Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Applied Surface Science 399 (2017) 229–236

Contents lists available at ScienceDirect

Applied Surface Science


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

Full length article

Characterization and diffusion model for the titanium boride layers


formed on the Ti6Al4V alloy by plasma paste boriding
Mourad Keddam a,∗ , Sukru Taktak b
a
Laboratoire de Technologie des Matériaux, Faculté de Génie Mécanique et Génie des Procédés, USTHB, B.P. No. 32, 16111 El-Alia, Bab-Ezzouar, Algiers,
Algeria
b
Metallurgical and Materials Engineering, Faculty of Technology, Afyon Kocatepe University, ANS Campus, 03200, Afyonkarahisar, Turkey

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

Article history: The present study is focused on the estimation of activation energy of boron in the plasma paste borided
Received 8 February 2016 Ti6Al4V alloy, which is extensively used in technological applications, using an analytical diffusion model.
Received in revised form Titanium boride layers were successfully produced by plasma paste boriding method on the Ti6Al4V alloy
18 September 2016
in the temperature range of 973–1073 K for a treatment time ranging from 3 to 7 h. The presence of both
Accepted 29 November 2016
TiB2 top-layer and TiB whiskers sub-layer was confirmed by the XRD analysis and SEM observations. The
Available online 11 December 2016
surface hardness of the borided alloy was evaluated using Micro-Knoop indenter. The formation rates
of the TiB2 and TiB layers were found to have a parabolic character at all applied process temperatures.
Keywords:
Titanium borides A diffusion model was suggested to estimate the boron diffusivities in TiB2 and TiB layers under certain
Plasma paste boriding assumptions, by considering the effect of boride incubation times. Basing on own experimental data on
Growth kinetics boriding kinetics, the activation energies of boron in TiB2 and TiB phases were estimated as 136.24 ± 0.5
Incubation time and 63.76 ± 0.5 kJ mol−1 , respectively. Finally, the obtained values of boron activation energies for Ti6Al4V
Activation energy alloy were compared with the data available in the literature.
© 2016 Elsevier B.V. All rights reserved.

1. Introduction surface modification of Ti6Al4V alloy in order to improve tribolog-


ical properties using numerous surface methods [6–9].
Titanium and its alloys are widely used in marine, sport industry, Boriding, which is one of the surface modification methods, is a
bioengineering, aviation and space technology due to their excel- thermo chemical surface treatment process where boron atoms can
lent properties, i.e. high yield strength, high strength to weight readily diffuse into a metal substrate and form one or more hard
ratio, good corrosion resistance and exceptional biocompatibility boride layer on the metal surface [10]. Boriding of titanium and its
[1]. In particular, Ti-6Al-4V, which constitutes more than 50% of alloys has been studied in many research groups for the last cou-
all alloys in use today, is most popular ␣ + ␤ alloy. This alloy has ple of years in an effort to create hard TiB2 layer and TiB whiskers
high tensile strength, and fatigue resistance and has been widely on the surface of Ti [11,12]. The growth of the phases is controlled
employed as orthopaedic and aeronautical applications [1–3]. The by diffusion of boron. Titanium boride layers can be obtained by
addition of alloying elements to titanium enables it to have a wide boriding of titanium and its alloys through diffusion of boron at
range of properties because Al tends to stabilize the ␣-phase and 700–1100 ◦ C for various methods [12–17]. The plasma paste borid-
vanadium tends to stabilize the ␤-phase, lowering the tempera- ing process, which is the effective way to form titanium borides
ture of the transformation from ␣ to ␤.The alpha phase promotes on the Ti surfaces, presents significant benefits such as reduced
good weldability, excellent strength characteristics and oxidation boriding temperatures, eliminating toxic and explosive gases and
resistance. The ␤-phase can precipitate by an ageing heat treatment corrosion medium compared with conventional and plasma borid-
[3,4]. Ti6Al4V alloy, nevertheless, has poor wear-friction character- ing processes [17–19].
istics and shear strength, making them less desirable for tribological From a kinetic point of view, the modeling of the boriding kinet-
applications. They also tend to gall or seize in sliding contact with ics is a relevant tool used to select the adequate process parameters
itself or another metal [5]. Many studies have been undertaken on in order to get optimum boride layer thicknesses according to the
industrial applications of borided materials. A few attempts have
been reported on the modeling of diffusional growth and multi-
∗ Corresponding Author. phase diffusion of intermediate layers in binary titanium-boron
E-mail address: keddam@yahoo.fr (M. Keddam). couples in the case of pack-boriding process [20,21]. Till now, no

http://dx.doi.org/10.1016/j.apsusc.2016.11.227
0169-4332/© 2016 Elsevier B.V. All rights reserved.
230 M. Keddam, S. Taktak / Applied Surface Science 399 (2017) 229–236

Fig. 1. Schematic illustration of the practical procedure used to estimate the boride layers thicknesses.

study was especially devoted to the modeling of the growth kinetics


of titanium borides produced by the plasma paste boriding process
in the Ti6Al4V alloy. Recently, Keddam et al. [22] have suggested
a kinetic model for estimating the boron diffusivities in titanium
borides, formed by plasma paste boriding (PPB) on Cp-Ti substrate,
in the temperature range of 973–1073 K. In their study, Keddam
et al. [22] have ignored the effect of boride incubation times during
the formation of titanium borides on Cp-Ti substrate.
In the current study, a diffusion model inspired form the refer-
ence work [22] was suggested to estimate the diffusion coefficient
of boron in each titanium boride layer in the temperature range
of 973–1073 K. This model was based on solving the mass balance
equations at the two (TiB2 /TiB) and (TiB/Ti) interfaces by consid-
ering the effect of boride incubation times during the formation of
TiB2 and TiB layers.

2. Experimental procedures

2.1. Plasma paste boriding and characterization


Fig. 2. Schematic representation of boron depth-concentration through the
The substrate material, Ti6Al4V alloy (Grade 5), essentially con- (TiB2 /TiB) bilayer growing on the substrate of Ti6Al4V alloy.
tained: 6 wt.% Al, 4 wt.% V, 0.83 wt.% Mn, 0.29 wt.% Si, 0.18 wt.%
Mo, 0.05 wt.% S and 0.016 wt.% P. The samples had a cylindrical
ple with about 12 mm length. Layer thicknesses and the standard
shape and were 16 mm in diameter and 6 mm in length. Before
deviations were determined by these multiple measurements. The
plasma paste boriding treatment, the samples were ground up to
X-ray diffractograms were obtained using a copper tube source
1200 mesh emery paper and polished with 0.1 ␮m Al2 O3 pow-
in the conventional Bragg–Brentano ( − 2–2) technique hav-
der. The substrates were washed in distilled water, ultrasonically
ing symmetric geometry with monochromatized radiation (Cu Ka,
degreased in acetone and rinsed in alcohol. All the samples were
␭ = 0.15418 nm). The hardness of the layers was measured on the
covered by borax (Na2 B4 O7 ) paste with a thickness of about 1 mm.
top surfaces of the samples using Micro-Knoop indenter (Shimadzu
The plasma paste boriding process was performed in a dc plasma
HMV-2) with 25 g loads.
system. The description of dc plasma system was explained in a
previous paper [18]. The samples covered by borax paste were
placed in the vacuum container and the system was evacuated 3. Diffusion model
to 2.5 × 10−2 mbar pressure for 1 h. The process was carried out
at 973, 1023 and 1073 K for 3, 5 and 7 h in a gas mixture of The diffusion model considers the (TiB2 /TiB) bilayer grown on
70%H2 –30%Ar under a constant pressure of 5 mbar. Temperature the Ti6Al4V alloy whose substrate being saturated with boron
of the samples was measured using a chromel–alumel thermocou- atoms as displayed in Fig. 2. Cup TiB2. (=31.10 wt.%B) and C TiB2.
low
ple, placed in a hole with depth of about 0.3 mm at the bottom of the (=30.10 wt.%B) are the upper and lower boron concentrations in
treated samples. During the experiment, deviation of the tempera- TiB TiB
TiB2 while Cup (=18.50 wt.%B) and Clow (=18 wt.%B) are the upper
ture read from thermocouple was about ±5 ◦ C. The glow discharge and lower boron concentrations in TiB, respectively. These corre-
was operated with a potential difference of 400–800 V to obtain the sponding values of upper and lower limits of boron concentrations
prescribed boriding temperature. The boriding temperatures were for each titanium boride phase were taken from the references
controlled by changing the power input. [20,21]. According to Sarma et al. [20,21], the TiB layer, is not a solid
The microstructures of polished and etched cross-sections of monolith, but is made up of growing TiB whiskers and the Ti matrix
the specimens were observed by a Leo 1430 VP scanning elec- among the TiB whiskers. These authors have observed by SEM that
tron microscope (SEM) and energy dispersive X-ray spectroscopy the Ti matrix appears to be in contact with the continuous TiB2
(EDS). Fig. 1 schematically illustrates the practical procedure used layer at the surface of the borided samples. So, the assumption that
to estimate the titanium boride layer thicknesses. The thicknesses considers TiB as a layer is reasonable. It is assumed in the diffusion
of TiB2 layer and TiB whiskers formed on the Ti6Al4V alloy were model that the Ti-matrix regions located among the TiB whiskers
measured precisely at 12 different equally spaced locations of a are not considered as an additional path for the diffusion of boron
SEM micrograph. These measurements were performed on at least atoms. The term Cads represents the adsorbed boron concentration
three different micrographs throughout cross-section of the sam- or the effective boron concentration during the boriding process
M. Keddam, S. Taktak / Applied Surface Science 399 (2017) 229–236 231

[23]. The variable u the position of the (TiB2 /TiB) interface while v CTi (x(t = t) = v) = C0 (9)
represents the position of the (TiB/substrate) interface as a func-
The mass balance equations [30,31] are then described by Eqs.
tion of the treatment time. C0 represents the boron solubility in the
(10) and (11).
substrate.
TiB2 ), imposed by
 
The upper boron content in the TiB2 phase (Cup TiB
(Cup 2 − Clow2 )
TiB TiB − C TiB )
(Cup
du low
the boriding medium, gives rise to the two titanium borides TiB2 wTiB2 ( ) = DBTiB2 − DBTiB (10)
dt u (v − u)
and TiB. From a thermodynamic point of view; the both phases (TiB2
and TiB) exhibit a narrow composition range (of about 1 at.% B) as  TiB − C TiB )

stated by the authors [24,25]. In fact, a linear concentration − profile dv du (Cup
wTiB ( ) + w ( ) = DBTiB low
(11)
through the TiB2 and TiB layers was assumed due to the narrow dt dt (v − u)
homogeneity ranges in both phases. In addition, a homogeneity
with
range for TiB around 49–50 at.% B was reported in the literature by  
TiB2 TiB2 TiB2 TiB TiB
several authors [24,26–28]. Furthermore, the composition range for wTiB2 = 0.5 × (Cup − Clow ) + (Clow − Cup ) ; wTiB = 0.5 × (Cup −
the TiB2 phase was also found around 65.5–67 at.% B according to

the reference [29], 65.2–66.3 at.% B from Rudy and Windisch [24]
TiB
Clow TiB
) + (Clow − C0 ) ; w = 0.5 × Cup
TiB TiB
− Clow
and 65.5–67.6 at.% B from Thebault et al. [28].
The following assumptions are made in the diffusion model:
The TiB2 layer thickness u grows in a parabolic manner following
Eq. (12), where k1 represents the parabolic growth constant at the
– The kinetics is dominated by the diffusion-controlled mechanism
(TiB2 /TiB) interface.
– A linear concentration – profile is assumed through the TiB2 and
0.5
TiB layers u = k1 [t − t0nB2 (T )] (12)
– The growth of boride layers is a consequence of the boron diffu-
The distance, v, is the location of the (TiB/substrate) interface
sion perpendicular to the sample surface
and k2 its parabolic growth constant (Eq. (13)) and the difference
– The diffusion of boron in the Ti matrix is not considered
(l = v-u) denotes the layer thickness of TiB (Eq. (14)).
– The incubation times for titanium borides are taken into account
– TiB is a layer that is made up of the growing TiB whiskers embed- v = k2 [t − t0 (T )]0.5 (13)
ded in the Ti matrix. 0.5
– The boride layer is thin in comparison to the sample thickness l = v − u = k2 [t − t0 (T )] − k1 [t − t0TiB2 (T )]0.5 (14)
– Local equilibrium is attained at the moving interfaces with k2  k1 and t0 (T )  t0TiB2 (T )
– Planar morphology is assumed for the phase interfaces The boron diffusion coefficient in each titanium boride can be
– The volume change during the phase transformation is not con- obtained as follows:
sidered
– The diffusion coefficient of boron in each titanium boride is inde- (wTiB2 + w )k12 ˚(T ) + wTiB k1 k2
DBTiB2 = (15)
TiB TiB
pendent of concentration 2˚(T )(Cup 2 − Clow2 )
– The effect of Ti diffusion on layer growth is ignored.
– A uniform temperature is assumed throughout the sample. and
– The effect of alloying elements is not considered during the boron [(w k1 ˚(T ) + wTiB k2 ][k2 ˚(T ) − k1 ]
DBTiB = (16)
diffusion TiB − C TiB )
2˚(T )(Cup low

The initial conditions of the diffusion problem are set up as: with
       1/2
CTiB2 x(t  0) = 0 = 0; CTiB x(t  0) = 0 = 0; CTi x(t  0) = 0 =0 (1)
t − t0 (T )
˚(T ) = (17)
The boundary conditions are given by the following equations: t − t0
TiB2
(T )
 
CTiB2 x(t = t0TiB2 (T )) = 0 TiB2
= Cup for Cads  30.10wt.%B (2)
Fig. 3 describes the temperature dependence of the ˚(T ) param-
  eter. It is noticed that this parameter is linearly dependent on the
CTiB2 x(t = t0TiB2 (T )) = 0 TiB2
= Clow for Cads ≺ 30.10wt.%B and with TiB2 phase boriding temperature and can be approximated by Eq. (18) using a
linear fitting of experimental data with a correlation coefficient of
(3)
0.98:

˚(T ) = 0.00024T + 0.73348 (18)


  TiB
CTiB x(t = t0 (T )) = 0 = Cup ; for 18wt.%B ≺ Cads ≺ 30.10w For estimating the boron diffusivities in the TiB2 and TiB lay-
ers, the diffusion model uses the following parameters as input
t.%B and without TiB2 phase (4) data: (the boriding temperature, the treatment time, the upper and
lower limits of boron concentration in each titanium boride, the
experimental values of parabolic growth constants and the ˚(T )
  TiB
parameter).
CTiB x(t = t0 (T )) = 0 = Clow ; for Cads ≺ 18wt.%B an

d without TiB2 phase (5) 4. Results and discussion

4.1. Characterization
TiB2
CTiB2 (x(t = t) = u) = Clow (6)
Titanium borides were successfully formed on the Ti6Al4V alloy
TiB
CTiB (x(t = t) = u) = Cup (7) using Na2 B4 O7 paste in the H2 -Ar plasma environment. SEM micro-
graphs of the plasma paste borided Ti6Al4V alloy at temperatures
TiB
CTiB (x(t = t) = v) = Clow (8) of 973 and 1073 K for 7 h are presented in Fig. 4. The metallographic
232 M. Keddam, S. Taktak / Applied Surface Science 399 (2017) 229–236

Table 1
Surface microhardness values of plasma paste borided Ti6Al4V alloy.

Plasma paste boriding Microhardness, HK0.025


temperature, K

Plasma paste boriding time, h

3 5 7

973 2077 ± 78 2115 ± 92 2240 ± 105


1023 2104 ± 95 2134 ± 88 2295 ± 97
1073 2121 ± 108 2225 ± 96 2356 ± 115

with the increase in temperature. It has been reported that TiB2


could form first due to lower negativity of the free energy than that
of TiB. Ti and TiB2 could further react to form TiB [32]. TiB whiskers
were found to penetrate deeper at some angles into the substrate
after higher temperatures and treatment times compared to that
seen in lower temperature and treatment times. In this process,
H+ ions formed in the plasma reacted with borax to produce active
Fig. 3. Evolution of the ˚(T )parameter as a function of boriding temperature.
boron (B+ ) and then boron hydride (BxHy). The formation of atomic
boron took place due to the decomposition of the boron hydride
and it became the ionic boron, B+1 in the plasma. Diffusion of ionic
boron into interstitial sites of the titanium occurred to grow tita-
nium boride compounds in consequence of concentration gradient
of boron and plasma driving forces [17].
The distribution of Al, V and Ti elements obtained using EDS line
analysis in the Ti6Al4V alloy plasma paste borided at 1073 K for 7 h
from surface to interior is indicated in Fig. 5. It can be seen that the
main alloying elements (Al and V) of the Ti6Al4V were not detected
throughout the boride layers. After the layers, the concentration
of alloying elements was in tendency to increase over the inte-
rior. Line microanalysis results revealed that the aluminum mainly
diffused to the interior more than vanadium during the plasma
paste boriding. Fig. 6 shows the XRD patterns of Ti6Al4V samples
plasma paste borided at 973 and 1073 K for 5 h. The presence of
phases of boride layers were TiB2 and TiB with additional diffrac-
tion peaks from titanium substrates. Microstructural findings were
confirmed by XRD analysis. X-ray patterns revealed that increas-
ing the temperature caused an increase of the peaks of TiB2 phase
and reduction in the ␣-Ti peak. The XRD analysis, which is in accor-
dance with the results of previous studies [11,12,20,21], verified
that both titanium borides were formed even at low temperature
such as 973 K. Because all plasma paste boriding temperatures stud-
ied were below the ␣→␤ transformation temperature (1163 K),
the substrate still had non-transformed ␣+␤ structure. The optical
micrograph of the Ti6Al4V alloy plasma paste borided at 1073 K for
3 h are illustrated in Fig. 7. As shown from the micrograph, there
is no appearance of secondary-␣ like acicular, lamellar, platelike
or Widmanstatten structures, which are proof of heat treatment
at a temperature of above ␤ transition. The microstructure was
consisted of equiaxed ␣ (bright) with undefined boundaries and
intergranular ␤ (dark) structure.
Table 1 shows the microhardness values of plasma paste borided
Ti6Al4V alloy obtained on the top surfaces for various temperatures
and times. As can be seen in Table 1, plasma paste boriding, which
was performed at 800 ◦ C for 7 h, produced the maximum surface
hardness value of around 2356 HK0.025 when compared to value
Fig. 4. SEM micrographs of the plasma paste borided Ti6Al4V alloy at the two tem-
peratures for 7 h : (a) 973 K and (b) 1073 K. of 384HK0.1 in the as-received Ti6Al4V alloy. The surface hard-
ness value obtained at lower temperature and time (700 ◦ C −3 h)
decreased to around 2077 HK0.025 . The surface hardness values
study of the cross-section of the samples showed that two distinct are consistent with values obtained by various boriding processes
titanium borides were formed on the surface of the alloy. The com- [11,12,14,16]. On the boride surface, the decrease in the surface
pact and flat TiB2 layer was located near the surface and TiB grew in hardness with decreasing process temperature and time could be
the form of whisker in the vicinity of titanium matrix [11,20,21,32]. explained as a consequence of the substrate effect due to the fact
TiB2 layer bonds tightly with TiB whisker and it was slightly thicker that the penetration depth of the Knoop indenter was higher than
M. Keddam, S. Taktak / Applied Surface Science 399 (2017) 229–236 233

Fig. 5. EDS line analysis from surface to interiorin the Ti6Al4V alloy plasma paste borided at 1073 K for 7 h.

Fig. 6. XRD patterns of Ti6Al4V samples plasma paste borided at 973and 1073 K for 5 h.

one-tenth of the layer thickness for the low layer thickness (<4 ␮m) Table 2
Experimental values of parabolic growth constants at the (TiB2 /TiB) and
obtained at lower temperatures.
(TiB/substrate) interfaces with the corresponding boride incubation times.

Temperature (K) k1 (␮m/s0.5 ) t0TiB2 (T )(s) k2 (␮m/s0.5 ) t0 (T )(s)


4.2. Results of the diffusion model
973 0.00762 950 0.02097 1400
1023 0.01273 500 0.02730 1150
Fig. 8 describes the time dependence of the square of TiB2 and 1073 0.01988 420 0.03548 520
(TiB + TiB2 ) layers thicknesses with the occurrence of boride incu-
bation times. Table 2 gives the experimental values of parabolic
growth constants at the (TiB2 /TiB) and (TiB/substrate) interfaces shown that the higher boriding temperatures involved the shorter
along with the values of the corresponding boride incubation times. incubation times as displayed in Table 2. According to certain
The parabolic growth constants were obtained from the slopes authors [31,33–36], the boride incubation time is decreased when
of the curves giving the square of boride layer thickness versus the boriding temperature rises.
the boriding time (see Fig. 8). The boride incubation times can The estimated values of boron diffusivities (using Eqs. (15) and
be deduced from Fig. 8 for a null boride layer thickness. It was (16)) for TiB2 and TiB layers are given in Table 3 for an upper boron
234 M. Keddam, S. Taktak / Applied Surface Science 399 (2017) 229–236

Table 3
Estimated values of boron diffusivities in the TiB2 and TiB layers for an upper boron
content of 31.10 wt.% in TiB2 .

Temperature (K) DBTiB2 (m2 /s) DBTiB (m2 /s)


−15
973 1.90 × 10 5.08 × 10−15
1023 4.32 × 10−15 7.25 × 10−15
1073 9.13 × 10−15 10.18 × 10−15

content of 31.10 wt.% in TiB2 . The boron diffusivity values obtained


in the present study are consistent with the values reported in
the literature [20,21]. The temperature dependence of boron dif-
fusion coefficient in each titanium boride can be expressed by the
Arrhenius equation as follows:

QBi
DBi = D0i exp(− ) (18)
RT
where: i = (TiB2 or TiB) and D0i is a pre-exponential constant, QBi
represents the boron activation energy of the i-phase, T denotes
Fig. 7. The optical micrograph of the Ti6Al4V alloy plasma paste borided at 1073 K
the absolute temperature in Kelvin and R = 8.314 J mol −1 K −1 is the
for 3 h showing non transformed ␣ + ␤ microstructure at below the ␣ → ␤ transfor-
mation temperature. ideal gas constant.
The value of boron activation energy in each boride layer was
determined by the slopes obtained in the plot of ln DBi as a func-
tion of the reciprocal temperature. Fig. 9 shows the temperature
dependence of boron diffusion coefficients in the TiB2 and TiB layers
according to the Arrhenius relationship. The value of boron activa-
tion energy in each boride layer can be readily obtained from the
slope of the corresponding curve. As a result, the boron diffusion
coefficients in the TiB2 and TiB layers are respectively given by Eqs.
(19) and (20):

(136.24 ± 0.5) kJmol−1
DBTiB2 = 3.918 × 10−8 exp − (19)
RT

and

(63.76 ± 0.5) kJmol−1
DBTiB = 1.334 × 10−11 exp − (20)
RT

The estimated values of boron activation energies in the TiB2


and TiB phases for the plasma paste boriding process were about
136.24 ± 0.5 and 63.76 ± 0.5 kJ mol−1 , respectively. The higher
boron activation energy for the TıB2 layer is a result of the contin-
uous growth of TiB2 as a monolithic layer. However, the low value
of boron activation energy in the TiB layer could be attributed to
its needle-like morphology extending into the substrates. Fan et al.
[37] have pointed out that the estimated diffusivity of boron in TiB
whiskers along the whisker direction is about 45 times that in TiB2
in the temperature range 870–970 ◦ C due to the fact that the fastest-
growing direction of TiB whiskers was normal to planes containing
both metal and B atoms. Furthermore, it was found that the TiB
whisker was associated with the preferential diffusion of boron
through the [010] direction in the crystal lattice of titanium [11],
which indicates that the boron activation energy into interstitial
lattice sites is low in this crystallographic direction.
Table 4 lists the values of the activation energies for the growth
of boride layer for Ti and Ti6Al4V alloy [17,22,31,37,38]. It is impor-
tant to note that the activation energies were found to be dependent
on the boriding method, the chemical composition of titanium and
process temperature, as well as on the concentration gradient of
boron atoms on the surface. Most of these methods used to calcu-
late the Q values rely on the parabolic growth law that is controlled
by boron diffusion perpendicular to the titanium surface. However,
in this study, the diffusion model used to estimate the boron activa-
tion energies in the titanium borides is based on the mass balance
equations at the growing interfaces (TiB2 /TiB) and (TiB/substrate)
Fig. 8. Evolution of the square of boride layer thickness as a function of the boriding by considering the upper and lower boron concentrations deter-
time: (A) for TiB2 layer, (B) for total boride layer (TiB2 + TiB). mined in the Ti–B phase diagram. As can be seen from Table 4,
M. Keddam, S. Taktak / Applied Surface Science 399 (2017) 229–236 235

Table 4
Activation energies of boron in the titanium borides for comparing the experimental results from this work with the data from the literature.

Substrate Boriding medium Temperature range Activation energy, Q Calculation method Refs.
(K) (KJ mole−1 )

Ti6Al4V Pack boriding 1273–1373 65.23(total layer) Parabolic law [32]


Cp-Ti CRTD-Bor (electrolyse liquid boriding) 1173–1373 189.9 (TiB2 ) Parabolic law [38]
Ti6Al4V Ti6Al4V/Sigma fiber composites 1143–1243 187 (TiB2 ) 190 (TiB) Parabolic law [37]
Cp-Ti Plasma paste boriding 973–1073 93.61(total layer) Parabolic law [17]
Cp-Ti Plasma paste boriding 973–1073 136.24 (TiB2 )63.76 (TiB) Multiphase analytical diffusion model [22]
Ti6Al4V Plasma paste boriding 973–1073 99.79(total layer) Parabolic law [17]
Ti6Al4V Plasma paste boriding 973–1073 136.24 (TiB2 )63.76 (TiB) Multiphase analytical diffusion model Present study

the (␣/␤) phase transition temperature as reported by Sarma et al.


[20]. In our study, the anomalous diffusion phenomenon was not
observed when plasma paste boriding of Ti6Al4V alloy in the tem-
perature range of 973–1073 K. This can be attributed to the fact that
the mechanism of boron diffusion in Ti6Al4V alloy is different from
the pack boriding method as investigated by Sarma et al. [20]. In
comparison with the previous diffusion model suggested by Sarma
et al. [20], the effect of boride incubation times was considered
during the formulation of the present diffusion model.

5. Conclusions

In this work, the boron diffusion into Ti6Al4V alloy surface


resulted in two layer consisted of an outer monolithic TiB2 top-
layer and TiB whiskers sub-layer growing roughly normal to the
surface. The thicknesses of boride layers were found to be about in
the range of 1–3.2 ␮m for TiB2 , which was extremely hard (2356
HK0.025 ), and in the range of 0.45-2.53 ␮m for TiB whiskers after
plasma paste boriding in the temperature range of 973–1073 K for
3–7 h.
A diffusion model was proposed to estimate the boron diffusiv-
ities in the TiB2 and TiB layers grown on the Ti6Al4V substrate.
The diffusion model included the effect of boride incubation
times. The derived mass balance equations at the (TiB2 /TiB) and
(TiB/substrate) interfaces were used to obtain the values of boron
activation energies for Ti6Al4V alloy. As a result, the boron activa-
tion energies of TiB2 and TiB layers were respectively estimated as
136.24 ± 0.5 and 63.76 ± 0.5 kJ mol−1 , for an upper boron concen-
tration of 31.10 wt.% in TiB2 . The obtained values of boron activation
energy were compared with the data available in the literature.

References

[1] R. Filip, K. Kubiak, W. Ziaja, J. Sieniawski, The effect of microstructure on the


mechanical properties of two-phase titanium alloys, J. Mater. Process.
Technol. 133 (1–2) (2003) 84–89.
[2] R.R. Boyer, An overview on the use of titanium in the aerospace industry,
Mater. Sci. Eng. A213 (1996) 103–114.
[3] C. Leyens, M. Peters, Titanium and Titanium Alloys, Fundamentals and
Applications, Wiley-VCH Press, Weinheim, 2003.
Fig. 9. An Arrhenius relationship between the boron diffusion coefficient and the [4] C. Oldani, A. Dominguez, Titanium as a biomaterial for implants, in: S. Fokter
(Ed.), Recent Advances in Arthroplasty, InTech publisher, 2012, pp. 149–162,
boriding temperature: (A) for TiB2 phase, (B) for TiB phase.
ISBN: 978-953-307-990-5.
[5] K.G. Budinski, Tribological properties of titanium alloys, Wear 151 (1991)
203–217.
the values of boron activation energies obtained by pack and elec- [6] Ş. Taktak, H. Akbulut, Diffusion kinetics of explosively treated and plasma
trolyse liquid boriding were higher than those estimated from the nitrided Ti6Al4V alloy, Vacuum 75 (3) (2004) 247–259.
[7] O. Kessler, H. Surm, F. Hoffmann, P. Mayr, Enhancing surface hardness of
plasma paste boriding process. The values of boron activation ener- titanium alloy Ti-6Al-4V by combined nitriding and CVD coating, Surf. Eng. 18
gies in the TiB2 and TiB phases decreased by about 1.38 and 3 times, (4) (2002) 299–304.
respectively compared to Q values calculated in Refs. [37,38]. The [8] X. Yao, X. Zhang, Y. Ma, A. Fan, L. Tian, B. Tang, Surface modification of Ti6Al4V
by plasma copper and nickel alloying process, Rare Metal Mater. Eng. 43 (3)
lower value of activation energy of this study could be due to the (2014) 559–562.
fact that a large quantity of active boron was generated by plasma [9] H. Huang, P.H. Lan, Y.Q. Zhang, X.K. Li, X. Zhang, C.F. Yuan, X.B. Zheng, Z. Guo,
mechanism, which accelerates the diffusion of boron atoms to the Surface characterization and in vivo performance of plasma-sprayed
hydroxyapatite-coated porous Ti6Al4V implants generated by electron beam
titanium network structure. melting, Surf. Coat. Technol. 283 (2015) 80–88.
The boriding temperatures used in this work were 973, 1023 [10] A.K. Sinha, Boriding (Boronizing) of steels. ASM handbook, in: Heat Treating,
and 1073 K. These temperatures are lower than 1183 K which is 10th ed., ASM International, Materials Park, Ohio, 1991, pp. 437–447.
236 M. Keddam, S. Taktak / Applied Surface Science 399 (2017) 229–236

[11] N.M. Tikekar, K.S.R. Chandran, A. Sanders, Nature of growth of dual titanium [27] R.G. Fenish, Phase Relationships in the Titanium–Boron System, NRM-138,
boride layers with nanostructured titanium boride whiskers on the surface of 1964, pp. 1–37.
titanium, Scr. Mater. 57 (2007) 273–276. [28] J. Thebault, R. Pailler, G. Bontemps-Moley, M. Bourdeau, R. Naslain, Chemical
[12] K.G. Anthymidis, D.N. Tsipas, E. Stergioudis, Boriding of titanium alloys in a compatibility in boron fiber-titanium composite materials, J. Less-Common
fluidized bed reactor, J. Mater. Sci. Lett. 20 (2001) 2067–2069. Metals 47 (1976) 221–233.
[13] M. Kulka, N. Makuch, P. Dziarski, A. Piasecki, A. Miklaszewski, Microstructure [29] Y. Hai-yan, W. Qun, C. Shao-mei, G. Ping, A first-principle calculation of
and properties of laser-borided composite layers formed on commercially structural, mechanical and electronic properties of titanium borides, Trans.
pure titanium, Opt. Laser Technol. 56 (2014) 409–424. Nonferrous Met. Soc. China 21 (2011) 1627–1633.
[14] E. Atar, E.S. Kayali, H. Cimenoglu, Characteristics and wear performance of [30] M. Keddam, Simulation of the growth kinetics of the (FeB/Fe2 B) bilayer
borided Ti6Al4V alloy, Surf. Coat. Technol. 202 (2008) 4583–4590. obtained on a borided stainless steel, Appl. Surf. Sci. 257 (2011) 2004–2010.
[15] G. Kartal, S. Timur, M. Urgen, A. Erdemir, Electrochemical boriding of titanium [31] M. Keddam, M. Kulka, N. Makuch, A. Pertek, L. Małdziński, A kinetic model for
for improved mechanical properties, Surf. Coat. Technol. 204 (2010) estimating the boron activation energies in the FeB and Fe2 B layers during the
3935–3939. gas-boriding of Armco iron: effect of boride incubation times, Appl. Surf. Sci.
[16] P. Kaestner, J. Olfe, K.T. Rie, Plasma-assisted boriding of pure titanium and 298 (2014) 155–163.
TiAl6V4, Surf. Coat. Technol. 142–144 (2001) 248–252. [32] F. Li, X. Yi, J. Zhang, Z. Fan, D. Gong And, Z. Xi, Growth kinetics of titanium
[17] V. Ataibis, S. Taktak, Characteristics and growth kinetics of plasma paste boride layers on the surface of Ti6Al4V, Acta Metall. Sin. 23 (4) (2010)
borided Cp?Ti and Ti6Al4V alloy, Surf. Coat. Technol. 279 (2015) 65–71. 293–300.
[18] I. Gunes, S. Ulker, S. Taktak, Plasma paste boronizing of AISI 8620, 52100 and [33] I. Campos-Silva, D. Bravo-Bárcenas, A. Meneses-Amador, M. Ortiz-Dominguez,
440C steels, Mater. Design 32 (4) (2011) 2380–2386. H. Cimenoglu, U. Figueroa-López, R. Tadeo-Rosas, Growth kinetics and
[19] I. Gunes, S. Ulker, S. Taktak, Kinetics of plasma paste boronized AISI 8620Steel mechanical properties of boride layers formed at the surface of the ASTM F-75
in borax paste mixtures, Prot. Met. Phys. Chem. Surf. 49 (5) (2013) 567–573. biomedical alloy, Surf. Coat. Technol. 237 (2013) 402–414.
[20] B. Sarma, N.M. Tikekar, K.S. Ravi Chandran, Kinetics of growth of superhard [34] I. Campos-Silva, M. Ortiz-Domínguez, C. Tapia-Quintero, G. Rodriguez-Castro,
boride layers during solid state diffusion of boron into titanium, Ceram. Int. M.Y. Jimenez-Reyes, E. Chavez-Gutierrez, Kinetics and boron diffusion in the
38 (2012) 6795–6805. FeB/Fe2 B layers formed at the surface of borided high-alloy steel, J. Mater.
[21] B. Sarma, K.S. Ravi Chandran, Accelerated kinetics of surface hardening by Eng. Perform. 21 (2012) 1714–1723.
diffusion near phase transition temperature: Mechanism of growth of boride [35] C.M. Brakman, A.W.J. Gommers, E.J. Mittemeijer, Boriding of Fe and Fe–C,
layers on titanium, Acta Mater. 59 (2011) 4216–4228. Fe–Cr, and Fe–Ni alloys; boride-layer growth kinetics, J. Mater. Res. 4 (1989)
[22] M. Keddam, S. Taktak, S. Tasgeriten, A Diffusion Model for the Titanium 1354–1370.
Borides on Pure Titanium, Surf. Eng. 32 (11) (2016) 802–808. [36] M. Keddam, Z. Nait Abdellah, M. Kulka, R. Chegroune, Determination of the
[23] L.G. Yu, X.J. Chen, K.A. Khor, FeB/Fe2 B phase transformation during SPS diffusion coefficients of boron in the FeB and Fe2 B layers formed on AISI D2
pack-boriding: boride layer growth kinetics, Acta Mater. 53 (2005) steel, Acta Phys. Pol. A 128 (2015) 740–745.
2361–2368. [37] Z. Fay, Z.X. Guo, B. Cantor, The kinetics and mechanism of interfacial reaction
[24] E. Rudy, S. Windisch, Ternary Phase Equilibria in Transition in sigma fibre-reinforced Ti MMCs, Compos. A: Appl. Sci. Manuf. 28A (1997)
Metal–Boron–Carbon–Silicon Systems. Part I. Related Binary System, Ti–B 131–140.
System, vol. VII, Technical Report No. AFML-TR-65-2, 1966. [38] G. Kartal, S. Timur, Growth kinetics of titanium borides produced by
[25] X. Ma, C. Li, Z. Du, W. Zhang, Thermodynamic assessment of the Ti–B system, CRTD-Bor method, Surf. Coat. Technol. 215 (2013) 440–446.
J. Alloy Compd. 370 (2004) 149–158.
[26] B.F. Decker, J.S. Kasper, The crystal structure of TiB, Acta Crystallogr. 7 (1954)
77–80.

You might also like