You are on page 1of 10

Surface & Coatings Technology 284 (2015) 182–191

Contents lists available at ScienceDirect

Surface & Coatings Technology

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

Numerical evaluation of scratch tests on boride layers


A. Meneses-Amador a,⁎, L.F. Jiménez-Tinoco a, C.D. Reséndiz-Calderon a, A. Mouftiez b,
G.A. Rodríguez-Castro a, I. Campos-Silva a
a
Instituto Politécnico Nacional, Grupo Ingeniería de Superficies, SEPI-ESIME, U.P. Adolfo López Mateos, Zacatenco, México D.F., 07738, México
b
ICAM Lille, Matériaux, 6 Rue Auber, B.P. 10079, 59016 Lille, Cedex, France

a r t i c l e i n f o a b s t r a c t

Article history: An experimental and numerical study of the scratch test on FeB/Fe2B bilayers is presented. The boride layers were
Received 1 April 2015 formed at the surface of AISI 304 steels by developing the powder-pack boriding process at temperatures of
Revised 6 June 2015 1223 K for 2, 6 and 10 h of exposure times. From the set of experimental conditions of the boriding process,
Accepted in revised form 14 June 2015
scratch tests were performed with a linearly-increasing load mode of 1 to 90 N on 7 mm in length to determinate
Available online 18 August 2015
the most effective and informative testing conditions and to estimate the critical load (Lc) at which the boride
Keywords:
layer fails. The damage in the boride layer was examined by high resolution SEM. Experiments tests indicated
Scratch test that at a critical load the boride layer fails through brittle fracture. Numerical calculations considering the residual
Boriding stress field generated by the scratch load showed that at this load the tensile stresses inside the boride layer be-
Finite element method come large enough to cause brittle failure. The residual stress fields generated by the scratch load were analyzed
and related with the failure mechanisms observed by the experimental tests.
© 2015 Elsevier B.V. All rights reserved.

1. Introduction steel including carbon steel, low alloy steel, tool steel and stainless steel.
Borided steel components display excellent performance in several tri-
One of the most important functional requirements in the bological applications in mechanical engineering and automotive indus-
manufacturing process is the adhesive strength of the coating with tries. Borided steels exhibit high hardness (about 2000 HV), high wear
the substrate because the performance and life of the coating are limit- resistance, and improved corrosion resistance [8–10].
ed by the adhesion strength [1]. Scratch testing is a characteristic tech- Scratch tests have been carried out to evaluate the coating–substrate
nique for systems with hard coatings [2]. The test consists of drawing a system of boriding processes by different techniques [11–14]. The
hard stylus along the material surface under an increasing normal load, scratch test was applied to evaluate TiB2/TiB coatings on high speed
until coating failure. The induced normal and tangential forces as well as steel produced by physical vapor deposition (PVD) where experimental
the resulting scratch morphology are analyzed. The load at which at results showed that the failure mechanisms observed were spalling at
least one of certain failure modes occurs at the coating and/or substrate the scratch flank for a critical load of 35 N [15]. On the other hand in
is referred to as the critical load. Material with uniform cross-sectional [16] a scratch test was developed on XC38 borided steels which were
hardness shows a lineal relationship between the critical load of hard exposed to molten borax salts. The type of coating formed at the surface
coating obtained by the scratch test and the substrate hardness [3]. of the steel was a single (Fe2B) or duplex (FeB + Fe2B) boride layer ac-
The properties of both substrate and coating, the metallurgical bond- cording to different agents used (B4C, Al and SiC). Three different types
ing between coating and substrate, the thickness of the coating and the of damage were observed: cracks that propagated in depth along the
geometry of the indenter determine the type of failure. If during a scratch trails (Hertzian fracture), cracks that developed on the scratch
scratch test the coating is very soft compared to the substrate the coat- side and propagated away from the scratching zone, and cohesive scal-
ing is scraped off exposing the substrate and if indeed the coating is ing that appeared on the sides at the extreme end of the scratch zone
harder than the substrate the failure modes result in spallation and when the applied load becomes relatively high.
buckling, finally when both the coating and the substrate are hard the Because of the high number of parameter involved in the experi-
failure mode is observed to be chipping [4–6]. mental test, the scratch test becomes a complex system. In practice
Furthermore, among the processes for modifying the surface of me- the results of the scratch test are only used to rank different layer/sub-
chanical components is the boriding in which the surface of a specimen strate system. However the finite element method can be used to ex-
is saturated by boron [7]. The boriding can be applied to a wide range of tend the knowledge of the failure mechanisms that occurred on the
coating during the scratch test, based on the study of the influence of
⁎ Corresponding author. each parameter separately [17–20]. The aim of this study is to analyze
E-mail address: amenesesa@ipn.mx (A. Meneses-Amador). by the finite element method the failure mechanisms through the stress

http://dx.doi.org/10.1016/j.surfcoat.2015.06.088
0257-8972/© 2015 Elsevier B.V. All rights reserved.
A. Meneses-Amador et al. / Surface & Coatings Technology 284 (2015) 182–191 183

distribution in the boride layer from borided AISI 304 steels during the Table 1
scratch test. Thickness of FeB/Fe2B layer obtained from the experimental conditions of borided AISI 304
steel.

2. Experimental procedure Exposure time Thickness [μm]


[h]
FeB FeB + Fe2B
2.1. The powder-pack boriding process
2 8.6 ± 0.9 13.4 ± 0.6
6 23.2 ± 2.3 30.3 ± 1.1
The powder-pack boriding process was carried out on commercial 10 34.6 ± 1.9 44.4 ± 1.2
square samples of AISI 304 stainless steel at a temperature of 1223 K,
and 2, 6 and 10 h of exposure times. The thermochemical treatment
was developed by embedding the samples in a closed-container con- radiation of λ = 0.154 nm wavelength. The XRD pattern, depicted in
taining B4C Ekabor II powder mixture. The container was placed into a Fig. 2, confirmed that the boride layer formed with 10 h of exposure
furnace in the absence of inert gases. After the furnace process, the con- was composed of a FeB/Fe2B bilayer.
tainer was removed from the furnace, and slowly cooled to room Because the thickness of the FeB phase is around 35 μm, the peaks
temperature. determined of the Fe2B phase in the diffractogram correspond to crys-
Metallographic sections were prepared to observe morphological tals dissolved in the FeB, since the FeB phase grows from the transfor-
details by the use of optical microscopy with a GX51 Olympus instru- mation of the Fe2B. Even this coating is defined as a FeB-base phase
ment. Fifty measurements were performed from a fixed reference on since it generally contains high boron products in the outermost part
different sections of the borided samples to determine the thickness of [21].
the boride layer. The layer thickness measurements were obtained by
selecting the needles that reached the deepest into the substrate.
Fig. 1 depicts the growth of the FeB and Fe2B layers at the surface of 2.2. Instrumented indentation testing
the borided AISI 304 steel for a treatment temperature of 1223 K. The
thickness of the boride layer for each treatment condition is shown in The borided AISI 304 steels were evaluated using a commercial
Table 1. nanoindenter (TTX-NHT, CSM Instruments) with a Berkovich diamond
The presence of borides on the surface of the borided AISI 304 steel tip. The nanoindentations were performed with a constant applied
was verified by XRD technique (GBC MMA instrument) using CuKα load of 100 mN along the depth of the boride layer (FeB + Fe2B) and

Fig. 1. Cross-sectional views of the borided AISI 304 steel for a boriding temperature of 1223 K and various exposure times: (a) 2 h, (b) 6 h and (c) 10 h.
184 A. Meneses-Amador et al. / Surface & Coatings Technology 284 (2015) 182–191

Fig. 2. XRD pattern obtained on the surface of the borided AISI 304 steel. The exposure Fig. 4. Behavior of the Young's modulus along the depth of the boride layer.
time was 10 h.

the maximum depth (hr/hmax) by the Eq. (3) where C = 0.9358,


diffusion zone according to the ISO 14577-1-2002 standard. The load- D = −1.6781 and E = 0.9931 [24]
displacement curves were analyzed by the Oliver–Pharr procedure [22].
   2
In addition, to avoid the influence of the free surface, the first inden- hr hr
n¼CþD þE : ð3Þ
tation was performed to 5 μm of the surface-region of the borided steel hmax hmax
(Fig. 3). The Young's modulus and hardness values were obtained at
each distance from the surface of the borided steels (Figs. 4 and 5), The value of the strain hardening coefficient was 0.23 ± 0.005 for
whereas the Poisson ratio of the boride layer was assumed of 0.25. the three treatment conditions.
The hardness values (H) were estimated as follows:
2.3. Scratch test
Pmax Pmax
H¼ ¼ 2
ð1Þ
Ac 24:5hc Scratch tests were performed on the layer/substrate system using a
CSM Revetest Xpress Instruments. A Rockwell C-type conical indenter
where Ac is the projected area of the Berkovich indenter and hc is the with 200 μm tip radius was used on the polished top surface of samples.
contact depth. The test was carried out operating in a progressive mode applying 1 N
A bilinear elastic–plastic material model was used to model the con- preload, increasing to 90 N. The average scratch length was measured
stitutive behavior of the boride layer and substrate. Therefore the yield to be 7 mm in all cases whereas the scratch speed was of 1.42 mm/min.
strength (σy) was estimated along the depth of the boride layer based The scratch process was developed as follows: The indenter penetrated
on Johnson cavity model by the Eq. (2) [23] (Fig. 6). the surface with a very small normal load and record the starting surface
  profile after normal and friction forces were recorded during the devel-
pm E opment of scratch and finally the indenter was removed. For each sample
¼ A þ Bln k εrep ð2Þ
σy σy three scratches were carried out. Obtained scratch tracks were examined
with optical and microscopy and SEM (Fig. 7). The mean values of critical
where E is the Young's modulus, εrep is the representative strain, pm is load and the failure mechanism obtained for each set of experimental
the mean pressure and the constant values are A = 4/3, B = 2/3 and conditions are summarized in Table 2.
k = 5/3.
The strain hardening coefficient (n) of the boride layer was deter- 3. Numerical analysis
mined in terms of the ratio of the depth of the residual impression to
3.1. Failure criteria

The failure modes in the scratch tests of hard layers can be classified
into three categories: Through — thickness cracking (cracks behind the
indenter, cracks as the coating is bent into the scratch track and Hertzian

Fig. 3. Berkovich indentations performed across the boride layer with a constant load of
100 mN. The boriding temperature was 1223 K for 10 h of exposure. Fig. 5. Behavior of the hardness along the depth of the boride layer.
A. Meneses-Amador et al. / Surface & Coatings Technology 284 (2015) 182–191 185

Table 2
Failure mode results for the progressive load scratch tests.

Exposure time [h] Critical load [N] Failure modes

2 Lc1 8.6 ± 0.3 Hertz cracking


Lc2 35.7 ± 2.5 Chipping
6 Lc1 17.3 ± 1.1 Hertz cracking
Lc2 43.4 ± 3.6 Chipping
10 Lc1 27.6 ± 1.0 Spalling

3.2. Finite element method

The finite element method (FEM) was used to evaluate the failure
mechanics of the boride layer from borided AISI 304 steels during the
Fig. 6. Behavior of the yield strength along the depth of the boride layer. scratch test (Fig. 8). The commercial ABAQUS 6.13 program was
employed to develop the numerical simulation of the scratch test on bo-
ride layers. The material properties of the FeB and Fe2B layers were
cracks), spallation (compressive spallation ahead of the indenter, buck- characterized by a bilinear elastic–plastic model with isotropic strain
ling spallation ahead of the indenter or elastic recovery induced spall- hardening. In addition the thermal residual stresses occurred in the bo-
ation behind the indenter) and chipping. The von Mises yield criteria ride layer were obtained by the Eq. (5) and considered in the numerical
was used to estimate the initiation of the yield which stated that yield- model.
ing begins when the distortional strain-energy density at a point equals
the distortional strain-energy density at yield in uniaxial tension or El
σ th ¼ ðα −αs ÞðTb −To Þ ð5Þ
compression. During the plastic deformation, the von Mises stress 1−vl l
(σM), is given by

where El is the Young's modulus of the layer, α is the thermal expansion


σ M ¼ σ y þ Eh εp ð4Þ
coefficient (αFeB = 11.5 × 10−6 C−1and α Fe2 B ¼ 8:7  10−6 C−1 )
[25], Tb is the temperature at the beginning of the stress develop-
where σy is the yield strength, Eh is the plastic hardening modulus ment, T o is the ambient temperature and v l is the Poisson ratio of
(Eh = nE) and εp is the plastic strain. the layer.

Fig. 7. SEM micrographs of scratch tracks. Exposure times: a) 2 h – Lc1, b) 6 h – Lc1, c) 10 h –Lc1 and d) 2 h – Lc2.
186 A. Meneses-Amador et al. / Surface & Coatings Technology 284 (2015) 182–191

Fig. 8. Numerical model of the scratch test on the borided AISI 304 steel. Fig. 9. Von Mises stress during the scratch test on the borided AISI 304 steel (2 h of treatment
time).

Several assumptions were taken into account in the numerical


simulations:

a) The Rockwell C indenter was considered as a rigid body with radius


of the spherical tip of 200 μm.
b) The sample was modeled with dimensions: 2 mm × 1 mm × 10 mm.
c) The morphology of the boride layer was flat. The thickness of the bo-
ride layers was varied according to the obtained results by the treat-
ment conditions. The bi-layer system was rigidly attached to the
substrate.
d) The layer-substrate system was modeled from a deformable solid. The
phases FeB, Fe2B and substrate were defined through sections by
assigning the characteristic properties of each zone.
e) For the sample with 10 h of exposure time, a diffusion zone was
modeled with mechanical properties of E = 240GPa, σy = 2.6GPa
and n = 0.23. However, the difference between the stress values
was less that 5%. Therefore, the diffusion zone was not considered in
the numerical model to reduce the process time of the simulation.
f) The model was meshed with C3D8R elements, where a fine mesh was
applied in the vicinity of the contact zone between the indenter and
the deformable body, and a gradually coarse mesh was used away
from the contact zone to evaluate the larger stress and strain gradients
with a high degree of accuracy, as its mesh coverage was able to obtain
319,200 elements.
g) The plastic part of the deformation is based on the von Mises yield
criterion and flow rule.
h) The bottom surface of the deformable body is constrained in all
directions.
i) The scratch test parameters were: sliding distance 7 mm, load
increases linearly from 1 N preload before sliding starts to 90 N at
7 mm sliding distance.
j) The sliding velocity is not included in the model (the model is time
independent).
k) The contact between the indenter and the mesh is frictionless.

4. Results and discussion

4.1. Overall friction coefficient, tangential force and scratch depth

The simulations of the scratch test were performed for each experi-
mental condition of the borided AISI 304 steels (Fig. 9). The scratch
depth is plotted as a function of the scratch distance during the load
increment and after removing the indenter. Fig. 10 shows the elasto-
plastic deformation of the boride layer formed for 2, 6 and 10 h of expo-
sure times during the scratch test. It is observed that the scratch depth
experimental presents variation during the test while the numerical
depth is linear, this is because the numerical model is considered with
incremental linear displacement. Likewise at the beginning, the varia- Fig. 10. Scratch depth during the scratch test for different treatment times: a) 2 h, b) 6 h,
tion of the scratch depth is greater, this is because the application of and c) 10 h.
A. Meneses-Amador et al. / Surface & Coatings Technology 284 (2015) 182–191 187

the load on the numerical model was by controlled displacement. How- The numerical model was considered frictionless because the
ever at the end of the test, the numerical and experimental depths pres- friction between indenter and surface has a considerable effect
ent excellent approximation for the scratch depth as the residual depth. only at the beginning of the scratch test when the indenter has not
The scratch depth was slightly higher for the thickness produced for 2 h penetrated the coating. This consideration mainly reduces both the
of treatment time (≈31 μm) and was decreasing with increasing expo- magnitude of the overall friction coefficient and the tangential
sure time which may be due to the influence of the residual thermal force in the numerical model. However, this simplification does not
stresses generated in the boride layer. For all three cases in the begin- significantly affect the results according to experimental test as
ning, the experimental scratch depth does not change significantly for seen in Figs. 11 and 12.
a distance of scratch that is in function with the layer thickness. Example The tangential force and the overall friction coefficient decreased as
for the model of 6 h of treatment time the scratch depth does not change increasing the treatment time which may be due to a shorter exposure
from the beginning to a scratch distance of approximately 1.5 mm time suggesting that for the condition of 2 h the scratch depth is greater
where the applied load exceeds the cohesive strength of the boride and therefore the resistance of the layer to be plowed is higher (Figs. 11
layer. Immediately the scratch depth presents an increase accompanied and 12). In addition the critical loads can be identified during the scratch
for a pronounced fluctuation that can be attributed to the fragility of the distance with the fluctuation of these parameters.
FeB layer, since when it reaches a scratch depth of ≈23 μm (thickness of On the other hand, the boride layer formed for 10 h of exposure time
FeB layer); at a scratch distance of ≈5 mm the scratch depth decreases exhibits the best load carrying capacity since it requires a higher maxi-
due to the compressive residual stress which is a characteristic of the mum normal load before the indenter/layer system starts to operate in
Fe2B layer then increases almost linearly indicating that the indenter the plowing mode (Fig. 10).
is in the Fe2B layer which presents a lower brittleness. Finally at a
scratch distance of ≈6.8 mm the scratch depth increases steeply sug-
gesting that the indenter is in the substrate. 4.2. Stress field at the layer-indenter contact surface
The overall friction coefficient μ in scratch test is the ratio of the max-
imum tangential force to the maximum normal force (FT/FN) and con- The stress distribution at the layer-indenter interface during the
sists of two parts: metallurgical bonding term (μa) and plowing term scratch test was represented by means of the maximum principal stress.
(μp) Eq. (6). The μa might be a function of the asperity condition and Fig. 13 shows the static mode in which only a normal load is applied by
the relative velocity between the contact surfaces while the μp reflects the indenter on the surface. The maximum principal stress is tensile in
the resistance as the indenter plows into the material. all three cases and it is located on the surface of the FeB layer outside
the contact circle. Although this stress is responsible for the formation
FT of Hertzian cone cracks, as the applied load in this step of the scratch
μ¼ ¼ μa þ μp: ð6Þ
FN test is low (≈ 1 N), the maximum principal stress is not sufficient to

Fig. 11. Tangential force during the scratch test for different treatment times: a) 2 h, b) 6 h, and c) 10 h.
188 A. Meneses-Amador et al. / Surface & Coatings Technology 284 (2015) 182–191

4.3. Stress field in the boride layer in the scratching direction

Fig. 15 shows the distribution of the tangential stress σxx in the mid-
dle plane of the thickness of the boride layer. The distribution of the σxx
is recorded since the beginning of the test until the location of the crit-
ical loads. It is important to aim that after the first failure mechanism, an
amount of energy is dissipated for the formation of cracks, so that the
stress magnitudes will vary in relation to the stress estimated by
means of a numerical model in which the dissipation of energy is not
considered. For all three cases, the stress pattern in the location of
critical loads is similar.

4.3.1. First critical load


For the first critical load, it is observed that a compressive stress field
ahead of the indenter which increases as the scratch distance is longer,
this compressive stress is followed by tensile stress as a result of the plas-
tic deformation of the FeB layer caused for the penetration of the indent-
er into the layer which is independent of the thickness of the layer. The
stress generated behind the indenter is a tensile stress and it is higher for
the smallest thickness which induces the formation of cracks in the bo-
ride layer for the lowest exposure times. It can be observed that for the
greatest thickness, the first failure mechanics was localized to a scratch
distance of ≈2.1 mm where the load capacity of the boride layer has
been exceeded. The spalling of the boride layer for 10 h of exposure
time might be caused for the amount of energy accumulated in the coat-
ing since there is no energy dissipation as a result of cracking in the layer.

4.3.2. Second critical load


It is observed that for the conditions of 2 h of exposure time, the
chipping of the boride layer occurred at a scratch distance of approxi-
mately 2.8 mm (distance larger than used for the spalling of the thick-
ness formed for 10 h of exposure time) where the scratch depth
reaches the substrate. For the boride layer formed with 6 h of exposure
time, a scratch distance of approximately 3.5 mm is required to cause
the chipping of the layer, which can be the result of the stress developed
from the interaction between the FeB and Fe2B layers.

4.4. Stress field in the boride layer in the direction perpendicular to the
scratching direction

During the scratch test, the boride layer is subjected to a bending de-
formation perpendicular to the scratch direction (stress component in
σzz and σyz) as consequence of the displacement of the layer from the
bottom of the scratch groove to the side of the groove indenter. Fig. 16
shows the σzz for the three layer thicknesses in the critical load. The
magnitude of the stresses for the condition of 2 h and 6 h is higher
Fig. 12. Coefficient of friction during the scratch test for different treatment times: a) 2 h,
b) 6 h, and c) 10 h. than the development for 10 h which explains the cohesive failure for
the thicknesses smallest before the adhesive failure. While for the thick-
ness greater (10 h of exposure time) the compressive stress σxx in front
of the indenter is higher than stress σzz therefore it is more likely that
cause cracking in the boride layer. Likewise stress field beneath the there is an adhesive failure for this thickness.
indenter is influenced by the thickness of the boride layer. For the
model developed for 2 h (Fig. 13a) and 6 h (Fig. 13b) of exposure 4.5. Interfacial shear stress
times, the substrate was reached by the stress field however as the
value was low (≈360 MPa) it did not cause plastic deformation. The distribution of the interfacial shear stresses σzx in the direction
During the scratching stage, for the numerical model with perpendicular to the scratch direction is shown in Figs. 17 and 18. The
the smallest thickness, the tension stresses behind the indenter σzx at the scratch distance of the first critical load for 2 h and 6 h of ex-
were formed at a lower scratch distance (≈ 0.7 mm). In addition posure times is shown in Fig. 17. It can be observed that the stress is
Fig. 14 shows the maximum principal stress to the same distance symmetric where the maximum stress is recorded for the smaller thick-
(≈ 1 mm) for the three different thicknesses of the boride layer. It ness (2 h of exposure time) which indicates that there is a greater influ-
can be observed that the magnitude of the maximum principal stress ence of substrate on the boride layer, however the magnitudes are low
decreases as the thickness of the boride layer increases, which (≈0.15 GPa) therefore it should not contribute to any damage between
suggests that the boride layer formed for 2 h of exposure time is the layer and the substrate. Fig. 18 shows the σzx at the scratch distance
more likely to suffer cracking on its surface at a lower load of scratch. of the second critical load for 2 h and 6 h of exposure times and the σzx at
Likewise the layer with the smallest thickness is most affected by the the distance of the first critical load for the condition of 10 h. The max-
deformation of the substrate. imum stress is observed for the condition of 2 h which decreases as the
A. Meneses-Amador et al. / Surface & Coatings Technology 284 (2015) 182–191 189

Fig. 13. Distribution of maximum principal stresses with applied normal load for different layer thicknesses: a) 13 μm, b) 30 μm, and c) 44 μm.

treatment time is greater. For the boride layer formed for 2 h and 6 h of • The boride layer formed for 10 h of exposure time exhibited the best
exposure times the interfacial shear stress σzs contributes to the failure load carrying capacity.
of the interface because the value is high (≈1 GPa). • The magnitude of the maximum principal stress on the surface
layer decreases as the thickness of the boride layer increases,
5. Conclusions indicating that the smallest thickness is more likely to suffer
cracking on its surface at a lower load of scratch (cohesive
In this work, the behavior of an AISI 304 steel submitted to thermo- failure).
chemical treatment of boriding was evaluated by the scratch test under
incremental sliding load using the finite element method. The following
The compressive stress field σxx ahead of the indenter increases
conclusions can be drawn:
when the scratch distance is longer, while the stress σxx behind the
• The tangential force and the overall friction coefficient decreased as indenter was tensile and is higher for the smallest thickness. This
coating thickness increased. suggests that for larger thickness the cracking behind the indenter
190 A. Meneses-Amador et al. / Surface & Coatings Technology 284 (2015) 182–191

Fig. 14. Distribution of maximum principal stresses with incremental sliding load for different layer thicknesses: a) 13 μm, b) 30 μm, and c) 44 μm.

did not occur, causing the failure of the layer in front of the indenter Acknowledgements
(spalling).
For the smallest thickness there is more influence of substrate This work was supported by research grants 20151312 and SIP-SNI-
according to results obtained by the interfacial shear stress σzs which 2011/15 Instituto Politécnico Nacional and CONACYT 253968 and
significantly contributes to the interfacial failure. 183836 in Mexico.

Fig. 17. Distribution of the interfacial stress σzx at the direction perpendicular to the
Fig. 15. Distribution of the stress σxx along the scratch track for different layer thicknesses scratch (cohesive failure).
according to treatment time.

Fig. 16. Distribution of the stress σzz at the direction perpendicular to the scratch for Fig. 18. Distribution of the interfacial stress σzx at the direction perpendicular to the
different layer thicknesses according to treatment time. scratch (failure of metallurgical bonding).
A. Meneses-Amador et al. / Surface & Coatings Technology 284 (2015) 182–191 191

References [13] F. Bidev, O. Baran, E. Arslan, Y. Totik, I. Efeoglu, Surf. Coat. Technol. 215 (2013) 266.
[14] M. Berger, M. Larsson, Surf. Eng. 16 (2000) 122.
[1] M. Zouari, M. Kharrat, M. Dammak, M. Barletta, Surf. Coat. Technol. 263 (2015) 27. [15] P. Kaestner, J. Olfe, K.-T. Rie, Surf. Coat. Technol. 142–144 (2001) 248.
[2] S.T. Gonczy, Int. Appl. Ceram. Technol. 2 (2005) 422. [16] O. Allaoui, N. Bouaouadja, G. Saindernan, Surf. Coat. Technol. 201 (2006) 3475.
[3] A. Rodrigo, P. Perillo, H. Ichimura, Surf. Coat. Technol. 124 (2000) 87. [17] J.L. Bucaille, E. Felder, G. Hochstetter, Wear 249 (2001) 422.
[4] P. Benjanim, C. Weaver, Proc. R. Soc. Lond. 254 (1960) 177. [18] K. Holmberg, A. Laukkanen, H. Ronkainen, K. Wallin, S. Varjus, J. Koskinen, Surf. Coat.
[5] S.J. Bull, Surf. Coat. Technol. 50 (1991) 25. Technol. 200 (2006) 3810.
[6] S.J. Bull, Mater. High Temp. 13 (1995) 169. [19] A. Laukkanen, K. Holmberg, J. Koskinen, H. Ronkainen, K. Wallin, S. Varjus, Surf. Coat.
[7] V. Sista, O. Kahvecioglu, G. Kartal, Q.Z. Zeng, J.H. Kim, O.L. Eryilmaz, A. Erdemir, Surf. Technol. 200 (2006) 3824.
Coat. Technol. 215 (2013) 452. [20] N. Aleksy, G. Kermouche, A. Vautrin, J.M. Bergheau, Int. J. Mech. Sci. 52 (2010) 455.
[8] A.K. Sinha, Boriding (boronizing), ASM Int. Handbook, vol. 4, The Materials [21] C. Martini, G. Palombarini, M. Carbucicchio, J. Mater. Sci. 9 (2004) 933.
International Society, 1991 437–447. [22] W.C. Oliver, G.M. Pharr, J. Mater. Res. 7 (1992) 1564.
[9] M. Carbucicchio, G.P. Palombarini, J. Mater. Sci. Lett. 6 (1987) 1147. [23] K.L. Jhonson, Contact Mechanics, Cambridge University Press, 1985.
[10] M.B. Mann, Wear 208 (1997) 125. [24] O. Casals, J. Alcalá, Acta Mater. 53 (2005) 3545.
[11] P. Sricharoenchai, N. Panich, P. Visuttipitukul, P. Wangyao, Mater. Trans. 51 (2010) [25] X. Cong-Xin, O. Meng-Lan, Wear 137 (1990) 151.
246.
[12] Y. Totik, E. Arslan, I. Efeoglu, I. Kaymaz, Surf. Rev. Lett. 16 (2009) 329.

You might also like