ArticlePDF Available

Composite materials based on a ceramic matrix of polycarbosilane and iron-containing nanoparticles

Authors:
  • State Atomic Energy Corporation Rosatom

Abstract

Composite materials comprised of a ceramic matrix with metal-containing nanoparticles were prepared by sintering iron (II) oxalate and polycarbosilane. The chemical composition of the material can be controlled by a sintering process. Sintering in inert atmosphere leads to reduction of the sample and metal iron formation (a-Fe, carbide). Formation of iron oxides requires calcination procedure in series (argon and air) for removing by products. The air-sintering materials consist mainly of oxide phases, but also contain metal iron. The prepared samples were characterized by the: SEM, TEM, XRD and EMR techniques and the Mössbauer spectroscopy. It was shown unique behavior that iron containing particles after calcination in air decreased from 5-30 nm to 2–5 nm due to interaction with matrix under air atmosphere.
Ceramics International 48 (2022) 37410–37422
Available online 13 September 2022
0272-8842/© 2022 Elsevier Ltd and Techna Group S.r.l. All rights reserved.
Composite materials based on a ceramic matrix of polycarbosilane and
iron-containing nanoparticles
G. Yu Yurkov
a
,
b
, D.A. Pankratov
c
, Yu.A. Koksharov
d
, Ye.A. Ovtchenkov
d
, A.V. Semenov
a
,
*
,
R.A. Korokhin
a
, G.I. Shcherbakova
e
, L.V. Gorobinskiy
b
, E.A. Burakova
f
, A.V. Korolkov
b
,
D.S. Ryzhenko
b
, V.I. Solodilov
a
a
N.N. Semenov Federal Research Center of Chemical Physics, Russian Academy of Sciences, 119334, Moscow, Russia
b
Bauman Moscow State Technical University, BMSTU, 2-nd Baumanskaya, 5, Moscow, 105005, Russia
c
Faculity of Chemistry, M.V. Lomonosov Moscow State University, 119991, Moscow, Russia
d
Faculity of Physics, M.V. Lomonosov Moscow State University, 119991, Moscow, Russia
e
State Research Institute for Chemistry and Technology of Organoelement Compounds, Moscow, 105005, Russia
f
Tambov State Technical University, Tambov, 392000, Russia
ARTICLE INFO
Keywords:
Chemical synthesis
Nanostructured materials
Metal matrix composite
ABSTRACT
Composite materials comprised of a ceramic matrix with metal-containing nanoparticles were prepared by
sintering iron (II) oxalate and polycarbosilane. The chemical composition of the material can be controlled by a
sintering process. Sintering in inert atmosphere leads to reduction of the sample and metal iron formation (a-Fe,
carbide). Formation of iron oxides requires calcination procedure in series (argon and air) for removing by
products. The air-sintering materials consist mainly of oxide phases, but also contain metal iron. The prepared
samples were characterized by the: SEM, TEM, XRD and EMR techniques and the M¨
ossbauer spectroscopy. It was
shown unique behavior that iron containing particles after calcination in air decreased from 5-30 nm to 25 nm
due to interaction with matrix under air atmosphere.
1. Introduction
Materials based on iron oxide are widely used in different applica-
tions. Several recent tendencies for applications on Fe
x
O
y
particles are:
environmental (as an example, removing of dyes from waste-water) [1],
agro chemistry (as a mineral fertilizing of cultivated plants) [2], medical
(boron-neutron capture therapy) [3], catalysis [4], composite materials
with controllable electrodynamic properties [5]. Iron-based complex
oxides or ferrites are the most important class of magnetic materials and
are widely discussed among researchers. The materials have got various
practical applications as magnets with high coercivity and remnant
magnetization [6]. The electrodynamic properties of the materials are
proven to be dependent both on the concentration of ferromagnetic
particles in the composite and their size, as well as on the properties of
the matrix [5]. The polymer composition is used for the creation of
nanosized ceramic which is durable at high temperatures with oxidation
tolerance. Ceramic shave next types: threads, matrix, complex special
barrier coatings, powders. The advantages of the application of polymer
precursors may be presented here: the negligible amount of uncontrolled
impurities, high adhesion on the edge bre-matrix, possibility of the
ceramic modelling of micro and macro structures on the stage of the
ceramic precursor, ability to adjust the physical form of the nal product
without ultra-high temperatures and pressures [7]. Several reports are
used polymer as a matrix: phenol-formaldehyde resin of the novolac
type [5], high- and low pressure polyethylene [8,9]. The iron oxide
nanoparticles can be produced by various techniques: hydrothermal
method [10,11], precipitation [11], ultrasound assisted green synthesis
[12], co-precipitation method [13,14], sonochemical method [15],
microwave-assisted synthesis [16], mechanochemical methods [17,18],
and others [19,20].
Silicon carbide is an important non-oxide ceramic which has diverse
industrial applications due to its outstanding properties, such as very
high hardness and strength, chemical, and thermal stability, high
melting point, oxidation resistance, high erosion resistance, excellent
thermal shock resistance. The material may be produced by pyrolysis
from preceramic polymer, for example, polycarbosilane [21,22]. The
* Corresponding author.
E-mail address: cemen9856@gmail.com (A.V. Semenov).
Contents lists available at ScienceDirect
Ceramics International
journal homepage: www.elsevier.com/locate/ceramint
https://doi.org/10.1016/j.ceramint.2022.09.096
Received 8 March 2022; Received in revised form 13 June 2022; Accepted 8 September 2022
Ceramics International 48 (2022) 37410–37422
37411
material based on polycarbosilane ceramic matrix can stabilize cobalt
ferrite nanoparticles [23].
The aim of this work is the possibility of preparing the material with
combine properties of silicon-based ceramic and magnetic properties of
iron oxide compounds under certain temperatures and annealing time.
Its expected to have a variety of potential applications, such as mag-
netic data storage, electromagnetic solutions, and so on.
2. Experimental
2.1. Ceramic preparation
The method for PCS synthesis was previously described in the liter-
ature [23]. It requires readily available precursors and typical chemical
laboratory equipment.
Polycarbosilane (PCS) as produced by "State Research Institute for
Chemistry and Technology of Organoelement Compounds". The process
lasts 3040 h. The rst stage is the decomposition and rearrangement of
polydimethylsilane (PDMS) at 350430С (with a controlled
temperature rise) and 0.4 MPa resulting in formation of raw PCS. The
second stage is the polycondensation at 390425С and residual pres-
sure of 0.20.4 kPa with removal of low-boiling components in accor-
dance with the scheme. PCS was used as a 50 % solution at toluene.
Other compounds were purchased with high chemical purity and used
without any additional purication.
Ceramic samples of nanoparticles of iron ferrite (Fe
3
O
4
) were pre-
pared by the thermal decomposition method. Solution of iron (II) oxa-
late at the water - C
2
H
5
OH mixture was added to the solution of PCS in
toluene (50 %) at 240C at argon atmosphere with the removal of sol-
vent and reaction by-products at ambient pressure. The ratio of FeC
2
O
4
to the PCS was 1:5 by weight correspondingly. The composite was
divided into four parts: unsintered one, sintered at 800C for 2 h in air,
sintered at 800C for 2 h in argon, the last part of the composite was
sintered at the same temperature and the same time in argon followed by
air. The temperature rate and time of sintering were chosen on the basis
Fig. 1. TEM micrographs of individual iron oxide nanoparticles in the ceramic
matrix before sintering.
Fig. 2. TEM micrograph nanosized iron oxide particles in the ceramic matrix
sintered in air.
Fig. 3. TEM micrograph nanosized iron oxide particles in the ceramic matrix
sintered in argon.
Fig. 4. SEM micrographs of individual iron oxide nanoparticles in the ceramic
matrix before sintering.
G.Y. Yurkov et al.
Ceramics International 48 (2022) 37410–37422
37412
of our research [23] where similar materials based on PCS and cobalt
ferrite (CoFe
2
O
4
) were synthesized.
2.2. Sample characterization
Transmission electron microscopy (TEM) was used to conrm the
presence of nanoparticles in the material and determine their sizes. The
studies were performed using a JEOL JEM 1011 microscope, with an
accelerating voltage of 80 kV and instrument constant of 80 cm.
SEM micrographs and EDS spectra of the samples prepared in this
study were recorded on a JEOL JSM 6380-LA scanning electron micro-
scope (with a 1 nA current and 20 kV accelerating voltage) equipped
with a JED 2300 energy dispersive X-ray analyzer.
X-ray diffraction patterns were recorded on a Shimadzu diffractom-
eter equipped with a GP-13 powder diffraction accessory and BSV21Cu
X-ray tube. The Cu K
α
1
(1.540598 Å) line was extracted from the radi-
ation spectrum using a Ni-lter. The PDF2 database by ICDD (2010) was
used as a reference.
M¨
ossbauer absorption spectra were recorded on a МS1104ЕM ex-
press M¨
ossbauer spectrometer produced by CJSC Kordon(Rostov-on-
Don). The source of γ-radiation was
57
Со in the metallic Rh matrix, with
the activity of 30 mCi, produced by RITVERC GmbH (Saint Petersburg).
Blended powder samples enclosed in a plastic cell were placed into a
vacuum cryostat. The spectra were recorded at room temperature (298
±3 K), and at temperature of liquid nitrogen (77.5 ±0.5 K). The
reproducibility of the spectra was ensured by comparing the spectra of
the same sample recorded at room temperature before vacuuming and
after the freezing-heating loop. The spectra were recorded with a noise/
signal ratio from 2.5 to 1.1%. Mathematical processing of M¨
ossbauer
spectra was performed for high-resolution spectra (1024 points) using
SpectRelax 2.8 software. The spectra were decomposed into symmetric
doublets and sextets with a controlled ratio between line widths and
intensities. The chemical shifts are provided relative to
α
-Fe.
EMR spectra were recorded at room temperature using a Varian-E4
X-band spectrometer. Magnetic measurements were performed using
EG&G PARC 155 vibrating sample magnetometer. The dependencies M
(H) were measured at room temperature in the elds up to 0.5 T.
3. Results and discussion
3.1. TEM
TEM micrograph of the unsintered sample made of iron ferrite and
ceramic matrix (Fig. 1) contains particles with sizes ranging from 5 to
Fig. 5. SEM micrograph nanosized iron oxide particles in the ceramic matrix
sintered in air.
Fig. 6. SEM micrograph nanosized iron oxide particles in the ceramic matrix
sintered in argon.
Fig. 7. Element composition of iron oxide nanoparticles in the ceramic matrix before sintering, according toTable 1.
G.Y. Yurkov et al.
Ceramics International 48 (2022) 37410–37422
37413
Fig. 8. Element composition of iron oxide nanoparticles in the ceramic matrix after sintering in air,according toTable 2.
Fig. 9. Element composition of iron oxide nanoparticles in the ceramic matrix after sintering in argon, according toTable 3.
Fig. 10. X-Ray diffraction pattern of iron oxide nanoparticles in the ceramic matrix before sintering.
G.Y. Yurkov et al.
Ceramics International 48 (2022) 37410–37422
37414
30 nm as well as larger then conglomerates formed by stacked particles.
TEM micrograph of the sample sintered in air (Fig. 2) contains
multiple small (25 nm) particles scattered uniformly throughout the
matrix as well as separate particles with a size up to 25 nm.
TEM micrograph of the sample sintered in argon contains nano-
particles with sizes ranging from 5 to 15 nm evenly dispersed in the
matrix (Fig. 3). It was shown a decrease in the average nanoparticle size
after sintering.
3.2. SEM
The TEM data comply with the morphology determined by SEM
(Figs. 46). As follows from X-ray spectral data (Figs. 79), the relative
content of Fe and C changes after sintering in argon or air. Sintering in
air results in the removal of the remaining solvent and low-molecular
materials from the blend as they are burned out. The molar ratio C/Fe
drastically reduces from 68 to 4 after sintering in air. The sintering in
Fig. 11. X-Ray diffraction pattern of iron oxide nanoparticles in the ceramic matrix after sinteringargon.
Fig. 12. X-Ray diffraction pattern of iron oxide nanoparticles in the ceramic matrix after sintering in air.
Fig. 13. X-Ray diffraction pattern of iron oxide nanoparticles in the ceramic matrix after sintering in argon and air.
G.Y. Yurkov et al.
Ceramics International 48 (2022) 37410–37422
37415
argon changes the C/Fe ratio in much less amount from 68 to 55.
3.3. XRD
X-ray diffraction is used to determine the composition of the samples.
X-ray diffraction patterns of the samples are shown (Figs. 1013) for
unsintered samples, sintered in different atmospheres: argon, argon
followed by sintering air, air respectively. The diffraction patterns of the
unsintered sample (Fig. 10) reveal the peaks attributable to Iron(III)
Oxalate - Fe
2
(C
2
O
4
)
3
, - Iron(II) Oxalate FeC
2
O
4
, Iron(II) Oxalate Hydrate
- FeC
2
O
4
*2H
2
O (ICDD PDF: 14762, 14807, 22635 respectively). It
indicates that the synthesis temperature is too lower for the total
decomposition of iron oxalate.
The X-ray pattern of the sample sintered in argon (Fig.11) contains
peaks attributable to different compounds: metal alpha iron
α
Fe;
austenite - solid solution of iron (with an alloying element carbon) Fe
+C; - iron silicide - Fe
3
Si; iron carbide - Fe
3
C (ICDD PDF: 6696,
31619, 35519, 35772). The chemical composition of the sample
shows that iron is reduced by the excess of the organic matrix, and is not
oxidized due to lack of oxygen.
The sample sintered in air (Fig.12) has some reduced form of iron.
The X-ray pattern of the sample shows several peaks correspond to the
different states of oxidation: iron -
α
Fe; magnetite Fe
3
O
4
; hematite
α
Fe
2
O
3
; Fayalite - Fe
2
SiO
4
(ICDD PDF: 6696, 19629, 33664, 34178).
One calcination procedure is not enough for complete oxidation, prob-
ably, due to kinetic factors.
The X-ray pattern of the sample sintered in argon followed by sin-
tering in air (Fig.13) contains peaks corresponding to the oxidized form
of iron: magnetite Fe
3
O
4
; and hematite
α
Fe
2
O
3
(ICDD PDF: 19629,
33664) [24]. Calcination in air of the sintered sample leads to deep
oxidation of the material.
3.4. M¨
ossbauer spectroscopy
M¨
ossbauer spectra of the unsintered sample have three resonance
lines at the two temperatures. The lines are different by intensities and
width. The intensities are strongly increased by the lowering of the
temperature. Such behavior is typical for metalorganic coordination
compounds with molecular crystal lattice [25,26].
The experimental spectrum can be satisfactorily described as a su-
perposition of two independent distribution functions of quadrupole
splitting and isomer shifts for doublets (Fig. 14). It corresponds to iron
atoms at different oxidation states (Table. 4). Iron(II) compounds at the
high spin state and octahedral oxygen environment [27] are described
by three modal distribution functions (Fig. 14). The most intense mode
of the function corresponds to a dehydrated form of iron(II) oxalate [28]
(Table. 4). The medium mode with a peak area around 7 % (Table 1)
obviously relates to original iron(II) oxalate dehydrate [29,30]. The
third mode with minimal area, probably, corresponds to the interme-
diate of the iron oxalate dehydration process [31]. Iron(III) compounds
[32] are presented by the second distribution - one, but asymmetric
mode with high dispersion. The mode describes initial products of
thermal decomposition of iron(II) oxalate [3335] which may be the
different hydrated form of iron(III) oxalate [36,37]. The presence of iron
oxalates in oxidation states of +2 and +3 are completely correlated with
the X-ray data of the sample described above.
M¨
ossbauer spectra of the sintered in argon sample have a large set of
resonance lines with different intensives (Fig. 15). The spectra can be
described as a superposition of ten subspectra of four different iron
contained phases. The three phases correspond to metal iron (Table. 5).
The most intense sextet attributes to metal iron -
α
-Fe [38]. The inner
part of the sextet is distorted by the presence of satellite subspectra iron
atoms at the lattice of alloy with composition:
α
-Fe
1-x
A
x
. The crystal
lattice of the alloy has iron atoms for which surrounded by 1, 2, or more
iron atoms by atoms of another element are replaced [3941]. Spectrum
parameters for these iron atoms are similar to subspectrumN1 param-
eters. The difference can be explained as a chemical shift and the
effective magnetic eld value. They are differenced from the original
values by n*dδ and -n*dH (where n is the number of the satellite, dF is
the value difference for M¨
ossbauer parameter F (Table. 5). The ratio of
satellite intense values is dened as binominal distribution dependent
on x (part of impurities) [42]. (Lower in Table 5 and Fig. 16 satellite with
an intense value higher than 0,1 % are shown only). In the present work,
the inuence of the changing number of iron atoms only at the rst
coordination sphere (it has eight atoms) is taken into account during the
experimental spectra investigation. Implementation of the model with a
second coordination sphere [43] (it contains 6 atoms) does not lead to a
decrease in normal functional at chi-square
χ
2
and does not change
parameter x (the parameter which describes the number of impurities) -
α
-Fe
0.8
A
0.2
. That is why the second sphere is not described in the current
investigation. It is possible to guess, that the alloy is a solid solution of Si
in metal Fe [44,45] -
α
-Fe
0.8
Si
0.2
. The occurrence of the iron silicide is
X-ray powder diffraction is shown as impurities to the main component
iron. It is possible that iron silicide is not an absolutely stoichiometric
compound, it has some defects with an ultimate form of a solution of Si
in Fe (see Table 4).
The third metal phase is shown as a high intense resonance line in the
spectrum center and it is shown as the singlet #10 (Table. 5). The metal
Fig. 14. M¨
ossbauer spectra and the distribution functions of the quadrupole
splitting at 78 and 298 K for the unsintered sample.
G.Y. Yurkov et al.
Ceramics International 48 (2022) 37410–37422
37416
phase corresponds to the metastable face-centered cubic lattice of metal
iron - γ-Fe [46,47].
The rest 40 % of the spectrum is described as a couple of sextets. They
describe iron atoms in the two different crystalline lattice positions of
θ-Fe
3
C [5,42,48].
M¨
ossbauer spectra of the sample sintered in argon and air and the
sample sintered only at air have complicated combined types (Fig. 16). A
higher number of the resonance line with smaller widths are occurred
for the sample sintered only at air compare to the sample sintered in
argon and air. However, a larger part of the spectra can be described by
components from the common phases. A subspectrum N1 related to
α
-Fe
2
O
3
[49,50] appears in the both samples (Tabl. 6). The higher width
of the resonance line and absence of Morin transition at hematite for the
sample sintered in series in argon and air is proof of high defectiveness
or the phase [38]. A subspectra N24 (Table 6) group describes solid
state solution magnetite-maghemite (Fe
3
O
4
- γ-Fe
2
O
3
) for the both
samples [51]. The solution can be presented as formula Fe
3-δ
O
4
, where
0 δ 1/3 is the coefcient of nonstoichiometry. The coefcient shows
the degree of the partial oxidation state of magnetite [38,52]. For the
low temperature spectra of the sample sintered only at air a higher value
of the isomer shifts of subspectra shows the smaller degree of the partial
Fig. 15. M¨
ossbauer spectra obtained at 298 K and 78 K of the sintered in argon sample and their model description with the numbering of subspectra according
to Table 5.
Table 1
Element composition of iron oxide nanoparticles in the ceramic matrix before
sintering.
Element/transition Energy, keV Weight, % Atomic, %
C K 0.277 62.20 74.73
O K 0.525 17.85 16.10
Si K 1.739 15.71 8.07
Fe K 6.398 4.24 1.10
Total, % 100.00 100.00
Table 2
Element composition of iron oxide nanoparticles in the ceramic matrix after
sintering in air.
Element/transition Energy, keV Weight, % Atomic, %
C K 0.277 11.60 18.63
O K 0.525 48.43 58.88
Si K 1.739 25.82 17.73
Fe K 6.398 13.76 4.75
Total,% 100.00 100.00
Table 3
Element composition of iron oxide nanoparticles in the ceramic matrix after
sintering in argon.
Element/transition Energy, keV Weight, % Atomic, %
C K 0.277 51.74 67.46
O K 0.525 16.23 15.89
Si K 1.739 27.67 15.43
Fe K 6.398 4.37 1.22
Total,% 100.00 100.00
G.Y. Yurkov et al.
Ceramics International 48 (2022) 37410–37422
37417
oxidation state. It is needed to relate this phase the doublet #7 (Tabl. 3)
too. The intensity of the doublet has strong temperature dependence.
This dependence is characteristic of the superparamagnetic state of iron
oxides.
A higher difference of the spectrum for the sample sintered in air is
the presence of subspectra correspond to the metal iron (
α
-Fe) and iron
silicate
α
-Fe
2
SiO
4
[53]. A smaller degree of the oxidation state of
magnetite may be explained by the existence of a kinetic barrier. The
barrier has occurred during the sintering procedure of iron oxalate in the
ceramic matrix. During the procedure, some gases (such as CO and
others) are involved in chemical reactions with iron compounds. The
reactions lead to the formation of iron with oxidation state +2 and
0 (metal iron). In contrast, the gas byproducts are removed on the rst
step (sintering in argon) during the sintering of the sample sintered in
argon and air. The next sintering in air leads to the oxidation of iron
compounds.
3.5. Magnetic properties
The eld dependence of magnetization M(H) measured at room
temperature also demonstrates that behaviour is different for the sam-
ples based on iron nanoparticles in the matrix before sintering, sintering
either in air or argon (Fig. 17). The sample before sintering is para-
magnetic. The sintered samples are ferromagnetic compounds. The co-
ercive force (
μ
0
H
c
) of the sample sintered in air is 0.01 T, and the
remnant magnetization (M
r
) is ca. 1.5 A*m
2
/kg. The saturation takes
place at magnetic induction B =0.5 T, the achieved saturation magne-
tization (Ms) is 9 A*m
2
/kg. The coercive force (
μ
0
H
c
) of the sample
sintered in argon is 0.01 T, and the remnant magnetization (M
r
) is ca. 3
A*m
2
/kg. A weak saturation of the M(H) curve for this sample may
indicate the presence of a large proportion of superparamagnetic
particles.
It is shown, that the magnetic properties of particles of Fe
3
O
4
are size
dependent. Nanoparticles of magnetite with sizes less than 20 nm are
superparamagnetic. The emergence of coercive force is shown for par-
ticles with sizes higher than 20 nm [54]. The Fe
3
O
4
particles with di-
mensions higher than 20 nm or other compounds may be in charge of
ferromagnetic behaviour in our case. The results of studying the eld
dependences of the magnetization M(H) are in good agreement with the
results obtained in M¨
ossbauer spectroscopy experiments. The data ob-
tained show that the unsintered sample contains mainly paramagnetic
iron phases, which are transformed into magnetic phases upon further
sintering. The sintering in argon produces large amounts of magnetic
particles.
3.6. EMR spectra
In order to obtain an analytical representation of the spectra, they
were decomposed into components described by the Tsallis function
(tsallian) [55,56].
Y’ ={2A/(q1)}{(2
q1
1) (HH
R
)/Γ
2
}{1+(2
q1
1) ((HH
R
)/Γ)
2
}
q/
(1q)
+
+{2A/(q1)}{(2
q1
1) (H +H
R
)/Γ
2
}{1+(2
q1
1) ((H +H
R
)/Γ)
2
}
q/
(1q)
H)
where A is the absorption line amplitude, H
R
is the resonance eld
strength, Γ is a relaxation rate parameter, q is a parameter which de-
termines the resonance line shape. Special cases of the tsallian are q =1
(gaussian) and q =2 (lorentzian). Lines with q >2 are called super-
lorentzians.
EMR spectra of three types of samples and decomposition into tsal-
lians are shown in (Figs. 1821).
The method used herein for the EMR spectra decomposition into
tsallians is described in more detail elsewhere [23]. It is based on
minimization of the discrepancy between experimental and theoretical
spectra using a hybrid technique combining gradient descent and
Monte-Carlo methods.
Fig. 19 contains spectra of the sintered in argon followed by sintering
in air. The spectra can be decomposed into 3 components. The narrowest
(ΔH
pp
190 Oe) one has g
eff
=1.99 and it likely corresponds to the
paramagnetic Fe
3+
ions. The wider component (ΔH
pp
680 Oe) with
g
eff
=2.02 is typical for small or amorphous nanoparticles of maghemite
(Fe
2
O
3
, γ-Fe
2
O
3
) (a maghemite phase with complex composition will
have a similar component). The most intense component (super-Lor-
entzian) has ΔH
pp
780 Oe) with g
eff
=2.42. It probably relates to
magnetite (Fe
3
O
4
) particles or the magnetite phase.
The sintered in air sample has higher magnetization for maghemite
and magnetite components. Decomposition of the EMR spectra leads to a
shift of components to lower elds Fig. 20.
The EMR spectrum of the sintered in argon sample has a new very
intense and broad (ΔH
pp
5660 Oe) component with high magnetiza-
tion (Fig. 21). The component can be explained by the presence of Fe (0).
The maghemite phase is presented also on spectra by line with small
intensity.
The chemical compositions of prepared samples are in line with
known data for stabilized iron oxide nanoparticles. For example, the
chemical composition of iron contained nanoparticles is complex after
sintering without oxygen. An iron citrate sample was interacted with a
portion of phenol-formaldehyde resin, heated without oxygen access at
800C. The resulted mix has several components: carbon, Fe
3
O
4
, Fe
3
C,
Fe different types [3]. In our case, all mentioned iron compounds are
presented with additional FeSi compound, after sintering in argon. FeSi
is presented because we used polycarbosylane with Si.
The chemical forms of iron in the sintered in air sample are by
M¨
ossbauerspectra:
α
-Fe (2 at%) also seen at X-ray powder diffraction,
FeSiO
4
(10 at%) - also seen at X-ray powder diffraction,
α
-Fe
2
O
3
(11 at
%) - also seen at X-ray powder diffraction, Fe
3
O
4
(75 at%) - also seen at
X-ray powder diffraction, Fe
+3
SPM
(2 at%) impurities.
The chemical forms of iron in the sintered argon followed by
Table 4
Parameters of the M¨
ossbauer spectra recorded at 78 and 298 K for unsintered sample.
Temperature, K Subspectrum Mode Phase δ
da
Δ
d
=2
ε
d
Γ S
d
ΣS
d
mm/s %
298 1 1 Fe
2
(C
2
O
4
)
3
xH
2
O 0.372 ±0.003 0.97 ±0.04 0.290 ±0.005 31.2 ±0.2 31.2 ±0.2
2 1 Fe(C
2
O
4
) 1.217 ±0.002 2.21 ±0.01 58.9 ±0.6 68.8 ±0.2
2 Fe(C
2
O
4
)2H
2
O 1.254 ±0.004 1.74 ±0.04 6.9 ±0.7
3 Fe(C
2
O
4
)xH
2
O 1.287 ±0.005 1.30 ±0.02 3.3 ±0.5
78 1 1 Fe
2
(C
2
O
4
)
3
xH
2
O 0.456 ±0.005 0.99 ±0.03 0.317 ±0.004 30.4 ±0.3 30.4 ±0.3
2 1 Fe(C
2
O
4
) 1.323 ±0.003 2.62 ±0.01 56.9 ±0.8 69.6 ±0.3
2 Fe(C
2
O
4
)2H
2
O 1.398 ±0.007 2.21 ±0.03 7.7 ±0.8
3 Fe(C
2
O
4
)xH
2
O 1.52 ±0.01 1.56 ±0.04 5.1 ±0.6
a
δ
d
- isomer shift, Δ
d
=2
ε
d
- quadrupole splitting and S
d
- relative area for maximum of mode distribution functions, Γ - line width.
G.Y. Yurkov et al.
Ceramics International 48 (2022) 37410–37422
37418
Table 5
M¨
ossbauer spectra parameters and iron phase distribution in sintered in argon sample.
Temperature,
K
78 298
# Phase δ
a
Dδ Δ =2
ε Г
exp
H
eff
dH S# Σ S# δ dδ Δ =2
ε Г
exp
H
eff
dH S# Σ S#
mm/s kOe % mm/s kOe %
1
α
-Fe 0.126 ±
0.001
0.006 ±
0.003
0.332 ±
0.006
338.6 ±
0.1
29.2 ±0.8 29.2 ±
0.8
0.003 ±
0.002
0.010 ±
0.005
0.39 ±
0.01
330.0 ±
0.2
27 ±1 27 ±1
2
α
-Fe
1-
x
A
x
0.126 ±
0.001
0.016 ±
0.003
0.01 ±0.01 0.32 ±
0.03
338.6 ±
0.1
14.4 ±
0.8
2.2 ±0.5 13.5 ±
0.6
0.003 ±
0.002
0.046 ±
0.005
0.05 ±
0.01
0.32 ±
0.03
330.0 ±
0.2
16
±1
3.2 ±0.9 16.5 ±
0.9
3 0.141 ±
0.003
324.2 ±
0.8
4.51 ±
0.03
0.049 ±
0.005
314.2 ±
1
5.84 ±
0.05
4 0.157 ±
0.007
310 ±2 3.98 ±
0.02
0.096 ±
0.009
298 ±2 4.62 ±
0.04
5 0.17 ±
0.01
295 ±2 2.00 ±
0.01
0.14 ±0.01 282 ±3 2.09 ±
0.02
6 0.19 ±
0.01
281 ±3 0.630 ±
0.004
0.19 ±0.02 266 ±4 0.589 ±
0.005
7 0.20 ±
0.02
267 ±4 0.127 ±
0.001
0.23 ±0.02 250 ±5 0.106 ±
0.001
8 θ-Fe
3
C 0.315 ±
0.001
0.008 ±
0.002
0.266 ±
0.006
252.6 ±
0.2
13.0 ±0.1 13.0 ±
0.1
0.200 ±
0.002
0.020 ±
0.003
0.37 ±
0.01
210.5 ±
0.4
24 ±1 24 ±1
9 0.315 ±
0.001
0.008 ±
0.002
0.394 ±
0.007
242.8 ±
0.2
25.9 ±0.1 25.9 ±
0.1
0.200 ±
0.002
0.020 ±
0.003
0.34 ±
0.02
197.4 ±
0.6
13 ±2 13 ±2
10 γ-Fe 0.039 ±
0.001
0.437 ±
0.004
18.4 ±0.1 18.4 ±
0.1
0.069 ±
0.002
0.512 ±
0.005
19.4 ±0.2 19.4 ±
0.2
x
b
0.20 ±0.02 0.18 ±0.02
a
δ – chemical shift; Δ =2
ε
quadrupole splitting;
Г
exp
line width; H
eff
hyperne magnetic eld; S site # area.
b
x - parameter for
α
-Fe
1-x
A
x
.
G.Y. Yurkov et al.
Ceramics International 48 (2022) 37410–37422
37419
Fig. 16. M¨
ossbauer spectra obtained at 78 K (a, c) and 298 K (b, d) of samples the sintered in Ar and air (a, b) or sintered in only air (c, d) and their model description
with the numbering of subspectra according to Table 6.
Table 6
M¨
ossbauer spectra parameters of samples the sintered in Ar and air or sintered in only air.
Temperature,К 78 298
Sample # Phase δ
a
Δ =2
ε Г
exp
H
eff
S# δ Δ =2
ε Г
exp
H
eff
S#
mm/s kOe % mm/s kOe %
Ar +air 1
α
-Fe
2
O
3
0.48 ±
0.01
0.03 ±0.01 0.42 ±
0.01
544.9 ±
0.4
19 ±1 0.40 ±
0.01
0.08 ±
0.01
0.45 ±
0.02
510.7 ±
0.4
12.2 ±
0.9
2 Fe
3-δ
O
4
0.49 ±
0.01
0.09 ±
0.01
0.48 ±
0.02
525.0 ±
0.7
25 ±2
3 0.37 ±
0.01
0.01 ±0.01 0.37 ±
0.01
507.1 ±
0.3
25 ±2 0.27 ±
0.01
0.00 ±0.01 0.49 ±
0.02
485.4 ±
0.2
31 ±3
4 0.42 ±
0.01
0.06 ±
0.02
0.82 ±
0.05
486 ±2 20 ±2 0.33 ±
0.01
0.03 ±
0.01
1.46 ±
0.09
455 ±4 40 ±2
7 Fe
+3
SPM
0.40 ±
0.02
1.19 ±0.04 1.32 ±
0.08
11.4 ±
0.4
0.34 ±
0.01
0.76 ±0.01 0.77 ±
0.05
17 ±2
Air 1
α
-Fe
2
O
3
0.49 ±
0.01
0.20 ±0.02 0.38 ±
0.02
537.3 ±
0.5
10.7 ±
0.6
0.38 ±
0.01
0.20 ±
0.01
0.31 ±
0.01
512.0 ±
0.1
19.6 ±
0.6
2 Fe
3-δ
O
4
0.48 ±
0.01
0.18 ±
0.01
0.31 ±
0.01
528.3 ±
0.2
19 ±1
3 0.44 ±
0.01
0.04 ±
0.01
0.63 ±
0.02
508.9 ±
0.7
20.8 ±
0.9
0.35 ±
0.01
0.16 ±
0.01
0.54 ±
0.02
491.5 ±
0.5
15.9 ±
0.8
4 0.53 ±
0.01
0.11 ±
0.02
1.36 ±
0.04
456 ±1 35 ±1 0.39 ±
0.01
0.10 ±
0.01
1.28 ±
0.03
452 ±1 33 ±1
5
α
-Fe 0.14 ±
0.01
0.04 ±
0.02
0.30 ±
0.01
339.4 ±
0.8
2.4 ±0.4 0.00 ±
0.01
0.07 ±
0.02
0.31 ±
0.01
328.9 ±
0.8
2.6 ±0.4
6
α
-Fe
2
SiO
4
1.28 ±
0.01
3.07 ±0.01 0.30 ±
0.01
10.2 ±
0.2
1.15 ±
0.01
2.80 ±0.01 0.35 ±
0.01
13.9 ±
0.2
7 Fe
+3
SPM
0.48 ±
0.02
0.85 ±0.03 0.47 ±
0.06
2.1 ±0.3 0.39 ±
0.01
0.91 ±0.01 0.45 ±
0.01
14.9 ±
0.2
a
δ – chemical shift; Δ =2
ε
quadrupole splitting;
Г
exp
line width; H
eff
hyperne magnetic eld; S site # area.
G.Y. Yurkov et al.
Ceramics International 48 (2022) 37410–37422
37420
sintering in air sample are by M¨
ossbauer spectra:
α
-Fe
2
O
3
(19 at%) - also
seen at X-ray powder diffraction and EMR spectra, Fe
3
O
4
(70 at%) - also
seen at X-ray powder diffraction and EMR spectra, Fe
+3
SPM
(11 at%)
impurities.
The observed reduction of iron during the heating in argon or air is
not unique for thermal treatment of iron nanoparticles. The formation of
metal iron and mostly iron carbide occurs during heating of iron nitrate
and chitosan matrix at 600 C. If the reaction mixture treated at lower
temperatures (300 500 C) than oxidized forms or iron are formed
[57].
The size of iron contained particles is dependent on the type of
particles and the way it is synthesized and stabilized. Several ranges of
unsintered iron oxide particles are known: 2090 nm [1], 1040 nm [2],
30150 nm [3]. The matrix stabilization leads to smaller particle size:
35 nm [4], 530 nm current job. It was shown in current work, that
calcination in air in matrix was able to produce very small particles
around 25 nm. The size of the particles is smaller than known ultrane
doped Fe
2
O
3
25 nm compare to 4,5 nm [5]. The samples sintered in air
or argon proved to be very stable and kept small particle sizes, in
contrast to resin matrix. The iron contained sample with resin matrix
after sintering in inert atmosphere has particle sizes 10100 nm with
average size of 6080 nm for iron compounds prepared from iron nitrate
[6].
Decreasing of the iron particle size during sintering in air is not
common for nanoparticles. It is known that iron nanoparticles increase
size from 4 nm to 25 nm after heating at 800C [57], but in our case the
particle size decrease to 25 nm. The reason for different nanoparticle
behaviour can be kinetic factors such as chemical reactions, mass and
heat transfer effects.
Fig. 17. Hysteresis loop M(H) recorded at 298 K for the sample comprised of
iron oxide nanoparticles in the ceramic matrix before sintering and after sin-
tering in either air or argon. Specic magnetization M is normalized to the
iron content.
Fig. 18. Experimental EMR spectra of the samples after sintering in air, argon,
or argon followed by sintering in air.
Fig. 19. Decomposition of the EMR spectrum of the sample sintered in argon
followed by sintering in air.
Fig. 20. Decomposition of the EMR spectrum of a sample after sintering in air.
G.Y. Yurkov et al.
Ceramics International 48 (2022) 37410–37422
37421
4. Conclusions
Methods for the preparation of composite materials based on iron
nanoparticles stabilized in the matrix by pyrolysis of iron precursor
polycarbosilane was worked out under certain temperatures and
annealing time. The size and product composition depend on the sin-
tering procedure. The thermal treatment, especially in argon, increases
remnant magnetization of the composites, evidenced by a larger hys-
teresis loop area. It was shown, that EMR spectra, X-ray diffraction
analysis are sensitive to structural changes and magnetic properties. The
methods can be recommended for samples investigation. It is shown that
sintering in argon leads to the reduction of iron compounds. The full
oxidation of iron to iron oxides needs at least two sintering procedures,
for example, sintering in argon followed by sintering in air; the one
sintering procedure in air is not enough for total oxidation of iron due to
kinetic factors (decomposition of the ceramic matrix provides an excess
of reducing agent such as hydrogen, carbon monoxide and so on ). The
completed work describes the thermal transformation of the composite
material in the different atmospheres, and the variation of the material
structure lets create a technology of composite material preparation
with designed particle formulation.
Preparing the material under different technological parameters will
be investigated.
Declaration of competing interest
The authors declare that they have no known competing nancial
interests or personal relationships that could have appeared to inuence
the work reported in this paper.
Acknowledgements
This work was performed as part of a State Task from the RF Ministry
of Science and Higher Education.
References
[1] D. Manojlovi´
c, K. Lelek, G. Rogli´
c, D. Zherebtsov, V. Avdin, K. Buskina,
C. Sakthidharan, S. Sapozhnikov, M. Samodurova, R. Zakirov, D.M. Stankovi´
c,
Efciency of homely synthesized magnetite: carbon composite anode toward
decolorization of reactive textile dyes, Int. J. Environ. Sci. Technol. 17 (2020),
https://doi.org/10.1007/s13762-020-02654-8.
[2] M.M. Anuchina, D.A. Pankratov, D.P. Abroskin, N.A. Kulikova, D.T. Gabbasova, D.
N. Matorin, D.S. Volkov, I.V. Perminova, Estimating the Toxicity and Biological
Availability for Interaction Products of Metallic Iron and Humic Substances, vol.
74, Moscow University Soil Science Bulletin, 2019, https://doi.org/10.3103/
S0147687419050028.
[3] D.I. Tishkevich, I.V. Korolkov, A.L. Kozlovskiy, M. Anisovich, D.A. Vinnik, A.
E. Ermekova, A.I. Vorobjova, E.E. Shumskaya, T.I. Zubar, S.V. Trukhanov, M.
V. Zdorovets, A.V. Trukhanov, Immobilization of boron-rich compound on Fe3O4
nanoparticles: stability and cytotoxicity, J. Alloys Compd. 797 (2019), https://doi.
org/10.1016/j.jallcom.2019.05.075.
[4] T.N. Rostovshchikova, M.S. Korobov, D.A. Pankratov, G.Y. Yurkov, S.P. Gubin,
Catalytic conversions of chloroolens over iron oxide nanoparticles 2.
Isomerization of dichlorobutenes over iron oxide nanoparticles stabilized on the
surface of ultradispersed poly(tetrauoroethylene), Russ. Chem. Bull. 54 (2005)
14251432, https://doi.org/10.1007/S11172-005-0422-1.
[5] D.S. Klygach, M.G. Vakhitov, D.A. Pankratov, D.A. Zherebtsov, D.S. Tolstoguzov,
Z. Raddaoui, S. el Kossi, J. Dhahri, D.A. Vinnik, A.V. Trukhanov, MCC: specic of
preparation, correlation of the phase composition and electrodynamic properties,
J. Magn. Magn Mater. 526 (2021), https://doi.org/10.1016/j.jmmm.2020.167694.
[6] A.V. Trukhanov, D.A. Vinnik, E.A. Tromov, V.E. Zhivulin, O.V. Zaitseva, S.
V. Taskaev, D. Zhou, K.A. Astapovich, S.V. Trukhanov, Y. Yang, Correlation of the
Fe content and entropy state in multiple substituted hexagonal ferrites with
magnetoplumbite structure, Ceram. Int. 47 (2021), https://doi.org/10.1016/j.
ceramint.2021.03.088.
[7] Z. Ren, C. Gervais, G. Singh, Fabrication and characterization of silicon
oxycarbidebre-mats via electrospinning for high temperature applications, RSC
Adv. 10 (2020), https://doi.org/10.1039/D0RA04060F.
[8] A.S. Fionov, G.Yu Yurkov, V.v. Kolesov, D.A. Pankratov, E.A. Ovchenkov,
YuA. Koksharov, Composite material based on iron-containing nanoparticles for
applications in the problems of electromagnetic compatibility, J. Commun.
Technol. Electron. 57 (2012), https://doi.org/10.1134/S1064226912040079.
[9] G.Y. Yurkov, A.S. Fionov, A.V. Kozinkin, Y.A. Koksharov, Y.A. Ovtchenkov, D.
A. Pankratov, O.V. Popkov, V.G. Vlasenko, Y.A. Kozinkin, M.I. Biryukova, N.
A. Taratanov, V.M. Bouznik, Synthesis and physicochemical properties of
composites for electromagnetic shielding applications: a polymeric matrix
impregnated with iron- or cobalt-containing nanoparticles, J. Nanophotonics 6
(2012), https://doi.org/10.1117/1.JNP.6.061717.
[10] P. Bhavani, C.H. Rajababu, M.D. Arif, I.V.S. Reddy, N.R. Reddy, Synthesis of high
saturation magnetic iron oxide nanomaterials via low temperature hydrothermal
method, J. Magn. Magn Mater. 426 (2017), https://doi.org/10.1016/j.
jmmm.2016.09.049.
[11] A. Lassoued, M.S. Lassoued, B. Dkhil, S. Ammar, A. Gadri, Synthesis,
photoluminescence and Magnetic properties of iron oxide (
α
-Fe2O3) nanoparticles
through precipitation or hydrothermal methods, Phys. E Low-dimens. Syst.
Nanostruct. 101 (2018), https://doi.org/10.1016/j.physe.2018.04.009.
[12] K. Sathya, R. Saravanathamizhan, G. Baskar, Ultrasound assisted phytosynthesis of
iron oxide nanoparticle, Ultrason. Sonochem. 39 (2017), https://doi.org/10.1016/
j.ultsonch.2017.05.017.
[13] A. Lazzarini, R. Colaiezzi, M. Passacantando, F. DOrazio, L. Arrizza, F. Ferella,
M. Crucianelli, Investigation of physico-chemical and catalytic properties of the
coating layer of silica-coated iron oxide magnetic nanoparticles, J. Phys. Chem.
Solid. 153 (2021), https://doi.org/10.1016/j.jpcs.2021.110003.
[14] K. Petcharoen, A. Sirivat, Synthesis and characterization of magnetite
nanoparticles via the chemical co-precipitation method, Mater. Sci. Eng., B 177
(2012), https://doi.org/10.1016/j.mseb.2012.01.003.
[15] A. Hassanjani-Roshan, M.R. Vaezi, A. Shokuhfar, Z. Rajabali, Synthesis of iron
oxide nanoparticles via sonochemical method and their characterization,
Particuology 9 (2011), https://doi.org/10.1016/j.partic.2010.05.013.
[16] T.A. Lastovina, A.P. Budnyk, S.P. Kubrin, A.v. Soldatov, Microwave-assisted
synthesis of ultra-small iron oxide nanoparticles for biomedicine, Mendeleev
Commun. 28 (2018), https://doi.org/10.1016/j.mencom.2018.03.019.
[17] P.A. Calder´
on Bedoya, P.M. Botta, P.G. Bercoff, M.A. Fanovich, Magnetic iron
oxides nanoparticles obtained by mechanochemical reactions from different solid
precursors, J. Alloys Compd. 860 (2021), https://doi.org/10.1016/j.
jallcom.2020.157892.
[18] B. Medina, M.G. Verd´
erioFressati, J.M. Gonçalves, F.M. Bezerra, F.A. Pereira
Scacchetti, M.P. Mois´
es, A. Bail, R.B. Samulewski, Solventless preparation of
Fe3O4 and Co3O4 nanoparticles: a mechanochemical approach, Mater. Chem.
Phys. 226 (2019), https://doi.org/10.1016/j.matchemphys.2019.01.043.
[19] A.G. Roca, L. Guti´
errez, H. Gavil´
an, M.E. Fortes Brollo, S. Veintemillas-Verdaguer,
M. del P. Morales, Design strategies for shape-controlled magnetic iron oxide
nanoparticles, Adv. Drug Deliv. Rev. 138 (2019), https://doi.org/10.1016/j.
addr.2018.12.008.
[20] C. Song, W. Sun, Y. Xiao, X. Shi, Ultrasmall iron oxide nanoparticles: synthesis,
surface modication, assembly, and biomedical applications, Drug Discov. Today
24 (2019), https://doi.org/10.1016/j.drudis.2019.01.001.
[21] S. Kaur, R. Riedel, E. Ionescu, Pressureless fabrication of dense monolithic SiC
ceramics from a polycarbosilane, J. Eur. Ceram. Soc. 34 (2014), https://doi.org/
10.1016/j.jeurceramsoc.2014.05.002.
[22] F. Süß, T. Schneider, M. Frieß, R. Jemmali, F. Vogel, L. Klopsch, D. Koch,
Combination of PIP and LSI processes for SiC/SiC ceramic matrix composites, Open
Ceramics 5 (2021), https://doi.org/10.1016/j.oceram.2021.100056.
[23] G.Y. Yurkov, K.A. Shashkeev, S.v. Kondrashov, O.v. Popkov, G.I. Shcherbakova, D.
v. Zhigalov, D.A. Pankratov, E.A. Ovchenkov, Y.A. Koksharov, Synthesis and
magnetic properties of cobalt ferrite nanoparticles in polycarbosilane ceramic
Fig. 21. Decomposition of the EMR spectrum of a sample after sintering
in argon.
G.Y. Yurkov et al.
Ceramics International 48 (2022) 37410–37422
37422
matrix, J. Alloys Compd. 686 (2016) 421430, https://doi.org/10.1016/J.
JALLCOM.2016.06.025.
[24] E.M. Kostyukhin, Synthesis of magnetite nanoparticles upon microwave and
convection heating, Russ. J. Phys. Chem. A 92 (2018), https://doi.org/10.1134/
S0036024418120233.
[25] D.A. Pankratov, V.D. Dolzhenko, R.A. Stukan, Y.F. al Ansari, E.v. Savinkina, Y.
M. Kiselev, Investigation of iron(III) complex with crown-porphyrin, Hyperne
Interact. 222 (2013), https://doi.org/10.1007/s10751-011-0378-5.
[26] L.I. Rodionova, N.E. Borisova, A.v. Smirnov, V.v. Ordomsky, A.A. Moiseeva, D.
A. Pankratov, Binuclear iron complexes with acyclic Schiff bases based on 4-tert-
butyl-2,6-diformylphenol: synthesis, properties, and use in catalytic partial
oxidation of isobutane, Russ. Chem. Bull. 62 (2013), https://doi.org/10.1007/
s11172-013-0164-4.
[27] D.A. Pankratov, M¨
ossbauer study of oxo derivatives of iron in the Fe2O3-Na2O2
system, Inorg. Mater. 50 (2014), https://doi.org/10.1134/S0020168514010154.
[28] M. Hermanek, R. Zboril, M. Mashlan, L. Machala, O. Schneeweiss, Thermal
behaviour of iron(ii) oxalate dihydrate in the atmosphere of its conversion gases,
J. Mater. Chem. 16 (2006), https://doi.org/10.1039/b514565a.
[29] M.C. DAntonio, A. Wladimirsky, D. Palacios, L. Coggiolaa, A.C. Gonz´
alez-Bar´
o, E.
J. Baran, R.C. Mercader, Spectroscopic investigations of iron(II) and iron(III)
oxalates, J. Braz. Chem. Soc. 20 (2009), https://doi.org/10.1590/S0103-
50532009000300006.
[30] M.A. Gabal, Non-isothermal decomposition of NiC2O4FeC2O4 mixture aiming at
the production of NiFe2O4, J. Phys. Chem. Solid. 64 (2003), https://doi.org/
10.1016/S0022-3697(03)00163-X.
[31] D. Xue, F. Li, Y. Kong, J. Yang, Phase transformation of FeC2O4 2H2O heat
treated in, J. Phys. Chem. Solid. 57 (1996), https://doi.org/10.1016/0022-3697
(95)00249-9.
[32] F.F.H. Arag´
on, J.A.H. Coaquira, S.W. da Silva, R. Cohen, D.G. Pacheco-Salazar, L.C.
C.M. Nagamine, Fe content effects on structural, electrical and magnetic properties
of Fe-doped ITO polycrystalline powders, J. Alloys Compd. 867 (2021), https://
doi.org/10.1016/j.jallcom.2021.158866.
[33] M.A. Gabal, A.A. El-Bellihi, S.S. Ata-Allah, Effect of calcination temperature on Co
(II) oxalate dihydrateiron(II) oxalate dihydrate mixture, Mater. Chem. Phys. 81
(2003), https://doi.org/10.1016/S0254-0584(03)00137-8.
[34] P.K. Gallagher, C.R. Kurkjian, A study of the thermal decomposition of some
complex oxalates of iron(III) using the M¨
ossbauer effect, Inorg. Chem. 5 (1966),
https://doi.org/10.1021/ic50036a013.
[35] B.S. Randhawa, A. Pal Singh, R.P. Sharma, P.S. Bassi, Comparative study of solid
state reactivity between ferrous oxalate dihydrate and unsubstituted/substituted
aniline hydrochlorides, J. Radioanal. Nucl. Chem. 128 (1988), https://doi.org/
10.1007/BF02166646.
[36] H. Ahouari, G. Rousse, J. Rodríguez-Carvajal, M.-T. Sougrati, M. Sauban`
ere,
M. Courty, N. Recham, J.-M. Tarascon, Unraveling the structure of iron(III) oxalate
tetrahydrate and its reversible Li insertion capability, Chem. Mater. 27 (2015),
https://doi.org/10.1021/cm5043149.
[37] P.Yu Tyapkin, S.A. Petrov, A.P. Chernyshev, A.I. Ancharov, L.A. Sheludyakova, N.
F. Uvarov, Structural features of hydrate forms of iron(III) oxalate, J. Struct. Chem.
57 (2016), https://doi.org/10.1134/S0022476616060111.
[38] D.A. Pankratov, M.M. Anuchina, Nature-inspired synthesis of magnetic non-
stoichiometric Fe3O4 nanoparticles by oxidative in situ method in a humic
medium, Mater. Chem. Phys. 231 (2019), https://doi.org/10.1016/j.
matchemphys.2019.04.022.
[39] W.E. Sauer, R.J. Reynik, Electronic and magnetic structure of dilute iron-base
alloys, J. Appl. Phys. 42 (1971), https://doi.org/10.1063/1.1660357.
[40] D. Grudinskyv, D. Zinoveev, A. Kondratiev, D. Lubyanoi, D. Pankratov, Reductive
smelting of neutralized red Mud for iron recovery and produced pig iron for heat-
resistant castings, Metals 10 (2019), https://doi.org/10.3390/met10010032.
[41] I. Vincze, I.A. Campbell, Mossbauer measurements in iron based alloys with
transition metals, J. Phys. F Met. Phys. 3 (1973), https://doi.org/10.1088/0305-
4608/3/3/023.
[42] P. Grudinsky, D. Zinoveev, D. Pankratov, A. Semenov, M. Panova, A. Kondratiev,
A. Zakunov, V. Dyubanov, A. Petelin, Inuence of sodium sulfate addition on iron
grain growth during carbothermic roasting of red Mud samples with different
basicity, Metals 10 (2020), https://doi.org/10.3390/met10121571.
[43] S.M. Dubiel, J. Cieslak, Short-range order in iron-rich Fe-Cr alloys as revealed by
M¨
ossbauer spectroscopy, Phys. Rev. B 83 (2011), https://doi.org/10.1103/
PhysRevB.83.180202.
[44] M.B. Stearns, Internal magnetic elds, isomer shifts, and relative abundances of the
various Fe sites in FeSi alloys, Phys. Rev. 129 (1963), https://doi.org/10.1103/
PhysRev.129.1136.
[45] H. Tokoro, S. Fujii, Y. Kobayashi, S. Muto, The growth of carbon coating layers on
iron particles and the effect of alloying the iron with silicon, J. Alloys Compd. 509
(2011), https://doi.org/10.1016/j.jallcom.2010.10.124.
[46] J. Ding, H. Huang, P.G. McCormick, R. Street, Magnetic properties of martensite-
austenite mixtures in mechanically milled 304 stainless steel, J. Magn. Magn
Mater. 139 (1995), https://doi.org/10.1016/0304-8853(95)90034-9.
[47] K. Haneda, Z.X. Zhou, A.H. Morrish, T. Majima, T. Miyahara, Low-temperature
stable nanometer-size fcc-Fe particles with no magnetic ordering, Phys. Rev. B 46
(1992), https://doi.org/10.1103/PhysRevB.46.13832.
[48] G. G.LeCaer, J.M. Dubois, J.P. Senateur, Etude par spectrom´
etrieM¨
ossbauer des
carbures de Fer Fe3C et Fe5C2, J. Solid State Chem. 19 (1976), https://doi.org/
10.1016/0022-4596(76)90145-6.
[49] S.J. Oh, D.C. Cook, H.E. Townsend, Characterization of iron oxides commonly
formed as corrosion products on steel, Hyperne Interact. 112 (1998), https://doi.
org/10.1023/A:1011076308501.
[50] A.Yu Romanchuk, S.N. Kalmykov, A.v. Egorov, Y.v. Zubavichus, A.A. Shiryaev, O.
N. Batuk, S.D. Conradson, D.A. Pankratov, I.A. Presnyakov, Formation of
crystalline PuO2+nH2O nanoparticles upon sorption of Pu(V,VI) onto hematite,
Geochem. Cosmochim. Acta 121 (2013), https://doi.org/10.1016/j.
gca.2013.07.016.
[51] D.A. Pankratov, M.M. Anuchina, F.M. Spiridonov, G.G. Krivtsov, Fe3 – δO4
nanoparticles synthesized in the presence of natural polyelectrolytes, Crystallogr.
Rep. 65 (2020), https://doi.org/10.1134/S1063774520030244.
[52] G. Niraula, J.A.H. Coaquira, F.H. Aragon, B.M. GaleanoVillar, A. Mello, F. Garcia,
D. Muraca, G. Zoppellaro, J.M. Vargas, S.K. Sharma, Tuning the shape, size, phase
composition and stoichiometry of iron oxide nanoparticles: the role of phosphate
anions, J. Alloys Compd. 856 (2021), https://doi.org/10.1016/j.
jallcom.2020.156940.
[53] I. Choe, R. Ingalls, J.M. Brown, Y. Sato-Sorensen, M¨
ossbauer studies of iron silicate
spinel at high pressure, Phys. Chem. Miner. 19 (1992), https://doi.org/10.1007/
BF00202313.
[54] S. Klomp, C. Walker, M. Christiansen, B. Newbold, D. Griner, Y. Cai, P. Minson,
J. Farrer, S. Smith, B.J. Campbell, R.G. Harrison, K. Chesnel, Size-dependent
crystalline and magnetic properties of 5100 nm Fe O nanoparticles:
superparamagnetism, verwey transition, and FeOFe O coreshell formation,
IEEE Trans. Magn. 56 (2020), https://doi.org/10.1109/TMAG.2020.3018154.
[55] D.F. Howarth, J.A. Weil, Z. Zimpel, Generalization of the lineshape useful in
magnetic resonance spectroscopy, J. Magn. Reson. 161 (2003), https://doi.org/
10.1016/S1090-7807(02)00195-7.
[56] YuA. Koksharov, Application of Tsallis functions for analysis of line shapes in
electron magnetic resonance spectra of magnetic nanoparticles, Phys. Solid State
57 (2015) 20112015, https://doi.org/10.1134/S1063783415100121.
[57] A.A. Vasilev, D.G. Muratov, G.N. Bondarenko, E.L. Dzidziguri, M.N. Emov, G.
P. Karpacheva, Synthesis of iron and cobalt nanoparticles in an IR-pyrolyzed
chitosan matrix, Russ. J. Phys. Chem. A 92 (2018), https://doi.org/10.1134/
S0036024418100369.
G.Y. Yurkov et al.
... Mössbauer spectroscopy (spectrometer MS1104EM, Cordon, Rostovon-Don, Russia) was used to analyze the compositions of the initial tablets and those exposed to moist incubation for 1, 3, and 5 months were analyzed using. The analysis involved the acquisition of Mössbauer spectra at both room temperature (296 K) and low temperature (78 K), following established procedures from our previous studies (Yurkov et al., 2016(Yurkov et al., , 2022Grudinsky et al., 2020;Valeev et al., 2021). To gain a more complete understanding of the material composition and corrosion products, including those lacking long-range order (X-ray amorphous), we used low-temperature Mössbauer spectroscopy. ...
... In contrast, the nano-iron-biochar composite exhibited a more diverse Fig. 2. Mössbauer spectra at 296 K (a, c, e) and 78 К (b, d, f) illustrating the initial nano-iron-biochar (Cn0, a, b), its corrosion products after one month of humidity exposure in cellulose (Cn1, c, d) and in peat (Pn1, e, f). composition, containing α-Fe alongside γ-Fe (Yurkov et al., 2022), and cementite (θ-Fe 3 C) (Cn0, Fig. 2a and b) (Yurkov et al., 2016). The presence of γ-Fe and cementite likely results from the reduction of iron during synthesis in the presence of excess carbon. ...
... The ability to use the high strength of thin fibrous reinforcing fillers under various types of external influences appears to be due to the adhesive strength at which external loads are transferred to the fibers [21][22][23][24][25][26][27]. ...
Article
Full-text available
The pull-out method was used to study the adhesive strength τ of “fiber–thermoset” systems with wide variations in area. Studied binders were based on resins that had different chemical natures (epoxy, epoxy phenol, orthophthalic, polyphenylsiloxane, and phenol–formaldehyde). Shear adhesive strength was determined for systems with two fiber types (glass and steel fibers). It was shown that strength τ depended on scale (area). Formation of τ occurred during the curing process and the system’s subsequent cooling to the measurement temperature T. It was found that interface strength depended on measurement temperature across a wide temperature range that covered the highly elastic and the glassy state of the adhesive. The influence of residual stresses τres, acting at the “binder–fiber” interface, on the nature of the curves describing the dependence of the adhesive strength on the studied factor was experimentally shown. A qualitative explanation of the observed regularities is proposed.
... The phase composition of the particles was determined using a Mössbauer spectrometer MS1104EM with 57 Co in Rh matrix, ensuring a noise/signal ratio of no more than 2 %. A detailed procedure for this can be found in our previous study (Yurkov et al., 2022). Additionally, X-ray diffraction diagrams were recorded using an X-ray diffractometer Diffray-401 (Scientific Instruments, Russia) with Bragg-Brentano focusing and Cr-Kα radiation (wavelength 0.22909 nm) at room temperature. ...
Article
Full-text available
A significant portion of the current knowledge regarding the use of iron nanoparticles for remediating metal-contaminated soils is derived from laboratory experiments, leaving several unanswered questions. This article presents a field experiment comparing the efficacy of magnetite nanoparticles and microparticles for the immobilization of metals and the growth of plants in metal-contaminated soils. This study aimed to investigate the effects of magnetite particle size on metal immobilization and plant growth in soils exposed to airborne pollution from the Middle-Urals Copper Smelter in the southern taiga subzone near Revda, Russia, 50 km from Ekaterinburg. Magnetite nano- and microparticles were added to forest litter at a 4 % w/w dose. The total metal contents in litter from the study plots were 1-2 orders of magnitude higher than background metal concentrations. The magnetite nanoparticle treatment was found to decrease the concentration of exchangeable copper in soil and improve the growth of red fescue (Festuca rubra L.) on polluted soil compared to the control. In contrast, magnetite microparticles did not show any statistically significant effects. These findings are in line with laboratory results that demonstrated the superior metal adsorption properties of magnetite nanoparticles compared to microparticles. However, this study was limited in duration (2 months), and longer field studies would be necessary to confirm the role of iron particle size in the rehabilitation of metal-contaminated soils. Keywords heavy metals; iron oxides; nanomaterials; nanotechnology; nano; micro
... This method of obtaining composites is called matrix isolation [26,27]. Varieties of this method include stabilizing the polymer suspension (microencapsulation) or forming distributed nanoparticles during thermal treatment of a solid joint solution of metal salt and polymer, which is accompanied by structuring and nanocapsulation of the metal-containing phase in a carbon matrix [28][29][30][31]. The nature of the reagent and the conditions of the catalytic reaction determine the methods of obtaining the composite [32,33]. ...
Article
Full-text available
Targeted synthesis of C/composite Ni-based material was carried out by the method of matrix isolation. The composite was formed with regard to the features of the reaction of catalytic decomposition of methane. The morphology and physicochemical properties of these materials have been characterized using a number of methods: elemental analysis, scanning electron microscopy (SEM), transmission electron microscopy (TEM), X-ray diffraction (XRD), Fourier transform infrared spectroscopy (FTIR), Raman spectroscopy, temperature programmed reduction (TPR-H2), specific surface areas (SSA), thermogravimetric analysis, and differential scanning calorimetry (TGA/DSC). It was shown by FTIR spectroscopy that nickel ions are immobilized on the polymer molecule of polyvinyl alcohol, and during heat treatment, polycondensation sites are formed on the surface of the polymer molecule. By the method of Raman spectroscopy, it was shown that already at a temperature of 250 °C, a developed conjugation system with sp2-hybridized carbon atoms begins to form. The SSA method shows that the formation of the composite material resulted in a matrix with a developed specific surface area of 20 to 214 m2/g. The XRD method shows that nanoparticles are essentially characterized by Ni, NiO reflexes. The composite material was established by microscopy methods to be a layered structure with uniformly distributed nickel-containing particles 5–10 nm in size. The XPS method determined that metallic nickel was present on the surface of the material. A high specific activity was found in the process of catalytic decomposition of methane—from 0.9 to 1.4 gH2/gcat/h, XCH4, from 33 to 45% at a reaction temperature of 750 °C without the stage of catalyst preliminary activation. During the reaction, the formation of multi-walled carbon nanotubes occurs.
... According to an analysis of the data presented in [38], the degree of substitution (x) in alumina hematite-α-(Fe1-xAlx)2O3 can be expressed in terms of the effective magnetic field at room temperature as: The paramagnetic part of the high-temperature Mössbauer spectra may have a dual nature. On the one hand, 7-10% of the area of this pair of doublets can be related to the superparamagnetic fractions of the Al-hematite and Al-goethite described above [40]. The remaining part, as shown in [16], refers to akageneite substituted by aluminum-β-Fe 1−x Al x O(OH, Cl), which, when cooled to the boiling point of nitrogen, transforms into a broadened sextet ( Table 2). ...
Article
Full-text available
Bauxite residue (BR), also known as red mud, is a byproduct of the alumina production using the Bayer process. This material is not used to make iron or other iron-containing products worldwide, owing to its high content of sodium oxide and other impurities. In this study, we investigated the hydrochemical conversion of goethite (FeOOH) to magnetite (Fe3O4) in high-iron BR from the Friguia alumina refinery (Guinea) by Fe2+ ions in highly concentrated alkaline media. The simultaneous extraction of Al and Na made it possible to obtain a product containing more than 96% Fe3O4. The results show that the magnetization of Al-goethite and Al-hematite accelerates the dissolution of the Al from the iron mineral solid matrix and from the desilication product (DSP). After ferrous sulfate (FeSO4·7H2O) was added directly at an FeO:Fe2O3 molar ratio of 1:1 at 120 °C for 150 min in solution with the 360 g L−1 Na2O concentration, the alumina extraction ratio reached 96.27% for the coarse bauxite residue size fraction (Sands) and 87.06% for fine BR obtained from red mud. The grade of iron (total iron in the form of iron elements) in the residue can be increased to 69.55% for sands and 58.31% for BR. The solid residues obtained after leaching were studied by XRD, XRF, TG-DTA, VSM, Mössbauer spectroscopy, and SEM to evaluate the conversion and leaching mechanisms, as well as the recovery ratio of Al from various minerals. The iron-rich residues can be used in the steel industry or as a pigment.
... The paramagnetic part of the high-temperature Mössbauer spectra may have a dual nature. On the one hand, 7-10% of the area of this pair of doublets can be related to the superparamagnetic fractions of the above described Al-hematite and Al-goethite [40]. The remaining part, as shown in [16] refers to akageneite substituted by aluminum -β-Fe1-xAlxO(OH, Cl), which, when cooled to the boiling point of nitrogen transforms into a broadened sextet ( Table 2).The remaining minor components in the low-temperature spectra in the form of doublets correspond to Fe 3+ ions in the high-spin state and octahedral oxygen environment [41], which can isomorphically replace Al 3+ ions [36] in the crystal lattice of, for example, sodalite [42]. ...
Preprint
Full-text available
Bauxite residue (BR), also known as red mud, is a by-product of the production of alumina via the Bayer process. Because of the high sodium oxide and other impurities content, this material is not used to obtain iron or other iron-containing products. In this paper, the hydro-chemical conversion of goethite (FeOOH) to magnetite (Fe3O4) in high-iron BR from the Friguia alumina refinery (Guinea) by Fe2+ ions in highly concentrated alkaline media was studied. The simultaneous extraction of Al and Na made it possible to obtain a product containing more than 96% Fe3O4. The results show that the magnetization of Al-goethite and Al-hemetite accelerates the dissolution of the Al from the iron mineral solid matrix and from the desilication product (DSP). After ferrous sulfate (FeSO4·7H2O) was added directly at the FeO:Fe2O3 molar ratio of 1:1 at 120 °C for 150 min in the solution with the 360 g L-1 Na2O concentration, the alumina extraction ratio reached 96.27% for the coarse bauxite residue size fraction (Sands) and 87.06% for fine BR obtained from red mud. The grade of iron (total iron in the form of iron element) in the residue can be increased to 69.55% for Sands and 58.31% for BR. The solid residues obtained after leaching were studied by XRD, XRF, TG-DTA, VSM, Mössbauer spectroscopy and SEM to evaluate the conversion and leaching mechanisms and the recovery ratio of Al from different minerals. The iron-rich residues can be used in the steel industry or as a pigment.
Article
Molybdenum-containing composite nanomaterials are synthesized by the thermal decompositionof molybdenum hexacarbonyl in a solution-melt of polyethylene in mineral oil. The concentration of a metalcontainingfiller in the composite materials varied from 1 to 20 wt %. A technique for preparing film samplesfor spectroscopic studies is developed, and the samples obtained are studied by UVI, IR, and Raman spectroscopy.It is found that additional absorption bands appear in the IR range, whose intensity depends on theconcentration of molybdenum-containing nanoparticles in the composite materials. The spectral characteristicsof Raman scattering show that all samples are characterized by the stretching of the C–C bond. In thevisible light region, the spectrum of nanocomposites has a flat edge of its own absorption located in the regionof wave numbers (18–31) × 103 cm–1.
Article
Full-text available
Molybdenum-containing composite nanomaterials are synthesized by the thermal decomposition of molybdenum hexacarbonyl in a solution-melt of polyethylene in mineral oil. The concentration of a metal-containing filler in the composite materials varied from 1 to 20 wt %. A technique for preparing film samples for spectroscopic studies is developed, and the samples obtained are studied by UV, IR, and Raman spectros-copy. It is found that additional absorption bands appear in the IR range, whose intensity depends on the concentration of molybdenum-containing nanoparticles in the composite materials. The spectral characteristics of Raman scattering show that all samples are characterized by the stretching of the CC bond. In the visible light region, the spectrum of nanocomposites has a flat edge of its own absorption located in the region of wave numbers (18-31) × 10 3 cm-1 .
Article
Full-text available
Silicon carbide fiber-reinforced silicon carbide matrix composites (SiC/SiC CMCs) are promising candidates for hot gas components in jet engines. Three common manufacturing routes are chemical vapor infiltration, reactive melt infiltration (RMI) and polymer infiltration and pyrolysis (PIP). A combination of the processes seems attractive: the remaining porosity after PIP process can be closed by subsequent siliconization, resulting in a dense material. This work describes a new approach of a combined PIP and RMI process. SiC/SiC CMCs were manufactured by PIP process using Hi-Nicalon Type S fibers. An additional RMI was carried out after a reduced number of PIP cycles. Microstructure was examined via μCT, SEM and EDS. Bending strength was determined to 433 MPa; strain to failure was 0.60 %. The overall processing time was reduced by 55 % compared to standard PIP route. The hybrid material contained 70 % less unreacted carbon than material produced by LSI process alone.
Article
Full-text available
Electrospinning is an emerging technique for synthesizing micron to submicron-sized polymer fibre supports for applications in energy storage, catalysis, filtration, drug delivery and so on. However, fabrication of electrospun ceramic fibre mats for use as a reinforcement phase in ceramic matrix composites or CMCs for aerospace applications remains largely unexplored. This is mainly due to stringent operating requirements that require a combination of properties such as low mass density, high strength, and ultrahigh temperature resistance. Herein we report fabrication of molecular precursor-derived silicon oxycarbide or SiOC fibre mats via electrospinning and pyrolysis of cyclic polysiloxanes-based precursors at significantly lower weight loadings of organic co-spin agent. Ceramic fibre mats, which were free of wrapping, were prepared by a one-step spinning (in air) and post heat-treatment for crosslinking and pyrolysis (in argon at 800 °C). The pyrolyzed fibre mats were revealed to be amorphous and a few microns in diameter. Four siloxane-based pre-ceramic polymers were used to study the influence of precursor molecular structure on the compositional and morphological differences of cross-linked and pyrolyzed products. Further thermal characterization suggested the potential of electrospun ceramic mats in high temperature applications.
Article
Full-text available
Red mud is an iron-containing waste of alumina production with high alkalinity. A promising approach for its recycling is solid-phase carbothermic roasting in the presence of special additives followed by magnetic separation. The crucial factor of the separation of the obtained iron metallic particles from gangue is sufficiently large iron grains. This study focuses on the influence of Na2SO4 addition on iron grain growth during carbothermic roasting of two red mud samples with different (CaO+MgO)/(SiO2+Al2O3) ratio of 0.46 and 1.21, respectively. Iron phase distribution in the red mud and roasted samples were investigated in detail by Mössbauer spectroscopy method. Based on thermodynamic calculations and results of multifactorial experiments, the optimal conditions for the roasting of the red mud samples with (CaO+MgO)/(SiO2+Al2O3) ratio of 0.46 and 1.21 were duration of 180 min with the addition of 13.65% Na2SO4 at 1150°C and 1350°C followed by magnetic separation that led to 97% and 83.91% of iron recovery, as well as 51.6% and 83.7% of iron grade, respectively. The mechanism of sodium sulfate effect on iron grain growth was proposed. The results pointed out that Na2SO4 addition is unfavorable for the red mud carbothermic roasting compared with other alkaline sulfur-free additives.
Article
Hexagonal ferrites with a magnetoplumbite structure (M-type hexaferrites or M-HF) and multiple substitution of Fe³⁺ ions with Ti⁴⁺, Al³⁺, Mn³⁺, Co²⁺, In³⁺, Ga³⁺, Cr³⁺ have a varying nominal chemical composition [Ba,Sr,Ca,La]{Fex(TiAlMnCoInGaCr)12-x}O19 (1.5<х<9) or ([BSCL]{Fex(TAMCIGC)12-x}O19). Ceramics samples were synthesized by solid-state reactions. Three batches of [BSCL]{Fex(TAMCIGC)12-x}O19 hexaferrites were produced at temperatures of 1300 °C, 1350 °C and 1400 °C. The correlation of chemical composition, phase content, structural and magnetic properties was investigated depending on the Fe content by EDX, XRD, SEM and PPMS. Samples synthesized at 1300°C were non-single-phase. Samples [BSCL]{Fex(TAMCIGC)12-x}O19 with 4.5<х<9, synthesized at 1300°C and 1400°C, were single-phase and can be described by the magnetoplumbite structure. The entropy state (ΔSconf) was calculated as a function of the Fe content. It was confirmed that single-phase multiple substituted M-HF with high ΔSconf can be obtained with Ti⁴⁺, Al³⁺, Mn³⁺, Co²⁺, In³⁺, Ga³⁺, Cr³⁺ substituents even if ΔSconf is noticeably less than 1.5R.
Article
Silica-coated iron oxide magnetic nanoparticles to be employed as the starting core for easily-recoverable heterogeneous catalysts supports, were synthesized. We tuned both the properties of the magnetic core and of the covering shell, with particular emphasis on the optimization of the silica layer thickness. Afterwards, structural-sensitive techniques (XRD, BET surface area analysis), surface-sensitive techniques (ATR-IR, XPS), imaging (SEM, TEM), elemental analyses (ICP-MS) and magnetization analyses (AGFM), were employed to unravel the features of the synthesized materials. The oxidation of cis-cyclooctene with tert-butyl hydroperoxide in presence of MNPs was used as probe reaction to check the reactivity of the NPs and the effectiveness of the silica shield. The results of these tests showed its chemical inertia confirming the absence of any detrimental iron oxide catalyzed side-reactions or magnetic core modification.
Article
Fe-doped indium tin oxide (ITO) is an exciting material because it combines the host matrix's good electrical conductivity with the magnetic properties coming from the most earth-abundant transition metal, Fe. In this regard, a single-pot synthesis route based on a polymeric precursor method has been used to produce high-quality undoped and iron-doped ITO with iron content up to 13.0 mol%. The crystal formation in the bixbyite-type structure of all samples is confirmed by X-ray diffraction data analysis. A monotonous decrease of the lattice parameters with the increase of the Fe content is determined, which is consistent with of Fe ions with an oxidation state of 3+ in agreement with the ionic radii difference between In³⁺ and Fe³⁺. Raman spectroscopy confirms the bixbyite structure formation and provides evidence of a high surface disorder. ¹¹⁹Sn Mössbauer spectroscopy reveals the formation of only Sn⁴⁺ ions. Meanwhile, ⁵⁷Co Mössbauer spectroscopy suggests the presence of Fe³⁺ ions in a paramagnetic state. DC magnetization characterization of the Fe-doped ITO nanoparticles confirms the compound's paramagnetic character. The sheet resistance (R/□) measurements provide a lower value for the undoped ITO sample (~0.26 Ω/sq) than the one of commercial bulk material. It has been determined that the sheet resistance increases with the Fe content, suggesting the decrease of the conduction electrons density as the iron content is increased.
Article
Metal-carbon composites (MCC) were produced by the dissolving of the Fe, Co and Ce nitrates in phenol-formaldehyde resin with further pyrolysis (under heating up to 800°C). Concentrations of the metal compositions in final MCC were 2, 7 and 15 wt.%. Phase composition of the final MCC samples was determined using XRD and Mössbauer spectroscopy. It was established phase separation in metallic compounds under pyrolysis process due to chemical interaction with phenol-formaldehyde decomposition products. The main phases for composites were: Fe3O4 (for MCC-Fe); Co (for MCC-Co) and CeO2 (for MCC-Ce). Average crystal size for metal-based fillers was 28, 24-35 and 3.5 nm for Fe-; Co and Ce-based composites respectively. The electrodynamic properties of the MCCs in the Extremely High Frequency range (30-50 GHz) were studied as function of the filler concentration. The frequency dependences of the permittivity and permeability (real and imaginary parts) of the investigated composites were measured. The nature of the MCC electrodynamic properties was discussed in detail in terms of polarization losses.
Article
In this study, iron oxide nanoparticles (NPs) were successfully prepared from three different reactive systems by using mechanochemical synthesis. The effect of using iron salts with different oxidation state and the addition of metallic Fe to the mixture reactive was investigated. The influence of different mechanochemical treatments times on the structural and magnetic properties of the obtained iron oxide NPs was also studied. Composition, crystal structure and morphology of the nanoparticles were analyzed by XRD, Raman spectroscopy, SEM, TEM and DLS techniques. The obtained crystalline NPs exhibited mean sizes of about 8–10 nm and agglomerate in clusters of about 300 nm. Also, magnetic properties as a function of temperature and applied field were determined for the obtained iron oxide NPs, showing high magnetization in the whole temperature range. The results indicated that the presence of metallic Fe in the starting mixture plays a crucial role in the formation of spinel magnetic phases (magnetite/maghemite). Structural and magnetic results are consistent with the formation of maghemite in the studied samples.
Article
This work describes a microwave synthetic approach for the controlled assembly of α-Fe2O3 nanosystems with defined morphologies, such as hollow nanotubes (NTs), solid nanorods (NRs) and nanodisks (NDs). The morphological control is aided during the crystallization processes by using phosphate anions as key surfactants in solution. Furthermore, the thermal reduction under H2 atmosphere of these NTs, NRs and NDs α-Fe2O3 systems to the correspondent Fe3O4 nanomaterials preserved their initial morphologies. It was observed that the concentration of phosphate anions and volume of solvent had significant impact not only on controlling the shapes and sizes, but also phase composition and stoichiometry of the NTs, NRs and NDs nanoparticles. X-ray Rietveld refinement analysis of the NTs, NRs and NDs systems, after reduction in H2, revealed the presence of zero-valent iron (Fe0) in the final materials, with Fe0 fractions that decreased gradually in % from NTs (∼16%), NRs (∼11%) to NDs (∼0%) upon increasing amount of phosphate anions. Bulk magnetic susceptibility measurements showed clear alterations of the Verwey transition temperatures (TV) and the development of unusual magnetic phenomena, such as magnetic vortex states in NDs, which was subsequently verified by micro-magnetic simulations. From the combination of XRD analysis, bulk magnetic susceptibility and Mössbauer results, we provide herein a detailed mechanistic description of the chemical processes that gated the development of shape-controlled synthesis of NTs, NRs and NDs and give a detailed correlation between specific morphology and magneto-electronic behaviors.
Article
Due to their non-toxicity and their ability to be functionalized, magnetite (Fe <sub xmlns:mml="http://www.w3.org/1998/Math/MathML" xmlns:xlink="http://www.w3.org/1999/xlink">3</sub> O <sub xmlns:mml="http://www.w3.org/1998/Math/MathML" xmlns:xlink="http://www.w3.org/1999/xlink">4</sub> ) nanoparticles (NPs) are good candidates for a variety of biomedical applications. To better implement their applications, it is crucial to well understand the basic structural and magnetic properties of the NPs in correlation with their synthesis method. Here, we show interesting properties of Fe <sub xmlns:mml="http://www.w3.org/1998/Math/MathML" xmlns:xlink="http://www.w3.org/1999/xlink">3</sub> O <sub xmlns:mml="http://www.w3.org/1998/Math/MathML" xmlns:xlink="http://www.w3.org/1999/xlink">4</sub> NPs of various sizes ranging from 5 to 100 nm and the dependence of these properties on particle size and preparation method. One synthetic method based on heating Fe(acac) <sub xmlns:mml="http://www.w3.org/1998/Math/MathML" xmlns:xlink="http://www.w3.org/1999/xlink">3</sub> with oleic acid consistently gives 5 ± 1 nm NPs. A second method using the thermal decomposition of Fe(oleate) <sub xmlns:mml="http://www.w3.org/1998/Math/MathML" xmlns:xlink="http://www.w3.org/1999/xlink">3</sub> in oleic acid led to larger NPs, greater than 8 nm in size. Increasing the amount of oleic acid caused the average NP size to slightly increase from 8 to 10 nm. Increasing both the reaction temperature and the reaction time caused the NP size to drastically increase from 10 to 100 nm. Powder X-ray diffraction and electron-microscopy imaging show a pure single crystalline Fe <sub xmlns:mml="http://www.w3.org/1998/Math/MathML" xmlns:xlink="http://www.w3.org/1999/xlink">3</sub> O <sub xmlns:mml="http://www.w3.org/1998/Math/MathML" xmlns:xlink="http://www.w3.org/1999/xlink">4</sub> phase for all NPs smaller than 50 nm and spherical in shape. When the NPs get larger than 50 nm, they notably tend to form faceted, FeO core–Fe <sub xmlns:mml="http://www.w3.org/1998/Math/MathML" xmlns:xlink="http://www.w3.org/1999/xlink">3</sub> O <sub xmlns:mml="http://www.w3.org/1998/Math/MathML" xmlns:xlink="http://www.w3.org/1999/xlink">4</sub> shell structures. Magnetometry data collected in various field-cooling conditions show a pure superparamagnetic (SPM) behavior for all NPs smaller than 20 nm. The observed blocking temperature, $T_{B}$ , gradually increases with NP size from about 25–150 K. In addition, the Verwey transition is observed with the emergence of a strong narrow peak at 125 K in the magnetization curves when larger NPs are present. Our data confirm the vanishing of the Verwey transition in smaller NPs. Magnetization loops indicate that the saturating field drastically decreases with NP size. While larger NPs show some coercivity ( $H_{c}$ ) up to 30 mT at 400 K, NPs smaller than 20 nm show no coercivity ( $H_{c} = 0$ ), confirming their pure SPM behavior at high temperature. Upon cooling below $T_{B}$ , some of the SPM NPs gradually show some coercivity, with $H_{c}$ reaching 45 mT at 5 K for the 10 nm NPs, indicating emergent interparticle couplings in the blocked state.