You are on page 1of 7

Wear 362-363 (2016) 1–7

Contents lists available at ScienceDirect

Wear
journal homepage: www.elsevier.com/locate/wear

Tribological behavior of borided AISI 1018 steel under linear


reciprocating sliding conditions
R. Carrera-Espinoza a,n, U. Figueroa-López a,n, J. Martínez-Trinidad b, I. Campos-Silva b,
E. Hernández-Sánchez c, Amir Motallebzadeh d
a
Tecnologico de Monterrey, Campus Estado de México, Carretera Lago de Guadalupe km. 3.5, Margarita Maza de Juárez, 52926 Atizapán de Zaragoza, México
b
Instituto Politécnico Nacional, Grupo Ingeniería de Superficies, SEPI-ESIME, U.P. Adolfo López Mateos, Zacatenco, México D.F. 07738, México
c
Instituto Politécnico Nacional, UPIBI, Av. Acueducto s/n Barrio La Laguna Ticomán, México D.F. 07340, México
d
Istanbul Technical University, Department of Metallurgical and Materials Engineering, Maslak, 34469 Istanbul, Turkey

art ic l e i nf o a b s t r a c t

Article history: In this study, the wear behavior of AISI 1018 borided steel is characterized. The boriding thermochemical
Received 29 December 2015 treatment was carried out at 1273 K for 6 h, and different thicknesses of the boriding agent surrounding
Received in revised form the samples were examined. The layer structure was composed of FeB and Fe2B layers.
29 April 2016
The tribological properties at the surface of borided and unborided steel were evaluated employing
Accepted 6 May 2016
Available online 14 May 2016
the ball-on-flat method with sliding reciprocating wear tests, using an Al2O3 ball as the counterpart. The
applied load and the sliding speed remained constant at 5 N and 5 mm/s, respectively, with sliding
Keywords: distances of 81 m, 108 m and 135 m.
Sliding wear The coefficients of friction (CoFs) on the boride layers from the sliding reciprocating wear tests for dry
Reciprocating wear test
and lubricated conditions were evaluated. The CoFs are independent of the boriding potential expressed
Diffusion treatments
as the powder thickness. Surface wear for longer sliding distances in lubricated conditions was
Steel
Coefficient of friction neglected; there was no difference between the wear track and the original surface roughness in the
profilometer results. The wear scar diameter and the worn surfaces of borided and unborided steels in
both experimental conditions were characterized by SEM to understand the wear mechanisms.
& 2016 Elsevier B.V. All rights reserved.

1. Introduction single Fe2B layer or a FeB/Fe2B bilayer, with flat or saw-tooth


morphology. The latter morphology is commonly formed on low-
Sliding wear of metals in machinery components is costly to carbon and low-alloy steels. The former is mainly produced on
industry and is often associated with adhesive contact [1,2]. Sur- high-carbon and highly alloyed steels. The development of one or
face treatments, such as carburizing, nitriding and boriding, have another structure depends on different parameters such as the
been used to reduce adhesive wear, but boriding has advantages chemical composition of the substrate, temperature/time, the
over the former two processes as it improves the surface hardness, boron potential at the surface, the activator used in the boriding
adhesive wear and high-temperature resistance [3–5]. The com- agent or media and the treatment method [14]. However, con-
sidering the mechanical properties, saw-toothed morphology is
bination of such surface properties makes this process an excellent
the preferred process, as its structure strongly adheres to the
candidate for specific industrial applications, such as gear boxes
substrate [14,15].
for turbines, cam-shafts, weapons, parts of agricultural machinery,
Boride coatings can be produced by liquid, gaseous or solid
etc., that require high tribological performance [6–10].
procedures, which have their respective advantages. For industrial
Boriding is a thermochemical treatment that involves the dif-
applications, boriding in solid state is more desirable than the
fusion of boron atoms at the surface of different ferrous and non- other two methods due to its ease of treatment, simplicity of the
ferrous materials at temperatures ranging from 1123 K to 1273 K required equipment and cost-effectiveness [16,17].
and with exposure time ranging from 1 to 10 h or more [11–13]. The superior contribution of the boride layer to the wear
The result of the treatment for an iron-based alloy consists of a resistance of some materials under dry sliding conditions has been
well-established. However, tribological research on borided sur-
n
Corresponding authors. Tel.: þ 52 55 5864 5555x2436; fax: þ52 55 5864 5751.
faces under the lubricated condition has received less attention.
E-mail addresses: rafaelcb05@gmail.com (R. Carrera-Espinoza), Some of these studies were conducted on borided AISI 52100 steel
ufiguero@itesm.mx (U. Figueroa-López). and Ti6Al4V alloy under dry and lubricating conditions [18,19].

http://dx.doi.org/10.1016/j.wear.2016.05.003
0043-1648/& 2016 Elsevier B.V. All rights reserved.
2 R. Carrera-Espinoza et al. / Wear 362-363 (2016) 1–7

The borided materials exhibited superior tribological performance polished with 1 mm diamond paste for both conditions. A 6 mm
in terms of remarkable wear resistance and low coefficients of diameter Al2O3 ball was utilized as the counterpart at ambient
friction compared to those without surface modification. The temperature (24 7 2 °C and 367 2% RH) under dry sliding for 81,
combination of surface roughness and high hardness could have 108 and 135 m sliding distances. Three wear tests were performed
contributed to decrease the coefficient of friction. Although the for each condition. Previous studies showed that the wear
boride layers of 1018 steel are well-known to be compact, hard mechanism and the friction coefficient of such borided steels
and adherent [20]. The present study was conducted to evaluate remain constant until the boride layer is removed [8,13]. There-
the tribological performance of boride and unborided low-carbon fore, we chose a short sliding distance to be sure that we remained
1018 steel. On the sliding reciprocating wear method, especially on the boride layer. In addition, in this reciprocating tribometer
the tribo-pair boriding steel and alumina ball under lubricated and there is a limit on the wear test speed. To increase the sliding
dry sliding conditions. Thus, the boride layers formed on low- distance and conduct the wear test in short exposure times, we
carbon steel under lubricated sliding conditions could be applied, had to increase the sliding speed. However, increasing the sliding
for example, to the manufacture of bearings for universal joints speed was beyond the tribometer's specifications. Steady-state
and other automotive applications where sliding wear prevails. friction values remain constant in short and long sliding distances
until the boride layer is removed. Wear data for the treated sur-
faces could not be usefully compared in experiments during which
2. Experimental procedure the boride layer was removed. The stroke length and the sliding
speed of the ball were 5 mm and 0.005 m/s, respectively. These
2.1. Powder-pack boriding process parameters were chosen according to the limits of the tribometer
and the size samples. A previous study showed that increasing the
The boriding treatment was carried out at a temperature of wear temperature to 500 °C (523 K) under dry sliding does not
1273 K for 6 h on AISI 1018 low-carbon steel samples without using affect the tribological behavior of borided steel. During the wear
controlled atmosphere; a conventional furnace brand Lindberg that testing, frictional force data were continuously recorded. The test
reaches a working temperature of 1273 K was used. The chemical load of dry and lubricated sliding wear was 5 N, which corre-
composition of AISI 1018 steel is shown in Table 1. Work pieces sponded to maximum contact pressures of about 1.28 GPa and
12.7 mm in diameter and 5 mm thick were employed. They were 1.45 GPa for the unborided and borided samples, respectively.
placed inside an AISI 304 stainless steel multicontainer; 6 samples However, when the wear tests were conducted under different
can be included at the same temperature and time. The boron loads (1, 3, 5, 6, 7 and 8 N), light loads such as 1 and 3 N did not
potential was varied through the different mixture thicknesses (from form a clear and obvious wear track for wear loss measurements.
5 to 30 mm). It was made with 70% SiC, 20% B4C and 10% KBF4. In addition, at higher loads such as 6, 7 or 8 N under the dry sliding
condition, the friction coefficient increased, and due to safety
2.2. Structural examinations conditions, the tribometer stopped. Alumina is widely used as a
counterpart because the system (alumina vs. hard coating) has
The presence of the FeB and Fe2B layers was verified with X-ray already been proven to show higher wear resistance and sig-
diffraction (XRD). XRD patterns were obtained with a PANalytical nificant differences in material behavior during sliding. This may
diffractometer model X’PER PRO MRD equipped with CuKα be related to the high hardness, high oxidation resistance and high
radiation at λ ¼ 0.154 nm. The growth of the iron boride layers at surface chemical inertness of the Al2O3 ball [21]. Sliding tests
the surface of the AISI 1018 steel sample was analyzed using a JEOL under lubrication were carried out using a fully formulated syn-
JSM-6360LV scanning electron microscope (SEM). thetic engine oil HX7 10W-40, which was smeared on the surface
of the samples to simulate boundary lubrication [19]. The results
2.3. Hardness and surface roughness of the wear tests were evaluated by calculating the area of the
wear tracks formed on the surface of the samples. Width and
Vickers microindentation tests were conducted at 50, 100, 150, depth measurements were taken using an optical profilometer. In
210, 260, 380 and 560 μm from the surface of the borided layers to addition, wear tracks were examined using a scanning electron
the substrate using a Wolpert 402 MVD apparatus. The values are microscope (SEM-HITACHI TM-100, Japan). Coated steels under
the average of five indentations with a constant indentation load lubricated conditions were sonically cleaned in acetone for 10 min
of 100 g and a dwell time of 15 s according to the ASTM-E384 and before the microscope observations. This procedure was also uti-
ASTM C1327 procedures. In addition, the surface roughness (Ra) lized for cleaning the counterparts. The contact surfaces of the
was measured on the borided and unborided samples by using a alumina ball were examined using a light optical microscope
Veeco optical profilometer model DEKTAK 6M – STYLUS PROFILER (LOM-Leica DM750M, Germany).
and DEKTAK 32 software.

2.4. Reciprocating wear tests 3. Results

Sliding friction and wear tests of the borided and unborided A cross-section SEM micrograph of the borided steel is shown
steel samples under dry and lubricated conditions were carried in Fig. 1. Borided layers exhibited a dual-layer structure, the FeB
out on a ball-on-disk tribometer with reciprocating motion. Prior and Fe2B layer, as shown with XRD in Fig. 2. Saw-toothed mor-
to the tribological experiments, the coated samples were slightly phology at the interfaces of the FeB/Fe2B and Fe2B/substrate was
grounded with abrasive paper with a grit size of 2500 and then observed. This morphology has been verified and reported
[11,12,16,22–25]. The average thickness of the total layer was
Table 1 about 215 μm. The measurement of individual layer thicknesses
Chemical composition of AISI 1018 steel (wt. %). showed that the FeB layer thickness is a function of the boron
potential, with values ranging from 80 to 108 μm.
C Si Mn S P Fe Hardness on the cross section of the borided AISI 1018 steel
0.15–0.20 0.15–0.35 0.60–0.90 0.050 (max) 0.040 (max) Balance
samples with different boron potentials is shown in Fig. 3. The
profiles show that the FeB phase the hardness is in the range of
R. Carrera-Espinoza et al. / Wear 362-363 (2016) 1–7 3

mechanism on the dry and lubricated unborided samples was


similar, exhibiting similar features, which can distinguished by the
surface roughness appearance because of the deformation asso-
FeB layer ciated with the plowing effects of the counterpart of the Al2O3 ball.
Fig. 5 shows typical profilometry plots measured across the
wear tracks of the borided and unborided samples of AISI 1018
steel under dry and lubricated conditions. The unborided samples
Fe2B layer are deeply worn in either dry or lubricated conditions. However,
there was sensible wear on the borided sample when it was tested
in dry conditions. When the sample was combined with lubrica-
tion, the loss of material was negligible, and the wear track was
recorded by the profilometer where it included roughness effects,
can be considered negligible.
The results of the wear tests are measured in terms of the CoF
Substrate and the wear rate (WR). As the wear tests were carried out on
different sliding distances, such as 81 m, 108 m and 135 m, the
WRs were calculated after the wear track area of the wear tracks
formed on the surface of the borided and unborided samples of
AISI 1018 steel for the sliding distances. First, the CoF behavior was
Fig. 1. Cross-sectional view of AISI 1018 borided steel exposed at a temperature of determined from the steady state of the friction curves after cer-
1273 K with 6 h of exposure and 10 mm thick mixture.
tain fluctuant periods during the test. Fig. 6 shows the friction
curves of (a) borided and (b) unborided samples tested under dry
sliding conditions and (c) borided and (d) unborided samples
tested under lubricated sliding conditions. The common features
of the friction curves of the unborided samples are slight fluc-
tuations on either dry or lubricated sliding tests. Lubrication
caused an abrupt reduction in the fluctuations in the borided
samples.
The CoF obtained in reciprocating wear tests for the tribo-pair
borided steel surface–Al2O3 ball ranged from 0.088 70.003 for the
lubricated sliding conditions. The CoFs for the dry sliding condi-
tions were higher, about 0.6 7 0.06.
The WRs for dry sliding conditions presented behavior similar
to that of the CoF, given that the borided sample exhibited a higher
wear rate than the unborided sample.
Under the lubricated sliding conditions, the borided steel
sample presented considerably lower WRs, 0.667 0.01  10–6
Fig. 2. X-ray diffraction pattern obtained at the surface of AISI 1018 borided steel
with 10 mm of mixture thickness. (mm3/N m), when compared to those for the unborided samples,
6.32 70.11  10–6 (mm3/N m). Even when the CoFs for the two
FeB Fe B Diffusion Zone Substrate samples were similar, the high WRs on the unborided samples
1400
were associated with the high plastic deformation that occurred
5 mm on the contact surface due to their relative softness against the
1200 10 mm Al2O3 ball (see Fig. 4c).
1000 15 mm Surface hardness has a significant effect on the abrupt reduc-
Hardness (HV)

tion of the WR and the low CoF of the borided samples when there
20 mm
800 is lubricated sliding. The boride layers showed very limited wear
25 mm (see Fig. 4d). In addition, Fig. 7 shows the behavior of the WRs of
600
30 mm the borided samples in dry and lubricated sliding conditions,
400 respectively, and the considerable reduction in the wear rates for
200 the tests in the lubricated sliding conditions in comparison with
the values obtained for the dry sliding condition is evident.
0 Wear scars were generated on the surface of the Al2O3 ball by
0 100 200 300 400 500 600
Distance from the surface (µm) contact with the borided steel sample under the dry sliding con-
ditions. Fig. 8b shows the plowing effect without the adhesion of
Fig. 3. Cross-section hardness profiles of the borided AISI 1018 steel with varied debris particles of the counterpart. The presence of the plowing
boron potential.
effect can be correlated with the high surface hardness of the
borided samples, and the parallel grooves are common features of
950–1100 HV100, whereas, in the compact zone of the Fe2B phase,
wear among ceramic materials [26–29]. However, only the plow-
the hardness reaches a maximum value of 1200 HV100, and in the
ing effect with the adhesion of the base material was observed on
substrate, the hardness is about 135 HV100. the surface of the Al2O3 ball when it was in contact with the
The surface roughness of the borided samples was higher than unborided steel, as shown in Fig. 8a.
that of the unborided samples and other borided steels [17]. The Similar to the tested tribo-pair borided steel against an Al2O3
estimated surface roughness was 0.02 μm and 0.23–0.33 μm for ball under dry sliding conditions, the surfaces of the Al2O3 ball
the unborided and borided samples, respectively. present a bright aspect after the lubricated sliding tests [19]. In this
Worn surface SEM micrographs of the borided and unborided case, the bright aspect of the Al2O3 ball could be attributed to the
AISI 1018 steel samples are presented in Fig. 4. The adhesive separation of the surfaces as a consequence of the lubricant film.
4 R. Carrera-Espinoza et al. / Wear 362-363 (2016) 1–7

Testing
AISI 10
018 unborided
d steel AISI 10
018 borided stteel
conditions
Plastic Deformaation Adhhesion of debris

Adhesion of material
m
Dry
sliding

Cracking

Teest Load = 5N
N Tesst Load = 5N
Contact Pressure = 1.2
28 GPa Pressure = 1.455 GPa
Contact P
Ploughiing Polishing

Lubricated
sliding
Plastic Defformation Porosiity

Teest Load = 5N
N Tesst Load = 5N
Contact P
Pressure = 1.2
28 GPa Contact P
Pressure = 1.455 GPa
Fig. 4. Worn surface SEM micrographs of AISI 1018 borided and unborided steel. Dry sliding test (a) unborided and (b) borided steel; lubricated sliding test (c) unborided and
(d) borided steel.

Even when the lubricant reduces the shear stresses and limits the Severe plastic deformation was observed on the worn surface
material transfer from the unborided steel to the Al2O3 ball, slight of the unborided samples under dry sliding conditions, sur-
plowing can be observed on the surface of the Al2O3 ball, as shown rounding the adhesion zones along with grooves parallel to the
in Fig. 8c. Fig. 8d shows a slight plowing effect with a little bit sliding direction. This behavior can be correlated with the ductility
adhesion of debris particles of the counterpart. of the substrate and the high hardness of the counterpart. In
lubricated sliding conditions, plastic deformation was not severe,
and the plowing effect was observed along the wear track.
4. Discussion The boride steel was slightly damaged under the dry sliding
conditions and showed only the plowing effect as a wear
The boriding process applied in this study generated dual- mechanism. These observations can be correlated with the phe-
phase boride layers, FeB and Fe2B layers at the outer and inner nomenon of stick-slip at the boride layer–alumina ball contact
sections, respectively. The absence of alloy elements combined zone, which encouraged cracking (Fig. 4b). As the result of plow-
with the low-carbon content in this steel promotes saw-tooth ing, the wear mechanism was altered to micro-abrasion; fine
layer morphology (at the interfaces of FeB/Fe2B and Fe2B/sub- groves detected in the wear tracks of the surface of the boride
strate), and the layers grow faster due to the absence of obstacles layer can be correlated with the third body abrasive effect of the
across the flow of active boron. This kind of morphology leads to wear debris [13]. Even at considerably high contact pressures, the
better adhesion between the borided layers and the substrate. shift of wear mechanisms from plowing to polishing for the tests
The X-ray diffraction pattern showed evidence of iron borides under lubricated sliding conditions on the borided samples is
FeB and Fe2B corresponding to the dual-layer structure generated clear. The wear tracks on the boride layer under the lubricated
in AISI 1018 steel. The hardness of the cross section of the borided sliding conditions did not present evidence of cracking, delami-
AISI 1018 steel shown in Fig. 3 ranged from 950 to 1100 HV100 in nation or spalling, only the polishing and shallow plowing effect,
the FeB phase, but the maximum value of 1200 HV100 was reached without forming a measurable wear track (Fig. 4d).
in the zone of the Fe2B phase. The lower hardness in FeB is The CoF of about 0.7 could be associated with various factors,
explained by its porosity–surface zone affecting the hardness such as the surface roughness, porosity, as well as slight adhesion
values. This behavior is common in the iron boride layers formed of the debris particles detached from the borided steel [32]. The
on the surface of low-carbon steels when they are compared with CoF is related mainly to surface roughness, hardness, lubricating
high-alloyed steels or high-carbon steel [17,30,31]. Furthermore, in conditions and other surface effects, such as local chemical inter-
this study the hardness values did not show dependence on the actions at the interface of the tribo-pair [32–37]. In the present
boron potential. work, the CoFs are independent of the boron potential expressed
R. Carrera-Espinoza et al. / Wear 362-363 (2016) 1–7 5

Testing
AISI 1018 unborided steel AISI 1018 borided steel
conditions

Depth of the wear track (µm)


Depth of the wear track (µm)
2 2
0 0
-2 -2
-4 -4
-6 -6
Dry
-8 -8
sliding -10 -10
-12 -12
0 500 1000 1500 2000 0 500 1000 1500 2000
Width of the wear track (µm) Width of the wear track (µm)
Depth of the wear track (µm)

2 2

Depth of the wear track (µm)


0 0
-2 -2
-4 -4
-6 -6
Lubricated
-8 -8
sliding
-10 -10
-12 -12
0 500 1000 1500 2000 0 500 1000 1500 2000
Width of the wear track (µm) Width of the wear track (µm)

Fig. 5. Profilometry graphs of wear tracks of unborided and borided samples of AISI 1018 steel under dry and lubricated sliding conditions. Dry sliding test (a) unborided and
(b) borided steel; lubricated sliding test (c) unborided and (d) borided steel.

25
Dry Lubricated
20
Specific wear rate
x10 -6 [mm 3/Nm]

15

10

0
81 108 135
Sliding distance (m)
Fig. 7. Specific wear rate graph of borided samples in dry and lubricated sliding
conditions, respectively.

in the CoF. Furthermore, the wear rates in these sliding conditions


depict about 89.6% reduction in the WRs (Fig. 7).
The lubricating conditions benefit the wear rate of the borided
steel because of the very low CoF (0.086–0.091) but also because
polishing was the main wear mechanism in these conditions.
Fig. 6. Friction curves of (a) borided and (b) unborided samples tested under the However, the CoFs for the dry sliding tests for the tribo-pair
dry sliding conditions; (c) unborided and (d) borided samples tested under the borided steel-Al2O3 ball were high, which can be attributed to the
lubricated sliding conditions.
roughness of the borided layers, which was much higher than that
on the unborided surfaces. CoFs are mainly related to surface
as powder thickness and are strongly influenced by surface effects hardness and roughness [32–37]. Roughness and porosity play an
and lubricating conditions. important role when the CoF is measured, so it is important to
The lubrication for the tribo-pair borided steel–Al2O3 ball estimate it; higher roughness and porosity increase the CoF [32].
provides a considerable benefit on the wear resistance of the Moreover, high CoFs for this kind of tribo-pair (borided steel
borided layers formed on the surface of the substrate, generating against alumina ball) under dry sliding conditions had been
CoFs in the range of 0.086–0.091 that depict about 85.7% reduction reported [13].
6 R. Carrera-Espinoza et al. / Wear 362-363 (2016) 1–7

Testing
AISI 10
018 unborided
d steel AISI 10
018 borided stteel
conditions
Surface of Surface of
Al2O3 ball Al2O3 ball

Dry
Adhesion of
sliding debris

Cracking

Surface of Surface of
Al2O3 ball Al2O3 ball

Lubricated
sliding
Adhesion of
debris

Fig. 8. Worn surface appearance of Al2O3 ball. Dry sliding test on AISI 1018 (a) unborided and (b) borided steel; lubricated sliding test on AISI 1018 (c) unborided and
(d) borided steel.

The borided steel exhibited a higher tribological performance 5. Conclusions


than the unborided steel even for higher sliding distances, as well
as higher contact pressure. Furthermore, the fact that the depth of The effects of boride treatments on the microstructure, hard-
the wear track was about 7 μm, shows that only was damaged the ness, and wear of AISI 1018 steel sliding under both dry and
FeB layer and there was no effect observed of the Fe2B layer in the lubricated conditions against alumina were investigated. The fol-
CoFs or WRs. Lubrication serves as a third body during the sliding lowing conclusions can be made:
tests of the borided steel, considerably reducing the friction among
the tribo-pair and preventing the development of wear tracks. a) The system obtained was of FeB and Fe2B layers, only the outer
The plowing effect with the adhesion of the base material as layer FeB thickness had a dependency of the boron potential
well as slight cracking of the adhered particles was observed on expressed as the powder thickness surrounding the surfaces of
the surface of the Al2O3 ball when it was in contact with the the samples.
unborided steel. This particular structure of the wear scar on the b) The CoFs and WRs for the tribo-pair Al2O3 ball and borided
Al2O3 ball can be explained because softer material fragments, or samples under the lubricated sliding conditions showed a
reduction of approximately 85.7% and 89.6% respectively, in
iron debris from the unborided steel substrate, are pulled out and
comparison with the dry sliding conditions.
attached by adherence to the surface of the Al2O3 ball. Because
c) The fact that the depth of the wear track on the borided
sliding contact between both materials continues during the test,
sample under dry sliding conditions was about 7 μm, shows
some of the debris particles crack, giving a network of cracks on
that only was damaged the FeB layer and there was no effect
the debris adhered to the surface of the Al2O3 ball. observed of the Fe2B layer in the CoFs or WRs.
Slight plowing and shallow grooves were observed on the d) The borided layers on AISI 1018 steel exhibited superior tri-
surface of the Al2O3 ball after lubricated sliding of Al2O3 ball on the bological performance against Al2O3 ball by remarkably low-
unborided steel. Although direct contact did not occur, the sur- ering the wear rate 0.66 70.01  10–6 (mm3/N m) and the
faces of the Al2O3 ball still experienced large stresses. In this case, friction coefficient μ ¼0.088 70.003, polishing was the prin-
wear fragments were easily removed from the surface of the cipal wear mechanism that appeared on the worn surfaces
unborided steel due to its lower wear resistance. With the evo- under the lubricated sliding conditions.
lution of the sliding on the wear track, the hard iron oxide parti-
cles formed parallel scratches aligned in the sliding direction of
the surface of the Al2O3 ball. This wear mechanism did not take Acknowledgments
place on the surface of the ball used with the borided steel.
Because of the high hardness of the borided sample surface, This work was supported by research grant 45799 and a
deformation was limited, the debris production was insignificant CONACyT, Mexico, visiting student scholarship. The authors wish
and the lubricated film limited the adhesion. Nevertheless, micro- to thank the Department of Metallurgical and Materials Engi-
plowing mechanisms, as well as slight adhesion of the debris neering Laboratories of the Istanbul Technical University, the
particles detached from the borided steel, were observed [18,19]. Instituto Tecnológico y de Estudios Superiores de Monterrey
R. Carrera-Espinoza et al. / Wear 362-363 (2016) 1–7 7

Campus Estado de México and the Nanosciences Center and [19] E. Atar, E. Sabri Kayali, H. Cimenoglu, Characteristics and wear performance of
Micro-Nano Technologies Laboratory of the Instituto Politécnico borided Ti6Al4V alloy, Surf. Coat. Tech. 202 (2008) 4583–4590.
[20] I. Campos-Silva, E. Hernández-Sánchez, G. Rodríguez-Castro, H. Cimenoglu, J.
Nacional for their cooperation. L. Nava-Sánchez, A. Meneses-Amador, R. Carrera-Espinoza, A study of inden-
tation for mechanical characterization of the Fe2B layer, Surf. Coat. Technol.
232 (2013) 173–181.
[21] R. Ramadoss, N. Kumar, S. Dash, D. Arivuoli, A.K. Tyagilnt, Wear mechanism of
References CrN/NbN superlattice coating sliding against various counterbodies, Int. J.
Refract. Met. Hard Mater. 41 (2013) 547–552.
[1] E. Rabinowicz, Friction and Wear of Materials, second ed., John Wiley & Sons, [22] I. Campos, O. Bautista, G. Ramírez, M. Islas, J. de la Parra, L. Zuñiga, Effect of
Inc, New York, 1995. boron paste thickness on the growth kinetics of Fe2B boride layers during the
[2] F.P. Bowden, D. Tabor, The Friction and Lubrication of Solids, Oxford Classic boriding process, Appl. Surf. Sci. 243 (2005) 429–436.
Series, Oxford, 2001. [23] A.J. Ninham, I.M. Hutchings, On the morphology of thermochemically pro-
[3] T.S. Eyre, Effect of boronising on friction and wear of ferrous metals, Wear 34 duced Fe2B/FeB interfaces, J. Vac. Sci. Technol. A 4 (1986) 2827–2831.
(1975) 383–397. [24] S. Taktak, Tribological behavior of borided bearing steels at elevated tem-
[4] J. Rus, C. Luis de Leal, D.N. Tsipas, Boronizing of 304 steel, J. Mater. Sci. Lett. 4 peratures, Surf. Coat. Technol. 201 (2006) 2230–2239.
(1985) 558–560. [25] V. Jain, G. Sundararajan, Influence of the pack thickness of the boronizing
[5] P. Gopalakrishnan, P. Shankar, R.V. Subba Rao, M. Sundar, S.S. Ramakrishnan, mixture on the boriding of steel, Surf. Coat. Technol. 149 (2002) 21–26.
Laser surface modification of low carbon bonded steels, Scr. Mater. 44 (5) [26] P.A. Dearnley, K.L. Dahm, H. Çimenoglu, The corrosion–wear behaviour of
(2001) 707–712. thermally oxidised CP-Ti and Ti–6Al–4V, Wear 256 (5) (2004) 469–479.
[6] S. Sen, U. Sen, C. Bindal, Tribological properties of oxidised boride coatings [27] V.J. Colangelo, F.A. Heiser, Analysis of Metallurgical Failures, John Wiley, New
grown on AISI 4140 steel, Mater. Lett. 60 (29–30) (2006) 3481–3486. York, 1974.
[7] A. Greco, K. Mistry, V. Sista, O. Eryilmaz, A. Erdemir, Friction and wear beha- [28] I.M. Hutchings, Tribology: Friction and Wear of Engineering Materials, Edward
viour of boron based surface treatment and nano-particle lubricant additives Arnold, London, 1992.
for wind turbine gearbox applications, Wear 271 (2011) 1754–1760. [29] B. Bhushan, Principles and Application of Tribology, Wiley-Interscience Pub-
[8] W. Fichtl, Boronizing and its practical applications, Mater. Eng. 2-6 (1981) lication, New York, 1999.
276–286. [30] M. Carbucicchio, E. Zecchi, G. Palombarini, G. Sambogna, Phase composition
[9] I. Ozbek, C. Bindal, Mechanical properties of boronized AISI W4 steel, Surf. and structure of boride layers grown on laboratory-cast low-chromium alloys,
Coat. Technol. 154 (2002) 14–20. J. Mater. Sci. 18 (1983) 3355–3362.
[10] Ü. Er, B. Par, The abrasive wear behaviour of boronized and untreated AISI [31] C. Badini, C. Gianoglio, G. Pradelli, The effect of carbon, chromium and nickel
1008 and AISI 1045 steels, Il, Uluslar. Sempozyumu (2004) 207–212. on the hardness of boride layers, Surf. Coat. Technol. 30 (1987) 157–170.
[11] A. G. Matuschka, Boronizing, Carl Hansen Verlag München, Wien, 1980. [32] B. Selcuk, R. Ipek, M.B. Karamis, A study on friction and wear behaviour of
[12] ASM Handbook, Heat Treating, Vol. 4, ASM International Handbook Commit- carburised, carbonitrided and borided AISI 1020 and 5115 steels, J. Mater.
tee, Ohio, 1995. Process. Technol. 141 (2003) 189–196.
[13] H. Cimenoglu, E. Atar, A. Motallebzadeh, High temperature tribological [33] C.C. Viafara, A. Sinatora, Influence of hardness of the harder body on wear
behaviour of borided surfaces based on the phase structure of the boride layer, regime transition in a sliding pair of steels, Wear 267 (2009) 425–432.
Wear 309 (2014) 152–158. [34] C.C. Viafara, A. Sinatora, Unlubricated sliding friction and wear of steels: an
[14] C. Martini, G. Palombarini, G. Poli, D. Prandstraller, Sliding and abrasive wear evaluation of the mechanism responsible for the T 1 wear regime transition,
behavior of boride coatings, Wear 256 (2004) 608–613. Wear 271 (2011) 1689–1700.
[15] H.J. Hunger, G. Trute, Boronizing to produce wear-resistant surface layers, Heat [35] G.E. Yi-cheng, Y. Mao-zhong, L. Li-ya, Influence of load on sliding tribology of
Treat. Met. 2 (1994) 31–39. C/C composite with 40Cr steel couple coated by Cr, Trans. Nonferrous Met.
[16] M. Keddam, S.M. Chentouf, A diffusion model for describing the bilayer growth Soc. China 17 (3) (2007) 570–574.
(FeB/Fe2B) during the iron powder-pack boriding, Appl. Surf. Sci. 252 (2) [36] K. De Moerlooze, F. Al-Bender, On the relationship between normal load and
(2005) 393–399. friction force in pre-sliding frictional contacts, part 2: experimental investi-
[17] S. Sahin, Effects of boronizing process on the surface roughness and dimen- gation, Wear 269 (3–4) (2010) 183–189.
sions of AISI 1020, AISI 1040 and AISI 2714, J. Mater. Process. Technol. 209 [37] E.A. Gallardo, R. Lewis, Twin disc assessment of wheel/rail adhesion, Wear 265
(2009) 1736–1741. (9–10) (2008) 1309–1316.
[18] E. Garcia-Bustos, M.A. Figueroa-Guadarrama, G.A. Rodríguez-Castro, O.
A. Gómez-Vargas, E.A. Gallardo-Hernández, I. Campos-Silva, The wear resis-
tance of boride layers measured by the four-ball test, Surf. Coat. Technol. 215
(2013) 241–246.

You might also like