You are on page 1of 15

Engineering Failure Analysis 134 (2022) 106080

Contents lists available at ScienceDirect

Engineering Failure Analysis


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

Microstructure and tribological performance of boride layers on


ductile cast iron under dry sliding conditions
Dongya Zhang a, b, *, Yue Li a, Xin Du a, Hongwei Fan b, Feng Gao a
a
Key Laboratory of Manufacturing Equipment of the Shaanxi Province, School of Mechanical and Precision Instrument Engineering, Xi’an University
of Technology, Xi’an, China
b
Shaanxi Key Laboratory of Mine Electromechanical Equipment Intelligent Monitoring, Xi’an University of Science and Technology, Xi’an, China

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

Keywords: Stamping die is prone to wear failure due to mechanical friction. In this work, high-hardness
Ductile cast iron boride layers were obtained on ductile cast iron as stamping die, by a boronizing process at
Boride layer 800–950 ℃ for 2–6 h with the addition of rare-earth oxide La2O3. A sawtooth-like morphology of
Hardness
the boride layer was observed using scanning electron microscopy (SEM). The greatest boride
Rare-earth
Wear mechanism
layer micro-hardness of the boride layers of 1648 HV0.1 was obtained at 900 ℃ and 6 h with 4%
La2O3, and X-ray diffraction (XRD) analysis showed that the boride layers were a component of
Fe2B as adding La2O3. The tribological tests showed that the wear resistance of the boride layers
was enhanced with increasing boronizing temperature and time, where the ductile cast iron
substrate corresponding to 900 ℃ and 6 h exhibited the best wear resistance. Compared with that
of the boride layer without La2O3, the thickness and wear resistance of the boride layer with 4%
La2O3 was increased by 37.2% and 31%, respectively. The thickness of the boride with 6% La2O3
decreased by 27.6%, and the wear rate was highest, indicating that excessive La2O3 inhibited the
enhancement effect. The wear mechanism of the boride layers transformed from mild adhesive
wear to fatigue spalling as the normal load and sliding speed increased.

1. Introduction

The service life of the mold plays an important role in the surface quality and production efficiency of parts [1]. If the mold is
frequently replaced due to the failure of a process, the normal production process will be seriously disturbed, and even cause the
fluctuation of product quality. Therefore, it is necessary to prolong the service life of the mold. Wear is the main form of die failure and
an important factor affecting the service life. Owing to excellent machinability, good vibration-damping behavior and good me­
chanical properties, ductile cast iron has been gradually used in stamping die [2]. Friction and wear occur in the presence of relative
movement between moving parts, and the graphite phase in ductile cast iron forms a lubricating transfer film on the friction surface,
and can reduce the friction and wear of the contact surface [3]. Nevertheless, the wear of ductile cast iron becomes increasingly severe
under harsh working conditions such as high speeds, high strengths and heavy loads. Hence, seeking an advanced surface modification
technique is imperative to strengthen the wear resistance. In recent years, to enhance the wear resistance of ductile cast iron, re­
searchers have developed various surface modification technologies, such as laser surface hardening and high hardness coating,

* Corresponding author at: Key Laboratory of Manufacturing Equipment of the Shaanxi Province, School of Mechanical and Precision Instrument
Engineering, Xi’an University of Technology, Xi’an, China.
E-mail address: dyzhang@xaut.edu.cn (D. Zhang).

https://doi.org/10.1016/j.engfailanal.2022.106080
Received 10 November 2021; Received in revised form 28 December 2021; Accepted 18 January 2022
Available online 21 January 2022
1350-6307/© 2022 Elsevier Ltd. All rights reserved.
D. Zhang et al. Engineering Failure Analysis 134 (2022) 106080

including TiN [4] and diamond-like carbon (DLC) [5] coatings.


Surface boronizing is a thermochemical treatment process in which boron atoms diffuse into iron, titanium or other alloy substrates
to form a high-hardness boride layer [6–8]. Through pack boronizing treatment of an iron-based alloy, ferroboron compounds can be
formed on the substrate [9]. Owing to their good wear-resistant surface, boride layers can improve corrosion resistance [10] and high-
temperature oxidation resistance [11]. Kul [12] investigated the morphology of the boride layer on ductile cast iron boronized at
950 ◦ C for 4 h. The thickness and hardness of the boride layer depended on the substrate material, boronizing temperature–time and
boronizing agent.
Mustafa [13] fabricated boride layer on Mustafa AISI H13 hot-work tool steel using powder pack-boriding method, and also
observed that the thickness of the boride layer on ductile iron increased depending on the boronizing time and temperatures, and the
wear volume decreased with the increase of boride layer thickness. Yusuf [14] investigated the adhesion and wear properties of boro-
tempered ductile iron, and wear tests indicated that boro-tempering heat treatment increased the wear resistance of ductile iron. Omar
[15] enhanced the EN-GJS-400–15 nodular cast iron hardness by pack-borided treatment, and the results showed that thickness of the
boride layers obtained at 1000 ◦ C and 6 h was 2.5 times higher than that of obtained at 900 ◦ C and 2 h. Sen [16] fabricated a boride
layer on ductile iron at 850 and 950 ◦ C for 2–8 h. The results showed that borides of FeB and Fe2B were formed, and the hardness of the
boride layers were in the range of 1160–2140 HV. However, the fracture toughness of the boride layer declined with increasing
boronizing temperature. The expansion coefficients of FeB and FeB2 are obviously different, a harmful residual stress is generated, and
cracks initiate under high temperature, and the boride layer easily peels off. Although FeB can enhance the hardness, the boride layers
are prone to fatigue spalling during use due to the low toughness of Fe2B [17], which has a negative effect on the wear resistance,
therefore, a single-phase Fe2B for the boride layer is preferred.
Recently, rare earth elements have been introduced into the boronizing process to alleviate the brittleness and spalling propensity
of the boride layers [18]. The diffusion activation energy of boron atoms in the matrix can be significantly reduced, and the inward
diffusion rate can be accelerated by rare earth infiltration. This is an effective method to prepare single-phase Fe2B on a steel substrate.
Yuan [19] fabricated Cr-rare earth-boronized layers on 45 steel at 650 ◦ C for 6 h, and the boride layers were denser and more uniform
while the wear resistance was significantly enhanced. According to the abovementioned research, the rare-earth boride layer showed
superior wear resistance because rare earth elements could increase the thickness and hardness of the boride layer. However, there is
an optimal content range for the addition of rare earth elements, and the performance of the boride layer is degraded if the rare earth
content exceeds the optimal content. In addition, the optimal rare earth content varies for steels with different carbon contents.
Due to the high yield point and high hardness of stainless steel, and serious adhesion phenomenon will occur in the forming process
[20], which makes the die wear more severe, and bringing a great challenge to the service life of the stamping die. In this paper, the
effect of boronizing modification on wear of nodular cast iron was studied by using nodular cast iron as stamping die, and the effects of
rare earth oxide (La2O3) content, boronizing temperature and boronizing time on the microstructure and composition of the boride
layer on ductile cast iron were investigated. Furthermore, the tribological properties were analyzed under dry test conditions, and the
possible wear mechanism was discussed.

2. Materials and methods

2.1. Fabrication of the boride layer

Ductile cast iron (QT450-10, micro-hardness 265 HV0.49, E = 160 GPa. v = 0.27), which was 30 mm in diameter and 5 mm in
thickness, was used as the substrate material. This ductile cast iron has a tensile strength of 450 MPa and yield strength of 310 MPa. The
chemical composition of QT450-10 ductile iron is shown in Table 1. Prior to the boronizing process, the specimens were polished
sequentially with 600, 800 and 1200 grit metallographic sandpapers, and then cleaned in an ultrasonic cleaner with absolute ethanol
for 10 min.
The commercial boronizing agent was mainly composed of B4C (as the boron source), KBF4 (as the activator), SiC (as the transport
medium), and its compositional contents are listed in Table 2. La2O3 (as catalyst) contents of 0, 2, 4 and 6 wt% were added to the
boronizing agent. La2O3 has been proven to increase the boride nucleation rate during the formation of boride layers and enhance the
diffusion of boron atoms, thereby reducing surface defects [21]. The specimens and the boronizing agent were put into a sealed
container, and heated to 200 ℃ in a vacuum furnace for 60 min. Then, the temperature increased to the boronizing temperature, which
was held for the preset boronizing time. Eventually, the specimens were cooled to room temperature in the furnace. The process
parameters were as follows: (i) When the boronizing time was 4 h, the boronizing temperatures were set as 800, 850 and 900 ℃, and
the corresponding specimen codes were Q-800–4, Q-850–4 and Q-900–4. (ii) When the boronizing temperature was 900 ℃, the
boronizing times were 2 h, 4 h and 6 h, and the corresponding specimen codes were Q-900–2, Q-900–4 and Q-900–6. After boronizing,
the boronizing agent on the specimen surfaces was polished with 1200 grit sandpaper, and then the specimens were ultrasonically
cleaned with absolute ethanol for 10 min and dried in a vacuum oven. Additionally, a non-boronized specimen was marked as Q-0.

Table 1
Chemical composition of ductile cast iron.
Elements C Si Mn P S Cu Mg Fe

Content (wt.%) 3.58 2.65 0.22 0.04 0.03 0.01 0.043 Bal.

2
D. Zhang et al. Engineering Failure Analysis 134 (2022) 106080

Table 2
Compositional contents of the boronizing agent.
Composition B4C KBF4 SiC La2O3

Content (wt.%) 10% 10% 80% 0, 2%, 4%, 6%

In this paper, the boronizing time and boronizing temperature were evaluated under the boronizing agent without La2O3. The
optimal La2O3 content on the performance of the boride layer was evaluated at the optimum boronizing process parameters.

2.2. Boride layer properties

An electron microscope (JSM-6700F) was utilized to observe the morphology and cross-sectional microstructure of the boride
layers and to measure the boride layer thickness. Additionally, the elemental distribution on the boride layer surfaces was analyzed by
energy dispersive spectroscopy (EDS). In addition, a PHILIPS X’ pert MPD RRO type X-ray diffraction (XRD-7000) apparatus was used
to determine the phase compositions on top surface of the boride layers, and the scan range was from 20◦ to 90◦ at a scanning rate of
0.4◦ /s.
The cross-section micro-hardness of the boride layer were measured by using a micro-Vickers hardness tester (TMVS-1), with a load
of 100 g for a holding time of 15 s, and five points were tested at different locations on the surface to determine the mean value.

2.3. Tribological tests

The tribological properties of the boride layers were conducted in a UMT-Tribolab tribometer with a reciprocating model under dry
friction conditions, where the upper specimen was a SUS 304 steel ball (a diameter of 10 mm, hardness of 280 HV0.49, E = 193 GPa. v =
0.3), and the lower specimen was a boride layer and a control specimen with a diameter of 30 mm and thickness of 5 mm. The
tribological test parameters were as follows (shown in Table 3): applied loads of 5 N and 10 N (the maximum contact pressure was 0.75
GPa and 0.95 GPa when the SUS 304 steel ball is in contact with ductile cast iron according the Hertz contact theory), reciprocating
stroke of 5 mm, sliding frequency of 2 Hz and 6 Hz, and a total sliding distance of 216 m. Prior to and after each test, the ball and disc
specimens were ultrasonically cleaned with absolute ethanol then acetone each for 10 min. The wear track profiles of the disc
specimens were measured with a surface profile-meter, and the computational formula for the wear rate k is:
V
k= (1)
F×S
where V denotes the wear volume (mm3), F denotes the applied load (N), and S is the sliding distance (m).
The microstructure and composition of worn surfaces were analyzed by SEM and EDS, and the effects of the test conditions on the
wear mechanism of the boride layers were explored.

3 Results and discussions

3.1. Characterization of the boride layers

Fig. 1 presents the SEM micrographs of the substrate and boride layers. As shown in Fig. 1a, a substantial distribution of spheroidal
graphite was observed on the surface of non-borided ductile cast iron, which could improve the lubricating performance during the
friction process [22]. Fig. 1b-e show the boride layers obtained under different boronizing parameters. As is clear, the boride layers had
some pores, whose sizes increased markedly with increasing boronizing temperature. The graphite was moved into the substrate by B
atoms during the boronizing process, thereby resulting in significantly elevated B elemental content and significantly decreased C
element content in the boride layers (as shown in Fig. 2f) [23]. Furthermore, the migration of C atoms promoted the flow of vacancies
toward the subsurface based on the diffusion mechanism, thereby leading to the pore formation on the surface [24].
The cross-sectional morphologies of the boride layers are illustrated in Fig. 2. Under the same boronizing time, the thickness of the
boride layers increased with increasing boronizing temperature, exhibiting increases from 13.3 μm to 64.8 μm as the temperature rose
from 800 ℃ to 900 ℃. Clearly, increasing the boronizing temperature could markedly increase the boride layer thickness. The reason
was that the mobility of B atoms was enhanced with the temperature increase, which accelerated their bonding to the Fe atoms and
facilitated their diffusion into the substrate, consequently leading to gradually increased thickness of the boride layers [25]. When the
boronizing temperature was 900 ℃ and the boronizing time was prolonged from 2 h to 6 h, the thickness of the boride layers increased

Table 3
Experimental parameters.
Load (N) Sliding frequency (Hz) Test time (h)

Test 1 5 2 3
Test 2 10 2 3
Test 3 10 6 1

3
D. Zhang et al. Engineering Failure Analysis 134 (2022) 106080

Fig. 1. SEM micrographs of the (a) substrate, and boride layers of (b) Q-800-4, (c) Q-850-4, (d) Q-900-2, (e) Q-900-4 and (f) Q-900-6.

Fig. 2. Cross-sectional morphologies of the boride layers: (a) Q-800-4, (b) Q-850-4, (c) Q-900-2, (d) Q-900-4 and (e) Q-900-6.

from 40 μm to 67.8 μm. The thickness of boride layers increased with prolonged diffusion time. Additionally, a distinct sawtooth-like
morphology was found at the junction between boride layers and the substrates. This morphology became increasingly evident with a
prolongation of the boronizing time. The sawtooth-like morphology was conducive to enhancing the adherence between the substrate
and the boride layers, which is a typical feature of preferred orientation growth for borides [26].
Fig. 3 shows the cross sectional morphologies of the boride layers obtained at different La2O3 contents, in which the formation of
boride layers consisting of three regions was clearly observed: (i) an outer layer including some cracks, (ii) a uniform dense inner layer
with no cracks was below the outer layer, (iii) a typical sawtooth-like morphology between the boride layer and substrate was also

4
D. Zhang et al. Engineering Failure Analysis 134 (2022) 106080

Fig. 3. Cross-sectional morphologies of the boride layers of Q-900-6 with La2O3 content:(a) 0%, (b) 2%, (c) 4%, and (d) 6%.

clearly observed, which were consistent with a previous work [27]. The thickness of the boride layers could also be obtained from the
SEM images, and the thickness of the boride layer without La2O3 was 64.8 μm. With the increase of La2O3 content, the thickness of the
boride layer increased first and then decreased. For the content La2O3 of 4%, the thickness reached the maximum of 74.6 μm, which
increased by 15.1% compared with that of the boride layer without adding La2O3. It indicates that La2O3 could promote the formation
of borides at this content. However, when the La2O3 content increased to 6%, the thickness of the boride layer decreased significantly,
its thickness was only 54.3 μm, which decreased by 16.2% compared with that of layer without adding La2O3, indicating that a higher

2000
Q-800-4
1800
Q-850-4
1600 Q-900-2
Q-900-4
1400
Micro-hardness (Hv)

Q-900-6
1200
1000
800
600
400
200
0
0 10 20 30 40 50 60 70 80 90 100 110 120 130
Depth of the layer (μm)

Fig. 4. Gradient hardness of the boride layers obtained at different boronizing temperatures and time.

5
D. Zhang et al. Engineering Failure Analysis 134 (2022) 106080

rare earth content impeded the formation of the boride layer.


Fig. 4 depicts the gradient hardness of the boride layer cross sections at various temperatures. The hardness values gradually
decreased as the depth of the boride layers increased. Under a boronizing time of 4 h, the maximum hardness of the boride layers
increased from 485 HV0.1 (800 ℃) to 1472 HV0.1 (900 ℃). By the action of high temperature, B atoms combined with Fe atoms to
enhance the hardness of the boride layers. At a boronizing temperature of 900 ℃, the maximum hardness increased from 1400 HV0.1 to
1648 HV0.1 as the boronizing time extended from 2 h to 6 h. With the extension of boronizing time, the amount of boride compounds
increased, consequently enhancing the hardness of the boride layers [28]. According to the foregoing analysis results, the average
hardness values of the boride layers were all higher than the hardness of the substrate, which is conducive to improving the wear
resistance.
Fig. 5 depicts the hardness of the boride layers obtained at different La2O3 contents. The hardness of the boride layer was in range of
950–1628 HV after boronizing process, which is significantly higher than that of the substrate. With increasing La2O3 content to 4%,
the hardness was higher than that without La2O3. The reason was that the rare-earth elements can make the distribution of iron-boron
component more uniform, and the sawtooth-like structure be dense, and as a result the hardness were enhanced [29]. When the
content was 6%, the surface hardness was 950 HV, which is 30% lower than that of the boride layer without La2O3. It indicates that the
rare earth can increase the diffusion rate of active boron atoms, which improved the hardness of boride layer. However, by further
increasing the content of rare earth in boronizing agent, the promoting effect gradually changed to inhibition, and the hardness of
boride layer was gradually decreased [30].
Fig. 6 shows the X-ray diffraction spectrum results of the boride layers obtained by the boronizing agent with and without rare earth
La2O3 at 900 ℃ for 6 h. The boronizing layer without the addition of rare earth elements was mainly composed of Fe2B and FeB phases
(as shown in Fig. 6a), and the diffraction peak of the Fe3(C, B) phase also appeared. During the boronizing process, the B atoms initially
replaced the C atoms in Fe3C to form Fe3(C, B), and Fe2B and FeB formed as the reaction continued [31]. Fe3(C, B) and cementite are
orthogonal crystal systems, and the radii of the boron and carbon atoms are similar, so the carbon atom in Fe3C can be replaced by a
boron atom to form boron-containing cementite Fe3(C, B).
For the boronizing layer with 2% La2O3 (as shown in Fig. 6b), the boride layer was composed of single-phase Fe2B without the FeB
phase. This finding indicates that the rare earth content can increase the diffusion coefficient of B atoms, and improve the diffusion
velocity of B atoms, and the growth rate of Fe2B was the fastest along the [2 0 0] direction. The addition of La2O3 promoted the
preferred orientation of Fe2B and made the crystal orientation of the Fe2B phase more obvious. However, the peak value of Fe2B was
the weakest as the La2O3 content increased to 6%, which indicates that the rare earth atoms accumulated on the surface and could not
quickly diffuse into the matrix when the rare earth content was too high. Furthermore, the diffusion of boron atoms was impeded and
prevented the growth of the boride layer. This phenomenon was consistent with the morphology of the boride layer with different
thicknesses, as seen in Fig. 3.

3.2. Tribological performance

3.2.1. Effect of boronizing temperature


Fig. 7a details the friction coefficients of the boride layers under a load of 5 N and a sliding frequency of 2 Hz. For Q-0, the friction
coefficient fluctuated gently at approximately 0.18, which is low and stable. The graphite embedded in ductile cast iron could be
transferred to the subsurface and formed a self-lubricating film, which inhibits the direct contact between the mating surfaces and
reduces the friction coefficient [32]. The friction coefficient of Q-800–4 gradually increased from 0.15 to 0.32 in the initial 5400 s, and
then rapidly increased to 0.43 accompanied by drastic fluctuations. The friction coefficient of Q-850–4 gradually increased to 0.37 in
the initial 2260 s and remained stable. The friction coefficient of Q-900–4 rapidly increased from 0.15 to 0.42 over 480 s and then
reached stability. For Q-800–4, Q-850–4 and Q-900–4, (i) the average friction coefficients were 0.34, 0.37 and 0.42, respectively,

2000
1800 1628
1530
1600 1360
Surface hardness (Hv)

1400
1200 950
1000
800
600
400
200
0
0 2 4 6
La2O3 content (%)

Fig. 5. Microhardness of the top surface of the boride layers obtained with different La2O3 contents.

6
D. Zhang et al. Engineering Failure Analysis 134 (2022) 106080

Fig. 6. XRD pattern of the boride layers obtained (a) without and (b) with La2O3 at 900 ℃ for 6 h.

0.8
(a) Load: 5 N, frequency: 2 Hz Q-0
Q-800-4
0.7
Q-850-4
Q-900-4
0.6
Friction coefficient

0.5

0.4

0.3

0.2

0.1

0.0
0 1800 3600 5400 7200 9000 10800
Sliding time(s)

0.8 0.8
Q-0 (c) Load:10 N, frequency: 6 Hz Q-0
(b) Load:10 N, frequency:2 Hz
Q-800-4 Q-800-4
0.7 0.7
Q-850-4 Q-850-4
Q-900-4 0.6 Q-900-4
0.6
Friction coefficient
Friction coefficient

0.5 0.5

0.4 0.4

0.3 0.3

0.2 0.2

0.1 0.1

0.0 0.0
0 1800 3600 5400 7200 9000 10800 0 600 1200 1800 2400 3000 3600
Sliding time (s) Sliding time(s)

Fig. 7. Friction coefficients of Q-0, Q-800-4, Q850-4 and Q-900-4 under different test conditions.

indicating an increase in the friction coefficient with increasing boronizing temperature. (ii) The running-in times were 5400 s, 2260 s
and 480 s, suggesting that the time required for the boride layers to reach frictional stability decreased with increasing boronizing
temperature.
An interesting phenomenon was found that the friction coefficient of the Q-800–4 slowly raised for a very long time and then
increased to a certain value. The maximum cross section hardness of Q-800–4 was 500 Hv0.1, which was much lower than that of

7
D. Zhang et al. Engineering Failure Analysis 134 (2022) 106080

boride layers obtained at other conditions. The micro-hardens curve (as shown Fig. 4) speculates that boron iron compound and
spheroidal graphite both existed on the top surface of the Q-800–4. Thus, the friction coefficient of the Q-800–4 was slowly increased to
0.2 under the lubrication effect of spheroidal graphite, which had a similar trend with Q-0. However, the spheroidal graphite was worn
off with the sliding time prolonged to 900 s. By this time, the content of boron iron compound with high hardness in the subsurface
layer was increased, and began participates in friction. Thus, the friction coefficient quickly reached to as high as that of other boride
layers as sliding continued.
Fig. 8b presents the friction coefficients under 10 N and 2 Hz condition. The friction coefficient of Q-0 was 0.10 at the initial stage,
which then slowly increased, reaching a maximum of 0.21 at 1250 s. Subsequently, the friction coefficient gradually decreased, and
stabilized at approximately 0.17 after 2300 s. The friction coefficient of Q-800–4 increased rapidly to 0.36 at 2300 s from the initial
0.10 and then slowly increased to 0.39. Q-850–4 and Q-900–4 exhibited identical variation trends in terms of their friction coefficient
stability was achieved after 570 s, and the corresponding stable-stage friction coefficients were 0.38 and 0.40. Compared to the results
under a load of 5 N, the average friction coefficients under 10 N were 0.34, 0.37 and 0.39, respectively, and the running-in times were
1250 s, 570 s and 570 s. This result suggests that the friction coefficient of the boride layers is more stable under higher applied load,
and the time required for running-in is also shorter.
When the test condition intensified to 10 N and 6 Hz (Fig. 8c), the friction coefficient of Q-0 fluctuated gently between 0.21 and
0.25 throughout the running process. Hence, ductile cast iron maintained a relatively low and stable friction coefficient with the
intensification of the test conditions. Nevertheless, the friction coefficient of Q-800–4 gradually increased to reach a maximum of 0.64
after sliding for 1360 s. Subsequently, the friction coefficient slowly decreased and reached a stable stage at 1900 s, showing a value of
0.51 in the stable stage. For Q-850–4, the friction coefficient increased to a maximum of 0.61 from the initial 0.10 after 1100 s of
running-in, and then slowly decreased and reached a normal wear phase. The friction coefficient of Q-900–4 rapidly increased to 0.63
in the initial 420 s and decreased to 0.46 at 490 s and obtained a steady state. The average friction coefficients of Q-800–4, Q-850–4
and Q-900–4 were 0.44, 0.41 and 0.48, respectively, showing markedly higher values than those under the conditions of 5 N-2 Hz and
10 N-2 Hz.
The tribological behavior of the coatings was affected by the coating thickness, surface roughness and hardness [33]. In the initial
friction stage, the friction coefficients of the boride layers reached a high value in a short time. The asperity on the two contact surfaces
played a pinning effect, and a large static friction resistance was generated at the moment of relative movement [34]. The asperity on
the surface of the steel ball was more prone to fracture under higher load and the wear debris formed and increased the tangential
resistance. When the two contact surfaces reached a balanced state, the friction decreased slowly and tended to be stable. Interestingly,
the surface hardness of Q-900–4 was high, and finer wear debris from the steel ball was generated. The two contact surfaces quickly
reached the dynamic balance state, inducing a short running-in period.
Fig. 8 describes the wear rates of the boride layers fabricated at different boronizing temperatures. Under test conditions of 5 N-2 Hz
and 10 N-2 Hz, the wear rates of the boride layers gradually decreased with increasing boronizing temperature. The highest wear rate
was observed with Q-800–4, whereas the lowest wear rate was observed with Q-900–4. However, under the 10 N-6 Hz condition, the
wear rates of the boride layers initially increased and then decreased with increasing boronizing temperature. Q-850–4 exhibited the
highest wear rate (10.19 × 10− 6 mm3 /Nm), which was 2.29 times and 1.86 times that of Q-800–4 and Q-900–4, respectively, and was
64.28% that of ductile cast iron. The boride layer hardness increased from 468 HV to 1200 HV within a boronizing temperature range
of 800–850℃, showing insignificant increases, and the corresponding improvement in wear resistance was indistinct. At 900 ℃, the
boride layer hardness exceeded 1400 HV0.1, and the wear resistance was remarkably improved due to the high hardness of the surface.

3.2.2. Effect of boronizing time


Fig. 9a-b present the friction coefficients of the boride layers (Q-900–2, Q-900–4 and Q-900–6) under loads of 5 N and 10 N and at a
sliding frequency of 2 Hz. As is clear, the friction coefficient curve trends of the three boride layers were similar. With the prolongation

35
5N 2Hz
28.01 10N, 2Hz
30 10N, 6Hz
Wear rate×10-6(mm3/ N·m)

25

17.74
20
14.26
15 12.28 10.8
10.98
10.32
10 8.15 8.24
6.55 6.72
5.00
5

0
0 800 850 900
Boronizing temperatre (oC)

Fig. 8. Wear rates of the boride layers obtained under different boronizing temperatures for 4 h.

8
D. Zhang et al. Engineering Failure Analysis 134 (2022) 106080

of sliding time, the friction coefficients gradually increased and then stabilized. Moreover, the times required for these boride layers to
reach a stable stage were similar as well. Under 5 N, the average friction coefficients were 0.45, 0.42 and 0.40. When the load increased
to 10 N, the average friction coefficients were 0.41, 0.37 and 0.3. It is clear that the average friction coefficients of the boride layers
tended to decrease with increasing boronizing time and test load.
When the applied load increased to 10 N and the sliding frequency increased to 6 Hz (as shown in Fig. 9c), the friction coefficient
curves of Q-900–2, Q-900–4 and Q-900–6 showed significant differences. The friction coefficient of Q-900–2 reached a maximum of
0.60 at 210 s and then gradually decreased and reached a stable value of 0.49 after sliding for 240 s, accompanied by drastic fluc­
tuations. The friction coefficient of Q-900–4 reached a maximum of 0.62 at 390 s, and then rapidly decreased to 0.49. The friction
coefficient of Q-900–6 increased to a maximum of 0.63 in the initial 250 s and then entered a stable friction stage. For Q-900–2 and Q-
900–4, friction coefficients with large fluctuations were observed in the normal wear stage. The hardness and peeling off resistance of
Q-900–2 and Q-900–4 were weak, and were obtained under short boronizing time. Such boride layers gradually cracked and were
abrasively worn under higher contact pressure. In addition, the temperature increase at the contact area accelerated the softening of
the steel ball with an increase in the sliding frequency, which elevated the proportion of adhesive wear to drastically alter the friction
coefficient. However, Q-900–6 had a smaller average friction coefficient that fluctuated little during the normal wear stage. This result
indicates that the boride layer with a long boronizing time could improve the moving system stability when tested under high-speed
and heavy-load conditions.
Fig. 10 presents the wear rates of the boride layers (Q-900–2, Q-900–4 and Q-900–6). Under the 5 N-2 Hz condition, the wear rates
of Q-900–2 and Q-900–6 were 7.19 × 10-6 mm3/N⋅m and 6.05 × 10-6 mm3/N⋅m, respectively. Compared to Q-0, the wear rate of Q-
900–6 was reduced by 50.75%. When under 10 N-2 Hz, the wear resistances of Q-900–2 and Q-900–6 were 68.1% and 77.8% higher
than that of Q-0. The wear resistance of ductile cast iron increased drastically by boronizing. Such an improvement in the wear
resistance of the boride layers was more remarkable as the boronizing time was prolonged. Nevertheless, the wear rates of the boride
layers initially increased and then decreased with the prolongation of the boronizing time under 10 N-6 Hz. The wear rate of Q-900–4
was 1.32 times that of Q-900–6. With the extension of boronizing time, the thickness and hardness of the boride layers are increased,
which is helpful to improve the wear resistance of the boride layers.

0.8
(a) Load:5 N, Frequency:2 Hz Q-900-2
0.7 Q-900-4
Q-900-6
0.6
Friction coefficient

0.5

0.4

0.3

0.2

0.1

0.0
0 1800 3600 5400 7200 9000 10800
Sliding time (s)
0.8 0.8
(b) Load:10 N, Frequency:2 Hz Q-900-2 (c) Load:10 N, Frequency: 6 Hz
0.7 Q-900-4 0.7
Q-900-6
0.6 0.6
Friction coefficient

Friction coefficient

0.5 0.5

0.4 0.4

0.3 0.3

0.2 0.2
Q-900-2
0.1 0.1 Q-900-4
Q-900-6
0.0 0.0
0 1800 3600 5400 7200 9000 10800 0 600 1200 1800 2400 3000 3600
Sliding time (s) Sliding time (s)

Fig. 9. Friction coefficients of Q-900-2, Q-900-4 and Q-900-6 under different test conditions.

9
D. Zhang et al. Engineering Failure Analysis 134 (2022) 106080

35
5N, 2Hz
30 28.2
10N, 2Hz
10N, 6Hz

Wear rate×10 (mm3/ N·m)


25

20 17.74

-6
15 12.28
10.8
10 8.1
7.19 6.72 7.2
5.66 6.15
5.2
3.94
5

0
0 2 4 6
Boronizing time (h)

Fig. 10. Wear rates of Q-900-2, Q-900-4 and Q-900-6 under different test conditions.

The boride layer is composed of a loose layer, a hardened layer and a transition layer (as shown in Fig. 2). The outermost loose layer
with defects such as looseness and holes, which provided a way for the initiation and propagation of cracks, causing the outermost
layer to the crack and peel during friction. Therefore, wear of loose layer occurs mainly under low load, so the boride layer did not
show good wear resistance. When the load increased appropriately, the hardened layer began to participate in the contact after the
loose layer was worn through. The denser Fe-B structure can delay and prevent plastic deformation and crack propagation. Therefore,
the boride layer shown a better resistance under high condition. However, as the load continued to increase, the friction heat generated
increased sharply, and the expansion coefficients of Fe2B phase (FeB = 23 × 10-6/℃, Fe2B = 7.85 × 10-6/℃) were very different [35],
resulting in thermal stress between the phases, which intensified separation and abrasion of the boride layers.

3.2.2. Effect of La2O3 contents


Fig. 11 shows the friction coefficient and wear rate of the boride layers obtained at different La2O3 contents under test condition of
10 N and 6 Hz after sliding for 60 min. For La2O3 contents of 0%, 2% and 4%, the friction coefficient curves had similar tendencies: they
reached peak values at 560 s, 340 s and 1074 s and then tended to be stable with prolonged test time. The corresponding average
friction coefficients were 0.56, 0.61 and 0.57. It is interesting to note that the friction coefficient of the boride layer obtained at 6%
La2O3 content was obviously lower than that of other La2O3-boride layers, the average friction coefficient was 0.52, and the running-in
time was 1430 s.
The cross-sectional profiles of the wear tracks on the boride layer after the friction and wear test are shown in Fig. 12a. The
maximum wear track depth of the boride layer obtained by 6% La2O3 was 7.34 μm, which is approximately 1.93 times greater than the
boride layer obtained by 4% La2O3. The wear track depths of those boride layers were much lower than their corresponding layer
thicknesses, indicating that the layers were not worn out under 10 N and 6 Hz, and the boride layer had a high wear resistance. The
corresponding wear rate was calculated and compared in Fig. 12 b. As shown, the wear rates of the boride layers obtained with 2% and
4% La2O3 were 6.8 mm3/N⋅m and 7.2 mm3/N⋅m, which were obviously lower than that of the boride layer obtained with 0% La2O3.
The main reasons were as follows: (i) The thickness of the boride layers were increased and the boron teeth were arranged more closely
as adding La2O3, which befitted for the improvement of the wear resistance. (ii) The FeB phase disappeared and Fe2B phase increases in

0.8
Load:10 N, Frequency: 6 Hz
0.7

0.6
Friction coefficient

0.5

0.4

0.3
Q900-6-0% La2O3
0.2 Q900-6-2% La2O3
Q900-6-4% La2O3
0.1 Q900-6-6% La2O3

0.0
0 600 1200 1800 2400 3000 3600
Sliding time (s)

Fig. 11. Effect of the La2O3 content on the friction coefficient of the boride layers under test condition of 10 N and 6 Hz.

10
D. Zhang et al. Engineering Failure Analysis 134 (2022) 106080

(a) 0 0

-2 w=1660 μm -2 w=1450 μm
h=4.8 μm h=4.49 μm
-4 -4

-6 -6

-8 -8

Wear depth (μm)


Q-0% La2O3 Q-2% La2O3
-10 -10
0 500 1000 1500 2000 0 500 1000 1500 2000
0 0
w=1370 μm w=1520 μm
-2 -2 h=7.34 μm
h=3.79 μm
-4 -4

-6 -6

-8 -8
Q-4% La2O3 Q-6% La2O3
-10 -10
0 500 1000 1500 2000 0 500 1000 1500 2000
Length (μm)
35
(b) Load:10 N, Frequency: 6 Hz
30
Wear rate×10-6(mm3/N.m)

25

20

15 13.06
9.86
10 7.2 6.8

0
0 2 4 6
La2O3 content (%)

Fig. 12. (a) Wear track profiles and (b) wear rates of the boride layers fabricated by different La2O3 contents under test condition of 10 N and 6 Hz.

the boride layers as adding La2O3, the internal stress micro-crack caused by the different expansion coefficient of FeB and Fe2B was
reduced. However, the boride layer obtained with 6% La2O3 (with a lower hardness of 950 Hv) had highest wear rate. Generally
speaking, a higher hardness reflected a stronger resistance to plastic deformation and led to high resistance to abrasive and adhesive
wear [36]. As a result, it can be said that the wear resistance of the boride layer could be increased by adding a moderate La2O3 content.

3.3. Worn surface analyses

Fig. 13 presents the SEM morphologies of the worn surfaces of the ductile cast iron and the boride layers after the dry sliding wear
test at 10 N-2 Hz. As shown in Fig. 13a, the worn surface was covered by substantial wide furrows for the Q-0 specimen, and adherence
of wear debris was observed in local regions. The primary wear form was furrowed abrasive wear, which was accompanied by slight
adhesive wear. The furrowed abrasive wear was greatly correlated with the surface hardness [37]. In view of the higher hardness of the
steel ball than the ductile iron substrate, the micro-asperities on the ball were compressed into the substrate under contact stress. At the
initiation of friction, a shearing action was produced by the compressed micro-asperities under frictional force, thereby leading to
formation of furrows on the friction surface and the generation of wear debris. Partial wear debris was retained between the friction
pair, which was gradually crushed and pulverized into wear particles. The wear particles, as third-body particles, scratched the metal
surface along the sliding direction under the frictional force, which left substantial cut marks on the worn surfaces, ultimately forming
the characteristic morphology of abrasive wear.
Fig. 13b-c depict the wear morphologies of Q-800–4 and Q-850–4. Furrows parallel to the sliding direction were present on the
worn surfaces, and the depths were markedly reduced. The boride layers effectively reduced abrasive wear by resisting the intrusion of
wear particles and weakening the cutting action [38]. Nevertheless, the boride layer wear tracks exhibited local spalling features, with
non-uniform distribution of Fe2B valence electron bonds and rather weak bond energy of B-B bonds, which were prone to fracture
under external forces, thereby resulting in a significantly increased brittleness [39]. Furthermore, massive flaky adhesives appeared on
the surfaces, indicating the transfer of steel ball material to the boride layers during the friction process and the formation of a metal
film with a certain thickness. Fig. 13d-f present the wear morphologies of Q-900–2, Q-900–4 and Q-900–6. For a boronizing time of 2 h,
many shallow furrows appeared on Q-900–2, that were also accompanied by local fatigue spalling. The predominant wear mechanism
was slight abrasion combined with local fatigue spalling. After extending the boronizing time to 4 h, the worn surface became
smoother, where only slight spalling was present. When the boronizing time extended to 6 h, no apparent fatigue spalling was found,

11
D. Zhang et al. Engineering Failure Analysis 134 (2022) 106080

Fig. 13. SEM images of wear scars on (a) Q-0, (b) Q-800-4, (c) Q-850-4, (d) Q-900-2, (e) Q-900-4, and (f) Q-900-6 under test condition of 10 N-2 Hz.

and Q-900–6 exhibited good wear resistance. This result suggests that there were fewer defects with the boride layers at 900 ℃, and Q-
900–6 was endowed with a higher hardness and a good resistance to plastic deformation. With the prolongation of boronizing time and
the elevation of boronizing temperature, the boride layers had a higher thickness and micro-hardness, and the layer tended to be dense
and uniform. Hence, the plastic deformation was mild under the effects of contact stress, and their peel off areas gradually decreased,
the wear resistances were improved.
To further reveal the wear mechanism of the boride layers, the wear forms of Q-900–6 under different test conditions were
elucidated based on SEM and EDS results, as shown in Fig. 14 and Fig. 15. Clearly, the wear morphologies varied obviously by test
condition.
The worn surface of Q-900–6 was smooth, and the wear debris on the contact interface was crushed under load to form an irregular

12
D. Zhang et al. Engineering Failure Analysis 134 (2022) 106080

film under 5 N-2 Hz. According to the EDS results, Cr, and Mn contents of 13.94% and 2.1% were detected in the film zone (Fig. 15a),
suggesting that the main component of the wear debris was transferred from the SUS 304 steel. A large amount of the wear debris was
crushed by the steel ball, and the debris became uniform and fine particles after repeated rolling. Then, a layer of steel transfer film was
formed on the friction surface by linking small wear particles, which reduced the friction coefficient and improved the wear resistance
of the friction pair. However, this kind of transfer film had a poor bond strength and might be destroyed under higher contact pressure.
Under dry sliding conditions, an increase in the sliding speed usually leads to an increase in the interface temperature, which affects
the friction and wear properties of coatings. Furthermore, the applied load could significantly affect the contact stress of the pair
surface, thus affecting the tribological properties of the friction pairs [40]. Under 10 N-6 Hz condition, the wear morphology of the Q-
900–6 became relatively rough and was severely worn, showing substantial lamellar exfoliation. In addition, the EDS results in Fig. 15b
exhibit much lower Cr and Mn contents than those under 5 N-2 Hz, where a metal transfer film was seldom observed. This finding
suggests that the metal film was easily peeled off and destroyed under high sliding speeds and loads, and thus had no antifriction effect.
Nevertheless, given the high thickness and hardness of the boride layer, the abrasive and adhesive wears were effectively reduced,
which was reflected in the presence of only slight and narrow wear tracks on the worn surface, endowing the surface with good wear
resistance. The main wear mechanism of the boride layer was slight abrasive wear combined with adhesive wear [41].
In addition, Fig. 14a2-b2 display SEM images of the steel ball wear morphologies. At 5 N-2 Hz, the steel ball had a wear track
diameter of 1192 μm and a rather smooth worn surface, where fine furrows were present, accompanied by the accumulation of wear
particles. At 10 N-6 Hz, the wear track diameter of the steel ball enlarged to 2202 μm, which is 1.8 times that under 5 N-2 Hz.
Furthermore, the worn surface roughness increased, and obvious plastic deformation was present. When the steel ball slid against the
boride layers, the asperity on the high-hardness boride surfaces pressed the softer stainless steel surface. The main wear types were
plowing and fracture for the elastic contact between two sliding surfaces and plowing and brittle spalling for the plastic contact, of
which the latter wear was more severe. With an increase in the applied load, the contact state changed from elastic to plastic, which led
to an intensified wear of the sliding ball.

4. Conclusions

In this paper, the boride layers were obtained on ductile cast iron through boronizing treatment. The effects of boronizing

Fig. 14. SEM images of the wear scars on Q-900-6 and the steel balls under test conditions of (a) 5 N-2 Hz and (b) 10 N-6 Hz.

13
D. Zhang et al. Engineering Failure Analysis 134 (2022) 106080

Fig. 15. EDS spectra of Q-900-6 under test conditions of (a) 5 N-2 Hz and (b) 10 N-6 Hz.

parameters on the morphology, thickness and hardness of the boride layers were explored, and the tribological properties of the boride
layers were investigated under different test conditions. The main conclusions are as follows:
(1) The boride layers were comprised primarily of FeB and Fe2B phases when the boronizing agent without adding La2O3. The
thicknesses of the boride layers increased significantly with increasing boronizing temperature and increased markedly with prolonged
boronizing time. In comparative terms, the temperature had a greater effect on the boride layer thickness than the boronizing time. A
higher temperature resulted in stronger activity of B atoms, and the thickness and hardness of the boride layers are also higher.
(2) Under low-speed light-load and low-speed heavy-load, the wear resistances of the boride layers were enhanced with increasing
boronizing temperature and time. When under a high-speed or heavy-load, the wear resistances were first enhanced and then
weakened with increases in the boronizing temperature and time. The tribological tests indicate that the boride layer has the best wear
resistance when borided at 900 ℃ for 6 h.
(3) The boride layers were comprised primarily of Fe2B phase as adding La2O3 into the boronizing agent. When the La2O3 content
was 4% during the boronizing process at 900 ◦ C for 4 h, the thickness and wear rate of the boride layer were 72.6 μm and 6.8 mm3/Nm,
respectively, which were clearly superior to those of the layer without La2O3. However, the thickness and wear resistance of boronizing
layer first increased and then decreased with increasing of La2O3 in the boronizing agent.

CRediT authorship contribution statement

Dongya Zhang: Investigation, Writing – original draft, Funding acquisition, Formal analysis. Yue Li: Data curation, Methodology.
Xin Du: Data curation, Methodology. Hongwei Fan: Conceptualization, Supervision, Resources. Feng Gao: Formal analysis, Project
administration.

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to
influence the work reported in this paper.

Acknowledgements

Acknowledgments: this work was support by the National Natural Science Foundation of China (No.52175186), Chinese post­
doctoral Project (No.2017 M613171), Postdoctoral foundation of Shaanxi province (No.2018BSHEDZZ09). Shaanxi key laboratory of
mine electromechanical equipment intelligent monitoring (No. -MEEIM201908).

References

[1] D. Papageorgiou, C. Medrea, N. Kyriakou, Failure analysis of H13 working die used in plastic injection moulding, Eng. Fail. Anal. 35 (2013) 355–359.
[2] P. Papadopoulos, M. Priest, W.M. Rainforth, Investigation of fundamental wear mechanisms at the piston ring and cylinder wall interface in internal combustion
engines, P I Mech. Eng. J-J Eng. 221 (3) (2007) 333–343.
[3] Y. Du, X. Gao, X. Wang, X. Wang, Y. Ge, B. Jiang, Tribological behavior of austempered ductile iron (ADI) obtained at different austempering temperatures,
Wear 456-457 (2020) 203396, https://doi.org/10.1016/j.wear.2020.203396.
[4] D.A. Colombo, A.D. Mandri, M.D. Echeverría, J.M. Massone, R.C. Dommarco, Mechanical and tribological behavior of Ti/TiN and TiAl/TiAlN coated
austempered ductile Iron, Thin Solid Films 647 (2018) 19–25.
[5] D.B. Salvaro, R.O. Giacomelli, R. Binder, C. Binder, A.N. Klein, J.D.B. de Mello, Assessment of a multifuncional tribological coating (nitride+DLC) deposited on
grey cast iron in a mixed lubrication regime, Wear 376–377 (4) (2017) 803–812.
[6] A. Erdogan, Boriding temperature effect on micro-abrasion wear resistance of borided tool steel, ASME. J. Tribol. 141 (12) (2019), 121702.
[7] P. Dziarski, N. Makuch, M. Tuliński, Thermal-oxidation behavior of single-phase boride layers produced on pure iron, Corros. Sci. 195 (2022) 109985, https://
doi.org/10.1016/j.corsci.2021.109985.

14
D. Zhang et al. Engineering Failure Analysis 134 (2022) 106080

[8] R.C. Morón, G.A. Arellano-Ortiz, G.A. Rodríguez-Castro, A. Meneses-Amador, A. Cruz-Ramírez, J.V. Méndez-Méndez, I. Campos-Silva, Wear performance under
dry and lubricated conditions of post boriding heat treatment in 4140 steel, ASME, J. Tribol. 143 (2) (2021), https://doi.org/10.1115/1.4047789.
[9] E. Medvedovski, M. Antonov, Erosion studies of the iron boride coatings for protection of tubing components in oil production, mineral processing and
engineering applications, Wear 452-453 (2020) 203277, https://doi.org/10.1016/j.wear.2020.203277.
[10] J. Jiang, Y.i. Wang, Q. Zhong, Q. Zhou, L. Zhang, Preparation of Fe2B boride coating on low-carbon steel surfaces and its evaluation of hardness and corrosion
resistance, Surf. Coat. Tech. 206 (2-3) (2011) 473–478.
[11] M. Keddam, R. Chegroune, A model for studying the kinetics of the formation of Fe2B boride layers at the surface of a gray cast iron, Appl. Surf. Sci. 256 (16)
(2010) 5025–5030.
[12] M. Kul, K.O. Oskay, A. Temizkan, B. Karaca, L.C. Kumruoğlu, B. Topçu, Effect of boronizing composition on boride layer of boronized GGG-60 ductile cast iron,
Vacuum 126 (4) (2016) 80–83.
[13] M.S. Gök, Y. Küçük, A. Erdoğan, M. Öge, E. Kanca, A. Günen, Dry sliding wear behavior of borided hot-work tool steel at elevated temperatures, Surf. Coat.
Tech. 328 (2017) 54–62.
[14] Y. Kayali, Y. Yalçin, Ş. Taktak, Adhesion and wear properties of boro-tempered ductile iron, Mater. Des. 32 (8-9) (2011) 4295–4303.
[15] O. Azouani, M. Keddam, O. Allaoui, A. Sehisseh, Characterization of boride coatings on a ductile cast iron, Prot Met Phys Chem Surf 53 (2) (2017) 306–311.
[16] U. Sen, S. Sen, S. Koksal, F. Yilmaz, Fracture toughness of borides formed on boronized ductile iron, Mater. Des. 26 (2) (2005) 175–179.
[17] A. Erdoğan, Investigation of high temperature dry sliding behavior of borided H13 Hot work tool steel with nanoboron powder, Surf. Coat. Tech. 357 (2019)
886–895.
[18] F. Xie, X. Wang, J. Pan, Accelerate pack boriding with reused boriding media by simultaneously employing al and alternating current field, Vacuum 141 (7)
(2017) 166–169.
[19] X. Yuan, B. Xu, Y. Caib, Study on Cr-rare Earth-boronizing of the steel 45 at low temperature, Phys. Procedia 50 (6) (2013) 82–87.
[20] S.M. Lee, H.M. Chow, F.Y. Huang, B.H. Yan, Friction drilling of austenitic stainless steel by uncoated and PVD AlCrN-and TiAlN-coated tungsten carbide tools,
Int. J Mach. Tool Manu. 49 (1) (2009) 81–88.
[21] Y.S. Zhu, Y.X. Yin, J. Wu, Y.F. Liu, W. Lu, D. Zuo, H. Xiao, D. Cao, T.J. Ko, Effect of RE on accelerating the kinetics of boride layer growth on titanium alloy,
J. Alloy. Compd. 844 (2020) 156091, https://doi.org/10.1016/j.jallcom.2020.156091.
[22] Y. Du, X. Wang, D. Zhang, X. Wang, C. Ju, B. Jiang, A superior strength and sliding-wear resistance combination of ductile iron with nanobainitic matrix,
J. Mater. Res. Technol. 11 (2021) 1175–1183.
[23] H.P. Yang, X.C. Wu, Y.A. Min, T.R. Wu, J.Z. Gui, Plasma boriding of high strength alloy steel with nanostructured surface layer at low temperature assisted by air
blast shot peening, Surf. Coat. Tech. 228 (2013) 229–233.
[24] A. Erdogan, B. Kursuncu, A. Günen, M. Kalkandelen, M.S. Gok, a new approach to sintering and boriding of steels “Boro-sintering”: Formation, microstructure
and wear behaviors, Surf. Coat. Tech. 386 (2020) 125482, https://doi.org/10.1016/j.surfcoat.2020.125482.
[25] I. Uslu, H. Comert, M. Ipek, O. Ozdemir, C. Bindal, Evaluation of borides formed on AISI P20 steel, Mater. Des. 28 (1) (2007) 55–61.
[26] H. Aytekin, Y. Akçin, Characterization of borided Incoloy 825 alloy, Mater. Des. 50 (2013) 515–521.
[27] A. Günen, K.M. Döleker, M.E. Korkmaz, M.S. Gök, A. Erdogan, Characteristics, high temperature wear and oxidation behavior of boride layer grown on nimonic
80A Ni-based superalloy, Surf. Coat. Tech. 409 (2021) 126906, https://doi.org/10.1016/j.surfcoat.2021.126906.
[28] J. Lentz, A. Röttger, W. Theisen, Hardness and modulus of Fe2B, Fe3(C, B), and Fe23(C, B)6 borides and carboborides in the Fe-C-B system, Mater. Charact. 135
(2018) 192–202.
[29] S.P. Sharma, D.K. Dwivedi, P.K. Jain, Effect of La2O3 addition on the microstructure, hardness and abrasive wear behavior of flame sprayed Ni based coatings,
Wear 267 (5-8) (2009) 853–859.
[30] C. Wang, J. Li, X. Hu, Applications of rare-earth thermal diffusion technology in plastic injection mould cavity surface treatment, Second International
Conference on Digital Manufacturing & Automation (2011).
[31] Y. Kayali, İ. Güneş, S. Ulu, Diffusion kinetics of borided AISI 52100 and AISI 440C steels, Vacuum 86 (10) (2012) 1428–1434.
[32] P.K. Rohatgi, S. Ray, Y. Liu, Tribological properties of metal matrix-graphite particle composites, Int. Mater. Rev. 37 (1) (1992) 129–152.
[33] K. Holmberg, A. Mathews, Coatings tribology: A concept, critical aspects and future directions, Thin Solid Films 253 (1–2) (1994) 173–178.
[34] S.A. Humphry-Baker, W.E. Lee, Tungsten carbide is more oxidation resistant than tungsten when processed to full density, Scripta Mater. 116 (2016) 67–70.
[35] A. Motallebzadeh, E. Dilektasli, M. Baydogan, E. Atar, H. Cimenoglu, Evaluation of the effect of boride layer structure on the high temperature wear behavior of
borided steels, Wear 328–329 (2015) 110–114.
[36] İ. Türkmen, E. Yalamaç, M. Keddam, Investigation of tribological behaviour and diffusion model of Fe2B layer formed by pack-boriding on SAE 1020 steel, Surf.
Coat. Tech. 377 (2019) 124888, https://doi.org/10.1016/j.surfcoat.2019.08.017.
[37] B. Zheng, J. Xing, W. Li, X. Tu, Y. Jian, Effect of chromium-induced (Fe, Cr)3C toughness improvement on the two-body abrasive wear behaviors of white cast
iron, Wear 456-457 (2020) 203363, https://doi.org/10.1016/j.wear.2020.203363.
[38] J. Jin, J. Sun, G. Wang, Effect of Mo content on microstructure and wear resistance of Mo-Fe-B claddings, Int. J. Refract. Met. H. 81 (2019) 233–241.
[39] B. Xu, M. Li, Effect on decreasing the eigen brittleness of boride layer on its friction and wear behavior, Chin. J. Mech. Eng. 38 (11) (2002) 131–134.
[40] D. Zhang, J.K.L. Ho, G. Dong, H. Zhang, M. Hua, Tribological properties of Tin-based Babbitt bearing alloy with polyurethane coating under dry and starved
lubrication conditions, Tribol. Int. 90 (2015) 22–31.
[41] A. Gåård, N. Hallbäck, P. Krakhmalev, J. Bergström, Temperature effects on adhesive wear in dry sliding contacts, Wear 268 (7-8) (2010) 968–975.

15

You might also like