Você está na página 1de 11

Surface & Coatings Technology 260 (2014) 369379

Contents lists available at ScienceDirect

Surface & Coatings Technology


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

Inuence of substrate pre-treatments on residual stresses and


tribo-mechanical properties of TiAlN-based PVD coatings
Tobias Sprute a,, Wolfgang Tillmann a, Diego Grisales a, Ursula Selvadurai a, Gottfried Fischer b
a
b

Institute of Materials Engineering, Technische Universitt Dortmund, Germany


Rif e.V. - Institut fr Forschung und Transfer, Joseph-von-Fraunhofer-Str. 20, 44227 Dortmund, Germany

a r t i c l e

i n f o

Available online 30 September 2014


Keywords:
PVD
Substrate pre-treatment
TiAlN PVD
Multilayer coatings
Residual stresses
Tribo-mechanical behavior

a b s t r a c t
Residual stresses in the substrate and in the PVD coating have a signicant inuence on the coating adhesion and
lifespan of machining as well as forming tools. Therefore, the understanding and control of the system's residual
stresses will lead to a better performance of the coated components. Moreover, although investigations were
conducted in the eld of stress analysis of PVD coatings, they do not focus on interdependencies of residual
stresses in the substrate and in the coating.
In this investigation, three different metallographically prepared substrates were used. SiC grinding, diamond
grinding, and SiC grinding and plasma nitriding preparations were selected, due to the substantial differences
in their nal residual stress states. Additionally, a Ti/TiAlN multilayer coating and a reference TiAlN monolayer
were deposited on each pre-treated substrate.
Their initial and nal residual stress states were measured by means of X-ray diffraction. In addition to the
residual stress analyses, tribo-mechanical tests, such as nano-indentation, ball-on-disc, and scratch tests were
performed in order to correlate the results with these residual stress states.
2014 Elsevier B.V. All rights reserved.

1. Introduction
In order to ensure a sufcient protection for industrially used tools
against wear, a functionalization of component surfaces, PVD protective
layers are applied. Depending on the requirements of the application,
single or multilayer PVD coatings can be used for this purpose.
Ti-based ceramic coating systems, such as TiN, TiCN, or TiAlN, are widely
used in the industry due to their high wear resistance [18]. In
particular, TiAlN possesses outstanding coating properties with a
maximum hardness of about 2832 GPa [9]. In addition, the aluminum
content in the coating ensures a high temperature resistance up to
800 C [10]. The formation of a thin, dense and well adhering protective
layer of aluminum oxide, which acts as a diffusion barrier and thus minimizes diffusion-induced wear, is responsible for this behavior [9,11,12].
Due to the possibility to produce near net-shaped coatings utilizing the
PVD technique, forming tools with a very high surface quality can be
coated. Besides the mostly low surface roughness of the tools,
unfavorable residual stresses in the composite system make it more
complicated to achieve a good adhesion. The residual stresses in
the coating and in the substrate affect the adhesive and cohesive damage processes at the coating/substrate interface, which can either
promote or prevent a failure of the coating [13]. Suitable substrate
Corresponding author at: Institute of Materials Engineering, Technische Universitt
Dortmund, Leonhard-Euler-Str. 2, 44227, Dortmund, Germany.
E-mail address: tobias.sprute@tu-dortmund.de (T. Sprute).

http://dx.doi.org/10.1016/j.surfcoat.2014.08.075
0257-8972/ 2014 Elsevier B.V. All rights reserved.

pre-treatments, selective coating architectures as well as adjustments


of the deposition process enable to consciously alter the system
properties and thus increase the life of PVD coated tools.
The substrate pre-treatments are used for the modication of the
upper substrate region and include various sequential processing
steps. While mechanical pre-treatments are primarily responsible for
cleaning the substrate surface (elimination of possible oxide layers),
and adjusting the surface topography, structural changes of the
substrate material are achieved by means of thermal- or plasmachemical processes, such as plasma nitriding. Thus, grinding with geometrically undened cutting edges generally causes the induction of
compressive residual stresses in the boundary layer of the substrate material due to supercial plastic deformations. However, friction-induced
temperature increases can also promote the formation of tensile residual stresses, leading to a reduction of compressive stresses. The use of
harder grain materials can nevertheless reduce the heat generation in
the contact zone since less wear occurs in the grains [14,15]. Plasma
nitriding is a common method to harden the outer region of the
substrate, and thus to provide a sufcient supporting effect of the base
material for a subsequently deposited layer [16,17]. It has been demonstrated in many studies that the plasma nitriding process optimizes the
adhesion and signicantly improves the wear resistance as well as the
friction behavior [1825]. Furthermore, it could be proven that the increase of the compressive residual stresses in the substrate is better
for a good adhesion [26]. Selvadurai et al. found out that the residual
stresses in the base material can be changed by the subsequent

370

T. Sprute et al. / Surface & Coatings Technology 260 (2014) 369379

deposition of a PVD layer. The previous high compressive stresses are


converted into tensile residual stresses after coating with a Ti/TiAlN
multilayer system. Here, it is assumed that a shift of the compensatory
tensile residual stresses towards the coating/substrate interface occurred in order to compensate the high compressive residual stresses
in the multilayer coating system [27,28].
In addition to the pre-treatments of the substrate, the choice of the
coating design is further crucial for the performance of the coating
system. In contrast to monolayers, multilayer coating systems typically
have an increased toughness and thus extend the life span of the
compound system [29,30]. Metallic interlayers allow the energy adsorption by plastic deformation [31] and consequently promote relaxation
processes, which can degrade the high residual stresses [32,33] and
simultaneously prevent delaminations of the coating [34]. At the same
time, the boundaries of the different individual layers inhibit crack
growths and crack propagations by deecting the cracks in the layer
transitions [4,29,35]. For example Castanho and Vieira investigated the
inuence of the numbers of interlayers on the mechanical properties
and residual stresses of alternating Ti/TiAlN systems and found out
that even one interlayer of titanium can halve the residual stresses in
the coating [36]. In addition, it was shown that the adhesion of the
PVD coating is improved with an increasing Ti interlayer thickness or
a decreasing proportion of ceramic [24], while the wear has increased
as a result of the declining hardness [26,37].
In order to further investigate these correlations, TiAlN monolayer
and Ti/TiAlN multilayer coatings are deposited on different, pretreated steel substrates and examined regarding their tribomechanical behavior as well as residual stresses.
2. Material and methods
Three different pre-treatments were performed on the hot work
tool steel substrate X37CrMoV5-1, used for this research. The rst
pre-treatment consisted of a metallographic preparation of the
substrate using SiC grinding papers; the second pre-treatment was
a similar metallographic preparation, which, however, uses diamond
grinding discs instead of SiC grinding papers. Finally, a plasma nitriding process was carried out on the rst pre-treatment in order to obtain the third pre-treatment. Herein, the different pre-treatments
will be called SiC, Diamond and Nitrided, respectively. The
plasma nitriding process was carried out by an Arc PVD device
(Metaplas, Germany), in which the samples were heated to 560 C
under a bias voltage of 650 V and a controlled atmosphere with a
constant pressure of 200 Pa for 8 h. The composition of the atmosphere consisted of 75 vol.% hydrogen and 25 vol.% nitrogen held
constant during the complete nitriding process.
The dimensions of the samples were kept constant throughout the
investigation and discs with a diameter of 40 mm and a thickness of
4 mm were used.
The deposition process of the PVD coatings was carried out by an
industrial magnetron sputtering device (CemeCon MLsinox800,
Germany). The samples rotated in the center of the chamber under a constant heating power of 5 kW during the deposition. This heating output
equates a temperature of about 400 C. The inert gases argon
(295 sccm) and krypton (200 sccm) were used as plasma gases during
the coating processes. The gas ow of the reactive gas nitrogen was controlled at a constant gas pressure of 580 mPa. Moreover, for all the deposition processes, a constant bias voltage of 100 V was applied. Three
alloyed TiAl and one pure Ti targets were mounted on the PVD device
as cathodes. Furthermore, at the time of the deposition, a target power
of 9.5 kW and 4.0 kW, respectively, was applied.
Monolayers of TiAlN and multilayers of Ti/TiAlN ((50 nm + 500 nm)
5), with a total thickness of 3000 and 2750 nm respectively, were deposited on the three different substrates and their thickness was evaluated by using a scanning electron microscope (FE-JSEM 7001 JEOL, Japan).
In order to ensure a good adhesive bonding of the coating and an excellent

wear protection, the multilayer starts with a Ti-layer on the surface of the
substrate and ends with a hard TiAlN layer. Moreover, the topography and
morphology of the deposited coatings were evaluated by using a scanning
electron microscope as well. An additional EDX detector was used in order
to analyze the chemical composition of the ceramic layers.
Additionally, mechanical properties such as the coating hardness
and Young's modulus were determined by means of nanoindentation
tests (G200 Agilent Technology, USA). A depth controlled penetration
was performed at room temperature and, in order to avoid the inuence
of the substrate on the properties of the thin lm, the results were evaluated in a range from 10 to 15% of the total coating thickness, which was
in this case between 100 and 400 nm.
With the purpose of determining the tribological properties of the
coatings, ball on disc tests at room temperature were conducted using
a tribometer (high temperature tribometer CSM, Switzerland) equipped
with a WC/Co ball as counterpart to the rotating samples with the aim of
analyzing the wear coefcient and with 51CrV4 pins as a counterpart to
obtain the friction coefcients of the different systems. 51CrV4 is a steel
used for quenching and tempering according to DIN EN 10083, and it is
usually used in the automotive and mechanical engineering industry
on components formed of metal such as gear parts, pinions, and shafts.
During the experiment, which consists of 8000 rotations, no external
lubricant was used and both the normal force and linear velocity of the
rotating discs were kept constant at 5 N and 40 cm/s, respectively. The
wear coefcient was evaluated by analyzing the wear tracks with an optical 3D surface analyzer (Innite Focus Alicona, Austria) that consists of
a confocal microscope connected to an image analyzer software, and the
wear mechanisms evidenced by SEM.
The scratch tester Revetest (CSM, Switzerland) was utilized to examine the adhesion between the PVD coatings and the steel substrates
at room temperature. For this, scratch tracks were generated with a
total length of 10 mm. The force was steadily and linearly increased
from 0 to 100 N and the results were analyzed using SEM in combination with an EDX-detector (Oxford Instruments, UK).
The residual stress evaluation was executed by means of a diffractometer (Bruker Advance D8), using the sin2 method [38]. Fe-K radiation was used instead of the usually used Cu-K radiation, in order to
avoid the uorescence radiation from the substrate [39].
Before proceeding with the residual stress measurements, phase
analyses of the metallic substrate, TiAlN monolayer, and Ti/TiAlN multilayer were performed to determine the present phases and to establish
the 2 angles related to the Bragg law, which thereafter are used for the
determination of the residual stresses. Consequently, the Fe 220 peak
found at an angle of 2 equaling 145 was selected to measure the residual stresses in the substrate. This specic reection was chosen instead
of other Fe reections, due to the high value of the angle 2 which benets the sensitivity of the sin2 method, where small changes in the lattice spacing, d, result in a corresponding shift of the diffraction angle 2.
For the analysis of the residual stresses in the coating systems, the
TiAlN 220 peak was selected and Fe-K radiation was used as well.
The reections were scanned and afterwards tted by the Pearson function to determine the positions of the peak. The X-ray elastic constants
of Fe were calculated by elastic single crystal constants. For the TiAlN
coatings, the XEC were calculated after Voigt by utilizing the experimentally determined macroscopic elastic moduli of these coatings.
To analyze the residual stresses in the substrates after the deposition
of monolayer and multilayer coating systems with X-ray diffraction,
several conditions need to be considered. In homogeneous materials,
the following relationship between the beam intensities at the moment
when the X-ray enters and leaves the specimen (hereinafter denoted as
I0 and If, respectively) holds true:
I f z I0  exp  k  z:

Here, stands for the linear attenuation coefcient of the material


and z is dened as the depth. The variable k denotes a geometry factor

T. Sprute et al. / Surface & Coatings Technology 260 (2014) 369379

371

Fig. 1. Residual stress depth proles in the substrate for the different substrate pretreatments.

which can be calculated from the incidence () and exidence () angle


or from the goniometer angles , and the Bragg angle 2. In thin
multilayer coatings, combining two materials, the penetration depth is
inuenced by the attenuation coefcient of the selected materials and
the thicknesses of the individual layers. Hence, special formulas were
developed by Fischer et al. to calculate the penetration depth in the
compound systems, depending on the goniometer angles [40]. The
following formulas were used to calculate the penetration depth in
the substrate, coated with a monolayer (Eq. (2)) and with a multilayer
(Eq. (3)), respectively:



1

a 1  ta
c  k
c

1
n 
c  k






a
b
1  ta
1  tb :
c
c

Eqs. (2) and (3) stand for the phases a, b and c, respectively.
Here, n is the total number of layers of phases a and b and t stands
for the thickness of the single sub-layers. In this particular case, a
corresponds to TiAlN, b to Ti metallic interlayers, and c to the
substrate material.
Afterwards, using Eqs. (2) and (3), the penetration depths of the
X-rays, for the reection Fe 220, in the different systems used for this

research at the different angles, were calculated. Here, attenuation


coefcients of Fe-K radiation in Fe was 552.6 cm 1 [41], in Ti was
1731.9 cm1 [41] and in ceramic layer of TiAlN, determined proportionally from TiN and AlN values, was 800.084 cm1 [42].
The maximum penetration depth of the uncoated substrates was
equal to 8.64 m. After the deposition of the monolayer, this depth is
reduced to a value of 4.3 m from the interface monolayer/substrate, a
value obtained by using Eq. (2). In the end, the X-ray penetration depths
for the multilayer/substrate systems were also computed by applying
Eq. (3). Here, the X-rays reach a maximum depth of 4.24 m, which is
a lower value when compared to the previously obtained range of the
monolayer/substrate compound, due to the high attenuation coefcient
of Ti metallic interlayers. These X-ray penetration depths can be seen in
Fig. 1. Moreover, the residual stress depth proles were obtained by
electrolytic polishing of the surface of the substrate (LectroPol-5 Struers,
Denmark) and measuring the electro-polished depths with an optical
3D surface analyzer (Innite Focus Alicona, Austria).
3. Results and discussion
3.1. Substrate pre-treatments and their inuence on residual stresses
The aim of this work is to study the inuence of different substrate
pre-treatments on the tribo-mechanical coating properties as well as
on the residual stresses in composite systems. This includes the

Fig. 2. Coating thickness from TiAlN monolayer and Ti/TiAlN multilayer systems.

372

T. Sprute et al. / Surface & Coatings Technology 260 (2014) 369379

Fig. 3. a. SEM images of the TiAlN monolayer system deposited on the different pretreated substrates. b. SEM images of the Ti/TiAlN multilayer system deposited on the different pretreated
substrates.

induction of different residual stress states in the substrate as well. In


this case, it has to be taken into account that conventional pretreatments are used, which provide reproducible results as well as comparable surface qualities and roughnesses. Therefore, two different
grinding processes and a combination of grinding and nitriding were selected. The different abrasive materials cause varying degrees of mechanical machining or deformations in the surface of the substrate on
the one hand, while generating an uneven heat input on the other
hand. This results in two different residual stresses in the substrate.
Further pre-treatment includes an additional nitriding process, which
is responsible for the hardening of the substrate and for the extremely
high compressive stresses in the upper substrate region. The different
residual stress states are shown in Fig. 1 as residual stress depth proles.
Compared to the diamond grinded substrates, the SiC grinded

substrates exhibit signicantly lower compressive residual stresses. In


addition, the inuence zone of the machining process is not as deep as
in the diamond prepared substrates. This can be mainly explained by
the choice of the abrasive material. In general, the mechanical machining of the fringe results in the formation of compressive residual stresses
due to supercial plastic deformations. However, friction-induced temperature increases can lead to the reduction of compressive stresses.
Softer grain materials, such as SiC abrasive grains, wear out faster
(blunting effect) and thus induce more heat into the contact zone
while harder grain materials (diamond grains) lead to a deeper mechanical deformation and consequently induce compressive residual
stresses [14]. During the nitriding process, the crystal lattice is extremely distorted and clamped in the diffusion zone. Finely dispersed nitrides
and dissolved nitrogen atoms in the iron lattice cause compressive

Fig. 4. Arithmetic average roughness Ra of the substrates before coating deposition and after deposition of both mono- and multilayer.

T. Sprute et al. / Surface & Coatings Technology 260 (2014) 369379

373

Fig. 5. Topography SEM pictures of the mono- and multilayer deposited on different pretreated substrates.

residual stresses in the substrate material [13,17,43,44]. The area of


inuence (diffusion zone), which is also visible in Fig. 1, is much deeper
compared to the grinded substrates. Compensating tensile residual
stresses can only be observed beginning at a depth of approximately
100 m.
3.2. Coating thickness and structure
Measured thicknesses of all the different systems as well as the
deviations of the desired coating thicknesses are shown in Fig. 2. It is
clearly visible that the compound systems, except from the monolayer
deposited on the SiC grinded substrate, achieve the planned lm thicknesses. Nevertheless, the outlying thickness value obtained for the
monolayer system deposited onto SiC is denitely in the range of tolerance accepted, using an industrial magnetron sputtering device.
In order to analyze the rst layer/substrate bonding as well as the
coating structure, fracture patterns were evaluated with a scanning
electron microscope. This procedure further allows the detection of
Table 1
Chemical composition of the TiAlN ceramic layer by mean of an EDX analysis.
Element

at.-% with N

at.-% without N

Ti
Al
N

19.98 0.43
25.18 0.35
54.84 0.78

43.66 0.21
56.34 0.21

possible defects in the coating. The facture patterns of the monolayer


and of the multilayer are shown in Fig. 3a and b. In both gures, the
left columns display the coatings in the secondary electron mode in
order to examine structural and morphological differences, whereas
the right columns show fractured surfaces in the backscattered electron
mode to show different material contrasts. All TiAlN monolayer systems
in Fig. 3 show the same structure due to the same parameter during the
deposition processes. The coatings are very dense and free of defects.
Furthermore, there are no spallations or delaminations recognizable at
the interfaces, which probably indicate a remarkable adhesion of the
coatings. The morphologies exhibit a predominantly glass-like structure
with a very slight orientation in the growth direction. However, a clear
columnar structure cannot be observed.
The coating design as well as the coating structure of the multilayer
systems can be seen in Fig. 3b. Especially the images with different material contrasts clearly visualize the alternating metallic and ceramic
layers. The thin Ti interlayers have constant thicknesses within the
complete coating system. In the structural images, left column, the
metallic interlayers are only shown as breaks of the TiAlN layers,
which are probably responsible for the inhibition of crack growth and
propagations [24,45,46]. In addition, all multilayer systems also show
no imperfections in the layer or spallations and delaminations at the
interfaces. Although all coatings have very dense structures, supposed
differences in terms of orientations are identied. The multilayer deposited onto SiC grinded and nitrided substrates shows a randomly
oriented structure like the monolayer coatings, while the multilayer

374

T. Sprute et al. / Surface & Coatings Technology 260 (2014) 369379

Fig. 6. X-ray phase diffractogram of the Ti/TiAlN multilayer system, Fe-K radiation.

deposited on the diamond grinded substrate presents a slightly columnar structure. However, the selected process parameters were constant
for all systems.
In addition, the comparable structures and morphologies of all
deposited coatings can also be attributed to the similar nucleation on
the surface of the substrate. It can be assumed that the surface roughnesses of the differently treated substrates prior to the deposition are
responsible for this behavior [4749]. The values are in the range of
Ra = 101 nm and Ra = 149 nm (Fig. 4). Due to the slight roughening
during the nitriding process by sputtering the substrate material, the
nitrided substrate has the highest roughness with Ra = 149 nm [43].
The SiC and diamond grinded substrates possess lower roughness
values of Ra = 116 nm and Ra = 101 nm, respectively. The trend observed for the different roughness values is constant both before and
after the deposition of the coatings. Thus, it can be noted that the roughness of the substrates before the deposition affects the roughness of the
different coatings. However, in general, the roughness values increase
after the coating deposition of mono- and multilayer systems due to
the sum of the substrate roughness and the intrinsic roughness of the
coating itself [47].
The initial roughness of the surface of the substrates is not only
reected in the roughness values after the coating application, but also
in the topography images of the systems (see Fig. 5). Here, the same
trend as in the roughness values can be observed. The coated surface
of the nitrided base material has a coarser texture than that of the
grinded and coated samples.
The chemical composition of TiAlN was determined by an EDX analysis. The chemical composition of the monolayer and the ceramic layers
of the multilayer system are shown in Table 1. The alternating Ti layers
used in the multilayers consist of pure Ti. Due to the fact that light

elements are difcult to determine quantitatively by EDX, the composition is examined with and without N. According to the literature,
sputtered coatings with an Al/Ti ratio of about 3/2, generally exhibit
fcc NaCl-type crystallographic structure with excellent tribomechanical properties [9,50]. Nonetheless, different investigations as
the ones performed by Makino [51] and Greczynski [52] should be
considered, since both authors have attributed the obtaining of cubic
or hexagonal phases to the use of different deposition techniques,
instead of different Al contents in the TiAlN structure.
3.3. Development of residual stresses
Residual stress depth proles for each substrate pre-treatment were
developed in order to determine the residual stress state at different
depths from the surface of the substrate and to evaluate their gradient,
Fig. 1. Hereby, the substrate was electrolytically polished one step after
the other to a depth of approximately 120 m, and residual stress evaluations were conducted for every step. In the detailed graph, shown in
Fig. 1, it can be observed that the residual stress states for the three different substrate pre-treatments present a very low gradient in depths
between 0 and 10 m. The penetration depth of the X-rays during the
different residual stress measurements in the substrate of these samples
reaches up to 10 m, being located within the above described region.
The residual stress measurements of the three different pre-treated
substrates before and after deposition, in this case, do not depend on
the deposited coating and thus are comparable.
Moreover, in order to measure the residual stresses in the coatings,
phase analyses were performed on the mono- and multilayers to dene
the most suitable reection. Fig. 6 shows a phase diffractogram of a Ti/
TiAlN multilayer system, deposited onto the substrate which was

Fig. 7. Residual stress evolution for the different substrates before and after deposition of TiAlN monolayer, Fe 220 reection.

T. Sprute et al. / Surface & Coatings Technology 260 (2014) 369379

375

Fig. 8. Residual stress evolution for the different substrates before and after deposition of Ti/TiAlN multilayer, Fe 220 reection.

metallographically prepared with SiC grinding discs. The reection


TiAlN 220 at the 2 angle equal to 82, showing a sufciently high intensity and no overlapping, was chosen for the stress evaluation.
Figs. 7 and 8 show the evolution of the residual stresses in the
substrate for each substrate pre-treatment. The residual stresses were
measured prior to the metallographic preparation (here named
unprepared), after each preparation (identied with the name of each
pre-treatment), and after the deposition of the mono- or multilayer
coatings.
Before the substrate pre-treatments, the anisotropic character of the
residual stresses in the substrate is evident due to the difference
between the values of the stress tensor 11 and 22. Thus, 11 presents
compressive residual stresses due to the transversal contraction derived
from the preferential cutting direction of the samples, while 22 exhibits
an absence of stresses. Moreover, after the pre-treatments conducted on
the different metallic substrates, no big differences are found between
11 and 22, corroborating the isotropic behavior of the resulting
residual stresses after each preparation. Subsequently, three different
residual stress states were identied: SiC with low compressive residual
stresses, diamond with slightly higher compressive residual stresses, and
lastly, nitrided samples presenting the highest value of compressive residual stresses. These results were correlated to each other and are
shown in ascending order, from left to right in Figs. 7 and 8.
In contrast to the uncoated substrates, the residual stresses in all the
substrates after depositing the different coating systems are reduced,
this behavior was also found by Tnshoff et al. [53]. Nitrided substrate
samples show the greatest reduction of residual stresses after the

deposition of the coatings whereas diamond and SiC grinded substrates


show the smallest reduction which is comparable with the results obtained by Selvadurai et al. [28]. Moreover, a slightly greater reduction
can be evidenced in the substrates deposited with TiAlN monolayer systems than for those deposited with Ti/TiAlN multilayer systems. This reduction cannot be only attributed to a single factor, as different
parameters affect the induction of residual stresses in the compound
systems, coating/substrate. One of these reasons is the temperature of
approximately 400 C, at which the deposition process is carried out approximately, and which can cause a relaxation of residual stresses in the
substrate [13]. Other reasons which are supposed to affect residual
stresses are for instance the different thicknesses between the monoand multilayers [54], and the lattice mist between the substrate and
the growing coating [13].
Because of the previously described randomly orientated structure
in the deposited coatings, a strong linear regression is obtained between
sin2 the lattice spacing d which is the base for the residual stress calculations in the coatings [38]. In Fig. 9, the inuence of the substrate preparations on the residual stresses of the coatings is evident. Hence, the
same coating material, which is supposed to have the same intrinsic residual stresses, presents higher residual stresses in those samples in
which the substrate shows higher residual stresses, caused by the pretreatment of the substrate. At the same time, clarifying that both TiAlN
and Ti/TiAlN deposited lms follow the trend described above, higher
compressive stresses were found in the multilayer systems deposited
on substrates with the same pre-treatment. This behavior was also evidenced by Selvadurai et al. [28]. The presence of higher compressive

Fig. 9. Residual stresses in the coatings, TiAlN 220.

376

T. Sprute et al. / Surface & Coatings Technology 260 (2014) 369379

Fig. 10. Mechanical properties of the TiAlN monolayer and Ti/TiAlN multilayer deposited on different pretreated substrates.

residual stresses in the multilayer systems can also be attributed to the


presence of thermal stresses, which can be found in PVD coatings due to
temperature changes during the deposition, and differences on the
thermal expansion coefcients of the coating system constituents [13,
55]. For instance, the thermal expansion coefcient for X37CrMoV5-1
steel is equal to 13 10 6 C 1 [56], for Ti interlayers it is 8.5
106 C 1 [57], and 7.37.5 10 6 C 1 for TiAlN ceramic layers
[5860]. Compressive residual stresses are developed in the material
with the lower thermal expansion coefcient [13]. This rule applies for
metallic Ti interlayers and ceramic TiAlN layers, where titanium exhibits
a higher thermal expansion coefcient, increasing the compressive residual stresses in TiAlN layers of the multilayer coatings. Additionally,
as the multilayer systems present more dissimilar material interfaces
(substrate/Ti and Ti/TiAlN), the dislocation density is increased in
these areas as well, causing an increase of the compressive residual
stresses at the proximity to the interfaces [13].
Finally, analyzing Figs. 7, 8, and 9 as a single entity, the following hypothesis can be assumed: The compressive residual stresses in the outer
regions have to be compensated by tensile residual stresses in the material core (substrate or substrate/coating system). After the deposition of
a PVD coating, a displacement of the tensile residual stress region to a
position closer to the interface substrate/coating occurs, causing the
translation of the stresses between the coating and the substrate.
The same statement has been made by Denkena and Breidenstein
[61], Tnshoff [53] and Bunshah [62], who have conrmed the relief of
compressive stresses in the substrate in order to balance the residual
stress state in the compound coating/substrate.

3.4. Mechanical properties


Hardness and Young's modulus values are presented in Fig. 10. As
described in Material and methods section, the nanoindentation experiments were performed with a Berkovich tip; and the data were analyzed at a depth of 1015% of the overall coating thickness.
Consequently, the hardness and Young's modulus values are constant
for all TiAlN monolayers and thus for all Ti/TiAlN coatings deposited
onto the three different pre-treated substrates. It has been previously
reported by different researches that TiAlN monolayers, deposited by
means of magnetron sputtering, present a hardness value around
30 GPa, which is comparable to the values obtained for this paper [63,
64]. According to Hrling et al., Ti1 xAlxN monolayers with x = 0.66,
have as a consequence of the grain size (HallPetch relation),
deposition-induced point defects, and precipitation hardening from
compositional inhomogeneities [65] a hardness equaling 33.1 GPa.
However, higher values for both hardness and Young's modulus are
found in the monolayer coatings when compared with the multilayer
systems for the same substrate pre-treatment. This behavior can be explained by the high metallic content of Ti in the multilayer bi-layer system, in this case 10% of the adjacent TiAlN layer. Moreover, the total
amount of hard ceramic material is also reduced in the multilayer system in relation to the TiAlN monolayer. In further investigations, the inuence of the Ti interlayer thickness on the hardness has to be
evaluated.
Additionally, the relation H/E is shown in Fig. 11. This relation represents the plasticity index or elastic strain to failure and it is a suitable

Fig. 11. H/E ratio of the TiAlN monolayer and Ti/TiAlN multilayer deposited on different pretreated substrates.

T. Sprute et al. / Surface & Coatings Technology 260 (2014) 369379

377

Fig. 12. Critical load of the TiAlN monolayer and Ti/TiAlN multilayer deposited on different pretreated substrates.

parameter to predict the wear resistance and to explain the deformation


properties of materials by considering the elastic rebound [66]. Thus, for
the present research, high H/E values for the Ti/TiAlN multilayer are
observed on all pre-treated substrates, which guarantee a higher toughness of the system and might be an indication for a better tribological
protection.
Critical load is an evidence of a better adhesion of the coatings to the
substrate [66]. Fig. 12 shows the critical loads for the investigated systems. Here, Lc3 represents the load value where the coating is completely removed from the center of the scratch. SiC and diamond prepared
metallic substrates presented the lowest adhesion for both mono- and
multilayer coating systems. During the scratch test, coatings deposited
on substrates with a low hardness result in a plastic deformation of
the substrate, promoting the crack formation and fragmentation of the
coating [63]. In contrast, a higher critical load value for those coatings
deposited onto substrates with higher hardness, nitride substrates, is
evidenced (Fig. 7). Values of Lc3 equal to 48.6 and 56.1 N for TiAlN and
Ti/TiAlN, respectively, were obtained. According to the abovementioned
results, an improvement of the adhesion of the coatings is observed due
to the presence of a metallic interlayer in the Ti/TiAlN multilayers [26,
32]. The great inuence of the substrate conditions on coating scratch
behavior can be observed as well. Nonetheless, the inuence of the
roughness of the substrate on the critical load of the coatings before
deposition should be considered. For substrates with high roughness
(in this case a nitrided substrate), the contact area between the
substrate and the coating at the interface substrate/coating is larger
than for substrates with a low roughness. This implies an enhanced

mechanical and physical bonding between the coating and the substrate, subsequently increasing the critical load value [67]. However,
the various hardness results of the substrate materials exhibit greater
differences than the roughness values. Thus, a major effect on the
critical load value is rather attributed to the hardness than to the roughness differences, even though these two parameters are closely linked in
this case.
3.5. Tribological behavior
Fig. 13 shows the wear coefcient obtained for the different coating/
substrate systems evaluated within this research. Hereby, the abrasive
wear mechanism was identied as a consequence of the high hardness
of the tribological counterpart used, WC/Co. Moreover, both the TiAlN
monolayer and Ti/TiAlN multilayer have shown a reduction of the
wear coefcient with increasing hardness and compressive residual
stresses in the substrate. In addition, the wear resistance correlates
with critical load and thus adhesive bonding of the coating on the substrate material. Therefore, the highest wear values were obtained at
the coatings deposited on SiC prepared substrates and the lowest values
onto nitride substrates.
Furthermore, Ti/TiAlN coatings have shown a poor tribological
behavior during wear resistance tests, which is evident in the higher
wear coefcient than this obtained for TiAlN monolayers, deposited on
the same type of substrate. This trend might result from the Ti metallic
interlayer thickness which, despite stopping the crack propagation, further reduces the system stability, as Ti interlayers can be plastically

Fig. 13. Wear coefcient of the TiAlN monolayer and Ti/TiAlN multilayer deposited on different pretreated substrates.

378

T. Sprute et al. / Surface & Coatings Technology 260 (2014) 369379

Fig. 14. Friction coefcient of the TiAlN monolayer and Ti/TiAlN multilayer deposited on different pretreated substrates using WC/Co and 51CrV4 counterparts.

deformed [26,37]. Furthermore, it has to be mentioned that the Ti/TiAlN


multilayer is thinner (2750 nm) than the monolayer (3000 nm) and
possesses a lower amount of total ceramic material, a fact that undoubtedly inuences the total wear coefcient.
Moreover, higher residual stress values were found in Ti/TiAlN multilayer systems when compared to TiAlN monolayers, Fig. 9, promoting
partial spallations, which are reected in higher wear coefcients of
these coating systems.
Friction coefcients of the coatings were evaluated with two different material counterparts, WC/Co and 51CrV4 tool steel (Fig. 14). The
experiments carried out using 51CrV4 pins have presented a relatively
constant value for all the substrate pretreatments and no trend derived
from the substrate pre-treatment was recognizable since the external
layer of the coatings always consisted of TiAlN. Nevertheless, friction
values obtained for the Ti/TiAlN multilayer systems were slightly higher
than those for TiAlN monolayers. Due to the low hardness of the pin
counterpart, the wear mechanism for this analysis was identied as
the adhesion of the pin material on the coating, also called micro-cold
welding. Consequently, friction values obtained using 51CrV4 were
highly inuenced by the metallic material, derived from the pins, deposited onto the coating systems.
Moreover, the experiments performed using WC/Co balls showed
that substrates possessing higher compressive residual stresses also
presented lower friction values. Such a behavior, comparable to the
results of the wear test, can probably be attributed to the big difference
of the hardness between the non-nitrided substrates and the coatings
[26]. Unquestionably, high elastic and plastic deformations of the soft
substrates limit a good tribological performance of the hard TiAlN and
Ti/TiAlN coatings and assist the plowing effect on the coating systems
[63]. Finally, in the present research, no strong correlation between
the roughness and tribological behavior of the coatings can be seen.
4. Conclusions
Different substrate pre-treatments were performed and their residual stresses before and after the deposition of TiAlN monolayers and
Ti/TiAlN multilayers analyzed. This research was carried out in order
to correlate the initial and nal residual stress states with the tribomechanical properties of the deposited coatings. The main conclusions
of this investigation are:
Different residual stress states and gradients are found depending on
the different metallic substrate pretreatments.
Nitrided substrates present the highest compressive residual stresses, and their hardness results are benecial for the tribo-mechanical
behavior of the coatings.
Low residual stresses in the substrate lead to high elastic and plastic
deformations of the substrate and, consequently, have a negative

effect on the tribo-mechanical properties of the coatings.


Residual stresses in the coatings are highly inuenced by the residual stresses in the substrate before the deposition process. Therefore,
coatings deposited onto nitrided substrates present the highest
compressive residual stresses. Moreover, the highest reduction of
the residual stress values of the substrate after deposition are
found in nitrided substrates, ranging from high compressive residual
stresses to a value close to zero.
Ti/TiAlN ((50 nm + 500 nm) 5) multilayer systems, despite all
expectations, present higher residual stresses than TiAlN monolayer
systems and therefore, an inferior tribo-mechanical response than
the monolayer.
The toughness of the multilayer system is higher than the toughness
of the TiAlN monolayer according to the obtained H/E ratio.
The adhesion of the coatings is improved by increasing hardness of
the substrate which is also related to an increase of compressive residual stresses. Ti metallic interlayers of the Ti/TiAlN multilayer
system inhibit crack propagations and enhance the critical load.
Moreover, the small disparity between the residual stresses of the
substrate before deposition and the residual stresses of the coating
results in a benecial adhesion.
With the aim of continuing the residual stress investigations and in
order to nd answers to different behaviors which cannot be explained
at this point, further investigations are planned:
Fatigue tests of different substrate pre-treatments of coating systems
should be conducted.
Different coating architectures have to be investigated, among them
different are Ti thicknesses in multilayers, graded systems, nanocomposites and nanolayers, only to mention a few.
Nanoindentation with spherical indents and an analysis of the load
displacement curves should be performed in order to obtain more
information about the elasto-plastic properties of the compound
systems. Moreover, conclusions about the toughness and crack
propagation in the coating can be obtained by analyzing cross
sections through the spherical indents.
Residual stresses after the deposition and their behavior under an
induced tensile load should be studied, investigating mini tensile
specimens by means of X-ray analysis.

Acknowledgment
The authors gratefully acknowledge the DFG (German Science
Foundation) for the nancial support for this work within the projects
Ti 343/34-1 and Fi 686/8-1.

T. Sprute et al. / Surface & Coatings Technology 260 (2014) 369379

References
[1] W.D. Sproul, R. Rothstein, Thin Solid Films 126 (1985) 257263.
[2] B. Navinek, A. abkar, Vacuum 36 (1986) 111115.
[3] E. Ertrk, O. Knotek, W. Burgmer, H.-G. Prengel, H.-J. Heuvel, H.G. Dederichs, C.
Stssel, Surf. Coat. Technol. 46 (1991) 3946.
[4] O. Knotek, F. Lfer, G. Krmer, Surf. Coat. Technol. 59 (1993) 1420.
[5] S. Novak, M. Sokovi, B. Navinek, M. Komac, B. Praek, Vacuum 48 (1997) 107112.
[6] V. Deringer, H. Brndle, H. Zimmermann, Surf. Coat. Technol. 113 (1999) 286292.
[7] H.A. Jehn, Surf. Coat. Technol. 131 (2000) 433440.
[8] K.-D. Bouzakis, G. Skordaris, S. Gerardis, G. Katirtzoglou, S. Makrimallakis, M. Pappa,
E. LilI, R. M'Saoubi, Surf. Coat. Technol. 204 (2009) 10611065.
[9] S. PalDey, S. Deevi, Mater. Sci. Eng. A 342 (2003) 5879.
[10] M. Kawate, A. Kimura Hashimoto, T. Suzuki, Surf. Coat. Technol. 165 (2003)
163167.
[11] O. Knotek, T. Leyendecker, J. Solid State Chem. 70 (1987) 318322.
[12] T. Leyendecker, O. Lemmer, S. Esser, J. Ebberink, Surf. Coat. Technol. 48 (1991)
175178.
[13] G.E. Totten, Maurice A.H. Howes, T. Inoue, Handbook of Residual Stress and Deformation of Steel, ASM International, Materials Park, OH, 2002. (accessed 4 June 2014).
[14] B. Scholtes, Eigenspannungen in mechanisch randschichtverformten Werkstoffzustnden:
Ursachen, Ermittlung und Bewertung, DGM-Informationsges, Oberursel, 1991.
[15] J. Maldaner, Verbesserung des Zerspanverhaltens von Werkzeugen mit HartmetallSchneidelementen durch Variation der Schleifbearbeitung, Kassel Univ. Press, Kassel, 2008. (Online-Ausg.).
[16] B. Podgornik, J. Viintin, Vacuum 68 (2002) 3947.
[17] K. Bobzin, in: Auage (Ed.), Oberchentechnik fr den Maschinenbau, 1, WileyVCH, Weinheim, Bergstr, 2012.
[18] M. van Stappen, M. Kerkhofs, C. Quaeyhaegens, L. Stals, Surf. Coat. Technol. 62
(1993) 655661.
[19] J. Batista, C. Godoy, V. Buono, A. Matthews, Mater. Sci. Eng. A 336 (2002) 3951.
[20] O. Kayser, Vak. Forsch. Prax. (2002) 156160.
[21] P. Novk, D. Vojtch, J. erk, V. Knotek, B. Brtov, Surf. Coat. Technol. 201 (2006)
33423349.
[22] F. Quesada, A. Mario, E. Restrepo, Surf. Coat. Technol. 201 (2006) 29252929.
[23] E. De Las Heras, D.A. Egidi, P. Corengia, D. Gonzlez-Santamara, A. Garca-Luis, M.
Brizuela, G.A. Lpez, M.F. Martinez, Surf. Coat. Technol. 202 (2008) 29452954.
[24] W. Tillmann, E. Vogli, S. Momeni, Vacuum 84 (2009) 387392.
[25] W. Tillmann, E. Vogli, S. Momeni, Surf. Coat. Technol. 205 (2010) 15711577.
[26] E. Vogli, W. Tillmann, U. Selvadurai-Lassl, G. Fischer, J. Herper, Appl. Surf. Sci. 257
(2011) 85508557.
[27] R. Menig, L. Pintschovius, V. Schulze, O. Vhringer, Scr. Mater. 45 (2001) 977983.
[28] U. Selvadurai, W. Tillmann, G. Fischer, T. Sprute, MSF 768769 (2013) 264271.
[29] Z. Burghard, L. Zini, V. Srot, P. Bellina, Peter A. van Aken, J. Bill, Nano Lett. 9 (2009)
41034108.
[30] S.J. Bull, A.M. Jones, Surf. Coat. Technol. 78 (1996) 173184.
[31] K. Holmberg, A. Matthews, H. Ronkainen, Tribol. Int. 31 (1998) 107120.
[32] M. Bromark, M. Larsson, P. Hedenqvist, S. Hogmark, Surf. Coat. Technol. 90 (1997)
217223.
[33] J. Lackner, L. Major, M. Kot, Bull. Pol. Acad. Sci. Tech. Sci. 59 (2011).
[34] G.S. Kim, S.Y. Lee, J.H. Hahn, B.Y. Lee, J.G. Han, J.H. Lee, Surf. Coat. Technol. 171
(2002) 8390.
[35] H. Holleck, V. Schier, Surf. Coat. Technol. 7677 (1995) 328336.
[36] J. Castanho, M. Vieira, J. Mater. Process. Technol. 143144 (2003) 352357.
[37] K. Steinhoff, H.-J. Maier, D. Biermann, Functionally Graded Materials in Industrial
Mass Production, Verl. Wiss, Scripten, Auerbach, 2009.
[38] B. Eigenmann, E. Macherauch, Mater. Wiss. Werkstofftech. 26 (1995) 148160.

379

[39] M. Birkholz, P.F. Fewster, C. Genzel, Thin Film Analysis by X-Ray Scattering,
Wiley-VCH, Weinheim, 2006. (Chichester: John Wiley, distributor, accessed 5
June 2014).
[40] G. Fischer, U. Selvadurai, J. Nellesen, T. Sprute, W. Tillmann, J. Appl. Crystallogr. 47
(2014) 335345.
[41] V. Hauk, H. Behnken, Structural and Residual Stress Analysis by Nondestructive
Methods: Evaluation, Application, Assessment, Elsevier, Amsterdam, New York,
1997. (http%3A//www.worldcat.org/oclc/162130648, accessed 13 August 2014).
[42] C. Chantler, K. Olsen, R. Dragoset, J. Chabg, A. Kishore, S. Kotochigova, D. Zucker,
Detailed Tabulation of Atomic Form Factors, Photoelectric Absorption and Scattering
Cross Section, and Mass Attenuation Coefcients for Z = 192 from E = 110 eV to
E = 0.41.0 MeV, NIST, Physical Measurement Laboratory, 2009. (http://www.nist.
gov/pml/data/ffast/index.cfm).
[43] M. Bussmann, Beitrag zur Grenzchenkonditionierung im Plasmanitrier-Arc-PVDHybridprozess, Werkstofftechnologische Schriftenreihe Bd, 5, Verl. Praxiswissen,
Dortmund, 2000.
[44] E. Macherauch, H.-W. Zoch, in: Au (Ed.), vollst. berarb. und erw, 11, Studium,
Vieweg et Teubner, Wiesbaden, 2011.
[45] K.J. Ma, A. Bloyce, T. Bell, Surf. Coat. Technol. 7677 (1995) 297302.
[46] K. Bobzin, E. Lugscheider, C. Pinero, Mater. Wiss. Werkstofftech. 35 (2004) 851857.
[47] D.M. Mattox, Surf. Coat. Technol. 81 (1996) 816.
[48] P. Panjan, M. ekada, M. Panjan, D. Kek-Merl, Vacuum 84 (2009) 209214.
[49] J. Olofsson, J. Gerth, H. Nyberg, U. Wiklund, S. Jacobson, Wear 271 (2011)
20462057.
[50] A. Cavaleiro, J.T.M. de Hosson, Nanostructured Coatings, Nanostructure Science and
Technology, Springer, New York, 2006. ().
[51] Y. Makino, S. Miyake, Trans. JWRI 30 (2001).
[52] G. Greczynski, J. Lu, M.P. Johansson, J. Jensen, I. Petrov, J.E. Greene, L. Hultman, Surf.
Coat. Technol. 206 (2012) 42024211.
[53] H. Tnshoff, B. Karpuschewski, A. Mohlfeld, H. Seegers, Surf. Coat. Technol. 116119
(1999) 524529.
[54] H. Edongu, Dnnschichtplastizitt und Wechselwirkung von Gitterversetzungen
mit Film/Substrat Grenzche, 150, Universitt Stuttgart, Stuttgart, 2004, (accessed
5 October 2013).
[55] V. Teixeira, Thin Solid Films 392 (2001) 276281.
[56] D. Klobar, J. Tuek, B. Taljat, Mater. Sci. Eng. A 472 (2008) 198207.
[57] J. Gerth, U. Wiklund, Wear 264 (2008) 885892.
[58] A. Srivastava, V. Joshi, R. Shivpuri, R. Bhattacharya, S. Dixit, Surf. Coat. Technol.
163164 (2003) 631636.
[59] C.V. Falub, A. Karimi, M. Ante, W. Kalss, Surf. Coat. Technol. 201 (2007) 58915898.
[60] D.-Y. Wang, C.-L. Chang, K.-W. Wong, Y.-W. Li, W.-Y. Ho, Surf. Coat. Technol.
120121 (1999) 388394.
[61] B. Denkena, B. Breidenstein, MSF 524525 (2006) 607612.
[62] R.F. Bunshah, Handbook of Hard Coatings: Deposition Technologies, Properties and
Applications, Noyes Publications; William Andrew Pub., Park Ridge, N.J., Norwich,
N.Y., 2001. (http://www.worldcat.org/oclc/41090847).
[63] S. Sveen, J.M. Andersson, R. M'Saoubi, M. Olsson, Wear 308 (2013) 133141.
[64] D.M. Devia, E. Restrepo-Parra, P.J. Arango, A.P. Tschiptschin, J.M. Velez, Appl. Surf.
Sci. 257 (2011) 61816185.
[65] A. Hrling, L. Hultman, M. Odn, J. Sjln, L. Karlsson, Surf. Coat. Technol. 191 (2005)
384392.
[66] P.M. Martin, Handbook of Deposition Technologies for Films and Coatings:
Science, Applications and Technology, 3rd ed. Elsevier, Amsterdam, Boston,
2010. (accessed 2 July 2014).
[67] K.-D. Bouzakis, N. Michailidis, S. Hadjiyiannis, K. Efstathiou, E. Pavlidou, G. Erkens, S.
Rambadt, I. Wirth, Surf. Coat. Technol. 146147 (2001) 443450.

Você também pode gostar