You are on page 1of 31

Journal Pre-proof

Towards an explanation of liquid metal embrittlement cracking in


resistance spot welding of dissimilar steels

Zhanxiang Ling, Min Wang, Liang Kong, Ke Chen

PII: S0264-1275(20)30590-6
DOI: https://doi.org/10.1016/j.matdes.2020.109055
Reference: JMADE 109055

To appear in: Materials & Design

Received date: 17 June 2020


Revised date: 5 August 2020
Accepted date: 10 August 2020

Please cite this article as: Z. Ling, M. Wang, L. Kong, et al., Towards an explanation
of liquid metal embrittlement cracking in resistance spot welding of dissimilar steels,
Materials & Design (2020), https://doi.org/10.1016/j.matdes.2020.109055

This is a PDF file of an article that has undergone enhancements after acceptance, such
as the addition of a cover page and metadata, and formatting for readability, but it is
not yet the definitive version of record. This version will undergo additional copyediting,
typesetting and review before it is published in its final form, but we are providing this
version to give early visibility of the article. Please note that, during the production
process, errors may be discovered which could affect the content, and all legal disclaimers
that apply to the journal pertain.

© 2020 Published by Elsevier.


Journal Pre-proof
Towards an explanation of liquid metal embrittlement cracking in resistance
spot welding of dissimilar steels
Zhanxiang Linga, Min Wanga*, Liang Konga*, Ke Chena

a
Shanghai Key Laboratory of Materials Laser Processing and Modification, Shanghai Jiao Tong University,
Shanghai, 200240, China

Zhanxiang Ling, first author, PhD student, Email: zxling@sjtu.edu.cn


Min Wang, corresponding author, PhD, Professor, Email: wang-ellen@sjtu.edu.cn

of
Liang Kong, co-corresponding author, PhD, Associate research fellow, Email: ingerkongliang@sjtu.edu.cn
Ke Chen, PhD, Associate Professor, Email: chenke83@sjtu.edu.cn

ro
*Corresponding author: Min Wang and Liang Kong
-p
E-mail address: wang-ellen@sjtu.edu.cn, ingerkongliang@sjtu.edu.cn
re
Tel.: +86-021-34202679
lP

Postal address: Room 403, F building, School of Materials Science and Engineering, Shanghai Jiao Tong
University, No.800 Dongchuan Road, Minhang District, Shanghai, China, 200240
na
ur
Jo
Journal Pre-proof

Towards an explanation of liquid metal embrittlement cracking in resistance


spot welding of dissimilar steels

Abstract
Automotive advanced high-strength steels and ultra-high-strength steels, such as Q&P980 steel with a
Zn coating, are prone to liquid metal embrittlement (LME) surface cracking during resistance spot welding,
while traditional low carbon steels (LCSs) do not appear to undergo LME. In this work, galvanized Q&P980
steel sheets were welded with LCS sheets. LME cracks were found on both the Q&P980 steel and LCS sides.

of
However, the severity of LME cracking on the Q&P980 side was much higher than that on the LCS side,
indicating that Q&P980 steel has a higher LME susceptibility than LCS, which was also proven by a hot

ro
tensile test. The higher LME susceptibility of Q&P980 was attributed to its lower reactivity with liquid Zn,
-p
higher achievable stress and increased number of LME-susceptible grain boundaries. Micro-analysis
revealed that Zn most likely entered the steel matrix via diffusion and not via liquid penetration, and
re
stress-assisted diffusion is the likely mechanism for LME.
lP

Key words: Resistance spot welding; Liquid metal embrittlement; Q&P980 steel; Low carbon steel.
na

1 Introduction
ur

Liquid metal embrittlement (LME) of Zn-coated steels during resistance spot welding (RSW) is a major
Jo

concern among materials scientists and in the automobile industry. LME refers to a phenomenon that
describes the loss of ductility of a normally ductile material while it is in contact with a liquid metal, and
brittle fracture may occur if a tensile stress is applied [1]. During RSW of galvanized steels, the zinc coating
on the steel surface melts and makes contact with the solid steel, and the electrode force and rapid thermal
cycling during RSW generate a tensile stress on the steel surface and cause LME, which is manifested by
LME cracks on the surface of weld spots [2]. These cracks potentially deteriorate the mechanical
performance of spot-welded joints [3-5].
The susceptibility to LME differs among different galvanized steels [6]. The type of coating can
influence LME behaviour, as studies have shown that hot-dipped galvanized (GI) steels demonstrated higher
LME susceptibility than electrogalvanized (EG) and galvannealed (GA) steels [7,8]. In addition, Kalashami
et al. [9] found that internal oxides formed during galvanizing of a high strength steel could assist LME
Journal Pre-proof
crack formation during RSW. Nevertheless, understanding the intrinsic susceptibility of different steel types
to LME is important. It is widely considered that the LME susceptibility increases with increasing strength
level of the steel. In fact, LME has mostly been reported in the RSW of advanced high-strength steels
(AHSSs) and ultra-high-strength steels (UHSSs), such as transformation-induced plasticity (TRIP) steels
[10-13], dual phase (DP) steels [9,14,15], twinning-induced plasticity (TWIP) steels [16,17], and quenching
and partitioning (Q&P) steels [18]. These steels are important lightweight materials in the automobile
industry. The restricted application of these steels because of LME is the main reason that the topic has
received a substantial amount of attention in recent years. In contrast, to the best knowledge of the authors,
LME has never been reported in traditional mild steels. Béal et al. [19] hypothesized that there was a critical
stress for LME to occur, and the risk of achieving the critical stress is higher in harder steels. The constituent

of
microstructure of the steel also influences its LME behaviour. Sigler et al. [20] claimed that a microstructure
that has fully transformed to austenite during RSW is essential for the occurrence of LME, leading to the

ro
suspicion that austenite might be more prone to LME than ferrite. However, Jung et al. [21] found that the
-p
occurrence of LME was not related to the microstructure, as LME was observed in both the ferrite and
austenite temperature ranges in a hot tensile test. Moreover, Tumuluru [22] compared the LME behaviour of
re
GEN3 steel with different retained austenite contents in the original microstructure and found that retained
lP

austenite was not required for LME to occur. The alloying elements in steel also play a role. Tumuluru [22]
found that increasing the Si content in the steel aggravated LME as Si modified the grain boundary
na

characteristics and made them prone to LME. Hong et al. [23] also found that Si promoted LME, but they
suggested that Si delayed the reaction between the steel matrix and Zn coating. From the above, we can see
ur

that the different LME behaviours of different steels were attributed to a complex combination of the
Jo

strength level, microstructure, and alloying elements, to name a few, and the phenomenon still remains
unclear.
The mechanism of LME is the key to understanding it. Several models [24] have been proposed since
the discovery of LME, and uncertainty still exists today. As the LME-induced fracture in Zn-coated steels is
mainly intergranular, the grain boundary penetration model was frequently cited to explain the LME of
Zn-coated steels [25]; however, it is unclear how Zn penetrates the steel grain boundaries (GBs). Some
researchers believed that Zn penetrates the GBs by liquid percolation [17], while others believed that it is by
solid diffusion [9,18,26]. The lack of direct evidence makes it difficult to make a judgement, which has
hindered the understanding of LME of Zn-coated steels during RSW.
Most research on LME during RSW has focused on the two-sheet similar steel stack-up configuration,
while few studies have been done on the three-sheet dissimilar steel stack-up configuration [27]. Benlatreche
et al. [28] studied the LME of several AHSSs and welded the AHSS to either a DP980 steel sheet or two low
Journal Pre-proof
carbon steel (LCS) sheets and found that LME cracking in the AHSS was more severe in the three-sheet
stack-up than that in the two-sheet stack-up owing to the higher heat input in three-sheet stack-up. However,
LME cracking on the LCS side was not mentioned in their work. Recently, WorldAutoSteel presented a
three-year study report on LME during RSW, and the three-sheet stack-up configuration was frequently used
in this work [29]. Moreover, the three-sheet stack-up (AHSS/LCS/LCS) has been designated as the test
combination to evaluate the weldability of AHSSs in certain automobile industry standards [30], indicating
that such a stack-up may be encountered in practice. Thus, studying LME behaviour in the three-sheet
stack-up configuration is of great interest and necessary. In the present work, we studied the LME cracking
behaviour in a resistance spot welded Q&P980/LCS/LCS dissimilar steel joint. The LME cracking
behaviour on both the Q&P980 steel and LCS sides was studied and compared. By using various

of
characterization methods, we propose the fundamental reasons for the different LME susceptibility of
different steels and provide key evidence to illuminate the LME mechanism, both of which are needed to

ro
understand the LME phenomenon during RSW. -p
re
2 Materials and methods
The materials used for welding in the present work were a 1.4 mm-thick galvanized Q&P980 steel, a 2.0
lP

mm-thick galvanized LCS and a 0.7 mm-thick galvanized LCS. The thicknesses of the steel sheets were
selected based on the standard [30]. The thicknesses of the Zn coatings on all steels were approximately 10
na

μm. Q&P980 steel is a third-generation UHSS composed of martensite (M), ferrite (F) and retained austenite
(RA), and the LCS was composed of ferrite and perlite (P), as shown in Fig. 1. The original grain size in the
ur

LCS was much larger than that in the Q&P980 steel. The chemical compositions of the Q&P980 steel and
Jo

LCS are listed in Table 1.


The stack-up is shown in Fig. 2. Three sheets were placed sequentially between electrodes. Three
different welding schedules, namely, a single-pulse welding schedule and two multiple-pulse welding
schedules, were used, which are schematically shown in Fig. 3. The welding experiment was conducted on a
medium-frequency direct current resistance spot welder with a cooling water rate of 4 L/min. The Cu-Cr-Zr
electrode was dome-radius type with a tip diameter of 8 mm, and its chemical composition is listed in Table
1. The welding parameters for all three schedules are listed in Table 2, and the parameters were the same for
different schedules with the exception of the welding time. The welding current was selected as the
expulsion current, which was defined as the minimum welding current at which all welding samples showed
expulsion. Expulsion is well known to facilitate LME cracking [18,19,27]. Six samples were welded under
each welding schedule to observe and statistically analyse the LME cracking.
Journal Pre-proof
After welding, all weld spots were wire cut along the centerline and made into metallurgical samples via
standard metallographic procedures. Then, the depth and number of LME cracks were observed and counted
with optical microscopy (OM, ZEISS AX10). The weld spots were scanned by X-ray microscopy (XRM,
Xradia 520 Versa) and reconstructed in commercial software Dragonfly to reveal the three-dimensional (3D)
characteristics of the LME cracks. Typical LME cracks were selected and observed via scanning electron
microscopy (SEM, VEGA-3-XMU) on an instrument equipped with energy dispersive spectrometry (EDS,
AZtec X-MaxN80). Electron back-scatter diffraction (EBSD) analysis was conducted by a GAIA3 GMU
scanning electron microscope equipped with focus iron beam (FIB) capability. Specimens for EBSD
analysis were prepared by an additional vibration polishing process. Samples for transmission electron
microscopy (TEM) were prepared by FIB, and the observations were conducted via a Talos F200X G2

of
transmission electron microscope equipped with EDS.
A Gleeble 3800 thermal simulator was used to carry out hot tensile tests on the bare/galvanized

ro
Q&P980 steel and LCS to observe the LME. The bare steels were prepared by immersing galvanized steels
-p
in a hydrochloric acid solution to remove the Zn coating. The tests were conducted from 600-1000 ℃ with a
heating rate of 200 ℃/s. A tensile force was immediately applied to the specimens with a strain rate of 0.1/s
re
when the temperature reached the target temperature. The dimensions of the specimens were in accordance
lP

with the work of Beal et al. [31], and the thicknesses of Q&P980 steel and LCS used in the hot tensile tests
were both 1.4 mm.
na

Table 1 Chemical composition of materials used in the present work (wt%).


ur

Material Mn Si C Al S P Fe Cr Zr Cu
Jo

Q&P980 2.248 1.774 0.1792 0.0418 0.0002 0.0075 Bal. - - -


LCS 0.10 - 0.012 0.04 0.03 0.03 Bal. - - -
Cu-Cr-Zr - - - - - - - 0.55 0.12 Bal.
Journal Pre-proof

of
ro
Fig. 1 Microstructures of the (a) Q&P980 steel and (b) LCS observed with SEM and EBSD inverse pole
-p
figure (IPF) maps of the (c) Q&P980 steel and (d) LCS.
re
lP
na
ur
Jo

Fig. 2 Schematic of three-sheet stack-up for resistance spot welding.

Fig. 3 Schematic diagrams of welding schedules with (a) one pulse, (b) two pulses and (c) four pulses.
Journal Pre-proof
Table 2 Welding parameters used in the present work.

Welding parameter Value

Electrode force 5 kN
Squeezing time 2000 ms
Holding time 300 ms
Welding time 480 ms/2×240 ms/4×120 ms
Cooling time between pulses 20 ms
Welding current 12 kA

3 Results and discussion

of
3.1 Hot tensile test

ro
The LME of galvanized steels has been widely studied by hot tensile tests, which can be used to
-p
interpret the LME susceptibility very well [31-35]. Thus, hot tensile tests of the Q&P980 steel and LCS
were conducted to demonstrate their LME susceptibilities. Fig. 4a and Fig. 4b are the representative tensile
re
curves of the bare/galvanized Q&P980 steel and bare/galvanized LCS at 800 ℃, respectively. At 800 ℃, the
lP

ductility of the galvanized Q&P980 steel was severely reduced compared with that of the bare Q&P980 steel.
In contrast, only a slight reduction in the ductility was observed in the LCS. The occurrence of LME can be
na

determined by the value of the relative reduction of energy (RRE), which can be calculated as follows:
𝐸𝐺 = ∫ 𝜎𝐺 𝑑𝜀𝐺 (1)
ur

𝐸𝐵 = ∫ 𝜎𝐵 𝑑𝜀𝐵 (2)
𝐸𝐺 −𝐸𝐵
Jo

𝑅𝑅𝐸 = × 100 % (3)


𝐸𝐵

where EG and EB are the fracture energies of the galvanized steel and bare steel, respectively; σG and σB are
the stresses in the galvanized steel and bare steel, respectively; and εG and εB are the strains in the galvanized
steel and bare steel, respectively. The RRE curves for the Q&P980 steel and LCS at different temperatures
are shown in Fig. 4c, and an RRE value less than -10 % indicates that LME has occurred [32]. For the
Q&P980 steel, LME occurred in the temperature range from 700-950 ℃, and the LME was extremely
severe in the temperature range from 800-850 ℃. In contrast, for the LCS, only a slight LME occurred in
the temperature range from 800-850 ℃. The results confirmed that the Q&P980 steel had a much higher
intrinsic LME susceptibility than the LCS.
Journal Pre-proof

of
ro
-p
re
lP
na

Fig. 4 Tensile curves of the (a) bare/galvanized Q&P980 steel and (b) bare/galvanized LCS at 800 ℃ and (c)
ur

relative reduction of energy of the Q&P980 steel and LCS at different temperatures.
Jo

3.2 Statistics of LME cracking during RSW


LME cracks formed on both the Q&P980 steel side and the 0.7 mm-thick LCS side in the spot-welded
joints, and they were found only on the outside surface that contacted the electrode. Fig. 5 demonstrates
where the cracks were located and how their depths were measured. The depth of the cracks was measured
from the surface to the crack root, similar to what was reported by Wintjes et al. [3]. These cracks were
further categorized by five different depth ranges, namely, less than 20 μm, 20-50 μm, 50-100 μm, 100-300
μm and greater than 300 μm. The number of cracks in each range was counted separately, and the results of
the total number in all six samples for each welding schedule are shown in Fig. 6. For cracking on the
Q&P980 steel side, most cracks were shallow with depths less than 50 μm, and their numbers were quite
similar among the different welding schedules. However, the multiple-pulse samples possessed more deep
cracks, especially those with depths beyond 300 μm, as no such deep crack was found in the single-pulse
Journal Pre-proof
samples. These deep cracks are believed to have potentially detrimental effects on the mechanical
performance of resistance spot welded joints [5,36]. For cracking on the LCS side, the multiple-pulse
samples possessed more cracks in all depth ranges, and no crack with a depth beyond 300 μm was found in
the single-pulse samples either. The cracking behaviour is governed by the temperature and stress states
during welding. It is reasonable to conjecture that the deep cracks formed under the multiple-pulse welding
schedule were facilitated by the cooling time between pulses, as additional cooling and thermal stresses were
introduced compared with the single-pulse welding schedule. Fig. 6c shows a comparison of the cracking on
the Q&P980 steel and LCS sides. The number of cracks was the sum from all welding schedules, and it can
be seen that there were many more cracks on the Q&P980 steel side than on the LCS side.

of
ro
-p
re
lP
na

Fig. 5 Joint profile of a resistance spot welded joint and magnifications of the crack locations.
ur
Jo
Journal Pre-proof

of
ro
-p
re
lP

Fig. 6 Statistics of LME cracks in the (a) Q&P980 steel and (b) LCS for all welding schedules and (c)
na

comparison of LME cracks in the Q&P980 steel and LCS. The numbers are the total from all samples.
ur

It has been widely reported that high strength steels are very prone to LME during RSW, and generally,
Jo

the higher the strength level is, the higher the susceptibility. This knowledge is consistent with the present
results. The different LME cracking behaviour of Q&P980 steels and LCSs during RSW can be attributed to
their intrinsic susceptibility to LME, as demonstrated in the former section. As the LME temperature range
was narrow and the embrittlement was not severe, LME cracks did not easily form during RSW of the LCS.
In fact, to the best knowledge of the authors, LME during RSW of LCSs has not been reported before. To
further confirm that the cracks in the LCS in the present work were caused by LME, several LCS sheets
were immersed in the hydrochloric acid solution to remove their Zn coating and were subsequently welded
under the same conditions. No cracking was found on the LCS side in these samples.

3.3 3D reconstruction of LME cracks


To reveal the overall distribution of the LME cracks, a circular spot weld on the Q&P980 steel side was
Journal Pre-proof
wire cut from the welding specimen and scanned by XRM. The scanned file was then imported into
Dragonfly image processing and 3D reconstruction software, within which the 3D scanning image was
established. The results are shown in Fig. 7. Fig. 7a shows the reconstructed cracks together with the weld
spot, which clearly demonstrates the distribution of cracks in the weld spot. Fig. 7b shows the thickness
distribution of the reconstructed cracks. As the pixel resolution during XRM scanning was set as 12 μm,
only cracks with thicknesses greater than that were distinguished from the matrix. The thicknesses of most
cracks were between 24 and 28 μm. Using the processing software, the volume of each crack was also be
measured, as shown in Fig. 7c. It should be mentioned that nearly no cracks were observed in the LCS by
XRM under the same resolution.

of
ro
-p
re
lP
na
ur
Jo

Fig. 7 3D reconstruction of the LME cracking in Q&P980. (a) Reconstructed cracks and welding spot, (b)
reconstructed cracks with colour rendering showing the thickness distribution and (c) histogram showing the
volume distribution of cracks.

3.4 Zn state during LME cracking


LME cracks are formed by an interaction between the liquid Zn on the steel surface and the steel matrix,
and cracks nucleate and propagate after the interaction, so a concentration of Zn is expected along the cracks.
In addition, liquid Zn can react with the Cu electrode at high temperatures; therefore, Cu is also likely to
enter the matrix along with Zn. Fig. 8 shows a typical LME crack formed on the Q&P980 steel side. Back
scattered electron (BSE) images were captured, and element distribution maps were obtained. The chemical
Journal Pre-proof
compositions of the analysis positions marked in the figure are listed in Table 3. Fig. 8a shows a crack at the
shoulder; the cracks there were usually deep [18]. A concentration of Zn can be seen along the crack, but its
weight percentage was 3-16 %, which indicates that the liquid phase did not exist inside the crack. No
concentration of Cu was observed because the electrode did not reach this position. Fig. 8b shows a crack at
the indentation center, and cracks there were usually deep [18]. Concentrations of Zn and Cu can be seen at
the beginning of the crack, with weight percentages of 44.76 % and 10.79 %, respectively. Deep within the
crack, Zn and Cu became rare, but a small amount of Zn was detected at the crack tip. The majority of
molten Zn was ejected by the electrode during welding, so liquid Zn was scarce here. It is very likely that
the existence of Zn deep within the crack was due to an atom diffusion mechanism rather than a liquid phase
penetration mechanism. To further confirm that hypothesis, a TEM sample was obtained by the FIB method

of
at the crack tip, as indicated by the white rectangle for FIB-1. Chemical and phase analysis were conducted
at the micro-scale, and the results are discussed later. Fig. 8c shows a crack at the indentation edge; the

ro
cracks there were usually shallow [18]. From the BSE image, we can clearly see that a different phase
-p
existed inside the crack, and element maps show that both Zn and Cu were concentrated inside the crack.
The weight percentages of Zn at the beginning and middle positions of the crack were over 70 %, which is
re
strong evidence that a liquid Zn-rich phase existed inside the crack during welding, as indicated in the
lP

literatures [25,34]. The weight percentage of Zn decreased at the crack tip; however, this may be because the
magnification was not high enough. To further confirm whether the liquid phase reached the crack tip, a
na

TEM sample was obtained by the FIB method at the crack tip, as indicated by the white rectangle for FIB-2.
A detailed analysis of the sample is discussed later. Fig. 8d shows several small cracks at the indentation
ur

edge; unlike the crack shown in Fig. 8c, no evidence of a liquid phase inside the crack was found, as
Jo

indicated by the EDS element maps and point analysis results. Another interesting result from the point
analysis is that the Si and Al contents were abnormally high inside the deep crack, and one may be curious if
the Si and Al played a role during the propagation of the crack. However, their contents were low inside the
shallow cracks, especially the cracks filled with the liquid phase. Thus, the Si and Al inside the deep hollow
cracks may have been derived from the sandpaper or the polishing paste, which contained SiC and Al2O3.
They could have become trapped inside the hollow cracks during the polishing procedure and might lead to
the misleading impression that the Si and Al in the matrix segregated to the cracking area. Whether
segregation occurred should be studied at a more microscopic scale such as using TEM.
Journal Pre-proof

of
ro
-p
re
lP

Fig. 8 LME cracks on the Q&P980 side. (a) BSE image of a deep crack at the shoulder and EDS maps of Fe,
na

Zn and Cu; (b) BSE image of a deep crack at the indentation center and EDS maps of Fe, Zn and Cu; (c)
BSE image of a shallow crack at the indentation edge and EDS maps of Fe, Zn and Cu; and (d) BSE image
ur

of shallow cracks at the indentation edge and EDS maps of Fe, Zn and Cu.
Jo
Journal Pre-proof
Table 3 EDS results with analysis positions taken from Fig. 8.

Source Analysis position Fe wt% Zn wt% Cu wt% Mn wt% Si wt% Al wt%

Spectrum 1 54.56 13.01 0.58 1.65 28.77 1.44


Spectrum 2 59.87 3.21 0.62 1.72 13.96 7.72
Fig. 8a
Spectrum 3 66.74 16.00 1.81 4.69 8.43 2.33
Spectrum 4 62.54 15.1 1.57 2.51 12.83 5.44
Spectrum 5 34.69 44.76 10.79 2.33 3.17 4.26
Spectrum 6 76.39 14.57 1.20 1.73 4.49 1.61
Fig. 8b
Spectrum 7 80.51 2.11 0.08 2.36 14.09 0.85
Spectrum 8 64.38 4.63 0.97 1.80 25.05 3.17

of
Spectrum 9 9.18 72.67 18.06 0.09 0 0

ro
Spectrum 10 14.85 70.88 13.78 0.40 0.06 0.03
Fig. 8c
Spectrum 11 36.47 44.91 -p 16.46 1.22 0.08 0.06
Spectrum 12 68.57 24.27 3.16 2.50 1.49 0
re
Spectrum 13 83.42 10.14 0.08 2.11 3.37 0.07
Spectrum 14 84.53 2.12 4.00 2.20 6.96 0.18
lP

Fig. 8d
Spectrum 15 77.02 9.53 2.40 1.65 9.24 0.17
Spectrum 16 60.49 20.77 16.01 1.86 0.86 0.00
na
ur

Fig. 9 shows the TEM results of the FIB-1 sample marked in Fig. 8. The crack tip was hollow and not
Jo

filled with any substances. Element mapping of the rectangular area shows no obvious concentration of Zn
around the crack. Selected area diffraction (SAD) and chemical composition analysis were conducted at
positions d-f. The SAD patterns demonstrated bcc iron phase at all positions. Zn was also detected at these
positions; the content of Zn decreased with increasing distance from the crack, and the highest content of Zn
was 9.42 % at position d. The low content of Zn indicated that no liquid phase reached the crack tip, and the
existence of Zn there was a result of Zn diffusion. In addition, the contents of Si, Mn, and Al also decreased
with increasing distance from the crack. Note that their contents are higher than those in the matrix at
positions d and e and lower than that in the matrix at position f. It seems that these alloying elements
segregated to the crack. Due to the unique preparation process of FIB, the sample was lifted out beneath the
surface and thinned by an iron beam, and the effect of impurity introduction during polishing should be
excluded. Thus, the segregation of alloying elements revealed by chemical composition analysis at the TEM
Journal Pre-proof
scale should be convincing.

of
ro
-p
Fig. 9 TEM results of sample FIB-1 marked in Fig. 8. (a) Bright-field (BF) image of the cracked zone, (b)
re
and (c) EDS maps of Fe and Zn in the rectangular area in (a) and (d)-(f) SAD patterns of and element
compositions of position d-f marked in (a).
lP

Fig. 10 shows the TEM results of the FIB-2 sample marked in Fig. 8. The crack tip was filled by a
na

different phase. Element mapping shows an obvious Zn concentration inside the whole crack. Line scanning
across the crack and matrix again shows a concentration of Zn inside the crack, and there is a clear
ur

concentration of Si at the phase boundary. Fig. 10e and Fig. 10f show the SAD patterns and chemical
Jo

composition analysis results at positions e and f. The Γ-Fe3Zn10 phase and bcc iron phase were distinguished
at position e, and the phase composition inside the crack was mainly composed of the Γ phase that was
mixed with the bcc iron phase, which explains why the Zn content was lower than the Γ phase inside the
crack. Position f was predominantly the bcc iron phase, with a very small amount of Zn that was likely cause
by diffusion. Fig. 10g shows a high-resolution TEM (HRTEM) image of the phase boundary, and a
magnification of the rectangular areas shows the atomic arrangement of the two phases. The interplanar
spacing was in accordance with the bcc iron and Γ phases. Fig. 10j shows the HRTEM image of the phase
boundary on the other side at the same incident angle of the electron beam. The atomic arrangement of the Γ
phase was still distinguishable, but the bcc iron phase was no longer distinguishable. This phenomenon
indicated that the grain orientation on both sides of the Γ phase was different, which provided strong
evidence that the crack was intergranular.
Journal Pre-proof

of
ro
-p
re
lP
na

Fig. 10 TEM results of sample FIB-2 marked in Fig. 6. (a) Bright-field (BF) image of the cracked zone, (b)
and (c) EDS maps of Fe and Zn, (d) EDS line profile marked in (a), (e)-(f) SAD patterns of and element
ur

compositions of positions marked in (a), (g) high-resolution TEM micrograph of the interphase area, (h)-(i)
high-resolution TEM micrograph of the area labelled in (g), and (j) high-resolution TEM micrograph of the
Jo

interphase area on the other side.

Fig. 11 shows typical LME cracks formed on the LCS side. BSE images were captured, and element
distribution maps were obtained. The chemical compositions of the analysis positions marked in the figure
are listed in Table 4. Fig. 11a shows a crack at the indentation center, and the crack was very deep and wide.
The BSE image shows that a different phase filled the coating and Cu electrode and likely entered the crack
in a liquid state. The high content of Cu indicates that the temperature here was very high during welding, as
the higher the content of Cu is, the higher the melting point of the Cu-Zn alloy. Fig. 11b shows another crack
at the indentation center; the crack was shallow and narrow compared with the shape of the former one, but
the phases inside the cracks were similar, as their chemical compositions were close. Fig. 11c shows several
shallow cracks at the indentation edge; different phases were not observed inside the crack, and
Journal Pre-proof
concentrations of Cu and Zn were not found. The Zn and Cu inside the crack may be lost during sample
preparation. However, a very small amount of Zn was still detected inside the crack, indicating that a
diffusion phenomenon occurred. In addition, Si and Al were found inside the hollow cracks, as Si and Al in
the LCS matrix and coating were extremely rare, which further proves that the Si and Al concentrations
detected under the SEM were due to the introduction of impurities during the polishing procedure. Although
deep cracks formed at the shoulder on the Q&P980 steel side, no cracks at a similar position were found on
the LCS side.

of
ro
-p
re
lP
na
ur

Fig. 11 LME cracking on the LCS side. (a) BSE image of a deep crack at the indentation center and EDS
Jo

maps of Fe, Zn and Cu; (b) BSE image of a shallow crack at the indentation center and EDS maps of Fe, Zn
and Cu; and (c) BSE image of a shallow crack at the indentation edge and EDS maps of Fe, Zn and Cu.
Journal Pre-proof
Table 4 EDS results with analysis positions taken from Fig. 11.

Source Analysis position Fe wt% Zn wt% Cu wt% Mn wt% Si wt% Al wt%

Spectrum 17 2.24 36.49 61.22 0.01 0.00 0.04


Spectrum 18 2.04 36.33 61.44 0.04 0.05 0.09
Fig. 11a
Spectrum 19 3.01 41.88 55.03 0.00 0.01 0.07
Spectrum 20 7.14 42.32 50.4 0.10 0.02 0.01
Spectrum 21 9.28 46.82 43.57 0.20 0.13 0.00
Fig. 11b Spectrum 22 15.11 42.71 41.61 0.35 0.18 0.03
Spectrum 23 18.91 47.02 33.95 0.03 0.10 0.00
Spectrum 24 97.08 0.68 0.55 0.00 1.50 0.19
Fig. 11c

of
Spectrum 25 98.26 0.17 0.00 0.17 1.11 0.16

ro
Based on the former results, we knew that in some crack tips, the liquid phase existed, but in some crack
-p
tips, it did not. Thus, the existence of the liquid phase at the crack tip should not be a necessary condition to
re
promote LME cracking. It is more likely that the cracks were generated via a stress-assisted diffusion
mechanism, which was first proposed by Gordon and An [37]. After the crack opened, the liquid phase
lP

entered the crack via a capillary effect, but it depended on the availability of the liquid phase on the surface
of the steel.
na

3.5 Reasons for the different LME susceptibilities of the Q&P980 steel and LCS
ur

3.5.1 Availability of liquid metal


Jo

One of the prerequisites for LME is the existence of liquid metal. For the LME of galvanized steels,
liquid metal is derived from the melting of the Zn coating at high temperatures. The steel can react with
liquid Zn at high temperatures, forming various reaction products, such as Γ and α-Fe(Zn). These products
have a higher melting point than pure Zn; therefore, the amount of liquid metal is reduced. As the chemical
compositions of Q&P980 steels and LCSs are different, they are assumed to have different reactivities with
liquid Zn. The main difference between the two steels is that Q&P980 steels have high amounts of Si and
Mn. Recently, Hong et al. [23] found that electrogalvanized TWIP steel with 1.5 % Si was more susceptible
to LME than that without Si; they claimed that Si in the steel delayed the reaction of Fe and liquid Zn, which
facilitated Zn penetration. To confirm the different reactivities of the two steels with liquid Zn, we analysed
the residual Zn coating on the steel surfaces of both resistance spot welded samples and hot tensile test
samples, and the results are shown in Fig. 12. Obviously, the contents of Zn in the residual coatings were
Journal Pre-proof
high in the resistance spot welded and hot tensile test Q&P980 steel samples. As the thermal cycles during
the hot tensile test at the studied locations for the Q&P980 steel and LCS were completely the same, it is
clear that the LCS reacted faster with the Zn and reduced the amount of liquid Zn at high temperatures. This
can be one of the reasons that the LCS was less susceptible to LME than the Q&P980 steel.

of
ro
-p
re
Fig. 12 Chemical compositions of the residual coating (a) at the indentation edge of the Q&P980 steel after
welding, (b) near the fracture location and (c) 8 mm away from fracture location of Q&P980 steel after hot
lP

tensile test at 800 ℃, (d) at the indentation edge of the LCS after welding, (e) near the fracture location and
(f) 8 mm away from fracture location of the LCS after hot tensile test at 800 ℃.
na

3.5.2 Magnitude of the tensile stress


ur

Tensile stress is another prerequisite because it prompts the opening and propagation of a crack. Béal
Jo

[19] suggested that LME cracking occurs only when the stress exceeds a critical stress. As the Q&P980 steel
had a much higher tensile strength than the LCS, a higher stress can be achieved in Q&P980 steels before
LME occurs, which can be seen from the hot tensile test results (Fig. 4). The higher stress increased the risk
of local failure. Nevertheless, the stress states in the two steels during RSW should be determined. As it is
impossible to monitor the stress during RSW, we carried out finite element simulation using the commercial
software ANSYS for a three-sheet resistance spot weld. Detailed descriptions of the modelling method can
be found in our previously published works [38,39]. The feasibility of the model for the three-sheet stack-up
configuration was validated by comparing the experimental and simulated nugget diameter (D), as shown in
Fig. 13. It should be mentioned that the model could not simulate the expulsion during RSW, and expulsion
resulted in a decrease in the nugget size; thus, the simulation and experiment were both carried out under a
welding current of 11 kA without expulsion using the double-pulse welding schedule.
Journal Pre-proof

Fig. 13 Comparation of the nugget sizes obtained from the experiments and simulation.

The stress histories along the Q&P980 steel and LCS surfaces at different locations experienced during
RSW were obtained from the simulation results, and the time with tensile stress is highlighted with blue
rectangles, as shown in Fig. 14. At all locations, the tensile stress on the Q&P980 surface was higher than

of
that on the LCS surface during RSW. It is also one of the reasons that LME cracking was more likely to

ro
occur in the Q&P980 steel. In addition, some other interesting conclusions can be drawn from Fig. 14. First,
the time that the surface suffered tensile stress at indentation center and shoulder is longer than that at
-p
indentation edge, which may be the main reason that deep cracks are always located at indentation center
re
and shoulder instead of indentation edge. Second, the cracks at indentation center and shoulder may generate
or propagate after the welding current was shut down, while cracks at indentation edge can only generate
lP

during welding time.


na
ur
Jo

Fig. 14 Comparation of stress histories during RSW at (a) indentation center, (b) indentation edge and (c)
shoulder in the Q&P980 steel and LCS.

3.5.3 Crystallographic characteristics


Fig. 15 shows the IPF maps of the crack areas on both the Q&P980 steel and LCS sides. The grains
were distinguished by different colours, which indicated different grain orientations. The grains on the LCS
side were very large compared with those on the Q&P980 steel side. This was because the original grain size
in the Q&P980 steel was much smaller than that in the LCS, as shown in Fig. 1. Along the cracking path, the
Journal Pre-proof
grain orientation on both sides of the crack was different, indicating an intergranular characteristic. However,
in a shallow crack on the Q&P980 steel side shown in Fig. 15d, the crack transferred from intergranular to
transgranular at the tip. Bhattacharya et al. [40] reported transgranular LME cracking in a hot tensile test of
galvanized steels; they claimed that when the crack encountered a grain with a grain boundary parallel to the
stress direction, the grain boundary could not be opened, and the crack would become transgranular,
although the LME crack always tended to propagate in an intergranular way. In the present work, it was
clear that the crack was arrested after it transformed into a transgranular crack. Unlike during the hot tensile
test, the steel surface experienced a tensile stress for a very short time during RSW; therefore, it was difficult
for transgranular LME cracks to propagate during RSW. The grain boundaries were the weak sites and the
preferred cracking path.

of
ro
-p
re
lP
na

Fig. 15 EBSD IPF maps (bcc phase) of (a) and (b) crack areas on the LCS side, (c) and (d) crack areas on
the Q&P980 side.
ur

Crystallographic characteristics are an important factor that determines the LME cracking behaviour.
Jo

GBs with high misorientation angles, especially those higher than 15°, were less stable than those with low
angles, and they were found to be weak sites and preferred paths for intergranular crack propagation [41].
Recently, Razmpoosh et al. [42] discovered that LME cracks tend to propagate along GBs with high-index
planes. In addition, Razmpoosh et al. [43] also proved that LME cracking occurred along high-angle GBs
(θ>15°). As LME cracks form at high temperatures, the characteristics of GBs in cracking areas at high
temperatures should be reconstructed to analyse the interaction between the GBs and cracks. Owing to the
very low hardenability of LCSs, no phase transformation occurred in the cracking areas during cooling.
Thus, high-angle GBs at room temperature could represent the real situation at high temperatures well.
However, the hardenability of Q&P980 steels is very high, leading to martensite transformation upon
cooling in cracking areas [18]. Thus, room temperature GBs in cracking areas contain GBs formed during
phase transformation, which should be excluded from analysis. It is easy to conjecture that LME cracking in
Journal Pre-proof
Q&P980 steels should propagate along austenite GBs at high temperatures; thus, prior austenite GBs should
be reconstructed in cracking areas in Q&P980 steels. We referred to the reconstruction method suggested by
Razmpoosh et al. [26]; that is, the GBs formed during martensite transformation should either be low-angle
GBs or high-angle GBs with angles in the range from 50 to 63°, prior austenite GBs could be reconstructed
by removing those GBs at room temperature. The blue lines in Fig. 16 show the reconstruction results of
high-angle GBs on the LCS side and prior austenite GBs on the Q&P980 steel side. It can be seen that the
GB size in the LCS was much larger than that in the Q&P980 steel, which was mainly due to the difference
in their original grain sizes (Fig. 1). The commercial software ImageJ was used to measure the proportion of
high-angle GBs in the LCS and prior austenite GBs in the Q&P980 steel, and the results are shown in the
top right corner in each figure. Note that the blue lines do not represent the real thickness of the GBs, so the

of
proportions can only be used to compare relevant differences among figures. We can see that the proportions
of GBs in the Q&P980 steel were much larger than those in the LCS, indicating that there were more

ro
susceptible GBs in the Q&P980 steel for the propagation of LME cracks than that in the LCS. It is another
-p
reason that Q&P980 steel is more susceptible to LME than LCS.
re
lP
na
ur
Jo

Fig .16 EBSD image quality maps with high-angle GBs of (a)-(c) cracking areas on the LCS side and EBSD
image quality maps with prior austenite GBs of (d)-(g) cracking areas on the Q&P980 steel side. Numbers
on the top right corner of each figure show the proportion of GBs (blue parts) in the whole figure.
Journal Pre-proof
Fig. 17 demonstrates the kernel average misorientation (KAM) maps of the cracking areas on both the
Q&P980 steel and LCS sides. It was reported that the KAM maps could measure the deformation-induced
local orientation gradients inside grains, they have a unique linear correlation with true plastic strain and can
be an appropriate parameter to evaluate local plastic deformation [9]. Cracking areas on the Q&P980 steel
side show higher KAM values than those on the LCS side. The high KAM in the Q&P980 steel was mainly
derived from the martensite transformation during cooling [9]. The relative deep cracks in the Q&P980 steel
(Fig. 17d and Fig. 17e) had a higher overall KAM than the relative shallow cracks (Fig. 17f and Fig. 17g),
indicating that more martensite formed around the deep cracks in the Q&P980 steel. As a result, the
microstructure around these deep cracks should be completely austenitized, and the temperature there during
welding was higher than that around the shallow cracks. Jung et al. [44] showed that plastic strain facilitated

of
LME cracking. The higher plastic strain in Q&P980 may have contributed to the propagation of LME
cracks.

ro
-p
re
lP
na
ur
Jo

Fig. 17 EBSD KAM maps of (a)-(c) cracking areas on the LCS side and (d)-(g) cracking areas on the
Q&P980 steel side.

Conclusions
In the present work, 1.4 mm-thick Q&P980 steel, 2.0 mm-thick LCS and 0.7 mm-thick LCS were
welded via RSW, and LME cracks that formed during RSW were characterized by various methods. Hot
tensile tests were also carried out on the Q&P980 steel and LCS to demonstrate their LME behaviour. The
Journal Pre-proof
main findings can be summarized as follows.
(1) During hot tensile test, under a strain rate of 0.1/s, the galvanized Q&P980 steel was embrittled via
LME at a temperature range from 700-950 ℃, and the embrittlement was extreme at approximately 800 ℃,
while the galvanized LCS was only slightly embrittled at approximately 800 ℃.
(2) In Q&P980/LCS/LCS dissimilar RSW joints, LME cracking was found on both the Q&P980 steel
and LCS sides, but many more cracks existed on the Q&P980 side. LME cracking was more severe under
the multiple-pulse welding schedule compared with the single-pulse welding schedule.
(3) A trace of liquid Zn in LME cracks at high temperatures was only found in some cracks, while some
cracks only possessed a small amount of Zn, which could only derive from diffusion behaviour. This
strongly indicated that LME cracks were formed via a stress-assisted diffusion mechanism.

of
(4) The LME cracks were mostly intergranular and were arrested when they transformed into
transgranular cracks, which indicated that GBs were the preferred path for cracks to propagate.

ro
(5) The LME susceptibility of the Q&P980 steel was higher than that for the LCS, which could be due
-p
to several reasons. First, the LCS was more reactive with liquid Zn than the Q&P980 steel, resulting in a
reduction in the liquid Zn at high temperatures. Second, a higher stress could be generated in the Q&P980
re
steel. Third, more LME-susceptible GBs were present in the Q&P980 steel.
lP

Data availability
na

The raw/processed data required to reproduce these findings cannot be shared at this time due to technical or
time limitations.
ur
Jo

CRediT authorship contribution statement


Zhanxiang Ling: Investigation, Methodology, Data curation, Writing original draft. Min Wang: Writing -
review & editing. Liang Kong: Funding acquisition, Writing - review & editing. Ke Chen: Writing - review
& editing.

Acknowledgements
This work was supported by the [National Key Research and Development Program of China] under Grant
[number 2017YFB0304403]; and the [National Natural Science Foundation of China] under Grant [number
51871154].

References
[1] Nicholas, M. G., & Old, C. F. (1979). Liquid metal embrittlement. Journal of Materials Science, 14(1),
Journal Pre-proof
1-18.
[2] Ling, Z., Wang, M., & Kong, L. (2018). Liquid metal embrittlement of galvanized steels during
industrial processing: A review. In Transactions on Intelligent Welding Manufacturing (pp. 25-42).
Springer, Singapore.
[3] Wintjes, E., DiGiovanni, C., He, L., Biro, E., & Zhou, N. Y. (2019). Quantifying the link between crack
distribution and resistance spot weld strength reduction in liquid metal embrittlement susceptible steels.
Welding in the World, 63(3), 807-814.
[4] DiGiovanni, C., Biro, E., & Zhou, N. Y. (2019). Impact of liquid metal embrittlement cracks on
resistance spot weld static strength. Science and Technology of Welding and Joining, 24(3), 218-224.
[5] DiGiovanni, C., Han, X., Powell, A., Biro, E., & Zhou, N. Y. (2019). Experimental and numerical

of
analysis of liquid metal embrittlement crack location. Journal of Materials Engineering and Performance,
28(4), 2045-2052.

ro
[6] Bhattacharya, D. (2018). Liquid metal embrittlement during resistance spot welding of Zn-coated
-p
high-strength steels. Materials Science and Technology, 34(15), 1809-1829.
[7] Ashiri, R., Haque, M. A., Ji, C. W., Salimijazi, H. R., & Park, Y. D. (2015). Supercritical area and
re
critical nugget diameter for liquid metal embrittlement of Zn-coated twining induced plasticity steels.
lP

Scripta Materialia, 109, 6-10.


[8] Tolf, E., Hedegård, J., & Melander, A. (2013). Surface breaking cracks in resistance spot welds of dual
na

phase steels with electrogalvanised and hot dip zinc coating. Science and technology of welding and
joining, 18(1), 25-31.
ur

[9] Kalashami, A. G., DiGiovanni, C., Razmpoosh, M. H., Goodwin, F., & Zhou, N. Y. (2020). The Role of
Jo

Internal Oxides on the Liquid Metal Embrittlement Cracking During Resistance Spot Welding of the
Dual Phase Steel. Metallurgical and Materials Transactions A, 1-12.
[10] Kim, Y. G., Kim, I. J., Kim, J. S., Chung, Y. I., & Du, Y. C. (2014). Evaluation of surface crack in
resistance spot welds of Zn-coated steel. Materials Transactions, M2013244.
[11] Ashiri, R., Mostaan, H., & Park, Y. D. (2018). A Phenomenological Study of Weld Discontinuities and
Defects in Resistance Spot Welding of Advanced High Strength TRIP Steel. Metallurgical and Materials
Transactions A, 49(12), 6161-6172.
[12] Choi, D. Y., Sharma, A., Uhm, S. H., & Jung, J. P. (2019). Liquid metal embrittlement of resistance spot
welded 1180 TRIP steel: effect of electrode force on cracking behavior. Metals and Materials
International, 25(1), 219-228.
[13] He, L., DiGiovanni, C., Han, X., Mehling, C., Wintjes, E., Biro, E., & Zhou, N. Y. (2019). Suppression
of liquid metal embrittlement in resistance spot welding of TRIP steel. Science and Technology of
Journal Pre-proof
Welding and Joining, 24(6), 579-586.
[14] Frei, J., & Rethmeier, M. (2018). Susceptibility of electrolytically galvanized dual-phase steel sheets to
liquid metal embrittlement during resistance spot welding. Welding in the World, 62(5), 1031-1037.
[15] Frei, J., Biegler, M., Rethmeier, M., Böhne, C., & Meschut, G. (2019). Investigation of liquid metal
embrittlement of dual phase steel joints by electro-thermomechanical spot-welding simulation. Science
and Technology of Welding and Joining, 24(7), 624-633.
[16] Ashiri, R., Shamanian, M., Salimijazi, H. R., Haque, M. A., Bae, J. H., Ji, C. W., ... & Park, Y. D. (2016).
Liquid metal embrittlement-free welds of Zn-coated twinning induced plasticity steels. Scripta
Materialia, 114, 41-47.
[17] Lee, H., Jo, M. C., Sohn, S. S., Kim, S. H., Song, T., Kim, S. K., ... & Lee, S. (2019). Microstructural

of
evolution of liquid metal embrittlement in resistance-spot-welded galvanized TWinning-Induced
Plasticity (TWIP) steel sheets. Materials Characterization, 147, 233-241.

ro
[18] Ling, Z., Chen, T., Kong, L., Wang, M., Pan, H., & Lei, M. (2019). Liquid Metal Embrittlement
-p
Cracking During Resistance Spot Welding of Galvanized Q&P980 Steel. Metallurgical and Materials
Transactions A, 50(11), 5128-5142.
re
[19] Béal, C. (2011). Mechanical behaviour of a new automotive high manganese TWIP steel in the presence
lP

of liquid zinc (Doctoral dissertation, Lyon, INSA).


[20] Sigler, D. R., Schroth, J. G., Gayden, X. Q., Yang, W., Jiang, C., Sang, Y., & Morin, P. J. (2008, May).
na

Observations of liquid metal assisted cracking in resistance spot welds of zinc-coated advanced high
strength steels. In Sheet Metal Welding Conference Sheet Metal Welding Conf. XIII.
ur

[21] Jung, G., Woo, I. S., Suh, D. W., & Kim, S. J. (2016). Liquid Zn assisted embrittlement of advanced
Jo

high strength steels with different microstructures. Metals and Materials International, 22(2), 187-195.
[22] Tumuluru, M. (2019). Effect of Silicon and Retained Austenite on the Liquid Metal Embrittlement
Cracking Behavior of GEN3 and High-Strength Automotive Steels. Welding Journal, 98(12),
351S-364S.
[23] Hong, S. H., Kang, J. H., Kim, D., & Kim, S. J. (2020). Si effect on Zn-assisted liquid metal
embrittlement in Zn-coated TWIP steels: Importance of Fe-Zn alloying reaction. Surface and Coatings
Technology, 125809.
[24] Fernandes, P. J. L., & Jones, D. R. H. (1997). Mechanisms of liquid metal induced embrittlement.
International materials reviews, 42(6), 251-261.
[25] Kang, H., Cho, L., Lee, C., & De Cooman, B. C. (2016). Zn penetration in liquid metal embrittled TWIP
steel. Metallurgical and Materials Transactions A, 47(6), 2885-2905.
[26] Razmpoosh, M. H., Macwan, A., Biro, E., Chen, D. L., Peng, Y., Goodwin, F., & Zhou, Y. (2018).
Journal Pre-proof
Liquid metal embrittlement in laser beam welding of Zn-coated 22MnB5 steel. Materials & Design, 155,
375-383.
[27] Wintjes, E., DiGiovanni, C., He, L., Bag, S., Goodwin, F., Biro, E., & Zhou, Y. (2019). Effect of
multiple pulse resistance spot welding schedules on liquid metal embrittlement severity. Journal of
Manufacturing Science and Engineering, 141(10). 101001.
[28] Benlatreche Y., Dupuy T., Ghassemi-armaki H., & Lucchini L. (2019) Methodology for Liquid Metal
Embrittlement (LME) evaluation of coated steels during spot welding. In: The 72nd annual assembly of
the International Institute of Welding (IIW). Bratislava, III-1959-19M.
[29] COMBINED REPORTS – AHSS IMPLEMENTATION SOLUTIONS: LIQUID METAL
EMBRITTLEMENT STUDY, WorldAutoSteel, 2020,

of
https://www.worldautosteel.org/projects/liquid-metal-embrittlement/.
[30] APPROVAL PROCEDURE FOR THE STEEL LINES WELDABILITY IN SPOT WELDING,

ro
MXP_CEB06_0118, PSA PEUGEOT CITROEN, 2013. -p
[31] Beal, C., Kleber, X., Fabregue, D., & Bouzekri, M. (2011). Liquid zinc embrittlement of a
high-manganese-content TWIP steel. Philosophical magazine letters, 91(4), 297-303.
re
[32] Beal, C., Kleber, X., Fabregue, D., & Bouzekri, M. (2012). Liquid zinc embrittlement of
lP

twinning-induced plasticity steel. Scripta Materialia, 66(12), 1030-1033.


[33] Beal, C., Kleber, X., Fabregue, D., & Bouzekri, M. (2012). Embrittlement of a zinc coated high
na

manganese TWIP steel. Materials Science and Engineering: A, 543, 76-83.


[34] Razmpoosh, M. H., Biro, E., Chen, D. L., Goodwin, F., & Zhou, Y. (2018). Liquid metal embrittlement
ur

in laser lap joining of TWIP and medium-manganese TRIP steel: The role of stress and grain boundaries.
Jo

Materials Characterization, 145, 627-633.


[35] Kim, D., Kang, J. H., & Kim, S. J. (2018). Heating rate effect on liquid Zn-assisted embrittlement of
high Mn austenitic steel. Surface and Coatings Technology, 347, 157-163.
[36] Choi, D. Y., Uhm, S. H., Enloe, C. M., Lee, H., Kim, G., & Horvath, C. (2017). Liquid metal
embrittlement of resistance spot welded 1180TRIP steel-effects of crack geometry on weld mechanical
performance. Mater Sci Technol, 454-462.
[37] Gordon, P., & An, H. H. (1982). The mechanisms of crack initiation and crack propagation in
metal-induced embrittlement of metals. Metallurgical Transactions A, 13(3), 457-472.
[38] Wan, Z., Wang, H. P., Wang, M., Carlson, B. E., & Sigler, D. R. (2016). Numerical simulation of
resistance spot welding of Al to zinc-coated steel with improved representation of contact interactions.
International Journal of Heat and Mass Transfer, 101, 749-763.
[39] Ling, Z., Chen, T., Wang, M., & Kong, L. (2020). Reducing liquid metal embrittlement cracking in
Journal Pre-proof
resistance spot welding of Q&P980 steel. Materials and Manufacturing Processes, 1-8.
[40] Bhattacharya, D., Cho, L., Van der Aa, E., Ghassemi-Armaki, H., Pichler, A., Findley, K. O., & Speer, J.
G. (2020). Transgranular cracking in a liquid Zn embrittled high strength steel. Scripta Materialia, 175,
49-54.
[41] Lin, H., & Pope, D. P. (1993). The influence of grain boundary geometry on intergranular crack
propagation in Ni3Al. Acta metallurgica et materialia, 41(2), 553-562.
[42] Razmpoosh, M. H., Macwan, A., Goodwin, F., Biro, E., & Zhou, Y. (2020). Crystallographic study of
liquid-metal-embrittlement crack path. Materials Letters, 267, 127511.
[43] Razmpoosh, M. H., Macwan, A., Goodwin, F., Biro, E., & Zhou, Y. (2020). Role of Random and
Coincidence Site Lattice Grain Boundaries in Liquid Metal Embrittlement of Iron (FCC)-Zn Couple.

of
Metallurgical and Materials Transactions A, 1-7.
[44] Jung, G., Woo, I. S., Suh, D. W., & Kim, S. J. (2016). Liquid Zn assisted embrittlement of advanced

ro
high strength steels with different microstructures. Metals and Materials International, 22(2), 187-195.
-p
re
lP
na
ur
Jo
Journal Pre-proof
CRediT authorship contribution statement
Zhanxiang Ling: Investigation, Methodology, Data curation, Writing original draft. Min Wang: Writing -
review & editing. Liang Kong: Funding acquisition, Writing - review & editing. Ke Chen: Writing - review
& editing.

of
ro
-p
re
lP
na
ur
Jo
Journal Pre-proof
 Liquid metal embrittlement cracking occurred in dissimilar resistance spot welding of

galvanized Q&P980 steel and galvanized low carbon steel.

 Q&P980 steel is more susceptible to liquid metal embrittlement cracking than low

carbon steel.

 Stress-assisted diffusion is the mechanism of liquid metal embrittlement in resistance

spot welding rather than liquid penetration.

of
ro
-p
re
lP
na
ur
Jo

You might also like