Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Journal of Magnetism and Magnetic Materials 394 (2015) 470476

Contents lists available at ScienceDirect

Journal of Magnetism and Magnetic Materials


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

Effects of microstructure and CaO addition on the magnetic and mechanical properties of NiCuZn ferrites
Sea-Fue Wang n,1, Yung-Fu Hsu, Yi-Xin Liu, Chung-Kai Hsieh
Department of Materials and Minerals Resources Engineering, National Taipei University of Technology, Taiwan, ROC

art ic l e i nf o

a b s t r a c t

Article history:
Received 18 September 2014
Received in revised form
23 May 2015
Accepted 12 July 2015
Available online 13 July 2015

In this study, the effects of grain size and the addition of CaCO3 on the magnetic and mechanical
properties of Ni0.5Cu0.3Zn0.2Fe2O4 ceramics were investigated. The bending strength of the ferrites increased from 66 to 84 MPa as the grain size of the sintered ceramics decreased from 10.25 m to
7.53 m, while the change in hardness was insignicant. The addition of various amounts of CaCO3
densied the Ni0.5Cu0.3Zn0.2Fe2O4 ceramics at 1075 C. In the pure Ni0.5Cu0.3Zn0.2Fe2O4 ceramic, second
phase CuO was segregated at the grain boundaries. With the CaCO3 content Z 1.5 wt%, a small amount of
discrete plate-like second phase Fe2CaO4 was observed, together with the disappearance of the second
phase CuO. The grain size of the Ni0.5Cu0.3Zn0.2Fe2O4 ceramic dropped from 7.80 m to 4.68 m, and the
grain size distribution widened as the CaCO3 content increased from 0 to 5 wt%. Initially rising to 807
after CaCO3 addition up to 2.0 wt%, due to a reduced grain size, the Vickers hardness began to drop as the
CaCO3 content increased. The bending strength grew linearly with the CaCO3 content and reached twice
the value for the Ni0.5Cu0.3Zn0.2Fe2O4 ceramic with an addition of 5.0 wt% CaCO3. The initial permeability
of the Ni0.5Cu0.3Zn0.2Fe2O4 ceramic decreased substantially from 402 to 103 as the addition of CaCO3 in
ferrite increased from 0 to 5 wt%, and the quality factor of the Ni0.5Cu0.3Zn0.2Fe2O4 ceramic was maximized at 95 for 1.0 wt% CaCO3 addition.
& 2015 Elsevier B.V. All rights reserved.

Keywords:
Ferrites
NiCuZn
Additive
Magnetic properties
Mechanical properties

1. Introduction
There has been a growing interest in applying NiCuZn ferrites
to manufactured multilayer-type chips mainly because the oxides
can be sintered at relatively low temperatures with a wide range of
compositions [14]. Adding Cu into the ferrite composition in
particular has been known to play a crucial role in reducing the
ring temperature. Moreover, with better properties at high frequencies than MnZn ferrite and lower densication temperatures
than NiZn ferrites, NiCuZn ferrites are critical components in a
wide spectrum of electronic devices, such as cellular phones, video
cameras, notebook computers, and oppy drives [5,6].
In the past decade, several approaches have been reported in
the literature to tailor the physical properties and/or lower the
sintering temperatures of NiCuZn or NiZn ferrites, including using
various synthesis routes [5,7,8] and sintering proles [911] or
adding dopants [5,7,1216] and ux agents [1720]. Some of these
methods are used to modify the microstructural evolution through

changing the densication mechanism (liquid phase sintering) and


tailoring grain boundary properties and grain size. Some are
adopted to alter intrinsic properties like the magnetocrystalline
anisotropy constant, YafetKittel angle, AOB super-exchange
reaction, and defect structure. All, however, aim to trigger impacts
on the ferrites magnetic properties. Although the magnetic
properties of the ferrites have been investigated by several researchers, the mechanical properties of the ferrites remain understudied while it is well known that ferrite materials should be
mechanically robust enough so as to prevent damaging parts,
especially chipping their edges or corners, during machining and
assembly [2125].
In this study, the effects of the grain size and the addition of
CaCO3 on the magnetic and mechanical properties of the
Ni0.5Cu0.3Zn0.2Fe2O4 ceramics were investigated. The mechanical
properties, including Vickers hardness and bending strength and
the magnetic properties including the initial permeability and
quality factor of the sintered parts, were measured, analyzed and
discussed.

Corresponding author. Fax: 88 6 2 2731 7185.


E-mail address: sfwang@ntut.edu.tw (S.-F. Wang).
1
Present address: Department of Materials and Mineral Resources Engineering, National Taipei University of Technology 1, Sec. 3, Chung-Hsiao E. Rd., Taipei
106, Taiwan, ROC.
http://dx.doi.org/10.1016/j.jmmm.2015.07.037
0304-8853/& 2015 Elsevier B.V. All rights reserved.

2. Experimental procedure
In this study, Ni0.5Cu0.3Zn0.2Fe2O4 was prepared by a solid state

S.-F. Wang et al. / Journal of Magnetism and Magnetic Materials 394 (2015) 470476

reaction with high purity NiO, CuO, ZnO, CaCO3 and Fe2O3 (Fisher
Scientic, reagent grade) as the raw materials. Oxides, based on
the compositions of Ni0.5Cu0.3Zn0.2Fe2O4, were mixed and milled
in DI water using polyethylene jars and iron balls ( 5 mm) for
12 hours and then dried at 80 C in an oven overnight. After drying, the powders were calcined at 900 C for 2 h, re-milled in DI
water for 24 h, and then dried. A spinel structure was conrmed
based on X-ray diffraction analysis (XRD, RIGAKU DMX-2200) with
CuK radiation ( 1.5418 ), operating at 40 kV/30 mA. In order
to improve the mechanical properties, the grain sizes of the
Ni0.5Cu0.3Zn0.2Fe2O4 ceramics were monitored when the particle
sizes of the initial powders underwent changes. The calcined
Ni0.5Cu0.3Zn0.2Fe2O4 powders were again milled for different periods of time up to 24 h and then dried. The particle sizes of the
milled powders were analyzed by light scattering (Zeta 1000). To
investigate the effect of CaCO3 addition on the physical properties,
the calcined Ni0.5Cu0.3Zn0.2Fe2O4 powders were mixed and milled
with different amounts of CaCO3 for 24 h and then dried. Both the
Ni0.5Cu0.3Zn0.2Fe2O4 powders with different particle sizes and
those with CaCO3 addition were then added with a 3 wt% of 15%PVA solution and pelletized into disc-shaped samples with a diameter of 10 mm and toroid-shaped samples with an internal diameter of 21.97 mm, an external diameter of 35.84 mm, and a
height of 6.0 mm, using uniaxial pressure of 5 t/cm2. The samples
were then heat treated at 550 C for 4 hours to eliminate the PVA,
followed by sintering at temperatures from 1000 C to 1150 C for
3 hours (heating and cooling rates 5 C /min).
The densities of the specimens were measured based on the
liquid displacement method. Phase identication was performed
on the polished surfaces of the sintered disc samples using XRD.
The microstructure of the sintered surfaces was investigated using
scanning electron microscopy (SEM, Joel JSM-6510LV) at 15 KV,
coupled with energy dispersive X-ray analysis (EDAX). The grain
sizes of the samples were calculated from the SEM micrographs
with the line intercept method. To examine changes in magnetic
properties, the initial permeability () and quality factor (Q) were
measured using an HP4285 impedance analyzer at frequencies
ranging from 100 KHz to 8 MHz on the toroid-shaped samples. The
hysteresis loop and saturation magnetization of the sintered toroids were measured using a Vibrating Sample Magnetometer
(VSM) with a maximum magnetic eld strength of 71.2 T. The
measurements of magnetic properties were conducted at room
temperature. The hardness of the ferrites was measured using a
Vickers hardness tester (Akashi MVK-H1) on the polished surfaces
of the sintered samples with a load of 1 kg for 10 s to create cracks
with a suitable shape under indentation (a minimum of 10 indentations per sample). The Vickers hardness value (Hv) was determined according to the following equation [26]:

HV =

P
d2

where P was the applied load (N), d was the average of the diagonal length (m), and was the angle between the opposite faces of
the indenter (136). The exural strength (sc) at room temperature
was determined using a standard three-point method for testing
bending strength (ASTMC674) using a computer-assisted Cometech QC-506A2 universal testing machine, based on the following
equation [27]:

c =

3P L
2b d 2

where P was the applied load (N), L was the sample length (mm), b
was the sample width (mm), and d was the sample thickness
(mm).

471

Table 1
Physical characteristics of the Ni0.5Cu0.3Zn0.2Fe2O4 ceramics prepared from the
powders milled for different periods of time and pelletized and sintered at 1075 C
for 2 h.
Milling time (h)
D10(m)
D50(m)
D90(m)
Dmean(m)
Sintering condition (C/3 h)
Sintered density (g/cm3)
Grain size (m)
Vickers hardness (HV)
Bending strength (MPa)

0
0.61
1.10
1.91
1.19
1075
5.24
10.25
755
66

6
0.46
0.8
1.42
0.88
1075
5.26
9.45
763
72

12
0.40
0.63
0.99
0.67
1075
5.23
8.21
753
80

24
0.35
0.54
0.81
0.56
1075
5.26
7.53
770
84

3. Results and discussion


The particle sizes of the Ni0.5Cu0.3Zn0.2Fe2O4 powders after
milling in DI water for various lengths of time are listed in Table 1.
The milled powders emerged to show a narrow particle size distribution, and the mean particle size ranged from 1.19 m to
0.56 m as they were subjected to a second milling for 024 h. The
powders were pelletized, binder burned-out, and sintered at
temperatures between 1000 C and 1150 C. From the sintered
density results, maximum densication was achieved at the sintering temperature of 1075 C, and a difference in the sintered
densities of the ceramics was not evident. Fig. 1 shows the SEM
micrographs of the Ni0.5Cu0.3Zn0.2Fe2O4 ceramics prepared from
the powders ball-milled for various durations and sintered at
1075 C. The microstructures of the sintered Ni0.5Cu0.3Zn0.2Fe2O4
ceramics were marked with a signicant grain growth and the
presence of a second phase located along the grain boundaries.
Based on EDS analysis, the second phase appeared to be copper
oxide, an observation consistent with one reported in a previous
study [6]. Grain sizes of 10.25, 9.45, 8.21 and 7.53 m, determined
from the SEM micrographs, were obtained, respectively, for
Ni0.5Cu0.3Zn0.2Fe2O4 ceramics prepared from the powders after
milling for 0, 6, 12 and 24 h. Compared to the particle size of the
initial powders, the crystalline size of the Ni0.5Cu0.3Zn0.2Fe2O4
ceramics increased by more than an order in dimension. A low
densication temperature and huge grain growth observed in the
Ni0.5Cu0.3Zn0.2Fe2O4 ceramics may be attributed to the large content of Cu2 in the chemical formula [6,9], which induced the
formation of a liquid phase of copper compound at high temperatures and triggered liquid phase sintering.
Also listed in Table 1 are the mechanical properties of the
Ni0.5Cu0.3Zn0.2Fe2O4 ceramics prepared from the powders ballmilled for various durations and sintered at 1175 C. The Vickers
hardness of the ferrites, ranging from 753 to 770, reported no
signicant difference, primarily due to their similar sintered densities and grain sizes [23,24]. On the other hand, the bending
strength of the Ni0.5Cu0.3Zn0.2Fe2O4 ceramics rose from 66 to
84 MPa as the grain size of the sintered ceramics dropped from
10.25 to 7.53 m (i.e. as the milling time of the initial powders rose
from 0 to 24 h). A slight change in the microstructure would signicantly alter the bending strength but not the hardness of the
Ni0.5Cu0.3Zn0.2Fe2O4 ceramics. To evaluate the effects of CaCO3 on
the mechanical properties of the Ni0.5Cu0.3Zn0.2Fe2O4 ceramics, the
initial powders were milled with the additive for 24 h to signicantly reduce their particle size.
According to Fig. 2, which shows the sintered densities of the
Ni0.5Cu0.3Zn0.2Fe2O4 ceramics with various amounts of CaCO3
sintered at different temperatures, the maximum sintered density
occurred at the same sintering temperature (1075 C), suggesting
that the additive did not alter the densication mechanism. The
maximum density value declined slightly from 5.37 to 5.18 g/cm3

472

S.-F. Wang et al. / Journal of Magnetism and Magnetic Materials 394 (2015) 470476

Fig. 1. SEM micrographs of Ni0.5Cu0.3Zn0.2Fe2O4 ceramics prepared from powders ball-milled for (a) 0 h, (b) 6 h, (c) 12 h, and (d) 24 h and sintered at 1075 C for 2 h.

Fig. 2. Sintered densities of the Ni0.5Cu0.3Zn0.2Fe2O4 ceramics with various


amounts of CaCO3 sintered at different temperatures for 2 h.

as the CaCO3 content rose from 0 to 5.0 wt%, as listed in Table 1.


This is due to the fact that the density of CaO (3.34 g/cm3) is apparently less than that of the Ni0.5Cu0.3Zn0.2Fe2O4 ceramic
(5.26 g/cm3). The additive CaCO3 decomposed at high temperatures and produced CaO and CO2 gases. A decrease in the sintered
density associated with the addition of CaO was also observed in
the study on NiZn ferrite conducted by Ali et al. [28].
Fig. 3 presents the XRD patterns of Ni0.5Cu0.3Zn0.2Fe2O4 ceramics with various amounts of CaCO3 sintered at 1075 C. For the
pure Ni0.5Cu0.3Zn0.2Fe2O4 ceramic and those with additions of 0.5
and 1.0 wt% CaCO3, only a spinel phase structure (JCPDS #77-0012)
was obtained, and there was no other visible peak caused by the
second phase, though a small amount of second phase CuO was
present in the pure Ni0.5Cu0.3Zn0.2Fe2O4 ceramic, as shown in Fig. 1
(d). When the CaCO3 content was Z1.5 wt%, a small amount of
second phase Fe2CaO4 (JCPDS #74-2136) with an orthorhombic
structure was detected, which is different from the second phase
CaFeO3 observed in the study on NiZn ferrite by Reszlescu et al.

Fig. 3. XRD patterns of the Ni0.5Cu0.3Zn0.2Fe2O4 ceramics with various amounts of


CaCO3 sintered at 1075 C for 2 h.

S.-F. Wang et al. / Journal of Magnetism and Magnetic Materials 394 (2015) 470476

Fig. 4. Lattice constants of the Ni0.5Cu0.3Zn0.2Fe2O4 ceramics with various amounts


of CaCO3 sintered at 1075 C for 2 h.

[29]. Fig. 4 shows the lattice constant of Ni0.5Cu0.3Zn0.2Fe2O4


ceramics with additions of various amounts of CaCO3 based on
their XRD patterns; the values are also listed in Table 1. The lattice
constant increased from 8.3959 to 8.4168 for the pure
Ni0.5Cu0.3Zn0.2Fe2O4 ceramic with the addition of 2.0 wt% CaCO3.
With a further increase in the CaCO3 content, the lattice constant
began to drop and ended up at 8.3989 for the addition of 5.0 wt%

473

CaCO3. The initial increment of the lattice constant was mainly


because of the larger ionic radius of the Ca2 (IV; 1.0 ) ion
compared to those of Fe3 (VI; 0.645 ), Ni2 (VI; 0.69 ), Cu2
(VI; 0.73 ), Zn2 (IV; 0.60 ), and Fe3 (IV; 0.49 ). With the
CaCO3 content 4 1.5 wt%, the formation of second phase Fe2CaO4
drew the Ca2 ions away from the spinel lattices and led to a
decrease in the lattice constant [30]. The concentration of the Ca2
ions at the grain boundaries became greater than the concentration inside the grains.
Fig. 5 shows the SEM micrographs of Ni0.5Cu0.3Zn0.2Fe2O4
ceramics with various amounts of CaCO3 sintered at 1075 C. All
samples demonstrated a dense microstructure marked with little
trapped porosity. Compared to the microstructure of the pure
Ni0.5Cu0.3Zn0.2Fe2O4 ceramic (Fig. 5(a)), there was a signicant
variation in the microstructure associated with the increase in the
CaCO3 content. Similar to other additives, such as SiO2and Nb2O5,
CaO was observed in the NiZn and MnZn ferrites segregated at
the grain boundaries to form high-resistivity layers, which inuence the kinetics of grain boundary motion and pore mobility
[29,31,32]. As the CaCO3 addition increased from 0% to 5 wt%, the
grain size distribution was widened and the grain size of the
Ni0.5Cu0.3Zn0.2Fe2O4 ceramics was reduced from 7.80 m to
4.68 m (Fig. 6). In the pure Ni0.5Cu0.3Zn0.2Fe2O4 ceramic, the
second phase CuO was segregated at the grain boundaries and
formed a continuous phase surrounding the grains. The thickness
of the CuO second phase seemed to be reduced with the addition
of CaCO3, possibly due to the larger gain boundary area associated
with the smaller grain size. The precise values of the solubilities of

Fig. 5. SEM micrographs of the Ni0.5Cu0.3Zn0.2Fe2O4 ceramics with (a) 0 wt%, (b) 0.5 wt%, (c) 1.0 wt%, (d) 1.5 wt%, (e) 3.5 wt%, and (f) 5.0 wt% CaCO3 sintered at 1075 C for 2 h.

474

S.-F. Wang et al. / Journal of Magnetism and Magnetic Materials 394 (2015) 470476

Fig. 8. Vickers hardness and bending strength of the Ni0.5Cu0.3Zn0.2Fe2O4 ceramics


with various amounts of CaCO3 sintered at 1075 C for 2 h.
Fig. 6. Average grain size of the Ni0.5Cu0.3Zn0.2Fe2O4 ceramics with various
amounts of CaCO3 sintered at 1075 C for 2 h.

CaO in NiZn as well as in NiCuZn ferrites remain unknown [29].


However, it was observed that the addition of CaCO3 in amounts
greater than 1.5 wt% triggered the formation of discrete plate-like
second phase Fe2CaO4 accompanied with the disappearance of the
CuO network. The EDS results shown in Fig. 7 verify the rich calcium element along the second phase at the grain boundaries of
the Ni0.5Cu0.3Zn0.2Fe2O4ceramic with 5 wt% CaCO3, while absent at
the grain boundary of pure Ni0.5Cu0.3Zn0.2Fe2O4ceramics.
Fig. 8 shows the hardness and bending strength of
Ni0.5Cu0.3Zn0.2Fe2O4 ceramics with various amounts of CaCO3
sintered at 1075 C. The Vickers hardness increased from 756.6 to
807.0 as the CaCO3 addition rose from 0 to 2.0 wt% mainly due to
the reduced grain size (from 7.80 to 5.23 m). With a further

increase in the CaCO3 content, the hardness began to drop although the grain size continued to decrease. This indicated that
the formation of second phase Fe2CaO4 weakened the ferrite.
Different from the hardness, the bending strength escalated almost linearly from 66 to 140 as the CaCO3 content in the ferrites
rose from 0 to 5.0 wt%, reaching a value twice as large as the ferrite
without CaCO3 addition. This increase associated with the CaCO3
addition demonstrated that the bending strength was strongly
dependent on the grain size of Ni0.5Cu0.3Zn0.2Fe2O4 ceramics.
Fig. 9 reveals the initial permeability and quality factor versus
frequency of Ni0.5Cu0.3Zn0.2Fe2O4 ceramics with various amounts
of CaCO3 sintered at 1075 C, and their values at 1 MHz are also
listed in Table 2. The frequency of the ferrimagnetic resonance
slightly increased with the addition of CaCO3. As the frequency
rose, the initial permeability declined slightly before dropping to

Fig. 7. SEM micrographs of the Ni0.5Cu0.3Zn0.2Fe2O4 ceramics with (a) 0 wt% and (b) 5.0 wt% CaCO3 sintered at 1075 C for 2 h, and the associated EDS analysis along the
second phase at the grain boundary.

S.-F. Wang et al. / Journal of Magnetism and Magnetic Materials 394 (2015) 470476

475

Ni0.5Cu0.3Zn0.2Fe2O4 ceramics experienced a substantial change


with the addition of CaCO3, which is similar to the observations
reported in the study on NiZn ferrites [28]. The nonmagnetic
Ca2 ions in Ni  Zn spinel ferrite were proposed to preferentially
occupy A-sites [29,33]. The site distribution of Ca2 ions in the
Ni0.5Cu0.3Zn0.2Fe2O4 ceramic can be expected to be:

Zn Fe 3 +

3+
2
2+ 2+
0.2 0.8 x Ca x tet Ni0.5Cu0.3 Fe1.2 + x oct O4

Fig. 9. (a) Initial permeability and (b) quality factor versus frequency of the
Ni0.5Cu0.3Zn0.2Fe2O4 ceramics with various amounts of CaCO3 sintered at 1075 C
for 2 h.

obviously lower values at higher frequencies because of ferrimagnetic resonance [28]. The initial permeability of the
Ni0.5Cu0.3Zn0.2Fe2O4 ceramic decreased substantially from 402 to
103 as the addition of CaCO3 in ferrite increased from 0 to 5 wt%.
The decrease in the initial permeability of the NiZn ferrites can be
explained using the following equation [33]:

MS2 D

When Ca2 entered the ferrite lattice, it forced a larger fraction


of Fe3 ions to occupy B-sites, thus resulting in the magnetic dilution of the A-sites. As the fraction of the nonmagnetic ions (Zn2
and Ca2 ) on A-sites increased to a certain value, the AB interactions were much weaker than the BB interactions, and the
weakened interactions between the octahedral and tetrahedral
site lattices led to a division of the octahedral lattice into two
sublattices with a canted magnetic moment, which in turn resulted in a decrease in the saturation magnetization [34]. Except
for the reduction in grain size and saturation magnetization, the
decrease in the initial permeability of the ferrites with an escalated content of CaCO3 may be partly due to the presence of an
Fe2CaO4 compound, which diluted the initial permeability value.
The quality factors decreased rapidly with the rise in the
measurement
frequency.
The
quality
factor
of
the
Ni0.5Cu0.3Zn0.2Fe2O4 ceramic read 70 at 1 MHz and increased to 85
and 95, respectively, with the addition of 0.5 and 1.0 wt% CaCO3.
The quality factor then declined considerably with a further increase in the CaCO3 content, falling to 71, 62, 60, and 54 for the
Ni0.5Cu0.3Zn0.2Fe2O4 ceramics with the additions of 1.5, 2.0, 3.5 and
5.0 wt% CaCO3, respectively. The trend of the quality factor versus
the CaCO3 content is similar to the one reported by Ali et al. in
their study on NiZn ferrite (Ni0.3Zn0.7Fe2O4) [28], which, however,
showed a quality factor maximized at a 0.3 wt% CaO addition. The
increase in the quality factor may be due to the fact that CaCO3
addition increased the electrical resistivity of the second phase
layers around the Ni0.5Cu0.3Zn0.2Fe2O4 grains, and the limited solubility of Ca2 in the spinel lattices acted as domain wall pinning,
thereby decreasing the contribution of the eddy current loss.
Further increases in the CaCO3 content led to the formation of a
discrete plate-like Fe2CaO4 compound together with the disappearance of the second phase, which triggered the decline in
quality factor.
Based on the results discussed above, the Ni0.5Cu0.3Zn0.2Fe2O4
ceramic with the addition of 1.02.0 wt% CaCO3 demonstrated the
best properties summarized as follows: an initial permeability of
179-276, a quality factor of 62-95, a Vickers hardness of 782.3
807.0, and a bending strength of 8390 MPa. While retaining acceptable mechanical properties, these ferrites possess good magnetic properties for high frequency applications.

K1

where i is the initial permeability, MS is the saturation magnetization, D is the average grain size, and K1 is the magnetocrystalline
anisotropy constant. The average grain size (Table 2) of

4. Conclusions
In this study, the grain size of Ni0.5Cu0.3Zn0.2Fe2O4 ceramics was

Table 2
Physical characteristics of the Ni0.5Cu0.3Zn0.2Fe2O4ceramics with various amounts of CaCO3 sintered at 1075 C for 3 h.
Additive content Sintered density (g/cm3) Grain size (m) Lattice parameter () Saturation magnetization (emu/g) Permeability(1 MHz) Quality factor (1 MHz)
0 wt% CaCO3
0.5 wt% CaCO3
1.0 wt% CaCO3
1.5 wt% CaCO3
2.0 wt% CaCO3
3.5 wt% CaCO3
5.0 wt% CaCO3

5.37
5.35
5.32
5.30
5.25
5.19
5.18

7.80
6.88
6.10
5.50
5.23
4.77
4.68

8.3959
8.4078
8.4123
8.4139
8.4168
8.3965
8.3989

71.2
70.2
69.4
68.8
66.9
61.5
57.4

402
325
276
220
179
139
103

70
85
95
71
62
60
54

476

S.-F. Wang et al. / Journal of Magnetism and Magnetic Materials 394 (2015) 470476

monitored as the particle sizes of the initial powders underwent


changes. When the grain size of the sintered ceramics declined
from 10.25 to 7.53 m, the bending strength of the ferrites increased from 66 to 84 MPa, and the change in hardness was insignicant. All Ni0.5Cu0.3Zn0.2Fe2O4 ceramics with the addition of
CaCO3 reached densication at 1075 C. The microstructure of the
Ni0.5Cu0.3Zn0.2Fe2O4 ceramic was marked with second phase CuO
surrounding the grain boundaries at a lower CaCO3 content, and a
small amount of discrete plate-like second phase Fe2CaO4 was
observed with CaCO3 content greater than 1.5 wt%. The Vickers
hardness reached its maximum (807) with the addition of 2.0 wt%
CaCO3, and the bending strength registered twice the value when
5.0 wt% CaCO3 was added to the Ni0.5Cu0.3Zn0.2Fe2O4 ceramic. The
initial permeability of the Ni0.5Cu0.3Zn0.2Fe2O4 ceramic decreased
substantially from 402 to 103 and the quality factor varied from 54
to 95 as the CaCO3 content escalated from 0 to 5.0 wt%.

References
[1] J. Mrbe, J. Tpfer, NiCuZn ferrite for low temperature ring: I. ferrite
composition and its effect on sintering behavior and permeability, J. Electroceram. 15 (2005) 215221.
[2] H. Su, H. Zhang, X. Tang, Y. Jing, Y. Liu, Effect of composition and sintering
temperature on properties of NiZn and NiCuZn ferrites, J. Magn. Magn. Mater.
310 (2007) 1721.
[3] L.Z. Li, L. Peng, X.H. Zhu, D.Y. Yang, Effect of Cu and Co substitution on the
properties of NiZn ferrite thin lms, J. Electron. Sci. Technol. 10 (2012) 8892.
[4] M. Sugimoto, The past, present, and future of ferrites, J. Am. Ceram. Soc. 82
(1999) 269280.
[5] M.A. Gabal, A.M. Asiri, Y.M. AlAngari, On the structure and magnetic properties
of La-substituted NiCuZn ferrites prepared using egg-white, Ceram. Int. 37
(2011) 26252630.
[6] I.Z. Rahman, T.T. Ahmed, A study on Cu substituted chemically processed Ni
ZnCu ferrites, J. Magn. Magn. Mater. 290291 (2005) 15761579.
[7] K. Praveena, K. Sadhana, S. Srinath, S.R. Murthy, Effect of TiO2 on electrical and
magnetic properties of Ni0.35Cu0.12Zn0.35Fe2O4 synthesized by the microwavehydrothermal method, J. Phys. Chem. Solids 74 (2013) 13291335.
[8] A. Xia, C. Jin, D. Du, G. Zhu, Comparative study of structural and magnetic
properties of NiZnCu ferrite powders prepared via chemical coprecipitation
method with different coprecipitators, J. Magn. Magn. Mater. 323 (2011)
16821685.
[9] N.N. Cheng, Z. Wang, T.T. Liu, Improved magnetic softness for NiCuZn ferrite by
two-step sintering, IEEE Trans. Magn. 49 (2013) 41884191.
[10] M.P. Reddy, I.G. Kim, D.S. Yoo, W. Madhuri, N.R. Reddy, K.V.S. Kumar, R.
R. Reddy, Characterization and electromagnetic studies on NiZn and NiCuZn
ferrites prepared by microwave sintering, Mater. Sci. Appl. 3 (2012) 628632.
[11] H. Su, X. Tang, H. Zhang, Y. Jing, F. Bai, Z. Zhong, Low-loss NiCuZn ferrite with
matching permeability and permittivity by two-step sintering process, J. Appl.
Phys. 113 (2013) 17B301.
[12] B.D. Ingale, M.A. Barote, Preparation and characterization of Mn doped NiCuZn
ferrite, J. Chem. Biol. Phys. Sci. 3 (2013) 28012805.
[13] H. Su, H. Zhang, X. Tang, X. Xiang, High-permeability and high-Curie

temperature NiCuZn ferrite, J. Magn. Magn. Mater. 283 (2004) 157163.


[14] T.Y. Byun, S.C. Byeon, K.S. Hong, Factors affecting initial permeability of Cosubstituted Ni-Zn-Cu ferrites, IEEE Trans. Magn. 35 (1999) 34453447.
[15] A. Lucas, R. Lebourgeois, F. Mazaleyrat, E. Labour, Temperature dependence of
spin resonance in cobalt substituted NiZnCu ferrites, Appl. Phys. Lett. 97
(2010) 182502.
[16] J. Kato, K. Ono, Y. Matsuo, High frequency magnetic properties of Bi and Si
oxides-doped NiCuZn ferrite, Mater. Sci. Eng. 18 (2011) 092012.
[17] Y. Wang, H. Zhang, L. Li, Y. He, W. Ling, Effect of CaOB2O3SiO2 glass on the
magnetic and dielectric properties of NiCuZn ferrites, J. Magn. Magn. Mater.
324 (2012) 471474.
[18] R. Lebourgeois, S. Duguey, J.P. Ganne, J.M. Heintz, Inuence of V2O5 on the
magnetic properties of nickel-zinc-copper ferrites, J. Magn. Magn. Mater. 312
(2007) 328330.
[19] J.H. Jean, C.H. Lee, Processing and properties of low-re NiCuZn ferrite with
V2O5, Jpn. Soc. Appl. Phys. 40 (2001) 22322336.
[20] L.S. Chen, C.S. Hsi, S.L. Fu, J.Y. Lin, Cosintering of NiZnCu ferrite with lowtemperature cored ceramic substrates, Jpn. J. Appl. Phys. 39 (2000) 150154.
[21] G. Uzma, Micro-structural and microhardness study of NiZn ferrite using Si
additive, J. Sci. Res. 5 (2013) 415420.
[22] K. Praveena, K. Sadhana, S.R. Murthy, Elastic behavior of Sn doped NiZn
ferrites, Int. J. Sci. Res. Publ. 3 (2013) 14.
[23] R.V. Mangalaraja, S.T. Lee, K.V.S. Ramam, S. Ananthakumar, P. Manohar, Mechanical characterization of Ni1  xZnxFe2O4 prepared by non-conventional
methods, Mater. Sci. Eng. A 480 (2008) 266270.
[24] S. Beseniar, M. Dorfenik, T. Kosma, V. Kraevec, Magnetic and mechanical
properties of ZrO2 doped NiZn ferrites, IEEE Trans. Magn. 24 (1988)
18381840.
[25] A. Dias, R.L. Moreira, Chemical, mechanical and dielectric properties after
sintering of hydrothermal nickel-zinc ferrites, Mater. Lett. 39 (1999) 6976.
[26] D.R.R. Lazar, M.C. Bottinoa, M. Ozcan, L.F. Valandro, R. Amaral, V. Ussui, A.H.
A. Bressiani, Y-TZP ceramic processing from coprecipitated powders: a comparative study with three commercial dental ceramics, Dent. Mater. 24 (2008)
16761685.
[27] A. Akbari-Fakhrabadi, R.V. Mangalaraja, M. Jamshidijam, S. Ananthakumar, S.
H. Chan, Mechanical properties of GdCeO2 electrolyte for SOFC prepared by
aqueous tape casting, Fuel Cells 13 (2014) 682-668.
[28] N.J. Ali, J. Rahman, M.A. Chowdhury, The inuence of the addition of CaO on
the magnetic and electrical properties of NiZn ferrites, Jpn. Soc. Appl. Phys.
39 (2000) 33783381.
[29] E. Rezlescu, Effect of substitution of divalent ions on the electrical and magnetic properties of NiZnMe ferrites, IEEE Trans. Magn. 36 (2000)
39623967.
[30] H. Shokkrollahi, K. Janghorban, Inuence of additives on the magnetic Properties, Microstructures, and Densication of MnZn Soft Ferrites, Mater. Sci.
Eng. B 141 (2007) 91107.
[31] E. Rezlescu, N. Rezlescu, C. Pasnicu, M.L. Craus, D.P. Popa, The inuence of
additives on the properties of Ni-Zn ferrite used in magnetic heads, J. Magn.
Magn. Mater. 117 (1992) 448454.
[32] X.L. Fu, Q.K. Xing, Z.J. Peng, C.B. Wang, Z.Q. Fu, L.H. Qi, H.Z. Miao, Microstructure and electromagnetic properties of MnZn ferrites with low melting
point nonmagnetic Sb3 Ions, Int. J. Mod. Phys. 27 (2013) 1350003 (12 pages).
[33] M.S.R. Prasad, B.B.V.S.V. Prasad, B. Rajesh, K.H. Rao, K.V. Ramesh, Magnetic
properties and DC electrical resistivity studies on cadmium substituted nickelzinc ferrite system, J. Magn. Magn. Mater. 323 (2011) 21152121.
[34] G.K. Joshi, A.Y. Khot, S.R. Sawant, Magnetisation, Curie temperature and YK
angle studies of Cu substituted and non substituted NiZn mixed ferrites,
Solid State Commun. 65 (1988) 15931595.

You might also like