You are on page 1of 9

Materials Science and Engineering A 540 (2012) 226234

Contents lists available at SciVerse ScienceDirect

Materials Science and Engineering A


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

Residual stress and deformation mechanism of friction stir welded aluminum


alloys by nanoindentation
C.A. Charitidis a, , D.A. Dragatogiannis a , E.P. Koumoulos a , I.A. Kartsonakis a,b
a
b

National Technical University of Athens, School of Chemical Engineering, 9 Heroon Polytechniou Str., Zografos, Athens GR-157 80, Greece
Sol-Gel Laboratory, IMS, NCSR DEMOKRITOS, 15310 Agia Paraskevi, Greece

a r t i c l e

i n f o

Article history:
Received 19 November 2011
Received in revised form 30 January 2012
Accepted 31 January 2012
Available online 8 February 2012
Keywords:
Nanoindentation
Aluminum alloys
Friction stir welding
Residual stress

a b s t r a c t
Residual stress during Friction Stir Welding (FSW) of lightweight alloys is of major concern, due to their
functionality and applications in transport and industry elds. Several conventional techniques which
are used to measure and characterize welded aluminum alloys are destructive. This drawback has raised
nanoindentation as the non-destructive alternative technique with many advantages, such as easy preparation and high spatial resolution. In this study a methodology was brought forward to investigate the
applicability of nanoindentation method in order to overcome limitations of this technique and measure
residual stress in two of most commonly used aluminum alloys in transport and industrial applications.
2012 Elsevier B.V. All rights reserved.

1. Introduction
In transport and industrial applications the lightweight alloys
undergo signicant mechanical loads under different applied loads
and time dependent conditions. More specically, high-demanding
mechanical properties and high accuracy to strength-to-weight
ratio of such materials have led to an increased research. Aluminum
alloys offer a high potential for weight reduction in automotive
and other transportation vehicle construction. Aluminum alloy
AA6082-T6 (AA: provided by the Aluminum Association) is a
high-strength AlMgSi alloy containing manganese to increase
ductility and toughness. The T6 condition is obtained through
articial ageing at a temperature of 170200 C, mainly for
welding applications [1]. Aluminum alloy AA5083-H111 is an
AlMg alloy which can be classied as wrought alloy product with excellent corrosion resistance. The H111 condition is
obtained through both cold hardening and partial annealing. In
particular, both alloys are often applied in shipbuilding industry
[2].
The increasing relevance of aluminum alloys in transportation
requires research on more efcient and reliable joining processes.
Friction Stir Welding (FSW) is a solid-state joining process, which
emerged as an alternative technique to be used in high strength
alloys that were difcult to join with conventional joining techniques. This welding process was initially developed for aluminum

Corresponding author. Tel.: +30 2107724046; fax: +30 2107722339.


E-mail address: charitidis@chemeng.ntua.gr (C.A. Charitidis).
0921-5093/$ see front matter 2012 Elsevier B.V. All rights reserved.
doi:10.1016/j.msea.2012.01.129

alloys, but since then FSW has also been utilized on joining a large
range of materials. In FSW the interaction of a non-consumable and
rotating tool with the work pieces being welded creates a welded
joint through frictional heating and plastic deformation at temperatures below the melting temperature of the alloys being joined.
The contacting shoulder applies frictional heat to the weld region
and prevents highly plasticized material from being expelled during the welding operation. The combined frictional heat from the
pin and the shoulder creates a plasticized condition around the
immersed pin and the contacting surface of the shouldered region
of the work piece top surface. Material ows around the tool and
coalesces behind it while relative traverse between substrate and
the rotating tool occurs. Notwithstanding the widespread interest
in the possibilities offered by FSW, data concerning the mechanical behavior of joints obtained using this process is still scarce.
The heat input (most of it generated by the friction between
tool shoulder and materials to be joined) increases by increasing tool rotational speed and by decreasing welding speed. Low
heat input causes the intermittent material ow and improper stirring action around the tool pin due to insufcient plasticization
of the materials under the tool shoulder. On the other side, high
heat input causes turbulent material ow around the tool pin due
to excess plasticization of material under the tool shoulder. Further details and analysis of the process are reported elsewhere
[3].
The frequent failure of structural components by time dependent degradation in severe operating environments has recently
become a concern. In particular, pipeline operating conditions are made more severe by cryogenic contents and many

C.A. Charitidis et al. / Materials Science and Engineering A 540 (2012) 226234

inhomogeneous welded joints, so that the welded joint is typically


the initiation point of fracture because of the microstructural and
mechanical inhomogeneities [4]. Thus safety assessment based
on precise mechanical properties of local regions is indispensable.
One of the challenges in studying the local mechanical properties
in a small volume of welded materials having microstructural
gradients is that the traditional methods, e.g. tensile test and
microhardness test, need bulky standard samples to evaluate
the mechanical behavior of materials and as a consequence are
not applicable, contrary to nanoindentation. In general conventional nondestructive techniques for welding residual stress
measurement have many disadvantages in this eld because of
poor repeatability, large scatter in data, complex procedures,
inaccurate results, etc. [5]. For instance, it is almost impossible
to use non-destructive methods to assess residual stresses in
weldments including heat-affected zones (HAZs), which have
very rapid microstructural gradients [4]. Also materials such as
high strength steel (e.g. HQ130 steel) are highly amenable to
crack formation during welding so with the normally available
instrumentation and measuring facilities, it is quite difcult to nd
out the instantaneous residual stress in the weld zone [6].
OliverPharr (O&P) analysis for nanoindentation has proven to
be an effective and convenient method of determining the mechanical properties of solids, most notably elastic modulus (E) and
hardness (H) [7]. The method relies on the analysis of the initial slope of the unloading load-displacement response which is
assumed to be elastic, even if the contact is elasticplastic.
Recently, residual stress determination using the analytical
methods based on nanoindentation technique has attracted extensive attention. A general methodology to obtain residual stress from
nanoindentation based on the analysis of contact area difference
between a residual stress-free material and the same material with
residual stress is proposed from Suresh and Giannakopoulos [8]. A
stress-analysis technique for sharp nanoindentation by considering
the stress interaction between the residual stress and the contact
pressure from the viewpoint of the shear plasticity is explored by
Lee and Kwon [9]. It is also known for various alloys, that internal uniaxial or biaxial stresses affect nanomechanical measurement
results, proved experimentally as well as by nite element simulation [10].
Residual stress plays an important role in mechanics of materials as it is a key quantity for the mechanical behavior and lifetime
of constructions, especially for welded materials in which welded
joint is related with the fatigue crack propagation (residual stress
state can have a large impact on the lifetime of the weld). The
evaluation of residual stress based on nanoindentation methods is
essential, as the instrumented indentation is widely used to probe
mechanical properties of small scales and further mechanical characterization. In order to estimate the effect of residual stresses and
develop models for its description, to the comparison of indentation depth, indentation force, unloading behavior, contact area,
pile-up or hardness with and without residual stress is essential.
For all the aforementioned reasons evaluation of residual stress on
welded structures by nanoindentation poses a complex mechanical
system.
This work emphasizes on analysis of the precise determination of true contact area during nanoindentation which is crucial
to making accurate residual stress determination. The common
used O&P method assumes that the contact periphery between
indenter and material sinks in, which limits its application to
materials that pile-up. It is clearly shown through Scanning Probe
Microscopy (SPM) that pile-up during nanoindentation of AA 5083H111 alloy is negligible in contrast with AA6082-T6, where the
resultant pile-up is signicant. The residual stresses in AA 6082T6 and AA5083-H111 aluminum alloys were determined by the
nanoindentation method.

227

Fig. 1. Geometry of FSW process, also indicating the tool transverse direction and
cross section region.

2. Experimental details
2.1. Materials used and welding conditions
Single-pass friction stir square butt joint welds were produced
using a tool (Fig. 1) made from heat treated steel; the necessary
clamping arrangement of the specimens to be welded was also
designed and manufactured. The welding direction was parallel
to the rolling direction of the plates whereas no inclination angle
concerning the welding tool was used. The welding process was
performed with rotating of the tool at 375 rpm and at a feed rate of
85 mm/min. These optimum conditions were obtained from a large
number of welding procedures [11].

2.2. Microstructural characterization


The chemical composition of AA6082-T6 (high SiMg containing
aluminum alloy) and AA5083-H111 (high Mg containing aluminum
alloy) are listed in Table 1, determined by SEM using a PHILIPS
Quanta Inspect (FEI Company) microscope with W (tungsted) lament 25 kV equipped with EDAX GENESIS (Ametex Process and
Analytical Instruments).
Qualitative analysis of the microstructure components of
AA6082-T6 alloy was performed by X-ray diffraction (XRD) using
a Powder diffractometer (SIEMENS D-500 equipped with a CuK
The results of XRD analysis in
lamp with wavelength of 1.5418 A).
diffractogram form is shown in Fig. 2(a) and (b).
For as-cast magnesium alloy, the coarse -Mg phase and Al12 Mg17 intermetallic compound disappeared after FSW [1214].
Further, Nakata et al. [12], Lee et al. [14], and Park et al. [15] reported
that the grain size in the weld nugget became larger with increasing
tool rotation rate and decreasing traverse speed due to increasing
heat input, which promoted the growth of recrystallized grains.
This observation is consistent with that in FSW aluminum alloys
[1621].

Table 1
Composition of both aluminum alloys in weight percent through EDAX analysis.
Element

AA6082-T6 (%, w/w)

AA5083-H111 (%, w/w)

Si
Fe
Cu
Mn
Mg
Zn
Ti
Cr
Al

1.13
0.34

0.30
0.56

97.67

0.22
0.26
0.61
4.95
0.20

0.14
93.63

228

C.A. Charitidis et al. / Materials Science and Engineering A 540 (2012) 226234
1000

(a)

Al<200>

800

AA 6082-T6
base metal

2 4
23
3 5 6
5 7

Load ()

1000
750

AlFeSi

Al6Mn
50

60

70

80

3 (-3) cm (1)
-1.5 cm (2)
-1 cm
(3)
-0.8 cm (4)
0 cm
(5)
0.5 cm (6)
1.3 cm (7)

1250

500
250

0
40

2 cm

1500

-Mg2Si
-Al12(Fe,Mn)3Si

Intensity

200

2000
1750

600

400

Retreating side (+)

Advancing side (-)


Al<111>

(a)

90

100

110

2 Theta / degree

20

40

60

80

100

120

140

160

180

200

180

200

Displacement (nm)
Al<200>

1000

Al<111>

(b)

(b)

800

AlFeSi

Retreating side (+)

4 5 6

23

2250
2000

20 mm

1750

Load (N)

-Mg2Si
-Al12(Fe,Mn)3Si

-15.5 mm (1)
-10 mm (2)
- 8mm (3)
0 mm
(4)
4 mm
(5)
11 mm (6)

1500
1250
1000

Al6Mn

Intensity

600

200

2500

AA 6082-T6
welding zone

400

Advancing side (-)

2750

750
500

250
40

50

60

70

80

90

100

110

2 Theta / degree

20

40

60

80

100

120

140

160

Displacement (nm)
Fig. 2. XRD analysis in diffractogram form of base metal and welding zone for
AA6082-T6.

3. Analytical modeling
3.1. Nanoindentation analysis
Nanoindentation testing was performed with Hysitron TriboLab Nanomechanical Test Instrument. Details about the instrument
and the experimental setup have been presented elsewhere [2].
The surface of the samples was prepared by grinding and polishing
(with 1 and 0.1 m agglomerated -alumina suspension). Based

Compressive
stress state

Stress free state

Load ()

LC

Tensile
stress state
L0

Fig. 4. Nanoindentation load-displacement curves at 200 nm indentation depth for


(a) the AA 6082-T6 aluminum alloy and for (b) the AA 5083-H111 aluminum alloy
(transverse). Hatched area indicates the welding zone.

on the half-space elastic deformation theory, H and E values can be


extracted from the experimental data (load displacement curves)
using the O&P method [7], where derived expressions for calculating the elastic modulus from indentation experiments are based on
Sneddons elastic contact theory [22]:

S 
Er =
(1)

2 Ac
where S is the unloading stiffness (initial slope of the unloading
load-displacement curve at the maximum displacement of penetration (or peak load)), Ac is the projected contact area between the
tip and the substrate and is a constant that depends on the geometry of the indenter ( = 1.167 for Berkovich tip) [3,4]. Conventional
nanoindentation hardness refers to the mean contact pressure; this
hardness, which is the contact hardness, (Hc ), is actually dependent
upon the geometry of the indenter (Eqs. (2)(4)).
Hc =

F
Ac

(2)

where

LT

A(hc ) = 24.5hc + a1 hc + a1/2 hc

1/2

+ + a1/16 hc

1/16

(3)

and

Displacement (nm)
Fig. 3. Typical nanoindentation load-displacement curves for different stress states.

hc = hm

Pm
Sm

(4)

C.A. Charitidis et al. / Materials Science and Engineering A 540 (2012) 226234

Retreating side (+)

Advancing side (-)

(a)

2500
2250

1 2

2000

45

1750

Load (N)

229

2 cm
-2 cm
-1 cm
0 cm
1 cm
1.2 cm
3 cm

1500
1250
1000

(1)
(2)
(3)
(4)
(5)
(6)

750
500
250
0
0

20

40

60

80

100

120

140

160

180

200

Displacement (nm)

(b)

Retreating side (+)

Advancing side (-)

2000

2 3 4 5 6

1750
1500

20 mm
Load ()

1250

-15.5 mm
-10.5 mm
-8.5 mm
0 mm
4.5 mm
8.5 mm

1000
750

(1)
(2)
(3)
(4)
(5)
(6)

500
250
0
0

20

40

60

80

100

120

140

160

180

200

Displacement (nm)
Fig. 5. Nanoindentation load-displacement curves at 200 nm indentation depth for (a) the AA 6082-T6 aluminum alloy and for (b) the AA 5083-H111 aluminum alloy (cross
section). Hatched area indicates the welding zone.

where hm is the total penetration displacement of the indenter at


peak load, Pm is the peak load at the indenter displacement hm ,
and is an indenter geometry constant, equal to 0.75 for Berkovich
indenter [23].
3.2. Residual stress measured by nanoindentation
A surface residual stress is assumed to be in an equibiaxial tensile state and uniform in the near-surface region (taken as about
three times the indentation depth) [8,9]. If an arbitrary indentation
state (ht , L0 ) is attained in an unstressed state and if the tensile
inplane stress res is applied to the loading state at a xed penetration depth ht , the indentation load L0 is reduced to a load LT
due to the decrease in surface penetration resistance (in Fig. 3 typical nanoindentation load-displacement curves for different stress
states are presented). The stress tensor is separated into mean
stress  M (hydrostatic stress) and plastic-deformation-sensitive

shear deviator stress  D . The surface-normal deviator stress in zdirection (normal to the surface of indentation) zD is 2res /3 by
removing the hydrostatic stress 2res /3 from the surface residual stress res and is added to the contact pressure [9]. LT L0 is
dened as a product of the selected deviator stress component and
its corresponding contact area ATC . Thus, an equation for the equibiaxial residual stress is derived in terms of the indentation load and
contact area as:
res = 3

(L0 LT )
2ATC

(5)

Here, ATC in the tensile stress state is calculated from LT A0C /L0
because the contact hardness Hc or L0 /A0C = LT /ATC is independent
of the elastic residual stress. This measurement is based on the initial elastic unloading part of the load-displacement curve which is
unaffected by the residual stresses for materials with isotropic or

230

C.A. Charitidis et al. / Materials Science and Engineering A 540 (2012) 226234

(a)

1000

(a)

Welding zone

AA6082-T6

Welding zone

AA6082-T6

Retreating side

Advancing side

Retreating side

Advancing side

800

400

Residual Stress (MPa)

Residual Stress (MPa)

600

600

400

200

200

-200

0
-400

-6

-5

-4

-3

-2

-1

-4

-3

-2

3000

AA5083-H111
2500

(b)

Welding zone

AA5083-H111

Welding zone

Retreating side

Advancing side

2000

Residual Stress (MPa)

1000

Retreating side

Advancing side

Residual Stress (MPa)

(b)

-1

Distance from weld center (cm)

Distance from weld center (cm)

1500
1000
500

500

-500

0
-500

-1000

-24 -20 -16 -12

-8

-4

12

16

20

24

-24

Distance from weld center (mm)

no strain hardening [8]. In order to measure the actual contact area


A0C with pile-up or sink-in information from an O&P curve analysis,
prior to indentation, an empirical calibration for instrumental stiffness and the area function of the indenter tip was performed by
making preliminary indentation tests on a fused silica, a standard
material for this purpose [3].
This model presented below is applicable in the specic stress
state of equibiaxial tensile or compressive state. This impedes wide
application of the instrumented indentation technique to complex
biaxial stress states in actual structures [24]. In order to analyze
general stress states this model needs modication. If we denote
one major stress component of the biaxial residual stress as res,x
and the other as a minor stress component res,y then res,y can
be expressed as pres,x using the stress ratio p or res,y /res,x . The
inuence of biaxial stress on the indentation plasticity also can
be analyzed through a similar hydrostatic-stress-removal method.
The deformation -sensitive deviator stress component is given as
(1 + p)res,x
3

(6)

in this case. Thus, if information on p is given, individual principal


stresses can be calculated from the instrumented indentation test
using (7).
res,x =

3(L0 LT )
(1 + p)ATC

-16

-12

-8

-4

12

16

20

24

Distance from weld center (mm)

Fig. 6. Residual stress distribution for (a) the AA 6082-T6 aluminum alloy and for (b)
the AA 5083-H111 aluminum alloy (transverse). Hatched area indicates the welding
zone.

zD =

-20

(7)

Lee and Kwon [25] showed the validity of (5) by empirical indentation tests on biaxially strained specimens. The stress ratio became

Fig. 7. Residual stress distribution for (a) the AA 6082-T6 aluminum alloy and for
(b) the AA 5083-H111 aluminum alloy (cross section). Hatched area indicates the
welding zone.

an important issue in stress characterization, and Underwoods


observation of a nonsymmetrical contact deformation in uniaxial
stress [26] supplied a clue to extracting the surface stress directionality. Lee et al. [27] tried to estimate the stress ratio by analyzing
the pile-up heights along the two principal biaxial stress axes. The
ratio of stress-induced pile-up shifts was linearly proportional to
the stress ratio. Several preliminary observations on the welded
joints yielded a stress ratio of about 0.33, and this value was used
in subsequent biaxial stress analyses [28].

4. Results and discussion


4.1. Residual stress distribution across the friction stir-welded
joint
According to the geometry of the welding process, the largest
stresses are expected parallel to the welding direction and close
to the weld zone. As a result, residual stresses measurements are
focused on this region [29]. The instrumented nanoindentation
curves of AA 6082-T6 and AA 5083-H111 for the different regions
across the friction stir welded joint including the HAZ, the thermo
mechanical affected zone (TMAZ), the base metal (BM) and the
reference curve far away from the welded joint is presented in
Figs. 4 and 5, for transverse and cross section respectively. The load
decreased due to surface stress at given indentation depth means
that the residual stress has positive sign (tensile). On the other hand,

C.A. Charitidis et al. / Materials Science and Engineering A 540 (2012) 226234
120

Welding zone

AA6082-T6

100

80

90

60
40
2.0

Hnano (GPa)

AA5083-H111
AA6082-T6
y=55.25x-6.71

110

100

Hmicro (HV0.3)

Hmicro (HV0.3)

(a)

231

1.8

80
70

1.6

60

1.4
1.2

50

1.0
-5

-4

-3

-2

-1

Distance from weld center (cm)

Hnano (GPa)

Hmicro (HV0.3)

(b)

90

1.0

1.2

1.4

1.6

1.8

2.0

2.2

Hnano (GPa)

Welding zone

AA5083-H111

Fig. 9. Distribution of micro- and nano-hardness for both alloys [43].

80
70
60
1.8
1.6
1.4
1.2
1.0
0.8
0.6
0.4
0.2
0.0
-16 -14 -12 -10 -8

-6

-4

-2

10 12 14 16

propagate downward and toward the surface as the loading continues and then the crack becomes unstable [32].
In the measured residual stresses shown in Fig. 6(a) for AA6082T6 the stresses appear to rise to a tensile stress of 800 MPa at
1 cm from the center, before reversing in trend to form a tensile region near the weld line of about 600 MPa. It can be seen
that this tensile region peaks at around 1 cm from the weld line,
a distance corresponding to the edge of the tool shoulder, before
dropping slightly at the weld centre as it is shown in Fig. 6(a). Also,
in the measured residual stresses shown in Fig. 6(b) for AA5083H111 the stresses from the parent material to the welding zone
appear to decrease linearly to a compressive stress steady state of

Distance from weld center (mm)


Fig. 8. Transverse distribution (comparison) of Hmicro and Hnano for both alloys (a)
AA 6082-T6 and (b) AA 5083-H111. Hatched area indicates the welding zone [43].

the load increased due to surface stress at given indentation depth


means that the residual stress has negative sign (compressive).
Fig. 6 shows the transverse stresses within the weld as a function of distance from the weld line. The distribution of residual
stresses measured by nanoindentation is in agreement with the
distribution measured by other researchers [29] based on X-ray
method. The weld zone is under tension and the parent material far away from the weld zone is under compression. The strain
elds associated with individual dislocations result in broadening
as can short range stresses caused by the dislocation arrays and
sub-grain boundaries of the type commonly found in the parent
material. Both of these factors affect peak width due to the different grain sizes, the number of subgrains and their size and the
differences in dislocation densities in the parent material and the
weld zone.
Fig. 7 shows the residual stresses in a plane perpendicular
to welding direction which is dened as cross section; the peak
stresses are somewhat lower than in the transverse direction for
both alloys.
The loaddisplacement curves for all applied loads exhibit many
discontinuities (Figs. 5 and 6), especially for the case of AA5083H111 cross section (Fig. 6b), where a sudden increase in the
displacement is revealed in the loading part without any appreciable change in the load. This phenomenon is called pop-in
behavior, noticed in many brittle thin lms due to the cracking and
chipping events [30,31]. The pop-in phenomenon in the loaddepth
curves generated by Berkovich indenter suggests that a crack
begins growing at the very early stage of the loading, continues to

Fig. 10. Pile-up of aluminum (a) AA 6082-T6 [3] and (b) AA 5083 alloys through
SPM imaging.

232

C.A. Charitidis et al. / Materials Science and Engineering A 540 (2012) 226234

(a)

hc/hm

AA6082-T6
AA5083-H111
100

10

1000

(b)

10

hc/hm

Displacement, hm (nm)

AA6082-T6
AA5083-H111
0.1
0.015

0.02

0.025

0.03

0.035

H/E*
Fig. 11. Normalized pile-up/sink-in height for both alloys.

about 500 MPa near the weld line. This form of prole has been
observed previously for aluminum AA5083-H111 by Peel et al. [29].
The dip around the pin position may be attributed to the high
temperatures associated with this region, which would limit the
capacity of the material to support the generated load. Alternatively, it could be a consequence of stress relief that occurred when
the plate was cut up. For both alloys, a small plateau region appears
near the welding line; yet, the peak stresses are much higher in the
case of AA5083-H111.
In the measured residual stresses shown in Fig. 7(a) for AA6082T6 the stresses appear to rise to a tensile stress of 250 MPa at 1 cm
from the center, before reversing in trend to form a compressive
region near the weld line. It can be seen that this tensile region
peaks at around 1 cm from the weld line, a distance corresponding to the edge of the tool shoulder, before dropping slightly at
the weld centre. Also, in the measured residual stresses shown in
Fig. 7(b) for AA5083-H111 the stresses appear to decrease linearly
to a compressive stress steady state of about 500 MPa near the
weld line.

4.2. Nano- and micro-hardness distribution across the weld for


both alloys
Transverse distribution of Hmicro and Hnano presented the same
shape [3,33] and the same decrease of values between base metal
and welding zone (20% for AA5083 and 33% for AA6082), as
shown in Fig. 8.
In Fig. 9 the distribution of micro- and nano-hardness for both
alloys is presented. The linear regression equation for this relationship is:
Hmicro (HV ) = 55.26 Hnano (GPa) 7.28

(8)

where Hmicro represents the microhardness and Hnano represents


the nanohardness.
The square of the correlation coefcient R2 for (8) is 90, indicating that approximately 90% of the total variation is explained by
this linear regression equation. The Vickers and Berkovich indenters have similar geometric effects on indentation geometry, so it
is reasonable that Vickers testing and nanoindentation Hc values

C.A. Charitidis et al. / Materials Science and Engineering A 540 (2012) 226234

exhibit reasonable statistical relationship. The indentation shape


produced by the Berkovich indenter (used for the Hnano tests in this
study), is designed to be similar to the indentation created by the
Vickers indenter. The experimental results show that Hnano results
from a nanomechanical test instrument with a Berkovich indenter
can be reliably compared with Hmicro (Vickers) test results. Thus,
the selection of an appropriate H test should be based on the problem being investigated. The results of this investigation show that
Hnano test results from a nanomechanical test instrument with a
Berkovich indenter can be used to reliably analyze Hnano distribution on a friction stir welded alloy. The enhancement of elasticity
for weaker loads may be a real physical effect of superelastic behavior of materials under mN scale forces [34], due to the inactivation
of dislocations. It may also be an artefact due to the piling-up of the
surface during indentation.

4.3. Pile-up/sink-in deformation


The contact area is inuenced by the formation of pile-ups
and sink-ins during the indentation process. To accurately measure the indentation contact area, pile-ups/sinks-ins should be
appropriately accounted for. The presence of creep during nanoindentation has an effect on pile-up, which results in incorrect
measurement of the material properties. FischerCripps observed
this behaviour in aluminum where the measured elastic modulus was much less than expected [35]. Rar et al. [36] observed
that the same material when allowed to creep for a long duration produced a higher value of pile-up/sink-in indicating a switch
from an initial elastic sink-in to a plastic pile-up [36]. Creeping time is reported to have no impact on the pile-up/sink-in of
rate sensitive aluminum alloy [37]. In Fig. 10, SPM imaging of AA
6082-T6 [3] and AA 5083 alloys for 5000 N of applied load is presented; AA 6082-T6 exhibits signicant pile-up compared to AA
5083.
In Fig. 11, the normalised pile-up/sink-in height hc /hm is plotted vs. displacement hm and the normalized hardness H/E*. Higher
stresses are expected in high H/E*, hard materials, and high stress
concentrations develop towards the indenter tip, whereas in case
of low H/E*, soft materials, the stresses are lower and are distributed more evenly across the cross section of the material [35].
Rate sensitive materials experience less pile-up compared to rate
insensitive materials due strain hardening. Cheng and Cheng [38]
reported a 22% pile-up for a work hardening exponent. This is
consistent with the fact that when hc /hm approaches 1 for small
H/E*, deformation is intimately dominated by pile-up [39,40]. On
the other hand when hc /hm approaches 0 for large H/E* it corresponds to purely elastic deformation and is apparently dominated
by sink-in in a manner prescribed by Hertzian contact mechanics
[41].
An advantage of extracting residual stress by nanoindentation
is that there is no prerequisite of the knowledge of the stress state
of a reference sample or any particular mechanical properties of
the test materials; nanoindentation gives the opportunity of continuous measurement of residual stress and strain based on the
continuous measurement of stiffness [42]. However, as the method
relies on an accurate determination of the hc /hm ratio and as a
consequence of the contact area, experimental factors such as surface roughness, creep and pile-up may result in large error in the
experimental determination of the hc /hm ratio from the unloading curve of indentation. In nanoindentation measured areas that
exhibit pile-up the contact area calculated by the O&P method will
be overestimated, which leads to an underestimation of hardness
of both free and residual stress state region. Also, many uncertainties are connected with creep, since O&P method is unable to take it
into account time dependent deformation. These drawbacks pose

233

a both interesting and complicated problem for the estimation of


residual stress by nanoindentation.
5. Conclusions
Friction stir welding, contrary to classical welding techniques,
has the advantage that most of the welding parameters can be
controlled in a precise manner, thus controlling the energy input
into the system. However, the effect of different welding parameters on the weld properties remains an area of uncertainty. In
this work, the results of microstructural, mechanical integrity and
residual stress investigations of two friction stir welds (AA6082-T6
and AA5083-H111 aluminum alloys (in both the longitudinal (parallel to tool travel) and transverse (perpendicular to tool travel)
directions) produced under varying conditions), are reported. It
was found that the weld properties were dominated by the thermal input rather than the mechanical deformation by the tool.
It is clearly from this work that nanoindentation could arise
like an alternative technique to estimate and describe residual
stresses distribution. It is found that residual stress distribution
in the longitudinal direction of AA5083-H111 is in agreement
in trend with classical methods such as synchrotron residual
stress analysis. Increasing the traverse speed (and hence reducing the heat input) narrows this weld zone. Further investigation
includes extended pile-up investigation and synchrotron residual
stress analysis, which is reported to indicate whether the weld
zone is in tension in both the longitudinal and transverse directions.
Acknowledgements
Authors would like to thank D.I. Pantelis and N.M. Daniolos for
providing the FSW aluminum alloys.
References
[1] A. Scialpi, L.A.C. De Filippis, P. Cavaliere, J. Mater. Des. 28 (2007) 11241129.
[2] R. Nandan, T. DebRoy, H.K.D.H. Bhadeshia, Prog. Mater. Sci. 53 (August (6))
(2008) 9801023.
[3] E.P. Koumoulos, C.A. Charitidis, N.M. Daniolos, D.I. Pantelis, Mater. Sci. Eng. B
176 (2011) 15851589.
[4] J.Y. Kim, K.W. Lee, J.S. Lee, D. Kwon, Surf. Coat. Technol. 201 (2006) 42784283.
[5] J. Jang, D. Son, Y.H. Lee, Y. Choi, D. Kwon, Scr. Mater. 48 (2003) 743748.
[6] L. Yajiang, W. Juan, C. Maoai, S. Xiaoqin, Indian Acad. Sci. 27 (2004) 127132.
[7] W.C. Oliver, G.M. Pharr, J. Mater. Res. 7 (1992) 15641583.
[8] S. Suresh, A.E. Giannakopoulos, Acta Mater. 46 (1998) 57555767.
[9] Y.-H. Lee, D. Kwon, Scr. Mater. 49 (2003) 459465.
[10] N. Huber, J. Heerens, Acta Mater. 56 (2008) 62056213.
[11] S.K. Chionopoulos, C.H.I. Sarafoglou, D.I. Pantelis, V.J. Papazoglou, Proc. Int. Conf.
Manuf. Eng. (ICMEN), Chalkidiki, 2008.
[12] K. Nakata, S. Inoki, Y. Nagano, T. Hashimoto, S. Johgan, M. Ushio, J. Jpn. Inst.
Light Met. 51 (2001) 528533.
[13] W.B. Lee, J.W. Kim, Y.M. Yeon, S.B. Jung, Mater. Trans. 44 (2003) 917923.
[14] W.B. Lee, Y.M. Yeon, S.B. Jung, Mater. Sci. Technol. 19 (2003) 785790.
[15] S.H.C. Park, Y.S. Sato, H. Kokawa, T. Tsukeda, in: S.A. David, T. DebRoy, J.C.
Lippold, H.B. Smartt, J.M. Vitek (Eds.), Proceedings of the 6th International
Conference on Trends in Welding Research, ASM International, USA, 2003, pp.
267272.
[16] Y. Li, L.E. Murr, J.C. McClure, Mater. Sci. Eng. A 271 (1999) 213223.
[17] Z.Y. Ma, R.S. Mishra, M.W. Mahoney, Acta Mater. 50 (2002) 44194430.
[18] Y.J. Kwon, N. Saito, I. Shigematsu, Mater. Sci. Lett. 21 (2002) 14731476.
[19] Y.S. Sato, M. Urata, H. Kokawa, Metall. Mater. Trans. A 33 (2002) 625635.
[20] Y.J. Kwon, I. Shigematsu, N. Saito, Mater. Trans. 44 (2003) 13431350.
[21] Y.J. Kwon, I. Shigematsu, N. Saito, Scr. Mater. 49 (2003) 785789.
[22] I.N. Sneddon, Proc. Camb. Philos. Soc. 44 (1948) 492507.
[23] R.B. King, Int. J. Solids Struct. 23 (1987) 16571664.
[24] A.E. Giannakopoulos, J. Appl. Mech. 70 (2003) 638643.
[25] Y.H. Lee, D. Kwon, Acta Mater. 52 (2004) 15551563.
[26] J.H. Underwood, Exp. Mech. 13 (1973) 373380.
[27] Y.H. Lee, K. Takashima, Y. Higo, D. Kwon, Scr. Mater. 51 (2004) 887891.
[28] Y.H. Lee, J.Y. Kim, J.S. Lee, K.H. Kim, J.Y. Koo, D. Kwon, Philos. Mag. 86 (2006)
54975504.
[29] M. Peel, A. Steuwer, M. Preuss, P.J. Withers, Acta Mater. 51 (2003) 47914801.
[30] X.D. Li, B. Bhushan, Thin Solid Films 315 (1998) 214221.
[31] X.D. Li, B. Bhushan, Thin Solid Films 355 (1999) 330336.

234

C.A. Charitidis et al. / Materials Science and Engineering A 540 (2012) 226234

[32] P. Venkateswaran, Z.-H. Xu, X. Li, A.P. Reynolds, J. Mater. Sci. 44 (2009)
41404147.
[33] D.P.P. Booth, M.J. Starink, I. Sinclair, Mater. Sci. Technol. 23 (2007) 276284.
[34] G. Cross, A. Schirmeisen, A. Stalder, P. Grtter, M. Tschudy, U. Drig, Phys. Rev.
Lett. 80 (1998) 46854688.
[35] A.C. Fischer-Cripps, Mater. Sci. Eng. A 385 (2004) 7482.
[36] A. Rar, S. Sohn, W.C. Oliver, D.L. Goldsby, T.E. Tullis, G.M. Pharr, Fund. Nanoindentation and Nanotribology III. Symp, 2005, pp. 119124.
[37] S. Mandal, S. Kose, A. Frank, A.A. Elmustafa, Int. J. Surf. Sci. Eng. 2 (2008) 4151.
[38] Y.-T. Cheng, C.M. Cheng, Philos. Mag. Lett. 78 (1998) 115120.

[39]
[40]
[41]
[42]

R. Hill, B. Storakers, A.B. Zdunek, Proc. R. Soc. Lond. A 423 (1989) 301330.
S. Biwa, B. Storakers, J. Mech. Phys. Solids 43 (1995) 13031333.
H. Hertz, Miscellaneous Papers by H. Hertz, Macmillan, London, 1896.
Z.-H. Xu, X. Li, in: F. Yang, J.C.M. Li (Eds.), Micro and Nano Mechanical Testing
of Materials and Devices, Springer Verlag, New York, 2008, pp. 139154.
[43] E.P. Koumoulos, C.A. Charitidis, N.M. Daniolos, D.I. Pantelis, Determination of
onset of plasticity (yielding) and comparison of local mechanical properties of
friction stir welded aluminum alloys using the micro- and nano- indentation
techniques, Int. J. Struct. Integrity (2011) in press.

You might also like