You are on page 1of 12

See

discussions, stats, and author profiles for this publication at: http://www.researchgate.net/publication/257540562

Linear friction welding of a near- titanium


alloy
ARTICLE in ACTA MATERIALIA JANUARY 2012
Impact Factor: 4.47 DOI: 10.1016/j.actamat.2011.04.037

CITATIONS

DOWNLOADS

VIEWS

58

76

5 AUTHORS, INCLUDING:
Elvi Dalgaard

P. Wanjara

Pratt & Whitney Canada

National Research Council Canada

8 PUBLICATIONS 21 CITATIONS

118 PUBLICATIONS 532 CITATIONS

SEE PROFILE

SEE PROFILE

Xinjin Cao

John J. Jonas

National Research Council Canada

McGill University

155 PUBLICATIONS 1,236 CITATIONS

468 PUBLICATIONS 8,882 CITATIONS

SEE PROFILE

SEE PROFILE

Available from: Xinjin Cao


Retrieved on: 12 August 2015

Author's personal copy

Available online at www.sciencedirect.com

Acta Materialia 60 (2012) 770780


www.elsevier.com/locate/actamat

Linear friction welding of a near-b titanium alloy


E. Dalgaard a,b,, P. Wanjara b, J. Gholipour b, X. Cao b, J.J. Jonas a
b

a
Materials Engineering, McGill University, 3610 University Street, Montreal, QC, Canada
Aerospace Manufacturing Technology Centre, Institute for Aerospace Research, National Research Council Canada, 5145 Decelles Avenue,
Montreal, QC, Canada

Received 8 April 2011; accepted 18 April 2011


Available online 16 November 2011

Abstract
The linear friction welding (LFW) behaviour of near-b titanium alloy Ti5Al5V5Mo3Cr (Ti-5553) was investigated by varying the
processing conditions of frequency and axial pressure. The examined mechanical properties of the welded material included microhardness
and tensile properties. The maximum strains experienced by the material during LFW for each set of welding parameters were estimated
based on the process parameters and then evaluated using Aramis, a three-dimensional optical deformation measurement system. The
LFWed Ti-5553 was examined with electron backscatter diraction techniques to relate the texture and phase changes to the
thermomechanical conditions. Characterisation of the welds included analysis of the microstructural features of the weld region and
the thermomechanically aected zone in relation to the parent material.
Crown Copyright 2011 Published by Elsevier Ltd. on behalf of Acta Materialia Inc. All rights reserved.
Keywords: Linear friction welding; Ti-5553; Microstructure; Tensile properties

1. Introduction
Beta (b) and near-b titanium alloys are of increasing
interest in the aerospace industry due to their better formability and toughness as compared with the more common
a+b Ti6Al4V alloy. High-strength metastable b alloys
such as Ti-5553 have potential to replace steel as the preferred material for large components such as the landinggear truck beam on the latest generation of airframes [1].
Ti-5553 was introduced in 1997 by TIMET and has the
nominal composition 5 wt.% Al, 5 wt.% V, 5 wt.% Mo,
3 wt.% Cr, with the balance being Ti. Table 1 lists some
typical physical and mechanical properties of this alloy in
the solution treated and aged (STA) condition. The forging
behaviour of Ti-5553 is similar to that of Ti10V2Fe3Al
(Ti1023), though the higher b transus temperature of
the former (856 C) vs. the latter (800 C) allows higher
forging temperatures [2].
Corresponding author at: Materials Engineering, McGill University,
3610 University Street, Montreal, QC, Canada.
E-mail address: ecdalgaard@gmail.com (E. Dalgaard).

Weldability is a classic problem with Ti and its alloys.


The metal rapidly reacts with atmospheric gases in the molten state, requiring protective gas shielding in order to join
it successfully using fusion welding methods. The low thermal conductivity of Ti leads to longer weld times for low
energy density processes such as TIG welding. The long
weld times lead to slow cooling, requiring gas protection
for an extended time to protect the highly reactive Ti. High
energy density methods, such as laser welding or electronbeam welding, partially solve this problem by allowing a
rapid welding cycle (heating, melting and cooling) through
localised heat input, but still require protection (gas shielding for all areas above 350 C or vacuum) during the time
that the weld is molten. Clearly, Ti is a prime candidate for
the development and application of solid-state welding
methods.
Linear friction welding (LFW), a solid-state process
consisting of oscillating one part against a stationary part
while applying an axial load (see Fig. 1), eliminates the
necessity for a protective environment when welding, since
the material does not reach fusion temperatures. No axial
symmetry is required and the parts can have quite complex

1359-6454/$36.00 Crown Copyright 2011 Published by Elsevier Ltd. on behalf of Acta Materialia Inc. All rights reserved.
doi:10.1016/j.actamat.2011.04.037

Author's personal copy

E. Dalgaard et al. / Acta Materialia 60 (2012) 770780


Table 1
Physical and tensile properties of Ti-5553 in the STA condition [2].
Average b transus, C
Density, kg m3
Tensile elastic modulus, GPa
Compressive elastic modulus, GPa
Ultimate tensile strength (UTS), MPa
Yield strength (YS) @ 0.2%, MPa
Per cent elongation at fracture (%El.), with gauge length of 4D

856
4650
112
113
1236
1174
13

Frictional
Pressure

Stationary
(Top)
H Reciprocal
Motion (Bottom)
W

771

the materials are brought to rest after the desired


shortening has been attained. Once the materials have been
brought to rest and aligned, the axial pressure is increased
and the weld is consolidated [14].
Interest in manufacture and/or repair of aircraft components using emerging welding technologies of importance
to the aerospace industry has created an incentive for investigating the LFW of Ti-5553. This research work forms
part of a larger programme on the research and technology
development of LFW for the manufacture of near-a, ab
and near-b Ti alloys. Microstructure characterisation, texture analysis and mechanical property evaluation of the
welds were performed as part of the LFW process development trials to dene the optimum parametric window. The
intent of this paper is to report on the various ndings of
this research work on LFWed Ti-5553.
2. Experimental procedure

Horizontal
Fig. 1. Sample geometry and oscillation direction [6].

geometries involving curves. With heat generated directly


at the interface in friction welding, a high energy density
(low heat input) comparable to that developed in laser or
electron-beam welding can be achieved. This, combined
with the low thermal conductivity of Ti and its alloys, creates a very small heat-aected zone (HAZ) [3]. The process
was rst patented in 1969 [4,5], and in the early 1980s The
Welding Institute (TWI) demonstrated a working LFW
machine for metals. Since then the process has been developed for a variety of materials such as Ti alloys [68], titanium aluminides [9], and Ni-based superalloys [1013],
with a more detailed examination of the relationship
between parameters and properties. Nonetheless, to the
knowledge of the authors, little work has been done on
the LFW of b Ti alloys.
The LFW process can be divided into four distinct
phases [6]. The rst phase, known as the initial or contact
phase, begins with contact between the two pieces in order
to initiate the wear of surface asperities. The second phase,
known as the transition or conditioning phase, begins when
the large wear particles that were created during the rst
phase begin to be expelled from the interface. Frictional
heat creates a soft plasticized region that is no longer able
to support the applied axial load and begins to deform permanently. When moving into the third phase, known as the
equilibrium or burn-o phase, the ash begins to form [14].
The axial pressure is increased and oscillation continues as
in the prior phases. Frictional heat is conducted away from
the interface and the plastic zone develops further, extruding material away from the interface as ash. The last
phase is known as the deceleration or forge phase, where

The equipment used for welding was an MTS LFW process development system, comprised of two hydraulic actuators: the in-plane actuator that oscillates the lower work
piece horizontally; and the forge actuator that applies a
downward load through the top stationary workpiece.
The technical specications of the equipment are described
elsewhere [6]. Table 2 shows the experimental plan used in
this study as well as the calculated maximum strain rate
using the equation proposed by Vairis and Frost [15], the
measured welding time and the total maximum strain calculated from these two latter quantities. The baseline
(BL) values of fBL (frequency) and PBL (pressure) were
established based on optima reported for Ti6Al4V [6],
since there is little in the literature regarding the friction
welding of b or near-b Ti alloys and no specic parameters
have been published [16]. Forging pressure was maintained
at the same level as during oscillation in order to examine
the eect of oscillation on the growth of the HAZ without
the complication of expelling extra material during the
forging phase. In general the a and b phase content of a
Ti alloy is very pertinent to its properties and hence to
the selection of thermomechanical processing parameters.
In LFW, the welding behaviour is closely related to the
ow stress; in the case of Ti-5553, since the material is
already mainly b phase, the use of parameters which led
to ow in the b phase of the Ti6Al4V material [6] was felt
to be valid (since thermomechanical processing would
occur in the b phase in any case). Ti-5553 material in ingot
form was sectioned to obtain weld coupons of 13 mm in

Table 2
Samples prepared.
Sample Frequency
ID

Pressure

Estimated Welding Estimated


strain
time (s) max.
rate (s1)
strain

LFW-1 fBL
PBL
3.6
LFW-2 Low (0.6fBL) High (1.5PBL) 2.2

1.3
1.6

4.8
3.5

Author's personal copy

772

E. Dalgaard et al. / Acta Materialia 60 (2012) 770780

Fig. 2. Sectioning schematic for metallographic samples extracted from


welded coupons: WD, welding oscillation direction; TWD, direction
transverse to WD, in the plane of welding; ND, normal to the welding
plane.

width (W), 35 mm in height (H) and 26 mm in length (L),


as illustrated in Fig. 1. Prior to welding, the contact surfaces of the samples were ground and cleaned with alcohol.
LFWed samples were sectioned transverse to the oscillation direction through the weld zone (as shown in Fig. 2)
for metallographic and electron backscatter diraction
(EBSD) investigation (i.e. parallel to the plane formed by
the transverse and the normal directions). Conventional
polishing procedures were used for optical microscopy
and for preliminary preparation of the EBSD samples.
Final EBSD polishing was carried out in a vibratory polisher for 12 h. It is noteworthy that the material required
an extensive and sensitive polishing procedure, as electropolishing is not recommended in order to preserve the b
phase. For examination of the microstructure using optical
microscopy, etching was done using Krolls reagent.
Microstructural examination was then performed using
an inverted optical microscope (Olympus GX71) equipped
with digital image analysis software (AnalySIS Five). Backscatter imaging and EBSD mapping were performed at
20 kV on a Hitachi S-3000N VP-SEM equipped with an

Oxford (HKL) EBSD data acquisition system (polished


surface).
Microhardness was measured using a Struers Duramin
A300 machine with a fully automated testing cycle (stage,
load, focus, measure). A load of 300 g was applied using
a load cell with closed-loop circuit control, and hardness
proles were determined across the weld region using an
average of three measurements for each point, with an
indent interval of 0.2 mm and a dwell period of 15 s.
For each weld condition, three tensile specimens having
a standard sub-size geometry of 25 mm in gauge length,
6 mm in width and 4 mm in thickness were machined in
accordance with ASTM E8M-01. All specimens were tested
at room temperature using a 250 kN MTS 810 tensile
machine equipped with an Aramis 3-D deformation measurement system. Before executing tensile testing, each
sample was painted with a high-contrast random pattern
of black on a white background. The functionality of the
Aramis system depends on the quality of this speckle pattern. The quality of the pattern was veried before mechanical property evaluation to ensure strain recording along
the entire gauge length. After examination for pattern recognition, tensile property evaluation was conducted using
displacement control at a rate of 2 mm min1 up to the
yield point and then 8 mm min1 up to the rupture with
the Aramis system set at an acquisition rate of 2 frames
per second (fps).
3. Results and discussion
The microstructure of the as-received material revealed
by optical microscopy and scanning electron microscopy
(SEM) is shown in Fig. 3. At low magnication, extremely
large b grains averaging 100500 lm in diameter are visible. The SEM image at high magnication reveals an acicular microstructure within the b grains. One of the large b
grain boundaries can be seen crossing the corner of the
frame (Fig. 3b). This as-received microstructure is consistent with that expected for an ingot structure [1] without

grain boundary

(a)

(b)

Fig. 3. Parent material showing (a) a dark-eld image of the large equiaxed grains (Krolls reagent) and (b) high magnication backscatter electron image
of acicular substructure using compositional contrast (mirror polished).

Author's personal copy

E. Dalgaard et al. / Acta Materialia 60 (2012) 770780

773

Fig. 4. (a) As-welded LFW-1 showing (b) right and (c) left side ash cross-sections.

any secondary operations, such as thermomechanical


processing or solutionizing and/or ageing heat treatments.

structure that is characteristic of the as-received material


(Fig. 3b) can be seen clearly.

3.1. Microstructural evolution

3.2. Microtexture measurements

Visual inspection of the interface region in the LFWed


Ti-5553 shows an appreciable ash from all four sides of
the joint (Fig. 4a), suggesting that the weld is integral
[6,14]. The ash length was found to be larger in the direction of the oscillatory movement, i.e. parallel to the specimen length as compared to that along the specimen
width. However, unlike LFWed Ti6Al4V, which exhibited a series of ridges on the ash extruded in the direction
of the reciprocating motion, the LFWed Ti-5553 ash,
though rough on the outer surface, displayed no regular ripples under the process conditions examined in the present
work. As the ash layer consists of plastically deformed
material extruded during the welding process, the dierence
in the ow behaviours of the two alloys (a + b vs. near-b) is
almost certainly responsible for the dierence in ash morphology. While Ti6Al4V behaves in large part like an
anisotropic hexagonal close-packed material with very limited slip systems, Ti-5553 is for the most part a body-centred
cubic material with multiple slip systems activated even at
room temperature, and certainly at the temperature at
which ash is extruded. This leads to the very narrow
spread and strong directionality of the ash ow in
Ti6Al4V and the more vertical spread of the Ti-5553
ash.
The microstructure of a welded sample is illustrated in
Fig. 5. At high magnication using optical microscopy,
faint indications of grain boundaries were seen, but, overall,
the grains along the weld centre were not eectively revealed
through chemical etching. The low etching response is
attributed to the highly deformed and dynamically recrystallized microstructure in the weld zone. In Fig. 5b, the
region adjacent to the weld centre is presented; here a few
more equiaxed grains are visible, as well as some indication
of an acicular structure in certain regions between the
recrystallized grains. In the micrograph of Fig. 5c, representing the structure about 0.4 mm from the weld line, this
acicular structure is clearly revealed. Finally, in Fig. 5d, 1
mm from the weld line, the undeformed acicular micro-

To reveal the details of the weld microstructure, including delineation of the grain boundaries and phase divisions,
orientation data were obtained from the as-received and
LFWed samples. Fig. 6 depicts the inverse pole gure
and phase fraction map of the as-received material, revealing that grains are equiaxed and 100250 lm in diameter.
Also observed is that the a phase comprises approximately
3% of the surface, and by extension, volume of the material. This phase is mainly concentrated at grain boundaries.
In order to improve the accuracy of the phase fraction
data, points in the scan with a condence index below 0.2
were discarded before calculation.
In Fig. 7, an inverse pole gure map of sample LFW-1
can be seen. An image quality map is overlaid in order to
dene the grain boundary locations. Clearly, in the aswelded condition, this alloy undergoes complete dynamic
recrystallization in the weld zone (200 lm wide) under
the thermomechanical conditions employed, namely the
combination of straining at elevated temperatures and high
strain rates. Beyond the weld region, an abrupt transition
to a thermomechanically aected zone (TMAZ) consisting
of mixed recrystallized and deformed grains is observed,
followed by a further transition to the original large b
grains. While the large equiaxed and deformed grains display a variety of orientations, it can be seen that, within
the recrystallized zone, the grains are predominantly oriented with their h1 1 1i directions normal to the sample surface. This represents the weld oscillation direction. This
reorientation to a preferred texture may be due to the
alignment during oscillation of the primary slip system in
this phase, {1 1 0}h1 1 1i, to be parallel with the oscillation
direction.
Fig. 8 indicates an overall phase area fraction of 0.4% a
in sample LFW-1. The a fraction decreases to nearly none
(estimated at 0.05%) in the recrystallized weld zone. This is
in contrast to the 3% a observed in the as-received material,
and indicates that the fast cooling rate experienced by the
material did not permit an equilibrium phase fraction to

Author's personal copy

774

E. Dalgaard et al. / Acta Materialia 60 (2012) 770780

(a)

(b)

grain boundary

(c)

(d)

Fig. 5. LFW-1 (Krolls reagent): (a) weld centre at high magnication, (b) 100 lm from weld centre, (c) 400 lm from weld centre, and (d) 1 mm from weld
centre.

Fig. 6. EBSD maps of as-received Ti-5553: (a) inverse pole gure map; (b) phase fraction map.

develop and metastable b was retained in preference to the


formation of a. Once again, based on the accuracy of the
TSL OIM Analysis software suggested in the literature
[17,18], points with a condence index below 0.2 were
removed from the data set. It was observed that these
points were nearly always indexed as a, due perhaps to
the large number of lines available in the pattern, but that

manual examination of the Kikuchi pattern and the index


assigned to it led to the conclusion that the indexing was
erroneous.
The width of the recrystallized zone in sample LFW-2
(see Figs. 9 and 10) averages 380 lm at the narrowest point
(weld centre), approximately 50% wider than in LFW-1.
Once again, the recrystallized zone consists almost entirely

Author's personal copy

E. Dalgaard et al. / Acta Materialia 60 (2012) 770780

HAZ

TMAZ

Weld Center

TMAZ

775

HAZ

240 m

Fig. 7. LFW-1 image quality combined with inverse pole gure.

Fig. 8. Phase fraction map of LFW-1; points with condence index below 0.2 removed.

TMAZ

Weld Center

Fig. 9. Inverse pole gure of LFW-2 showing one side of weld zone with adjacent deformed grains.

of grains oriented with the h1 1 1i direction normal to the


surface. A close-up scan of this recrystallized region can
be seen in Fig. 11. The phase fraction map conrms the
earlier nding in the lower-resolution scan that there is little to no a present in these grains, as compared with the
TMAZ, which contained 0.51 vol.% a.

In Fig. 12, b pole gures are presented for the recrystallized regions of LFW-1 and LFW-2. No a pole gures are
presented due to the negligible amount of a phase observed
in the material. The texture intensity of these recrystallized
zones can be seen to be very high, approximately 1015
times random. This strong texture is based on a large

Author's personal copy

776

E. Dalgaard et al. / Acta Materialia 60 (2012) 770780

Fig. 10. LFW-2 phase map; step size 2 lm; points with condence index below 0.2 removed.

Fig. 11. High-resolution scan of equiaxed grain zone at step size 0.10 lm: (a) inverse pole gure combined with image quality map; (b) phase map.

number of small grains and is quite reliable. As expected


from the inverse pole gure maps, many of these recrystallized grains are aligned along the h1 1 1i direction. Due to
the very large grain size in the unrecrystallized regions of
the welded specimens, the textures will need to be measured
over larger scanned regions in order to be statistically
signicant.
3.3. Mechanical testing
The microhardness proles shown in Figs. 13 and 14
reveal that the weld region is somewhat softer than the
surrounding TMAZ, which in turn is softer than the surrounding parent material; this is consistent with the micro-

structural observations of a depletion in the weldment.


Sample LFW-1, welded at lower pressure and higher frequency, can be seen to display a pronounced hardness drop
in the weld with a sharp drop over the range 1.6 mm.
Meanwhile LFW-2, the higher-pressure/lower-frequency
sample, displays less softening in the weld centre and a
slightly less abrupt drop that, however, spans a wider region
as indicated by comparison with the parent metal hardness
range. It is noteworthy that as the indentation spacing
(200 lm to give ve indentation widths from centre to centre) is comparable to the weld width, more detailed analysis
of the eect of processing on the hardness distribution would
require further investigation with higher-resolution methods
such as nanoindentation. Nevertheless, the widths of the

Author's personal copy

E. Dalgaard et al. / Acta Materialia 60 (2012) 770780

777

Fig. 12. b Pole gures of the recrystallized zones in (a) LFW-1; (b) LFW-2. The welding direction (WD) is normal to the gure (not shown); TD indicates
the direction transverse to WD, and ND the direction normal to the welding plane.

Fig. 13. Microhardness prole across the weld line for sample LFW-1
(baseline pressure and frequency).

Fig. 14. Microhardness prole across the weld line for sample LFW-2
welded at low frequency and high pressure.

softened regions in the two samples are consistent with the


weld zone widths observed in the EBSD scans given the resolution of the hardness proles.
The drop in hardness is readily explained by the fact that
no post-weld heat treatment was performed on these samples. The solution treatment of Ti5553 is usually carried

out below the b transus, so that some globular a remains,


much of it on grain boundaries. Without this treatment,
the strengthening and grain boundary pinning eects (as
well as the detrimental eect on fracture toughness) of
the globular a that would have formed are absent.
Although an increase in strength in the weld region might
have been possible due to the grain rening eects of heavy
deformation combined with recrystallization at elevated
temperature, this increase was not observed, perhaps
obscured by the greater eect of softening due to the loss
of a.
Tensile testing was performed on both the as-received
and the welded specimens. The as-received values in Table
3 are for the material that was not heat-treated, and are
thus lower than literature values quoted for solutionized
and aged Ti-5553. The large variation in the %El. observed
is consistent with previous ndings for a billet structure
without STA, and Fanning has reported [2] lower variation
in the elongation when the microstructure of Ti-5553 is
homogenised using STA. As compared to the as-received
tensile properties, the YS, UTS and %El. values are lower
for the LFWed samples. The reduction in tensile strength
of the welded samples can be explained by the presence
of a softened region in the weldment with a microstructure
depleted of a. In addition, there is a distinct dierence in
tensile properties between sample LFW-1, welded at BL
axial pressure and frequency, and sample LFW-2, welded

Table 3
Tensile testing results.

As-received
LFW-1
LFW-2

YS (MPa)

UTS (MPa)

Uniform El. (%)

Total El. (%)

1046 13
1019 19
988 16

1108 25
1058 23
1013 10

7.4 3.0
3.0 0.5
2.0 0.1

11.2 6.5
4.0 1.0
2.9 0.9

Author's personal copy

778

E. Dalgaard et al. / Acta Materialia 60 (2012) 770780

(a)

Fracture surface
WZ

Fracture surface

WZ

Fig. 15. SEM images of fracture surfaces: (a) as-received, low magnication; (b) welded, low magnication; (c) as-received, high magnication; (d) welded,
high magnication.

(b)

Fig. 16. Polarized light micrographs of fractured tensile bars: (a) LFW-1
and (b) LFW-2.

at higher pressure and lower frequency, with LFW-1 displaying higher YS and UTS as well as greater elongation
(see Table 3). The better mechanical performance of
LFW-1 as compared to LFW-2 can be attributed to the
smaller amount of material aected thermomechanically
(weld zone and TMAZ), as indicated by the EBSD maps.
The fracture surfaces of the welded tensile coupons
revealed very large grains (see Fig. 15) that are consistent
with fracture occurring in the TMAZ adjacent to the
recrystallized weld zone. Both unwelded and welded samples displayed ductile fracture characteristics with some
areas of shear. In fact, all welded specimens fractured in
the TMAZ within 1 mm of the weld zone, as shown in
Fig. 16. Conversely, as-received (unwelded) tensile specimens fractured at random locations along their gauge
lengths. The consistent failure of the welded samples in
the TMAZ is most likely due to a depletion in the
coarse-grained microstructure, which would have a tendency to be weaker than the similarly depleted ne-grained
recrystallized weld zone.
An example of the strain distribution in the unwelded
condition is presented in Fig. 17. This should be compared
with that of the welded LFW-1 in Fig. 18. The region of
strain concentration indicated as the area of highest intensity in the welded specimen corresponds to the location of
the fracture that occurred immediately afterward. While it

Author's personal copy

E. Dalgaard et al. / Acta Materialia 60 (2012) 770780

779

Weld Line

TMAZ

TMAZ

Fig. 17. Strain distribution in as-received Ti-5553 just before fracture.

PM

PM

Fig. 18. Strain distribution in LFW-1 just before fracture.

can be seen that in the weld-free sample the average strain


is approximately 12%, rising to about double that at the
eventual fracture location, the localisation is much more
pronounced in the welded specimen, with an average strain
in the majority of the material of only 2%. This rises sharply to 5% just outside the TMAZ and thence steadily but
rapidly to the maximum of 9.3%, almost ve times the
average, at the fracture location adjacent to the weld line.
The resolution of the strain distribution image and the difculty of accurately positioning the weld line (which is not
altogether straight) in the centre of the specimen leads to
some ambiguity in Fig. 18 as to the exact position of the
fracture with respect to the weld zones. However, the fracture micrographs in Fig. 16 show clearly that the fracture
occurred next to the weld line in the TMAZ.
Both microhardness analysis and tensile evaluation of
the welded specimens lead to the conclusion that a softened
region exists at and on either side of the weld line. This soft
region, depleted of the strengthening a phase, as revealed by
the microstructural and textural examinations, represents
an area where the strain concentrates. An object is strongest
when the strain is evenly distributed over its area. Here, due
to the weld zone with its corresponding a phase depletion
and grain renement, and the TMAZ with only a phase
depletion, there is strain concentration caused by the inhomogeneity of the strengths of these dierent regions, and
therefore in the ow stress; this results in a localised increase
in the strain. Of future interest for welded Ti-5553 is the use
of a post-weld heat treatment to promote the precipitation
of the strengthening a phase. The present study indicates

that a solutionizing and ageing cycle would be benecial


in improving the properties of the weld.
4. Conclusions
Near-b alloy Ti5Al5V5Mo3Cr alloy was welded
using an MTS LFW process development system. Process
conditions were varied in order to examine the relationship
between process parameters and the microstructural, textural and mechanical properties. The following conclusions
can be drawn:
1. The as-received material displayed a large equiaxed b
grain structure of about 100500 lm diameter grains.
These grains contained an acicular substructure. The a
phase in the as-received material comprised approximately 3% of the total, mainly concentrated at grain
boundaries.
2. After welding, a very ne-grained (15 lm diameter)
recrystallized zone was observed in the weld centre,
ranging in width from 240 to 380 lm for the process
conditions tested. Within this recrystallized zone, the
grains were almost all oriented with their h1 1 1i directions normal to the sample surface, i.e. parallel to the
weld oscillation direction.
3. The TMAZ was observed to consist of deformed large b
grains with some recrystallization localised at the grain
boundaries. Also, less than 1% volume fraction of a
phase was observed in the TMAZ and weld zone following welding.

Author's personal copy

780

E. Dalgaard et al. / Acta Materialia 60 (2012) 770780

4. Microhardness testing revealed a softened area within


2 mm of the weld centre line.
5. Fracture of the welded specimens during tensile testing
occurred in the TMAZ within 1 mm of the recrystallized
weld zone.
6. The strain in the fracture zone of the welded specimens
was approximately ve times the average strain in the
tested specimen. By contrast, in the as-received material,
the strain in the vicinity of the fracture was about double
the average.

Acknowledgements
The authors are grateful to Standard Aero Limited for
the materials used in this study. Thanks are also due to
M. Guerin, X. Pelletier and D. Chiriac for extensive technical assistance.
References
[1] Jones NG, Dashwood RJ, Dye D, Jackson M. Metall Mater Trans A
2009;40A:194454.
[2] Fanning JC. J Mater Eng Perform 2005;14(6):78892.

[3] Wilhelm H, Furlan R, Moloney KC. Titanium 95: science and


technology. London: Institute of Materials; 1996. p. 6207.
[4] Threadgill P. Knowledge summary, TWI. <http://www.twi.co.uk/
content/ksplt001.html>.
[5] Caterpillar Tractor Co., inventor. Improvements in bonding metal
parts involving their rubbing contact. UK patent GB1161800 A; 1969.
[6] Wanjara P, Jahazi M. Metall Mater Trans A 2005;6A:214964.
[7] Dalgaard E, Coghe F, Rabet L, Jahazi M, Wanjara P, Jonas JJ. Adv
Mater Res 2009;8991:1249.
[8] Dalgaard E, Wanjara P, Gholipour J, Jonas JJ. Linear friction
welding of a forged near-a titanium alloy. Mater Sci Forum, in press.
[9] Godfrey S, Strangwood M, Threadgill P. Linear friction welding of a
gamma titanium aluminide alloy. In: SF2M/ONERA conference,
Paris; 1996.
[10] Mary C, Jahazi M. Adv Eng Mater 2008;10(6):5738.
[11] Chamanfar A, Jahazi M, Gholipour J, Wanjara P, Yue S. Metall
Mater Trans A 2010;42:72944.
[12] Karadge M, Preuss M, Withers PJ, Bray S. Mater Sci Eng A
2008;491:44653.
[13] Wanjara P, Dalgaard E, Gholipour J, Larose J. Linear friction
welding of a single crystal superalloy. Mater Sci Forum, in press.
[14] Vairis A, Frost M. Mater Sci Eng 2000;A292:817.
[15] Vairis A, Frost M. Wear 1998;217(1):11731.
[16] Attallah M, Preuss M, Bray S. Trends in welding research. In:
Proceedings of the 8th international conference; 2009. p. 48691.
[17] OIM Version 3.0. Online help. Draper, UT: TSL; 2001.
[18] Sharma H, van Bohemen SMC, Petrov RH, Sietsma J. Acta Mater
2010;58:2399407.

You might also like