Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Corrosion Engineering, Science and Technology

The International Journal of Corrosion Processes and Corrosion Control

ISSN: 1478-422X (Print) 1743-2782 (Online) Journal homepage: https://www.tandfonline.com/loi/ycst20

The galvanic effect of high-strength weathering


steel welded joints and its influence on corrosion
resistance

Chen Huang, Feng Huang, Hai-Xia Liu, Qian Hu & Jing Liu

To cite this article: Chen Huang, Feng Huang, Hai-Xia Liu, Qian Hu & Jing Liu (2019):
The galvanic effect of high-strength weathering steel welded joints and its influence
on corrosion resistance, Corrosion Engineering, Science and Technology, DOI:
10.1080/1478422X.2019.1636484

To link to this article: https://doi.org/10.1080/1478422X.2019.1636484

Published online: 01 Jul 2019.

Submit your article to this journal

Article views: 17

View Crossmark data

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=ycst20
CORROSION ENGINEERING, SCIENCE AND TECHNOLOGY
https://doi.org/10.1080/1478422X.2019.1636484

The galvanic effect of high-strength weathering steel welded joints and its
influence on corrosion resistance
Chen Huang, Feng Huang, Hai-Xia Liu, Qian Hu and Jing Liu
State Key Laboratory of Refractories and Metallurgy, Wuhan University of Science and Technology, Wuhan, People’s Republic of China

ABSTRACT ARTICLE HISTORY


The influence of the galvanic effects on the welded joints of A710 high-strength weathering steel with Received 14 April 2019
two different weld metals (80Ni1-H4 and WER70) under a simulated marine atmospheric environment Accepted 23 June 2019
was investigated by employing Scanning Kelvin Probe (SKP), galvanic current measurements, cyclic
KEYWORDS
wet/dry accelerated corrosion testing, Field-Emission Scanning Electron Microscopy (FE-SEM) and A710 steel; welded joints;
other approaches. The results indicate that the average galvanic currents for the HAZ-BM and HAZ- galvanic corrosion; SKP
WM couples of the two welded joints initially decreased and then gradually stabilised, which is
related to the formation of a stable rust layer. In addition, the WER70 welded joints presented a
higher average corrosion rate than for the 80Ni1-H4 welded joints. The significant galvanic effect of
the WER70 welded joints and the formation of a compact rust layer are the dominant effects during
the initial and later corrosion stages, respectively.

Introduction
the most common type of failure mode for welded joints
Welding processes are a very important and inevitable [13,14]. Therefore, there has been increasing attention
aspect of the application of weathering steels. Owing to given to these joints, especially considering their influen-
the influence of welding thermal cycle and different welding tial factors.
materials, the structures and chemical compositions of Zhang et al. [15] discovered that the galvanic corrosion
weathering steel welded joints are extremely inhomo- rates for different coupling pairs were positively correlated
geneous. This could significantly deteriorate the structural with the potential difference between the two different
integrity of the joints and result in probable changes to metals. Arya et al. [16] also noted that the corrosion cur-
the corrosion resistance [1–3]. In recent years, an increasing rent increased with the increasing cathode/anode area
number of researchers have focused their attention on the ratio, and, for a given ratio, the greater the separation dis-
corrosion behaviour of weathering steel welded joints. tance between the anode and cathode, the smaller the cur-
Yue-Lin et al. [3] found that the welded joint of a low car- rent density. Varela et al. [17] considered that the
bon Cu–P–Cr weathering steel prepared using electric slag corrosion current for carbon steel/stainless steel coupling
pressure welding exhibited a lower corrosion resistance was not positively correlated with temperature, especially
than the base steel. Xie et al. [4] studied the corrosion when around 100°C. However, limited studies have been
behaviour of SMA490BW weathering steel welded joints done with respect to the galvanic corrosion kinetics of
and noted that automatic Metal Active Gas (MAG) arc weathering steel welded joints and its effect on corrosion
welded joints showed higher corrosion resistances compared resistance, which have a direct impact on the service safety
with manual MAG welded joints. Gong et al. [5] also of weathering steel in marine environments. In addition,
confirmed that the welded joints from 09CuPCrNi weather- there is limited research on the galvanic corrosion kinetics
ing steel made using shielded metal arc welding with a of weathering steel made with different welding processes
J502NiCrCu welding wire had a lower corrosion rate than and filler wires.
the base metal at later stages. Therefore, welded joints do Here, the high-strength weathering steel denoted as A710
not necessarily have deteriorated corrosion resistances, steel was chosen as the base metal due to its excellent mech-
and a reasonable welding process and welding materials anical properties and weldability, and because it is expected
can even lead to improvements. to be widely applied in marine engineering. The main objec-
Most studies have shown that the inhomogeneity of the tive of this work is to investigate the galvanic corrosion kin-
microstructure and chemical composition of welded joints etics of A710 steel with different filler wires and welding
are the primary reasons for a deteriorated corrosion resist- processes. The influences on the corrosion resistance for
ance [6–8], which may cause localised corrosion, such as these welded joints were investigated by employing Scan-
galvanic corrosion, stress corrosion and others [9–12]. ning Kelvin Probe (SKP), dry/wet cycling galvanic current
The characteristics of localised corrosion include a measurements, electrochemical technologies, Field-Emission
high local corrosion rate, small structural loss rate and Scanning Electron Microscopy (FE-SEM) and others. This
strong concealment, which can lead to the early cata- work provides a theoretical basis and data accumulation
strophic failure of engineered structures and generate for the application of A710 steel in marine atmospheric
huge economic losses. Statistically, galvanic corrosion is environments.

CONTACT Feng Huang huangfeng@wust.edu.cn State Key Laboratory of Refractories and Metallurgy, Wuhan University of Science and Technology, Wuhan,
Hubei 430081, People’s Republic of China
© 2019 Institute of Materials, Minerals and Mining Published by Taylor & Francis on behalf of the Institute
2 C. HUANG ET AL.

Experimental procedures Table 2. The welding parameters.


Welded joint D (mm) I (A) U(V) V (cm min−1) HI (J mm−1)
Tested material and welding procedure 80Ni1-H4 φ1.6 240–260 28–30 23 2500
WER70 φ1.2 250–270 28–30 29 1500
The A710 high-strength weathering steel employed in this
research was supplied by the Wuhan Iron and Steel Cor-
poration, and two different filler wires (80Ni1-H4 and
WER70) were used for the welded joints and were created Table 3. The mechanical properties of the A710 weathering steel and its welded
using a gas shielded arc process. The chemical compo- joints.
sitions of the material and the welding parameters are R (MPa) Rm (MPa) A (%) Z (%)
shown in Tables 1 and 2, respectively, while Table 3 sum- A710 540 640 24 77
marises the mechanical properties of the base metal and 80Ni1-H4 / 668 / 80
WER70 / 654 / 78
the welded joints.
Based on the ASTM A710/A710M-2013 standard, the
yield and tensile strengths of the A710 steel meet the specified with distilled water and alcohol and then subjected to cyclic
standards (Table 3). During tensile testing, the fracture sites wet/dry accelerated corrosion testing. The cyclic period was
of the two welded joints were located in the base metal 60 min including immersion into a 3.5 wt-% NaCl solution
(A710 steel). The tensile strengths of the two welded joints at 40 ± 1°C for 12 min and drying at 42 ± 1°C with 60 ± 2%
were both higher than the lower yield strength (Rel) of the Relative Humidity for 48 min. The test procedure consisted
base metal, and the section shrinkage (Z ) was also larger of: (i) weighing the initial samples (60 × 40 × 5 mm); (ii)
than the base metal (Table 3). This shows that the tensile removing the samples after corroding for 1, 2, 3, 6, 12
test results for the two welded joints are qualified. In addition, and 24 days; (iii) washing the samples with deionised
due to the differences in the used filler wires, the welded joints water to prevent the accumulation of progressive salt; (iv)
are separately denoted as the 80Ni1-H4 and WER70 welded removing the rust from the samples in a solution of
joints. It is noted that all samples cut from the welded joints 500 mL hydrochloric acid + 500 mL distilled water + 3.5 g
included the base metal and the weld metal with an area ratio hexamine, where the blank sample was used to correct for
of approximately 4:1. defects and (v) re-weighing the samples after drying. The
corrosion rates of the samples were determined by consider-
ing the total weight loss.
Microstructure analysis
The two welded joint specimens for the A710 steel were wet
ground with silicon carbide papers up to a 1200 grit, and then Morphology observations and composition analysis of
polishing in 6 and 1 μm diamond suspensions. After that, the the rust layers
specimens were etched with 4% nital (4 mL nitric acid + The cross-section morphologies of the rust layers (corrosion
96 mL ethyl alcohol). The surfaces were then examined products) at different corrosion times were observed in the
under an optical microscope. locations of the WM by using an FE-SEM (Nova400 Nano
SEM). The corrosion products taken from the coupon surface
at different corrosion times were ground into fine powders
Volta potential distribution measurements
and were all equalised to have the same mass. Then X-ray
The Volta potential distributions for different regions of the diffraction (XRD; Shimadzu XRD-6000 diffractometer with
two A710 weathering steel welded joints were measured Co Kα radiation) was used for the phase composition analysis
using a VersaSCAN electrochemical workstation with the of the rust layers.
SKP technique. A tungsten wire probe with a diameter of
100 μm was invoked as the needle, whose amplitude and fre-
quency were 30 μm and 1 kHz, respectively. The test area Morphology observations for the welded joints after
included the base metal (BM), the heat affected zone (HAZ) removing corrosion products
and the weld metal (WM), and was 18 000 × 4000 μm with A schematic of the sample setup used for the measurements
a step size of 100 μm both in the X- and Y-directions, as illus- to quantitatively determine the depth differences in the two
trated in Figure 1. The speed was set to 100 μm s−1, and the welded joints before and after the corrosion testing is shown
average distance between the tip and the coupon was approxi- in Figure 2. The red zone on the left side of the sample is
mately 100 μm during the testing. the un-corroded area, which was covered with a water-
resistant silica gel to prevent corrosion during the immer-
sion stage, and the right side is the corroded area. The
Cyclic wet/dry accelerated corrosion testing
un-corroded area was the reference plane used for the cor-
After being sequentially ground by 180, 240, 400, 600, and rosion zone. After 8 days of alternating in the wet–dry mar-
800 grit silicon carbide abrasive papers, the two different ine environment, a topographic analysis was performed
welded joints using the A710 weathering steel were rinsed using a VHX-5000 stereo-microscope for the samples.

Table 1. Chemical composition of the materials used (wt-%).


C Si Mn P S Ni Cr Cu Mo Nb
A710 0.045 0.29 0.63 0.009 0.005 1.23 0.78 1.01 0.19 0.025
80Ni1-H4 0.005 0.32 1.20 0.014 0.009 0.93 0.15 0.35 / /
WER70 0.06 0.49 1.81 0.01 0.005 0.62 0.26 0.26 / /
CORROSION ENGINEERING, SCIENCE AND TECHNOLOGY 3

Figure 1. The test area for the A710 weathering steel welded joints: (a) WER70
welded joint and (b) 80Ni1-H4 welded joint.

Open Circuit Potential measurements


The open circuit potential (OCP) measurements for different
regions (BM, HAZ and WM) of two welded joints were con-
ducted using an Auto Lab PGSTAT204 electrochemical
workstation with platinum as the counter electrode, a KCl-
saturated calomel electrode (SCE) as the reference electrode,
the different regions (BM, HAZ and WM) as the working Figure 3. Schematic diagram used for the galvanic current measurement and of
electrode and a 3.5 wt-% NaCl solution as the electrolyte. the specimen (a: HAZ-WM couple; b: HAZ-BM couple).
The measurement time for the OCP was 5400 s.
Results

Galvanic current measurements Microstructure of the two welded joints

During the cyclic wet/dry accelerated corrosion testing, the Optical macrographs of the WER70 and the 80Ni1-H4 welded
galvanic current for the HAZ in the HAZ-BM and HAZ- joints are shown in Figure 4. The BM of these welded joints
WM couples for the two A710 steel welded joints was con- (Figure 4a) is primarily made up of relatively large ferrite
stantly measured using a CST500 Galvanic corrosion/elec- and granular bainite with good uniformity. The WM of the
trochemical noise measuring meter. The HAZ of the WER70 welded joint (Figure 4c) mainly consists of ferrite wid-
welded joints was used as the working electrode I, the BM mannstatten and a lesser amount of bainite. For the 80Ni1-H4
or WM was used as the working electrode II, and the welded joint, the WM mainly consists of acicular and relatively
KCl-SCE worked as the reference electrode. If the galvanic large ferrite with small amounts of bainite (Figure 4e). The
current measured by the CST500 is positive, the working similarities in both the welded joints in the WM consist of
electrode I is anodic; if not, the working electrode II is the inhomogeneous structure and the large grains.
cathodic. A schematic of the setup used to measure the gal- The HAZ in the WER70 welded joint (Figure 4b) is mainly
vanic currents for the specimens is illustrated in Figure 3. It composed of relatively large ferrite and small amounts of
is noted that the area ratios of BM, HAZ and WM were the granular bainite, which shows different grain sizes. However,
same as those for the cyclic wet/dry accelerated corrosion the HAZ of the 80Ni1-H4 welded joint (Figure 4d) mainly
tests, and the measurement time was 8 days, which included consists of granular bainite and small amounts of large ferrite
192 cycles. with different grain sizes. This means that the microstructures
have obvious variations for the different regions of the two
welded joints.

Potential distribution of the two welded joints


The distributions of the Volta potential for the two welded
joints measured via Kelvin probe are shown in Figures 5
and 6. From Figure 5(a), we see that the potential for differ-
ent regions of the WER70 welded joint progressively
decreased in the order of BM, HAZ and WM. The potential
differences with a distance of Y = 2.4 mm between the HAZ
and BM were approximately 160 mV for the HAZ and
100 mV for the MW, as shown in Figure 5(b). However,
Figure 2. Schematic diagram of the sample to determine the depth differences the ranked potentials for the different regions of the
before and after the corrosion testing. 80Ni1-H4 welded joint from largest to smallest are the
4 C. HUANG ET AL.

Figure 4. Microstructure morphologies: WER70 welded joint for (a) BM, (b) HAZ and (c) WM; and the 80Ni1-H4 welded joint for (d) HAZ and (e) WM.

HAZ, BM and WM (Figure 6a). Furthermore, the potential corrosion time, the OCP values tend to gradually stabilise.
differences between the HAZ and BM, and the HAZ and Furthermore, after stabilisation, the OCPs of the 80Ni1-H4
WM for the 80Ni1-H4 welded joint were approximately welded joints decrease in the order BM, HAZ and WM,
20 and 80 mV, respectively. The potential varied with dis- whereas the OCPs for the WER70 welded joints reduce as
tance at Y = 2.4 mm (Figure 6b) and was much lower HAZ, BM and WM. These results are consistent with the
than that for the WER70 welded joint. According to the SKP results from the two welded joints (Figures 5 and 6),
potential differences of the three regions, the WER70 indicating that there could be differences in the corrosion ten-
welded joint meets the thermodynamic conditions for galva- dency between different areas of the two welded joints.
nic corrosion to occur.
As is well known, there is a positive correlation between
the natural corrosion potential and the Volta potential, as The galvanic current for the two welded joints
seen in Equation (1) [18]. For the same metal, a smaller
To intuitively and quantitatively analyse the galvanic cor-
Volta potential leads to a lower natural corrosion potential.
rosion behaviour of the two A710 weathering steel welded
Ecorr = Dw + A (1) joints, the galvanic currents for the HAZ-BM and HAZ-
WM couples were measured, as shown in Figure 8. For the
where Ecorr is the natural corrosion potential (mV), Δϕ is the WER70 welded joint, the galvanic current of the HAZ-BM
Volta potential (mV) and A is a constant. couple has a positive value (Figure 8a), while the HAZ-WM
To accurately analyse the variations in the natural cor- couple is negative (Figure 8b). This indicates that the HAZ
rosion potential, the OCPs for different regions (BM, HAZ is an anode against the BM, while the WM plays the role of
and WM) of the two welded joints in a 3.5 wt-% NaCl sol- the anode against the HAZ. For the 80Ni1-H4 welded joint,
ution were measured, as shown in Figure 7. It is clearly the galvanic currents for the HAZ-BM and HAZ-BM couples
seen that in the initial immersion stage, the OCPs for the are both negative (Figure 8c and d), indicating that the HAZ
different regions shift negatively, and with the continued acts as the anode to both the BM and WM.
CORROSION ENGINEERING, SCIENCE AND TECHNOLOGY 5

Figure 5. The SKP results (mV) for the WER70 welded joint: (a) the area scan for the potential difference and (b) variations in the potential with distance at Y =
2.4 mm.

It is noted that every cycle includes the immersion and stability of rust layers, which is discussed in the following
drying stages, but the galvanic currents measured in the dry- sections.
ing stage are unstable. Therefore, the average galvanic current
values for the couples in the immersion stage as a function of
Average corrosion rates for the two welded joints
the number of wet/dry cycles are exhibited in Figure 9. It is
observed that the average change in the galvanic currents Figure 10 shows the average corrosion rates for the two
for the HAZ-BM and HAZ-WM couples of the two welded different A710 weathering steel welded joints as functions
joints with the number of wet/dry cycles can be divided of the wet/dry cycle time. Over the entire corrosion process,
into two stages: a declining stage and a stabilizing stage. the corrosion rate of the WER70 welded joint is always
Further observations suggest there is a transition point corre- much larger than that of the 80Ni1-H4 welded joint at
sponding to approximately 6 days between the first stage and the same corrosion time. The average corrosion rate versus
the second stage. After the transition point, the galvanic cur- the corrosion testing time for both joints presents the same
rents for both welded joints became lower and remained variation trend as decreasing with longer corrosion times.
nearly stable. This could be related to the formation and Furthermore, it can be seen that the entire corrosion
6 C. HUANG ET AL.

Figure 6. The SKP results (mv) for the 80Ni1-H4 welded joint: (a) the area scan for the potential difference and (b) variations in the potential with distance at Y =
2.4 mm.

process can be divided into two stages: a faster downward a compact inner rust layer was formed, but some cracks par-
trend for the first stage and a slower downward trend for allel to the surface appeared on the WER70 welded joint. In
the second stage. The transition points for these stages are particular, at the 24th day, the WER70 welded joint still
at approximately 6 days. According to the literature [19– had some small cracks connecting the outer rust layer and
21], the higher corrosion rate stage is related to the for- the substrate, whereas there were no tiny cracks observed
mation process of the corrosion products, and the lower on the 80Ni1-H4 welded joint.
corrosion rate is correspondingly attributed to the for- Usually, the compact inner rust layer can improve the cor-
mation of a stable rust layer. rosion resistance by hindering the invasion of corrosive ions
to the substrate surface. Thus, after 6 days of the wet/dry cor-
rosion, the barrier properties of the rust layers were
Morphologies of the rust layers
improved, which contributed to the slowing trend in the
The cross-sectional structure of the rust layers formed on the second stage, as shown in Figure 11. Furthermore, it was
two welded joints after the different corrosion cycles is shown observed that the rust layer of the 80Ni1-H4 welded joint
in Figure 13. On first and third days, the rust layers on these was more compact than that for the WER70 welded joint
two samples are loose and porous. On the 6th and 12th days, given the same cycle time.
CORROSION ENGINEERING, SCIENCE AND TECHNOLOGY 7

the more stable the rust layer is and the better its barrier
effect.
The values of α and α/γ* are shown in Table 4 to quanti-
tatively compare the relative amounts of these phases in the
rust layers. With the increase of the corrosion time, the rela-
tive amounts of α-FeOOH and the α/γ* values increase for
both welded joints. Furthermore, the values of α and α/γ*
for the 80Ni1-H4 welded joint are much larger than those
of the WER70 welded joint, especially during the second cor-
rosion stage.

Morphologies of two welded joints after removing the


rust layers
Figure 7. The OCPs for the different regions as functions of time.
Figure 13 provides 3D topographic images that show the cor-
rosion morphologies of the two welded joints with galvanic
interactions as well as the depth variation with distance for
Composition of the powdered rust
Y = 500 μm. It can be seen that the depth differences between
The composition of the rust layers formed on the two the un-corroded and corroded regions for different areas in
welded joints was analysed using XRD, as shown in Figure the WER70 welded joint exhibit a larger Δdmax in the WM
12. The rust layers for both welded joints are quite similar at 76 μm than that of the 80Ni1-H4 welded joint at a Δdmax
after 3, 12 and 24 days, and are mainly composed of α- of 39 μm. Furthermore, for the corroded regions, the depth
FeOOH, γ-FeOOH and Fe3O4. According to Ref. [18], in the 80Ni1-H4 welded joint was much shallower than for
the ratio of α to γ* is an important index of the rust the WER70 welded joint. This indicates that there was a sig-
layer protection performance, where α represents the con- nificant galvanic effect between the different areas taking
tent of α-FeOOH and the γ* is the total content of the γ- place during the wet/dry cyclic immersion of the WER70
FeOOH and Fe3O4 or Fe2O3. The greater the α/γ* value is, welded joint.

Figure 8. The galvanic currents for the HAZ-BM and HAZ-WM couples in the WER70 welded joint (a, b) and the 80Ni1-H4 welded joint (c, d) as functions of the
number of wet/dry cycles.
8 C. HUANG ET AL.

preferential local corrosion with high electrochemical activi-


ties [23]. Numerous studies have shown that the galvanic cor-
rosion effects are dominated by the potential difference [15],
cathode–anode area ratio [16], environmental medium [17]
and others.
As a result, the same area ratios were applied in the cyclic
wet/dry accelerated corrosion testing processing for the
WER70 and 80Ni1-H4 welding joints. Different degrees of
galvanic corrosion occurred for the two welded joints (Figure
13), where ‘large cathode and small anode’ structures (HAZ
and WM as the anode, BM as the cathode) were formed in
the WER70 welded joint, whereas ‘small cathode and large
anode’ structures (BM and WM as the anode, HAZ as the
cathode) were formed in the 80Ni1-H4 welded joint (Figures
8 and 9). The microstructures for the local regions (HAZ,
WM and BM) of the welded joints (Figure 4) caused by the
different filler wires and welding processes also resulted in
their different electrochemical activities (Figure 7). The larger
the potential differences between the different regions
(Figures 5 and 6) were responsible for a more serious galvanic
corrosion of the WER70 welded joint (Figure 13).
Further analysis found that the evolution process for the
average galvanic currents in the HAZ for the HAZ-BM and
HAZ-WM couples of the two welded joints as a function of
the number of cycles can be divided into two stages: initial
decline and gradual stabilisation (Figures 8 and 9). The tran-
sition point was at approximately 6 days, which is related to
the formation of steady corrosion products on the welded
joints, as also observed in the literature [24]. The galvanic
effect between different regions in the two welded joints was
Figure 9. The average galvanic currents for the HAZ-BM and HAZ-WM couples continuously inhibited due to the coverage of corrosion pro-
in the WER70 welded joint (a) and 80Ni1-H4 welded joint (b) as functions of the ducts on the substrate [24], which were responsible for
number of wet/dry cycles.
decreases in the galvanic currents in the HAZ-BM and
HAZ-WM couples at the initial corrosion stage (Figure 8).
Discussion During the later corrosion stage, the rust layer became increas-
ingly compact with the extended corrosion time, which signifi-
Behaviours of the galvanic corrosion kinetics for the cantly improved the shielding effect of the rust layer on the
two welded joints
steel substrate. Therefore, the dense rust layer greatly wea-
In a simplified sense, galvanic corrosion refers to the local kened the galvanic effect of the welded joints and led to a wea-
corrosion of different metals with different corrosion poten- kened and gradually stabilised galvanic current (Figure 9) [25].
tials in the same corrosive medium after contact or connec-
tion [22]. However, broadly speaking, a microgalvanic
The influence of galvanic effect of the welded joints
corrosion cell can also be formed due to local differences in
on its corrosion resistance
the electrochemical activity of the same metal, leading to
The galvanic effect between the different areas of the welded
joints affected its corrosion resistance [26,27]. In this work,
the average corrosion rate of the WER70 welded joint was sig-
nificantly higher than that of the 80Ni1-H4 welded joint
given the same corrosion time. Changes in the corrosion
rate were also divided into two stages of decline and stabilis-
ation, where the turning point was at approximately 6 days
(Figure 10). This is almost identical to the observed galvanic
current change law. This similarity might be related to the
galvanic effect and the formation of a compact rust layer on
the two welded joints.
In the initial corrosion stage (0–6 days), a discontinuous,
thin and loose corrosion product was formed on the surface
of both welded joints, which provided a weak shielding
effect for the substrate (Figure 11). The aggressive ions
(such as Cl–) in the solution could still directly penetrate to
the surface of the substrate through the rust layer. However,
Figure 10. The average corrosion rates for the two welded joints as functions of due to the partial coverage of the corrosion product, the active
the corrosion time. area of the electrode was reduced, causing the average
CORROSION ENGINEERING, SCIENCE AND TECHNOLOGY 9

Figure 11. Cross-sectional morphological evolution of the rust layers formed on the A710 weathering steel welded joints.

Table 4. The values of α and α/γ* after 3, 12 and 24 days.


WER70 welded joint 80Ni1-H4 welded joint
α-FeOOH α/γ* α-FeOOH α/γ*
3d 4.3 0.045 10 0.111
12d 16.5 0.198 34.5 0.527
24d 22.7 0.294 52.1 1.088

WER70 welded joint was greater than that of the 80Ni1-H4


welding joint.
In the later corrosion stage (6–24 days), as the corrosion
time progressed, the α-FeOOH content in the rust layer of
the two welded joints increased (Table 4), and a dense
inner rust layer formed on the substrate (Figure 11). This
led to a significant improvement in the shielding effect of
the rust layer. In addition, the formation of the dense rust
layer caused the galvanic effect of the two welded joints to
greatly weaken and gradually vanish [25]. Therefore, the
influence of the galvanic effect on the corrosion resistance
of the welded joints could be virtually neglected. Thus the
influence of the rust layer shielding effect on the corrosion
resistance of the welded joints was dominant. However, the
thickness and compactness of the rust layer in the 80Ni1-
H4 welded joint were obviously superior to those of the
WER70 welded joint (Figure 11). Therefore, the corrosion
rate of the 80Ni1-H4 welded joint in the later stages was
lower than that of the WER70 welded joint (Figure 10).
And as mentioned in previous paragraph, differences in the
microstructures and chemical composition resulted in the
‘small cathode and large anode’ structures of the 80Ni1-H4
welded joint, which are more conducive to practical engineer-
ing applications.
Figure 12. XRD patterns for the corrosion products formed on the two steel
welded joints at different wet/dry cycle times: (a) WER70 and (b) 80Ni1-H4
welded joints.
Conclusion

corrosion rate to decrease with the prolonged corrosion time. (1) Differences in the microstructures and chemical compo-
At the same time, the galvanic effect between different regions sition led to varying degrees of the galvanic effect
in the two welded joints decreased gradually with the cor- between the different regions of two A710 steel welded
rosion time (Figures 8 and 9), which could also result in a joints. There was a more serious galvanic effect between
decrease of the corrosion rate of the welded joints (Figure the different regions during the wet/dry cyclic immersion
10). However, due to the weak shielding performance of the for the WER70 welded joint compared with the 80Ni1-
rust layer during the initial stage, the influence of the galvanic H4 welded joint.
effect on the overall corrosion resistance of the two welded (2) Owing to the initial galvanic effect and the inhibition of
joints was dominant, and the resulting corrosion rate of the the denser rust layer on the galvanic interactions,
10 C. HUANG ET AL.

Figure 13. Depth profile at the surface of the A710 weathering steel welded joints after 8 days of wet/dry alternating in a marine environment: WER70 welded joint
for (a) BM, (b) HAZ and (c) WM; and 80Ni1-H4 welded joint for (d) BM, (e) HAZ and (f) WM.

variations in the average galvanic current for the


HAZ-BM and HAZ-WM couples in the two welded Disclosure statement
joints as functions of the number of wet/dry cycles No potential conflict of interest was reported by the authors.
were divided into two stages: initial decline and gradual
stabilisation.
(3) The corrosion resistance of the 80Ni1-H4 welded joint Funding
was superior to that of the WER70 welded joint. This
This work was supported by National Key R&D Program of China (No:
is mainly due to the serious galvanic corrosion effect in 2017YFB0304801) and the National Natural Science Foundation of
the WER70 welded joint at the initial corrosion stage China (No: 51871172).
and the more compact rust layer of the 80Ni1-H4 welded
joint in the middle and later corrosion stages. In addition,
the ‘small cathode and large anode’ structures of the References
80Ni1-H4 welded joint are more conducive to practical [1] Zhou F, Cheng R, Xingxiang W U, et al. Research on welding pro-
engineering applications. cess and microstructure and properties of welded joint of
CORROSION ENGINEERING, SCIENCE AND TECHNOLOGY 11

09CuPCrNi-A and 316L dissimilar steel. J Jiangsu Univ Sci [15] Zhang R, Chenliang R, Dapeng L I, et al. Galvanic corrosion
Technol. 2016;30(2):131–135. behavior of dissimilar metal coupling in highly acidic environ-
[2] Wang S, Ma Q, Li Y. Characterization of microstructure, mechan- ments. Corros Prot. 2017;38(9):671–688.
ical properties and corrosion resistance of dissimilar welded joint [16] Arya C, Vassie P RW. Influence of cathode-to-anode area ratio
between 2205 duplex stainless steel and 16MnR. Mater Des. and separation distance on galvanic corrosion currents of steel
2011;32(2):831–837. in concrete containing chlorides. Cem Concr Res. 1995;25
[3] Chao YL, Zhou YL, Wang QL, et al. Corrosion performance of (5):989–998.
welded joint and base steel of Cu–P–Cr weathering steel by cyclic [17] Varela FE, Kurata Y, Sanada N. The influence of temperature on
immersion in Na HSO3 solution. Trans Mater Heat Treat. 2015;36 the galvanic corrosion of a cast iron-stainless steel couple (predic-
(2):84–90. tion by boundary element method). Corros Sci. 1997;39(4):775–
[4] Xu X, Xiang-Yang W U, Zhi-Yi Z, et al. Effects of welding pro- 788.
cesses on corrosion behavior of SMA490BW weathering steel [18] Jin D. Study on microstructure properties and corrosion resistance
joints. Surf Technol. 2018;47(10):256–261. mechanism of weathering bridge steel. Northeastern University;
[5] Gong L, Xing Q, Wang H. Corrosion behaviors of weathering steel 2014.
09CuPCrNi welded joints in simulated marine atmospheric [19] Ahmadi M, Mirsalehi SE. Investigation on microstructure, mech-
environment. Anti-Corros Methods Mater. 2016;63(4):295–300. anical properties and corrosion behavior of AISI 316L stainless
[6] Liu P, Han S, Yi Y, et al. Corrosion behavior of welded joint of steel to ASTM A335-P11 low alloy steel dissimilar welding joints.
Q690 with CMT twin. Int J Corros. 2018;2018:1–9. Mater High Temp. 2015;32(6):627–635.
[7] Ahmadi M, Mirsalehi SE. Investigation on microstructure, mech- [20] Zhang X, Yang S, Zhang W, et al. Influence of outer rust layers on
anical properties and corrosion behavior of AISI 316L stainless corrosion of carbon steel and weathering steel during wet–dry
steel to ASTM A335-P11 low alloy steel dissimilar welding joints. cycles. Corros Sci. 2014;82(5):165–172.
Mater High Temp. 2015;32(6):627–635. [21] Aso T, Utsumi T. The corrosion behavior of weathering steel
[8] Wang S, Ma Q, Li Y. Characterization of microstructure, mechan- under different corrosive environments. Australasia and South-
ical properties and corrosion resistance of dissimilar welded joint East Asia Structural Engineering and Construction Conference;
between 2205 duplex stainless steel and 16MnR. Mater Des. 2014.
2011;32(2):831–837. [22] Zhang W-y. Progress in research on galvanic corrosion behavior
[9] Zhu J, Xu L, Feng Z, et al. Galvanic corrosion of a welded joint in and protection. Total Corros Control. 2018;32(12). 51–56+122.
3Cr low alloy pipeline steel. Corros Sci. 2016;111:391–403. [23] Wei J, Dong J, Ke W, et al. Influence of inclusions on early cor-
[10] Srinivasan B, Dietzel W, Zettler R, et al. Stress corrosion cracking rosion development of ultra-low carbon Bainitic steel in NaCl sol-
susceptibility of friction stir welded AA7075-AA6056 dissimilar ution. Corrosion. 2015;71(12):1467–1480.
joint. Mater Sci Eng A. 2005;392(1):292–300. [24] Sun W. Numerical simulation of the influences of deposited cor-
[11] Blasco-Tamarit E, Igual-Muñoz A, Antón J G. Effect of tempera- rosion products on localized corrosion behavior. Dalian
ture on the galvanic corrosion of a high alloyed austenitic stainless Universal of Technology; 2013.
steel in its welded and non-welded condition in LiBr solutions. [25] Wen S, Liu G, Wang L, et al. A mathematical model for modeling
Corros Sci. 2007;49(12):4472–4490. the formation of calcareous deposits on cathodically protected
[12] Wan H, Du C, Liu Z, et al. The effect of hydrogen on stress cor- steel in seawater. Electrochim Acta. 2012;78(9):597–608.
rosion behavior of X65 steel welded joint in simulated deep sea [26] Wang YP, Zuo XR, Li JL. Corrosion resistance of the welded joint
environment. Ocean Eng. 2016;114:216–223. of submarine pipeline steel with ferrite plus bainite dual-phase
[13] Baboian R. Corrosion, Galvanic. Encyclopedia of Materials: microstructure. Steel Res Int. 2015;86(11):1260–1270.
Science and Technology. 2001: 1689–1690. [27] Guilherme LH, Rovere CAD, Kuri SE, et al. Corrosion behavi-
[14] Fang N, Chen G, Zhu H, et al. Statistical analysis of leakage acci- our of a dissimilar joint TIG weld between austenitic AISI
dents of submarine pipeline. Oil Gas Storage Transp. 2014;33 316L and ferritic AISI 444 stainless steels. Weld Int. 2016;19
(1):99–103. (1):42–50.

You might also like