Sunteți pe pagina 1din 13

Corrosion Science 52 (2010) 2439–2451

Contents lists available at ScienceDirect

Corrosion Science
journal homepage: www.elsevier.com/locate/corsci

Corrosion resistance of welds in type 304L stainless steel made


with a nickel–copper–ruthenium welding consumable
D. Liang a, J.W. Sowards b, G.S. Frankel a,*, B.T. Alexandrov b, J.C. Lippold b
a
Department of Materials Science and Engineering, The Ohio State University, Columbus, OH 43210, USA
b
Welding Engineering, The Ohio State University, Columbus, OH 43210, USA

a r t i c l e i n f o a b s t r a c t

Article history: Ni–Cu–Pd welding consumables have been recently developed for 300-series austenitic stainless steels
Received 14 January 2010 such as Type 304L (SS304L) to reduce the amount of Cr(VI) in the welding fumes. In this study, a modified
Accepted 6 March 2010 filler metal that replaces Pd with Ru was evaluated. Initial tests conducted on button-melted samples and
Available online 12 March 2010
bead-on-plate welds indicated that Ni–Cu–Ru exhibited good corrosion properties. Actual Ni–Cu–Ru arc
welds made on SS304L were successfully produced and the corrosion performance was comparable to or
Keywords: better than that of Ni–Cu–Pd welds. These welds are a suitable replacement for welds made with stan-
A. Ni alloys
dard 300-series welding consumables, such as SS308L.
A. Stainless steel
C. Localized corrosion
Ó 2010 Elsevier Ltd. All rights reserved.
C. Welding

1. Introduction Any welding process for stainless steel must result in welds
with suitable mechanical and corrosion properties. From a corro-
Stainless steel structures are often fabricated by arc welding sion perspective, the criteria for designing a new filler metal for
processes that use Cr-containing welding consumables. Owing to welding SS304 were described previously [4,7] and specified that:
the hazards associated with Cr(VI) in weld fumes generated by (1) The corrosion potential of the weld metal should be slightly
welding with Cr-containing consumables, it is of interest to devel- higher than the corrosion potential of SS304 to minimize galvanic
op Cr-free consumables. In previous work, welds in SS304L made attack of the weld metal; (2) The breakdown potential (Eb) and
with Ni–Cu or Ni–Cu–Pd welding consumables were successfully repassivation potential (Erp) of the weld metal should be higher
produced and their corrosion properties were found to be compa- than the corrosion potential of SS304 to prevent localized corrosion
rable to those made with a standard SS308L consumable [1]. The of the weld metal.
addition of Pd to Ni–10Cu was beneficial for corrosion resistance, The prices of pure metals fluctuate with time, but ruthenium is
especially the resistance to localized corrosion. Pd improved the usually the least expensive platinum group metal with a cost that
repassivation potential, even when present in concentrations as is substantially lower than that of Pd. Additions of Ru have been
low as 1 wt.%. Pd concentrations in the weld consumable lower found to increase the corrosion resistance of passive metals and al-
than 1% were not as effective due to dilution with the SS304L base loys including stainless steels [8–11], Ti [12–14], and Cr [15].
metal [2]. Increasing Pd above 1% improved the corrosion proper- Remarkable improvement of the corrosion resistance of Ru-alloyed
ties of Ni–Cu welds on SS304L, but Pd is very expensive and adds Cr has been reported in various solutions, and comparable corro-
a considerable cost to the consumable even at a fraction of a per- sion resistance with Pd-alloyed Cr alloys was achieved in 40%
cent of total alloy concentration [3]. In this work, efforts were fo- sulfuric acid [16]. Comparable corrosion rates of Ru-alloyed and
cused on finding a more affordable replacement for Pd in the Pd-alloyed Cr were obtained over a wide range of temperatures
weld consumable alloy. The main effect of Pd on the corrosion per- and in different concentrations of sulfuric and hydrochloric acids
formance of Ni–Cu–Pd was found to be its catalytic influence on [17,18]. The corrosion resistance of stainless steel in various acidic
the Cu redeposition reaction [4–6], so any replacement for Pd solutions was also improved by alloying with Ru. Higginson found
should also have good catalytic properties. that Fe–40Cr alloyed with 0.2% or 0.1% Ru spontaneously passiv-
ated in both 0.5 M sulfuric and hydrochloric acids while 0.2% Ru
addition led to faster passivation [19]. Similar to the situation with
Cr, Ru alloying of stainless steel can result in comparable corrosion
* Corresponding author. Tel.: +1 614 688 4128; fax: +1 614 292 9857. properties to that of Pd-alloyed steels [20]. Stainless steel alloyed
E-mail addresses: liangdosu@gmail.com (D. Liang), sowards.18@osu.edu (J.W.
Sowards), frankel.10@osu.edu (G.S. Frankel), alexandrov.1@osu.edu (B.T. Alexan-
with Ru actually exhibited somewhat better corrosion resistance
drov), lippold.1@osu.edu (J.C. Lippold). than that of stainless steel alloyed with same amount of Pd.

0010-938X/$ - see front matter Ó 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.corsci.2010.03.005
2440 D. Liang et al. / Corrosion Science 52 (2010) 2439–2451

The goal of this study was to investigate the substitution of Ru 2.3. Generation III welds
for Pd in Cr-free Ni–Cu based consumables for welding stainless
steel. The effects of a Ru addition on the corrosion behavior of Four generations of welds were investigated. The first two gen-
Ni–Cu alloys have not been reported in the literature. Considering erations of welds were fabricated with Ni–Cu–Pd consumables, as
the similar properties of Ru and Pd and the good corrosion resis- described previously [1]. Generations III and IV welds were fabri-
tance reported for Ru-alloyed stainless steel in various solutions, cated using Ru-containing consumables.
Ru appeared to be a potential replacement for Pd in the Ni–Cu– The consumables for Generation III welds were produced from a
Pd consumable. cast ingot with a target composition (wt.%) of Ni–7.5Cu–1Ru–0.5Al–
0.5Ti–0.02C. Al and Ti were alloyed to prevent weld metal porosity as
2. Materials and methods they are effective at removing oxygen and nitrogen from molten
weld pools [21]. The fabricated Ni–Cu–Ru ingot contained 8.2% Cu
A variety of tests was performed on button-melted samples, and 1.36% Ru; the actual composition of the ingot is listed in Table
bead-on-plate spot welds, and actual welded samples. For the lat- 1. The ingots were reduced by hot forging to 12.7 mm diameter
ter samples, the use of ‘‘weld” when referring to the exposed area and then drawn to 3.2 mm wire. Some of this wire was further drawn
means both the weld and base metal, whereas ‘‘weld metal” means to 1.1 mm for welding by GTAW. GTA welds were performed using
that only the weld metal was exposed. bare wire consumables. All GTA welding was performed on a Jetline
Sidebeam Carriage with automatic arc voltage control (set to 10%
2.1. Button-melted samples sensitivity) using a Miller Dynasty 300 LX power supply. A 2.4 mm
thoriated tungsten electrode was used at 200 A, 13.5 V, 2.1 mm/s
Buttons with total weight of 10 g were made by gas tungsten travel speed and 25.4 mm/s wire feed speed.
arc melting pure Ni, Cu, and Ru feed material under Ar and then Electrochemical tests were conducted on button samples that
air cooling. The following compositions were fabricated: Ni–5Cu, were made by GTA melting of the Generation III welding wire of
Ni–10Cu, Ni–10Cu–1Ru, and Ni–10Cu–0.5Ru (all compositions primary concentration Ni–8.2Cu–1.36Ru. Dilutions of 25% or 50%
are given in wt.%). Since Ru was in the form of powder and could were achieved by melting 304L stainless steel chips together with
be easily blown away by the arc, the powder was placed beneath pieces of the Ni–Cu–Ru wire, with multiple remelts to ensure
a Ni–Cu button that was then remelted twice to generate Ni–Cu– homogeneity.
Ru samples. Dilution was achieved by melting SS304L chips with Ni–Cu–Ru GTA welds on SS304L base metal were fabricated
the available Ni–Cu–Ru buttons. Button dilution was calculated with the wires mentioned above and machined for cyclic potentio-
with the following formula: dynamic polarization (CPP) tests. The dilution for Ni–Cu–Ru GTA
welds was about 30% and the weld to base metal area ratio for
%Dilution ¼ Mass304L =ðMass304L þ MassNi—Cu—Ru Þ  100 the CPP experiments was about 1:2 with the weld section in the
middle.
A wire was soldered to the back of the button samples for elec- The 3.2 mm wire was later used for production of SMAW elec-
trochemical measurements. Then the samples were mounted in trodes after the addition of flux coatings by Special Metals Corp.
epoxy and ground to a 600 grit finish. The samples were left in a The composition of the coating was not provided. Three-layer mul-
desiccator for 24 h before the start of an experiment. Before each tipass SMA welds were made on SS304L substrates with the SMAW
experiment, the edge between the sample and epoxy was sealed electrodes at 120–130 A, 24–25 V, and 2.5 mm/s travel speed. Sam-
by black wax. The electrochemical testing was performed about ples were machined from each layer in the weld deposit for corro-
1.5–2 h after applying the black wax to allow it to dry and harden sion testing. Some porosity was detected during the welding
to the proper extent. process but corrosion testing was conducted in regions determined
Homogenized buttons of Ni–10Cu–1Ru and Ni–10Cu–0.5Ru to be porosity-free. The dilution levels were 22%, 7%, and 4% for the
with 0% and 20% dilution by SS304L were also investigated. first, second, and third layers, respectively, as determined from:
Homogenization was accomplished by cold rolling the buttons to %Dilution ¼ XFe-weld =XFe-SS304L  100
75% reduction in thickness, then heating at 1050 °C for 1 h in an
evacuated glass tube, followed by water quenching. where XFe-weld and XFe-SS304L are the Fe compositions measured by
EDS of the weld layer and base metal, respectively. Fe composition
2.2. Bead-on-plate (BOP) welds was used in the dilution calculation because the welding electrode
contained no Fe. The samples were soldered to an electrical wire
To further investigate the possible replacement of Pd by Ru, and mounted in epoxy. The weld to base metal area ratio was main-
simulated welds were made using the bead-on-plate (BOP) tech- tained at 1:2.
nique. Slugs of Ni–10Cu with 0.5% or 1% Ru were placed into holes
drilled in an SS304L plate and melted using the gas tungsten arc 2.4. Generation IV welds
welding process. These were simple spot welds with the voltage,
current, and weld time controlled to achieve the desired dilution. The porosity problem associated with Generation III SMA welds
The dilutions of these welds were estimated to be approximately led to the development of the Generation IV SMAW consumables.
50% based on energy dispersive spectroscopy (EDS) measurements Another Ni–Cu–Ru ingot was cast with increased Ti content. The tar-
performed in a scanning electron microscope (SEM). The area ratio get composition of this new ingot was Ni–7.5Cu–1Ru–0.5Al–4Ti–
of weld to base metal was about 1:2 for all corrosion samples sub- 0.02C. The actual composition of the ingot is listed in Table 2; it
jected to cyclic potentiodynamic polarization testing. contained 7.78% Cu and 1.11% Ru. The ingots were reduced by hot

Table 1
Measured compositions of Ni–Cu–Ru ingot used for fabrication of Generation III welding wire.

C N O S P Si Cu Ni Al Ti Ru
0.014 <0.001 0.0031 <0.001 <0.005 – 8.20 89.3 0.56 0.53 1.36
D. Liang et al. / Corrosion Science 52 (2010) 2439–2451 2441

Table 2 defects. Specimens were tested at a strain rate of 3  107 s1 both
Measured compositions of Ni–Cu–Ru ingot used for fabrication of Generation IV in air and in 25 wt.% NaCl solution at pH 1.5. After experiments, the
welding consumable core wire.
fracture surface was examined by SEM. All SSRT experiments were
C N O S P Si Cu Ni Al Ti Ru performed at room temperature, approximately 23 °C.
0.019 – – <0.002 <0.002 0.10 7.78 85.85 0.83 4.31 1.11 Atmospheric exposure tests were also performed on welded
samples. Samples 5  5 cm in size and ground to a 600 grit finish
were put on the roof of MacQuigg Laboratory on the Ohio State
University campus in Columbus, OH for exposure to a typical Mid-
forging to 12.7 mm diameter and then drawn to 3.2 mm wire. The
west urban environment.
electrodes were coated by Electrode Engineering with flux of compo-
sition listed in Table 3. Welds were produced with the Generation IV
consumable on SS304L base metal using the SMAW process. Weld 4. Results and discussion
current was maintained in the range of 120–130 A, arc voltage be-
tween 24 and 25 V, and weld travel speed at approximately 4.1. Button samples
2.5 mm/s.
Ni–Cu–Ru button samples of all compositions exhibited a cellu-
lar dendritic grain structure as shown in Fig. 1. This microstructure
3. Experimental methods is similar to the dendritic structure found in actual weld deposits,
and likely represents similar microsegregation patterns. Therefore,
Cyclic polarization tests were conducted in aerated 0.1 M NaCl it is expected that the button samples behaved similar to actual
solution using a Gamry reference 600 potentiostat at a scan rate weld deposits from a corrosion standpoint.
of 10 mV/min. A saturated calomel electrode (SCE) and a pure plat- The open circuit potential, breakdown potential, and repassiva-
inum mesh were used as reference and counter electrode, respec- tion potential were evaluated by cyclic polarization tests in aerated
tively. Air was bubbled through the solution for aeration. Prior to 0.1 M NaCl solution. Fig. 2 shows the results for four as-cast alloy
initiating the scan, the sample open circuit potential (OCP) was al-
lowed to stabilize for a period of 1–2 h.
Uncreviced and creviced long time exposure tests were per-
formed in aerated 0.1 M NaCl solution. Creviced long time immer-
sion tests were performed using crevice-forming washers made
from Teflon and machined according to ASTM G78 [22]. Samples
of dimension 2.5  5 cm were machined from the welds and a
0.6 cm diameter hole was drilled at the fusion line of the weld
for attachment of the crevice-forming washer. In this way, some
of the washer feet were on the weld and some on the base metal
for each sample. Teflon tape was used to cover the washer feet sur-
faces. A torque of 80 N m was applied to the crevice washers using
a bolt. Uncreviced long time immersion tests in aerated 0.1 M NaCl
were performed on weld samples 2.5  5 cm in size. The samples
surfaces were also ground to a 600 grit finish. A Pt wire was spot
welded onto the samples to hang the sample in the solution and
to provide electrical contact for electrochemical measurements.
The Pt wire and connection spot were covered with red lacquer
to prevent galvanic corrosion. OCP and polarization resistance
(Rp) were measured daily. Fig. 1. Microstructure of Ni–10Cu–1Ru button sample. Micron marker = 100 lm.
Slow strain rate testing (SSRT) was employed to assess the sus-
ceptibility of the samples to stress corrosion cracking (SCC). Round
un-notched tensile specimens with 4 mm diameter were made
according to ASTM standard E8 [23]. For welded samples, the ten-
sile specimens were fabricated transverse to the weld with the
weld in the center of the gauge section. The surface of the tensile
bar was polished to 1 lm to minimize crack-initiating surface

Table 3
Flux composition of Generation IV SMAW consumables.

Component Flux mix


Calcium carbonate 8.6
Cryolite 17.1
Rutile concentrate 13.3
Nephelene syenite 6.3
Feldspar 7.7
Strontium carbonate 15.4
Bentonite clay 1.5
Sodium CMC 0.7
Magnesium metal 1.5
Sodium alginate 0.5
Ferro silicon 4.6
Binder #6 22.8 Fig. 2. Cyclic polarization curves for as-cast Ni–Cu buttons with varying Ru levels in
aerated 0.1 M NaCl solution.
2442 D. Liang et al. / Corrosion Science 52 (2010) 2439–2451

buttons. The Ni–5Cu and Ni–10Cu buttons behaved similarly, minimum Ru composition is recommended to be 1 wt.% in the
although the repassivation potential of the Ni–10Cu alloy was welding consumable. Although not shown, buttons with Ti added
much higher. This agrees with prior work that showed the addition to an alloy containing Ru exhibited no additional improvement in
of Cu improved the repassivation potential due to the redeposition corrosion properties.
of dissolved Cu ions [4–6]. Although not shown, the addition of 1% To evaluate the effects of microsegregation in the button sam-
W increased the passive current and decreased the breakdown po- ples on corrosion properties, cyclic polarization tests were also
tential, and samples with 1–5% Ag exhibited many pores on the performed on homogenized button samples. Generally, conven-
polished cross-section, probably because the solubility of Ag in tional arc welding results in the redistribution of alloying elements
Ni–Cu alloys is very low [24]. In contrast, the addition of Ru im- in the fusion zone of the weld. As a result, microsegregation occurs
proved the corrosion behavior of the Ni–Cu weld metal even at a across dendrite arms during weld solidification [21]. For instance,
level of 0.5 wt.%. The open circuit potential (OCP, defined as the as the result of microsegregation, the Cu content of a Ni–65Cu alloy
zero-current potential on the upward scan), breakdown potential was found to vary from 75% in the dendrite boundary to 53% in the
(Eb, defined as the potential at which the current increased rapidly) dendrite core [25]. For Ni–30Cu, the maximum difference of Cu
and repassivation potential (Erp, defined as the zero-current poten- composition between Cu-rich region and Cu-depleted region was
tial on the reverse scan) all increased with Ru addition. about 11% [4]. The microsegregation decreased as the Cu content
Fig. 3 shows the OCP (Fig. 3a), Eb (Fig. 3b), and Erp (Fig. 3c) val- in Ni–Cu alloys decreased. Microsegregation can degrade the cor-
ues for Ni–10Cu–1Ru and Ni–10Cu–0.5Ru as-cast buttons with dif- rosion properties of the weld metal as suggested by the higher
fering dilution by SS304L. Also shown in these figures for breakdown potential of homogenized Ni–Cu alloy button samples
comparison are data for Ni–10Cu–1Pd [4]. For most conditions, compared to as-cast buttons [7]. In another study [26], a Ni–Mo al-
the alloy with 0.5 wt.% Ru exhibited a similar breakdown potential loy was more susceptible to localized corrosion at the dendrite
and much higher repassivation potential than the alloy with 1 wt.% boundary, which had higher composition of Mo than that of den-
Pd [4]. Fig. 3 indicates that increasing the Ru content to 1 wt.% pro- drite core due to microsegregation of Mo.
vided additional improvement, especially at higher dilutions. The breakdown potentials of the homogenized samples in aer-
Therefore, to maintain the beneficial alloying effect of Ru, the ated 0.1 M NaCl solution were higher than for the as-cast buttons

Fig. 3. Electrochemical data for as-cast Ni–Cu–Ru buttons in aerated 0.1 M NaCl solution at different dilutions with SS304L. Data for Ni–10Cu–1Pd button were added for
comparison [4]. (a) Corrosion potential; (b) Breakdown potential; (c) Repassivation potential.
D. Liang et al. / Corrosion Science 52 (2010) 2439–2451 2443

by only 20–40 mV for both 0% and 20% dilution. The small Table 4
improvement following homogenization suggests that only a min- Average potentials for Ni–Cu–Ru bead-on-plate welds with 50% dilution. Potentials
given in mVSCE.
or degree of Cu microsegregation was present in the as-cast but-
tons. Although Cu-enriched regions in Ni–Cu alloys were found Alloy OCP Eb Erp
to increase susceptibility to localized corrosion [4,7], limiting the Ni–10Cu–0.5Ru 133.1 240.4 136.3
Cu content to 10 wt.% in the tested buttons effectively minimized Ni–10Cu–1Ru 119.44 285.62 151.9
corrosion issues induced by microsegregation. Therefore, the
breakdown potential exhibited only a small improvement after
homogenization.
Cyclic polarization tests were conducted on as-cast Ni–10Cu–
1Ru and Ni–10Cu–0.5Ru buttons with dilutions of 0%, 20%, or
50% in aerated solutions containing 105, 350, 1050, 3500, 10,500,
or 35,000 ppm Cl by weight. As shown in Fig. 4, corrosion poten-
tials were lower than both the breakdown and repassivation
potentials, indicating good resistance to localized corrosion even
in the high chloride concentration electrolytes. Eb and Erp de-
creased with increasing Cl concentration and the differences be-
tween Eb and Erp and between these potentials and OCP
decreased with increasing Cl concentration. This is in agreement
with the trend of increased susceptibility to localized corrosion
with increasing concentration of aggressive ions [27]. The corro-
sion potential showed little dependence on Cl concentration and
a weak dependence on the dilution. The corrosion potentials of
Ni–10Cu–1Ru and Ni–10Cu–0.5Ru were higher than that of
SS304L (150 mVSCE [4]), and the repassivation potentials of the
Ni–10Cu–1Ru and Ni–10Cu–0.5Ru were much higher than that of
SS304L (91 mVSCE [4]). As mentioned previously, a higher OCP
than the base metal means that the weld will be galvanically pro-
Fig. 5. Corrosion, breakdown and repassivation potentials for as-cast buttons as a
tected by the base metal. Furthermore, a better resistance to local-
function of dilution in aerated 0.1 M NaCl solution, Ni–10Cu–0.5Ru: , Ni–10Cu–
ized corrosion can be expected from the higher repassivation 1Ru: filled triangles. Also shown are results for BOP welds at 50% dilution, Ni–10Cu–
potential of the new alloy. Although the breakdown potential of 0.5Ru: open diamonds, Ni–10Cu–1Ru: open squares.
SS304L (310–500 mVSCE [4]) was still higher than that of Ni–
10Cu–1Ru and Ni–10Cu–0.5Ru at low chloride concentration, it
(150 mVSCE [4]), it approached that of Ni–10Cu–1Ru and Ni– that Ni–10Cu–1Ru exhibited higher OCP, Eb, and Erp than Ni–10Cu–
10Cu–0.5Ru at higher chloride concentration. 0.5Ru. As described above, increased benefit with increased Ru
content was also observed in the experiments on button samples.
At 50% dilution, the BOP welds with 0.5 or 1 wt.% Ru exhibited bet-
4.2. Bead-on-plate samples ter corrosion behavior than the 1 wt.% Pd addition. The catalytic
activity of Ru is similar to Pd [28], which has been shown to en-
The corrosion properties of BOP samples having weld to base hance Cu2+ redeposition at breakdown sites and thereby ennoble
metal area ratio of 1:2 were also assessed by cyclic polarization Eb and Erp relative to Ni–Cu alloy [4]. Pits were found exclusively
tests in aerated 0.1 M NaCl solution. The dilution of the BOP sam- on the weld metal for both of the Ni–Cu–Ru BOP samples, which
ples was approximately 50%. Results of the polarization tests of Ni– is similar to Ni–Cu–Pd SMA welds. The pits formed first on the
Cu–Ru BOP samples are summarized in Table 4 and Fig. 5. It is clear weld metal of the BOP samples during potentiodynamic scanning
because the breakdown potential of Ni–Cu–Ru and Ni–Cu–Pd alloy
is lower than that of SS304L.

4.3. Generation III: Ni–Cu–Ru GTA and SMA welds

Fig. 6 shows the microstructure of the Generation III weld


deposited by GTAW. The Generation III weld metal exhibited a cel-
lular dendritic solidification morphology with secondary phases
along solidification grain boundaries (SGB) and subgrain solidifica-
tion boundaries (SSGB) [29]. Epitaxial nucleation from the SS304L
base metal was also observed.
Cyclic polarization testing in aerated 0.1 M NaCl solution was
performed on button samples with different dilutions (0%, 25%,
and 50%) made by gas tungsten arc melting of Generation III weld-
ing wire of primary concentration Ni–8.2Cu–1.36Ru with SS304L,
Fig. 7. The experimental data are summarized in Table 5. The
breakdown potential tended to increase with increasing dilution,
which might be due to an increase in Cr content. The repassivation
potential correspondingly decreased with increasing dilution.
Overall as the dilution by SS304L increased, the weld contained
Fig. 4. Effects of Cl concentration, Ru content, and dilution on corrosion, more of the elements of stainless steel and behaved more like
breakdown, and repassivation potentials in aerated NaCl solutions. stainless steel. The OCP of the buttons with 0% dilution was much
2444 D. Liang et al. / Corrosion Science 52 (2010) 2439–2451

Fig. 6. Weld microstructure of Generation III weld deposits on SS304L. (a) Micron marker = 50 lm; (b) Micron marker = 10 lm; (c) Micron marker = 10 lm. MGB denotes
migrated grain boundary.

Table 5
Critical potentials measured from multiple potentiodynamic polarization experi-
ments on as-cast buttons of Ni–Cu–Ru with varying levels of dilution by SS304L.
Potentials given in mVSCE.

Dilution (%) OCP Eb Erp


0 61.7 200 120
69.4 188.9 111.1
70 188.3 110
25 133.3 186.1 108.3
126.7 218.3 135
128.3 230.7 131.7
50 123.3 291.7 96.7
113.9 330.5 86.1
118.3 320.1 101.7

SS304L. Fig. 8 shows a comparison of critical potentials measured


for three types of samples: buttons made from the Ru-containing
welding wire ( symbols), buttons made from pure Ni, Cu, and
Ru at a ratio of Ni–10Cu–1Ru (triangle symbols), and bead-on-plate
welds made by melting pure elements into a hole in SS304L. The
results are very similar, which indicates that the effects of the min-
or alloying elements (Ti, Al, and C) in the weld wire on corrosion
Fig. 7. Cyclic polarization curves for as-cast buttons made from Ni–Cu–Ru bare
properties of the weld are limited.
wire consumable with varying dilutions of SS304L in aerated 0.1 M NaCl solution.
Uncreviced long time immersion tests were also performed on
the Ni–Cu–Ru button samples fabricated from the Generation III
higher than the rest of the buttons, which is consistent with previ- consumable. The button samples were cold rolled to a reduction
ous results that showed Ni–Cu–Ru alloys exhibit higher OCP than of 75% to increase the exposed area. Then they were connected
D. Liang et al. / Corrosion Science 52 (2010) 2439–2451 2445

Fig. 8. Critical potentials for Ni–Cu–Ru buttons containing different dilutions of Fig. 10. Cyclic polarization curves for Ni–Cu–Ru GTAW weld metal, base metal, and
SS304L. Data with  symbols are for buttons fabricated from Ni–Cu–Ru welding the entire weld including base and weld metal regions in aerated 0.1 M NaCl
wire consumable and filled triangles are for buttons fabricated from pure elements. solution.
Data previously reported for bead-on-plate (BOP) welds is shown for comparison.

was lower than the other welds. The OCP of the Ni–Cu–Ru GTA
to a Pt wire by spot welding and suspended in aerated 0.1 M NaCl weld was similar to those of the other welds. After the experiment,
solution for approximately 700 h. The corrosion potential and pits were found on both the weld and base metal. The relatively
polarization resistance, Rp, (inversely proportional to corrosion low repassivation potential was probably associated with pits in
rate) were measured daily. The results are shown in Fig. 9. The the SS304L base metal. This is different than what was observed
trend of OCP for these buttons with different dilutions is similar for Ni–Cu–Pd SMA weld and Ni–10Cu–1Ru button, where pits
to the OCP values from the potentiodynamic experiments on the exclusively occurred on the weld surface, and is evidence of the
buttons fabricated from the Generation III wire in which the OCP resistance of the weld material.
decreased slightly with increasing dilution (Fig. 8). The 25% dilu- Cyclic polarization tests were also performed on base metal and
tion sample exhibited an increase in OCP over the first several days. Ni–Cu–Ru GTA weld metal separately in aerated 0.1 M NaCl solu-
The Rp values were all similar in magnitude, with a tendency for tion. This was done to investigate why pits formed randomly on
higher Rp at higher dilution. These results were also consistent both weld and base metal for the entire weld sample. Only base
with the potentiodynamic polarization test results. metal or weld metal was exposed by masking with black wax.
Cyclic polarization tests were conducted on Generation III Ni– The cyclic polarization curves of both base and weld are included
Cu–Ru GTA welds (with dilution of approximately 30%) in aerated in Fig. 10 and a summary of OCP, Eb and Erp values is provided in
0.1 M NaCl solution. The results are labeled as entire weld in Table 6. The OCP values for base and weld metals were very simi-
Fig. 10. The breakdown potential of the Ni–Cu–Ru weld was found lar, which means galvanic interaction is minimized in this case.
to be higher than that of Ni–Cu–Pd SMA weld and also higher than Furthermore, the breakdown potential of weld metal for this Ru-
that of Ni–10Cu–1Ru button. However, the repassivation potential containing GTA weld was much higher than observed for previous

Fig. 9. Immersion tests of Ni–Cu–Ru button samples at different dilutions. Samples were immersed in aerated 0.1 M NaCl solution. (a) OCP; (b) Rp.
2446 D. Liang et al. / Corrosion Science 52 (2010) 2439–2451

Table 6
Critical potentials measured from replicate potentiodynamic polarization experi-
ments on the base metal and weld metal zones of Ni–Cu–Ru GTA welds, as well as for
whole welds including both base and weld metal. Potentials given in mVSCE.

OCP Eb Erp
Base 132 410 35
135 380 23
140 360 25
Weld 120 340 95
140 300 73
128 362 82
131 310 80
Whole 129 316 33
140 323 22
127 326 10

conditions, including Ru-containing buttons and BOP welds. As a


result, the weld metal breakdown potential was not far below that
Fig. 12. Cyclic polarization curves for three layers of Ni–Cu–Ru SMAW welds in
of the SS304L. The breakdown potential for the entire weld, con- aerated 0.1 M NaCl solution.
taining both weld and base metal, was also about the same value.
The similarity in base and weld metal breakdown potentials for
this weld explains why pits initiated randomly on both base and Table 7
weld metal as the potential was increased. The repassivation po- Critical potentials measured from multiple potentiodynamic polarization experi-
ments on different layers of Ni–Cu–Ru SMA welds. Potentials given in mVSCE.
tential for the weld metal was higher than that of both the
SS304L base metal and the entire weld. It is likely that the low OCP Eb Erp
repassivation potential for the Ni–Cu–Ru weld (including base me- Layer 1 63 170 146
tal) was due to the pits on the base metal surface. Since pits ap- 110 180 133
peared on both weld metal and base metal regions, the potential Layer 2 185 158 68
138 168 71
had to be stepped down to the low repassivation potential to stop
Layer 3 150 151 60
the growth of pits in the base metal. 166 128 54
Creviced long time immersion tests were also performed on the
Ni–Cu–Ru GTA weld in aerated 1000 ppm NaCl solution for
approximately 500 h. After the experiment, only one crevice site
the Erp were all lower than the values found for Ni–Cu–Ru GTA welds.
was found, and the attack was on the base metal, as shown in
The pits were found only on the weld metal, reflecting the higher cor-
Fig. 11. This result is an improvement over what was observed
rosion susceptibility of SMA welds compared to GTA welds. SMA
for Ni–Cu–Pd welds in terms of total number of crevice sites and
welds likely contain more features that can act as pit initiation sites
also the position of crevices [1]. In general, the creviced long time
such as inclusions and secondary phases including titanium–
immersion results are consistent with cyclic polarization test re-
carbonitrides.
sults in which Ni–Cu–Ru welds exhibited better corrosion proper-
ties than Ni–Cu–Pd welds.
The corrosion properties of each layer of the Ni–Cu–Ru multipass
4.4. Generation IV: Ni–7.78Cu–1.11Ru–4.31Ti–0.83Al SMA welds
SMA welds were assessed by cyclic polarization tests in aerated
0.1 M NaCl solution. The dilutions of layers 1, 2, and 3 were 22%,
As was found for Generation III weld metal, Migrated Grain
7%, and 4%, respectively. The OCP and Eb increased with dilution level
Boundaries (MGB) were also observed in Generation IV welds,
as expected, Fig. 12. The results are summarized in Table 7. The
Fig. 13. There was a difference in grain boundary migration behav-
repassivation potential also increased with dilution which is
ior of Generations III and IV deposits as the result of differences in
contrary to expectations based on previous data. The OCP, Eb, and
the Ti and C contents. The Generation IV weld deposits contained
approximately 1 wt.% Ti after dilution, which was significantly
higher than the Generation III deposit Ti level [30]. The increased
concentration of TiC particles pinned the MGBs, resulting in the
more tortuous boundaries for the SMA weld in Fig. 13 compared
to the GTA weld in Fig. 6. Higher concentrations of Ti and Al were
added to the core wire of the Generation IV weld consumable in an
attempt to reduce SMA weld porosity. SMA welds were success-
fully fabricated with no porosity using this consumable.
Electrochemical tests were performed on Generation IV Ni–Cu–
Ru–Ti–Al SMA welds with approximately 30% dilution in aerated
0.1 M NaCl solution. A typical cyclic polarization curve is shown
in Fig. 14 and the results of multiple experiments are shown in Ta-
ble 8. The OCP, Eb, and Erp values are similar to those of Generation
III Ni–Cu–Ru GTA welds. Also, the OCP and Eb values are higher
than multipass Generation III SMA welds, though the Erp is lower.
As discussed for the Ni–Cu–Ru GTA welds, the lower Erp is likely
associated with pits in the base metal because pits were found in
Fig. 11. SEM image of the crevice on the base metal of Ni–Cu–Ru GTA weld. Micron both weld and base metal. This random distribution of pits in the
marker = 200 lm. weld and base metals is an indication of the improvement of the
D. Liang et al. / Corrosion Science 52 (2010) 2439–2451 2447

Fig. 13. Weld microstructure of Generation IV SMA weld deposits on SS304L at three different magnifications. MGB denotes migrated grain boundary.

Table 8
Critical potentials measured from replicate potentiodynamic polarization experi-
ments on Ni–Cu–Ru SMA welds. Potentials given in mVSCE.

OCP Eb Erp
127 200 22
167 276 46
160 203 77
146 253 44
140 186 47
122 221 26
158 233 23

Uncreviced long time immersion tests in aerated 0.1 M NaCl


solution were conducted on Generation IV Ni–Cu–Ru–Ti–Al SMA
welds. OCP and Rp were measured daily for approximately 500 h.
The polarization resistance is shown as a function of immersion
Fig. 14. Cyclic polarization curves for Generation IV Ni–Cu–Ru SMA weld in aerated
time in Fig. 15. Oscillations in Rp values were observed during
0.1 M NaCl solution.
the whole exposure, but Rp gradually increased with time and re-
mained higher than that at the start, reflecting a gradual passiv-
Generation IV Ni–Cu–Ru SMA welds compared to Generation III ation process.
Ni–Cu–Ru SMA welds, which exhibited pits only in the weld metal. Creviced long time immersion tests were also performed on
The improvement of the corrosion properties can also be related to these Generation IV welds in 1000 ppm Cl containing solutions
the enhancement of the weldability since micro-porosity can serve for 21 days. Only one very shallow crevice was observed on the
as defect sites that not only impair the strength [21,30], but also weld metal, which is consistent with the high corrosion resistance
decrease the corrosion properties [31,32]. of this weld.
2448 D. Liang et al. / Corrosion Science 52 (2010) 2439–2451

Fig. 15. Polarization resistance of immersed sample in aerated 0.1 M NaCl solution.

4.5. Atmospheric corrosion


Fig. 17. Slow strain rate testing curves for Generation III Ni–Cu–Ru GTA welds in air
and solution, 25 wt.% NaCl solution at pH 1.5. Strain rate was 3  107 s1.
Atmospheric exposure tests were performed on a Generation III
Ni–Cu–Ru GTA weld with a dilution of approximately 30%. The sam-
ple was placed on the roof of MacQuigg Laboratory in Columbus, OH These experiments were conducted at room temperature in air and
for exposure to a typical Midwest urban environment from August in aerated 25 wt.% NaCl solution at pH 1.5. The stress–strain curves
2007 to May 2008. The sample was checked regularly and photo- for both conditions are shown in Fig. 17. The ductility in air for Gen-
graphed. However, the reflective surface was difficult to capture by eration III Ni–Cu–Ru GTAW welds, about 20%, was much lower than
photography. The Ni–Cu–Ru GTA weld showed little sign of corro- the values observed for SS304L (around 120%) and Generation II Ni–
sion after 9 months. In contrast, Ni–Cu and Ni–Cu–Pd welds started Cu–Pd welded samples (around 95%) [1]. This might be associated
to show discoloration after only 6 months of exposure [1]. The atmo- with defects in the welds, such as porosity, which ultimately led to
spheric corrosion performance of Ni–Cu–Ru GTA welds are also sim- the development of the Generation IV consumables. Nonetheless,
ilar to or better than SS308L welds, which also started to show both the tensile strength and the strain were lower in solution than
discoloration after 9 months. Close inspection of samples in Fig. 16 in air, which is a sign of SCC. SEM images of the fracture surfaces are
indicates that a small amount of reaction product can be observed shown in Fig. 18. Fig. 18a exhibits a typical fracture surface for Ni–Cu
on one side of the Ni–Cu–Ru samples after 9 months exposure. The and Ni–Cu–Ru welds in air; the surface exhibits dimples indicative of
atmospheric corrosion exposure results are consistent with cyclic ductile fracture. Fig. 18b–d exhibit the Ni–Cu–Ru weld fracture sur-
polarization tests as Ni–Cu–Ru GTA welds exhibited better perfor- face of the sample in solution. Most of the fracture surface was
mance than Ni–Cu and Ni–Cu–Pd welds. through the SS304L base metal as indicated by EDS, and this part
of the fracture exhibited cleavage instead of dimples (top right of
4.6. SCC susceptibility evaluation Fig. 18b and c). The fracture in the base metal was near the weld,
likely the heat affected zone (HAZ). Fig. 18d shows the small part
Generation III Ni–Cu–Ru GTA welds and Generation IV Ni–Cu–Ru of fracture surface that went through the weld metal as indicated
SMA welds were subjected to SSRT at a strain rate of 3  107 s1. by EDS. In this region the fracture was ductile. Therefore, despite

Fig. 16. Optical images of atmospheric corrosion samples after exposure to typical Midwest climate. (a) SS308L GTA weld after 16 months; (b) Ni–Cu–Ru SMA weld after
9 months.
D. Liang et al. / Corrosion Science 52 (2010) 2439–2451 2449

Fig. 18. SEM pictures of Generation III Ni–Cu–Ru GTA welds in air and aerated pH 1.5 25 wt.% NaCl solution. (a) Sample in air. Micron marker = 20 lm; (b) Sample in solution.
(c) Base metal region of fracture for sample in solution (top right of image b). (d) Weld metal region of fracture for sample in solution in (bottom left of image b). Micron
marker = 50 lm.

the relatively low ductility of the Generation III welds, SSRT in a cor-
rosive environment resulted in fracture at the SS304L HAZ as a result
of an embrittlement of this region.
Fig. 19 shows the SSRT curves for Generation IV Ni–Cu–Ru SMA
welds in air and in solution. Higher stresses and better ductility
were achieved for these welds than for the Generation III GTA
welds, probably because of lower porosity. However, the ductilities
were not as large as observed for Generation II welds [1] for rea-
sons that are not understood, but might be related to differences
in the SS304L base metal. Similar to Ni–Cu–Ru GTA welds, the frac-
ture in air was ductile and occurred in the weld region because the
weld metal has lower tensile strength compared to the base metal.
Fig. 19 indicates a lower ductility for the sample tested in solution,
but the decrease is less than for the other weld samples indicating
a smaller reduction in properties as a result of SCC. EDS analysis
indicated that the sample in solution fractured totally in the base
metal unlike the Generation III Ni–Cu–Ru GTAW sample in solu-
tion, which contained a small area of weld metal on the fracture
surface. Fig. 20 shows the fracture surface morphologies after the
SSRT in solution. Fig. 20a and b indicate that both intergranular
and transgranular fracture was observed. As shown in the low
magnification image of Fig. 20c and the higher magnification image Fig. 19. Slow strain rate testing curves for Generation IV Ni–Cu–Ru SMA welds in
in Fig. 20d, a small region of dimples was also observed on the frac- air and aerated 25 wt.% NaCl solution at pH 1.5. Strain rate was 3  107 s1.

ture surface. The ductile region was probably just the overload
fracture after the load bearing section decreased.
Table 9 exhibits the mechanical properties of the samples from the smallest decrease in ductility in solution relative to air, indicat-
SSRT. The largest change in properties from air to solution was for ing a relatively high resistance to SCC. Furthermore, all the
the SS304L sample, indicating a large susceptibility of this alloy to mechanical properties decreased compared to high strain rate ten-
SCC in this aggressive solution. The Generation IV weld exhibited sile test results [29].
2450 D. Liang et al. / Corrosion Science 52 (2010) 2439–2451

Fig. 20. SEM pictures of different morphologies on SSRT fracture surfaces for Generation IV Ni–Cu–Ru SMA welds tested in aerated pH 1.5 25 wt.% NaCl solution. (a)
Intergranular fracture. (b) Transgranular fracture. (c) Lower magnification image. (d) Ductile region at lower right of image c. Micron marker = 20 lm.

Table 9
Comparison of mechanical properties of SS304L and welded samples. Some data are reproduced from Ref. [1].

Weld metal Base metal Environment Failure location 0.2% Proof stress, MPa Ultimate tensile strength, MPa Elongation, %
N/A 304L Air 304L 550 750 117
Solution 300 380 20
Generation II Ni–Cu–Pd SMAW Air Weld metal 500 650 90
Solution Weld metal 260 410 46
Generation III Ni–Cu–Ru GTAW Air Weld metal 257 369 20
Solution Base & weld 257 351 14
Generation IV Ni–Cu–Ru SMAW Air Weld metal 296 446 23.8
Solution Base metal 281 416 20.6

5. Conclusions repassivation potential of the weld metal was higher than that of
the base metal. Pits were found both on base metal and weld metal
The corrosion properties of a variety of samples fabricated from because of the similar breakdown potentials. Creviced immersion
Ni–Cu–Ru welding consumables were investigated. In general, the tests on Ni–Cu–Ru GTA welds resulted in crevice sites only on
results indicate that SMA welds on SS304L exhibited good corro- the base metal, indicating that these welds have better resistance
sion properties and that this consumable is suitable for replacing to crevice corrosion than SS304L.
SS308L for welding of SS304L. The nominal composition of the (3) Experiments on multipass SMA welds in SS304L fabricated
weld metal is Ni–7.5Cu–1Ru–1Ti–1Al–0.02C. Specifically, the fol- using the last generation of NiCuRu consumable indicated excel-
lowing observations were made: lent corrosion resistance.
(1) Experiments on button and bead-on-plate (BOP) samples (4) Good resistance of the welds made with Ni–Cu–Ru to atmo-
suggest that Ni–Cu–Ru alloys exhibit corrosion properties equiva- spheric corrosion and stress corrosion cracking was exhibited.
lent to or better than Ni–Cu–Pd alloys, which were developed pre-
viously [1]. Acknowledgements
(2) Ni–Cu–Ru GTA welds in SS304L exhibited excellent corro-
sion properties in various corrosion tests relative to welds made This project was supported by the Strategic Environmental Re-
with SS308L. Cyclic polarization results showed that the base me- search and Development Program (SERDP) through project WP
tal and weld metal had similar breakdown potential while the 1415 under the direction of Bruce Sartwell.
D. Liang et al. / Corrosion Science 52 (2010) 2439–2451 2451

References [18] N.D. Tomashov, G.P. Chernova, E.N. Ustinsky, Platin. Met. Rev. 23 (1979) 143–
149.
[19] A. Higginson, R.C. Newman, R.P.M. Procter, Corros. Sci. 29 (1989) 1293–1318.
[1] D. Liang, G.S. Frankel, J.W. Sowards, B.T. Alexandrov, J.C. Lippold, Mat. Corr.
[20] J. Potgieter, A. Heyns, W. Skinner, Rev. Appl. Electrochem. 20 (1990) 711–715.
(2010), doi:10.1002/maco.200905583.
[21] S. Kou, Welding Metallurgy, second ed., John Wiley & Sons, Inc., Hoboken, New
[2] S. Kou, Welding Metallurgy, Wiley-Interscience, Hoboken, NJ, 2003.
Jersey, 2002.
[3] J.J. McEwan, P.V.T. Scheers, D. Knight, Newly developed ruthenium-modified
[22] G78, Standard guide for crevice corrosion testing of iron-base and nickel-base
stainless gives good service in acid environments, in: 13th International
stainless alloys in seawater and other chloride-containing aqueous
Corrosion Congress, 1996.
environments, in: Annual Book of ASTM Standards, ASTM, Philadelphia, PA,
[4] Y.H. Kim, Chromium free Consumable for Welding Stainless Steel, Ph.D. Thesis,
2001.
The Ohio State University, 2005.
[23] E8, Standard test methods for tension testing of metallic materials, in: Annual
[5] Y.H. Kim, G.S. Frankel, J. Electrochem. Soc. 154 (2007) 36–42.
Book of ASTM Standards, ASTM, Philadelphia, PA, 2004.
[6] Y.H. Kim, G.S. Frankel, J.C. Lippold, Effect of noble element alloying on passivity
[24] X.J. Liu, F. Gao, C.P. Wang, K. Ishida, J. Electronic Mat. 37 (2008) 210–217.
and passivity breakdown of Ni, in Passivity 9, Elsevier, Paris, 2005.
[25] W.F. Savage, E.F. Nippes, T.W. Miller, Welding J. (1976) 181–187.
[7] Y.H. Kim, G.S. Frankel, J.C. Lippold, G. Guaytima, Corrosion 62 (2006) 44–53.
[26] M.J. Perricone, J.N. DuPont, Laser welding of Ni–Mo alloys, in: Joining of
[8] I. Wolff, P. Park-Ross, A. Ball, Brit. Corr. J. 26 (1991) 202–208.
Advanced and Specialty Materials, ASM International, Indianapolis, IN, 2002.
[9] J. Potgieter, P. Ellis, A. Bennekom, ISIJ Int. 35 (1995) 197–202.
[27] D.A. Jones, Principles and Prevention of Corrosion, second ed., Prentice Hall,
[10] I.M. Wolff, L.E. Iorio, T. Rumpf, P.V.T. Scheers, J.H. Potgieter, Mat. Sci. Eng. A
New York, 1996.
A241 (1998) 264–276.
[28] G. Webb, Platin. Met. Rev. 8 (1964) 60–66.
[11] G. Myburg, K. Varga, W.O. Barnard, P. Baradlai, L. Tomcsanyi, J.H. Potgieter,
[29] J. Sowards, Development of a Chromium-free Consumable for Joining Stainless
C.W. Louw, M.J. van Staden, Appl. Surf. Sci. 136 (1998) 29–35.
Steels, Ph.D., The Ohio State University, 2009.
[12] E. Van der Lingen, R.F. Sandenbergh, Corr. Sci. 43 (2001) 577–590.
[30] G. Tani, A. Ascari, G. Campana, A. Fortunato, Appl. Surf. Sci. 254 (2007) 904–
[13] M. Stern, H. Wissenberg, J. Electrochem. Soc. 106 (1959) 759–764.
907.
[14] M. Stern, H. Wissenberg, J. Electrochem. Soc. 106 (1959) 755–759.
[31] M. Zhong, W. Liu, H. Zhang, Wear 260 (2005) 1349–1355.
[15] N. Greene, C. Bishop, M. Stern, J. Electrochem. Soc. 108 (1961) 836–841.
[32] A. Wang, X. Chang, W. Hou, J. Wang, Acta Metall. Sinica 42 (2006) 537–539.
[16] N.D. Tomashov, G.P. Chernova, E.N. Ustinsky, Corrosion 40 (1984) 134–138.
[17] J. Potgieter, Rev. Appl. Electrochem. 21 (1991) 471–482.

S-ar putea să vă placă și