., 2002, 245, 86–90.
58 S. Nie, E. Starodub, M. Monti, D. A. Siegel, L. Vergara, F. El 72 T. J. Daou, S. Begin-Colin, J. M. Grenèche, F. Thomas, A.
Gabaly, N. C. Bartelt, J. de La Figuera and K. F. McCarty, Derory, P. Bernhardt, P. Legaré and G. Pourroy, Chem.
J. Am. Chem. Soc., 2013, 135, 10091–10098. Mater., 2007, 19, 4494–4505.
59 C. J. Goss, Phys. Chem. Miner., 1988, 16, 164–171. 73 K. Sun, C. Sun and S. Tang, CrystEngComm, 2016, 18, 714–720.

This journal is © The Royal Society of Chemistry 2017 CrystEngComm, 2017, 19, 246–255 | 255

You might also like