Вы находитесь на странице: 1из 11

Journal of Materials Processing Technology 176 (2006) 168178

Pitting and galvanic corrosion behavior of laser-welded stainless steels


C.T. Kwok a , S.L. Fong a , F.T. Cheng b, , H.C. Man c
a

Department of Electromechanical Engineering, University of Macau, Taipa, Macau, China


Department of Applied Physics, The Hong Kong Polytechnic University, Hung Hom, Kowloon, Hong Kong, China
c Department of Industrial & Systems Engineering, The Hong Kong Polytechnic University, Hung Hom, Kowloon, Hong Kong, China
b

Received 18 October 2005; received in revised form 16 February 2006; accepted 21 March 2006

Abstract
Autogenous welded specimens of austenitic (S30400 and S31603), duplex (S31803) and super duplex (S32760) stainless steels were fabricated
by laser penetration welding (LPW) with a CW Nd:YAG laser in an argon atmosphere. The microstructure and the phases present in the resolidified
zone of the laser-welded specimens were analyzed by optical microscopy and X-ray diffractometry, respectively. The pitting and galvanic corrosion
behavior of the stainless steels in the laser-welded and unwelded conditions in 3.5% NaCl solution at 23 C was studied by means of electrochemical
measurements. X-ray diffraction analysis showed that the phases present in the weld metal depended on the composition of the base metal. While
the laser weld for S31603 retained the original austenitic structure, the laser weld of S30400 contained austenite as the major phase and -ferrite
as the minor phase. On the other hand, a slight change of -ferrite to austenite ratio was found in both the laser welds of S31803 and S32760, with
austenite present at the -ferrite grain boundaries. The welds exhibited passivity but their pitting corrosion resistance was in general deteriorated
as evidenced by a lower pitting potential and a higher corrosion current density compared with those of the unwelded specimens. The decrease
in pitting corrosion resistance of the welds was attributed to microsegregation in the weld zone of S31603, and to the presence of -ferrite in
S30400. For the duplex grades S31803 and S32760, the disturbance of the ferrite/austenite phase balance in the weld zone could be the cause of
the decrease in corrosion resistance. The initial free corrosion potentials of the unwelded specimens were considerably higher than those of the
corresponding laser welds, indicating that the welds were more active and were expected to act as anodes in the weldment. The ranking of galvanic
current densities (IG ) of the couples formed between the laser-welds (LW) and the as-received (AR) specimens with area ratio 1:1, in ascending
order, is: AR S31603/LW S31603 < AR S31803/LW S31803 < AR S32760/LW S32760 < AR S30400/LW S30400. The recorded IG in all couples
was low (in the range of nA/cm2 ).
2006 Elsevier B.V. All rights reserved.
Keywords: Laser welding; Stainless steels; Pitting corrosion; Galvanic corrosion; -Ferrite

1. Introduction
Owing to their excellent mechanical properties and corrosion
resistance, stainless steels are extensively used in many industrial and medical applications. Commercial products such as
razors, cigarette lighters, watch springs, motor and transformer
lamination, hermetic seals, battery and pacemaker cans, and
hybrid circuit packages require delicate welds with high quality and precision. A kilowatt laser beam can melt and vaporize
the material, and the pressure of the vapor displaces the molten
material so that a narrow and deep keyhole is formed. The keyhole supports the transfer of the laser energy into the material
and guides the laser beam deep into the material. Laser pene-

Corresponding author. Tel.: +852 2766 5691; fax: +852 2333 7629.
E-mail address: apaftche@polyu.edu.hk (F.T. Cheng).

0924-0136/$ see front matter 2006 Elsevier B.V. All rights reserved.
doi:10.1016/j.jmatprotec.2006.03.128

tration welding (LPW) can produce low-distortion and precise


weldments with minimal heat-affected zones [1,2]. In LPW of
stainless steels, phase transformation is common. The mechanical properties and corrosion resistance of laser-welded stainless
steels may be deteriorated due to microsegregation, unfavorable
phase content, presence of porosities, solidification cracking,
micro-fissures and loss of materials by vaporization. Galvanic
cell may also be set up between different parts of the weldment. Galvanic corrosion in weldments should not be overlooked
because it can lead to accelerated deterioration of the anodic
region especially in hostile environments. Pitting corrosion and
galvanic corrosion have been investigated in the couples between
dissimilar alloys such as 316L, Ti, Nb and Ta [3], CoCr and REX
734 [4], annealed and cold-worked 316L [5], and GTAW welded
and unwelded N08031 [6]. The pitting corrosion resistance of
several austenitic stainless steels welded by a CO2 laser has been
investigated by Vilpas [7]. However, reports on the galvanic

C.T. Kwok et al. / Journal of Materials Processing Technology 176 (2006) 168178

corrosion behavior in laser weldments of stainless steels are


scarcely reported. The aim of the present study is to investigate the pitting corrosion behavior of various stainless steels in
the laser-welded and unwelded conditions in 3.5% NaCl solution, and the galvanic effect in the laser weldments. The change
in the pitting corrosion behavior of the laser-welded specimens
and the presence of galvanic effect in the laser-weldments were
explained in terms of the change in microstructure and chemical
compositions. It must be pointed out that no attempt of optimization of processing parameters was made. The present study
thus would serve only as a preliminary baseline investigation for
reference in future studies on laser welding of stainless steels.
Moreover, detailed mechanical properties will be reported elsewhere.
2. Experimental details
Austenitic (UNS S30400 and S31603), duplex (UNS S31803) and super
duplex (S32760) stainless steels with different chemical compositions (Table 1)
were selected in the present study. The as-received (AR) stainless steels were in
the form of plates with a thickness of 1 mm. LPW was carried out using a highpower CW Nd:YAG laser with a power of 0.9 kW and a beam size of 0.5 mm in
diameter (6.1 105 W/cm2 ). The laser beam was transmitted by an optical fibre
and focused onto the specimen by a BK-7 lens with a focal length of 80 mm. The
flexible optical fibre delivery was controlled by an XY table. Argon flowing
at 20 l/min was used as the shielding gas. In order to reduce thermal distortion
of the workpiece, it was held in place by a clamping device. Beam scanning
speed of 35 mm/s was used. Such laser parameters were chosen for achieving
full penetration and minimal thermal distortion. The bead-on-plate seam welds
were made on the plates by line scanning of the focused laser beam.
The laser-welded specimens were sectioned, polished and etched. The
microstructure and phases in the resolidified zones were analyzed by optical
microscopy (OM) and X-ray diffractometry (XRD), respectively. The radiation
source of the X-ray diffractometer was Cu K with nickel filter and generated
at 1.2 kW and the scan rate was 0.25 s1 , with the scanning direction along the
weld. The respective volume fraction of -ferrite present in the stainless steels
was evaluated by the direct comparison method [8]. Due to rapid solidification,
the specimens exhibited a preferred orientation to some extent. The integrated
intensities of the (1 1 1), (2 0 0) and (2 2 0) diffraction peaks for -austenite and
the (1 1 0) and (2 0 0) peaks for -ferrite were taken into account. The volume
fraction of -ferrite (C ) was calculated using the following expression according to the ASTM Standard E974 [9]:

C = 1 +

= 1+

(I(1 1 1) /R(1 1 1) ) + (I(2 0 0) /R(2 0 0) ) + (I(2 2 0) /R(2 2 0) )


1.5((I(1 1 0) /R(1 1 0) ) + R(I(2 0 0) /R(2 0 0) ))
(I(1 1 1) /182.8) + (I(2 0 0) /81.6) + (I(2 2 0) /44.4)
1.5((I(1 1 0) /233.8) + (I(2 0 0) /31.9))

169

the hardness of the laser-welded specimens was determined using a Vickers


microhardness tester. The load applied was 200 g and the loading time was 15 s.
Cyclic potentiodynamic polarization scans were carried out using an EG&G
PARC 273 corrosion system according to ASTM Standard G61-94 [11] for
investigating the pitting corrosion behavior. Free corrosion potential (Ec ) measurement and galvanic corrosion test were conducted using the same corrosion
system conforming to ASTM Standard G71-81 [12]. The base stainless steels
and their welds for corrosion studies were cut into 6 mm 10 mm plates. Prior
to corrosion tests, the surface of all specimens was freshly ground and then
polished with 1 m-diamond paste in order to keep the surface roughness consistent. The specimens were embedded in epoxy resin with an exposed area
of 4 mm2 of the weld track or unwelded region. The specimens were cleaned,
degreased and dried before the polarization test in 3.5% NaCl solution, which
was kept at a constant temperature of 23 C and open to air. A saturated calomel
electrode (SCE) was used as the reference electrode and two parallel graphite
rods served as the counter electrode for current measurement. For the cyclic
potentiodynamic polarization test, all data were recorded after an initial delay of
30 min for the specimen to stabilize. The potential was increased from 200 mV
below the corrosion potential in the anodic direction at a scan rate of 5 mV s1 .
The scan was then reversed when an anodic current density of 5 mA cm2 was
reached and continued until the loop closed at the protection potential. Galvanic
corrosion test was performed using the built-in zero-resistance ammeter (ZRA)
in the potentiostat. The galvanic current density (IG ) and galvanic potential (EG )
in the couples formed by the as-received specimen and the laser weld were continuously monitored for 24 h. The exposed area ratio of anode to cathode of all
galvanic couples was 1:1.

3. Results and discussion


3.1. Microstructural and metallographic analysis
Full penetration was achieved in all specimens in the LPW
and the widths of welds were about 0.8 mm. Typical crosssectional view of laser-welded (LW) S31603 is shown in Fig. 1.

1

1

(1)

where I(h k l) and I(h k l) are the integrated intensities of a given crystallographic
plane (h k l) from the and phases, respectively, and the values of R(h k l)
and R(h k l) of and for various planes were obtained from Jatczak et al.
[10]. The chemical compositions of the resolidified microstructure after LPW
were analyzed by energy dispersion X-ray spectrometry (EDS). In addition,

Fig. 1. Cross-sectional view of laser-welded S31603.

Table 1
Nominal compositions (wt.%) of various stainless steels

S30400
S31603
S31803
S32760
a

Fe

Cr

Ni

Mo

Mn

Cu

Si

Creq /Nieq a

Balance
Balance
Balance
Balance

18.4
17.6
22.5
25.6

8.7
11.2
5.6
7.2

2.5
2.9
4.0

1.6
1.4
1.5
0.6

0.2
0.8

2.1
1.4
1.6
0.7

0.08
0.03
0.03
0.03

0.3
0.4
0.4
0.3

0.1

0.1

0.2

1.70
1.62
3.59
2.09

Creq = [Cr] + [Mo] + 1.5[Si] + 0.5[Cb]. Nieq = [Ni] + 0.5[Mn] + 30[C] + 30[N].

170

C.T. Kwok et al. / Journal of Materials Processing Technology 176 (2006) 168178

Fig. 2. XRD patterns of as-received stainless steels and laser welds: (a) S30400, (b) S31603, (c) S31803 and (d) S32760.

X-ray diffraction analysis showed that the phases present in the


weld metals depended on the composition of the base metal.
According to the XRD patterns in Fig. 2, the laser weld for
S31603 retained the original austenitic structure, while the laser
weld for S30400 contained austenite as the major phase and -

ferrite as the minor phase. On the other hand, both S31803 and
S32760 were mainly composed of -ferrite, with austenite as the
minor phase. The microstructures of the stainless steels before
and after laser welding are shown in Fig. 3. In LW S30400, the
skeletal network of residual -ferrite is present in the austenitic

C.T. Kwok et al. / Journal of Materials Processing Technology 176 (2006) 168178

matrix () as shown in Fig. 3(e). Austenitic dendrites in different orientations are observed in LW S31603 as shown in
Fig. 3(f). After laser welding, the grain size of LW S30400 and
LWS31603, which were predominantly austenitic, was refined
due to rapid solidification. On the contrary, the grain size in LW

171

S31803 and LW S32760 was increased and their microstructures


are shown in Fig. 3(g and h), respectively. The Widmanstatten
structure of semi-continuous dendritic austenite was present in
the columnar grain boundaries of ferrite. When the melt pool
of the weld zone of the duplex grade stainless steels solidifies,

Fig. 3. Microstructure of various stainless steels in unwelded (as-received, AR) condition: (a) AR S30400, (b) AR S31603, (c) AR S31803 and (d) AR S32760; and
in laser-welded (LW) condition: (e) LW S30400, (f) LW S31603, (g) LW S31803 and (h) LW S32760.

172

C.T. Kwok et al. / Journal of Materials Processing Technology 176 (2006) 168178

Fig. 3. (Continued ).

C.T. Kwok et al. / Journal of Materials Processing Technology 176 (2006) 168178
Table 2
Volume fraction of -ferrite and microhardness in as-received (AR) and laserwelded (LW) regions
Stainless
steels

C (AR)
(%)

C (LW)
(%)

Difference
in C (%)

Hv0.2
(AR)

Hv0.2
(LW)

S30400
S31603
S31803
S32760

0
0
60
54

14
0
62
50

+14
0
+2
4

176
179
268
290

301
194
316
314

the possible phase transformation sequence upon cooling may


be represented as:
Liquid liquid + ferrite ferrite ferrite + austenite
The degree of completion of the transformation and hence the
final phase structure of the weld metal depend on the composition of the base metal and the welding parameters [13].
Based on the XRD spectra, the volume fractions of -ferrite
(C ) of various laser welds were calculated using Eq. (1) and
shown in Table 2. -ferrite did not exist in LW S31603, whereas
14%, 62% and 50% of -ferrite were present in LW S30400,
LW S31803 and LW S32760, respectively. The effect of laser
welding on the microstructural change of various stainless steels

173

under the same laser processing conditions was different because


of their difference in chemical compositions. The ratios of the
chromium equivalent (Creq ) and the nickel equivalent (Nieq ) of
the stainless steels are shown in Table 1. As the Creq /Nieq ratio
increased, the ferrite-forming tendency of the stainless steels
increased. The highest volume fraction of -ferrite was observed
in S31803, which had the highest Creq /Nieq ratio. In addition, the
solid-state transformation of -ferrite to austenite is considered
to be diffusional. Thus, the high solidification rate typical in laser
processing would also suppress the ferrite-to-austenite transformation, resulting in a high fraction of -ferrite. The increase
in volume fraction of -ferrite in LW S31803 in the present
study was also reported by others [14,15] in the autogenous laser
welding of duplex stainless steels. The disturbance of the ferrite/austenite phase balance in the weld metal might be remedied
via the use of welding consumables having a more austenitic
composition, and/or the use of a shielding gas containing an
appropriate amount of N2 , which is an austenite stabilizer [16].
However, there is a higher probability of forming intermetallic
precipitates and nitrides in the weld metal, both of which would
decrease the corrosion resistance [2,17].
The hardness profiles along the depth and across the crosssection of the weld zones of various specimens are shown in
Fig. 4, and the results are summarized in Table 2. The hardness

Fig. 4. Hardness profiles of various laser-welded specimens: (a) along the depth of the cross-section and (b) across the cross-section.

174

C.T. Kwok et al. / Journal of Materials Processing Technology 176 (2006) 168178

Fig. 5. Ec vs. time of the as-received stainless steels and their corresponding weldments: (a) S30400, (b) S31603, (c) S31803 and (d) S32760 in 3.5% NaCl solution.

values in the welds were nearly constant and higher than those
of the base metals (hardness outside the welds) as shown in
Fig. 4(b). The hardness of the weld zone for S31603, S31803
and S32760 was higher by 818% as compared with that of the
base metals. The most significant change in hardness is observed
in LW S30400, with an increase of about 70%. The increase in
hardness could be attributed to the refinement of grains and also
to the presence of hard -ferrite.
3.2. Free corrosion potential
From measurements of the free corrosion potentials (Ec ) of
the base stainless steels and their corresponding laser welds
(Fig. 5), information about the dynamic behavior of the passive
oxide film might be obtained. The Ec of the unwelded specimens increased towards more noble values and became steady
at the end of the 2-h test. This reflects that the growth of passive
oxide was almost complete. The steady values of Ec after 2 h are
shown in Table 3. For unwelded S32760, the oxide film was the
most stable because it was highly alloyed with the elements Cr,
Mo and N, all of which could enhance passivity. It can be also
observed that the Ec of all unwelded stainless steels are considerably higher than those of their corresponding welds. For
instance, the Ec of as-received S32760 (3 mV SCE) is higher
than that of its laser weld (327 mV SCE).

3.3. Pitting corrosion behavior


Cyclic potentiodynamic polarization curves of various
unwelded stainless steels and their laser welds in 3.5% NaCl
solution at 23 C are shown in Fig. 6. The corrosion parameters,
including pitting potential (Epit ), protection potential (Eprot ) and
corrosion current density (icorr ), are summarized in Table 3. All
unwelded stainless steels and their laser welds exhibited passivity in 3.5% NaCl solution. However, there was a general and
substantial shift of the polarization curves towards higher current densities, indicating deterioration in corrosion resistance.

Table 3
Corrosion parameters of unwelded (AR) and laser-welded (LW) stainless steels
in 3.5% NaCl solution at 23 C, open to air
Specimens

Ec (mV)

Epit (mV)

Eprot (mV)

icorr (A/cm2 )

AR S30400
LW S30400
AR S31603
LW S31603
AR S31803
LW S31803
AR S32760
LW S32760

270
310
256
343
301
348
3
327

330
85
423
195
1170
671
1040
1040

39
11
76
39
1179
111
980
954

0.416
3.475
0.252
6.456
0.405
9.429
0.138
0.518

C.T. Kwok et al. / Journal of Materials Processing Technology 176 (2006) 168178

175

Fig. 6. Potentiodynamic polarization curves of the as-received stainless steels and their corresponding weldments: (a) S30400, (b) S31603, (c) S31803 and (d)
S32760 in 3.5% NaCl solution.

In addition, the pitting corrosion resistance and the repassivation capability of the laser welds were lower than those of the
unwelded specimens as reflected by lower values in both Epit
and Eprot . The ranking of the pitting corrosion resistance of the
specimens is as follows:
ARS30400 < ARS31603 < ARS32760 < ARS31803
and
LWS30400 < LWS31603 < LWS31803 < LWS32760.
Decrease in Epit is observed in the laser welds for S30400
(from 330 mV to 85 mV SCE), S31603 (from 423 mV to 195 mV
SCE) and S31803 (from 1170 mV to 671 mV SCE). On the other
hand, there is no significant change in Epit in the laser weld for
S32760. In addition, the corrosion current densities of all welds
increased.
While the results in the present study indicate a decrease in
corrosion resistance for stainless steels due to laser welding,
laser cladding (LC) of stainless steel on mild steel followed
by laser remelting resulted in increase of corrosion resistance
as reported by Li et al. [18,19]. This is not unexpected as in
laser surfacing the processing condition is chosen to yield a
homogeneous surface layer while in welding the primary aim is
to achieve joining. Some improvement in pitting resistance of

laser-surface melted S30400 [2023] and S31603 [20,24] was


reported by several authors due to the removal of MnS inclusions (for S30400 and S31603) and the trapping of sulfur in
the -ferrite (for S30400). For S32760, there is no significant
change in Epit and Eprot resulting from laser surface melting
[20], consistent with the present results.
The change in corrosion behavior due to laser welding could
arise from different causes depending on the base stainless steel.
Pan et al. attributed the deleterious effect of -ferrite on the
pitting corrosion resistance of S32100 to the galvanic effect
existing between -austenite and -ferrite [25]. In fusion welding, solidification from the melt pool in general results in local
compositional variations, which would in turn result in less stable passive film and hence lower corrosion resistance [26]. The
compositional heterogeneity in the weld metal could arise from
three main causes: microsegregation during weld metal solidification, element partition in solid-state transformation from
ferrite to austenite, and precipitation of intermetallic phases,
carbides and nitrides, leading to the formation of Cr-depleted
regions. The microstructure in a weld is a complex function
of the solidification parameters (solidification rate and temperature gradient at the solid/liquid interface), which in turn are
determined by the processing parameters and the alloy composition [27,28]. Though microsegregation is generally small in

176

C.T. Kwok et al. / Journal of Materials Processing Technology 176 (2006) 168178

Table 4
EDS composition analysis of different regions in laser welds
Specimen

Region

Cr (wt.%)

Ni (wt.%)

Mo (wt.%)

LW S30400

16.4
19.9

7.8
8.2

LW S31603

Dendrite core ()
Interdendritic ()

16.1
18.2

9.6
12.3

1.5
2.1

LW S31803

20
24

6.3
5.4

1.8
2.6

LW S32760

26.7
26.1

7.1
7.2

3.2
3.8

laser welding in comparison with conventional welding because


of a higher solidification rate in the former, some degree of
microsegregation still occurred in the laser welds of various
stainless steels in the present study. EDS compositional analysis of the different phases or regions in the laser welds is
shown in Table 4. For LW S31603, the difference in composition
between the dendritic cores and the interdendritic regions reveals
microsegregation during solidification, with a subsequent Cr and
Mo enrichment of the liquid phase when the solidification front
grew from the core to the boundary of the dendrites while Ni
segregated in the opposite direction, similar to that reported in
the literature [29]. For LW S30400, which exhibited primary
austenitic solidification, a small amount of ferrite was formed
from the melt between dendrites as result of microsegregation.
For LW S31803, which was predominantly ferritic, difference
in the composition between austenite and ferrite existed, similar to the case of LW S30400. The Cr, Ni and Mo contents
in the austenite and ferrite for LW S32760 were very close in
each phase, indicating minimal microsegregation. The impairment in the pitting corrosion resistance of the laser welds might
be attributed to microsegregation, and also to unfavorable ferrite/austenite phase content [30], in addition to the presence of
defects arising from welding. The corrosion behavior of LW
S32760 was relatively close to that of the base metal because
microsegregation was minimal.
3.4. Galvanic corrosion behavior
Plots of galvanic potentials (EG ) and galvanic current densities (IG ) of the galvanic couples between the base stainless
steels and their corresponding laser welds as a function of time
are shown in Fig. 7. Since the values of the Ec for all base stainless steels were higher than those of their welds (Fig. 5), the
welds were more active and are expected to act as the anode
when coupled to the corresponding base metals. From Fig. 7, it
can be observed that the current densities of all the couples were
changing with time in the initial stage and then reached different steady-state values at the end of the 24-h test. The values of
EG and IG for different galvanic couples attained after 24 h are
shown in Table 5. The ranking of IG in the galvanic couples in
ascending order is as follows:
ARS31603/LWS31603 < ARS31803/LWS31803
< ARS32760/LWS32760 < ARS30400/LWS30400

Fig. 7. Time dependence of (a) EG and (b) IG for various galvanic couples in
3.5% NaCl solution.

The galvanic corrosion rates (i.e. IG ) in all these couples were


low, only in the range of nA/cm2 . The galvanic corrosion rate in a
galvanic couple depends on the difference in corrosion potentials
(the driving force) of the members in the galvanic couple and on
their polarization characteristics (the resistance), both of which
in turn depend on the compositions and microstructures of the
members [31].
IG for AR S30400/LW S30400 was the highest due to a large
difference in volume fraction of -ferrite for members in this
couple (+14%). The amount of -ferrite in the austenite matrix
determines the Ec , and hence the IG in the couple. The values
of IG in AR S32760/LW S32760 and AR S31803/LW S31803
were much smaller since the difference in volume fraction of Table 5
Steady-state EG and IG of galvanic couples between the unwelded stainless steels
and their weldments
Galvanic couples

EG (mV)

IG (nA/cm2 )

AR S30400/LW S30400
AR S31603/LW S31603
AR S31803/LW S31803
AR S32760/LW S32760

311
330
304
347

78.6
26
8.6
35

C.T. Kwok et al. / Journal of Materials Processing Technology 176 (2006) 168178

ferrite was relatively small. IG was the lowest in AR S31603/LW


S31603 because there was no change in phase structure (100%
austenite) in this case, and galvanic corrosion was attributable
to the minor changes in microstructure. Nevertheless, the galvanic couples formed between the base stainless steels and their
welds exhibited negligible galvanic effect as evidenced by the
low values of IG , only of the order of nA/cm2 . There is no risk
of triggering the phenomenon of pitting corrosion, because the
Epit of the welds for various stainless steels were in the range
of 851040 mV versus SCE and much higher than the EG of
various couples (ranging from 347 mV to 304 mV SCE). In
addition, the welded surface of different stainless steels after the
galvanic corrosion tests did not reveal any surface damage such
as pits or discoloration.
4. Conclusions
Autogenous laser welding of two austenitic and two duplex
stainless steels in Ar atmosphere was attempted and the pitting
and galvanic corrosion behavior of the weldments were studied.
The following conclusions can be drawn:
1. The laser weld of S30400 was essentially austenitic with the
presence of a small amount of ferrite while the laser weld of
S31603 remained purely austenitic. On the other hand, the
ferrite/austenite phase balance for the two duplex stainless
steels was slightly disturbed.
2. The microhardness of the laser welds for various stainless
steels generally increased, possibly due to the increase in the
volume fraction of ferrite, or to grain refinement.
3. All laser welds for stainless steels exhibited passivity in
3.5% NaCl solution but their pitting resistance deteriorated
as evidenced by lower pitting potentials and higher corrosion
current densities compared with those of the base metals. It
is attributed to microsegregation (for all stainless steels studied), to the presence of -ferrite (S30400) or to incorrect
phase balance (S31803 and S32760).
4. Galvanic current densities in the couples formed between the
base stainless steels and their welds were very low (in range
of nA/cm2 ), indicating very small galvanic effect.
Acknowledgments
The authors wish to acknowledge the support from the infrastructure of the University of Macau and the Hong Kong Polytechnic University.
References
[1] J. Mazumder, Laser beam welding, ASM Handbook, vol. 6, 10th ed.,
Welding, Brazing, and Soldering, ASM International, Materials Park,
OH, USA, 1990, pp. 263268.
[2] D. Schuocker, Welding with lasers, in: High Power Lasers in Production
Engineering, Imperial College Press, UK, 1999, pp. 337370.
[3] J. Gluszek, J. Masalski, Galvanic coupling of 316L steel with titanium,
niobium, and tantalum in Ringers solution, Br. Corros. J. 27 (1992)
135138.

177

[4] L. Reclaru, R. Lerf, P.Y. Eschler, A. Blatter, J.M. Meyer, Pitting, crevice
and galvanic corrosion of REX stainless-steel/CoCr orthopedic implant
material, Biomaterials 23 (2002) 34793485.
[5] C.C. Shih, C.M. Shih, Y.Y. Su, S.J. Lin, Galvanic current induced by
heterogeneous structures on stainless steel wire, Corros. Sci. 47 (2005)
21992212.
[6] E. Blasco-Tamarit, A. Igual-Munoz, J. Garcia Anton, D. Garcia-Garcia,
Effect of aqueous LiBr solutions on the corrosion resistance and galvanic
corrosion of an austenitic stainless steel in its welded and non-welded
condition, Corros. Sci. 48 (2006) 863886.
[7] M. Vilpas, Prediction of Microsegregation and Pitting Corrosion Resistance of Austenitic Stainless Steel Welds by Modelling, VTT Publications, 390 VTT, Espoo, 1999, p. 70.
[8] B.D. Cullity, Elements of X-Ray Diffraction, Addison-Wesley, NY, 1978,
p. 411.
[9] ASTM Standard E975-03, Standard Practice for X-Ray Determination
of Retained Austenite in Steel with Near Random Crystallographic Orientation, ASTM International, Philadelphia, 2003.
[10] C.F. Jatczak, J.A. Larson, S.W. Shin, Retained Austenite and its Measurements by X-Ray Diffraction, SP-453, SAE, Philadelphia, 1980, p.
14.
[11] ASTM Standard G61-94, Standard Method for Conducting Cyclic Potentiodynamic Polarisation Measurements for Localised Corrosion Susceptibility of Iron-, Nickel-, or Cobalt-Based Alloys, Annual Book of ASTM
Standards, ASTM International, Philadelphia, 1994.
[12] ASTM Standard G71-81, Conducting and Evaluating Galvanic Tests in
Electrolytes, Annual Book of ASTM Standards, ASTM International,
Philadelphia, 1992.
[13] D.N. Noble, Selection of wrought duplex stainless steels, ASM Handbook, vol. 6, 10th ed., Welding, Brazing, and Soldering, ASM International, Materials Park, OH, USA, 1990, pp. 471481.
[14] J.M. Robinson, R.C. Reed, A. Camyab, CO2 laser welding of duplex
and super-duplex stainless steels (the effect of argonnitrogen assist gas
mixtures), in: Proceedings of the 4th International Conference on Trends
in Welding Research, TN, USA, 1995, pp. 499504.
[15] N.A. McPherson, H. Samson, T.N. Baker, N. Suarez-Fernandez, Steel
microstructure in autogenous laser welds, J. Laser Appl. 15 (2003)
200210.
[16] V. Muthupandi, P. Bala Srinivasan, S.K. Seshadri, S. Sundaresan, Corrosion behavior of duplex stainless steel weld metals with nitrogen
additions, Corros. Sci. Technol. 38 (2003) 303308.
[17] N.A. McPherson, K. Chi, T.N. Baker, Submerged arc welding of stainless steel and the challenge from the laser welding process, J. Mater.
Process. Technol. 134 (2003) 174179.
[18] R. Li, M.G.S. Ferreira, M.A. Anjos, R. Vilar, Localized corrosion of
laser surface cladded UNS S31254 superaustenitic stainless steel on
mild steel, Surf. Coat. Technol. 88 (1997) 9095.
[19] R. Li, M.G.S. Ferreira, M.A. Anjos, R. Vilar, Localized corrosion performance of laser surface cladded UNS S44700 superferritic stainless
steel on mild steel, Surf. Coat. Technol. 88 (1997) 96102.
[20] C.T. Kwok, H.C. Man, F.T. Cheng, Cavitation erosion and pitting corrosion of laser surface melted stainless steels, Surf. Coat. Technol. 99
(1998) 295304.
[21] E. McCafferty, P.G. Moore, Electrochemical behavior of laser-processed
metal surfaces, J. Electrochem. Soc. 133 (1986) 10901096.
[22] O.V. Akgun, T.O. Inal, Desensitization of sensitized 304 stainless steel
by laser surface melting, J. Mater. Sci. 30 (1992) 21472153.
[23] M. Carbucicchio, G. Palombarini, M. Rateo, G. Sambogna, Improvement of surface properties of AISI 304 stainless steel by laser melting,
Hyperfine Interact. 112 (1998) 1924.
[24] C. Carboni, P. Peyre, G. Beranger, C. Lemaitre, Influence of high power
diode laser surface melting on the pitting corrosion resistance of type
316L stainless steel, J. Mater. Sci. 37 (2002) 37153723.
[25] Q.Y. Pan, W.D. Huang, R.G. Song, Y.H. Zhou, G.L. Zhang, The
improvement of localized corrosion resistance in sensitized stainless steel
by laser surface remelting, Surf. Coat. Technol. 102 (1998) 245255.
[26] T.G. Gooch, Corrosion behavior of welded stainless steel, Weld. J. 75
(1996) 135s154s.

178

C.T. Kwok et al. / Journal of Materials Processing Technology 176 (2006) 168178

[27] S.A. David, S.S. Babu, J.M. Vitek, Welding: solidification and
microstructure, JOM 55 (2003) 1420.
[28] N. Suutala, Effect of solidification condition on the solidification mode
in austenitic stainless steels, Metall. Trans. A 14 (1983) 191197.
[29] W.R. Cieslak, D.J. Duquette, W.F. Savage, Pitting corrosion of stainless
steel weldments, in: S. David (Ed.), ASM Conference on Trends in
Welding Research in the United States, New Orleans, USA, November
1618, 1981, pp. 361379.

[30] A. Wahid, D.L. Olson, D.K. Matlock, Corrosion of weldments,


ASM Handbook, vol. 6, 10th ed., Welding, Brazing, and Soldering, ASM International, Materials Park, OH, USA, 1990, pp. 1065
1069.
[31] H.P. Hack, Galvanic corrosion, in: R. Baboian (Ed.), Corrosion Tests
and Standards: Application and Interpretation, ASTM STP 978, ASTM,
1995, pp. 186196.

Вам также может понравиться