Sie sind auf Seite 1von 10

Materials Science and Engineering A 527 (2010) 70997108

Contents lists available at ScienceDirect

Materials Science and Engineering A


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

Deformation behaviour of spot-welded high strength steels


for automotive applications
S. Brauser , L.A. Pepke, G. Weber, M. Rethmeier
BAM Federal Institute for Materials Research and Testing, Unter den Eichen 87, 12205 Berlin, Germany

a r t i c l e

i n f o

Article history:
Received 25 March 2010
Received in revised form 21 July 2010
Accepted 28 July 2010

Keywords:
Resistance spot welding
Deformation behaviour
Advanced high strength steel
TRIP steel
Similar and dissimilar material spot weld
EBSD
SEM
Strain eld

a b s t r a c t
Numerical simulation of component and assembly behaviour under different loading conditions is a main
tool for safety design in automobile body shell mass production. Knowledge of local material behaviour
is fundamental to such simulation tests. As a contribution to the verication of simulation results, the
local deformation properties of spot-welded similar and dissimilar material joints in shear tension tests
were investigated in this study for a TRIP steel (HCT690T) and a micro-alloyed steel (HX340LAD). For
this reason, the local strain distribution was calculated by the digital image correlation technique (DIC).
On the basis of the hardness values and microstructure of the spot welds, the differences in local strain
between the selected material combinations are discussed. Additionally, the retained austenite content
in the TRIP steel was analysed to explain the local strain values. Results obtained in this study regarding
similar material welds suggest signicant lower local strain values of the TRIP steel HCT690T compared
to HX340LAD. One reason could be the decrease of retained austenite in the welded area. Furthermore,
it has been ascertained that the local strain in dissimilar material welds decreases for each component
compared with the corresponding similar material weld.
2010 Elsevier B.V. All rights reserved.

1. Introduction
In the last decade a change in body shell mass production has
occurred in the automotive industry. In answer to the intensifying energy crisis and in order to meet customer requirements
for automobiles such as weight reduction for energy saving and
enhancement of passenger safety, new materials, e.g. advanced
high strength steels (AHSS) have to be applied. These materials
are gaining in popularity due to their high strength in combination with good ductility characteristics compared to traditional
high strength steels, for example micro-alloyed steels [13]. An
important AHSS representative is the so-called TRIP (TRansformation Induced Plasticity) steel dominated by a ferrite matrix with
retained austenite, bainite and martensite as dispersed phases,
offering excellent mechanical properties due to the transformation
of retained austenite into martensite during plastic straining [2,4].
As a result, both strength and uniform strain increase owing to the
appearance of a harder phase and to the additional local plastic
yielding of the surrounding grains related to the transformation
strain [5,6].
In the lightweight body shell mass production of automobiles,
resistance spot welding is the most important joining technique.

Corresponding author. Tel.: +493081044103; fax: +493081041557.


E-mail address: stephan.brauser@bam.de (S. Brauser).
0921-5093/$ see front matter 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.msea.2010.07.091

Typical vehicles contain about 30005000 spot welds. Therefore,


good resistance spot welding behaviour must be one of the key
characteristics of any steel grade to be used in automobile production [6].
For safe design of spot-welded body shell components, knowledge of the failure mechanism under static and fatigue loading is
of main interest [68]. Typically, three different failure modes can
occur in spot-welded structures, i.e. interfacial failure, plug failure and partial plug failure [6,9]. Of these, the plug failure is the
desired failure mode in automobile industry. For example, Zuniga
and Sheppard [10] as well as Lin et al. [11] studied the failure
modes of lap-shear specimens using optical micrographs. Lin et
al. [11] and Wung et al. [12,13] proposed failure criterions under
combined three resultant forces and three resultant moments and
under combined loads, respectively, based on experimental results.
The verication of such models is difcult because of the limited
available experimental test data. Further experimental work has
shown that in the plug failure mode, stress is concentrated at the
nugget circumference and leads to necking in the HAZ and base
metal, respectively [1416]. But, due to the localised joining zone
and, hence, localised stress/strain in the spot welds, characteristic material values for the welded area are not available till today,
especially for AHSS.
In order to investigate the failure mechanism in more detail,
the plastic strain and the stress distribution near the nugget must
be known [17]. However, for lap-shear specimens which are typi-

7100

S. Brauser et al. / Materials Science and Engineering A 527 (2010) 70997108

Fig. 1. FE-simulation of local strain behaviour of spot-welded shear tension samples via [6].

cal of spot-welded structures, analytical elasticplastic solutions


for stresses and plastic strains near the nugget are difcult to
obtain. Therefore, researchers usually apply numerical methods.
Numerous studies have been dedicated to elastic and elasticplastic
nite element analyses in order to characterise the fracture under
static and cyclic loading, i.e. the fatigue behaviour. Radakovic and
Tumuluru [6] proposed a simplied three-dimensional model of
lap-shear specimens and determined the local strain behaviour for
the interfacial failure and the plug failure, see Fig. 1. In the interfacial
failure mode, the maximum local strain occurs in the nugget. In the
plug failure mode, by contrast, the maximum local strain is found
to be at the inner surface of the sheet and decreases in direction to
the outer surface. These simulation results are in agreement with
the experimental work discussed above. Lin et al. [17], Kan [18], Pan
and Shepherd [19] as well as Satoh et al. [20] conducted two- and
three-dimensional nite element analyses to examine the fatigue
behaviour of spot-welded structures on the basis of plastic strain
distribution near the nugget. Particularly with a view to validating and optimising such numerical simulations, local strain values
are needed to calculate the real deformation behaviour and nally
build up a realistic nite element model to analyse the fracture
behaviour of spot-welded joints in automotive structures.
The objective of this study was to investigate the local surface
deformation behaviour of spot-welded similar and dissimilar material welds in a shear tension test and to calculate local material
data that can be used to validate numerical simulation of static and
cyclic loading of spot welds. For this purpose, an optical strain eld
measurement system with high resolution was used. To investigate
the effect of different stain values the fracture surface was characterised using scanning electron microscopy (SEM). Furthermore, on
the basis of EBSD (Electron Backscatter Diffraction) measurements
of the retained austenite content, explanation will be given for the
local strain values of TRIP steel HCT690T.

2. Experimental
In this study two different types of high strength steels
were selected, including micro-alloyed steel HX340LAD and AHSS
HCT690T. The micro-alloyed steel HX340LAD was chosen because
of its extensive use in the automotive industry for similar material welds (SMW) and above all with regard to its application for
dissimilar material welds (DMW), especially in conjunction with
HCT690T. Table 1 shows an extraction of the chemical composition
and the mechanical properties of the tested steels. Furthermore,
the carbon equivalent (CE) characterised based on Eq. (1) [21] is
listed, too. All steel grades offer a thickness of 1 mm and were hot
dip zinc coated with an average weight of 140 g m2 .

Fig. 2. Shear tension specimen dimension.

DMW dominate in body shell mass production. Therefore, in


addition to the base metal combinations HX340LAD/HX340LAD
and HCT690T/HCT690T, used as a reference, the combination
HX340LAD/HCT690T will also be investigated.
CE = %C + %Mn/6 + (%Cr + %V)/5 + %Si/15

(1)

In this investigation, two sheet samples, 105 mm long and


45 mm wide, were overlapped by 35 mm and single spot-welded
in the centre of the overlapped region, Fig. 2. These shear tension samples were used to calculate the local strain of conventional
spot-welded structures.
Following EN ISO 14329 [9], spot-weld failure may occur in
three modes: interfacial failure, plug failure and partial plug fail-

Table 1
Mechanical properties and an extract of the chemical composition of the base materials, measured via tensile test and Emission Spectrometry.
Steel grade

HX340LAD
HCT690T

Yield strength (MPa)

370
420

Tensile strength (MPa)

450
750

A (%)

32
30

Alloying elements (wt%)


C

Mn

Cr

Al

Si

Fe

CE

0.09
0.19

0.78
1.70

0.051
0.027

0.04
1.33

0.15
0.077

Balance
Balance

0.24
0.48

S. Brauser et al. / Materials Science and Engineering A 527 (2010) 70997108

Fig. 3. Examples for failure types in shear tension test.

ure. In the interfacial failure mode, the failure occurs through the
nugget, while in the plug and the partial plug failure mode, it occurs
by complete or partial withdrawal of nugget from one sheet, as
schematically shown in Fig. 3.
It is well established that the nugget diameter has an inuence
on the fracture behaviour of spot-welded joints. In simple terms, a
small nugget diameter often results in an interface failure while a
bigger weld nugget normally leads to a plug failure or to a partial
plug failure [15,16]. It is conjecturable that different failure types
lead to modied deformation characteristics. The interface failure
results in negligible deformation of the specimen surface while the
plug failure is characterised by a signicant deformation on the
specimen surface, see Fig. 2. Accordingly, to avoid an inuence of
weld size and fracture type, respectively, on the resulting deforma
tion, a constant nugget diameter of 4.5 t (t = sheet thickness, 1 mm)
was used which induces plug failures in all cases. The samples were
produced using an electromotive controlled welding gun with constant current control and direct-current operation. All specimens
were welded using medium frequency current and electrode caps
of type F16 attened to a face diameter of 5.5 mm. The welding

7101

parameters resulting always in a nugget diameter of 4.5 mm are


summarised in Table 2.
Due to the local deformation in spot-welded shear tension specimens only a qualitative description of the deformability of the
tested materials could be given for the shear tension test. Therefore,
the deformation behaviour of spot-welded specimens was analysed
using a digital image correlation system. This non-destructive optical method calculates the local strain by analysing the change of a
stochastically patterned surface coating. In this investigation the
specic strain x (strain in tensile direction, Fig. 4) is measured;
the local strain in the y- and z-directions are relatively low. Note
that the strain eld shows the strain distribution on the specimen
surface.
The measurement system consists of two 2-megapixle cameras
adjusted and linked to an image and data processing system which
calculates strain elds of specimens with stochastically patterned
surface coating, as schematically shown in Fig. 4. The frame rate of
the taken pictures was 1 Hz. Additionally, two stroboscopes were
used and synchronised with the camera system to ensure uniform
exposure of the patterned specimen surfaces. Normally, the deformation of one specimen surface can be measured. Following Lorenz
and Kannengiesser [22] where an optical system consisting of two
reectors was used, it was possible to observe the specimen from
the back side, front side and side, see Fig. 4. Hence, independently
of the specimen view, the maximum local strain of the specimen
can be appointed. The shear tension test was used to characterise
the mechanical properties and the deformation behaviour of the
welds. All tests were performed at room temperature with a strain
rate of 0.01 mm s1 . Metallographic tests were used to measure the
nugget diameter.
Additionally, to identify the deformation behaviour of spotwelded TRIP steel HCT690T, the content of retained austenite in the
area of crack propagation was analysed using EBSD and the Sorpas
FE-simulation software. The purpose of using the commercial FEsimulation software Sorpas is to determine the cooling times t8/5
in the HAZ fracture area and to correlate this information with the
change in austenite content which is calculated by EBSD. In EBSD,
a stationary electron beam strikes a tilted crystalline sample and
the diffracted electrons form a pattern on a uorescent screen. This
pattern is characteristic of the crystal structure and enables, among
other factors, the separation of face cubic centered (fcc) and body
cubic centered (bcc) lattice structures.

Table 2
Welding parameters.
Material

Electrode force (kN)

Welding current (kA)

Welding time in cycles

HX340LAD
HCT690T
HCT690T/HX340LAD

3.5
3.5
3.5

8.00
7.20
7.15

12
12
12

Pre/post holding time in cycles


3/5
3/5
3/5

Fig. 4. Schematic overview of strain eld measurement with stochastically patterned shear tension specimen.

7102

S. Brauser et al. / Materials Science and Engineering A 527 (2010) 70997108

Fig. 5. Similar and dissimilar material spot-welded cross-sections, (a) HX340LAD, (b) HX340LAD/HCT690T, (c) HCT690T.

Fig. 6. Hardness proles for similar and dissimilar material spot welds with schematic location of the indentations, (a) HX340LAD, (b) HX340LAD/HCT690T and (c) HCT690T.

Fig. 7. Shear tension test results, (a) loaddisplacement curves, (b) Failure load depending on base metal combination.

S. Brauser et al. / Materials Science and Engineering A 527 (2010) 70997108

7103

Fig. 8. Local strain x on front and back side of the specimens during shear tension test for (a) HX340LAD, (b) HCT690T and (c) HX340LAD/HCT690T.

3. Results and discussion


3.1. Microstructure and hardness distribution
Weld cross-sections of the tested steel combinations are shown
in Fig. 5. The three different regions, i.e. base metal (BM), heat
affected zone (HAZ) and fusion zone (FZ) are labelled.
The hardness distribution provides indirect information about
the strength and the deformation behaviour of spot-welded joints.
Normally, increasing hardness results in decreasing formability
with simultaneously increasing strength [23]. Typical weld nugget
hardness proles of the tested material combinations are given in
Fig. 6. Due to the high cooling rates in resistance spot welding, i.e.
300010,000 K/s [24], an increase in weld nugget hardness, in comparison to the base metal, is observed. It can also be seen that the
HAZ hardness decreases from the nugget edge towards the base
metal, which indicates the decrease of material strength in that
area. In agreement with the literature [15,16], the hardness of the
fusion zone (nugget) and of the HAZ of TRIP steel HCT690T is appreciably higher as a consequence of the higher carbon equivalent
(Table 1), in comparison to the micro-alloyed steel HX340LAD.
The hardness prole for the dissimilar material welds
HX340LAD/HCT690T shown in Fig. 6b takes a discontinuous
shape which is attributable to the different properties of the
welded materials. Examination of the HAZ hardness values of each
component (HX340LAD; HCT690T) does not reveal any signicant
increase in hardness compared to the corresponding SMW. Thus,

in the case of plug failure where the HAZ strength is of main


importance to the shear tension strength of the weld, the DMW
performance is expected to be similar to that of the SMW of
HX340LAD, based on the assumption that the fracture happens in
the weakest part of the weld.
However, concerning the weld nugget, a deviation in hardness
can clearly be identied which lies between the values obtained
for the similar material weld, with a shift to the harder material
(HCT690T). Consequently, in DMW the weld nugget hardness is
inuenced to a greater extent by the harder material.
3.2. Static shear tension test
The stability of spot welds is often characterised with the help
of shear tension tests. The loaddisplacement curves obtained
from shear tension tests for a nugget diameter dn of nearly 4.5 mm
are plotted in Fig. 7a. For SMW, a considerable increase in failure
load with ascending base metal strength is discovered. This can be
related to the fact that the strength of spot welds is determined,
among other parameters like sheet thickness and nugget size
(constant parameter in this study), by the nugget and especially
HAZ hardness.
Regarding the failure loads of DMW, comparable values as in
the case of SMW of HX340LAD can be achieved; however, there is
a signicant drop in displacement. This conrms the assumptions
regarding the strength of DMW based on the results of hardness
tests are seen below. Fig. 7b shows this trend of increasing peak

Fig. 9. Results of front and back side strain led measurement at the peak load (a) HX340LAD (b)/HCT690T/HX340LAD with schematically indicated nugget location.

7104

S. Brauser et al. / Materials Science and Engineering A 527 (2010) 70997108

Fig. 10. Maximum local strain l ,max versus shear tension load with visualisation of local strain (a) HX340LAD/HCT690T (b) HCT690T, (c) HX340LAD/HCT690T and (d)
HX340LAD.

shear tension loads with rising base metal strength. Previous work
concerning the inuence of the softer material part in DMW has
also shown that the softer material component leads to a decrease
of the failure load [25].
Furthermore, the decrease in displacement, shown in Fig. 7a,
points out that DMW of HX340LAD/HCT690T offer lower
deformability compared to the base metal combination of
HX340LAD/HX340LAD.

3.3. Strain measurement


In order to measure the surface strain of spot welds in the spot
weld area, the systematic error of the measuring system was rst
determined by analysing an unstressed specimen. As a result, the
local strain varies within a range of 0.25% and 0.3%.
The results of local strain measurements in the area of maximal surface strain are plotted in the diagrams depicted in Fig. 8.

Fig. 11. Cross-section with hardness values (a) HX340LAD, (b) HCT690T, (c) HX340LAD/HX340LAD.

S. Brauser et al. / Materials Science and Engineering A 527 (2010) 70997108

7105

Fig. 12. SEM-fractography of spot weld fracture areas, (a) HX340LAD, (b) HCT690T, HX340LAD/HCT690T.

Due to the identical microstructures as well as hardness values in


the HAZ and the fracture area, respectively, no signicant differences in strain values between front and back side of SMW could
be observed. Fig. 9a exemplies the strain elds of HX340LAD on
both sides to demonstrate that not only the strain values are nearly
identical but also the measured strain elds. In the following, the
specimen side where the maximum strain occurs will be considered
to discuss the strain behaviour.
As expected, the results of DMW (HCT690T/HX340LAD) reveal
material-dependent strain behaviour, Figs. 8c and Fig. 9b. After
exceeding 1% local strain the HX340LAD exhibits a faster increase of
local strain than the TRIP component. The comparison of the strain
behaviour of SMW and DMW shows only small differences concerning the maximum strain values (Fig. 8) as well as the measured
strain elds (Fig. 9).
By analysing the measured strain eld during the shear tension
test as well as the cross-section of the tested sheet metal combination, statements about the deformation behaviour can be made. In
Fig. 10, the maximum local strain x,max is plotted versus the shear
tension load. Moreover, the results of strain eld measurements are
visualised. For visualisation of the deformation behaviour, the peak
load Fs,max (shear tension strength) was used as reference point,
because the crack initiation and propagation was not in the focus
of this study.
The strain curve of spot-welded HX340LAD (Fig. 10b) shows
a large range of plastic deformation and necking which is characteristic of ductile material behaviour. The maximum strain is
concentrated in the HAZ/base metal transition zone (see Fig. 10d)
and reaches values up to 15% at the peak load. In contrast, the TRIP
steel HCT690T exhibits only a small range of plastic deformation
and fracture happened after achieving the peak load (11.4 kN). This
material behaviour is representative of a less ductile fracture. The
maximum deformation achieved at the surface happened also in
the HAZ/base metal transition zone (Fig. 10b) and reaches values
of nearly 5%. A reason for the lower local strain of the TRIP steel
HCT690T is the lower rotation of the nugget due to the higher
nugget hardness that hinders the deformation of the surface [14].
It should be noted that the local strain behaviour and the position
of the maximum local strain is in good agreement with the results
of Radakovic and Tumuluru [6] shown in Fig. 1b.
The respective loadstrain curves of DMW show runs comparable to those of the corresponding curves of spot-welded SMW.

However, since the softer material signicantly determines the failure load in DMW, the peak load of DMW is lower than in SMW
of TRIP steel HCT690T.At the maximum local strain (peak load),
the HX340LAD component offers values of up to 12% while the
HCT690T component exhibits a value of 3.5%, Fig. 10a and c. Therefore, a nearly similar decrease of the DMW local strain values of up
to 20% is found for both steel parts in comparison to the corresponding SMW. Consequently, the combination of different materials in
RSW results in a major decrease of the peak load associated with
a drop in maximum local strain relating to the stronger material
part. Concerning the softer material component, only the maximum local strain is reduced, however without any inuence on the
fracture behaviour.
The differences in local strain values between the tested sheet
metal combinations are reected in the fracture behaviour, Fig. 11.
The micrographs of the cross-section for the micro-alloyed steel
HX340LAD and the DMW (HX340LAD/HCT690T) show that fracture
happened in the HAZ/base metal transition zone approximately
1 mm away from the nugget circumference after signicant necking, Fig. 11a and c. In the DMW, fracture happened in the softer
material part (HX340LAD). Lin et al. [17] have performed nite element analyses of the failure modes of spot welds and have shown
that when necking failure occurs at the distance in the order of
the thickness away from the notch tip, the ductility of the material near the notch or crack along the nugget circumference is high.
This result corresponds well with the local strain values discussed
above.
Unlike the micro-alloyed steel HX340LAD where fracture happened only in the HAZ/base metal transition zone, the TRIP steel
HCT690T fracture started directly in the HAZ region with the maximum hardness gradient (nearly 526 HV 378 HV, Fig. 6b) and
propagated into the region with reduced hardness and strength,
Fig. 10b. Based on the outcomes of [17] this behaviour could be
attributed to lower ductility of the material near the notch which
leads to initiation of kinked cracks at the critical locations of the
notch seen in Fig. 10b.
To examine the deformation behaviour in more detail, SEM analyses of the fracture region were performed, Fig. 12. Due to the
fact that the shape of the dimples depends on the loading conditions, the plug failure under tensile loading predominantly results
in equiaxed dimples while shear loading will create elongated dimples [15]. Fig. 12a shows the results of SEM for the micro-alloyed

7106

S. Brauser et al. / Materials Science and Engineering A 527 (2010) 70997108

Fig. 13. Force distribution at nugget centreline and circumference during shear
tensile test [15].

steel HX340LAD (SMW). It can be seen that the dimples signicantly elongated indicate that the fracture happened under shear
load. This result is opposite to the work of Chao [7] who studied
failure mechanisms of pullout occurring in RSW during the shear
test. In contrast, the TRIP steel HCT690T exhibits equiaxed dimples
in the fracture zone which are typical of a tensile fracture mechanism Fig. 12b. In the case of DMW, elongated dimples are observed
(Fig. 12c) that implies similar behaviour to SMW of HX340LAD
(fracture under shear load).
The differences in fracture mechanisms (shear, tensile) between
TRIP steel HCT690T and micro-alloyed steel HX340LAD could be
explained by the simple model for stress distribution in spot welds
under shear tensile load seen in Fig. 13. Shear stresses are dominant at the interface. At the nugget circumference, the stress nature
is tensile shear at position A and compressive at position B [15].
Owing to the macroscopic rotation of the weld, not only tensile
shear force F| is produced at the spot weld but also cross-tension
force F, whereas with resign rotation angle the cross-tension
force F increases:
F = F sin

(2)

F = F sin

(3)

As a result, the softer material HX340LAD with higher rotation


angle (10 at the peak load) is subject to a higher crosstension force than the TRIP steel HCT690T (4 ). Therefore, the
micro-alloyed steel HX340LAD and the DMW HX340LAD/HCT690T
(rotation angle 9 ) show in contrast to the SMW of HCT690T a
shear fracture behaviour with elongated dimples. The difference in
rotation corresponds well with that described in the work of [14].
Concerning the deformability, ductile fracture characterised by
a dimple structure is observed in all cases (SMW, DMW). Even
though the deepness of the dimples is an indication of the material
ductility [26], the differences in local strain seen in Fig. 10 cannot
be detected by SEM analyses for comparison due to the elongation
of the dimples.

Fig. 14. Retained austenite (blue) in TRIP steel HCT690T base metal. (For interpretation of the references to colour in this gure legend, the reader is referred to the
web version of the article.)

3.4. Decrease in ductility of TRIP steel HCT690T


A main reason for the high ductility of TRIP steel HCT690T is the
retained austenite content which can transform into martensite
under stress [1,2]. Fig. 14 shows the austenite distribution (blue
marked area) in the base metal measured with the help of EBSD.
The ne dispersed austenite is arranged at the grain boundaries of
the fcc lattice structures (grey area; primarily ferrite) and reached
a content of 17%.
The following discussion regarding the low local ductility in
the spot-welded area of TRIP steel HCT690T focuses on the HAZ,
because in this investigation fracture starts in and propagates
through the HAZ. During resistance spot welding, the HAZ undergoes a temperature cycle which is characterised by a high cooling
rate compared to other welding procedures [24]. Fig. 15a shows
the temperature eld at the peak temperature of the tested TRIP
steel determined by FE software Sorpas. Furthermore, the crack
path seen in Fig. 11b is labelled. On the basis of the calculated
temperature eld, cooling-down-curves for three points near the
crack path and the corresponding cooling times t8/5 were investigated, Fig. 15b. The short process time of resistance spot welding
results in times t8/5 of 80 ms (point 1) and 100 ms (point 2), respectively. Based on continuous cooling transformation diagrams for
low alloyed TRIP steel [27,28], only martensite can develop in the
HAZ directly adjacent to the fusion zone.

Fig. 15. Temperature eld at the peak temperature of spot-welded TRIP steel HCT690T with spot weld diameter of 4.5 mm determined by FE software Sorpas (a), coolingdown-curves for three measurement points in the HAZ (b).

S. Brauser et al. / Materials Science and Engineering A 527 (2010) 70997108

7107

Fig. 16. Results of EBSD measurement near the crack path area; black and red marked areas correspond to fcc structure, blue marked areas correspond to bcc structure.

The EBSD analyses conrm the assumptions regarding the


reduction of austenite during spot welding. Fig. 16 shows the results
of EBSD measurement near the crack path area (HAZ) transcribed
from Fig. 11b. It can be seen on the one hand that the austenite
content in the region of crack propagation is negligibly small. On
the other hand, the austenite content increases with increasing
distance from the fusion zone. That is a result of the lower peak
temperature during welding (see Point 3 in Fig. 15b) which does
not lead to a complete austenite transformation into bcc structures (martensite, bainite). The black marked area (Fig. 16b and
c) corresponds to very ne fcc structures.
Summarising, resistance spot welding leads to a nearly complete
removal of retained austenite as well as a hardness increase in the
fracture region (HAZ) resulting in a considerable decrease of local
strain measured with digital image correlation technique.

4. Conclusions
In this study, deformation behaviour of SMW and DMW in shear
tension test was investigated using a system for optical strain led
analysis. This was accomplished applying hardness values of the
weld and HAZ, loaddisplacement curves as well as loadstrain
curves. SEM and EBSD analyses were carried out to characterise
the deformation behaviour in the fracture area.
The following essential conclusions can be drawn:
The failure load does not show a linear increase with the base
metal strength. Failure loads of dissimilar material welds are
located between the analysed similar material welds with a sig-

nicant shift to the softer material (HX340LAD) and a drop in


displacement.
Up to the point of uniform elongation, the local strain measurement of SMW shows no signicant differences between front and
back side.
The local strain observed for the micro-alloyed steel HX340LAD
reaches values of 15% while the TRIP steel offers values of only
5%.
In DMW, a nearly similar decrease of the local strain values
accounting for up to 20% for both steel parts is found compared
with the results of SMW.
With regard to the ductility behaviour, spot-welded joints are
characterised by a signicant loss of local strain accounting for
up to 25% for HX340LAD and 85% for HCT690T compared to the
base metal.
SEM results show no indication of reduced deformability of TRIP
steel compared to the micro-alloyed steel HX340LAD. But evidence of different loading conditions between HX340LAD and
HCT690T was found.
EBSD analysis of the retained austenite content reveals an
elimination of austenite in the fracture region of spot-welded
HCT690T. The austenite reduction and the hardness increase in
the fracture region (HAZ) results in low strain values compared
to the base metal.

The strain values obtained in this investigation can be integrated


into numerical simulation studies to help analyse the deformation
as well as the fracture behaviour of spot welds and are expected to
contribute to increased passenger safety in the long run.

7108

S. Brauser et al. / Materials Science and Engineering A 527 (2010) 70997108

References
[1] S. Maggi, M. Murgia, Weld. Int. 22 (2008) 610618.
[2] Advanced High Strength Steel (AHSS) Application Guidelines Version 4.1
http://www.worldautosteel.org/Projects/AHSS-Guidelines.aspx, 13.04.2010.
[3] M.D. Tumuluru, Weld. J. (2006) 3137.
[4] G. Lacroix, T. Pardoen, P.J. Jacques, Acta Mater. 56 (2008) 39003913.
[5] M.H. Saleh, R. Priestner, J. Mater. Proc. Technol. 113 (2001) 587593.
[6] J.D. Radakovic, M.D. Tumuluru, Proceeding of the International Sheet Metal
Welding Conference XIII, 2008.
[7] Y.J. Chao, J. Eng. Mater. Technol. 125 (2003) 125132.
[8] J.A. Davidson, SAE Technical Paper 830033, Society of Automotive Engineers,
Warrendale, PA.
[9] ISO 14329:2000: Resistance welding destructive tests of welds failure types
and geometric measurements for resistance spot, seam and projection welds,
2003.
[10] S. Zuniga, S.D. Sheppard, in: R.S. Piascik (Ed.), Fatigue and Fracture Mechanics,
vol. 27, ASTM STP 1296, 1997, pp. 469489.
[11] S.-H. Lin, J. Pan, S. Wu, T. Tyan, P. Wung, SAE Technical Paper 2001-01-0428,
Society of Automotive Engineering, Warrendale, 2001.
[12] P. Wung, T. Walsh, A. Ourchane, W. Stewart, M. Jie, Exp. Mech. 41 (2001)
100106.
[13] P. Wung, Exp. Mech. 41 (2001) 107113.

[14] H. Zhang, J. Senkara, Resistance Welding: Fundamentals and Applications, 1st


ed., CRC Press, 2005, pp. 107146.
[15] M. Pouranvari, H.R. Asgari, S.M. Mosavizadch, P.H. Marashi, M. Goodarzi, Sci.
Technol. Weld. J. 12 (2007) 217225.
[16] M.I. Khan, L.M. Kuntz, Y. Zhou, Sci. Technol. Weld. J. Vol.13 (2008) 294304.
[17] P.-C. Lin, S.-H. Lin, J. Pan, Eng. Frac. Mech. 73 (2006) 22292249.
[18] Y.R. Kan, Met. Eng. Quart. 16 (1976) 2636.
[19] N. Pan, S. Sheppard, Int. J. Fatigue 24 (2002) 519528.
[20] T. Satoh, H. Abe, K. Nishikawa, M. Morita, Trans. Jpn. Weld. Soc. 22 (1991) 4651.
[21] M.J. Cieslak, ASM Handbook, vol. 6, ASM International, Materials Park, Ohio,
1990, pp. 9495.
[22] S. Lorenz, T. Kannengiesser, Proceeding of the Fifteenth International Conference on the Joining of Materials (JOM 15), Helsingr-Denmark, 2009.
[23] DIN EN ISO 18265: Metallic materialsConversion of hardness values (ISO
18265:2003); German version EN ISO 18265:2003.
[24] J.E. Gould, S.P. Khurana, T. Li, Weld. J. 85 (2006) 111116.
[25] B. Hernandez, M.L. Kuntz, M.I. Khan, Y. Zhou, Sci. Technol. Weld. J. 13 (2008)
769776.
[26] G. Lange, (Hg.): Systematische Beurteilung Technischer Schadensflle, 5th ed.,
Wiley-VCH, Weinheim, 2001, ISBN 3-527-30417-7 (in German).
[27] M. Zhang, L. Lia, R.Y. Fu, D. Krizan, B.C. De Coomanc, Mater. Sci. Eng. A 438
(2006) 296299.
[28] M. Grajcar, Opiela, J. Ach. Mater. Manu. Eng. 29 (2008) 7178.

Das könnte Ihnen auch gefallen