You are on page 1of 13

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/325983265

Hydrothermal synthesis, morphology, magnetic properties and self-assembly


of hierarchical α-Fe 2 O 3 (hematite) mushroom-, cube- and sphere-like
superstructures

Article  in  Applied Surface Science · June 2018


DOI: 10.1016/j.apsusc.2018.06.224

CITATION READS

1 27

4 authors:

Djordje Trpkov Matjaž Panjan


Vinča Institute of Nuclear Sciences Jožef Stefan Institute
8 PUBLICATIONS   42 CITATIONS    55 PUBLICATIONS   795 CITATIONS   

SEE PROFILE SEE PROFILE

Lazar Kopanja Marin Tadic


Univerzitet ALFA Vinca Institute, University of Belgrade
17 PUBLICATIONS   163 CITATIONS    45 PUBLICATIONS   720 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

HiPIMS Studies View project

Reactive HiPIMS View project

All content following this page was uploaded by Marin Tadic on 23 December 2018.

The user has requested enhancement of the downloaded file.


Applied Surface Science 457 (2018) 427–438

Contents lists available at ScienceDirect

Applied Surface Science


journal homepage: www.elsevier.com/locate/apsusc

Full Length Article

Hydrothermal synthesis, morphology, magnetic properties and self-assembly T


of hierarchical α-Fe2O3 (hematite) mushroom-, cube- and sphere-like
superstructures

Djordje Trpkova, Matjaž Panjanb, Lazar Kopanjac,d, Marin Tadiće,
a
Vinca Institute of Nuclear Sciences, University of Belgrade, Mike Petrovica Alasa 12-14, 11001 Belgrade, Serbia
b
Jožef Stefan Institute, Jamova 39, 1000 Ljubljana, Slovenia
c
Faculty of Technology and Metallurgy, University of Belgrade, Belgrade, Serbia
d
Faculty of Mathematics and Computer Science, Alfa BK University, Palmira Toljatija 3, 11070 Belgrade, Serbia
e
Condensed Matter Physics Laboratory, Vinca Institute, University of Belgrade, P.O. Box 522, 11001 Belgrade, Serbia

A R T I C LE I N FO A B S T R A C T

Keywords: We report on glycine-free and glycine-assisted hydrothermal synthesis of microsized superstructures composed
Iron oxide of self-assembled hematite nanoparticles. An X-ray powder diffraction measurements of the samples confirm
Hematite (α-Fe2O3) good crystallization of the hematite nanoparticles with hydrothermal reaction time-dependent crystallite sizes in
Hydrothermal synthesis a range from ∼15 nm (45 h) to ∼26 nm (90 h). The FTIR and Raman spectroscopy confirm hematite structure,
Image analysis
whereas TEM measurements reveal nanoparticle sub-units (subparticles). The computational analyses of particle
Surface effects
Magnetic properties
shape show that the addition of glycine surfactant in hydrothermal reaction leads to more spherical shape of
hematite hierarchical structures and smaller sizes. We found strong coercivity increases (up to ∼3 times) in the
samples synthesized in the presence of glycine. The coercivity values from HC = 1305 Oe (mushroom-like shape
synthesized by glycine-free hydrothermal reaction) to HC = 3725 Oe (sphere-like shape synthesized by glycine-
assisted hydrothermal reaction) were obtained at 300 K. These results and their comparison with other described
in the literature (e.g. bulk, wires, urchin-like, rods, tubes, plates, star-like, dendrites, platelets, irregular, na-
nocolumns, spindles, disks hematites, etc.) reveal that the hematite superstructures possess good magnetic
properties. We propose that the glycine, oriented subparticles, exchange and dipole-dipole interactions may play
an important role in the development of magnetic properties.

1. Introduction space group R3c and lattice parameters a = 5.0356 Å, c = 13.7489 Å,


having Fe3+ ions occupy two thirds of its octahedral sites that are
The most common natural forms of iron oxides are magnetite confined by the nearly ideal hexagonal closed-pack O lattice. It is n-type
(Fe3O4), maghemite (γ -Fe2O3) and hematite (α-Fe2O3) [1–29]. During metal oxide semiconductor with bandgap around 2.1 eV [66]. Hematite
the past decades great attention has been paid to less known iron oxide in bulk form is paramagnetic at temperatures above its Curie tem-
polymorphs, β-Fe2O3 and ε-Fe2O3 [30–43]. Iron oxides have been ex- perature of TC ≈ 956 K. Below TC it is weakly ferromagnetic material
tensively studied as materials with applications such as microwave and undergoes a phase transition at TM ∼ 260 K (the Morin tempera-
absorbtion [44–51]. Hematite is an inexpensive metal oxide, non-toxic ture) to an antiferromagnetic state. The magnetic behavior of hematite
and environmentally friendly with good stability and magnetically re- depends on crystallinity, particle size, cation substitution, shape, and
coverable material [52]. It is widely used in diverse practical applica- exchange and dipole-dipole interactions [67]. In nanoparticles, the
tions, such as catalysis [53], environmental protection [54,55], gas magnetic properties that are influenced by particle size and surface
sensors [56,57], Li-ion batteries [58,59], and water splitting [60–62]. effects become more important as particle size decreases, due to the
Examples of biomedical applications include new ways of cancer increase of the surface to volume ratio. The surface crystallography and
treatment such as magnetic drug targeting [63] and magnetic hy- faceting are important in determining the spin structure and overall
perthermia [64], as contrast agents or tracers in magnetic resonance magnetization of the particles [68].
imaging [65]. Hematite is isostructural with corundum (α-Al2O3), with Much effort has been made in the design of α-Fe2O3 materials with a


Corresponding author at: Condensed Matter Physics Laboratory, Vinca Institute, P.O. Box 522, 11001 Belgrade, Serbia.
E-mail address: marint@vinca.rs (M. Tadić).

https://doi.org/10.1016/j.apsusc.2018.06.224
Received 12 December 2017; Received in revised form 14 May 2018; Accepted 24 June 2018
Available online 25 June 2018
0169-4332/ © 2018 Elsevier B.V. All rights reserved.
D. Trpkov et al. Applied Surface Science 457 (2018) 427–438

Table 1 desired structure and morphology such as zero-dimensional (0D) (na-


The chemicals used in the synthesis and reaction conditions. noparticles) [69,70], one-dimensional (1D) (nanowires [71], nanobelts
Sample Reagents FeCl3·6H2O Reaction Reaction [72], nanorods [73,74], nanotubes [75]), two-dimensional (2D) (na-
temperature [°C] time [h] norings [76], nanoflakes [77,78]) and three-dimensional (3D) (nano-
cubes [79], urchin-like nanostructures [80]). Hematite nanocrystals
S1 15 ml DI H2O, 1 ml 20 mmol 140 45
with different sizes, morphologies and magnetic properties have been
NH4OH (25%)
S2 15 ml DI H2O, 1 ml 6 mmol 140 90
often synthesized via a hydrothermal method. Reaction conditions and
NH4OH (25%) surfactants play crucial role in the formation of different structures and
S3 6 mmol glycine, 6 mmol 140 90 morphologies under hydrothermal conditions. It was reported that
15 ml DI H2O, 1 ml synthesized sample with molar ratio Fe3+/glycine = 1:1 (G1) in hy-
NH4OH (25%)
drothermal reaction has a wide open M−H loop with a coercivity of
S4 20 mmol glycine, 20 mmol 140 45
15 ml DI H2O, 1 ml 5575.6 Oe, whereas the coercivity of sample with Fe3+/glycine = 1:12
NH4OH (25%) (G4) has only 42.7 Oe. The high coercivity was attributed to the in-
crease in the resistance of the domain rotation. The reported remanent
and saturation magnetizations of G1 and G4 samples are

Fig. 1. Structural, morphological and magnetic characterization of mushroom-like particles (sample S1, glycine-free hydrothermal reaction and 45 h reaction time):
(a) X-ray diffraction pattern, (b–d) SEM micrographs, and (e) magnetic field dependence of magnetization M(H) at 300 K.

428
D. Trpkov et al. Applied Surface Science 457 (2018) 427–438

Fig. 2. Structural, morphological and magnetic characterization of cube-like particles (sample S2, glycine-free hydrothermal reaction and 90 h reaction time): (a) X-
ray diffraction pattern, (b–d) SEM micrographs, and (e) magnetic field dependence of magnetization M(H) at 300 K.

Mr = 0.1938 emu/g, MS = 0.4629 emu/g and Mr = 0.0168 emu/g, 417.0 Oe, whereas the remanent magnetization increase from 0.0178 to
Ms = 0.5235 emu/g, respectively [81]. Hematite synthesis using gly- 0.0244 emu/g and 0.0268 emu/g [86]. Magnetic characterization of
cine-assisted hydrothermal reaction is also mentioned in few reports in hierarchical polyhedral α-Fe2O3 microparticles indicated saturation
current decade [82–84], although without magnetic characterization of magnetization MS = 2.682 emu/g, remanent magnetization
materials. Yin, Minakshi et al. [82] obtained cubic α-Fe2O3 micro- Mr = 0.549 emu/g, and coercivity HC = 103.014 Oe [87]. Uniform
particles (side lengths = 0.3–1.3 µm). Chaudari, Kim et al. [83] re- mesoporous 3D hematite superstructures self-assembled using nano-
ported uniformly dispersed spheric hematite nanoparticles with an particles as building blocks have been successfully obtained in large
average particle size of 80–90 nm. Chen, Zhao et al. [84] synthesized quantities via an oleic acid-assisted one-pot solvothermal route [88].
porous α-Fe2O3 nanospheres (large mesopores (ca. 10 nm) in the core Lee and Kwak obtained tubular superstructures composed of super-
and small mesopores (< 4 nm) in the shell). Above mentioned samples paramagnetic α-Fe2O3 nanoparticles (around 10 nm) from pyrolysis of
were synthesized by a glycine-assisted hydrothermal method. Jayanthi metal − organic frameworks in a confined space [89].
et al. [85] reported weak-ferromagnetic properties of nanohematite In this work, we studied glycine-free and glycine-assisted hydro-
(particle size of aproximately 50 nm) with saturation magnetization thermal synthesis of hematite superstructures with intriguing particle
MS = 1.899 emu/g, remanent magnetization Mr = 0.659 emu/g, and morphologies, structures and magnetic properties. We found that the
coercivity HC = 414.98 Oe at 300 K. The changing shape of the hema- hydrothermal synthesis is a good method for the preparation of self-
tite nanoparticles from hexagonal microdisks to drum-like particles and assembled iron oxide hierarchical superstructures with controlled
spindles, the coercive force change from 942.5 to 647.5 Oe and morphology (mushroom-like, cube-like and sphere-like) and

429
D. Trpkov et al. Applied Surface Science 457 (2018) 427–438

Fig. 3. Structural, morphological and magnetic characterization of cube-like particles (sample S3, glycine-assisted hydrothermal reaction and 90 h reaction time): (a)
X-ray diffraction pattern, (b–d) SEM micrographs, and (e) magnetic field dependence of magnetization M(H) at 300 K.

dimensionality (1–5 μm). Magnetic measurements show that the parti- presented in Table 1. Ferric chloride hexahydrate (FeCl3·6H2O, p.a.
cles possess weak-ferromagnetic properties with coercivity in range purity ≥ 99%, Sigma-Aldrich, USA), glycine (p.a. purity 99.4%, Merck-
from 1305 Oe (mushroom-like morphology prepared by glycine-free Alkaloid, Macedonia), NH4OH (min 25%, Centrohem, Serbia), absolute
hydrothermal reaction) to 3725 Oe (sphere-like morphology prepared ethanol (min. 99.8, Zorka Pharma, Serbia) and deionised (DI) water
by glycine-assisted hydrothermal reaction). Both particle shapes show (18.2 MΩ, Purite Ltd. UK) were used. The solvents, additive and pre-
similar values for remanent (Mr ≈ 0.2 emu/g) and saturation magneti- cursor were mixed in proportion presented in Table 1 and stirred until
zation (MS ≈ 0.55 emu/g). The glycine in hydrothermal reactions in- the reagents were totally dissolved. The solutions were magnetically
fluences formation of samples with higher coercivity, more spherical stirred for 1 h at room temperature. Afterwards, the pH was measured
shape of the superstructures and their smaller sizes. showing values from 3 to 4 for all samples. The solutions were trans-
ferred into Teflon-lined stainless steel autoclaves of 33 ml capacity,
firmly sealed and heated at 140 °C for 45 or 90 h as presented in
2. Experimental section
Table 1. After the reaction finished, autoclaves were cooled naturally to
room temperature. The resulting slurry was transferred into 10 ml
2.1. Synthesis of hematite (α-Fe2O3)
plastic tubes, centrifuged, rinsed four times with absolute ethanol and
DI water, respectively. The products were obtained after drying at 60 °C
Hematite samples were synthesized by hydrothermal method. All
for 24 h. Finally, red solids were obtained and used for further analysis.
the reagents are of analytical grade and used without any purification.
Experimental conditions, solvents, additives and precursor are

430
D. Trpkov et al. Applied Surface Science 457 (2018) 427–438

Fig. 4. Structural, morphological and magnetic characterization of sphere-like particles (sample S4, glycine-assisted hydrothermal reaction and 45 h reaction time):
(a) X-ray diffraction pattern, (b–d) SEM micrographs, and (e) magnetic field dependence of magnetization M(H) at 300 K. Inset of Fig. 4(d) is crashed hematite
sphere-like particle.

2.2. Sample characterization spectrometer (IS 50) at room temperature in the wavenumber range of
4000–400 cm−1 with a resolution of 4 cm−1.
Crystal structure was analysed by X-ray powder diffraction (XRPD)
on a Rigaku RINT-TTRIII diffractometer operating with Cu Kα radiation. 3. Results and discussion
Measurements were performed in range 2θ = 10–80°. The particle
morphology and size were investigated using scanning electron mi- The structural, morphological and magnetic characterization of the
croscopy (SEM, JEOL JSM-7600F) and TEM (transmission electron sample S1 (glycine-free hydrothermal synthesis with reaction time
microscopy, JEOL 2010F). Morphology characterizations were eval- 45 h) is shown in Fig. 1. The structure and phase composition of the
uated with Matlab software (MathWorks) using circularity shape de- sample was determined by X-ray powder diffraction (XRPD) measure-
scriptor. Magnetic measurements were performed at vibrating-sample ments. The recorded and indexed diffraction peaks are presented in
magnetometer with the maximum field of 15 kOe at room temperature Fig. 1(a), where broad peaks can be observed, as expected for nanosized
(VSM, LakeShore 7400 Series). The Raman spectrum of the samples was particles. The positions of all the maxima coincide with the peaks
measured with a Horiba Jobin Yvon LabRAM HR spectrometer using characteristic of the hematite phase (JCPDS card 33-0664). No dif-
the 632.81 nm excitation line of a He-Ne laser. Fourier Transform fraction peaks corresponding to other iron oxide phases have been
Infrared (FTIR) spectra of samples were recorded on a Nicolet FTIR observed. Crystallite size (dXRPD) was estimated applying Scherrer

431
D. Trpkov et al. Applied Surface Science 457 (2018) 427–438

Fig. 5. Individual iron oxide microparticles of the samples (a) S1, (b) S2, (c) S3 and (d) S4 with corresponding black–white (binary) images after segmentation
procedure.

Fig. 6. TEM images of the mushroom-like particles (sample S1).

equation dXRPD = Kλ/β cos θ to (1 0 4) peak in the XRPD pattern, where S1 (Fig. 1(e)) were HC = 1305 Oe, Mr = 0.205 emu/g and
K = 0.9, λ = 1.5406 Å, β is FWHM (full width at half maximum) of MS = 0.592 emu/g. The structural, morphological and magnetic char-
(1 0 4) peak and θ is diffraction angle corresponding to peak maximum. acterization of the sample S2 is shown in Fig. 2. Sample S2 was also
Calculated crystallite size was dXRPD(S1) = 15.3 nm. The morphology synthesized with glycine-free hydrothermal reaction as sample S1 but
and particle sizes of the sample were studied in more detail by SEM. The reaction time for S1 was two times longer (90 h). The results of XRPD
SEM images (Fig. 1(b)–(d)) show uniform micrometer-sized complex measurements (Fig. 2(a)) show single-phase hematite structure in S2,
structures with diameter of DSEM(S1) ∼ 5 μm composed of self-as- with crystallite size dXRPD(S2) = 27.28 nm. Morphology of micro-
sembled nanoparticles as building blocks. The observed microparticles particles in sample S2 is distorted cube-like shape (Fig. 2 (b)–(d)), with
in Fig. 1(b–d) have uniform mushroom-like morphology. The M(H) hierarchical structures consisting of uniformly assembled nanoparticles
magnetization measurements of the sample S1 at 300 K is shown in as building blocks. The micrometer size DSEM(S2) ∼ 5 μm of hematite
Fig. 1(e)). The hysteresis curve is typical for a soft magnetic material superstructure was observed in the SEM micrographs (Fig. 2(b)–(d)).
and indicates weak-ferromagnetic properties of hematite [90]. The The measured coercivity HC = 1877 Oe (Fig. 2(e)) is higher in com-
measured coercivity, remanent and saturation magnetization of sample parison with S1 and slightly overcomes value for the bulk hematite

432
D. Trpkov et al. Applied Surface Science 457 (2018) 427–438

Fig. 7. TEM images of the distorted cube-like particles (sample S2).

Fig. 8. TEM images of the rounded cube-like particles (sample S3).

(HC = 1670 Oe) [90]. The sample S3 was synthesized by glycine-as- synthesis conditions of the sample S4 were in higher concentration of
sisted hydrothermal method (molar ratio Fe3+/glycine = 1) with same Fe3+ ions and glycine and shorter reaction time (45 h). The pure phase
reaction time as in the case of sample S2 (90 h). The single-phase he- of α-Fe2O3 is observed in the sample S4 (Fig. 4(a)) and the smallest
matite nanostructure can be observed in XRPD pattern (Fig. 3(a)). The crystallite size dXRPD(S4) = 15.08 nm which is similar as in the sample
crystallite size of the sample S3 was determined dXRPD(S3) = 24.98 nm, S1 (glycine-free hydrothermal reaction). The morphology of S4 micro-
which is slightly smaller than for sample S2. The morphology of the particle superstructures was spherical (DSEM(S4) ∼ 3 μm) with clearly
sample S3 was rounded cubic-like shapes (Fig. 3(b)–(d)). The observed visible nanoparticle composition (Fig. 4(b)–(d)). Sample S4 exhibited
microparticles (DSEM(S3) = 2.7 μm) are significantly smaller than in the the highest coercivity value HC = 3724 Oe, with Mr = 0.186 emu/g and
samples S1 and S2 (glycine-free hydrothermal synthesis) with nano- MS = 0.537 emu/g (Fig. 4(e)). Sample S4 broken microspheres (inset in
particle sub-units. Interestingly, the S3 microparticles possess much Fig. 4(d)) may explain formation mechanism of observed hierarchical
higher coercivity (Hc = 2947 Oe, Fig. 3(e)) in comparison to S1 and S2, nanoparticle superstructures. Formation mechanism probably consisted
with Mr = 0.195 emu/g and Ms = 0.554 emu/g. Modifications of the of two steps: nucleation and aggregation, as mentioned in previous

433
D. Trpkov et al. Applied Surface Science 457 (2018) 427–438

Fig. 9. TEM images of the sphere-like particles (sample S4).

Fig. 11. The coercivity as a function of the calculated crystallite diameter dXRPD
Fig. 10. The circularity (C) dependence of coercivity HC in hierarchical α-Fe2O3
related to two hydrothermal reaction times (red circles: t = 90 h, black squares:
nanostructures.
t = 45 h). (For interpretation of the references to colour in this figure legend,
the reader is referred to the web version of this article.)
Table 2
The morphology, crystallite size dXRPD, nanoparticle sub-units size dTEM, su-
reports, with similar morphologies and reaction conditions [91]. The
perstructure size DSEM and magnetic parameters of the samples S1-S4.
TEM studies of the samples showed microsized particles with nano-
Sample Morphology dXRPD dTEM DSEM HC [Oe] Mr Ms particle sub units in the range 20–80 nm as shown in Fig. 6 (sample S1),
[nm] [nm] [μm] [emu/g] [emu/g] 7 (sample S2), 8 (sample S3) and 9 (sample S4). The size distributions
S1 mushroom- 15.3 35 5.36 1305 0.205 0.592
α-Fe2O3 subparticles (nanoparticles) were obtained from an image
like analysis of the TEM micrographs measuring about 80 nanoparticles in
S2 distorted 27.28 40 5.2 1877 0.201 0.54 random fields of view. Since the α-Fe2O3 subparticles are not spherical
cube-like the area-equivalent diameter was calculated from the measured surface
S3 rounded 24.98 45 2.7 2947 0.195 0.554
area of the individual nanoparticles in the TEM images using the Di-
cube-like
S4 sphere-like 15.08 45 2.8 3724 0.186 0.537 gitalMicrograph™ software. The subparticles in the samples S2 and S3
are clearly elongated whereas subparticles in the sample S4 have irre-
gular shapes and large size distribution. The sample S1 are consisted of
Table 3 densely packed nanoparticles in the mushroom-like morphology and it
Area and circularity of the nanoparticles of the samples (Fig. 5). is difficult to measure single nanoparticles. A rough estimation of
subparticles sizes is around dTEM(S1) = 35 nm. The estimated sizes for
Sample S1 S2 S3 S4
other samples are dTEM(S2) = 40 ± 25 nm, dTEM(S3) = 45 ± 25 nm
Area (μm2) 17.5 21.4 4.9 7.6 and dTEM(S4) = 45 ± 25 nm. The corresponding selected area electron
Circularity (Cn=15) 0.4988 0.7534 0.8999 0.9675 diffraction (SAED) patterns are also presented in Figs. 6(d), 7(d), 8(d)

434
D. Trpkov et al. Applied Surface Science 457 (2018) 427–438

where μp, q (S ) represent normalized moments [94]. Such analysis has


been increasingly adopted in the literature as a quantitative description
of morphology [8,22,95–98]. In this paper we used the modified cir-
cularity measure
n
1
C n (S ) = ⎜⎛ ⎞

⎝ 2π (μ 2,0 (S ) + μ0,2 (S )) ⎠ (2)

which for a suitably chosen n ≥ 1 a minor changes in shape makes


numerically more visible [99]. Circularity measure determines how
much a particular particle shape differs from an ideal circle. This
measure provides the number which is from (0, 1]. For a maximum
value of 1, the particle shape is a perfect circle. Values closer to zero
means that shape very differs significantly from a perfect circle. Fig. 5
present isolated particles and images obtained after segmentation pro-
cedure. The correlation between the coercivity and the shapes (circu-
larity) of investigated superstructures is shown in Fig. 10. In general,
Fig. 12. The coercivity as a function of glycine quantity related to two hy-
drothermal reaction times (black squares: t = 90 h, red circles: t = 45 h). (For spherical shape reveals lower anisotropy and lower coercivity than
interpretation of the references to colour in this figure legend, the reader is other elongated shapes [100–103]. In this work, we observe that the
referred to the web version of this article.) coercivity increases with the circularity (Tables 2 and 3). Sample S4
posses the maximum value of the circularity (Cn=15(S4) = 0.9675,
sphere-like morphology) and coercivity (HC = 3725 Oe). Considering
Table 4 coercivity and particle shape anisotropy, we may conclude that the
Reported results of magnetic parameters for different hematite structures. shape anisotropy does not significantly influence the magnetic prop-
Morphology Ms [emu/g] Mr [emu/g] Hc [Oe] Reference erties (coercivity) of particles.
Fig. 11 presents coercivity as a function of calculated crystallite
Nanowires 13.18 0.06 32.27 [104] sizes of the samples. It should be noted that the hydrothermal reaction
Nanoflakes 78.36 0.1027 13.887 [105]
time of t = 45 h produced crystallites about two times smaller
Nanorods 0.343 0.085 483.29 [106]
Nanotubes 0.538 0.11 321.96 [106]
(∼15 nm) than obtained after t = 90 h (∼26 nm). The coercivity de-
Disks – 0.0178 942.5 [86] pendence on the glycine quantity is shown in Fig. 12. Evidently there is
Irregular particles – 0.0209 448.6 [86] a strong increase in the coercivity with addition of glycine in the hy-
Nanospindles – 0.0268 417 [86] drothermal reaction. The influence of the particle size, exchange/di-
Nanoplates – 0.1909 943.26 [21]
polar interactions, particle morphology and surface effects on the
Urchin-like 0.72 0.17 5300 [107]
Star-like – 0.569 156.08 [108] magnetic properties are reported in the literature [104–110]. It has
Dendrites – 0.05 224 [109] been demonstrated that the coercivity of iron oxides is mainly depen-
Nanocolumns 10.32 2.15 550 [110] dent on the magnetocrystalline anisotropy, shape anisotropy and sur-
Nanoplatelets 7.36 1.72 485 [110]
face anisotropy [2,111–117]. Table 4 presents literature reports of
magnetic parameters (HC, Mr and MS) for different hematite structures.
It should be noted that the hematite sphere-like particles presented in
and 9(d). The sample S1 show typical ring-shaped electron diffraction this work (Table 2) possess higher coercivity than in: hematite bulk,
pattern expected for structure composed of the nanoparticles with nanorods, nanotubes, nanotires, nanoplates, irregular, nanospindles,
random crystal orientations (Fig. 6(d)). The diffracted area of S2, S3 tubular superstructures, spherical 3D superstructures, clusters, hier-
and S4 (Figs. 7(d), 8(d) and 9(d)) show bright spots and less visible archical polyhedral structures and hexagonal microdisks (Table 4).
rings that are expected in samples with polycrystalline nanoparticles Moreover, the obtained coercivity (Table 2) is much larger than the
and particles with larger sizes. The diffraction rings are least visible in coercivity reported for individual α-Fe2O3 particles (∼20 times) [118].
the diffraction area of the sample S4 indicating well-crystallized par- The Raman spectrums of the samples are shown in Fig. 13. The
ticles (bigger grains) or possible oriented subparticles [92,93]. Particles spectrums are similar and typical of the α-Fe2O3 (hematite) phase,
can be polycrystalline, single crystal or amorphous. For example, a showing peaks at around 220, 288, 403, 490, 605, 654, 820, 1067, and
nanoparticle with 80 nm particle size can be made of: a single crystal 1306 cm−1. These peaks match well with the previously reported
(crystallite size ∼ particle size), many subparticles with grain sizes of Raman data for hematite [119–121]. The Raman peaks that appear at
about 10 nm (crystallite size ≪ particle size), or of amorphous material. around 220 cm−1 and 490 cm−1 have been assigned to the A1g mode,
The size of the crystallites is usually smaller than particle size observed and those at around 288 cm−1, 403 cm−1, and 605 cm−1 have been
by SEM or TEM microscope. These observations indicate that our na- assigned to the Eg modes [119–121]. The peak located at around
noparticles are polycrystalline. 654 cm−1 has been attributed to disorder effects and/or the presence of
Previously we proposed the circularity measure to quantitatively Fe2O3 nanocrystals, whereas the peak observed at around 1306 cm−1
describe morphological (shape) properties of synthesized particles. The has been assigned to hematite two-magnon scattering. The FTIR spec-
circularity measure (C) for an arbitrary shape (S) whose centroid co- trums of the samples were illustrated in Fig. 14. Two absorption peaks
incides with the origin is described by equation: at about 520 and 435 cm−1 are observed for all samples. These peaks
are corresponded to stretching and bending modes of the FeeO bond in
hematite [121–123]. No peaks corresponding to other phases (organic
1
C (S ) = or inorganic compounds) were observed, indicating pure hematite
2π (μ 2,0 (S ) + μ0,2 (S )) (1) phase.

435
D. Trpkov et al. Applied Surface Science 457 (2018) 427–438

Fig. 13. Raman spectrums of hematite superstructures: (a) S1, (b) S2, (c) S3 and (d) S4.

It has been reported for magnetic materials in nanosized form that 4. Conclusions
the controlled assembly of iron oxide nanoparticles enhanced the
magnetic properties [124,125]. It is generally believed that the coer- Hematite superstructures were obtained by glycine-free and glycine-
civity of magnetic materials is very sensitive to the assembly of the assisted hydrothermal synthesis. By applying different conditions and
magnetic nanoparticles in various structures. The origin of the magnetic surfactants into the hydrothermal reaction, hierarchical nanostructures
properties in the hierarchical magnetic nanoparticle superstructures is with various morphologies (mushroom-like, cube-like and sphere-like)
still controversy and subject of intensive research work. The hier- and sizes (2–6 μm) were selectively synthesized. The synthesized α-
archical magnetic nanoparticle superstructures are very complex for Fe2O3 superstructures are constructed of nanoparticle sub-units as
investigate thus it is very difficult to make clear conclusion about building blocks with a diameter in the range 20–80 nm. Therefore,
source of the magnetic properties. Namely, the magnetic properties presented synthetic methods provide direct solution growth of hier-
strongly depend on the size, size distribution, defect structures, shape archical nanostructures based on hematite nanoparticle building
(morphology) of the particles, surface effects, exchange and dipole blocks. The iron and glycine molar ratio and their concentrations effect
inter-particle interactions, surface-adsorbed ions, orientation of the the physical parameters, including morphology (more spherical shapes
subparticles, etc. It is known from the literature that the assembly of the are obtained in glycine-assisted hydrothermal synthesis), size (smaller
small and oriented subparticles into the superstructure results in the sizes are obtained in glycine-assisted hydrothermal synthesis), as well
change of the single domain to the multidomain, leading to the higher as magnetic properties. The hysteresis loops measured at 300 K reveal
remanent magnetization and coercivity [126–129]. The interaction ferromagnetic-like magnetic properties of the samples. The coercivity of
between hematite nanoparticles in close proximity can strongly influ- the samples synthesized by glycine-assisted hydrothermal method was
ence the magnetic properties of the α-Fe2O3 materials [130,131]. The much higher HC = 3725 Oe (sphere-like superstructure) than that of the
exchange and dipolar interactions in these structures may increase the samples synthesized by glycine-free hydrothermal reaction
effective anisotropy of the nanocrystals, leading to a “harder” reversal HC = 1305 Oe (mushroom-like superstructure). Hence, glycine in hy-
of the magnetization and an enhancement of coercivity [132]. We drothermal reaction strongly enhances coercivity value, up to 3 times.
propose that the magnetic coupling between the magnetic moments of The high values of HC may be attributed to the hierarchical super-
the oriented hematite subparticles results in enhanced coercivity va- structure, oriented hematite subparticles and exchange/dipolar inter-
lues. actions. The shape anisotropy and crystallite size dependence of coer-
civity did not dominate in the investigated samples.

436
D. Trpkov et al. Applied Surface Science 457 (2018) 427–438

Fig. 14. FTIR spectrums of hematite superstructures: (a) S1, (b) S2, (c) S3 and (d) S4.

Acknowledgement Surfaces B: Biointerf. 141 (2016) 537–545.


[14] A. Gholoobi, Z. Meshkat, K. Abnous, M. Ghayour-Mobarhan, M. Ramezani,
F. Homaei Shandiz, K.D. Verma, M. Darroudi, J. Mol. Struct. 1141 (2017)
The Ministry of Education and Science of the Republic of Serbia 594–599.
supported this work financially (Grant no. III 45015 and III 45005). The [15] A.K. Yury, E.L. Stanislav, M.R. Sergey, G.D. Gennadii, I.O. Elena, Mater. Res. Exp.
support by the Ministry of Higher Education, Science and Technology of 4 (2017) 035031.
[16] K. Davis, B. Cole, M. Ghelardini, B.A. Powell, O.T. Mefford, Langmuir 32 (2016)
the Republic of Slovenia within the National Research Program is ac- 13716–13727.
knowledged. We would also like to acknowledge Serbian-Slovenian [17] A.O. Baskakov, A.Y. Solov’eva, Y.V. Ioni, S.S. Starchikov, I.S. Lyubutin, I.I. Khodos,
bilateral project BI-RS/16-17-030 and Serbian-Slovakian bilateral pro- A.S. Avilov, S.P. Gubin, Appl. Surf. Sci. 422 (2017) 638–644.
[18] S. Hong, Z. Li, C. Li, C. Dong, S. Shuang, Appl. Surf. Sci. 427 (2018) 1189–1198.
ject SK-SRB 2016-0055.
[19] Q. Meng, K. Wang, Y. Tang, K. Zhao, G. Zhang, L. Zhao, J. Alloy. Compd. 722
(2017) 8–16.
References [20] A. Rufus, N. Sreeju, D. Philip, RSC Adv. 6 (2016) 94206–94217.
[21] H. Hao, D. Sun, Y. Xu, P. Liu, G. Zhang, Y. Sun, D. Gao, J. Colloid Interf. Sci. 462
(2016) 315–324.
[1] K. Isa, D.H. Rezagholipour, A. Mustafa, Mater. Res. Exp. 3 (2016) 095022. [22] M. Tadic, L. Kopanja, M. Panjan, S. Kralj, J. Nikodinovic-Runic, Z. Stojanovic,
[2] M. Iacob, D. Sirbu, C. Tugui, G. Stiubianu, L. Sacarescu, V. Cozan, A. Zelenakova, Appl. Surf. Sci. 403 (2017) 628–634.
E. Cizmar, A. Feher, M. Cazacu, RSC Adv. 5 (2015) 62563–62570. [23] U.C. Rajesh, V.S. Pavan, D.S. Rawat, RSC Adv. 6 (2016) 2935–2943.
[3] I. Karimzadeh, H.R. Dizaji, M. Aghazadeh, J. Magn. Magn. Mater. 416 (2016) [24] S. Krehula, M. Ristić, M. Reissner, S. Kubuki, S. Musić, J. Alloy. Compd. 695
81–88. (2017) 1900–1907.
[4] I. Karimzadeh, M. Aghazadeh, M.R. Ganjali, P. Norouzi, S. Shirvani-Arani, [25] S. Demirci, M. Yurddaskal, T. Dikici, C. Sarıoğlu, J. Hazard. Mater. 345 (2018)
T. Doroudi, P.H. Kolivand, S.A. Marashi, D. Gharailou, Mater. Lett. 179 27–37.
(2016) 5–8. [26] S. Nag, A. Roychowdhury, D. Das, S. Mukherjee, Mater. Res. Bull. 74 (2016)
[5] I. Karimzadeh, M. Aghazadeh, M.R. Ganjali, P. Norouzi, T. Doroudi, P.H. Kolivand, 109–116.
Mater. Lett. 189 (2017) 290–294. [27] M. Satheesh, A.R. Paloly, K.G. Suresh, M. Junaid Bushiri, J. Phys. Chem. Solids 98
[6] A. Zeleňáková, V. Zeleňák, J. Bednarčík, P. Hrubovčák, J. Kováč, J. Alloy. Compd. (2016) 247–254.
582 (2014) 483–490. [28] N.B. Kondrashova, A.S. Starostin, V.A. Val’tsifer, V.Y. Mitrofanov, S.A. Uporov,
[7] V. Zeleňák, D. Halamová, A. Zeleňáková, V. Girman, J. Porous Mater. 23 (2016) E. Bormashenko, Russ. J. Appl. Chem. 89 (2016) 1960–1968.
1633–1645. [29] M.V. Batygina, N.M. Dobrynkin, A.S. Noskov, Russ. J. Appl. Chem. 89 (2016)
[8] L. Kopanja, I. Milosevic, M. Panjan, V. Damnjanovic, M. Tadic, Appl. Surf. Sci. 362 1763–1768.
(2016) 380–386. [30] P. Brázda, J. Kohout, P. Bezdička, T. Kmječ, Cryst. Growth Des. 14 (2014)
[9] F. Pinakidou, M. Katsikini, K. Simeonidis, E. Kaprara, E.C. Paloura, M. Mitrakas, 1039–1046.
Appl. Surf. Sci. 360 (2016) 1080–1086. [31] G. Carraro, D. Barreca, C. Maccato, E. Bontempi, L.E. Depero, C. de Julian
[10] X. Zhu, J. Fan, Y. Zhang, H. Zhu, B. Dai, M. Yan, Y. Ren, J. Mater. Sci. 52 (2017) Fernandez, A. Caneschi, CrystEngComm 15 (2013) 1039–1042.
12717–12723. [32] K. Anil, S. Aditi, Nanotechnology 18 (2007) 475703.
[11] X. Chen, J. Wei, X. Wang, P. Wang, J. Alloy. Compd. 705 (2017) 138–145. [33] J.L. García-Muñoz, A. Romaguera, F. Fauth, J. Nogués, M. Gich, Chem. Mater. 29
[12] J.R. Jesus, J.G.S. Duque, C.T. Meneses, R.J.S. Lima, K.O. Moura, Ceram. Int. 44 (2017) 9705–9713.
(2018) 3585–3589. [34] S.S. Yakushkin, D.A. Balaev, A.A. Dubrovskiy, S.V. Semenov, K.A. Shaikhutdinov,
[13] F. You, G. Yin, X. Pu, Y. Li, Y. Hu, Z. Huang, X. Liao, Y. Yao, X. Chen, Colloids M.A. Kazakova, G.A. Bukhtiyarova, O.N. Martyanov, O.A. Bayukov, J. Supercond.

437
D. Trpkov et al. Applied Surface Science 457 (2018) 427–438

Novel Magn. 31 (2018) 1209–1217. [81] C.S. Biju, D.H. Raja, D.P. Padiyan, Chem. Phys. Lett. 610 (2014) 103–107.
[35] Y. Wang, J. Ma, S. Zuo-Jiang, K. Chen, Ceram. Int. 43 (2017) 16482–16487. [82] C.-Y. Yin, M. Minakshi, D.E. Ralph, Z.-T. Jiang, Z. Xie, H. Guo, J. Alloy. Compd.
[36] I. Shanenkov, A. Sivkov, A. Ivashutenko, Influence of oxygen concentration on 509 (2011) 9821–9825.
plasma dynamic synthesis product in Fe-O system, Solid State Phenom. 265 (2017) [83] N.K. Chaudhari, M.-S. Kim, T.-S. Bae, J.-S. Yu, Electrochim. Acta 114 (2013)
652–656. 60–67.
[37] C. Maccato, G. Carraro, D. Peddis, G. Varvaro, D. Barreca, Appl. Surf. Sci. 427 [84] H. Chen, Y. Zhao, M. Yang, J. He, P.K. Chu, J. Zhang, S. Wu, Anal. Chim. Acta 659
(2018) 890–896. (2010) 266–273.
[38] J. López-Sánchez, A. Muñoz-Noval, C. Castellano, A. Serrano, A.d. Campo, [85] S.A. Jayanthi, D.M.G.T. Nathan, J. Jayashainy, P. Sagayaraj, Mater. Chem. Phys.
M. Cabero, M. Varela, M. Abuín, J.D.l. Figuera, J.F. Marco, G.R. Castro, 162 (2015) 316–325.
O.R.d.l. Fuente, N. Carmona, J. Phys.: Condens. Matter. 29 (2017) 485701. [86] G. Xu, L. Li, Z. Shen, Z. Tao, Y. Zhang, H. Tian, X. Wei, G. Shen, G. Han, J. Alloy.
[39] J. López-Sánchez, A. Serrano, A. Del Campo, M. Abuín, O. Rodríguez de la Fuente, Compd. 629 (2015) 36–42.
N. Carmona, Chem. Mater. 28 (2016) 511–518. [87] N.V. Long, Y. Yang, C.M. Thi, L.H. Phuc, M. Nogami, J. Electron. Mater. 46 (2017)
[40] A. Sivkov, E. Naiden, A. Ivashutenko, I. Shanenkov, J. Magn. Magn. Mater. 405 3301–3308.
(2016) 158–168. [88] Z. Gan, A. Zhao, Q. Gao, M. Zhang, D. Wang, H. Guo, W. Tao, D. Li, E. Liu, R. Mao,
[41] S.S. Yakushkin, G.A. Bukhtiyarova, O.N. Martyanov, J. Struct. Chem. 54 (2013) RSC Adv. 2 (2012) 8681–8688.
876–882. [89] J. Lee, S.-Y. Kwak, Cryst. Growth Des. 17 (2017) 4496–4500.
[42] M. Tadic, I. Milosevic, S. Kralj, M. Mitric, D. Makovec, M.-L. Saboungi, L. Motte, [90] A.H. Hill, F. Jiao, P.G. Bruce, A. Harrison, W. Kockelmann, C. Ritter, Chem. Mater.
Nanoscale 9 (2017) 10579–10584. 20 (2008) 4891–4899.
[43] G.A. Bukhtiyarova, M.A. Shuvaeva, O.A. Bayukov, S.S. Yakushkin, [91] Z. An, J. Zhang, S. Pan, G. Song, Powder Technol. 217 (2012) 274–280.
O.N. Martyanov, J. Nanopart. Res. 13 (2011) 5527. [92] S. Anila, A. John, J.K. Thomas, S. Solomon, J. Mater. Sci.: Mater. Electron. 28
[44] L. Zhu, X. Zeng, X. Li, B. Yang, R. Yu, J. Magn. Magn. Mater. 426 (2017) 114–120. (2017) 18497–18507.
[45] M. Yoshikiyo, A. Namai, M. Nakajima, K. Yamaguchi, T. Suemoto, S.-I. Ohkoshi, J. [93] P. Li, X. Yan, J. Ji, Y. Wu, J. Hu, Y. Wang, H. Jiang, W. Zhang, CrystEngComm 19
Appl. Phys. 115 (2014) 172613. (2017) 1926–1932.
[46] A. Sawada, Y. Masubuchi, T. Motohashi, M. Sasaki, S. Kikkawa, Mater. Lett. 115 [94] J. Žunić, K. Hirota, P.L. Rosin, Pattern Recogn. 43 (2010) 47–57.
(2014) 198–200. [95] L. Kopanja, S. Kralj, D. Zunic, B. Loncar, M. Tadic, Ceram. Int. 42 (2016)
[47] Z. Su, J. Tao, J. Xiang, Y. Zhang, C. Su, F. Wen, Mater. Res. Bull. 84 (2016) 10976–10984.
445–448. [96] J. Zunic, K. Hirota, P.L. Rosin, J. Electron. Imaging 23 (2014) 029701.
[48] W. Wu, C.Z. Jiang, V.A.L. Roy, Nanoscale 8 (2016) 19421–19474. [97] J. Žunić, K. Hirota, Iberoamerican Congress on Pattern Recognition, Springer,
[49] A. Sood, V. Arora, J. Shah, R.K. Kotnala, T.K. Jain, Mater. Sci. Eng., C 80 (2017) Berlin, Heidelberg, 2008, pp. 94–101.
274–281. [98] D. Žunić, J. Žunić, Appl. Math. Comput. 226 (2014) 406–414.
[50] S. Karthi, G.A. Kumar, D.K. Sardar, G.C. Dannangoda, K.S. Martirosyan, [99] L. Kopanja, Z. Kovacevic, M. Tadic, M.C. Žužek, M. Vrecl, R. Frangež, Histochem.
E.K. Girija, Mater. Chem. Phys. 193 (2017) 356–363. Cell Biol. (2018), http://dx.doi.org/10.1007/s00418-018-1670-0.
[51] M.F. Casula, E. Conca, I. Bakaimi, A. Sathya, M.E. Materia, A. Casu, A. Falqui, [100] J.-G. Li, G. Fornasieri, A. Bleuzen, M. Gich, A. Gloter, F. Bouquet, M. Impéror-
E. Sogne, T. Pellegrino, A.G. Kanaras, PCCP 18 (2016) 16848–16855. Clerc, Small 12 (2016) 5981–5988.
[52] A.K. Patra, S.K. Kundu, A. Bhaumik, D. Kim, Nanoscale 8 (2016) 365–377. [101] D. Toulemon, Y. Liu, X. Cattoën, C. Leuvrey, S. Bégin-Colin, B.P. Pichon, Langmuir
[53] X. Mou, X. Wei, Y. Li, W. Shen, CrystEngComm 14 (2012) 5107–5120. 32 (2016) 1621–1628.
[54] Z.-H. Ruan, J.-H. Wu, J.-F. Huang, Z.-T. Lin, Y.-F. Li, Y.-L. Liu, P.-Y. Cao, Y.- [102] L. Wang, X. Lu, J. Wang, S. Yang, X. Song, J. Alloy. Compd. 681 (2016) 50–56.
P. Fang, J. Xie, G.-B. Jiang, J. Mater. Chem. A 3 (2015) 4595–4603. [103] S.-I. Ohkoshi, K. Imoto, A. Namai, S. Anan, M. Yoshikiyo, H. Tokoro, J. Am. Chem.
[55] A. Berendjchi, R. Khajavi, A.A. Yousefi, M.E. Yazdanshenas, Appl. Surf. Sci. 367 Soc. 139 (2017) 13268–13271.
(2016) 36–42. [104] N. Yahya, S. Qureshi, Z.u. Rehman, B. Alqasem, C. Fai Kait, J. Magn. Magn. Mater.
[56] L. Sun, X. Han, K. Liu, S. Yin, Q. Chen, Q. Kuang, X. Han, Z. Xie, C. Wang, 428 (2017) 469–480.
Nanoscale 7 (2015) 9416–9420. [105] B. Alqasem, N. Yahya, S. Qureshi, M. Irfan, Z. Ur Rehman, H. Soleimani, Mater.
[57] C. Hua, Y. Shang, Y. Wang, J. Xu, Y. Zhang, X. Li, A. Cao, Appl. Surf. Sci. 405 Sci. Eng., B 217 (2017) 49–62.
(2017) 405–411. [106] J. Jayashainy, P. Sagayaraj, J. Alloy. Compd. 626 (2015) 323–329.
[58] J.S. Cho, Y.J. Hong, J.-H. Lee, Y.C. Kang, Nanoscale 7 (2015) 8361–8367. [107] B. Cui, T. Li, J. Liu, M. Marinescu-Jasinski, Mater. Res. Express 2 (2015) 045011.
[59] B. Tian, J. Światowska, V. Maurice, S. Zanna, A. Seyeux, P. Marcus, Appl. Surf. Sci. [108] F. Song, J. Guan, X. Fan, G. Yan, J. Alloy. Compd. 485 (2009) 753–758.
353 (2015) 1170–1178. [109] F. Beshkar, H. Jahangiri, M. Mouasavi-Kamazani, J. Mater. Sci.: Mater. Electron.
[60] X. Qi, G. She, X. Huang, T. Zhang, H. Wang, L. Mu, W. Shi, Nanoscale 6 (2014) 27 (2016) 12869–12875.
3182–3189. [110] Q.-J. Sun, X.-G. Lu, G.-Y. Liang, Mater. Lett. 64 (2010) 2006–2008.
[61] M. Wang, M. Pyeon, Y. Gonullu, A. Kaouk, S. Shen, L. Guo, S. Mathur, Nanoscale 7 [111] T. Wen, Y. Li, D. Zhang, Q. Zhan, Q. Wen, Y. Liao, Y. Xie, H. Zhang, C. Liu, L. Jin,
(2015) 10094–10100. Y. Liu, T. Zhou, Z. Zhong, J. Colloid Interf. Sci. 497 (2017) 14–22.
[62] A. Subramanian, E. Gracia-Espino, A. Annamalai, H.H. Lee, S.Y. Lee, S.H. Choi, [112] E. Wetterskog, A. Castro, L. Zeng, S. Petronis, D. Heinke, E. Olsson, L. Nilsson,
J.S. Jang, Appl. Surf. Sci. 427 (2018) 1203–1212. N. Gehrke, P. Svedlindh, Nanoscale 9 (2017) 4227–4235.
[63] R. Mejías, S. Pérez-Yagüe, L. Gutiérrez, L.I. Cabrera, R. Spada, P. Acedo, C.J. Serna, [113] R. Moulin, G. Fornasieri, M. Impéror-Clerc, E. Rivière, P. Beaunier, A. Bleuzen,
F.J. Lázaro, Á. Villanueva, M.D.P. Morales, D.F. Barber, Biomaterials 32 (2011) ChemNanoMat 3 (2017) 833–840.
2938–2952. [114] K. Brymora, F. Calvayrac, J. Magn. Magn. Mater. 434 (2017) 14–22.
[64] K. Maier-Hauff, F. Ulrich, D. Nestler, H. Niehoff, P. Wust, B. Thiesen, H. Orawa, [115] A. Zeleňáková, V. Zeleňák, I. Mat'ko, M. Strečková, P. Hrubovčák, J. Kováč, J.
V. Budach, A. Jordan, J. Neuro-Oncol. 103 (2011) 317–324. Appl. Phys. 116 (2014) 033907.
[65] T.-H. Shin, Y. Choi, S. Kim, J. Cheon, Chem. Soc. Rev. 44 (2015) 4501–4516. [116] D. Peddis, C. Cannas, A. Musinu, A. Ardu, F. Orrù, D. Fiorani, S. Laureti,
[66] N.M. Abdul Rashid, C. Haw, W. Chiu, N.H. Khanis, A. Rohaizad, P. Khiew, S. Abdul D. Rinaldi, G. Muscas, G. Concas, G. Piccaluga, Chem. Mater. 25 (2013)
Rahman, CrystEngComm 18 (2016) 4720–4732. 2005–2013.
[67] A.S. Teja, P.-Y. Koh, Prog. Cryst. Growth Charact. Mater. 55 (2009) 22–45. [117] S. Singh, N. Khare, Appl. Surf. Sci. 364 (2016) 783–788.
[68] S. Xu, A.H. Habib, S.H. Gee, Y.K. Hong, M.E. McHenry, J. Appl. Phys. 117 (2015) [118] S. Okada, K. Takagi, K. Ozaki, Mater. Lett. 140 (2015) 135–139.
17A315. [119] Y.-S. Hu, A. Kleiman-Shwarsctein, A.J. Forman, D. Hazen, J.-N. Park,
[69] H. Xia, C. Hong, B. Li, B. Zhao, Z. Lin, M. Zheng, S.V. Savilov, S.M. Aldoshin, Adv. E.W. McFarland, Chem. Mater. 20 (2008) 3803–3805.
Funct. Mater. 25 (2015) 627–635. [120] A.M. Jubb, H.C. Allen, ACS Appl. Mater. Interf. 2 (2010) 2804–2812.
[70] J. Tomaszewska, S. Jakubiak, J. Michalski, W. Pronk, S.J. Hug, K.J. Kurzydłowski, [121] Y. Yan, H. Tang, J. Li, F. Wu, T. Wu, R. Wang, D. Liu, M. Pan, Z. Xie, D. Qu, J.
Appl. Surf. Sci. 366 (2016) 529–534. Colloid Interf. Sci. 495 (2017) 157–167.
[71] M. Cao, T. Liu, S. Gao, G. Sun, X. Wu, C. Hu, Z.L. Wang, Angew. Chem. Int. Ed. 44 [122] E. Darezereshki, Mater. Lett. 65 (2011) 642–645.
(2005) 4197–4201. [123] E.M. El Afifi, M.F. Attallah, E.H. Borai, J. Environ. Radioact. 151 (2016) 156–165.
[72] P. Basnet, G.K. Larsen, R.P. Jadeja, Y.-C. Hung, Y. Zhao, ACS Appl. Mater. Interf. 5 [124] M. Tadic, S. Kralj, M. Jagodic, D. Hanzel, D. Makovec, Appl. Surf. Sci. 322 (2015)
(2013) 2085–2095. 255–264.
[73] S. Lian, E. Wang, Z. Kang, Y. Bai, L. Gao, M. Jiang, C. Hu, L. Xu, Solid State [125] D. Toulemon, M.V. Rastei, D. Schmool, J.S. Garitaonandia, L. Lezama, X. Cattoen,
Commun. 129 (2004) 485–490. S. Begin-Colin, B.P. Pichon, Adv. Funct. Mater. 26 (2016) 2454–2462.
[74] Z. Jia, Q. Wang, D. Ren, R. Zhu, Appl. Surf. Sci. 264 (2013) 255–260. [126] J. Lian, X. Duan, J. Ma, P. Peng, T. Kim, W. Zheng, ACS Nano 3 (2009) 3749–3761.
[75] T. Mushove, T.M. Breault, L.T. Thompson, Ind. Eng. Chem. Res. 54 (2015) [127] L.-P. Zhu, H.-M. Xiao, S.-Y. Fu, Crystal Growth Design 7 (2007) 177–182.
4285–4292. [128] J. Ma, J. Lian, X. Duan, X. Liu, W. Zheng, J. Phys. Chem. C 114 (2010)
[76] C.-J. Jia, L.-D. Sun, F. Luo, X.-D. Han, L.J. Heyderman, Z.-G. Yan, C.-H. Yan, 10671–10676.
K. Zheng, Z. Zhang, M. Takano, J. Am. Chem. Soc. 130 (2008) 16968–16977. [129] Y. Zhu, W. Zhao, H. Chen, J. Shi, J. Phys. Chem. C 111 (2007) 5281–5285.
[77] R. Rajendran, Z. Yaakob, M. Pudukudy, M.S.A. Rahaman, K. Sopian, J. Alloy. [130] M. Tadić, D. Marković, V. Spasojević, V. Kusigerski, M. Remškar, J. Pirnat,
Compd. 608 (2014) 207–212. Z. Jagličić, J. Alloy. Compd. 441 (2007) 291–296.
[78] X. Su, C. Yu, C. Qiang, Appl. Surf. Sci. 257 (2011) 9014–9018. [131] M. Tadic, V. Kusigerski, D. Markovic, I. Milosevic, V. Spasojevic, J. Magn. Magn.
[79] T. Wang, S. Zhou, C. Zhang, J. Lian, Y. Liang, W. Yuan, New J. Chem. 38 (2014) Mater. 321 (2009) 12–16.
46–49. [132] B. Aslibeiki, P. Kameli, H. Salamati, J. Appl. Phys. 119 (2016) 063901.
[80] B. Wang, J.S. Chen, X.W.D. Lou, J. Mater. Chem. 22 (2012) 9466–9468.

438

View publication stats

You might also like