Mehdi 2019

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Journal of Magnetism and Magnetic Materials 491 (2019) 165597

Contents lists available at ScienceDirect

Journal of Magnetism and Magnetic Materials


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

Research articles

Non-oriented electrical steel with core losses comparable to grain-oriented T


electrical steel
Mehdi Mehdia,b, Youliang Hea,⁎, Erik J. Hilinskic, Narayan C. Kard, Afsaneh Edrisyb
a
CanmetMATERIALS, Natural Resources Canada, Hamilton, ON L8P 0A5, Canada
b
Department of Mechanical, Automotive, and Materials Engineering, University of Windsor, Windsor, ON N9B 3P4, Canada
c
Tempel Steel Co., Chicago, IL 60640-1020, USA
d
Department of Electrical and Computer Engineering, University of Windsor, ON N9B 3P4, Canada

ARTICLE INFO ABSTRACT

Keywords: A 3.2 wt% Si non-oriented electrical steel (NOES) was processed using conventional rolling and annealing routes,
Electrical steels i.e. hot rolling, hot band annealing, cold rolling, and final annealing. The evolution of texture during all the
Texture thermomechanical processing stages was investigated using electron backscatter diffraction (EBSD) techniques.
Core loss It was found that the final annealing temperature and holding time had a profound effect on the texture. At
Magnetic properties
higher temperatures and shorter holding times, the desired 〈1 0 0〉//ND (normal direction) texture was pro-
Magnetic Barkhausen noise
Recrystallization
moted, while the magnetically unfavorable 〈1 1 1〉//ND texture was weakened. The magnetic properties of the
Rolling steel sheets were measured by Epstein frame method and magnetic Barkhausen noise (MBN) analysis, and it was
Magnetocrystalline anisotropy found that the AC (alternating current) core losses of the NOES annealed at 850 °C for 60 min were comparable
EBSD to grain-oriented electrical steel (GOES) with the same silicon content and the same sheet thickness, even in the
rolling direction (RD). The DC (direct current) core losses and the MBN decreased with increasing grain size, due
to the decrease of the pinning sites (grain boundaries). The good magnetic properties of the processed NOES
were attributed to a combination of an optimized texture and an appropriate grain size, which was, in turn, the
result of proper hot band annealing, fast heating during annealing and a good selection of the annealing tem-
perature and holding time.

1. Introduction recrystallization (nucleation and grain growth) [6,7]. Although many


studies have been carried out to investigate the effect of thermo-
Non-oriented electrical steels (NOES) are the most widely used core mechanical processing on the crystallographic texture, and several
material for electric motors, generators and alternators, due to their models [6,7] and theories have been proposed to elucidate the under-
high magnetic permeability, high magnetization saturation, low core lying mechanisms, very limited improvement has been made to opti-
loss, and relatively low cost [1–5]. The magnetic properties of non-or- mize the final texture of non-oriented electrical steels using conven-
iented electrical steels are highly dependent on the silicon content, the tional processing schemes [7]. The final textures in both cold-rolled and
cleanliness of the steel, the thickness of the steel sheets, and the grain annealed non-oriented electrical steel sheets are normally dominated by
size and texture of the final laminations. While most of these properties the magnetically unfavorable 〈1 1 1〉//ND fibre (γ-fibre) [7]. On the
can be effectively controlled during the steel making and processing other hand, a number of unconventional processing techniques, e.g.
stages, the formation of the desired crystallographic texture in these two-stage cold rolling or cross-rolling [7,8], columnar grain growth
steels, i.e. the 〈1 0 0〉//ND fibre (θ-fibre), is proven to be challenging. [9–12], asymmetrical warm rolling [13], phase transformation an-
This is because all the thermomechanical processing procedures used to nealing [14,15], inclined rolling [16,17], skew rolling [18], etc., have
manufacture the final thin sheets will alter the texture, and the me- been developed in laboratory to optimize the texture of steels, but these
chanisms that govern the formation of textures in some processes (e.g. techniques are either very difficult to be implemented in industrial
recrystallization and phase transformation) are still not completely processes for mass production, or the results have not been validated
understood. The texture undergoes significant alterations during soli- under industrial conditions.
dification, phase transformation, plastic deformation as well as In this paper, a 3.2 wt% Si non-oriented electrical steel was


Corresponding author.
E-mail address: youliang.he@canada.ca (Y. He).

https://doi.org/10.1016/j.jmmm.2019.165597
Received 12 March 2019; Received in revised form 9 July 2019; Accepted 16 July 2019
Available online 17 July 2019
0304-8853/ Crown Copyright © 2019 Published by Elsevier B.V. All rights reserved.
M. Mehdi, et al. Journal of Magnetism and Magnetic Materials 491 (2019) 165597

processed using conventional manufacturing routes, i.e. casting, hot 2.2. Characterization of microtexture
rolling, hot band annealing, cold rolling and final annealing. The evo-
lution of texture and microstructure during these processes was char- The microtextures of the steel after all the thermomechanical pro-
acterized using electron backscatter diffraction (EBSD) techniques. The cessing stages were characterized by EBSD in a field emission gun (FEG)
steel after hot rolling was annealed at a relatively high temperature scanning electron microscope (Nova NanoSEM, FEI) equipped with an
(860 °C) for an extended time (24 h) to produce a microstructure that EDAX Orientation Imaging Microscopy (OIM 6.2) system. The EBSD
was believed to promote the formation of the θ-fibre texture during the samples were prepared using conventional metallographic procedures
subsequent cold rolling and annealing. The final annealing was con- plus a final polishing step using a 0.05 µm colloidal silica solution fol-
ducted at different temperatures (750 °C– 1150 °C) for different holding lowed by thorough ultrasonic cleaning. The EBSD scans after hot
times (3 h to 2 min) to vary the final microstructure and texture. rolling, hot-band annealing and cold rolling were performed on the RD-
The magnetic properties of the processed NOES and a commercial ND cross sections, which essentially covered the entire thickness of the
GOES were measured by standard Epstein frame method to obtain the plates/sheets. On the other hand, due to the large grain sizes after
bulk permeability and core loss, and by a relatively new technique, annealing, the EBSD scans on the final annealed samples were per-
magnetic Barkhausen noise (MBN) analysis [19–22], to characterize the formed on the middle-thickness planes of the steel sheets (i.e. RD-TD
local magnetic response. Magnetic Barkhausen noise is a signal origi- section), with approximately 3 mm × 3 mm scanned areas, in order to
nated from the discontinuous movements of magnetic domain walls obtain a good statistical representation of the texture. Orientation dis-
when ferromagnetic material is magnetized by an alternating external tribution functions (ODFs) were calculated using a harmonic series
magnetic field, which reveals the interaction between the domain walls expansion method with a series rank of 22 and a Gaussian half-width of
and the stress configurations and/or microstructures of the material 5°. The φ2 = 45° section of the Euler space (Bunge notation) was used
[22]. The microstructural features, e.g. phases, precipitates, inclusions, to represent the typical bcc texture fibres and components. The volume
dislocations, grain boundaries, textures, etc., will affect the MBN signal. fractions of the three major texture fibres, i.e. θ-fibre (〈1 0 0〉//ND), α-
MBN has been used to characterize the magnetic properties of both fibre (〈1 1 0〉//RD) and γ-fibre (〈1 1 1〉//ND) were calculated from the
GOES and NOES [23,24], and is especially useful for the evaluation of ODFs with a tolerance of 15°. The grain sizes were calculated using the
local magnetic properties, which is of particular interest for non-or- OIM software based on the grain orientation data [26].
iented electrical steels used in rotating machines [21,25]. Based on the
results of magnetic property measurements and microstructure/texture
2.3. Magnetic property measurements by Epstein frame method
characterization, the relationships between the grain size/texture of the
steel and the magnetic properties (e.g. core loss, permeability, satura-
Overall (bulk) magnetic properties of the final annealed steel were
tion magnetic induction, MBN, etc.) were discussed.
measured using standard Epstein frame method [27]. For each material,
a total of 20 rectangular strips (305 mm × 30 mm) were cut from the
steel sheets, with the longitudinal axis parallel to the rolling direction.
2. Material and experimental procedure
For the commercial GOES, an extra set of 20 strips were cut along the
transverse direction (TD), in order to measure and compare the mag-
2.1. Material and processing
netic properties in the TD. The bulk magnetic properties were measured
using a SMT-700 Soft Magnetic Material Testing System (Magnetic In-
The chemical composition of the non-oriented electrical steel in-
strumentation, Inc. [28]). Both DC and AC tests were conducted to
vestigated in this study is given in Table 1. The steel was melted in a
characterize the material’s magnetic behavior. For DC measurements, a
vacuum induction furnace and cast into ingots of
maximum current of 12 A, and a maximum induction of 2.5 T were
200 mm × 100 mm × 265 mm (width × thickness × length). The in-
utilized. The DC core loss, saturation induction, and maximum per-
gots were then reheated to 1050 °C and hot rolled to a thickness of
meability were obtained. The AC tests were carried out at similar
20 mm (80% thickness reduction) in a reversing rolling mill in six
testing conditions with frequencies varying from 3 Hz to 300 Hz. A
passes. The hot rolling starting temperature was ~945 °C and the fin-
summary of the testing parameters used in the DC and AC tests is given
ishing temperature was ~865 °C (the temperatures were measured on
in Table 2.
the surface of the steel plates using infrared thermometers). A second
hot rolling (also reheated to 1050 °C) was applied to reduce the thick-
ness from 20 mm to ~3.3 mm (~85% reduction) in six passes as well. In 2.4. Magnetic Barkhausen noise measurements
this case the starting and finishing temperatures were ~925 °C and
~680 °C, respectively. The steel plates were then pickled in a hydro- Local magnetic response of the electrical steel sheets was evaluated
chloric acid solution to remove surface oxides. by magnetic Barkhausen noise analysis using a Rollscan 350 MBN
The hot-rolled plates were subsequently annealed at 860 °C for 24 h Analysis System (Stresstech Oy, Finland [29,30]). The MBN is measured
in argon-protected atmosphere and furnace cooled. The annealed plates as a voltage signal (in mV) by placing a pick-up coil close to the surface
were cold rolled to a final thickness of 0.35 mm (~90% reduction). of the material while the material is magnetized by an external ex-
These were then annealed at various temperatures: 750, 850, 950, 1050 citation field with varying direction and/or intensity. The recorded
and 1150 °C, for different times: 180, 60, 30, 10 and 2 min, respectively.
The samples were inserted into a tube furnace that had been heated to Table 2
the designated temperatures so that the samples could be quickly he- Parameters used for the DC and AC Epstein frame tests.
ated up. The shorter holding times used for higher temperatures were Strip dimensions (length × width) 305 mm × 30 mm
intended to control the grain size so that the grain sizes of all the
Total Mass ~500 g
samples did not vary very significantly.
Density 7.65 g/cm3
H coil turns 700
Table 1 B coil turns 74
B coil resistance 0.6 Ω
Chemical composition of the investigated non-oriented electrical steel (wt%).
B step size 0.02 T
C Mn P S Si Al N O Fe Frequency (AC only) 3 Hz–300 Hz
AC waveform Sinusoidal
0.004 0.40 0.0095 0.0027 3.20 0.58 0.0034 0.0037 Balance Form factor 1.11

2
M. Mehdi, et al. Journal of Magnetism and Magnetic Materials 491 (2019) 165597

Fig. 1. Schematic illustration of the MBN measurements on the Epstein frame strip: (a) measuring locations along the rolling direction, (b) schematic of the MBN
sensor and the magnetization of the sample.

MBN signal is then processed and stored as digital data for further 3. Results
statistical analysis. To obtain a good representation of the magnetic
properties, 10 measurements were performed on one of the 20 Epstein 3.1. Hot rolling texture and microstructure
frame strips for each steel at 10 equally-spaced locations along the
central line in the RD (Fig. 1). The measured results were then averaged The cross-section (RD-ND) EBSD inverse pole figure (IPF) map of the
and compared to the DC core losses obtained from Epstein frame steel after hot rolling is illustrated in Fig. 2a. The microstructure and
method (both measurements were in RD only). All the MBN measure- crystal orientations show significant heterogeneity across the thickness
ments were conducted at a magnetizing frequency of 125 Hz and a peak of the plate. The surfaces (approximately 1/3 of the thickness) are
voltage of 5 V (default setting values of the MBN system). The MBN composed of small, equiaxed grains (average grain size ~ 13 µm), while
analyzing depth was estimated to be ~150 µm based on the resistivity the rest 2/3 of the plate consists of large, elongated (deformed) grains.
(6.41 µΩ cm) and relative permeability (~4000) of the 3.20 wt% Si Different grain orientations can be noticed in both regions: while the
steel [31,32]. Each MBN measurement consists of 20 bursts of MBN surface grains show a variety of different orientations (colors), the
signal, which was filtered to a frequency range of 0.1–1000 kHz. elongated grains only illustrate a few colors, i.e. essentially red
(〈0 0 1〉//ND), green (〈0 1 1〉//ND), and blue (〈1 1 1〉//ND). The
elongated grains can be further divided into two sub-areas: the central
area with mostly 〈0 0 1〉//ND (red) and 〈1 1 1〉/ND (blue) grains, and
the transition areas (between the surface and the central area) with
mainly 〈0 1 1〉//ND (green) grains.

Fig. 2. Microstructure and microtexture of the 3.2 wt% Si NOES after hot rolling: (a) EBSD IPF map of the RD-ND cross section, (b) texture of the surface regions, (c)
texture of the central region, (d) texture of the transition areas, (e) overall texture across the thickness. φ2 = 45° sections of the ODFs (Bunge notation).

3
M. Mehdi, et al. Journal of Magnetism and Magnetic Materials 491 (2019) 165597

Fig. 3. Microstructure and microtexture after hot rolling and annealing: (a) EBSD IPF map of the cross section (RD-ND plane), (b) φ2 = 45° section of the ODFs
(Bunge notation) showing the overall texture.

The texture of the surface regions (Fig. 2b) is mainly composed of deformed microstructure, the growth rate of these grains is fast since
copper ({1 1 2} 〈1 1 1〉 ) , brass ({1 1 0} 〈1 1 2〉 ) and Goss their growth is accomplished by consuming the neighboring deformed
({1 1 0} 〈0 0 1〉 ) components, which are common orientations resulted matrix, the latter having high stored energy, and the driving force being
from the friction (shear deformation) between the rolls and the plate large [37].
[33]. The central region (slightly more than 1/3 of the thickness) fea- The texture of the steel plate after hot band annealing is shown in
tures a strong θ-fiber (〈1 0 0〉//ND) (including cube and rotated cube) Fig. 3b. The texture is substantially altered after annealing: the θ-fibre
and a weak γ-fibre (〈1 1 1〉//ND) texture (Fig. 2c). The texture of the (〈1 0 0〉//ND) in the hot-rolled steel (Fig. 2e) essentially disappears,
transition area (Fig. 2d) is mainly composed of a strong Goss, a copper while a strong {1 1 4} 〈2 6 1〉 component is produced, which is close to
and a brass. There is also a weak {0 0 1} 〈1 2 0〉 component on the θ- the α*-fibre ({1 1 h} 〈1 2 1/h〉) [38]. There is also a
fibre. The overall texture of the cross section (Fig. 2e) is featured by a θ- {1 1 3} 〈1 1 0〉 component on the α-fibre. The γ-fibre becomes con-
fibre and some weak components such as brass, copper, Goss and γ- tinuous after annealing. The copper, brass and Goss in the hot-rolled
fibre. The intensities of these components, however, are much lower steel have been significantly weakened or completely eliminated.
than those in the individual regions. Both the texture and the micro- It is worth noting that most of the recrystallized grains in the mi-
structure of the hot-rolled NOES are very similar to the rolling textures crostructure (Fig. 3a), particularly those in the central region, are not
and microstructures of interstitial free (IF) steels [33–36]. equiaxed, i.e. they show apparent lateral discrepancies in the RD and
TD (i.e. larger in the RD than TD), indicating that the grain growth
preferentially occurs in the rolling direction. In fact, the average in-
3.2. Microstructure and texture after hot band annealing tercept length measured from 200 horizontal lines parallel to the RD
(i.e. the average grain diameter in RD) is ~190 µm, while that mea-
After annealing at 860 °C for 24 h, the microstructure and micro- sured in the ND is only ~105 µm. This suggests that the newly formed
texture of the hot-rolled plate changed significantly (Fig. 3a). Both the nuclei in the central region more likely grow by consuming the elon-
surface and central regions are composed of recrystallized grains, but gated (deformed) grains in the rolling direction during grain growth. As
microstructural variation still exists across the thickness, i.e. the grains have been shown in Fig. 2a, these elongated grains usually have grain
near the surface regions are generally smaller than those at the central boundaries essentially parallel to the RD, thus the recrystallized grains
region. The average grain size is approximately 123 µm, but the in- grown within them usually also show longitudinal axes parallel to the
dividual grains show large deviations, i.e. grains in the range of tens of RD.
microns to ~500 µm can all be seen in the microstructure. The variation
in grain size is due to the fact that before annealing (Fig. 2a), the sur-
face grains have already recrystallized during hot rolling, and they es- 3.3. Cold rolling microstructure and texture
sentially only undergo grain growth during the subsequent annealing
(nucleation is essentially prohibited due to the lack of strain). The low The microstructure and microtexture after cold rolling are shown in
stored energy in these grains limited their growth since all the neigh- Fig. 4a. Cold rolling results in the elongation of the originally re-
boring grains already recrystallized and the driving force was low [37]. crystallized grains (Fig. 3a). In some grains, a large amount of shear
On the other hand, the central region is composed of deformed grains bands can be seen, and the accumulation of dislocations in these regions
after hot rolling, which have higher stored energy than the re- makes the indexing difficult, thus they may appear black in the IPF
crystallized grains near the surfaces. Once new grains nucleate from the map. In other grains, the formation of shear bands is not obvious, but

4
M. Mehdi, et al. Journal of Magnetism and Magnetic Materials 491 (2019) 165597

Fig. 4. Microstructure and texture of the cold-rolled 3.2 wt% Si electrical steel: (a) EBSD IPF map of the RD-ND cross section, (b) φ2 = 45° section of the ODFs (Bunge
notation) showing the overall texture across the thickness.

there are a lot of microbands showing different orientations (colors), cube and the α-fibre in the cold-rolled steel essentially disappear. Si-
which may be the result of grain fragmentation (intra-grain rotation) milarly, a near α*-fibre is seen to extend from the θ-fibre to the γ-fibre.
during heavy plastic deformation [38]. These locations are considered When the annealing temperature is 950 °C and the holding time is
as very important nucleation sites during the subsequent recrystalliza- 30 min, the average grain size increases to ~ 160 µm (Fig. 5c), with a
tion process. spread between ~ 30 µm and ~ 500 µm. The texture (Fig. 5h) is again
Although the crystal orientations in the annealed hot band show composed of a {0 0 1} 〈1 4 0〉 component (intensity 6.0) on the θ-fibre
large variations (Fig. 3a), the texture after ~85% cold rolling reduction and a {111 〈1 1 2〉 (intensity 3.2) on the γ-fibre. A weak Goss texture
(Fig. 4b) is essentially composed of three fibres, i.e. 〈0 0 1〉//ND (θ- ({1 1 0} 〈0 0 1〉 ) starts to appear. The rotated cube and α-fibre are es-
fibre), 〈1 1 0〉//RD (α-fibre) and 〈1 1 1〉//ND (γ-fibre), which are ty- sentially eliminated, and there also exists a near α*-fibre.
pical textures commonly observed in rolled bcc iron [34]. The max- If the annealing is conducted at 1050 °C for 10 min, the average
imum intensity (19.2) is observed at the rotated cube ({0 0 1} 〈1 1 0〉 ) grain size increases to ~200 µm (Fig. 5d), with a spread of 38–550 µm.
orientation, and a strong {1 1 1} 〈1 1 0〉 on the α-fibre (intensity 9.8) is The texture (Fig. 5i) is also featured by a {0 0 1} 〈1 4 0〉 on the θ-fibre
also noted. (intensity 5.0) , a {1 1 1} 〈1 1 2〉 on the γ-fibre (intensity 2.4), and an
α*-fibre. The α-fibre and the rotated cube in the cold-rolled steel are
completely eliminated. If the annealing temperature is further increased
3.4. Microstructure and texture after final annealing to 1150 °C and the time is reduced to 2 min, the average grain size drops
slightly to ~187 µm (Fig. 5e), with a spread of 35–525 µm. The de-
As has been seen in Fig. 4a, the microstructure after cold rolling crease of the average grain size as compared to annealing at 1050 °C for
consists of a lot of shear bands and microbands, which are preferred 10 min may be due to the short holding time, which limits the grain
nucleation regions during recrystallization. It has been shown [39,40] growth. The main texture is still the {0 0 1} 〈1 4 0〉 component, but the
that, within these regions, there are crystallites or crystal volumes that intensity drops to 4.3 (Fig. 5j). However, the γ-fibre is essentially
have lower stored energy than their surroundings, and these crystallites eliminated, and a cube texture (intensity 2.6) is observed, which is not
(volumes) form the initial nuclei, which grow by consuming the sur- seen in any of the samples annealed at other temperatures. The rotated
rounding high-energy domains. As the recrystallization continues, the cube and the α-fibre are eliminated. Again, there is also an α*-fibre.
growing grains also consume the deformed matrix. The driving force for Therefore, although the annealing temperature and holding time are
the growth of the newly formed grains is the reduction of the stored significantly different, the resulted recrystallization textures all contain
energy accumulated during plastic deformation. Depending on the an- a {0 0 1} 〈1 4 0〉 component, a {1 1 1} 〈1 1 2〉 component and a near α*-
nealing temperature and holding time, the annealed microstructure and fibre [38]. The main differences among the textures are the relative
texture show significant discrepancies (Fig. 5). intensities of the {0 0 1} 〈1 4 0〉 component on the θ-fibre and the
Fig. 5a shows the microstructure of the sample after annealing at {1 1 1} 〈1 1 2〉 component on the γ-fibre. Annealing at 750 °C for
750 °C for 180 min. The average grain size is ~82 µm, and there exists a 180 min produced the strongest γ-fibre and the weakest
large variation in the grain diameter, i.e. 10–250 µm. The texture {0 0 1} 〈1 4 0〉 component, while annealing at 850 °C for 60 min created
(Fig. 5f) is featured by a strong γ-fibre and a relatively strong {0 0 1} the strongest {0 0 1} 〈1 4 0〉 component and a relatively weak
〈1 4 0〉 component on the θ-fibre. Compared to the cold rolling texture {1 1 1} 〈1 1 2〉 component. Annealing at 1150 °C for 2 min can essen-
shown in Fig. 4b, the strong rotated cube and the α-fibre disappear, tially eliminate the magnetically unfavorable γ-fibre, although the in-
while the γ-fibre persists. A partial fibre close to the α*-fibre is noted tensity of the {0 0 1} 〈1 4 0〉 component is also decreased.
stretching from the θ-fibre to the γ-fibre. The variations of the volume fractions of the three major fibre
When the annealing temperature is increased to 850 °C and the textures (α, θ, and γ) with respect to the annealing temperature are
holding time is reduced to 60 min, the average grain size is increased to summarized in Fig. 6. It is seen that annealing at all the temperatures
~103 µm (Fig. 5b), also with a considerable spread (20–300 µm). The significantly reduced the volume fraction of the α-fibre (compared to
prominent texture is now the {0 0 1} 〈1 4 0〉 component on the θ-fibre, that after cold rolling), while the θ-fibre and γ-fibre may be strength-
with an intensity of 8.7 (Fig. 5g). The γ-fibre changes to only one weak ened or weakened depending on the annealing temperature. Annealing
{1 1 1} 〈1 1 2〉 component with an intensity of 2.8. Again, the rotated

5
M. Mehdi, et al. Journal of Magnetism and Magnetic Materials 491 (2019) 165597

Fig. 5. Microstructure and texture of the final annealed steel: (a)-(e) EBSD IPF maps showing the microstructure after annealing at 750 °C for 180 min, 850 °C for
60 min, 950 °C for 30 min, 1050 °C for 10 min and 1150 °C for 2 min, respectively, (f)-(j) φ2 = 45° sections of the ODFs (Bunge notation) showing the corresponding
textures. The EBSD scans were performed on the sheet plane at the middle thickness.

Fig. 7. Relations among the annealing temperature, average grain size and the
DC core loss.

Fig. 6. Variations of the common texture fibres (α, θ and γ) with respect to the
annealing temperature leads to a slight increase of the DC core loss.
annealing temperature. Thus the DC core loss generally decreases with the increase of the
average grain size.

at 850 °C and 1050 °C can promote the θ-fibre texture, while annealing
at other temperatures, the θ-fibre volume fraction essentially does not 3.5.2. AC magnetic properties
change, i.e. the same as the cold-rolled steel. The volume fraction of the AC core losses were measured at two frequencies (50 and 250 Hz),
γ-fibre, on the other hand, generally decreases with the increase of the and the steel strips were magnetized to saturation. Depending on the
annealing temperature. frequency, the saturation flux density is different: for the low frequency
(50 Hz), the saturation magnetic flux density reaches ~1.68 T, while
that of the high frequency (250 Hz) is only ~1.38 T. The results at a
3.5. Magnetic properties magnetic flux density of 1.0 T for both frequencies are shown in Fig. 8.
The AC core losses at low and high frequencies behave slightly differ-
3.5.1. DC core loss ently with respect to the annealing temperature: at a low frequency, a
The measured DC core losses for samples annealed at different minimum loss is observed at 850 °C, and further increasing the an-
temperatures are illustrated in Fig. 7, where the losses are also com- nealing temperature gradually increases the core loss; at a high fre-
pared to the average grain sizes of the samples. It is seen that with the quency, the minimum loss is observed at 750 °C, and it gradually in-
increase of the annealing temperature, the average grain size gradually creases with the annealing temperature. The AC losses at 250 Hz are
increases until 1050 °C, after which the grain size slightly drops when approximately an order of magnitude higher than those at 50 Hz.
the annealing temperature is further increased to 1150 °C. The DC core The saturation magnetic flux densities of all the annealed samples at
loss, on the other hand, gradually decreases with the annealing tem- DC and AC (50 Hz) magnetizing conditions are shown in Fig. 9a. Gen-
perature. At 1050 °C, a minimum is reached, and further increasing the erally, the DC saturation magnetization is consistently higher than the

6
M. Mehdi, et al. Journal of Magnetism and Magnetic Materials 491 (2019) 165597

of the NOES is ~103 µm (Fig. 5b). As can be seen from Fig. 10a, if the
magnetic flux density is smaller than ~1.25 T, the core losses of the
NOES are comparable to (slightly higher than) those of the commercial
GOES even in the rolling direction. When the magnetic flux density is
greater than ~1.25 T, the core losses of the NOES are apparently higher
than those of the GOES. It is also noted that the GOES has apparent
anisotropy between the rolling direction and the transverse direction,
i.e. the core losses in the TD are considerably higher than those in the
RD.
The core losses of both steels at different frequencies (under a fixed
magnetic flux density of 1.0 T) are quite close with each other
(Fig. 10b), i.e. the core losses of the NOES produced in this study are
also comparable to those of the commercial GOES at all the frequencies
if the magnetic flux density is relatively small (1.0 T). In fact, the core
losses are even slightly lower than those of the commercial GOES (RD)
when the frequency is higher than 225 Hz.

3.5.4. Magnetic Barkhausen noise


The average MBN root mean square (rms) value [21,30] of the 10
measurements along the rolling direction (Fig. 1a) of each steel strip is
Fig. 8. The variations of AC core losses at magnetizing frequencies of 50 Hz and
250 Hz with respect to annealing temperature. plotted against the annealing temperature in Fig. 11a, where the
MBNrms is also compared to the average grain size. It is seen that the
MBNrms drops significantly when the annealing temperature increases
AC saturation magnetization at all the annealing temperatures, but both from 750 °C to 1050 °C. After that, the MBNrms slightly increases with
show essentially the same trend when the annealing temperature in- the increase of the annealing temperature. This shows essentially the
creases. In each case, a minimum saturation magnetization is observed same trend as the DC core loss with respect to annealing temperature
at 850 °C, and at all the other temperatures, the saturation magnetic and grain size (Fig. 7). If the MBNrms is plotted against the square root
flux densities are essentially the same. Consequently, although the AC of the average grain size d (Fig. 11b), an approximate linear relation-
loss has a minimum after annealing at 850 °C for 60 min (with the ship is observed between the MBNrms and the d1/2, which is very similar
strongest {0 0 1} 〈1 4 0〉 texture), the saturation magnetization is also to the well-known Hall-Petch relationship between the yield strength
the minimum. On the other hand, the relative permeability (AC, 50 Hz) and the grain size observed in structural materials [24,37].
of the electrical steel (Fig. 9b) shows an opposite trend with respect to
the annealing temperature, i.e. the relative permeability has a max- 3.5.5. Magnetocrystalline anisotropy
imum at 850 °C, where the magnetically favourable The magnetocrystalline anisotropy of the electrical steel can be
{0 0 1} 〈1 4 0〉 texture has the highest intensity. characterized using a texture factor [7,21,36] defined as the minimum
angle between the magnetization direction M and the three easy axes
3.5.3. AC core loss compared to commercial grain-oriented electrical steel (〈1 0 0〉 ) of the crystal (Fig. 12), weighted by the intensity of the crystal
The magnetic properties (rolling direction) of the non-oriented orientation, which is expressed as:
electrical steel produced in this study (annealed at 850 °C for 60 min) A = f (g ) A (g ) dg (1)
are compared to those of commercially available grain-oriented elec-
trical steel in Fig. 10. Both steels contain essentially the same amount of where A (g ) is the minimum angle between the magnetization vector
silicon (3.2 wt%), and the thicknesses of the sheets are all ~ 0.35 mm. and the three easy 〈1 0 0〉 axes of a given orientation (g), θ is the angle
However, the microstructure and texture of the two steels are sig- between the magnetization direction and the sample coordinate system
nificantly different. The GOES has essentially a single Goss (RD in this study), and f(g) is the orientation distribution function. For
({0 1 1} 〈1 0 0〉 ) texture and very large grain sizes (i.e. in the range of each angle (θ), the texture factor can be readily calculated using Eq. (1)
centimeters), while the NOES shows the strongest texture at by rotating the ODF around the ND by the θ angle. The average texture
{0 0 1} 〈1 4 0〉 with an intensity of 8.7 (Fig. 5g). The average grain size factor is then calculated from the texture factors at all the directions

Fig. 9. The variations of maximum saturation flux density and relative permeability with respect to annealing conditions: (a) maximum saturation flux density at DC
(Bs) and AC (B50) magnetization, and (b) relative permeability (µr).

7
M. Mehdi, et al. Journal of Magnetism and Magnetic Materials 491 (2019) 165597

Fig. 10. Magnetic properties of the 3.2 wt% Si non-oriented electrical steel produced in this study compared to those of a commercial grain-oriented electrical steel:
(a) core losses at different magnetic flux densities (with a fixed frequency of 125 Hz), (b) core losses at various frequencies (with a fixed magnetic flux density of
1.0 T).

Fig. 11. The average MBNrms values measured in the rolling direction of the steel strips: (a) relation between MBNrms and annealing temperature, (b) relation
between MBNrms and the square root of the average grain size (d).

(θ = 0°–360°). Apparently, the smaller the texture factor, the closer the RD and TD, i.e. at 45°/225° and 135°/315°. The minimums were seen
crystals’ easy axes are to the magnetization vector. close to RD and TD, i.e. 0°/180° and 90°/270°. The GOES (Fig. 13f), on
Fig. 13 illustrates the texture factors (at all the directions in the the other hand, shows an hourglass shape with a narrow neck in the
sheet plane) of the NOES sheets annealed at different temperatures rolling direction. The very small texture factor (close to 0°) in the
compared to those of the GOES. It is seen that the texture factors rolling direction is due to the unique Goss texture ({0 1 1} 〈1 0 0〉 ) in
(Fig. 13a–e) for all NOES sheets illustrate a similar quatrefoil shape, this type of steel, i.e. the minimum angle between the magnetization
with the maximum values falling approximately half way between the vector (the RD) and the crystal 〈1 0 0〉 axes is zero. Apparently, the

Fig. 12. Schematic illustration of the texture factor Aθ (g) defined as the minimum angle between the magnetization vector (M ) and the crystal easy magnetization
axes [1 0 0], [0 1 0] and [0 0 1]. Since α1 is the minimum angle in this example (α1 < α2 < α3), Aθ (g) would be equal to α1.

8
M. Mehdi, et al. Journal of Magnetism and Magnetic Materials 491 (2019) 165597

Fig. 13. The texture factors at various directions to the RD (0°) for the NOES sheets annealed under different conditions: (a) 750 °C for 180 min, (b) 850 °C for 60 min,
(c) 950 °C for 30 min, (d) 1050 °C for 10 min, (e) 1150 °C for 2 min, and (f) commercial GOES.

NOES shows much less magnetocrystalline anisotropy than the GOES [48] reported that a higher heating rate led to grain refinement as well
since the variation of the texture factor in all the directions is much as the weakening of the 〈1 1 1〉//ND texture. A higher annealing tem-
smaller than that of the GOES. perature also promotes the nucleation of the θ-fibre grains. The results
shown above (Figs. 5 and 6) indicate that the volume fraction of the γ-
4. Discussion fibre decreased with the increase of the annealing temperature.
It is well known that the core loss (per hysteresis cycle) of electrical
It has been seen that the crystallographic texture of the 3.2 wt% Si steel at a given frequency f is composed of three major parts [3,5,49]:
NOES processed in this study has been considerably improved (i.e. the W (f )
θ-fibre is strengthened and the γ-fibre is weakened or eliminated) by = Wh + Wcl (f ) + Wexc (f )
f (2)
combining hot band annealing at 860 °C for 24 h and final annealing at
850 °C or higher. The prolonged hot band annealing leads to the for- where Wh is the quasi-static hysteresis loss, Wcl (f ) is the classical eddy
mation of a heterogeneous microstructure across the plate thickness current loss, and Wexc (f ) is the excess loss. The hysteresis loss Wh is
(Fig. 3a) and a randomized texture prior to cold rolling (Fig. 3b). Lee independent of the frequency (f), while the other two losses are closely
et al. [41] has reported that the coarse grains after hot band annealing related to the frequency.
can slow down the recrystallization process after cold rolling, which The hysteresis loss is directly related to the domain wall motion
retards the formation of the 〈1 1 1〉//ND orientations after final an- during the magnetization and demagnetization process under quasi-
nealing, and promotes the 〈0 0 1〉//ND and 〈1 1 3〉//ND textures. On static conditions, which can be evaluated using the Steinmetz empirical
the other hand, de Dafe et al. [42] reported that coarse grains after hot formula [5,49,50].
band annealing promote the formation of magnetically favourable
crystal orientations (i.e. 〈0 0 1〉//ND), due to the increased amount of Wh = HdB Kh (Bm ) x (J /m3) (3)
the shear bands in the deformed microstructure. It is well known that
where Bm is the peak flux density, and Kh and x are material-dependent
nucleation preferably starts at regions with microstructural hetero-
parameters.
geneities (high stored energies), e.g. shear bands, microbands, etc.
Eddy current loss is dependent on the thickness of the steel sheet, t,
[37,43–46]. This will suppress the nucleation at grain boundaries that
the electrical resistivity, , the peak flux density, Bm, as well as the
usually leads to the γ-fibre texture. Thus, increasing the amounts of
magnetizing frequency, f, which can be calculated as [3,5,49]:
shear bands, microbands or crystal fragmentations leads to the forma-
tion of the magnetically favorable θ-fibre components that are normally 2
Wcl (f ) = t 2Bm2 f (J / m3)
difficult to form from the nucleation at grain boundaries [41,42]. k (4)
On the other hand, the annealing in this study was conducted by
inserting the samples into a furnace already heated to the designated where k is a magnetization waveform dependent constant.
temperature, so the heating rate is faster than that if the sample was The origin of the excess loss remains a subject of debate in the lit-
heated up with the furnace [47]. Since the 〈0 0 1〉//ND grains have a erature. Some researchers have attributed the excess loss to the motion
lower stored energy than the 〈1 1 1〉//ND grains [37], their nucleation of large magnetic domains, and thus believe that Wh and Wexc have a
usually requires a higher temperature to occur. The rapid heating rate similar origin. The excess loss can be evaluated as:
allows the material to reach the required temperature quickly before f Bm3/2 (J / m3)
Wexc (f ) = kexc (5)
the deformed 〈0 0 1〉//ND grains are consumed by other recrystallized
grains, leading to the formation of the 〈0 0 1〉//ND nuclei. Wang et al. where σ = 1/ρ is the conductivity, and kexc is a material parameter

9
M. Mehdi, et al. Journal of Magnetism and Magnetic Materials 491 (2019) 165597

related to the domain structure and the structural properties of the evolution of microstructure and texture was characterized by EBSD.
material [3,5,49]. The magnetic properties of the processed steel were measured by
It has been pointed out in the literature [7,51,52] that the core loss Epstein frame method and magnetic Barkhausen noise analysis, which
of electrical steels is more dependent on the grain size than on the were compared to those of a commercial grain-oriented electrical steel.
texture. The DC core loss shown in Fig. 7 indicated that the DC core loss The findings of this research can be summarized as follows:
was indeed closely related to the grain size, i.e. with the increase of the
grain size, the DC core loss decreases. This may be attributed to the • The texture of the 3.2 wt% Si non-oriented electrical steel was im-
decreased grain boundaries, which resulted in the decrease of the pin- proved through hot band annealing at 860 °C for 24 h, fast heating
ning effect during the magnetization and demagnetization processes. during final annealing, as well as an appropriate combination of the
This trend was also reflected in the MBN response (Fig. 11), i.e. the annealing temperature and holding time.
fewer the grain boundaries, the fewer the pinning sites for domain wall • Annealing the steel at a temperature higher than 750 °C resulted in
motion, and the smaller the MBN. The texture, on the other hand, did the increase of the magnetically favourable θ-fibre and suppressing
not show an apparent effect on the DC core loss or the MBN. the magnetically detrimental γ-fibre. This was due to the preferred
Although Eq. (4) does not show an explicit relation between the nucleation at shear bands, microbands and fragmented grains,
eddy current loss and the grain size, it has been reported by a few re- especially after obtaining coarse grains during hot band annealing.
searchers [3,5,53–55] that the eddy current loss increases with the • The DC core losses showed a good correlation with respect to the
grain size. Because of the opposite relation between the hysteresis loss grain size, i.e. the larger the grain size, the smaller the core loss, due
and the grain size, there exists an optimal grain size at which the total to the decrease of grain boundaries that act as pinning sites during
loss induced by hysteresis and eddy current is minimum, e.g. the sample magnetization and demagnetization. The MBN response also
annealed at 850 °C for 60 min as shown in Fig. 8a. However, this is true showed a similar relation with respect to the grain size.
only for low frequency excitations, since at high frequencies, the eddy • The AC core loss at a low frequency (50 Hz) showed the lowest value
current loss (which is proportional to f 2 ) will dominate the total loss, in the sample annealed at 850 °C for 60 min, and the relative per-
thus a smaller grain size would result in a smaller core loss [3,5,53–55]. meability (at 1.0 T) also showed the highest value in this sample.
The sample annealed at 750 °C for 180 min has the smallest grain size These might be mainly due to an optimal grain size obtained in this
(~82 µm), thus it has the lowest core loss at 250 Hz. sample, although this sample also showed a strong 〈0 0 1〉//ND
Although it has been shown in Fig. 5 that, when the annealing texture.
temperature was increased, the texture was optimized, i.e. the θ-fibre • For high frequency applications, it was shown that the grain size
was promoted and the γ-fibre was weakened, the AC core loss did not was the dominant factor. So the lowest core loss was obtained in the
show apparent improvement with the texture optimization, i.e. the sample with the smallest grain size, i.e. the one annealed at 750 °C
variation of the core loss was quite small no matter what was the tex- for 180 min.
ture. As shown in Fig. 13, the texture factors (in RD) of the steels an- • The sample annealed at 850 °C for 60 min showed AC core losses
nealed at different temperatures vary from 26.0° to 29.3°, and the core comparable to (only slightly higher than) those of the commercial
losses of these steels vary from 0.94 W/kg to 1.025 W/kg. The lowest grain-oriented electrical steel at low frequencies (<225 Hz), and
texture factor does not correspond to the lowest core loss. In fact, it is even lower losses at higher frequencies (>225 Hz).
the opposite, i.e. the largest texture factor results in the lowest core loss
(annealing at 850 °C for 60 min). Apparently, the dominant factor here Acknowledgements
is not the crystallographic texture, i.e. other microstructural features
(e.g. grain size) may have dominated the core loss (e.g. the excess loss) Funding of this research was provided by the Program of Energy
[7,52,56,57]. Similarly, the permeability of the electrical steel is also Research and Development (PERD), Natural Resources Canada Natural
affected by both the grain size and texture. At low inductions (i.e. below Sciences, and by Natural Sciences and Engineering Research Council of
about 1.0 T), the permeability is predominately influenced by grain size Canada (NSERC). Peter Newcombe, Douglas McFarlan, Howard
instead of texture, while under high inductions (i.e. 1.5 T or more), the Webster, and David Saleh are acknowledged for casting the steel.
texture has a major effect [52]. Thus the largest permeability observed Michael Attard and Mason Thomas are thanked for carrying out the hot
in the sample annealed at 850 °C for 60 min (Fig. 9b) might also be due and cold rolling of the steel. Renata Zavadil and Jian Li are gratefully
to an optimal grain size. acknowledged for their assistance in EBSD measurements.
Nonetheless, it has been shown that the core losses of the non-or-
iented electrical steel annealed at 850 °C for 60 min were comparable Appendix A. Supplementary data
to, or even better than, those of the commercially available grain-or-
iented electrical steel if the magnetic flux density is smaller than 1.25 T. Supplementary data to this article can be found online at https://
It is well known that the grain-oriented electrical steel consists of es- doi.org/10.1016/j.jmmm.2019.165597.
sentially only one grain orientation (the Goss) with very large grain
sizes (millimeters to centimetres) [37]. Although it has very small References
texture factor in the rolling direction (which is beneficial to the core
loss), the extremely large grain size will simultaneously deteriorate the [1] B.D. Cullity, C.D. Graham, Introduction to Magnetic Materials, John Wiley & Sons,
core loss, especially when the frequency is high. This is because the Hoboken, New Jersey, USA, 2011.
[2] C.W. Chen, Magnetism and Metallurgy of Soft Magnetic Materials, Dover publica-
eddy current loss will dominate the total loss when the frequency is tions, New York, USA, 1986.
high, and the eddy current loss increases with grain size. A good [3] F. Fiorillo, G. Bertotti, C. Appino, M. Pasquale, Soft magnetic materials, in:
combination of the microstructure (grain size) and the crystallographic M. Peterca (Ed.), Wiley Encyclopedia of Electrical and Electronics Engineering,
Wiley, Hoboken, New Jersey, 2016.
texture of the non-oriented electrical steel produced in this study might [4] A.J. Moses, Electrical steels: past, present and future developments. IEE Proceedings
have resulted in the better core losses than the grain-oriented electrical A (Physical Science, Measurement and Instrumentation, Management and
steel. Education), 1990, 137 (5), 233–245.
[5] A.J. Moses, Energy efficient electrical steels: magnetic performance prediction and
optimization, Scrip. Mater 67 (6) (2012) 560–565.
5. Summary and conclusions [6] R.K. Ray, J.J. Jonas, Transformation textures in steels, Int. Mater. Rev. 35 (1)
(1990) 1–36.
[7] L. Kestens, S. Jacobs, Texture control during the manufacturing of non-oriented
In this study, a non-oriented electrical steel with 3.2 wt% of silicon
electrical steels, Texture, Stress, and Microstructure, Article ID 173083, 2008, 1–9.
was processed using conventional rolling and annealing routes, and the

10
M. Mehdi, et al. Journal of Magnetism and Magnetic Materials 491 (2019) 165597

[8] D. Vanderschueren, L. Kestens, P. van Houtte, E. Aernoudt, J. Dilewijns, U. Meers, [33] D. Raabe, Overview on basic types of hot rolling textures of steels, Steel Res. Int. 74
The effect of cross rolling on texture and magnetic properties of non oriented (5) (2003) 327–337.
electrical steels, Textur. Microstruct. 14–18 (1991) 921–926. [34] R.K. Ray, J.J. Jonas, R.E. Hook, Cold rolling and annealing textures in low carbon
[9] F. Kovac, M. Dzubinský, Y. Sidor, Columnar grain growth in non-oriented electrical and extra low carbon steels, Int. Mater. Rev. 39 (4) (1994) 129–172.
steels, J. Magnet. Magnet. Mater. 269 (3) (2004) 333–340. [35] M. Hölscher, D. Raabe, K. Lücke, Rolling and recrystallization textures of bcc steels,
[10] L. Xie, P. Yang, N. Zhang, C. Zong, D. Xia, W. Mao, Formation of 1 0 0 textured Steel Res. 62 (12) (1991) 567–575.
columnar grain structure in a non-oriented electrical steel by phase transformation, [36] J.J. Sidor, K. Verbeken, E. Gomes, J. Schneider, P.R. Calvillo, L.A. Kestens, Through
J. Magnet. Magnet. Mater. 356 (2014) 1–4. process texture evolution and magnetic properties of high Si non-oriented electrical
[11] T. Tomida, N. Sano, K. Ueda, K. Fujiwara, N. Takahashi, Cube-textured Si-steel steels, Mater. Characteriz. 71 (2012) 49–57.
sheets by oxide-separator-induced decarburization and growth mechanism of cube [37] F.J. Humphreys, M. Hatherly, Recrystallization and Related Annealing Phenomena,
grains, J. Magnet. Magnet. Mater. 254–255 (2003) 315–317. Elsevier, 2012.
[12] M. Džubinský, Y. Sidor, F. Kováč, Kinetics of columnar abnormal grain growth in [38] P. Gobernado, R.H. Petrov, L.A.I. Kestens, Recrystallized {311}〈136〉 orientation in
low-Si non-oriented electrical steel, Mater. Sci. Eng. A 385 (1–2) (2004) 449–454. ferrite steels, Script. Mater. 66 (9) (2012) 623–626.
[13] T. Nguyen Minh, J.J. Sidor, R.H. Petrov, L. Kestens, Texture evolution during [39] A. Samet-Meziou, A.L. Helbert-Etter, T. Baudin, Comparison between re-
asymmetrical warm rolling and subsequent annealing of electrical steel, Mater. Sci. crystallization mechanisms in copper and Ti-IF steel after a low amount of de-
For. 702–703 (2011) 758–761. formation, Mater. Sci. Eng.: A 528 (10–11) (2011) 3829–3832.
[14] N. Yoshinaga, L. Kestens, B.C. De Cooman, α→ γ→ α transformation texture for- [40] W.B. Hutchinson, Deformation substructures and recrystallisation, Mater. Sci.
mation at cold-rolled ultra low carbon steel surfaces, Materials Science Forum, Forum. 558 (2007) 13–22.
Trans Tech Publications, 2005, pp. 1267–1272. [41] K.M. Lee, M.Y. Huh, H.J. Lee, J.T. Park, J.S. Kim, E.J. Shin, O. Engler, Effect of hot
[15] T. Tomida, A new process to develop (100) texture in silicon steel sheets, J. Mater. band grain size on development of textures and magnetic properties in 2.0% Si non-
Eng. Perform. 5 (3) (1996) 316–322. oriented electrical steel sheet, J. Magnet. Magnet. Mater. 396 (2015) 53–64.
[16] Y. He, E. Hilinski, J. Li, Texture evolution of a non-oriented electrical steel cold [42] S.S.F. de Dafe, S. da Costa Paolinelli, A.B. Cota, Influence of thermomechanical
rolled at directions different from the hot rolling direction, Metallurg. Mater. Trans. processing on shear bands formation and magnetic properties of a 3% Si non-or-
A 46 (11) (2015) 5350–5365. iented electrical steel, J. Magnet. Magnet. Mater. 323 (24) (2011) 3234–3238.
[17] M. Mehdi, Y. He, E.J. Hilinski, A. Edrisy, Effect of skin pass rolling reduction rate on [43] J.T. Park, J.A. Szpunar, Evolution of recrystallization texture in nonoriented elec-
the texture evolution of a non-oriented electrical steel after inclined cold rolling, J. trical steels, Acta Mater. 51 (11) (2003) 3037–3051.
Magnet. Magnet. Mater. 429 (2017) 148–160. [44] D. Dorner, S. Zaefferer, D. Raabe, Retention of the Goss orientation between mi-
[18] Y. He, E.J. Hilinski, Skew rolling and its effect on the deformation textures of non- crobands during cold rolling of an Fe3%Si single crystal, Acta mater. 55 (7) (2007)
oriented electrical steels, J. Mater. Process. Technol. 242 (2017) 182–195. 2519–2530.
[19] H. Barkhausen, Gerausche bien ummagnetisieren von eisen, Physikalishe Zeitschr. [45] T. Nguyen-Minh, J.J. Sidor, R.H. Petrov, L.A.I. Kestens, Occurrence of shear bands
20 (1919) 401–403. in rotated Goss ({110}〈110〉) orientations of metals with bcc crystal structure,
[20] S. Tiitto, Influence of elastic and plastic strain on the magnetization process in Fe- Script. Mater. 67 (12) (2012) 935–938.
3.5%Si, IEEE Trans. Magnet. MAG-12 (1976) 855–857. [46] I.L. Dillamore, J.G. Roberts, A.C. Bush, Occurrence of shear bands in heavily rolled
[21] Y. He, M. Mehdi, E.J. Hilinski, A. Edrisy, Through-process characterization of local cubic metals, Metal Sci. 13 (2) (1979) 73–77.
anisotropy of Non-oriented electrical steel using magnetic Barkhausen noise, J. [47] M. Sanjari, M. Mehdi, Y. He, E.J. Hilinski, S. Yue, L.A. Kestens, A. Edrisy, Tracking
Magnet. Magnet. Mater. 453 (2018) 149–162. the evolution of annealing textures from individual deformed grains in a cross-
[22] C.G. Stefanita, Chapter 2 Barkhausen Noise as a Magnetic Nondestructive Testing rolled non-oriented electrical steel, Metall. Mater. Trans. A 48 (12) (2017)
Technique, in: R. Hull, J.R.M. Osgood, J. Parisi, J.H. Warlimont (Eds.), From Bulk to 6013–6026.
Nano, The Many Sides of Magnetism, Springer, Berlin, 2008, pp. 19–40. [48] J. Wang, L.I. Jun, X.F. Wang, J.J. Tian, C.H. Zhang, S.G. Zhang, Effect of heating
[23] M. Birsan, J.A. Szpunar, T.W. Krause, D.L. Atherton, Magnetic Barkhausen noise rate on microstructure evolution and magnetic properties of cold rolled non-or-
study of domain wall dynamics in grain oriented 3% Si-Fe, IEEE Trans. Magnet. 32 iented electrical steel, J. Iron Steel Res. Int. 17 (11) (2010) 54–61.
(2) (1996) 527–534. [49] J.B. Goodenough, Summary of losses in magnetic materials, IEEE Trans. Magnet. 38
[24] A.A. Samimi, T.W. Krause, L. Clapham, M. Gallaugher, Y. Ding, P. Ghosh, (5) (2002) 3398–3408.
A.M. Knight, Correlation between ac core loss and surface magnetic Barkhausen [50] C.P. Steinmetz, On the law of hysteresis, Proceed. IEEE 72 (2) (1984) 197–221.
noise in electric motor steel, J. Nondestruct. Evaluat. 33 (4) (2014) 663–669. [51] A. Chaudhury, R. Khatirkar, N.N. Viswanathan, V. Singal, A. Ingle, S. Joshi,
[25] K. Hartmann, A.J. Moses, T. Meydan, A system for measurement of AC Barkhausen I. Samajdar, Low silicon non-grain-oriented electrical steel: Linking magnetic
noise in electrical steels, J. Magnet. Magnet. Mater. 254 (2003) 318–320. properties with metallurgical factors, J. Magnet. Magnet. Mater. 313 (1) (2007)
[26] ASTM E2627-13, Standard Practice for Determining Average Grain Size Using 21–28.
Electron Backscatter Diffraction (EBSD) in Fully Recrystallized Polycrystalline [52] C.K. Hou, Effect of silicon on the loss separation and permeability of laminated
Materials, ASTM International, West Conshohocken, PA, 2013. steels, J. Magnet. Magnet. Mater. 162 (2–3) (1996) 280–290.
[27] ASTM A343/A343M-14, Standard Test Method for Alternating-Current Magnetic [53] G. Bertotti, G. Di Schino, A.F. Milone, F. Fiorillo, On the effect of grain size on
Properties of Materials at Power Frequencies Using Wattmeter-Ammeter-Voltmeter magnetic losses of 3% non-oriented SiFe, Le J. de Physi. Colloques 46 (C6) (1985)
Method and 25-cm Epstein Test Frame, ASTM International, West Conshohocken, C6–385.
PA, 2014. [54] J. Qin, P. Yang, W. Mao, F. Ye, Effect of texture and grain size on the magnetic flux
[28] KJS Associates/Magnetic Instrumentation Inc. Technical reference manual: com- density and core loss of cold-rolled high silicon steel sheets, J. Magnet. Magnet.
puter-automated magnetic hysteresigraph for testing of magnetically soft materials, Mater. 393 (2015) 537–543.
model SMT-700, 2015, 1–74. [55] N. Leuning, S. Steentjes, K. Hameyer, Effect of grain size and magnetic texture on
[29] Stresstech Group (2013). Rollscan 350 Operating Instructions, V1.0b. 1–39. iron-loss components in NO electrical steel at different frequencies, J. Magnet.
[30] Stresstech Group (2015). MicroScan 600 Operating Instructions, V.5.4b. 1–37. Magnet. Mater. 469 (2019) 373–382.
[31] Y. He, M. Mehdi, E.J. Hilinski, A. Edrisy, S. Mukundan, A. Mollaeian, N.C. Kar, [56] E. Stephenson, A. Marder, The effects of grain size on the core loss and permeability
Evaluation of local anisotropy of magnetic response from non-oriented electrical of motor lamination steel, IEEE Transact. Magnet. 22 (2) (1986) 101–106.
steel by magnetic barkhausen noise, IEEE Trans Magnet. 99 (2018) 1–5. [57] M. Shiozaki, Y. Kurosaki, The effects of grain size on the magnetic properties of
[32] Deveci, M. Stresstech Bulletin 2: The Properties of Barkhausen Noise. https://www. nonoriented electrical steel sheets, J. Mater. Eng. 11 (1) (1989) 37–43.
stresstech.com/download_file/view_inline/381/.

11

You might also like