Vous êtes sur la page 1sur 10

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/259517200

Microstructure and dynamic tensile behavior of


DP600 dual phase steel joint by laser welding
Article in Materials Science and Engineering A January 2014
Impact Factor: 2.57 DOI: 10.1016/j.msea.2013.11.047

CITATIONS

READS

73

6 authors, including:
Danyang Dong

Yang Liu

Northeastern University (Shenyang, China)

NASA

5 PUBLICATIONS 17 CITATIONS

443 PUBLICATIONS 4,180 CITATIONS

SEE PROFILE

SEE PROFILE

Yuling Yang

Jinfeng Li

Northeastern University (Shenyang, China)

Institute of Physics,Chinese Academy of Sci

18 PUBLICATIONS 144 CITATIONS

2 PUBLICATIONS 12 CITATIONS

SEE PROFILE

SEE PROFILE

Available from: Danyang Dong


Retrieved on: 30 April 2016

Materials Science & Engineering A 594 (2014) 1725

Contents lists available at ScienceDirect

Materials Science & Engineering A


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

Microstructure and dynamic tensile behavior of DP600 dual phase


steel joint by laser welding
Danyang Dong a,n, Yang Liu b, Yuling Yang a, Jinfeng Li a, Min Ma a, Tao Jiang a
a
b

College of Science, Northeastern University, No. 11, Lane 3, WenHua Road, HePing District, Shenyang 110819, China
Key Laboratory for Anisotropy and Texture of Materials, Ministry of Education, Northeastern University, Shenyang 110819, China

art ic l e i nf o

a b s t r a c t

Article history:
Received 27 June 2013
Received in revised form
12 October 2013
Accepted 14 November 2013
Available online 21 November 2013

Dual phase (DP) steels have been widely used in the automotive industry to reduce vehicle weight and
improve car safety. In such applications welding and joining have to be involved, which would lead to a
localized change of the microstructure and property, and create potential safety and reliable issues under
dynamic loading. The aim of the present study is to examine the rate-dependent mechanical properties,
deformation and fracture behavior of DP600 steel and its welded joint (WJ) produced by Nd:YAG laser
welding over a wide range of strain rates (0.0011133 s  1). Laser welding results in not only signicant
microhardness increase in the fusion zone (FZ) and inner heat-affected zone (HAZ), but also the
formation of a softened zone in the outer HAZ. The yield strength (YS) of the DP600 steel increases and
the ultimate tensile strength (UTS) remains almost unchanged, but the ductility decreases after welding.
The DP600 base metal (BM) and WJ are of positive strain rate sensitivity and show similar stressstrain
response at all studied strain rates. The enhanced ductility at strain rates ranging from 1 to 100 s  1 is
attributed to the retardation of the propagation of plastic strain localization due to the positive strain rate
sensitivity and the thermal softening caused by deformation induced adiabatic temperature rise during
dynamic tensile deformation. The tensile failure occurs in the inner HAZ of the joint and the distance of
failure location from the weld centerline decreases with increasing strain rate. The mechanism for the
changing failure location can be related to the different strain rate dependence of the plastic deformation
behavior of the microstructures in various regions across the joint. The DP600 WJ absorbs more energy
over the whole measured strain rates than that of the BM due to the higher strength at the same strain
when the deformation only up to 10% is considered.
& 2013 Elsevier B.V. All rights reserved.

Keywords:
Dual phase (DP) steel
Laser welding
Dynamic tensile
Microstructure
Strain rate

1. Introduction
Efforts to reduce weight and improve crash performance have
resulted in increased application of advanced high strength steels
(AHSSs) in automotive industry [1]. Ferrite-martensite dual phase
(DP) steels are one of the most common AHSSs which are
currently used in automotive components to meet enhanced
government regulations and safety standards. In this combination
of two phases, martensite contributes with high strength and
ferrite matrix provides good elongation that can produce a good
combination of strength and ductility for applications which
require good formability. This unique composite microstructure
offers other interesting mechanical properties such as continuous

Corresponding author. Tel.: 86 24 83687658; fax: 86 24 83691575.


E-mail addresses: dongdanyang@mail.neu.edu.cn (D. Dong),
liuyang@mail.neu.edu.cn (Y. Liu), yulingyang@mail.neu.edu.cn (Y. Yang),
lijinfengboda@163.com (J. Li), sharon6789@163.com (M. Ma),
tao.jiang906@yahoo.com (T. Jiang).
0921-5093/$ - see front matter & 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.msea.2013.11.047

yielding, low yield stress to tensile strength ratios and high initial
work-hardening rate [2].
During manufacturing of automotive components, welding is
the mostly used joining operation. Due to ease of automation and
exibility, laser welding has gained its popularity in metal joining
industry [3] and has been considered to potentially replace some
other joining processes such as resistance spot welding (RSW) and
friction stir welding (FSW).
The structural components of auto-bodies are often subjected
to dynamic loadings during their manufacturing processes, as well
as in an accident. The range of the strain rate is several tens to
hundreds per second in a common vehicle crash and the highest
strain rate can reach more than 1000 s  1 [46] at which the
dynamic response of steel sheets is different from static ones and
the strain rate effect becomes signicant. Thus, it is essential to
understand the mechanical behavior of AHSSs considering the
strain rate effect for the reliable design of the structural components. A lot of studies have been carried out on the laser
weldability of DP steels coupled with the effect of welding on
the tensile properties [3,712], fatigue properties [3,7,12] and

18

D. Dong et al. / Materials Science & Engineering A 594 (2014) 1725

formability [9,11,13]. Heterogeneity of microstructures and properties across a weld brought about by the laser welding process
may greatly change the deformation response of a weld, and thus
change the overall welded specimen properties, which would
create potential safety and reliability issues under the dynamic
loading. Most of the property characterizations for the AHSS WJ, as
mentioned above, have been performed under quasi-static loadings [3,713]. But very limited studies on the dynamic mechanical
behavior of the joint, especially the effect of microstructure
heterogeneity on the dynamic tensile properties, deformation
and fracture behavior and energy absorption have been reported.
This study was, therefore, aimed at investigating the strain rate
effects on the mechanical properties, deformation and fracture
behavior of DP steel joints produced by laser welding. The microstructure and microhardness change of the DP steel after welding
were examined. Quasi-static and dynamic tensile tests were performed on the BM and WJ of DP steel over a wide strain rate range
(0.0011133 s  1) to investigate the tensile properties and deformation behavior at various strain rates. The failure location, the tensile
fracture surface morphology and the deformed microstructure of
both BM and WJ were analyzed. The energy absorption at various
strain rates was also evaluated in the present study.

2. Experimental
The as-received DP600 steel sheet was used in the present
study and its thickness is 0.7 mm. The chemical composition is
listed in Table 1.
Laser welding was performed using a pulsed Nd:YAG laser
(JHM-1GY-400) with a maximum mean power of 400 W and a
wavelength of 1064 nm. Two identical DP600 steel sheets with
dimensions of 65 mm  45 mm were welded together in a butt
joint conguration. A schematic illustration of the welded blank is
shown in Fig. 1(a). High-purity argon (99.99%) shielding gas
Table 1
Chemical composition of the DP600 steel used in the present study in wt%.

was supplied to the top surface of the specimens at a ow rate


of 5 L/min. Welding trials were carried out and specimens were
extracted from various locations of the WJ and subjected to
macrostructure analysis. The optimized process parameters presented in Table 2 were used to fabricate the joints free of
volumetric defect and lack of penetration for further investigation.
Metallographic specimens were cut from the weld crosssection (Fig. 1(a)), prepared according to the standard procedure
and etched with 4% nital solution. The microstructure examination
was carried out using a JEOL JSM-6510A scanning electron microscope (SEM). The volume fraction of martensite in the BM was
measured using an image analysis tool in SEM. Microhardness
tests were performed at a load of 50 g with a holding time of 10 s
using a Vickers hardness tester (Wilson Wdpert 401MVD). All
indentations were adequately spaced to avoid any potential effect
of strain elds caused by adjacent indentations.
Uniaxial tensile tests were conducted on the BM and WJ of DP600
steel at strain rates ranging from 0.001 to 1133 s  1 at room
temperature. A material testing machine (MTS810) was used for the
testes at strain rates of 0.0011 s  1, and a servo-hydraulic high speed
tensile testing machine (Zwick HTM5020) was used for the tests at
higher strain rates. A schematic illustration of the dimension of tensile
specimens is shown in Fig. 1(b). All specimens were made so that the
tensile axis coincides with the rolling direction. For better comparability, the geometry and dimension of the specimens for the tensile
tests of the BM and WJ were identical at all strain rates. The high
speed tensile testing machine has experimental capability for the
maximum velocity of 20 m/s, maximum load of 50 kN and maximum
stroke (or maximum displacement of the crosshead) of 350 mm. The
dynamic load was measured by the strain gauge attached at the grip
section of the specimen to reduce the load ringing phenomenon, as
illustrated in Fig. 1(b). The strain gauge positioning was identical for
all dynamic tensile test specimens. For the strain measurement, a

Table 2
Parameters used for Nd:YAG laser welding DP600 steel in the present study.

Mn

Si

Cr

Al

Mo

Fe

Focal
length
(mm)

0.061

1.40

0.89

0.012

0.003

0.026

0.032

0.021

Bal.

100

Defocus
distance
(mm)

Pulse
duration
(ms)

Pulse
frequency
(Hz)

Welding
speed
(mm/s)

Pulse
current
(A)

Pulse
energy
(J/pulse)

15

3.0

14

1.5

300

4.82

Fig. 1. Schematic illustrations showing position and dimensions of the specimen used in the present study, (a) dimensions of welded blanks and sampling positions of
metallographic specimens and tensile specimens, (b) dimensions of tensile specimens at all strain rates.

D. Dong et al. / Materials Science & Engineering A 594 (2014) 1725

linear displacement transducer (LDT) was used to measure the


displacement of the crosshead, and the strain of the specimen was
calculated from this acquired displacement. The data synchronous
acquisition of dynamic load, displacement, strain and loading rate was
performed at a natural frequency of 1 MHz during the dynamic
tensile tests. The quasi-static and dynamic tensile properties were
characterized in terms of the yield stress (YS), ultimate tensile
strength (UTS), uniform elongation (UE) and total elongation (TE).
Note that the YS was determined by 0.2% offset method and the UE
was determined at the strain where the engineering stress is
maximized. The TE is the strain at failure in the stressstrain curve.
The energy absorption up to 10% engineering strain was calculated for
the BM and WJ of DP600 steel at various strain rates.
The failure locations of the DP600 WJ after tensile testing at
various strain rates were examined using a high precision measuring microscope (OLYMPUS STM6). The fracture surfaces of the BM
and WJ were examined using SEM. The fractured specimens were
sectioned along the mid-width in longitudinal direction for SEM
analysis of deformed microstructure near the fracture surface.

19

martensite is estimated to be 15%. The microhardness prole on the


weld cross-section is shown in Fig. 3. A laser beam traveling along
the joint has the effect of a transient thermal wave moving through
the material. The local microstructure and properties at any point
within the WJ are generally determined by the thermal cycle
experienced at that location, with the peak temperature decreasing
with increasing distance from the weld centerline [14]. In the FZ, the
peak temperature is above the melting point. As expected by the
presence of a fully martensite microstructure (Fig. 2(c)), the microhardness of the FZ becomes higher (maximum microhardness of
340 HV) than that of the BM (average microhardness of 210 HV). Just
outside the FZ, near the fusion boundary, the peak temperature
exceeds the critical temperature (Ac3 temperature), and 100% austenite
forms. This region is called the supercritical HAZ. Further from the

3. Results and discussion


3.1. Microstructure and microhardness
The overall view of weld cross-section of the DP600 WJ is
shown in Fig. 2(a). It is seen that the weld cross-section exhibited a
heterogeneous structure across the various regions of the WJ,
including fusion zone (FZ), inner heat-affected zone (HAZ), outer
HAZ and BM.
More detailed microstructure changes are shown in Fig. 2(b)(e).
The microstructure of the DP600 BM (Fig. 2(b)) was characterized by
martensite islands in the ferrite matrix, where the volume fraction of

Fig. 3. Microhardness prole across the DP600 WJ.

Fig. 2. SEM micrographs showing microstructure change of the DP600 WJ, (a) overall view of the cross-section, (b) BM, (c) FZ, (d) inner HAZ, and (e) outer HAZ (where M:
martensite, F: ferrite, and TM: tempered martensite).

20

D. Dong et al. / Materials Science & Engineering A 594 (2014) 1725

fusion boundary, the peak temperature is within the intercritical range


(temperatures between Ac1 and Ac3 temperatures), which results in
austenitization of the carbon-rich martensite phases, while large areas
of undissolved ferrite remain unchanged. The inner HAZ consists of
the supercritical and intercritical regions. The microstructure examination as shown in Fig. 2(d) reveals the presence of the ne
martensite and undissolved ferrite. With increasing distance from
the weld centerline, the volume fraction of undissolved ferrite
increases, while that of ne martensite decreases. Compared to the
BM, a higher volume fraction of hard phases are found in the FZ and
inner HAZ, which results in an increase in microhardness and the
formation of a hardened zone. Further from the fusion boundary, the
peak temperature is below Ac1 (i.e., subcritical) and the local tempering of the metastable martensite phase occurs, resulting in a reduction
in microhardness. A microhardness valley is observed in the weld
cross-section, in which the local microhardness drops below the BM
microhardness. This outer HAZ formed a softened zone (circled in
Fig. 3) which is adjacent to the unaffected BM.
HAZ softening is unavoidable in DP steels in the tempered or
subcritical HAZ [1416]. The lower martensite content in the
original DP600 BM and the relatively lower heat input of Nd:
YAG laser resulted in a shallower and narrower softened zone in
the outer HAZ of the WJ in the present study. The average width of
softened zone is about 40 m, and the microhardness drop is
10 HV. Only partial tempering of martensite occurs in the outer
HAZ of the DP600 WJ, as indicated in Fig. 2(e).
According to the description above, there are local changes in
microstructure and microhardness of the DP600 WJ fabricated by
Nd:YAG laser welding. Hardening in FZ and inner HAZ, and softening in outer HAZ result in a local strength and ductility change,
signicantly affecting the overall tensile properties, deformation
behavior and failure mechanism of the DP600 WJ over a wide strain
rate range. This will be discussed in the following sections.

3.2. Tensile properties


Fig. 4 describes the engineering stressstrain curves of the
DP600 BM and WJ obtained from the tensile tests at strain rates
ranging from quasi-static to dynamic. The strain rate dependent
plastic ow stress behavior of the DP600 BM and WJ in terms of
the YS, UTS, UE and TE were examined and illustrated as a function
of the strain rate in logarithmic scale in Fig. 5.
It can be noted that the DP600 BM and WJ exhibit continuous
yielding behavior at quasi-static strain rates, and develop a yield
point at the strain rates over 0.1 s  1, as indicated in Fig. 4. The
difference between the upper and lower yield points increases as
the strain rate increases. Similar results of the presence of yield
point phenomena in steels under high rate tensile loading were
reported for the DP steel [17], TRIP steel [1820] and IF steel [21],
respectively. In the case of the tensile specimen of WJ, the
presence of yield point phenomena is likely due to the interstitial
diffusion of solute atoms during laser welding process. The high
temperatures generated by the laser beam drove the carbon or
nitrogen atoms in iron to diffuse to the position of high energy just
below the extra plane of atoms in a positive edge dislocation. The
elastic interaction was so strong that the impurity atmosphere
became completely saturated and condensed into a row of atoms
along the core of the dislocation. When the welded specimen with
dislocations pinned by interstitials was loaded, a higher stress was
required to start the dislocation motion representing the onset of
plastic deformation. As a result, the YS of DP600 steel after laser
welding became higher, as shown in Fig. 5(a). This is consistent
with the results for the DP steel joint reported by Farabi et al.
where the testing was performed under static and quasi-static
strain rates [7].
The UTS of DP600 WJ is slightly lower than that of the BM at all
strain rates (Fig. 5(b)), indicating a high joint efciency (maximum

Fig. 4. Engineering stressstrain curves, (a) DP600 BM tested at strain rates from 0.001 to 1 s  1, (b) DP600 BM tested at strain rates from 14 to 1133 s  1, (c) DP600 WJ tested
at strain rates from 0.001 to 1 s  1, and (d) DP600 WJ tested at strain rates from 14 to 1000 s  1.

D. Dong et al. / Materials Science & Engineering A 594 (2014) 1725

21

reduction in the overall specimen elongation, as necking and


failure were observed to occur in the weld during tensile testing.
Both the DP600 BM and WJ exhibit a similar expected non-linear
strain rate dependence on YS and UTS, which is shallow at lower
strain rates and becomes steeper with increasing strain rate (Fig. 5
(a) and (b)). The increase of the strength with increasing strain rate
is well known for positive strain rate sensitive materials.
Fig. 5 also includes the strain-rate dependence of ductility of
the DP600 BM and WJ. The results show that the TE of the DP600
BM and WJ decreases with increasing strain rate from 0.001 to
1 s  1, and then increases up to the strain rate of 100 s  1. This is
somewhat different from the reported strain rate sensitivity
results for mild steel, where a typical reduction of ductility is
observed at higher strain rates. The enhanced ductility at higher
strain rates results from that the positive strain rate sensitivity
extends the post-uniform elongation due to the increased strength
of necking zone by higher strain rates. Since the strain rate of the
localized region is larger than that of the non-localized region, the
strength of the localized zone increases due to the positive strain
rate sensitivity, which in turn results in the increase of strength in
that region and eventually delay of the propagation of localization.
Furthermore, the increase in ductility at higher strain rates is also
related to the thermal softening of the matrix caused by deformation induced adiabatic temperature rise [22]. Similar behavior was
also observed in DP steels where the steel exhibits a better
ductility at high strain rates compared with that at quasi-static
loading condition [5,23,24]. These observations are also consistent
with the high rate sheet formability results presented by Seth et al.
[25] that high rate deformation can be quite effective in diffusing
deformation throughout a specimen leading to increased formability by stabilization against neck growth. The ductility of the
DP600 BM and WJ decreased slightly with further increasing strain
rate above 100 s  1. The plastic instability (or initiation of necking)
is determined by the competition of the effect of strain hardening
or strain rate hardening, and the combination effect of the
retardation of neck propagation at the original localized position
and the thermal softening. With increasing strain rate above
100 s  1, the loss of ductility resulting from the strain rate hardening is larger than the increase of ductility caused by the delay of
the propagation of strain localization and the thermal softening of
matrix. More detailed explanations can be found in our previous
study [26].
As for the DP600 WJ, laser welding led to an overall reduction
in the ductility. However, it exhibited a similar changing trend of
ductility to that of the BM with respect to the strain rate. This
experimental result is worthy of attention for the crashworthiness
of a vehicle, especially in terms of the fracture and tearing of
vehicle members. The results above also reveal that sheet metal
forming process at an adequate strain rate can enhance the
formability compared to the static forming process, especially
the formability of tailor-welded blanks.
3.3. Deformation behavior and failure mechanism
Fig. 5. Tensile properties of the DP600 BM and WJ as a function of the strain rate in
logarithmic scale, (a) YS, (b) UTS, and (c) ductility in terms of UE and TE.

value of 96%). Obviously, the strength of the DP600 WJ fabricated


by Nd:YAG laser welding did not decrease in spite of the presence
of softened zone. This is obviously related to the narrow softened
zone and small drop in microhardness of the present Nd:YAG laser
welded joint, as mentioned above. The ductility of the DP600 WJ is
much lower than that of the BM, as described in terms of the UE
and TE shown in Fig. 5(c). This is due to the fact that yielding
occurred rst in the weld and the subsequent plastic deformation
was predominantly concentrated there. This in turn leads to a

Typical top views of the failure location of the DP600 BM and


WJ are shown in Fig. 6. Generally, all the WJ failed at a zone

Fig. 6. Typical top views of the failure location, (a) DP600 BM tested at 553 s  1 and
(b) DP600 WJ tested at 558 s  1.

22

D. Dong et al. / Materials Science & Engineering A 594 (2014) 1725

adjacent to the FZ, i.e. in HAZ, after tensile testing at various strain
rates. Careful examinations of the failure location were performed
using the high precision measuring microscope and the results are
shown in Fig. 7. It indicates that the tensile failure occurred in the
inner HAZ of the joint, and the distance of failure location from the
weld centerline decreased with increasing strain rate. As shown in
Figs. 8 and 9, the fracture surfaces of the DP600 BM and WJ are
mainly characterized by the cup-like dimple rupture, in spite of

Fig. 7. Variation of tensile failure location of the DP600 WJ with the strain rate.

different fracture locations of the WJ at various strain rates. This


typical ductile fracture behavior did not change with increasing
strain rate in the measured strain rate range.
The typical deformed microstructures near the tensile fracture
surface for the BM and WJ of DP600 steel at various strain rates are
presented in Fig. 10. The fracture mechanisms both for the BM and
WJ are characterized as void nucleation, growth and coalescence.
For the DP600 BM, SEM analysis reveals that void nucleation
occurred mainly by decohesion at the ferrite-martensite interface
(indicated by black arrows in Fig. 10(a)) due to the incompatibility
of plastic deformation locally between the two phases. A few of
voids was also observed on the ferrite matrixes. In the case of the
DP600 WJ, the fracture surfaces exhibit a large amount of smaller
dimples (Fig. 9) compared to those of the BM (Fig. 8) over the
whole strain rate range, which was related to the ne martensite
structure found in inner HAZ. The fracture surfaces also display a
certain number of larger dimples corresponding to the undissolved ferrite during welding. The number of the larger dimples
decreases with increasing strain rate, and the dimples almost
disappear when the strain rate is increased to over 14 s  1.
It indicates that the tensile failure occurred at the interface
between the ne martensite and the undissolved ferrite in the
inner HAZ at the strain rates lower than 14 s  1, as indicated by
black arrows in Fig. 10(b) and (c). When the strain rate continues
to increase, the fracture surfaces of the WJ display smoother
appearances with smaller dimples (Fig. 9(f)(h)). From the above
results, it is seen that the nucleation of the voids can be attributed
to the localized cracking of ne martensite, as shown by white
arrows in Fig. 10(d) and (e).

Fig. 8. Fractographs of the DP600 BM at various strain rates, (a) 0.001 s  1, (b) 0.01 s  1, (c) 0.1 s  1, (d) 1 s  1, (e) 14 s  1, (f) 100 s  1, (g) 553 s  1, and (h) 1133 s  1.

D. Dong et al. / Materials Science & Engineering A 594 (2014) 1725

23

Fig. 9. Fractographs of the DP600 WJ at various strain rates, (a) 0.001 s-1, (b) 0.01 s  1, (c) 0.1 s  1, (d) 1 s  1, (e) 14 s  1, (f) 100 s  1, (g) 558 s  1 and (h) 1000 s-1.

The Nd:YAG laser welded joint of DP600 steel was a heterogeneous structure with large gradient variations in the properties.
The plastic deformation ability and failure mode of the DP600 WJ
is determined by the compatible deformation capability of the
microstructures of different zones across the joint and the strain
partition within the various microstructures throughout the tensile test. At lower strain rates, the stress concentration at the
ferrite-martensite interface is released and relieved through the
compatible deformation of the ferrite. The overall compatible
deformation capability of the inner HAZ decreased signicantly
resulting from the obviously decreasing amount of undissolved
ferrite and increasing amount of ne martensite formed in this
region. The non-uniform deformation rst occurs in the ferrite
microstructure of the inner HAZ, with the microcracks forming
primarily at the ferrite-martensite interfaces due to high degree of
stress concentration. The microcracks then quickly extend to the
fully ne martensite microstructure region in the inner HAZ
(Fig. 10(b) and (c)) until nal failure in this region. With increasing
strain rate, the propagation of strain localization is delayed, and
the strain hardening effect near the ferrite-martensite interface in
the inner HAZ is enhanced, which in turn results in the delay of
the non-uniform deformation of this region. Accelerated microcrack coalescence and propagation occur once there is stress
concentration at the fully martensite region in the inner HAZ
(Fig. 10(d) and (e)) due to its poor compatible deformation
capability. Consequently, the tensile failure occurs in the inner
HAZ adjacent to the FZ and the distance of failure location from
the weld centerline decreases when the strain rate rises above

14 s  1. Fracture occurrence at the softened HAZ has been reported


in other studies [4,5,9], however, this phenomenon is not observed
in the present study under all deformation conditions. This may be
related to the less extensive HAZ softening caused by the lower
heat input in the present Nd:YAG laser welding process and the
lower fraction of martensite in the DP600 BM.
The strain rate sensitivity of the tensile failure location has no
signicant effects on the overall strength and ductility of the DP600
WJ due to the small size of the FZ and HAZ formed during the Nd:
YAG laser welding process. The DP600 WJ exhibits a similar trend of
changes in tensile properties with that of the BM with increasing
strain rate over the whole strain rate range, as indicated in Fig. 5.
3.4. Energy absorption
Signicant effects of the strain rate on the structural energy
absorption and loading capacity have been reported elsewhere
[24,27,28]. For some crashworthiness models, stressstrain data
with as much as 10% strain are most important [24]. In many
applications, this in fact could be closer to the real case, where
only limited deformation is allowed because of passenger safety
reasons. Therefore, the areas under the determined engineering
stressstrain curves of the DP600 BM and WJ up to 10% engineering strain were calculated in this study. The results of absorbed
energy for the DP600 BM and WJ are shown in Fig. 11.
The results showed that the 10% strain-absorbed energy of the
DP600 BM and WJ gradually increases with increasing strain rate,
indicating that the faster the loading is applied, the more the

24

D. Dong et al. / Materials Science & Engineering A 594 (2014) 1725

Fig. 10. Deformed microstructures near the fracture surfaces, (a) DP600 BM tested at 0.001 s  1, (b) DP600 WJ tested at 0.001 s  1, (c) DP600 WJ tested at 14 s  1, (d) DP600
WJ tested at 100 s  1, (e) DP600 WJ tested at 1000 s  1 (where black arrows: occurrence of voids by decohesion at the ferrite-martensite interface, white arrows: occurrence
of voids by cracking of martensite).

4. Conclusions
Butt welded joints of DP600 steel was produced by Nd:YAG
laser welding and characterized with respect to the microstructure, microhardness, mechanical properties, deformation and
fracture behavior. In addition to the conventional quasi-static
tensile test at strain rates from 0.001 to 0.1 s  1, the uniaxial
tensile tests with the strain rate ranging from 1 to 1133 s  1 were
performed on the DP600 BM and WJ using the high speed tensile
testing machine. The strain rate effects on the tensile property,
failure location, fracture mode and energy absorption of the DP600
BM and WJ were also investigated. The main conclusions are
drawn as follows.

Fig. 11. Absorbed energy of the DP600 BM and WJ up to 10% strain at various
strain rates.

material will resist deformation. The DP600 WJ absorbed more


energy over the whole measured strain rates than that of the BM
due to the higher strength at the same strain.

(1) The DP600 WJ fabricated by Nd:YAG laser welding is a


heterogeneous structure with different material properties
due to the local hardening happened in FZ and inner HAZ,
and local softening in outer HAZ. The Nd:YAG laser welded
joint of DP600 steel has a much narrower softened zone with a
smaller microhardness drop due to the less heat input and
lower degree of martensite tempering. This results in higher
UTS with a joint efciency reaching as high as 96% while the

D. Dong et al. / Materials Science & Engineering A 594 (2014) 1725

YS is essentially unaffected and the ductility in terms of TE


becomes lower.
(2) The DP600 WJ exhibits continuous yielding at quasi-static
strain rates and develops a yield point at higher strain rates.
The DP600 WJ exhibits a similar stressstrain response to that
of the BM at all studied strain rates. The YS and UTS increase
with increasing strain rate, while the ductility decreases rst
with increasing strain rate from 0.001 to 1 s  1, and then
increases up to the strain rate of 100 s  1, but decreases again
as the strain rate is higher than 100 s  1. The enhanced
ductility in the proper strain rate range is attributed to the
retardation of the propagation of plastic strain localization due
to the positive strain rate sensitivity and the thermal softening
caused by deformation induced adiabatic temperature rise
during dynamic tensile deformation.
(3) The tensile failure occurs in the inner HAZ of the DP600 WJ,
and the distance of failure location from the weld centerline
decreases with increasing strain rate. The DP600 BM and WJ
show typical ductile fracture patterns characterized by dimples over the whole strain rate range. The mechanism for the
changing failure location at quasi-static and dynamic strain
rates can be related to the different strain rate dependence of
the plastic deformation behavior of the microstructures in
various regions across the joint.
(4) The 10% strain-absorbed energy of the DP600 WJ is higher
than that of the BM, and increases with increasing strain rate
in both cases.

Acknowledgments
This work was nancially supported by National Natural Science
Foundation of China (Grant no. 51101029). The authors would also
like to thank the nancial support from Specialized Research Fund
for the Doctoral Program of Higher Education of China (Grant no.
20110042120025) and Fundamental Research Funds for the Central
Universities of China (Grant no. N110405004).

25

References
[1] A.S. Khan, M. Baig, S.H. Choi, H.S. Yang, X. Sun, Int. J. Plasticity 3031 (2012)
117.
[2] M.S. Rashid, Ann. Rev. Mater. Sci. 11 (1981) 245266.
[3] W. Xu, D. Westerbaan, S.S. Nayak, D.L. Chen, F. Goodwin, E. Biro, Y. Zhou, Mater.
Sci. Eng. A 553 (2012) 5158.
[4] B.L. Boyce, M.F. Dilmore, Int. J. Impact Eng. 36 (2009) 263271.
[5] H. Huh, S.B. Kim, J.H. Song, J.H. Lim, Int. J. Mech. Sci. 50 (2008) 918931.
[6] W. Wang, M. Li, C.W. He, X.C. Wei, D.Z. Wang, H.B. Du, Mater. Des. 47 (2013)
510521.
[7] N. Farabi, D.L. Chen, J. Li, Y. Zhou, S.J. Dong, Mater. Sci. Eng. A 527 (2010)
12151222.
[8] M. Hazratinezhad, N.B. Mostafa Arab, A.R. Suzadeh, M.J. Torkamany, Mater.
Des. 33 (2012) 8387.
[9] C.Y. Kang, T.K. Han, B.K. Lee, J.K. Kim, Mater. Sci. Forum 539543 (2007)
39673972.
[10] R.S. Sharma, P. Molian, Mater. Des. 30 (2009) 41464155.
[11] U. Reisgen, M. Schleser, O. Mokrov, E. Ahmed, J. Mater. Process. Technol. 210
(2010) 21882196.
[12] W. Xu, D. Westerbaan, S.S. Nayak, D.L. Chen, F. Goodwin, Y. Zhou, Mater. Des.
43 (2013) 373383.
[13] M. Xia, N. Screenivasan, S. Lawson, Y. Zhou, Z. Tian, J. Eng. Mater. Technol.
ASME 129 (2007) 446452.
[14] S.K. Panda, N. Sreenivasan, M.L. Kuntz, Y. Zhou, J. Eng. Mater. Technol. ASME
130 (2008) 04100310410039.
[15] E. Biro, J.R. Mcdermid, J.D. Embury, Y. Zhou, Metall. Mater. Trans. A 41 (2010)
23482356.
[16] M.S. Xia, E. Biro, Z.L. Tian, Y. Zhou, ISIJ Int. 48 (2008) 809814.
[17] H.D. Yu, Y.J. Guo, X.M. Lai, Mater. Des. 30 (2009) 25012505.
[18] I. Choi, D.M. Bruce, D.K. Matlock, J.G. Speer, Met. Mater. Int. 14 (2008) 139146.
[19] X. Sun, A. Soulami, K.S. Choi, O. Guzman, W. Chen, Mater. Sci. Eng. A 541 (2012)
17.
[20] J. Van Slycken, P. Verleysen, J. Degrieck, J. Bouquerel, B.C. De Cooman, Mater.
Sci. Eng. A 460461 (2007) 516524.
[21] M. Kuroda, A. Uenishi, H. Yoshida, A. Igarashi, Int. J. Solids Struct. 43 (2006)
44654483.
[22] L.W. Meyer, N. Herzig, T. Halle, F. Hahn, L. Krueger, K.P. Staudhammer, J. Mater.
Process. Technol. 182 (2007) 319326.
[23] J.H. Kim, D. Kim, H.N. Han, F. Barlat, M.G. Lee, Mater. Sci. Eng. A 559 (2013)
222231.
[24] S. Curtze, V.-T. Kuokkala, M. Hokka, P. Peura, Mater. Sci. Eng. A 507 (2009)
124131.
[25] M. Seth, V.J. Vohnout, G.S. Daehn, J. Mater. Process. Technol. 168 (2005)
390400.
[26] D.Y. Dong, Y. Liu, L. Wang, L.J. Su, Acta Metall. Sin. 49 (2013) 159166.
[27] W.R. Wang, M. Li, C.W. He, X.C. Wei, D.Z. Wang, H.B. Du, Mater. Des. 47 (2013)
510521.
[28] X.C. Wei, R.Y. Fu, L. Li, Mater. Sci. Eng. A 465 (2007) 260266.

Vous aimerez peut-être aussi