You are on page 1of 8

Ceramics International 48 (2022) 29854–29861

Contents lists available at ScienceDirect

Ceramics International
journal homepage: www.elsevier.com/locate/ceramint

Influence of sintering temperature on heterogeneous-interface structural


evolution and magnetic properties of Fe–Si soft magnetic powder cores
Rui Wang a, YiHai He a, Hui Kong a, Jian Wang b, Zhaoyang Wu a, c, *, Haichuan Wang a, **
a
Key Laboratory of Green Fabrication and Surface Technology of Advanced Metal Materials (Anhui University of Technology), Ministry of Education, Ma’anshan,
243002, China
b
Institute of New Materials, Guangdong Academy of Sciences, Guangzhou, Guangdong, 510650, China
c
Ma’anshan Shenma Machinery Manufacturing Co., Ltd., Ma’anshan, Anhui, 243002, China

A R T I C L E I N F O A B S T R A C T

Keywords: Ceramic oxides have attracted considerable research attention as ideal coating layers for novel Fe-based soft
Soft magnetic powder cores magnetic powder cores (SMPCs). However, maintenance of the integrity and uniformity of Fe-based/MOx cor­
Core–shell heterostructure e–shell heterostructures is challenging. The mechanism underlying the evolution of the core–shell hetero­
Sintering temperature
structure is a key determinant of the performance of Fe-based SMPCs. Herein, the laws governing the evolution of
Magnetic performance
the core–shell structure of and heterogeneous interface in Fe–Si/SiO2 SMPCs with temperature and the influence
of this evolution on SMPCs performance were investigated. The results revealed that at the sintering temperature
of 1093–1183 K, the core–shell heterostructure gradually integrated, while SiO2 insulating coatings underwent
amorphous-to-crystalline state transformation. When the sintering temperature was >1243 K, Fe–Si particles
melted partially, and the core–shell heterostructure collapsed owing to the overheating induced by the gradient
temperature field during the hot-pressing sintering process. When the sintering temperature was 1153 K, the
core–shell heterostructure was intact, and the Fe–Si/SiO2 SMPCs had a saturation magnetisation of 245.5 emu/g,
resistivity of 0.42 mΩ cm, and total loss of 923.2 kW/m3 at 10 mT and 100 kHz. When the core–shell hetero­
structure was destroyed, the resistivity dropped drastically, and the total loss increased by approximately 36.7%
and 41.8%. Based on these results, the relationship among the core–shell heterostructure of Fe–Si/SiO2 SMPCs,
sintering temperature, and magnetic properties was established, which is instrumental in achieving high power
density, high efficiency, and miniaturisation in SMPCs.

1. Introduction permeability. Inorganic insulating materials have high thermal stability


and can be used as insulating coating agents as they can withstand the
With the rapid evolution of electromagnetic drives and development high-temperature moulding of SMPCs; thus, they are considered an ideal
of next-generation semiconductors, soft magnetic materials have been alternative to organic resins [6].
developed to achieve high power density, high efficiency, and minia­ Among various inorganic insulating materials, ceramic oxides (MOx,
turisation [1]. Consequently, exploration and design of new Fe-based M = metal/non-metal such as Mg [7], Zr [8], Fe [9], and B [10]) have
soft magnetic powder cores (SMPCs) have received considerable been widely used. Ceramic oxides exhibit high resistivity, good
research attention [2]. Currently, most insulating layers in commercial film-formability, and high heat resistance and have low fabrication cost,
Fe-based SMPCs are prepared using organic resins such as epoxy resins making them the most suitable inorganic insulating materials for
[3], phenolic resins [4], and silicone resins [5]. However, the Fe-based SMPCs. However, the insulating coating on the surface of soft
compressive and heat-resistant properties of organic insulating layers magnetic particles is prone to rupture during the moulding process. This
are low. Therefore, Fe-based SMPCs can be fabricated and heat-treated results in an uneven insulating coating and increases the eddy current
only at low temperatures, resulting in low density and magnetic loss of the SMPCs because of the hardness and brittleness of the ceramic

* Corresponding author. Key Laboratory of Green Fabrication and Surface Technology of Advanced Metal Materials (Anhui University of Technology), Ministry of
Education, Ma’anshan, 243002, China.
** Corresponding author.
E-mail addresses: ahutwzy@ahut.edu.cn (Z. Wu), which@ahut.edu.cn (H. Wang).

https://doi.org/10.1016/j.ceramint.2022.06.250
Received 12 April 2022; Received in revised form 20 June 2022; Accepted 23 June 2022
Available online 28 June 2022
0272-8842/© 2022 Elsevier Ltd and Techna Group S.r.l. All rights reserved.
R. Wang et al. Ceramics International 48 (2022) 29854–29861

oxides. To overcome these drawbacks of ceramic oxides, numerous chemical vapour deposition (FB-CVD) [19]. Then, 26 g of the
experimental approaches have been proposed. For example, Wang et al. as-obtained Fe–Si/SiO2 core–shell particles were placed in a graphite die
[11] used SiO2 to coat Fe–Si particles; the obtained SMPCs prepared at and compacted in a vacuum hot-pressing sintering furnace (Changjiang
2100 MPa and room temperature exhibited high bulk resistivity (>1.4 Ω Jinggong Material Technology Co., Ltd., China). The sintering temper­
m), and their eddy current losses accounted for a lower contribution ature was increased to the temperature range of 1093–1303 K in 10 min,
(20.2–22.1%) to the total core loss than those observed in SMPCs coated and the temperature was then maintained for another 10 min. The
only with an organic resin (61.9%). Sun et al. [12] achieved a uniform temperature was reduced to 50 K at the rate of 50 K/min. The pressure
Al2O3/SiO2 coating of carbonyl iron powders using a hydrothermal was maintained at 14 MPa till the target temperature was reached. After
method; the resultant SMPCs exhibited a low core loss of 786.8 kW/m3 sintering, the SMPCs were annealed at 923 K for 120 min. The final
and high magnetic flux density at 400 Hz. Although the previously re­ dimensions of the Fe–Si/SiO2 SMPCs were as follows: outer diameter 30
ported approaches have expanded the application scope of ceramic ox­ mm, inner diameter 20 mm, and height 5 mm.
ides to SMPCs, the obtained insulating coating still has the major
drawback of noncompaction, which was reflected in lower density
(porosity between particles) and uneven insulating layer. To overcome 2.3. Characterisation
this, hot-pressing sintering was introduced. Hot-pressing sintering is a
method that introduces external axial pressure to promote material Differential scanning calorimetry (DSC) was performed using a
densification at high temperature. It can prevent the growth of grains, Netzsch thermal analyser (STA-499C, calibrated using Al, Ni, and Au) to
effectively increase the contact surface of grains, and promote the evaluate the temperature and determine the corresponding thermal ef­
diffusion of atoms into the material, thereby promoting ceramics, which fects of phase changes in the Fe–Si/SiO2 core–shell particles. X-ray
could be conducted to increase the temperature, achieving desired diffraction (XRD) analysis (Bruker D8 Advance) was conducted using Cu
properties [13,14] during the moulding process, thereby compacting the Kα radiation to analyse the phase compositions of the Fe–Si/SiO2 cor­
insulting coating. e–shell particles and Fe–Si/SiO2 SMPCs. The morphologies and local
Furthermore, the sintering conditions (e.g., temperature and time) chemical homogeneities of the samples were investigated by scanning
are crucial to the performance of SMPCs. Core–shell heterostructures electron microscopy (SEM, Tescan MIRA3 XMU, 2.0 nm@15 kV) with
formed by Fe-based soft magnetic particles and ceramic oxides in SMPCs energy-dispersive spectroscopy (EDS, JED-2300). Changes in the elec­
are fundamental to limiting the induced current during AC magnet­ tronic structure of SMPCs were analysed by X-ray photoelectron spec­
isation inside the insulating and conductive particles, reducing the eddy troscopy (XPS, PHI-5000versaprobe). Hysteresis loops were measured at
current radius, and minimising the energy loss [15]. To obtain room temperature using a vibrating sample magnetometer (MPMS-3)
high-performance SMPCs, the core–shell heterostructures must not be within the magnetic field strength − 20000-+20000 Oe, with a step size
damaged during the subsequent high-temperature moulding process. of 50 Oe. The electrical resistivities of the Fe–Si/SiO2 SMPCs were tested
Therefore, in the preparation of SMPCs, moulding temperature is a key using a resistivity tester (ST2253y). The core loss was determined using
determinant of the material structure and electromagnetic properties. A a B–H analyser (SY-8258 IWATSU).
low moulding temperature results in low material density and imperfect
grain morphology, while a high moulding temperature can easily 3. Results and discussion
destroy the integrity of the core–shell heterostructure, weakening the
effect of induced current limitation. At present, reports that specifically 3.1. Microstructures of Fe–Si/SiO2 core–shell particles
discuss the influence of moulding conditions focus only on the effects of
material density and final magnetic properties [16–18]; very few reports XRD analysis was performed to investigate the chemical components
have discussed the influence of moulding conditions on the integrity of of the particle samples. The XRD patterns of the Fe–Si alloy particles
the core–shell heterostructures. before and after deposition are shown in Fig. 1(a). For the Fe–Si alloy
In this study, Fe–Si/SiO2 SMPCs were synthesised at different sin­ particles, three diffraction peaks are centred at 44.87◦ (110), 65.46◦
tering temperatures and their magnetic properties were analysed. The (200), and 82.87◦ (211). The peaks are attributed to the body-centred
effects of the sintering temperature on the microscopic characteristics of cubic α-Fe(Si) solid-solution (ICDD09-065-9130), whose crystal has a
core–shell heterostructures, crystalline of the insulating coating, and space group of Im3m (229). After FB-CVD, the XRD pattern shows a wide
magnetic properties of Fe–Si/SiO2 SMPCs were investigated. By amorphous diffraction peak at 22◦ , in addition to the aforementioned
observing the structural and interfacial changes in Fe–Si/SiO2 in SMPCs, α-Fe(Si)-phase diffraction peaks. This indicates that the SiO2 deposition
the evolution behaviour of the Fe–Si/SiO2 heterostructure and its rela­ process does not affect the phase structure of Fe–Si substrates, which is
tionship with the magnetic properties were studied in a high- consistent with the results reported in the literature [20,21]. Multiple
temperature environment. The relationship concluded herein will intermediates, such as OSi(OC2H5)2, O(Si(OC2H5)3)2, and (OC2H5)2Si
prove helpful in improving the performance, expanding the application (OH)2, are generated during the thermal decomposition of C8H20O4Si,
scope, and promoting the development of SMPCs. while SiO2 and C2H5OH are the final products under non-oxidising, at­
mospheric pressure conditions [22]. Thus, the formation of Fe–Si/SiO2
2. Experimental core–shell particles can be summarised by the following chemical re­
action [23].
2.1. Materials
C8H20O4Si (l) → Si(OH)4-x(OC2H5)x (g, x = 0–4, precursor) → SiO2 (s) +
C2H5OH (g) (1)
Commercial atomised Fe–Si particles (Si content: 6.62 wt%; purity:
>99.9%, average particle size = 50 μm) were supplied from Hualiu New To characterise the core–shell structure of the coated particles, SEM
Materials Co., Ltd. Tetraethyl orthosilicate (C8H20O4Si, 99.0%) was images and EDS spectra of the Fe–Si particles before and after deposition
provided by Zhiyuan Chemical Reagent Co., Ltd. Argon gas (Ar, 99.99%) were recorded; these images are displayed in Fig. 1(b–e). Despite the
was supplied from Tianze Gas Co., Ltd. presence of some wrinkles and pits, the surface of the Fe–Si particles is
relatively smooth. Furthermore, numerous insulating SiO2 particles,
2.2. Synthesis which were synthesised and deposited on the surface of Fe–Si substrate
by CVD, are layered and uniform in size. The SEM images reveal that the
In this study, Fe–Si/SiO2 core–shell particles, with Fe–Si particles as Fe–Si particles are tightly coated by an integrated, continuous insulating
the core and SiO2 as the shell, were synthesised by fluidised-bed coating. As illustrated in Fig. 1(f–m), the planar and cross-sectional SEM

29855
R. Wang et al. Ceramics International 48 (2022) 29854–29861

Fig. 1. XRD patterns (a), surface morphology SEM images before (b–c), and after deposition (d–e), and EDS results of the deposited Fe–Si particles (Fe in (g) and (k),
Si in (h) and (l), and O in (i) and (m)).

images and EDS spectra of the Fe–Si/SiO2 particles were mapped to coating and hetero-interface under the sintering pressure. Consequently,
demonstrate the formation of the core–shell structure. The elemental the collapse and destruction of the core–shell heterostructure begins.
distribution maps show that Si is distributed throughout the cross sec­ The partial melting of the Fe–Si particles mentioned above is caused by
tion of the particles, whereas O and Fe are mainly distributed on the hot-pressing sintering. During the hot-pressing sintering process, an
shell and in the core, respectively [9,24]. These results show that axial pressure is applied to the SMPCs, which promotes the diffusion of
Fe–Si/SiO2 particles with a fine microstructure and core–shell hetero­ atoms into the material, strengthens the dynamic conditions of sintering,
structure characteristics can be obtained using previously described increases the heating rate, and shortens the sintering time. The sintering
methods [25]. heat required for preparing SMPCs from Fe–Si/SiO2 core–shell hetero­
structure composites depends on the thermal energy of the Fe–Si
inter-particles and the thermal energy absorbed by the composite. The
3.2. Microstructural evolution of Fe–Si/SiO2 SMPCs
superposition of the two thermal energies at the heterogeneous interface
locally generates high temperature on the surface of the Fe–Si particles
Backscattered electron imaging(BSE) of the polished surfaces of the
and forms a gradient temperature field on the particles which decreases
Fe–Si/SiO2 SMPCs, sintered at different temperatures, are shown in
from the surface to the interior [27,28]. Based on this mechanism,
Fig. 2(a–h). The SMPCs sintered at 1093 K contain some pores, which
simultaneous high densification of the core (Fe–Si particles with a low
are mainly caused by the poor fluidity of the particles owing to the large
melting point) and shell (SiO2 insulating coating with a high melting
frictional force arising from the nano-SiO2 particles [26] and an insuf­
point) at different sintering temperatures can be achieved. Furthermore,
ficient pressing temperature during the moulding process; further, the
this temperature transfer gradient is highly extensive, resulting in
coated SiO2 accumulates in the interface of the particles. With an in­
localised melting (>1700 K) of the Fe–Si particles with an increase in the
crease in the sintering temperature, the SMPCs sintered at 1123, 1153,
sintering time.
and 1183 K (Fig. 2(b–d)) have fewer pores than that sintered at 1093 K
Fig. 2(i) shows the XRD patterns of Fe–Si/SiO2 SMPCs sintered at
(Fig. 2(a)). The concentration of the coated SiO2 gradually increases
different temperatures. Three diffraction peaks appear at 44.75◦ , 65.22◦ ,
along the edges of the Fe–Si particles, where the coatings are even,
and 82.59◦ , corresponding to the (110), (200), and (211) planes of the
owing to the increase in density and decrease in porosity with increasing
α-Fe (Si) phase of the body-centred cubic structure, respectively, where
temperature. Furthermore, the core–shell structure is completely pre­
the space group of the crystal is Im3m (229). To further investigate the
served in the temperature range of 1093–1213 K. Fig. 2(e) shows that
formation mechanism of the Fe–Si/SiO2 composites, DSC was per­
when the sintering temperature continues to increase up to 1243 K, an
formed, and the curves are shown in Fig. 2(j). The Fe–Si/SiO2 compos­
uncoated grey region appears and a part of the insulating coating dis­
ites have a characteristic peak from 902.75 to 1015.5 K, centred at
appears. With further increase in sintering temperature, black regions
968.6 K; this can be attributed to the crystallisation of the amorphous
begin to form in the grey region, which gradually becomes more
SiO2 phase obtained from C8H20O4Si [29,30].
disorderly, as shown in Fig. 2(f–h). In addition, the internal Fe–Si par­
A comprehensive analysis of the XRD patterns of the particles and
ticles start melting, owing to which the Fe–Si cannot support the SiO2

29856
R. Wang et al. Ceramics International 48 (2022) 29854–29861

Fig. 2. BSE images of the polished surfaces of Fe–Si/SiO2 SMPCs sintered at different temperatures: (a) 1093 K, (b) 1123 K, (c) 1153 K, (d) 1183 K, (e) 1213 K, (f)
1243 K, (g) 1273 K, and (h) 1303 K; XRD patterns of Fe–Si/SiO2 SMPCs (i); and DSC curves (j).

SMPCs showed that SiO2 existed in an amorphous state on the surface of group and the connection reduced the electron binding energy of the O
the Fe–Si particles; however, after high-temperature sintering, the broad atoms. These results imply the formation of interfacial covalent bonds
peak representing the amorphous SiO2 disappeared. This was because between the Fe–Si core and SiO2 shell [32]. As the temperature
the local overheating phenomenon promoted the transformation of SiO2 increased to 1273 K, the proportion of the Et3 group derived from Si
to the crystalline state during the high-temperature forming process, decreased to 33.13 at%, whereas the proportions of the Et3 group and
which can be inferred from the DSC curve. The combination of XRD and Fe–O group corresponding to O1s exhibited a significant increase
DSC results revealed that with an increase in the sintering temperature (56.11 at%) and a slight decrease (6.37 at%), respectively; this can be
beyond 1183 K, the crystallisation of SiO2 was promoted, the crystalline attributed to the breakage of the interfacial bonds owing to disruption of
characteristic peak of SiO2 (26.46◦ and 54.58◦ ) became highly pro­ the core–shell heterostructure.
nounced, and the peak intensity increased. This was because the Fe–Si
particles melted with a further increase in the sintering temperature, and
3.3. Effect of sintering temperature on magnetic properties
SiO2 was no longer confined at the interface of the particles but diffused
extensively.
Fig. 4(a) and (b) shows the hysteresis loops and the coercivity of the
To further analyse the reactions, surface chemistry, and composition,
SMPCs prepared at different sintering temperatures. All SMPCs exhibi­
XPS was performed on the Fe–Si/SiO2 SMPCs; the spectra are shown in
ted a high saturation magnetisation (Ms) when the coercivity was
Fig. 3. The peak positions and atomic ratios were obtained by curve
maintained at 10–30 Oe. However, owing to limitations of the testing
fitting, and the electron binding energy of the C1s peak was set to 284.8
instrument, each test step during the detection of the hysteresis loop
eV to correct the positions of the other peaks. The spectrum of Si2p
corresponded to 50 Oe, which is much larger than the actual value.
(Fig. 3(a)) contained four deconvolution curves, with their peak electron
Therefore, the coercivity value cannot be used as the basis for perfor­
binding energies at 100.53 (20.82 at%), 101.07 (9.42 at%), 102.27
mance comparison. Further, the Ms value scaled from 226.7 to 245.5
(44.79 at%), and 103.16 eV (24.97 at%), indicating the existence of four
emu/g with increasing temperature. This was because the density of the
Si radicals constituting a Si–O–Si structure [23]. The Etx group
SMPCs increased, the insulating SiO2 became more evenly distributed,
composition was represented as (Si(OC2H5)x(OH)4-x, x = 0–3), which
and the core–shell heterostructure became more complete with an in­
corresponds to the Et2 (100.53 eV), Et0 (101.07 eV), Et3 (102.27 eV),
crease in the sintering temperature, thereby reducing the porosity and
and Et1 (103.16 eV) groups, according to the literature [31]. Similarly,
increasing the total magnetic moment per unit volume. In the sintering
the relationship between the O1s binding state and the electron binding
temperature range of 1153–1183 K, Ms reached the highest value.
energy was derived from the O atoms of the Et2 (528.84 eV), Et0
However, as the sintering temperature continued to increase, the satu­
(529.73 eV), Et1 (530.7 eV), and Et3 (532.03 eV) groups and the atoms
ration magnetisation of the other samples did not change significantly
connected to Fe, whose binding energy was 532.86 eV, as shown in
because it depended on the total magnetic moment per unit volume [20,
Fig. 3(b). The total number of O atoms (42.75 at%) from the Et3 group
33]. This trend was observed because although the core–shell hetero­
(34.75 at%) and atoms connected with the Fe atoms (8 at%) was almost
structure was destroyed, the number of magnetic parts did not decrease.
equal to the number of Si atoms in the Et3 group (42.71 at%). This shows
After 1213 K, the pores generated by the destruction of the core-shell
that the O atoms connected to the Fe atoms also originated from the Et3
heterostructure led to a slight decrease in Ms [9].

29857
R. Wang et al. Ceramics International 48 (2022) 29854–29861

Fig. 3. XPS images of the Fe–Si/SiO2 SMPCs sintered at 1183 K (Si2p (a) and O1s (b)) and 1273 K (Si2p (c) and O1s (d)).

Fig. 4. Hysteresis curves (a) and saturation magnetisation (b) showing the coercivity and Ms of Fe–Si/SiO2 SMPCs sintered at different temperatures.

Quantifying the uniformity in the coating of SMPCs is challenging. material, e is the charge on an electron, and μ is the semiconductor
Qualitatively, better uniformity implies higher inter-particle electrical mobility [35]. After 1183 K, the amorphous SiO2 begins to crystallise; its
insulation, which is conducive to efficiently block the generation of crystal structure arrangement gradually becomes neat and long-range
inter-particle eddy currents as well as to reduce the magnetic weakening ordered, the vacancies that originally exist in the amorphous layer fill
caused by additional non-magnetic agglomeration [34]. Thus, resistivity up, and the carrier concentration is greatly reduced. In contrast, the
was introduced to characterise the electrical insulation of the prepared Fe–Si atomic structure tends to be uniform. Crystallisation of the
Fe–Si/SiO2 SMPCs. The trend of change of resistivity is shown in Fig. 5. insulting coating destroys the continuity, dense conductive network
When the sintering temperature is low, the SiO2 insulting coating is formed by the particles, thereby increasing the electron-tunnelling
amorphous with long-range disorder and uneven distribution, and there barrier and reducing the electron-tunnelling probability; as a result,
are many vacancy-like structures. With increase in sintering tempera­ the resistivity shows a sudden increase. However, the sintering tem­
ture, the internal pores disappear, and the resistivity decreases. perature continues to increase, and black spots began to appear at the
particles boundary, which was a sign of over-sintering, resulting in a
ρ = 1/neμ (2) drop in resistivity [36]. After 1243 K, owing to the accumulation of the
In Equation (2), n is the carrier concentration of the semiconductor two thermal effects of the sintering process, the temperature between

29858
R. Wang et al. Ceramics International 48 (2022) 29854–29861

decreases the resistivity again.


Fig. 6 shows the total loss distribution diagram of the Fe–Si/SiO2
particles formed in the SMPCs after sintering at different temperatures.
The total loss of SMPCs sintered at all temperatures increases with
increasing test frequency, and the total loss first decreases and then
increases with increasing sintering temperature. Under the conditions of
10 mT and 100 kHz, for the SMPCs sintered at 1273 and 1303 K, the total
losses are up to 1309.1 and 1262.2 kW/m3, while the total loss of the
SMPC sintered at 1183 K is 943.3 kW/m3, which is a decrease of 38.8%
and 33.8%, respectively. This shows that with increase in sintering
temperature, the pores gradually decrease, the core–shell hetero­
structure becomes complete, and the loss gradually decreases. As the
sintering temperature continues to increase, the Fe–Si particles melt,
which distorts the core–shell heterostructure. As a result, the magnetic
domain structure is destroyed, the potential barrier to the displacement
of the domain wall is increased, thereby increasing the total loss.
According to the classic Bertotti’s loss separation theory, the total
mass loss Pcv (kW/m3) can be separated into three different physical
parts: hysteresis loss—Physt, eddy current loss—Pec, and excess
loss—Pexc [37]. The model can be expressed as follows:
Fig. 5. Changing trend of resistivity of Fe–Si/SiO2 SMPCs sintered at different
temperatures. Pcv = Physt + Pec + Pexc (3)

The hysteresis loss is computed as the product of the area of the


the particles inside the SMPCs is much higher than the setting sintering quasi-static hysteresis loop multiplied and the frequency:
temperature, even exceeding the melting point of the Fe–Si particle.
Consequently, the particles melt, and the insulating coating loses its Physt = Chyst Bmα f (4)
support and gradually becomes scattered and disordered, which

Fig. 6. Total loss (a) of Fe–Si/SiO2 SMPCs sintered at different temperatures and the hysteresis loss (b), eddy current loss (c), and excess loss (d) after the loss
distribution of Fe–Si/SiO2 SMPCs sintered at 1183 and 1273 K.

29859
R. Wang et al. Ceramics International 48 (2022) 29854–29861

where Chyst represents the hysteresis coefficient, Bm represents the weakened, thus reducing the total loss. When the sintering temperature
maximum induction, f represents the frequency, and α represents a was 1183 K, the core–shell heterostructure remained intact, and the
simulation coefficient, usually between 1.6 and 2.2 for most ferromag­ total loss of SMPCs was minimum. As the sintering temperature
netic materials and alloys. In soft magnetic composite materials, the continued to increase, the local overheating during the sintering process
eddy current can be expressed as follows: led to a gradual rupture of the core–shell heterostructure. The internal
structure of the SMPCs tended to be chaotic, which increased the total
Pec = Pinter + Pintar
ec
/ ( 2 2
ec loss. Fig. 6(b–d) show the comparison of various losses of the SMPCs
= 1 6 π deff B2m f) ​ /[1 − 0.633(w/h)tanh(1.58h/w)]ρs prepared at 1183 and 1273 K before and after the destruction of the
+ π2 d2 B2m f 2 /β2 ρp (5) core–shell heterostructure. The main effect of the different sintering
temperatures on the SMPCs was the reduction in eddy current losses,
where Pinter which decreased by 47.1%.
ec and Pec represent the inter- and intra-particle eddy current
intar

coefficients, respectively. deff represents the effective eddy current The overall comparison of SMPCs sintered at different temperatures
dimension, which is the specimen thickness, and d is the particle size. ρs is presented in Table 2. Under 14 MPa sintering pressure for 10 min,
and ρp represent the bulk resistivity of the specimen and the particle, 1153 K is the optimum sintering temperature in terms of the magnetic
respectively. β1 represents the rectangular cross-sectional geometrical properties.
factor, and w and h represent the width and height of the rectangle,
respectively. β2 represents a granular geometrical factor with different 4. Conclusions
values for different geometries, e.g. for spheres, β2 = 20 [38]. Empiri­
cally, Pexc ∝ f1.5 is assumed to be an oversimplification; the excess loss In this study, the effects of sintering temperature on the microscopic
depends on several variables, such as the test frequency, applied mag­ characteristics of core–shell heterostructures, crystalline behaviour of
netic field, cross-sectional area of the material perpendicular to the insulating coating, and magnetic properties of Fe–Si/SiO2 SMPCs were
magnetic flux, and number of active magnetic objects under quasi-static investigated. The main conclusions are as follows:
and dynamic magnetisation [24]. Accordingly, the expression should be
modified as follows: 1) In the temperature range of 1093–1213 K, the core–shell hetero­
structures were well integrated and became more uniform with
Pexc = Cexc Bxm f y (6) increasing temperature. However, when the temperature exceeded
1243 K, the core–shell heterostructure was destroyed.
where Cexc represents the excess coefficient, and x and y represent the
2) As the saturation magnetisation increased, the resistivity and total
coefficients related to the applied magnetic field and frequency,
loss decreased with increasing sintering temperature, while the
respectively.
core–shell heterostructure became more complete. The resistivity
First, the Pcv/f versus f curve was simulated for different maximum
increased abruptly because of the partial crystallisation of SiO2. After
magnetic flux densities using a polynomial curve simulation method.
the core–shell heterostructure was destroyed owing to the exces­
Then, the intercepts representing the value of hysteresis loss under
sively high sintering temperature (>1243 K), the saturation mag­
quasi-static conditions were obtained by extrapolating the above-
netisation changed slightly, the loss increased, and the resistivity
mentioned simulation curve to zero frequency. By fitting the quasi-
gradually decreased.
static hysteresis loss under different external fields, the Chyst and α
3) The Fe–Si/SiO2 SMPCs sample exhibited the best magnetic proper­
values were obtained, and the eddy current loss coefficient was calcu­
ties at a sintering temperature of 1153 K. The sample had a Ms of
lated using Equation (5). The parameters obtained after fitting and
245.5 emu/g, resistivity of 0.42 mΩ cm, and total loss of 923.2 kW/
calculation were introduced in Equation (6) to fit the parameters Cexc, x,
m3 at 10 mT, 100 kHz, which decreased by 41.8% from the
and y [39]. Through the above derivation, the formula and diagram of
maximum.
total loss and loss separation can be obtained, the fitting process can be
seen in Fig. S1 and the final fitting results are listed in Table 1. To better
Funding
understand the influence of the existence of core–shell heterostructure
on various losses, the SMPCs sintered at 1183 K with an intact core–shell
This work was supported by the Chinese National Science Founda­
heterostructure and the SMPCs sintered at 1273 K with a damaged
tion (grant number 51904002), China; Natural Science Foundation of
core–shell structure were analysed, as shown in Fig. 6(b–d).
Anhui Province (grant number 1908085QE190), China; Anhui Provin­
Fig. 6(a) shows that the total loss of SMPCs first decreased and then
cial Key Research and Development Plan (202104b11020007), China;
increased with increasing sintering temperature. Initially, pore sizes in
Anhui Special Support Plan (T000609), China and Distinguished Pro­
the SMPCs gradually decreased, the core–shell heterostructure was
fessor of the Wanjiang Scholars Project, China.
progressively refined, and induced anisotropy caused by the impurity-
like diffusion and hindering effect between the magnetic domains was

Table 1
Coefficients of Chyst, Cec, Cexc and other fitting parameters of the Fe–Si/SiO2
Table 2
SMPCs at different sintering temperatures.
Overall comparison of magnetic properties at different sintering temperatures.
Sample sintering Hysteresis Eddy current Excess R-
Sample sintering Saturation Resistivity Total loss at 10
temperature (K) component component component square
temperature (K) magnetisation (emu/ (mΩ⋅cm) mT, 100 kHz
Chyst А Cec Cexc x y g) (kW/m3)

1093 30.94 1.95 0.25 0.89 1.98 1.84 0.9995 1093 229.2 0.61 1299.6
1123 47.56 2.03 0.37 0.65 1.81 1.78 0.9996 1123 226.7 0.47 1094.2
1153 35.92 2.05 0.41 0.48 1.74 1.58 0.9999 1153 245.5 0.42 923.2
1183 33.74 2.03 0.27 0.55 1.84 1.78 0.9999 1183 241.2 0.67 943.3
1213 32.88 1.99 0.38 0.61 1.84 1.69 0.9999 1213 220.6 0.53 1010.4
1243 32.77 1.98 0.45 1.04 1.95 1.67 0.9994 1243 209.2 0.38 1091.8
1273 23.97 1.92 0.50 1.13 2.02 1.82 0.9999 1273 217.4 0.35 1309.1
1303 17.33 1.82 0.49 1.60 2.16 1.83 0.9999 1303 208.6 0.36 1262.2

29860
R. Wang et al. Ceramics International 48 (2022) 29854–29861

Author Contributions [17] Z. Luo, X.a. Fan, W. Hu, F. Luo, G. Li, Y. Li, J. Wang, X. Liu, Effect of sintering
temperature on microstructure and magnetic properties for Fe-Si soft magnetic
composites prepared by water oxidation combined with spark plasma sintering,
Rui Wang, Writing - Original draft, Formal analysis, and Investiga­ J. Magn. Magn Mater. 491 (2019), 165615-165615.
tion; Yihai He, Investigation and Methodology; Hui Kong, Picture [18] A.M. Predescu, R. Vidu, A. Predescu, E. Matei, C. Pantilimon, C. Predescu,
rendering and Review, Jian Wang, Performance Test; Zhaoyang Wu, Synthesis and characterization of bimodal structured Cu-Fe3O4 nanocomposites,
Powder Technol. 342 (2019) 938–953.
Review, Editing and Project administration; Haichuan Wang, Project [19] Q. Zhang, S. Li, W. Zhang, K. Peng, Influence of processed parameters on the
administration. magnetic properties of Fe/Fe3O4 composite cores, J. Mater. Sci. Mater. Electron. 32
(2020) 1233–1241.
[20] A.H. Gemeay, B.E. Keshta, R.G. El-Sharkawy, A.B. Zaki, Chemical insight into the
Declaration of competing interest adsorption of reactive wool dyes onto amine-functionalized magnetite/silica core-
shell from industrial wastewaters, Environ. Sci. Pollut. Res. Int. 27 (2020)
The authors declare that they have no known competing financial 32341–32358.
[21] S.G. de Moura, T.C. Ramalho, L.C.A. de Oliveira, L.C.L. Dauzakier, F. Magalhães,
interests or personal relationships that could have appeared to influence Photocatalytic degradation of methylene blue dye by TiO2 supported on magnetic
the work reported in this paper. core shell (Si@Fe) surface, J. Iran. Chem. Soc. 19 (2021) 921–935.
[22] L. Liu, X.W. Liao, J.X. Jia, H. Kong, X.A. Fan, Z.Y. Wu, X.S. Wang, Temperature-
controlled conversion from Fe–Si particles to integrated Fe–Si/SiO2 core–shell
Appendix A. Supplementary data structure particles during fluidised bed chemical vapour deposition, Ceram. Int. 46
(2020) 3059–3065.
Supplementary data to this article can be found online at https://doi. [23] Z.Y. Wu, C. Xian, J.X. Jia, X.W. Liao, H. Kong, X.S. Wang, K. Xu, Silica coating of
Fe-6.5 wt%Si particles using fluidized bed CVD: effect of precursor concentration
org/10.1016/j.ceramint.2022.06.250.
on core–shell structure, J. Phys. Chem. Solid. 146 (2020), 109626.
[24] Q. Zhang, S. Li, W. Zhang, K. Peng, Influence of processed parameters on the
References magnetic properties of Fe/Fe3O4 composite cores, J. Mater. Sci. Mater. Electron. 32
(2020) 1233–1241.
[1] E. Poskovic, F. Franchini, L. Ferraris, E. Fracchia, J. Bidulska, F. Carosio, [25] Z. Wu, C. Xian, J. Jia, X. Liao, H. Kong, K. Xu, Formation process of the integrated
R. Bidulsky, M. Actis Grande, Recent advances in multi-functional coatings for soft core(Fe-6.5wt.%Si)@Shell(SiO2) structure obtained via fluidized bed chemical
magnetic composites, Materials 14 (22) (2021) 6844. vapor deposition, Metals 10 (4) (2020) 520.
[2] W. Yuan, K. Sun, J. Tian, Y. Li, Z. Wang, B. Liu, R. Fan, Improved magnetic [26] K.L. Zheng, P.F. Yan, X.S. Wei, B. Yan, Study of the nano-network structure in the
properties of iron-based soft magnetic composites with a double phosphate-SiO2 friction transfer film of the hybrid reinforced aluminum-based composite, Wear
shells structure, J. Mater. Sci. Mater. Electron. 32 (2021) 21472–21482. 494–495 (2022), 204268.
[3] C. Wu, M. Huang, D. Luo, Y. Jiang, M. Yan, SiO2 nanoparticles enhanced silicone [27] L. Zhou, J. Huang, X. Wang, G. Su, J. Qiu, Y. Dong, Mechanical, dielectric and
resin as the matrix for Fe soft magnetic composites with improved magnetic, microwave absorption properties of FeSiAl/Al2O3 composites fabricated by hot-
mechanical and thermal properties, J. Alloys Compd. 741 (2018) 35–43. pressed sintering, J. Alloys Compd. 774 (2019) 813–819.
[4] M. Streckova, R. Bures, M. Faberova, L. Medvecky, J. Fuzer, P. Kollar, [28] T. Schäfter, J. Burghaus, W. Pieper, F. Petzoldt, M. Busse, New concept of Si–Fe
A comparison of soft magnetic composites designed from different ferromagnetic based sintered soft magnetic composite, Powder Metall. 58 (2014) 106–111.
powders and phenolic resins, Chin. J. Chem. Eng. 23 (2015) 736–743. [29] L. Han, J. Song, C. Lin, J. Liu, T. Liu, Q. Zhang, Z. Luo, A. Lu, Crystallization,
[5] S. Wu, A. Sun, F. Zhai, J. Wang, Q. Zhang, W. Xu, P. Logan, A.A. Volinsky, structure and properties of MgO-Al2O3-SiO2 highly crystalline transparent glass-
Annealing effects on magnetic properties of silicone-coated iron-based soft ceramics nucleated by multiple nucleating agents, J. Eur. Ceram. Soc. 38 (2018)
magnetic composites, J. Magn. Magn Mater. 324 (2012) 818–822. 4533–4542.
[6] B. Kocsis, L.K. Varga, I. Zsoldos, Preparation of soft magnetic composite from Fe- [30] Z. Gao, J. Jia, Q. Zhao, H. Kong, Z. Wu, J. Li, Determination of a quantitative
6.9wt%Si by different heat treatment strategies, IOP Conf. Ser. Mater. Sci. Eng. 903 relationship between deposition duration and magnetic performance of soft
(2020), 012042. ferromagnetic composites via data-analysis and theoretical models, J. Magn. Magn
[7] X. Shi, X. Chen, K. Wan, B. Zhang, P. Duan, H. Zhang, X. Zeng, W. Liu, H. Su, Mater. 549 (2022), 168891.
Z. Zou, Y. Du, Enhanced magnetic and mechanical properties of gas atomized Fe-Si- [31] V. Shukla, Role of spin disorder in magnetic and EMI shielding properties of Fe3O4/
Al soft magnetic composites through adhesive insulation, J. Magn. Magn Mater. C/PPy core/shell composites, J. Mater. Sci. 55 (2019) 2826–2835.
534 (2021), 168040. [32] B. Bai, Y. Zhu, J. Miao, X. Wang, S. Bi, L. Kong, W. Liu, L. Zhang, Electromagnetic
[8] K. Geng, Y. Xie, L. Xu, B. Yan, Structure and magnetic properties of ZrO2-coated Fe wave absorption performance and mechanisms of geoploymer-based composites
powders and Fe/ZrO2 soft magnetic composites, Adv. Powder Technol. 28 (2017) containing core-shell SiO2@Fe3O4 nanoparticles, Ceram. Int. 48 (2022)
2015–2022. 2755–2762.
[9] C. Zhang, W. Zhang, W. Yuan, K. Peng, Preparation and magnetic properties of [33] M. Tajabadi, I. Rahmani, S.M. Mirkazemi, H. Goran Orimi, Insights into the
core–shell structured Fe-Si/Fe3O4 composites via in-situ reaction method, J. Magn. synthesis optimization of Fe@SiO2 Core-Shell nanostructure as a highly efficient
Magn Mater. 531 (2021), 167955. nano-heater for magnetic hyperthermia treatment, Adv. Powder Technol. 33
[10] L.L. Evangelista, G. Tontini, A.I. Ramos Filho, L.E. Machado, B.S. Silva, I.P.C. Silva, (2021), 103366.
G. Hammes, R. Binder, C. Binder, N.J. Batistela, A.N. Klein, V. Drago, [34] Z. Luo, X.a. Fan, B. Feng, Z. Yang, D. Chen, S. Jiang, J. Wang, Z. Wu, X. Liu, G. Li,
Developments on soft magnetic composites with double layer insulating coating: Y. Li, Highly enhancing electromagnetic properties in Fe-Si/MnO-SiO2 soft
synergy between ZnO and B2O3, J. Magn. Magn Mater. 497 (2020), 166023. magnetic composites by improving coating uniformity, Adv. Powder Technol. 32
[11] J. Wang, S. Song, H. Sun, G. Hang, Z. Xue, C. Wang, W. Chen, D. Chen, Insulation (2021) 4846–4856.
layer design for soft magnetic composites by synthetically comparing their [35] X. Wei, H. Qi, S. Zhu, X. Zhang, Y. Wang, X. Ouyang, W. Zheng, Extracting carrier
magnetic properties and coating process parameters, J. Magn. Magn Mater. 519 concentration of black c-BN single crystal by mid-infrared reflectance
(2021), 167496. spectroscopy, Vacuum 202 (2022), 111132.
[12] K. Sun, S. Feng, Q. Jiang, X. Li, Y. Li, R. Fan, Y. An, J. Wang, Intergranular [36] J. Fan, Y. Han, P. Li, Z. Sun, Q. Zhou, Micro/nano composited tungsten material
insulating reduced iron powder-carbonyl iron powder/SiO2-Al2O3 soft magnetic and its high thermal loading behavior, J. Nucl. Mater. 455 (2014) 717–723.
composites with high saturation magnetic flux density and low core loss, J. Magn. [37] M.M. Nell, B. Schauerte, T. Brimmers, K. Hameyer, Simulation of iron losses in
Magn Mater. 493 (2020), 165705. induction machines using an iron loss model for rotating magnetization loci in no
[13] W.Z. Mu, P.G. Jonsson, H. Shibata, K. Nakajima, Inclusion and microstructure electrical steel, COMPEL (2022) 600–614, ahead-of-print 41.
characteristics in steels with TiN additions, Steel Res. Int. 87 (2016) 339–348. [38] W. Li, H. Cai, Y. Kang, Y. Ying, J. Yu, J. Zheng, L. Qiao, Y. Jiang, S. Che, High
[14] W.Z. Mu, H.H. Mao, P.G. Jonsson, K. Nakajima, Effect of carbon content on the permeability and low loss bioinspired soft magnetic composites with nacre-like
potency of the intragranular ferrite formation, Steel Res. Int. 87 (2016) 311–319. structure for high frequency applications, Acta Mater. 167 (2019) 267–274.
[15] Y. Pan, J. Peng, L. Qian, Z. Xiang, W. Lu, Effects of compaction and heat treatment [39] Z. Guo, J. Wang, W. Chen, D. Chen, H. Sun, Z. Xue, C. Wang, Crystal-like
on the soft magnetic properties of iron-based soft magnetic composites, Mater. Res. microstructural Finemet/FeSi compound powder core with excellent soft magnetic
Express 7 (2020), 016115. properties and its loss separation analysis, Mater. Des. 192 (2020), 108769.
[16] M. Chen, M. Gao, F. Dang, N. Wang, B. Zhang, S. Pan, Tunable negative
permittivity and permeability in FeNiMo/Al2O3 composites prepared by hot-
pressing sintering, Ceram. Int. 42 (2016) 6444–6449.

29861

You might also like