You are on page 1of 14

Available online at www.sciencedirect.

com

ScienceDirect
Acta Materialia 83 (2015) 6174
www.elsevier.com/locate/actamat

In situ nitrided titanium alloys: Microstructural evolution during


solidication and wear

H. Mohseni,1 P. Nandwana,1 A. Tsoi, R. Banerjee and T.W. Scharf

Department of Materials Science and Engineering, University of North Texas, Denton, TX 76207, USA
Received 20 March 2014; revised 9 September 2014; accepted 12 September 2014
Available online 21 October 2014

AbstractSurface and subsurface structural evolution during sliding wear of two in situ nitrided titanium alloys, b Ti35Nb7Zr5Ta (TNZT) and
a/b Ti6Al4V (Ti64), was studied by cross-sectional transmission electron microscopy coupled with precession electron diraction. In situ nitriding
during the laser engineered net shaping (LENSe) process resulted in the formation of hard and wear-resistant Ti2N + TiN phases and nitrogenenriched a phase in TNZT and Ti64, respectively. Subsurface structural analyses of the worn nitrided TNZT revealed the tendency of a grains,
and to greater extent b grains, to undergo severe plastic deformation, forming a heavily grain-rened nanocrystalline a and b tribolayer. Corresponding precessionorientation imaging phase maps were used to determine the orientation and percentage of a and b-Ti in the worn nitrided TNZT. The
maps revealed that the nanocrystalline grains of soft/compliant b are much smaller (10100 nm) than hard/sti a grains (>100 nm). Wear reduction is
due to the combination of the above phases and the increase in the alignment of {0 0 0 2}-textured coarser a grains along the sliding direction in
absence of texture in the highly rened b grains. Conversely, nitrided Ti64 exhibited slightly increasing wear, despite higher hardness, due to the
change in sliding-induced deformation mechanism where shear bands formed and networked leading to brittle fracture and third body abrasive wear
particle generation.
2014 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
Keywords: Titanium alloys; Nitriding; Wear; Transmission electron microscopy (TEM); Precession electron diraction (PED)

1. Introduction
Titanium (Ti) alloys have long been used in many
diverse applications ranging from orthopedic biomedical
implants (femoral stems) to aerospace gas turbine engine
components (compressor blade roots) [14]. While Ti alloys
such as the a + b alloy Ti6Al4V (Ti64) and the b alloy
Ti35Nb7Zr5Ta (TNZT) are well known for their corrosion resistance, adequate fracture toughness and fatigue
strength, they exhibit high wear rates, resulting in galling
(adhesive) failure and fretting fatigue [48]. Thus, they
require surface modication treatments to improve their
wear behavior [412]. One such approach is surface nitriding, which produces a high hardness and a chemically inert
layer that overcomes the poor wear resistance of Ti alloys.
The process involves heat treating at elevated temperatures
(7001100 C) in a nitrogen-enriched atmosphere or
implantation of nitrogen ions into the surface of Ti alloys
[8,10,11,1318]. Surface nitriding is a fast, inexpensive
and simple treatment to form a hardened, wear-resistant
case layer on the surface. However, the penetration depth
of nitrogen into the bulk alloy is limited, thus restricting
the thickness of the nitride layer to mitigate wear, and the

Corresponding author; e-mail: Scharf@unt.edu


1

These authors contributed equally to this work.

high processing temperatures can result in alloy softening.


This paper focuses on an alternative approach that employs
laser engineered net shaping (LENSe) with in situ formation of a uniformly distributed wear-resistant nitrided layer.
Previously, the authors have directly introduced
titanium boride particles during LENSe processing of Ti
alloy powders to deposit in situ TiBTi composites [7,19].
However, the quality of such an interface between externally introduced reinforcements and the alloy matrix can
be rather dicult to control and results in TiB precipitate
pull-out during interfacial sliding that leads to accelerated
third body abrasive wear [7]. Thus, a reinforcement ideally
created as a product of an in situ gas reaction oers the
advantage of a thermodynamically and mechanically stable
interface with the matrix. Furthermore, by employing an
in situ nitride phase formation reaction, such as a nitrogen-enriched hard a-Ti phase, a more uniform and homogeneous distribution of the nitride phase can be
potentially achieved throughout the alloy matrix. To this
end, we have fabricated in situ nitride-reinforced TNZT
and Ti64 alloys using LENSe in the presence of nitrogen
gas to study the solidication structure, sliding wear properties and novel wear-induced deformation structures.
Mechanistic studies of the unworn and worn surfaces and
subsurfaces were conducted to fundamentally understand
the structural evolution during wear by using the
techniques of cross-sectional focused ion beam (FIB)

http://dx.doi.org/10.1016/j.actamat.2014.09.026
1359-6462/ 2014 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

62

H. Mohseni et al. / Acta Materialia 83 (2015) 6174

microscopy coupled with transmission electron microscopy


and precession electron diraction (TEM-PED). These
combined techniques will resolve, for the rst time, grain
(re)orientation and texture of wear-induced near-surface
ultrane grain (<100 nm) structures that determine nitrided
Ti alloy wear behavior. For this purpose, a grain orientation technique capable of resolving wear-induced nanoscopic grains unambiguously is required, since traditional
electron backscatter diraction (EBSD) cannot resolve
orientation and texture of highly strained wear-induced
ultrane grain (<50 nm) structures. However, PED in
conjunction with TEM (also known as EBSD-like TEM
analysis) has emerged as a versatile technique to determine
nanoscopic grain orientation and phases [2025], and will
be utilized for the rst time here to study the evolution of
wear-induced ultrane grain structures. The PED technique
was developed by Vincent and Midgley [26] to overcome
dynamic diraction concerns and produce diraction patterns that exhibit a more kinematical character [27]. In a
TEM-retrotted PED apparatus, the incident electron
beam is simultaneously scanned across the sample and precessed (tilted and rotated in a conical fashion) around the
optical axis of the TEM [20,21]. At each point, electron diffraction patterns, or more specically quasi-kinematical
PED patterns, are collected with a convergent nanoprobe
using an external CCD camera mounted in front of the
TEM screen. The acquired patterns are compared with
pre-calculated (simulated) diraction patterns using a
cross-correlation technique to nd the best match for all
possible material phases and orientations. A major advantage of PED over traditional EBSD is the much ner spatial
resolution of the TEM probe size that allows previously
inaccessible phase orientation mapping of nanostructural
materials.

in preparing cross-sectional TEM samples was to deposit Pt


by using electron-beam (e-Pt) and then ion-beam (i-Pt),
which protect the TEM sample against Ga ion-beam damage during milling. The second step includes milling and
nal polishing of the lifted-out sample using Ga ions at a
conservatively low voltage at 5 keV and current at 70 pA.
The TEM is retrotted with a PED unit (DigiSTAR
P1000 system, NanoMEGAS Sprl, Brussels, Belgium)
[20,21] that was used to acquire phase and crystal orientation maps at a step size of 5 nm with a beam diameter of
1 nm. The spot diraction patterns were collected with
a precession angle of 0.65. Data was exported into TSL
OIM Analysis 5.0 software for analysis of pole gures
and determination of phase concentrations.
Standard Vickers microhardness measurements were
conducted with a 300 g load held for 15 s with ve measurements made for each alloy. Sliding wear behavior was
determined with a ball-on-disc tribometer (Falex Implant
Science ISC 200). Unidirectional sliding wear tests on the
alloy discs were conducted against a Si3N4 counterface ball
(H = 22 GPa) of diameter 3.175 mm under a normal load
of 1 N that corresponds to an initial mean Hertzian contact
stress of 500 MPa. A minimum of three tests were conducted for each alloy at room temperature and relative
humidity (40%) over a total sliding distance of 200 m at
a speed of 2.1 cm s1. Following the sliding tests, a stylus
prolometer (Veeco Dektak 150) was employed to measure
wear depth, width and cross-sectional area for each alloy to
calculate wear factor (rate) based on the volume of material
loss (mm3) divided by the product of load (N) and total
sliding distance (m).

3. Results and discussion


3.1. Microstructure

2. Experimental methods
A LENSe 750 system (Optomec, Inc.) with an Nd:YAG
laser at a wavelength of 1.064 lm and a power of 760 W
was used to deposit blends of elemental powders under
100% Ar and 100% N2 atmospheres for deposition of base
and nitrided (TNZT-N and Ti64-N) alloys, respectively.
The TNZT and Ti64 alloy powders (manufactured by
TIMET, Inc.) were mixed with a twin-roller ball mill in
the ratio of Ti35Nb7Zr11Ta and Ti6Al4V (wt.%).
More details of the LENSe processing parameters are
reported elsewhere [7,28]. As-deposited base and nitrided
TNZT and Ti64 alloys were subsequently surface polished
using 1200-grade SiC grinding paper followed by ne polishing down to 200 nm arithmetic surface roughness with
0.05 lm colloidal silica. X-ray diraction (Rigaku Ultima
III diractometer) with a Cu Ka source was used for phase
identication. Surface microstructural characterization was
carried out in a eld emission gun scanning electron
microscope (FEI Nova Nano SEM230).
Site selective, cross-sectional FIB-SEM studies were
performed on the alloys with an FEI Nova 200 Nanolab
dual-beam microscope. Cross-sectional lift-outs acquired
outside and inside worn areas were then analyzed with
bright-eld TEM (BFTEM) and energy-dispersive X-ray
spectroscopy (EDS) line proles and maps. These analyses
were performed in a FEI Tecnai G2 F20 S-twin transmission electron microscope operated at 200 keV. The rst step

3.1.1. Nitrided Ti64


XRD patterns of LENSe deposited Ti64 and nitrided
Ti64 are shown in Fig. 1a. For both alloys, the reections
were indexed corresponding to hexagonal close-packed
(hcp) a and body-centered cubic (bcc) b phases. There do
not appear to be any distinct peaks attributable to either
TiN or TiN0.3 phases in nitrided Ti64, as previously
reported for other laser processing routes incorporating
nitrogen into the surface layers [15,29,30]. The presence
of the latter TiN0.3 phase cannot be unequivocally ruled
out since its lattice parameters are very close to the a-Ti
phase with the same hexagonal structure. However, nitriding did result in the a-Ti peaks shifting slightly to lower 2h
values, indicating an increase in lattice parameters (results
summarized in Table 1), which has been attributed to the
combined eect of nitrogen concentration and tensile residual stress due to rapid cooling during laser processing [15].
The c/a ratio also increased slightly from 1.59 to 1.60, indicating incorporation of nitrogen into the a-Ti phase possibly as random solid solution, since nitrogen, a strong a
stabilizer element, promotes a formation and has a solubility of up to 25 at.% in a-Ti [15,29].
The comparison of the microstructures of the LENSe
deposited Ti64 and nitrided Ti64 samples is shown in the
backscatter electron (BSE) micrographs presented in
Fig. 2ad. The microstructure of Ti64 appears to exhibit ne
scale a laths in the b matrix (Fig. 2a and b), typically

H. Mohseni et al. / Acta Materialia 83 (2015) 6174

Fig. 1. XRD patterns obtained from LENS deposited (a) Ti64 and
nitrided Ti64 (Ti64-N), and (b) TNZT and nitrided TNZT (TNZT-N)
alloys.

referred to as a basketweave microstructure in a/b Ti alloys,


containing multiple crystallographic variants of a intersecting within the same prior b grain. Such a basketweave a/b
microstructure can be attributed to the high cooling rates
involved in LENSe deposition due to the substrate acting
as an ecient heat sink. These ne scale a laths have been
attributed to martensitic decomposition of the b matrix
[15]. However, in the present case, the BSE image shown
in Fig. 2b clearly shows a signicant dierence in contrast
between the darker a laths and the brighter b matrix, indicative of the partitioning of alloying elements between these
two phases. Hence it is unlikely that these ne scale a laths,
observed in the LENSe deposited Ti64 sample, are purely

63

martensitic in nature. In contrast, the BSE images of the


nitrided Ti64 sample shown in Fig. 2c and d exhibit a very
dierent microstructure. Comparing the higher-magnication images for both alloys (Fig. 2b and d), there are two
principal dierences. First, Fig. 2d shows there is a more
pronounced contrast between the a precipitates (darker)
and the b matrix (lighter) in nitrided Ti64, indicating more
substantial atomic mass dierence between these two
phases, presumably arising from the partitioning of V into
the b matrix, while the lighter Al and N partition into the
a precipitates. Second, the morphology of a precipitates in
nitrided Ti64 is substantially dierent compared to base
Ti64. The a precipitates in nitrided Ti64 are coarser, exhibit
a smaller aspect ratio, and in some cases appear to be more
equiaxed in shape. There also appear to be two distinctly
dierent morphologies of a precipitates in nitrided Ti64
(Fig. 2c and d): a smaller volume fraction of larger elongated needle-like or acicular precipitates exhibiting complex
internal contrast variations (referred to as Type 1 in
Fig. 2d); and smaller-scale precipitates exhibiting a lower
aspect ratio (referred to as Type 2 in Fig. 2d).
Further investigation of these two types of precipitates
was carried out by mapping the elemental distribution with
EDS. Fig. 3a is a BSE image of the region that was mapped
and Fig. 3bd show the EDS maps of nitrogen, aluminum
and vanadium, respectively. Interestingly, the nitrogen content of the Type 1 needle-like precipitates is higher compared to the Type 2 precipitates. In contrast, these
needle-like precipitates have a lower Al content compared
to the Type 2 precipitates. The two Type 1 precipitates
denoted in Fig. 3a appear to have similar composition, thus
indicating they are the same phase, while the dierence in
their morphologies may arise from the sectioning eect
when viewed in the limited 2-D eld of view. It should be
noted that both types of precipitates are coarser compared
to the ne scale a laths observed in the LENSe deposited
base Ti64 alloy (Fig. 2b). Therefore, it can be concluded
that in the case of the nitrided Ti64 sample, the microstructure is not martensitic but rather involves diusional partitioning of the alloying elements and the introduction of
nitrogen into the a precipitates clearly changes their aspect
ratio and morphology, in agreement with previous reports
in the literature [28]. These nitrogen-enriched a precipitates
exhibiting a dierent morphology will subsequently be
referred to as a(Ti,N) precipitates in this paper. Therefore,
LENSe processing in a nitrogen atmosphere has a pronounced eect on the stabilization of the a(Ti,N) solidsolution phase, in agreement with the XRD results. An
enhancement in the precipitation of a phase has also been
observed on the surface of gas nitrided Ti64 [18]. The vanadium EDS map in Fig. 3d shows the surrounding b-Ti
phase since vanadium is a b stabilizer.
A cross-sectional BFTEM image of nitrided Ti64 is
shown in Fig. 3e. It is evident from this cross-section that
there is a combination of an a-lath and b-rib that conrms
the b phase is present despite its weak XRD signal in

Table 1. Lattice parameters and d spacings of signicant planes of a and b phases in Ti64, Ti64-N, TNZT and TNZT-N alloys.
)
)
)
)
)
Alloy
d(1010)a (A
d(0002)a (A
aa (A
ca (A
c/a
d(110)b (A
Ti64
Ti64-N
TNZT
TNZT-N

2.526
2.535

2.587

2.335
2.345

2.408

2.917
2.927

2.987

4.651
4.686

4.816

1.59
1.60

1.61

2.342
2.333

)
ab (A

3.312
3.30

64

H. Mohseni et al. / Acta Materialia 83 (2015) 6174

Fig. 2. (a) Low- and (b) high-magnication backscatter SEM images of LENS deposited Ti64 alloy. (c) Low- and (d) high-magnication backscatter
SEM images of LENS deposited nitrided Ti64 (Ti64-N) alloy. See text for explanation of Type 1 and 2 precipitates.

Fig. 1a. An EDS line prole (not shown) conrmed an


increase in the vanadium X-ray counts with a corresponding decrease in titanium counts across the b-rib. The dierence in the solubility of vanadium, a b stabilizer, in a-Ti
and b-Ti clearly indicates the border between the a-lath
and b-rib. The b-Ti has a higher solubility for vanadium and
a lower solubility for Al, which is an a stabilizer. From both
the TEM image and the EDS line prole, this b-rib is
5070 nm thick.
3.1.2. Nitrided TNZT
The structure of LENSe deposited TNZT is signicantly dierent from that of Ti64 as conrmed by the
XRD pattern in Fig. 1b which reveals the presence of peaks
corresponding to the b phase and negligible peaks of the a
phase. Such an observation was anticipated due to the large
amounts of b stabilizing alloying elements (Nb, Zr and Ta)
in TNZT. Conversely, the LENSe deposited nitrided
TNZT sample exhibits the precipitation of multiple phases
and
that were indexed as tetragonal e-Ti2N (a = 4.822 A
), cubic rocksalt d-TiN (a = 4.226 A
), a-Ti
c = 3.067 A
and b-Ti. Table 1 summarizes the calculated d-spacings
and lattice parameters for b-Ti in TNZT and nitrided
TNZT. The incorporation of nitrogen in TNZT shows that
nitrogen has an insignicant eect on the b phase since
).
there is no change in the lattice parameter (a = 3.30 A
This observation is expected since nitrogen is known to
be a strong a stabilizer and has limited solubility in the b
phase at ambient temperature. However, the XRD pattern
in Fig. 1b shows that signicant precipitation of the a phase
occurred due to nitrogen incorporation into LENSe
deposited TNZT. A similar eect has been observed on

the surface of nitrided TNZT [18]. The lattice parameters


of the a phase in nitrided TNZT were calculated to be
and c = 4.816 A
, which results in a c/a ratio
a = 2.987 A
of 1.61 compared to the ideal ratio of 1.59 for pure a-Ti.
The slight increase in the c/a ratio due to nitriding is consistent with the values reported in previous studies
[28,29]. This increase is an indication of saturation of the
a phase by nitrogen and corresponds to the solubility limit
of nitrogen (i.e. 2425 at.%) in a-Ti, shown by the TiN
binary phase diagram in Fig. 4 [31].
The inuence of nitrogen on the structure of the LENSe
deposited TNZT is also evident by comparing the BSE
images of base (Fig. 5a) and nitrided (Fig. 5b) TNZT.
The dendritic microstructure typical of single-phase TNZT
can be seen in Fig. 5a where b dendrites are separated by
darker interdendritic regions that show microsegregation
of alloying additions. Such microsegregation is a common
feature in alloys with a high concentration of b stabilizers
[3]. Fig. 5b shows a completely dierent microstructure
for nitrided TNZT where the phase with brighter contrast
represents the b matrix and the phases exhibiting the darker
contrast include TiN/Ti2N and a (grain boundary and
intragranular a precipitates). The higher-magnication
inset shown in Fig. 5b clearly shows these two dierent precipitate morphologies: ner scale intragranular a precipitates with platelet morphology, and coarser TiN/Ti2N
exhibiting a dendritic morphology. Fig. 5c shows a crosssectional STEM-HAADF image along with EDS elemental
maps of a site-specic sample prepared from the nitrided
TNZT sample. It is evident from the compositional maps
that the intragranular a precipitates (200 nm thick) are
decient in Nb and Ta (b stabilizers), whereas the b matrix

H. Mohseni et al. / Acta Materialia 83 (2015) 6174

65

Fig. 3. (a) Backscatter SEM image of nitrided Ti64 alloy with corresponding EDS maps of (b) nitrogen, (c) aluminum and (d) vanadium. (e) Crosssectional BFTEM image of nitrided Ti64.

is enriched in Nb and Ta. The Zr map was noisy due to the


lower counts and hence has not been included.
3.1.3. Evolution of microstructure in nitrided Ti64 and
TNZT
From the XRD, SEM and TEM/EDS analyses, the
eect of nitrogen addition on the microstructural evolution
is considerably dierent for LENSe deposited Ti64 and
TNZT. The TiN binary phase diagram in Fig. 4 can

provide some insights towards the probable sequence of


phase evolution during in situ nitriding of Ti64 and TNZT.
Since Ti64 is an a/b alloy, addition of nitrogen to this system can possibly expand the a phase stability region on the
phase diagram. The increased volume fraction of a phase is
capable of absorbing more nitrogen thereby preventing the
precipitation of nitride phases. Conversely, the eect of
nitrogen can be somewhat complicated in nitrided TNZT
since the high concentration of b stabilizing elements

66

H. Mohseni et al. / Acta Materialia 83 (2015) 6174

Fig. 4. Binary TiN phase diagram [31].

Fig. 5. Low-magnication backscatter SEM images of LENS deposited (a) TNZT and (b) nitrided TNZT (TNZT-N) alloys. (c) Cross-sectional
HAADF-STEM image of nitrided TNZT alloy with corresponding titanium, niobium and tantalum EDS maps.

(Nb, Zr and Ta) can counteract the a stabilizing eects of


nitrogen. Furthermore, the complex interplay of the multiple components in this alloy can possibly lead to changes in
the shapes and sizes of the various phase elds on the

multicomponent phase diagram in such a manner as to


nucleate primary TiN precipitates in the liquid phase followed by nucleation of b grains, as evident by the resulting
microstructure in Fig. 5b. Consequently, the primary TiN

H. Mohseni et al. / Acta Materialia 83 (2015) 6174

precipitates forming in the liquid exhibit a dendritic morphology, and can act as heterogeneous nucleation sites
for the b grains during solidication, resulting in more
equiaxed b grains (Fig. 5b) as compared to the dendritic
b grains observed in LENSe deposited TNZT in the
absence of nitrogen (Fig. 5a). Despite the prior formation
of the d-TiN dendrites, at the elevated temperatures where
the b grains nucleate, they have a certain solubility of nitrogen, as evident from the TiN binary phase diagram in
Fig. 4. Thus on subsequent cooling this retained nitrogen
will facilitate the precipitation of a phase at both heterogeneous grain boundary sites as well as homogeneously
within the b grains. Furthermore, as soon as a precipitates
within the b matrix, the retained nitrogen within the b partitions to the a phase, resulting in an increase in the c/a
ratio of this phase.
To better understand the possible phase formation
sequence, isothermal sections of the TiNbN ternary system were simulated using PANDATe developed by Computherm LLC and are shown in Fig. 6ac, corresponding
to temperatures of 2000, 1000 and 100 C, respectively.
Nb was chosen as the third component since it is the most
abundant element in the alloy after Ti. This ternary phase
diagram can provide some valuable insights on the competing nature of the b stabilizing elements and a stabilizing
eects of nitrogen on the phase stability in this complex
alloy. From Fig. 6a, it evident that fcc rocksalt d-TiN is a
high-temperature phase and exists at 2000oC in conjunction
with bcc b phase and liquid for a nominal composition of
35 wt.% Nb and 10 wt.% nitrogen. Fig. 6b shows for
a similar average composition that bcc b, d-TiN and Ti2N
are the stable phases at 1000oC, indicating the possible formation of the second nitride, Ti2N phase. Further lowering
the temperature to 100 C results in the precipitation of the
hcp a phase in addition to the bcc b, d-TiN and Ti2N
phases, as shown in Fig. 6c. The retention of d-TiN dendrites, in the LENSe deposited TNZT at room temperature, is explained by the non-equilibrium cooling involved

Fig. 6. Isothermal sections of ternary TiNbN phase diagram


showing Ti-rich compositions at (a) 2000 C, (b) 1000 C and (c)
100 C. Closed circle denotes the approximate ternary composition.

67

and the sluggish decomposition kinetics of this nitride


phase. Furthermore, it should be noted that the PANDATe predictions hold only for the equilibrium phases
and are dependent on the accuracy of the thermodynamic
models used and their calibration. It can be summarized
that the basic variations in the starting microstructures
and nature of alloying additions in Ti64 and TNZT are
responsible for the unambiguous dierences that nitrogen
addition has in modifying their respective microstructures.
3.2. Microhardness and wear behavior
The result of microhardness analysis of the LENSe
deposited base and nitrided Ti64 and TNZT are summarized in Table 2. Ti64 has a higher hardness (419 VHN)
compared to TNZT (288 VHN), since the b phase is softer
than a, which makes the Ti64, with a/b structure, harder
than TNZT. The averaged microhardness of both alloys
approximately doubled with nitriding: 1047 and 518
VHN for nitrided Ti64 and TNZT, respectively. The
increasing hardness with in situ nitriding can be attributed
to the formation of TiN/Ti2N phases in TNZT, and the
nitrogen-enriched a phase, i.e. increasing volume of
a(Ti,N) precipitates in Ti64. This increase in microhardness
is consistent with prior work on nitriding a/b and b Ti
alloys [16,28,29].
Sliding wear behavior comparisons of the four alloys are
summarized in Table 2. The values of the wear track depth
and width along with the wear factor were substantially
higher for Ti64 compared to TNZT despite the higher hardness of Ti64. These observations do not agree with Archards law for adhesive and abrasive sliding wear,
according to which lower wear volume loss corresponds
to a harder material. This suggests that hardness alone cannot account for the dierence in wear behavior. After
examining the Ti64 and TNZT wear surfaces by SEM, it
is evident that the Ti64 surface shown in Fig. 7a exhibits
large amounts of debris particulates on the surface. EDS
conrmed these are SiOx particles that were tribochemically
transferred from the sliding Si3N4 counterface to the wear
surface. A harder surface such as Ti64 would more easily
abrade the Si3N4 counterface, thus resulting in a larger area
of adhesively transferred SiOx wear debris. In comparison,
a much lower concentration of transferred SiOx particles is
present on the TNZT shown in Fig. 7b, since TNZT is
softer and less abrasive. There are also wear surface morphology dierences between the alloys. The SEM images
of the Ti64 surface in Fig. 7a exhibits extensive plastic
deformation, microplowing and cutting, all typical processes associated with an abrasive wear mechanism. Conversely, the TNZT wear surface in Fig. 7b shows less
severity in wear. It is apparent that the surface morphology
has undergone a plastic shear deformation mechanism
typically associated with less wear volume removal. The
SEM images of the TNZT surface exhibit ductile layering
and smearing of coarse wear platelets with varying sizes
(515 lm) aligned along the sliding direction. These
dierences in wear surface morphology are consistent with
previous sliding wear studies on Ti64 and TNZT [7,32].
Fig. 8 shows SEM images of the worn nitrided Ti64 and
TNZT surfaces. From these and other images, the averaged
wear track widths calculated in Table 2 were considerably
smaller for nitrided Ti64 and TNZT alloys compared to
the base alloys, and the nitrided TNZT exhibited slightly

68

H. Mohseni et al. / Acta Materialia 83 (2015) 6174

Table 2. Summary of averaged Vickers microhardness values, wear track depth, width and wear factor for Ti64, Ti64-N, TNZT and TNZT-N alloys.
For nitrided Ti64 and TNZT alloys, adhesive wear of Si3N4 counterface transferred to the wear track prevented accurate wear depth and factor
calculations.
Alloy

Microhardness (HV)

Wear track depth (lm)

Wear track width (lm)

Wear factor (mm3 (Nm)1)

Ti64
Ti64-N
TNZT
TNZT-N

419 32
1047 29
288 21
518 35

20.3 0.7

3.9 0.4

715 39
330 15
420 20
312 11

6.4 0.2  10

5.3 0.3  10

Fig. 7. Low- and high-magnication SEM images of (a) Ti64 and (b) TNZT wear tracks.

smaller width than nitrided Ti64. However, the wear track


depth for both nitrided alloys was not measurable since it
was within the noise of the surface prolometer (50 nm
depth resolution). Hence wear factors could not be calculated, although wear track depths for both nitrided alloys
were orders of magnitude less than the base alloys. These
observations indicate a reduction in the real area of contact
between the nitrided alloys and the Si3N4 counterface due
to increased hardness from nitriding that results in decreasing wear. Measuring the wear depth values was also dicult due to the transfer of SiOx debris from the Si3N4
counterface to the wear track thereby preventing accurate
wear calculations. An additional comparison between the
Fig. 8 SEM images of nitrided Ti64 and TNZT wear surfaces reveal that the nature of surface fracture appears different. Surface cracks in nitrided Ti64 occur at more of an
oblique angle to the sliding direction, denoted by arrows in
the higher-magnication image in Fig. 8a, while the cracks
in nitrided TNZT appear to be conned within the
near-surface platelet wear debris. These dierences in
microstructural features imply dierent wear mechanism(s)
are occurring between nitrided Ti64 and TNZT and will be

discussed in the next section. While the above observations


provide some insight into sliding-induced deformation,
they are conned to the surface and are of lower resolution.
Therefore, a more detailed mechanistic study of the surface
and subsurface structural evolution during sliding wear is
necessary and will be discussed below.
3.3. Wear-induced deformation mechanisms
3.3.1. Cross-sectional TEM analysis
To further investigate the sliding wear mechanisms of
nitrided Ti64 and TNZT, cross-sectional FIB/TEM analyses were performed inside the wear tracks. Such wear deformation analyses for nitrided Ti alloys are unknown, and the
use of PED in this context is particularly novel. Fig. 9 represents the cross-sectional TEM and PED orientation analysis directions for ball-on-disc sliding, namely SD = sliding
direction, TD = transverse direction and ND = normal
direction. During LENSe processing, the alloy deposition
is parallel to the ND. Therefore, the top wear surface corresponds to the ND of the frame, while for the wear subsurface viewed in cross-section, the alloy deposition is

H. Mohseni et al. / Acta Materialia 83 (2015) 6174

69

Fig. 8. Low- and high-magnication SEM images of nitrided (a) Ti64 and (b) TNZT wear tracks. Arrows in the higher magnication image in (a)
denote cracks.

perpendicular to the ND of the frame, i.e. along the TD.


Fig. 10 presents the wear-induced surface and subsurface
structural evolution in nitrided Ti64 acquired along the
TD. Both the low- (Fig. 10a) and higher- (Fig. 10b) magnication BFTEM images show that on the wear surface,
below the protective Pt, is a 430 nm thick layer of amorphous silica (a-SiOx), which was conrmed by an EDS line
scan. This layer transferred from the Si3N4 counterface
during the wear process as previously discussed in terms
of the SEM surface analysis in Fig. 8a, showing that silica
wear debris is present in some locations on the surface. The
subsurface BFTEM images in Fig. 10 further show that this
porous SiOx transfer layer exhibits cracking along the SD.
Below the SiOx transfer layer is a near-surface deformed
region that contains an extensive network of shear bands
with dislocation substructures (tangles/networks) in
between. The presence of shear bands, occurring to a

Fig. 9. Schematic representation of the cross-sectional TEM and PED


orientation analyses directions for ball-on-disc sliding wear. SD = sliding direction, TD = transverse direction, ND = normal direction.

subsurface depth of 2 lm based on this FIB liftout,


implies that wear-induced fracture processes are occurring
in nitrided Ti64. This process involves continuous formation of sliding-induced shear bands (some up to 1.5 lm
long) that branch, network and evolve into cracks, like
the ones shown on the surface in Fig. 8a. Subsequently,
Ti64 third-body abrasive wear particles of varying sizes
are generated from these networked cracks. These particles,
either by themselves or when combined with SiOx wear debris, result in microplowing and cutting, processes associated
with an abrasive wear mechanism, which is consistent with
the SEM surface imaging in Fig. 8a. In addition to the brittle behavior, there is evidence of wear-induced plasticity in
nitrided Ti64. The dislocated substructure shown in Fig. 10
surrounding the shear bands suggests a work-hardened
zone that would lead to further hardening in the worn area.
In comparison, the base Ti64 wear surface, shown in
Fig. 7a, did not exhibit signs of surface cracking despite
its high wear rate. Therefore, in situ nitriding resulted in
increasing hardness that improved the wear resistance of
Ti64; however, the enhancement in the precipitation of a
phase with nitriding is likely responsible for surface/
subsurface cracking.
Completely dierent wear-induced surface and subsurface structures were observed in nitrided TNZT. Fig. 11a
and b represent two regions of the nitrided TNZT wear
track, acquired along the TD, where the SiOx wear debris
is and is not present on the surface, respectively. This conrms that there is a heterogeneous distribution of the SiOx
debris on the wear surface. For the case where SiOx is present, below it is a non-uniform (400900 nm thickness)
heavily grain rened zone. This indicates that the depth
of plastic strain due to sliding is within this thickness range.

70

H. Mohseni et al. / Acta Materialia 83 (2015) 6174

Fig. 10. (a) Low- and (b) high-magnication cross-sectional BFTEM


images of worn nitrided Ti64 acquired along the TD.

The severely deformed zone is composed of equiaxed grains


varying from 10 to 200 nm size that are presumably a and
b grains and will be veried by PED in the next section.
Below this wear-induced plastically deformed zone,
denoted by the dashed line, is a transition to the undeformed bulk structure. Fig. 11b shows the same severely
deformed zone with nanocrystalline grain renement,
although with a more uniform thickness of 700 nm. The
presence of this zone without transferred SiOx on top suggests counterface wear is not necessary to induce grain
renement. Furthermore, there is no evidence of shear
bands that network into cracks, as previously shown for
nitrided Ti64, within this heavily grain rened zone or
between the interfaces. While the nitrided TNZT grain
rened zone is less prone to fracture processes, both alloy
surfaces are unstable due to localized shear instability during sliding. As a result, the near-surface material can generate ner loose wear debris particles that can also
mechanically mix with the counterface material, e.g. SiOx.

Fig. 11. Cross-sectional BFTEM images of worn nitrided TNZT (a)


with and (b) without SiOx counterface wear debris. The white dashed
lines represent the borders between the wear-induced zones. Images
acquired along the TD.

While the hardness of both alloys increases with nitriding,


it alone cannot account for the improvements in wear
resistance, e.g. the hardness of base Ti64 and nitrided

H. Mohseni et al. / Acta Materialia 83 (2015) 6174

TNZT are approximately the same despite their large dierences in wear values (Table 2). Thus, the dierences in surface and subsurface structural evolution during sliding
wear are also determining the wear behavior of these alloys.
Recrystallization due to sliding is unlikely since the computed interfacial ash temperature due to frictional heating
is low (40 C), due in part to the low contact stress and
sliding speed used in this study. Therefore, the wearinduced structural changes are a result of the high deformation strain during sliding, e.g. large plastic shear strains
exceeding 10 are common with sliding contact of metallic
materials [33]. In the case of nitrided TNZT, formation of
more stable nanocrystals of equiaxed a and b-Ti without
fracture, e.g. shear bands, indicates the potential to accommodate the interfacial shear stresses during sliding contact
that reduces wear. Detailed TEM-based orientation mapping of the severely deformed zone in nitrided TNZT will
be presented in the following section to explain what role
these rened a and b grains have on the wear behavior
and mechanism. The presence of nanocrystalline grain
renement has been determined to improve sliding wear
resistance in titanium when processed by a severe plastic
deformation (SPD) technique. For example, Stolyarov
et al. reported an improvement in the wear resistance of
commercially pure titanium after equal-channel angular
pressing (ECAP) that resulted in an ultrane-grained
(UFG) structure that was responsible for a decrease in
the adhesive component of friction and wear [34]. Similarly,

71

UFG (300 nm size) Ti prepared by ECAP showed that


grain renement was responsible for decreasing wear rates
in comparison to coarse-grained Ti [35]. In addition, creation of a nanocrystalline surface layer (160 lm thick) on
pure Ti using a surface mechanical attrition treatment
was determined to be an eective procedure to reduce wear
[36]. A recent review article on the tribological properties of
UFG Ti and other metallic materials processed by SPD
provides more details on the above ECAP studies [37].
Lastly, there was no evidence of titanium oxide, e.g.
TiO2, formation by tribo-oxidation on the wear surfaces
for both nitrided alloys, due in part to the aforementioned,
relatively low interfacial ash temperatures during sliding.
Under more severe testing conditions, e.g. higher contact
stresses and sliding speeds, frictional heating has been
shown to facilitate tribochemical formation of TiO2 surface
layers when sliding on ECAP UFG Ti that resulted in lower
friction and wear [35]. However, these sliding conditions,
especially the high contact loads/stresses, are more severe
than would be exhibited in certain applications, e.g. biomedical load-bearing applications such as hip implants.
Furthermore, such surface layers can mask the role and
importance of the dierent underlying wear-induced
structures.
3.3.2. Cross-sectional TEM-PED analysis
Fig. 12a and b show cross-sectional inverse pole gure
(IPF) orientation maps of b-Ti and a-Ti, respectively,

Fig. 12. Cross-sectional IPF maps of worn nitrided TNZT showing orientation map of (a) b-Ti, and (b) a-Ti and phase concentration maps of the (c)
top portion of the zone 1 worn region, and (d) bottom portion of the zone 2 unworn region. The insets show grain orientation color codes. The white
dashed lines represent the border between zone 1 and zone 2. In (c) and (d) phase maps, red represents b grains and green represents a grains taken
from zone 1 (yellow dashed line box) and zone 2 (brown dashed line box) regions. (For interpretation of the references to colour in this gure legend,
the reader is referred to the web version of this article.)

72

H. Mohseni et al. / Acta Materialia 83 (2015) 6174

acquired from PED analysis of worn nitrided TNZT. These


orientation IPF b-Ti and a-Ti grain maps, with their color
code in the inset, were acquired along the TD from the
same FIB liftout as the cross-sectional TEM image shown
in Fig. 11a. Based on the b-Ti and a-Ti grain size and distribution, there are two distinct zones in the worn cross-section, denoted zones 1 and 2 by the white dashed lines in
Fig. 12a and b. Zone 1 is the top portion with wear-induced
grain renement as previously observed in the cross-sectional BFTEM images in Fig. 11. Zone 1 is composed of
both ultrane b-Ti grains (10100 nm size) shown in
Fig. 12a and mostly larger grains of a-Ti shown in
Fig. 12b. Zone 2 is the bottom portion of the unworn
nitrided TNZT bulk structure that was not exposed to
wear-induced subsurface contact stresses.
The IPF maps in Fig. 12 were also used to observe grain
orientation changes and texture evolution resulting from
sliding-induced deformation. Fig. 12a shows that the b-Ti
grains in zones 1 and 2 are randomly orientated while the
majority of the a-Ti grains (unworn and worn) exhibit
(2
1
10) planes that are parallel to the ND, represented on

the IPF map in green in Fig. 12b. Not shown are IPF maps
rotated at 90 to the TD of the sample (viewed along the
ND) where the (0001) planes, red on the IPF map, are perpendicular to the ND. Therefore, the a-Ti (0002) planes are
parallel to the wear surface, i.e. lie in the sliding plane (SD
and TD). From these IPF maps, it is evident, at least based
on this particular TEM liftout, that a pronounced a-Ti texture of the basal type remains after sliding. More details of
this texture will be presented below with pole gure analysis. Furthermore, Fig. 12 shows there are clear dierences in
the b-Ti and a-Ti grain sizes in zone 1 compared to the
unworn grain sizes in zone 2, with the former exhibiting
more grain renement. This is due to the softer and more
compliant nature of the b grains in comparison to the surrounding hard/sti a grains [11], which results in signicant
b grain renement during the wear process. This grain size
comparison is more evident with viewing the combined b-Ti
and a-Ti phase maps of zones 1 (worn) and 2 (unworn) in
Fig. 12c and d, respectively. Clearly the b-Ti grains,
denoted in red, are more rened in zone 1, while the a-Ti
grains in green are less rened. Phase concentration

Fig. 13. Comparison of b-Ti and a-Ti pole gures acquired from worn (zone 1) and unworn (zone 2) nitrided TNZT viewed along ND. Units for the
texture scale bars are in multiples of random.

H. Mohseni et al. / Acta Materialia 83 (2015) 6174

calculations determined that zone 1 contains 60% a-Ti and


40% b-Ti, whereas this proportion changes to 47% and
53%, respectively, in zone 2. While this is not a detailed statistical phase analysis based on several FIB liftouts along
the wear track, it does show that more a-Ti is present in
the wear-induced zone 1 compared to unworn zone 2. This
distribution is likely a result of the aforementioned a-Ti
phase exhibiting higher hardness and stiness than b-Ti.
Thus, the a-Ti phase has a higher resistance to interfacial
shear stresses present during sliding wear, whereas the
softer/compliant b-Ti grains become more rened with an
increasing probability of becoming loose, ejected wear particles that would account for their deceasing concentration
in zone 1.
Corresponding b-Ti and a-Ti pole gures acquired from
worn (zone 1) and unworn (zone 2) nitrided TNZT are presented in Fig. 13. These pole gures are viewed down the
ND (rotated 90 to the TD) to facilitate interpretation of
the a-Ti (0002) texture evolution during wear. In comparing
the a-Ti worn and unworn < 0001 > pole gures, there is an
increase in the alignment of the a-Ti basal planes along the
SD due to the wear process. In contrast, the unworn aTi < 0001 > pole gure shows there are more spots with less
alignment along the SD. This sliding-induced alignment of
a-Ti basal planes would lower interfacial shear stress during
sliding thereby reducing wear. Basal slip of hcp metals along
the sliding direction has been determined to lower the resistance to shear in sliding contacts [38]. The b-Ti pole gures
in Fig. 13 show that the rened and unworn b-Ti grains do
not exhibit any specic texture, which is corroborated by the
lower spread in texture intensity values compared to a-Ti
values. Along with the higher degree of randomness of the
zone 1 b-Ti grains, there is also the possibility that these
softer/compliant b-Ti rened grains exhibit grain rotation
during the wear process.

4. Summary and conclusions


Detailed studies of processingstructureproperty relations were conducted for in situ nitrided titanium alloys
TNZT and Ti64 with comparisons to untreated TNZT
and Ti64. Dierences in microstructural evolution during
solidication and wear have been determined as follows:
1. Nitriding during LENSe deposition resulted in a substantial increase in the microhardness of TNZT and
Ti64 due to the precipitation of TiN/Ti2N phases and/
or nitrogen-enriched a phase, respectively. This phase
evolution with nitrogen addition during solidication
was relatively well predicted with thermodynamic calculation software.
2. Increased hardness in both nitrided alloys was partially
responsible for the reduction in real area of contact during interfacial sliding that resulted in improved wear
resistance. However, hardness alone cannot account
for the dierences in wear behavior with and without
nitriding since there were clear dierences in slidinginduced deformation mechanisms. In the case of nitrided
Ti64, strain-induced brittle fracture in the form of surface and subsurface shear bands resulted in cracking
and subsequent third-body abrasive wear debris generation. The networked shear bands were surrounded by a
dislocated substructure suggesting a work-hardened
zone had led to further hardening in the worn area. In

73

the case of nitrided TNZT, strain-induced plastic deformation resulted in the formation of a heavily grainrened tribolayer composed of both ultrane b-Ti grains
(10100 nm size) and larger a-Ti grains. Formation of
more stable nanocrystals of equiaxed a and b-Ti without
fracture, e.g. shear bands, indicates the potential to
accommodate the interfacial shear stresses during sliding
contact that reduces wear.
3. Orientation imaging analysis of nitrided TNZT using
TEM-PED revealed that the a-Ti grains exhibit basal
texture and become more aligned parallel to the sliding
direction during wear, while the softer and compliant
nanocrystalline b-Ti grains are randomly orientated
and likely prone to rotation during wear. These sliding-induced phenomena coupled with the hard phases
that form during solidication are responsible for
improvements in wear resistance with in situ nitriding.
Acknowledgements
The authors would like to acknowledge the support of the Air
Force Research Laboratory (AFRL)-sponsored UNT Institute for
Science and Engineering Simulation (ISES) with Grant FA865008-C-5226. We also acknowledge the UNT Center for Advanced
Research and Technology (CART), UNT College of Engineering
for sponsoring A.T. as part of its SUPER program, and Anchal
Sondhi for assistance with XRD acquisition.

References
[1] C. Leyens, M. Peters, Titanium and Titanium Alloys, WileyVCH, Weinheim, 2003.
[2] S. Nag, R. Banerjee, Fundamentals of medical implant
materials, in: R. Narayan (Ed.), ASM Handbook, Mater.
Med. Devices, Vol. 23, ASM International, Materials Park,
OH, 2012.
[3] D. Banerjee, J.C. Williams, Acta Mater. 61 (2013) 844.
[4] C.H. Hager Jr., J. Sanders, S. Sharma, A. Voevodin, Wear
263 (2007) 430.
[5] K.G. Budinski, Wear 151 (1991) 203.
[6] P.J. Blau, B.C. Jolly, J. Qu, W.H. Peter, C.A. Blue, Wear 263
(2007) 1202.
[7] S. Samuel, S. Nag, T.W. Scharf, R. Banerjee, Mater. Sci. Eng.
C 28 (2008) 414.
[8] D.G. Bansal, O.L. Eryilmaz, P.J. Blau, Wear 271 (2011) 2006.
[9] M. Long, H.J. Rack, Biomaterialia 19 (1998) 1621.
[10] M. Geetha, A.K. Singh, R. Asokamani, A.K. Gogia, Prog.
Mater. Sci. 54 (2009) 397.
[11] X. Liu, P.K. Chu, C. Ding, Mater. Sci. Eng. R 47 (2004) 49.
[12] M. Niinomi, Mater. Sci. Eng. A 243 (1998) 231.
[13] R.A. Buchanan, E.D. Rigney, J.M. Williams, J. Biomed
Mater. Res. A 21 (1987) 355.
[14] C.B. Johansson, J. Lausmaa, T. Rostlund, P. Thomsen, J.
Mater. Sci. Mater. Med. 4 (1993) 132.
[15] C. Hu, H. Xin, L.M. Watson, T.N. Baker, Acta Mater. 45
(1997) 4311.
[16] A. Czyrska-Filemonowicz, P.A. Buat, M. Lucki, T. Moskalewicz, W. Rakowski, J. Lekki, et al., Acta Mater. 53 (2005)
4367.
[17] A. Zhecheva, W. Sha, S. Malinov, A. Long, Surf. Coat
Technol. 200 (2005) 2192.
[18] M. Nakai, M. Niinomi, T. Akahori, N. Ohtsu, H. Nishimura,
H. Toda, et al., Mat. Sci. Eng. A 486 (2008) 193.
[19] R. Banerjee, P.C. Collins, A. Genc, H.L. Fraser, Mater. Sci.
Eng. A 358 (2003) 343.
[20] E.F. Rauch, J. Portillo, S. Nicolopoulos, D. Bultreys, S.
Rouvimov, P. Moeck, Z. Kristallogr. 225 (2010) 103.

74

H. Mohseni et al. / Acta Materialia 83 (2015) 6174

[21] P. Moeck, S. Rouvimov, E.F. Rauch, M. Veron, H. Kirmse,


I. Hausler, et al., Cryst. Res. Technol. 46 (2011) 589.
[22] K.J. Ganesh, A.D. Darbal, S. Rajasekhara, G.S. Rohrer, K.
Barmak, P.J. Ferreira, Nanotechnology 23 (2012) 135702.
[23] H. Mohseni, P.C. Collins, T.W. Scharf, Nanometer Energy 1
(2012) 318.
[24] J.S. Carpenter, X. Liu, A. Darbal, N.T. Nuhfer, R.J.
McCabe, S.C. Vogel, et al., Scr. Mat. 67 (2012) 336.
[25] J.G. Brons, G.B. Thompson, JOM 66 (2014) 165.
[26] R. Vincent, P.A. Midgley, Ultramicroscopy 53 (1994) 271.
[27] C.S. Own, A.K. Subramanian, L.D. Marks, Microsc. Microanal. 10 (2004) 96.
[28] T. Borkar, S. Gopagoni, S. Nag, J.Y. Hwang, P.C. Collins, R.
Banerjee, J. Mater. Sci. 47 (2012) 7157.
[29] J.-P. Bars, D. David, E. Etchessahar, J. Debuigne, Metall.
Trans. A 14 (1983) 1537.

[30] R.S. Razavi, M. Salehi, M. Monirvaghe, G.R. Gordani, J.


Mater. Process. Technol. 203 (2008) 315.
[31] H.A. Wriedt, J.L. Murray, Bull. Alloy Phase Diagrams 8
(1987) 378.
[32] M. Long, H.J. Rack, Wear 249 (2001) 157.
[33] D.A. Rigney, Mat. Res. Innov. 1 (1998) 231.
[34] V.V. Stolyarov, L.Sh. Shuster, M.Sh. Migranov, R.Z. Valiev,
Y.T. Zhu, Mater. Sci. Eng. A 371 (2004) 313.
[35] P.Q. La, J.Q. Ma, Y.T. Zhu, J. Yang, W.M. Liu, Q.J. Xue,
et al., Acta Mater. 53 (2005) 5167.
[36] M. Wen, C. Wen, P.D. Hodgson, Y.C. Li, Trib Lett. 45 (2012)
59.
[37] N. Gao, C.T. Wang, R.J.K. Wood, T.G. Langdon, J. Mater.
Sci. 47 (2012) 4779.
[38] J.P. Hirth, D.A. Rigney, Wear 39 (1976) 133.

You might also like