You are on page 1of 8

Progress in Natural Science: Materials International 30 (2020) 110–117

HOSTED BY Contents lists available at ScienceDirect

Progress in Natural Science: Materials International


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

Original Research

Microstructure and properties of the interlayer heat-affected zone in X80 T


pipeline girth welds
Fang Baia,d, Hongsheng Dinga, Lige Tongb,⁎, Liqing Panc
a
School of Mathematics and Physics, University of Science and Technology Beijing, Beijing, 100083, China
b
School of Energy and Environmental Engineering, University of Science and Technology Beijing, Beijing, 100083, China
c
College of Science, China Three Gorges University, Yichang, 443002, China
d
Pipeline Institute, China Petroleum Pipeline Engineering Co., Ltd., Langfang, 065000, China

ARTICLE INFO ABSTRACT

Keywords: The heat-affected zone between the layers of a weld metal is a sensitive and important component of girth welds
Girth weld in X80 pipelines. However, only a few studies have focused on the interlayer heat-affected zone (IHAZ). In this
Interlayer heat-affected zone work, thermal simulation was conducted to investigate the relationship between IHAZ microstructure and
Thermal simulation properties. Results showed that the toughness of the weld metal improved after a single welding thermal cycle
Granular bainite
but decreased after double welding thermal cycles, particularly when the weld metal was subjected to unaltered
Toughness
coarse grains of IHAZ. The relationship between the microstructure and properties of IHAZ was established to
present the correlation of toughness with granular bainite content, grain size, and high-angle grain boundary
content.

1. Introduction granular bainite (GB) and quasi-polygonal ferrite (QF), and grain size
may have a considerable effect on weld toughness [18,19].
Welding is crucially important for pipeline construction. The The microstructure of HAZ between base and weld metals differs
heating process makes the girth weld the most sensitive and weakest fromthe IHAZ. However, few studies have been conducted on the IHAZ
part of pipelines [1,2]. Multipass welding is commonly performed in of weld metals. The microstructural inhomogeneity of IHAZ is con-
various pipeline constructions [3]. In this method, each welded joint sidered the primary reason for performance variation [20,21]. The
comprises two or more welding layers, and the interlayer heat-affected original microstructure of IHAZ exhibits a columnar morphology and
zone (IHAZ) is formed between the welding layers. The low-tempera- poor toughness [22,23]. By contrast, the microstructure of HAZ be-
ture toughness of the welded joint is related to the microstructure of tween base and weld metals comprises acicular ferrite (AF), which
IHAZ [4–6]. Therefore, the relationship between the microstructure and demonstrates excellent toughness [24]. Wang et al. [25] reported that
properties of IHAZ should be studied to identify the key factors for the welding reheating process changes prior austenite morphology from
improving girth weld toughness. columnar to equiaxed structure. In addition, a brittle zone, called weld
However, research on IHAZ remains limited. Most studies have fo- metal subjected to intercritical coarse grains (WM–IC CG), is formed
cused on the heat-affected zone (HAZ) between base and weld metals. during the welding reheating process. Moreover, “necklace” M–A con-
Several studies have shown that the impact performance of HAZ is stituents distributed along austenite grain boundaries induce stress
discretized due to the welding thermal cycling on HAZ [7], and the concentration and reduce toughness. Wang et al. [26] showed that lath
coarse-grained HAZ (CGHAZ) is the weakest region in the girth weld ferrite or AF, bainite/martensite, and fine acicular-retained austenite
because of the large size of the prior austenite and the large amounts of are beneficial for the weld metal toughness of IHAZ. There is, however,
martensite–austenite (M–A) constituents in this zone [8,9]. Other stu- no systematic research carried out for the microstructure and properties
dies have reported that a local brittle zone forms when CGHAZ is re- of IHAZ, and the key factors that influence IHAZ toughness remain
heated to Ac1 and Ac3 [10], and the M–A constituents comprise the unclear.
major contributing factor to this brittle zone [11–15]. Moreover, a The microstructure and properties of IHAZ after single and double
decrease in prior austenite grain size and M–A constituent amount welding thermal cycles were investigated in this study through thermal
improves Charpy impact energy [16,17], whereas the phases of simulation. The influences of different phase fractions, grain sizes, high-


Corresponding author.
E-mail addresses: bf851218@163.com (F. Bai), tonglige@me.ustb.edu.cn (L. Tong), lpan@ctgu.edu.cn (L. Pan).

https://doi.org/10.1016/j.pnsc.2019.08.010
Received 20 February 2019; Received in revised form 13 August 2019; Accepted 14 August 2019
Available online 27 February 2020
1002-0071/ © 2020 Chinese Materials Research Society. Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/BY-NC-ND/4.0/).
F. Bai, et al. Progress in Natural Science: Materials International 30 (2020) 110–117

Table 1 thermal simulator to obtain different microstructures. Previous studies


Mechanical properties of the X80 pipeline steel. [27,28] have indicated that the reheated microstructure of weld metal
Material Yield strength Tensile Elongation (%) Impact energy subjected to coarse grains (WM–CG) always exhibits the poorest
(MPa) strength (MPa) −20 °C (J) toughness among the microstructure of weld metals. Hence, the heating
peak temperature (Tp,1) for a single thermal cycle was set to 1350 °C to
X80 555 715 27 130
obtain WM–CG, whereas the second heating peak temperature for the
double thermal cycles (Tp,2) was set to 1100 °C, 900 °C, and 750 °C.
Fig. 2 shows the typical thermal cycle curves used in this study [26].
angle grain boundaries, and M–A constituents on IHAZ toughness were
The microstructure and impact toughness of the thermal simulation
analyzed. Furthermore, the relationship between the microstructure
samples were tested. The Charpy impact test was conducted at −20 °C.
and properties of IHAZ was established.
The microstructures were characterized via optical microscopy (Zeiss
Upright Light Microscopes), scanning electron microscopy (JEOL-
2. Experiment 7000F), electron backscattered diffraction (EBSD), and transmission
electron microscopy (TEM; JEM-2200FS). The distribution and fraction
2.1. Materials and welding of M–A constituents were visualized by treating with LePera etchant
and analyzed using Image-Pro Plus 6.0 software. Microstructure grain
An X80 pipeline steel plate with a thickness of 18.4 mm was used in size was also analyzed with Image-Pro Plus 6.0.
this study. Table 1 provides the mechanical properties of the X80 pi-
peline steel.
The semiautomatic welding of a self-shielded flux-cored wire was 3. Results
adopted as the welding method in the experiment. The welding mate-
rials included a ϕ3.2 mm LB-52 welding rod for root welding and a 3.1. Impact toughness
ϕ2.0 mm self-shielded flux-cored JC-30 wire for filling and cap pass.
These materials have been commonly used in pipeline construction. Table 2 presents the impact toughness of different thermally simu-
Fig. 1(a) shows the welded joint during the thermal simulation test. The lated specimens. The original weld metal with columnar microstructure
thickness of the last welding pass was set to approximately 5 mm to has an average toughness value of 46.5 J. The WM–CG samples present
obtain the original microstructure of the weld metal with sufficient high toughness with an average value of 83.5 J. The toughness of the
thickness and uniform properties. Fig. 1(b) shows a thermal simulation weld metal decreased after double thermal cycles. The IHAZ obtained
specimen with dimensions of 5 mm × 10.5 mm × 71 mm that was when Tp,2 = 1100 °C was defined as weld metal subjected to unaltered
sampled from the welded joint. coarse grains (WM–UA CG) and had an average toughness of only 7.5 J.
The IHAZ obtained when Tp,2 = 900 °C was defined as weld metal
2.2. Experimental method subjected to supercritical coarse grains (WM–SC CG) and had an
average toughness of 45 J. The IHAZ obtained when Tp,2 = 750 °C was
Specific thermal cycles were performed by using a Gleeble-3500 defined as WM–IC CG and had an average toughness of 32.5 J. These

Fig. 1. Diagram of a sample for thermal simulation.

Fig. 2. Single and double thermal cycles.

111
F. Bai, et al. Progress in Natural Science: Materials International 30 (2020) 110–117

Table 2 grain boundaries and coarse grains. The WM–CG obtained after a single
Comparison of the impact toughness of weld metal samples after different thermal cycle was coarse-grained. Fig. 4(b) shows that the micro-
thermal cycles. structure of WM–CG was dominated by GB, bainite ferrite (BF), and
Thermal cycle Test temperature Sample Impact energy (J) M–A constituents. Different microstructures were obtained after double
numbers (°C) thermal cycles at different Tp,2 values. Fig. 4(c) indicates that the mi-
Single Average crostructure of WM–UA CG remained coarse-grained. However, the
value value
average grain size of the microstructure of WM–UA CG was slightly
None −20 Original 56.5 46.5 smaller than that of the microstructure of WM–CG. The microstructure
31.5 was dominated by BF and GF with small amounts of QF and M–A
51.5 constituents. Fig. 4(d) shows that the microstructure of WM–SC CG was
Single WM-CG 77.5 83.5
characterized by non-uniform fine grains and mostly composed of GB,
76.5
97.0 BF, QF, polygonal ferrite (PF), and blocky M–A constituents. Fig. 4(e)
Double WM-UA CG 6.5 7.5 indicates that the microstructure of WM–IC CG was dominated by BF
9.0 and GB. The microstructure of WM–IC CG had a smaller average grain
7.0 size and thinner and denser laths than that of WM–CG. The M–A con-
WM-SC CG 59.5 45
stituents of WM–IC CG were distributed discontinuously along the grain
39.5
36.0 boundaries, and several M–A constituents outlined the grain bound-
WM-IC CG 31.0 32.5 aries.
36.5
30.0
3.3. EBSD

The grain boundary misorientations of the thermally simulated


results show that the toughness of the weld metal improved after a
samples were analyzed through EBSD, as presented in Fig. 5.
single thermal cycle but decreased after double thermal cycles, parti-
Fig. 5(a)–5(d) show the Euler maps of the thermally simulated samples.
cularly when Tp,2 = 1100 °C.
The high-angle grain boundaries are marked in yellow and black in the
Fig. 3 shows the fracture surfaces of the samples. The fracture sur-
pictures, whereas the low-angle grain boundaries are marked in red.
face of the sample WM–CG (Fig. 3(b)) displays typical ductile fracture
Fig. 5(e) provides a comparison of the grain boundaries in the simulated
characteristics with many dimples. The other fracture surfaces ex-
samples at different angles. In this study, high-angle grain boundaries
hibited brittle cleavage fracture manner with small areas of dimples.
were defined as grain boundary angles greater than 15°, whereas grain
boundary angles within the range of 5–15° were defined as low-angle
3.2. Microstructure grain boundaries [29]. The WM–SC CG and WM–UA CG samples had
the highest (28.8%) and lowest (9.7%) contents of high-angle grain
Fig. 4(a) shows that the microstructure of the original weld metal boundaries, respectively. The WM–CG sample had the highest content
was columnar with AF and GB. This microstructure demonstrated clear of low-angle grain boundaries (17.1%), and the WM–UA CG sample had

Fig. 3. Impact fracture surfaces of the simulated samples: (a) original, (b) WM–CG, (c) WM–UA CG, (d) WM–SC CG, and (e) WM–IC CG.

112
F. Bai, et al. Progress in Natural Science: Materials International 30 (2020) 110–117

Fig. 4. Optical micrographs of the simulated samples: (a) original, (b) WM–CG, (c) WM–UA CG, (d) WM–SC CG, and (e) WM–IC CG.

Fig. 5. EBSD analysis of grain boundary misorientations: (a)–(d) Euler maps of WM–CG, WM–UA CG, WM–SC CG, and WM–IC CG, respectively; and (e) comparison
of grain boundaries at different angles in the simulated samples.

113
F. Bai, et al. Progress in Natural Science: Materials International 30 (2020) 110–117

Table 3
Average grain size of different simulated samples.
Samples Microstructure grain Original WM–CG WM–UA CG WM–SC CG WM–IC CG

Average grain diameter (μm) 106 75 49 32 51


Optical picture

the highest content of grains with boundary angles between 0° and 5° observed features of M–A constituents in the weld metal. TEM ob-
(80.6%). servation showed that several M–A constituents were in blocky and
sharp-angled shape (Fig. 6(a)), whereas others were in the form of a
3.4. Grain size stringer film (Fig. 6(b)). M–A constituents were typically in the hard
brittle phase and had a high carbon content; stress concentration can
Table 3 provides the average grain size of each simulated sample. form around M–A constituents in sharp-angled forms [26].
The original sample had an average grain diameter of approximately Fig. 7(a) presents the M–A constituent contents of the different si-
106 μm due to its coarse columnar microstructure, whereas the average mulated samples. The WM–CG and WM–UA CG samples had the highest
grain diameter of the WM–CG sample was 75 μm. The microstructure of (2.8%) and lowest (1.86%) M–A constituent contents, respectively.
the WM–UA CG sample remained coarse-grained, with an average grain Morphology analysis was performed to further understand the effect of
diameter of 49 μm. The average grain diameters of WM–SC CG and M–A constituents. M–A constituents with a length longer than 2 μm
WM–IC CG were 32 μm and 51 μm, respectively. WM–SC CG had the were defined as large, and those with a length-to-width ratio lower than
lowest average grain diameter among all the samples. 3 were considered as blocky. Fig. 7(b) shows the stringer and blocky
M–A constituent contents of the different samples. The WM–IC CG
3.5. M–A constituents sample exhibited the highest average stringer M–A constituent content
(20.9%), and the WM–SC CG sample had the highest average blocky
The size and morphology of the M–A constituents of the samples M–A constituent content (70.9%). Fig. 7(c) provides a comparison of
subjected to different thermal cycles were analyzed. Fig. 6 presents the the large and large blocky M–A constituent contents of the different

Fig. 6. Morphology of M–A constituents: (a) blocky and sharp-angled and (b) stringer film.

Fig. 7. Statistics for the M–A constituents of the simulated samples: contents of (a) M–A, (b) M–A with different morphologies, and (c) large and large blocky M–A.

114
F. Bai, et al. Progress in Natural Science: Materials International 30 (2020) 110–117

Fig. 8. Relationships between IHAZ microstructure and toughness: (a) GB content, (b) average grain diameter, (c) high-angle grain boundaries, (d) M–A constituents,
and (e) large blocky M–A constituents.

samples. The WM–CG sample presented the highest average large M–A was affected when the WM–CG sample was reheated to temperature
constituent content (12.1%), whereas the WM–SC CG sample had the above Ac3, and increments in GB lath width reduced the impact
highest content of large blocky M–A constituents (7.7%). toughness of the samples.
Fig. 8(b) illustrates the relationship between grain size and impact
4. Analysis and discussion toughness. With the exception of the WM–CG sample, the toughness of
the WM–UA CG, WM–SC CG, and WM–IC CG samples was inversely
Fig. 8 shows the relationships between IHAZ microstructure and proportional to microstructure grain size.
toughness. IHAZ microstructure is mostly composed of AF, GB, BF, M–A Fig. 8(c) shows the relationship between high-angle grain boundary
constituents, and small amounts of PF and QF. The presence of large contents and impact toughness. With the exception of the WM–CG
blocky GB reduces toughness, whereas lath BF improves toughness. sample, the toughness of the WM–UA CG, WM–SC CG, and WM–IC CG
Fig. 8(a) presents the relationship between the GB content and impact samples was proportional to the contents of high-angle grain bound-
toughness of the simulated samples. The WM–UA CG sample exhibited aries. The WM–UA CG sample exhibited low toughness due to its low
the worst low-temperature toughness and the highest GB content of content of high-angle grain boundaries. The low high-angle grain
75%. In the other samples, GB content was inversely related to tough- boundary content of the WM–UA CG sample may be attributed to the
ness. The toughness decreases as the GB content increases. The GB presence of a large amount of GB, which has no high-angle grain
morphology of the samples was also observed via TEM. The WM–UA CG boundary. The most effective barrier for crack propagation was high-
sample with the worst toughness had the highest GB content and a GB angle misorientation boundaries, which were advantageous for tough-
lath width of approximately 1.1 μm (Fig. 9(a)). The GB lath width of the ness [30].
WM–CG sample was approximately 0.55 μm (Fig. 9(b)). GB lath width Fig. 8(d)–(e) presents the relationship between M–A constituents

Fig. 9. Morphologies of GB: (a) WM–CG and (b) WM–UA CG

115
F. Bai, et al. Progress in Natural Science: Materials International 30 (2020) 110–117

and impact toughness. The contents of the M–A constituents exerted a low-transformation temperature weld, Mater. Des. 160 (2018) 384–394 https://doi.
complex effect on toughness. Fig. 8(d) shows that toughness was en- org/10.1016/j.matdes.2018.09.016.
[6] Junyan Ni, Xiaowei Wang, Jianming Gong, Magd Abdel Wahab, Thermal, me-
hanced by the high total contents of M–A constituents. However, tallurgical and mechanical analysis of circumferentially multi-pass welded P92 steel
toughness and the content of large blocky M–A constituents were un- pipes, Int. J. Press. Vessel. Pip. 165 (2018) 164–175 https://doi.org/10.1016/j.
related primarily because the effect of M–A constituents depends on ijpvp.2018.06.009.
[7] Lewei Tong, Lichao Niu, Shuang Jing, Liwen Ai, Xiao-Ling Zhao, Low temperature
their shape and distribution. The shape of M–A constituents was re- impact toughness of high strength structural steel, Thin-Walled Struct. 132 (2018)
markably affected by the crystallography of surrounding bainite [31]. 410–420 https://doi.org/10.1016/j.tws.2018.09.009.
M–A constituents do not decrease microstructure toughness but im- [8] Xueda Li, Yuran Fan, Xiaoping Ma, S.V. Subramania, Chengjia Shang, Influence of
Martensite–Austenite constituents formed at different intercritical temperatures on
prove toughness, particularly stringer M–A constituents that enhance toughness, Mater. Des. 67 (2015) 457–463 https://doi.org/10.1016/j.matdes.2014.
the toughness of CGHAZ [32–34]. However, Luo et al. [35] suggested 10.028.
that stringer M–A constituents are more harmful to toughness compared [9] You Yang, Chengjia Shang, Liang Chen, Sundaresa Subramanian, Investigation on
the crystallography of the transformation products of reverted austenite in inter-
with blocky M–A constituents. Studies showed that a decrease in crack
critically reheated coarse grained heat affected zone, Mater. Des. 43 (2013)
initiation absorbed energy is mostly attributed to coarse M–A con- 485–491 https://doi.org/10.1016/j.matdes.2012.07.015.
stituents [36–38]. M–A constituents cause the brittle behavior of hard [10] Xueda Li, Chengjia Shang, Xiaoping Ma, S.V. Subramanian, R.D.K. Misra,
zones [39]. Lan et al. [40] reported that stress concentration is present Jianbo Sun, Structure and crystallography of martensite–austenite constituent in
the intercritically reheated coarse-grained heat affected zone of a high strength
around M–A constituents and is primarily responsible for the reduction pipeline steel, Mater. Char. 138 (2018) 107–112 https://doi.org/10.1016/j.
in toughness. Accordingly, the effect of M–A constituents on toughness matchar.2018.01.042.
remains debatable. In the current study, the effect of the M–A con- [11] S. Moeinifar, A.H. Kokabi, H.R. MadaahHosseini, Influence of peak temperature
during simulation and real thermal cycles on microstructure and fracture properties
stituents on toughness are not analyzed because clear evidence for this of the reheated zones, Mater. Des. 31 (2010) 2948–2955 https://doi.org/10.1016/j.
issue remains unavailable. matdes.2009.12.023.
The preceding discussion shows that girth weld toughness is related [12] J.R. Yang, H.K.D.H. Bhadeshia, Continuous heating transformation of bainite to
austenite, Mater. Sci. Eng. A 131 (1991) 99–113 https://doi.org/10.1016/0921-
to many factors. The reciprocal of the square root of grain size was also 5093(91)90349-R.
used in this study to consider the expression of material strength and [13] F. Maresca, V.G. Kouznetsova, M.G.D. Geers, W.A. Curtin, Contribution of auste-
grain size in the Hall–Pech relation. The relationship of IHAZ toughness nite-martensite transformation to deformability of advanced high strength steels:
from atomistic mechanisms to microstructural response, Acta Mater. 156 (2018)
with the reciprocal of the square root of grain size, GB content, and 463–478 https://doi.org/10.1016/j.actamat.2018.06.028.
large-angle grain boundary content can be expressed as [14] Xueda Li, Xiaoping Ma, S.V. Subramanian, Chengjia Shang, R.D.K. Misra, Influence
1 of prior austenite grain size on martensite–austenite constituent and toughness in
IE = 313.89 860.81 × d 2 2.79 × CGB + 1.95 × CHA. (1) the heat affected zone of 700MPa high strength linepipe steel, Mater. Sci. Eng. A
616 (2014) 141–147 https://doi.org/10.1016/j.msea.2014.07.100.
where IE denotes the impact energy (J), d refers to the grain diameter [15] S. Moeinifar, A.H. Kokabi, H.R. MadaahHosseini, Role of tandem submerged arc
(μm), CGB is the GB content (%), and. CHA denotes the content of high- welding thermal cycles on properties of the heat affected zone in X80 microalloyed
pipe line steel, J. Mater. Process. Technol. 211 (2011) 368–375 https://doi.org/10.
angle grain boundaries (%). 1016/j.jmatprotec.2010.10.011.
Eq. (1) shows that IHAZ toughness is proportional to the reciprocal [16] Sadegh Moeinifar, Amir Hossein Kokabi, Hamid Reza Madaah Hosseini, Effect of
of the square root of grain size, whereas the content of large-angle grain tandem submerged arc welding process and parameters of Gleeble simulator
thermal cycles on properties of the intercritically reheated heat affected zone,
boundaries is inversely proportional to GB content. Mater. Des. 32 (2011) 869–876 https://doi.org/10.1016/j.matdes.2010.07.005.
[17] Kanwer Singh Arora, Sangeetha Ranga Pandu, Nikhil Shajan, Prashant Pathak,
5. Conclusions Mahadev Shome, Microstructure and impact toughness of reheated coarse grain
heat affected zones of API X65 and API X80 linepipe steels, Int. J. Press. Vessel. Pip.
163 (2018) 36–44 https://doi.org/10.1016/j.ijpvp.2018.04.004.
The microstructures and properties of IHAZ were studied through a [18] Seok Gyu Lee, Seok Su Sohn, Bohee Kim, Woo Gyeom Kim, Kyung-Keun Um,
thermal simulation test. The following major conclusions were drawn. Sunghak Lee, Effects of martensite-austenite constituent on crack initiation and
propagation in inter-critical heat-affected zone of high-strength low-alloy (HSLA)
steel, Mater. Sci. Eng. A 715 (2018) 332–339 https://doi.org/10.1016/j.msea.2018.
(1) The weld metal after a single thermal cycle (referred to as WM-CG 01.021.
in this paper) exhibited high impact energy. The toughness of weld [19] Wendong Zhang, Guoqun Zhao, Qianjin Fu, Study on the effects and mechanisms of
metal decreased after double thermal cycles, particularly that of induction heat treatment cycles on toughness of high frequency welded pipe welds,
Mater. Sci. Eng. A 736 (2018) 276–287 https://doi.org/10.1016/j.msea.2018.09.
WM–UA CG. 004.
(2) The corresponding relationship between the microstructure and [20] Xiong Zhang, Gaoyang Mi, Lingda Xiong, Chunming Wang, Effects of interlaminar
properties of IHAZ was established. The toughness of IHAZ was microstructural inhomogeneity on mechanical properties and corrosion resistance
of multi-layer fiber laser welded high strength low alloy steel, J. Mater. Process.
related to many factors. Toughness was primarily affected by GB Technol. 252 (2018) 81–89 https://doi.org/10.1016/j.jmatprotec.2017.09.012.
content. The smaller the grain size, the higher the content of high- [21] Y.L. Sun, G. Obasi, C.J. Hamelin, A.N. Vasileiou, T.F. Flint, J. Balakrishnan,
angle grain boundaries, and the better the toughness of weld metal. M.C. Smith, J.A. Francis, Effects of dilution on alloy content and microstructure in
multi-pass steel welds, J. Mater. Process. Technol. 265 (2019) 71–86 https://doi.
org/10.1016/j.jmatprotec.2018.09.037.
References [22] Z.J. Xie, G. Han, W.H. Zhou, X.L. Wang, C.J. Shang, R.D.K. Misra, A novel multi-step
intercritical heat treatment induces multi-phase microstructure with ultra-low yield
ratio and high ductility in advanced high-strength steel, Scr. Mater. 155 (2018)
[1] A.R.H. Midawi, C.H.M. Simha, A.P. Gerlich, Assessment of yield strength mismatch
164–168 https://doi.org/10.1016/j.scriptamat.2018.06.042.
in X80 pipeline steel welds using instrumented indentation, Int. J. Press. Vessel. Pip.
[23] Yang Peng, Chao Wu, Jieliang Gan, Jun Dong, Characterization of heterogeneous
168 (2018) 258–268 https://doi.org/10.1016/j.ijpvp.2018.09.014.
constitutive relationship of the welded joint based on the stress-hardness relation-
[2] Jiankai Tang, Zheng Liu, Shouwen Shi, Xu Chen, Evaluation of fracture toughness in
ship using micro-hardness tests, Constr. Build. Mater. 202 (2019) 37–45 https://
different regions of weld joints using unloading compliance and normalization
doi.org/10.1016/j.conbuildmat.2018.12.218.
method, Eng. Fract. Mech. 195 (2018) 1–12 https://doi.org/10.1016/j.
[24] Reza Khatib Zadeh Davani, Reza Miresmaeili, Mohammadreza Soltanmohammadi,
engfracmech.2018.03.022.
Effect of thermomechanical parameters on mechanical properties of base metal and
[3] Junyan Ni, Xiaowei Wang, Jianming Gong, Magd Abdel Wahab, Thermal, me-
heat affected zone of X65 pipeline steel weld in the presence of hydrogen, Mater.
tallurgical and mechanical analysis of circumferentially multi-pass welded P92 steel
Sci. Eng. A 718 (2018) 135–146 https://doi.org/10.1016/j.msea.2018.01.101.
pipes, Int. J. Press. Vessel. Pip. 165 (2018) 164–175 https://doi.org/10.1016/j.
[25] X.L. Wang, X.M. Wang, C.J. Shang, R.D.K. Misra, Characterization of the multi-pass
ijpvp.2018.06.009.
weld metal and the impact of retained austenite obtained through intercritical heat
[4] Y.L. Sun, G. Obasi, C.J. Hamelin, A.N. Vasileiou, T.F. Flint, J. Balakrishnan,
treatment on low temperature toughness, Mater. Sci. Eng. A 649 (2016) 282–292
M.C. Smith, J.A. Francis, Effects of dilution on alloy content and microstructure in
https://doi.org/10.1016/j.msea.2015.09.030.
multi-pass steel welds, J. Mater. Process. Technol. 265 (2019) 71–86 https://doi.
[26] X.L. Wang, Y.R. Nan, Z.J. Xie, Y.T. Tsai, J.R. Yang, C.J. Shang, Influence of welding
org/10.1016/j.jmatprotec.2018.09.037.
pass on microstructure and toughness in the reheated zone of multi-pass weld metal
[5] Huai Wang, Wanchuck Woo, Dong-Kyu Kim, VyacheslavEm, SooYeol Lee, Effect of
of 550 MPa offshore engineering steel, Mater. Sci. Eng. A 702 (2017) 196–205
chemicaldilution and the number of weld layers on residual stresses in a multi-pass
https://doi.org/10.1016/j.msea.2017.06.081.

116
F. Bai, et al. Progress in Natural Science: Materials International 30 (2020) 110–117

[27] M. Mohammadijoo, J. Valloton, L. Collins, H. Henein, D.G. Ivey, Characterization of martensite–austenite constituents in simulated welding conditions, J. Mater.
martensite-austenite constituents and micro-hardness in intercritical reheated and Process. Technol. 153–154 (2004) 87–92 http://doi.org/10.1016/j.jmatprotec.
coarse-grained heat affected zones of API X70 HSLA steel, Mater. Char. 142 (2018) 2004.04.021.
321–331 https://doi.org/10.1016/j.matchar.2018.05.057. [35] Xiang Luo, Xiaohua Chen, Tao Wang, Shiwei Pan, Zidong Wang, Effect of
[28] Nazmul Huda, Abdelbaset R.H. Midawi, James Gianetto, Robert Lazor, Adrian morphologies of martensite–austenite constituents on impact toughness in inter-
P. Gerlich, Influence of martensite-austenite (MA) on impact toughness of X80 line critically reheated coarse-grained heat-affected zone of HSLA steel, Mater. Sci. Eng.
pipe steels, Mater. Sci. Eng. A 662 (2016) 481–491 https://doi.org/10.1016/j.msea. A 710 (2018) 192–199 https://doi.org/10.1016/j.msea.2017.10.079.
2016.03.095. [36] Liangyun Lan, Chunlin Qiu, Dewen Zhao, Xiuhua Gao, Linxiu Du, Analysis of
[29] L. Shi, S.A. Alexandratos, N.P. O'Dowd, Prediction of prior austenite grain growth in martensite-austenite constituent and its effect on toughness in submerged arc
the heat-affected zone of a martensitic steel during welding, Int. J. Press. Vessel. welded joint of low carbon bainitic steel, J. Mater. Sci. 47 (2012) 4732–4742
Pip. 166 (2018) 94–106 https://doi.org/10.1016/j.ijpvp.2018.08.005. https://doi.org/10.1007/s10853-012-6346-x.
[30] Gaojun Mao, Cyril Cayron, Rui Cao, Roland Logé, Jianhong Chen, The relationship [37] Liangyun Lan, Chunlin Qiu, Hongyu Song, Dewen Zhao, Correlation of martensi-
between low-temperature toughness and secondary crack in low-carbon bainitic te–austenite constituent and cleavage crack initiation in welding heat affected zone
weld metals, Mater. Char. 145 (2018) 516–526 https://doi.org/10.1016/j.matchar. of low carbon bainitic steel, Mater. Lett. 125 (2014) 86–88 https://doi.org/10.
2018.09.012. 1016/j.matlet.2014.03.123.
[31] N. Takayama, G. Miyamoto, T. Furuhara, Chemistry and three-dimensional mor- [38] Xueda Li, Chengjia Shang, Xiaoping Ma, Baptiste Gault, S.V. Subramanian,
phology of martensite-austenite constituent in the bainite structure of low-carbon Jianbo Sun, R.D.K. Misra, Elemental distribution in the martensite–austenite con-
low-alloy steels, Acta Mater. 145 (2018) 154–164 https://doi.org/10.1016/j. stituent inintercritically reheated coarse-grained heat-affected zone of a high-
actamat.2017.11.036. strength pipeline steel, Scr. Mater. 139 (2017) 67–70 https://doi.org/10.1016/j.
[32] Jiming Zhang, Weihua Sun, Hao Sun, Mechanical properties and microstructure of scriptamat.2017.06.017.
X120 Grade high strength pipeline steel, J. Iron Steel Res. Int. 17 (2010) 63–67 [39] Julian A. Avila, Johnnatan Rodriguez, Paulo Roberto Mei, Antonio J. Ramirez,
https://doi.org/10.1016/S1006-706X(10)60185-9. Microstructure and fracture toughness of multipass friction stir welded joints of
[33] Z.J. Xie, S.F. Yuan, W.H. Zhou, J.R. Yang, C.J. Shang, Stabilization of retained API-5L-X80 steel plates, Mater. Sci. Eng. A 673 (2016) 257–265 https://doi.org/10.
austenite by the two-step intercritical heat treatment and its effect on the toughness 1016/j.msea.2016.07.045.
of a low alloyed steel, Mater. Des. 59 (2014) 193–198 https://doi.org/10.1016/j. [40] Liangyun Lan, Meng Yu, Chunlin Qiu, On the local mechanical properties of iso-
matdes.2014.02.035. thermally transformed bainite in low carbon steel, Mater. Sci. Eng. A 742 (2019)
[34] Emin Bayraktar, Dominique Kaplan, Mechanical and metallurgical investigation of 442–450 https://doi.org/10.1016/j.msea.2018.11.011.

117

You might also like