You are on page 1of 9

Fusion Engineering and Design 195 (2023) 113973

Contents lists available at ScienceDirect

Fusion Engineering and Design


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

Investigation on microstructure and mechanical properties of 160mm thick


316L stainless steel electron-beam-welded joint
Jianguo Ma a, c, d, Zhiyong Wang a, b, Zhihong Liu a, c, *, Haibiao Ji a, c, Xiaowei Xia b,
Chengwen Li a, b, Zhenfei Liu a, c, Zhongtao Zhang a, e, Jiefeng Wu a, c
a
Institute of Plasma Physics, Hefei Institutes of Physical Science, Chinese Academy of Sciences, Hefei 230031, China
b
University of Science and Technology of China, Hefei 230026, China
c
Anhui Province Key Laboratory of Special Welding Technology, Huainan 232063, China
d
Institute of Energy, Hefei Comprehensive National Science Center, Hefei 230031, China
e
Institutes of Physical Science and Information Technology, Anhui University, Hefei 230039, China

A R T I C L E I N F O A B S T R A C T

Keywords: The maximum thickness of flanging position on the vacuum vessel (VV) window collar of China Fusion Engi­
Electron beam welding neering Test Reactor (CFETR) is 160 mm. Electron beam welding can achieve one-time penetration of the thick
Microstructure plate without groove and the deformation after welding is quite small. The cross section of weld was sampled for
Mechanical properties
microscopic observation and mechanical properties tests to explore the difference in microstructure and per­
Solidification mode
Composition overcooling
formances of 160 mm thick 316 L stainless steel electron-beam-welded joints. The results showed that the weld
depth-width ratio reached up to 16:1, and the microstructure of weld was composed of ferrite and austenite with
different morphologies which was mainly coarse cellular crystal in heat affected zone (HAZ). The equiaxed grain
orientation in the root region (denoted as “R region”) was mostly [001] and the columnar grain orientation along
the temperature gradient was mostly [101]. The microstructure along the depth direction was quite different,
which was mainly related to the solidification mode and composition overcooling. The high strength
(YS=715.22 MPa) of R region was due to the smaller temperature gradient and the larger cooling rate, the
position of fracture suggested that the strength of weld was higher than that of the base metal. The micro­
hardness measurement results showed that the hardness increased gradually along the depth direction of weld
(Top:156.5 HV, Root: 238.7 HV) and fluctuated along the width direction. The tensile properties along the
thickness direction exhibited a trend of decreasing first and then increasing, which was mainly related to the
characteristics of electron beam deep penetration welding and the difference of grain structure along the
thickness direction.

1. Introduction replacing internal components and installing auxiliary heating and. etc.
As a connecting part between windows and the main body of VV, the
As a core safe component of China Fusion Engineering Test Reactor window collar has the characteristics of complex contour and large
(CFETR), the vacuum vessel (VV) provides a stable vacuum environment thickness as shown in Fig. 1(a). The flanging is a key position where
for plasma fusion reaction[1–3]. The VV serves in a complex connects the main body of VV, thus the control of welding deformation
mechanical-temperature-electromagnetic coupling environment, is more stringent. The maximum size of the window collar of the CFETR
requiring all welds to achieve full penetration. Besides, the arc welding VV is about 4.2 m × 2.7 m.
defect level realizes the ISO5817: 2014-B quality standard and the The vacuum EBW can realize the micro-deformation welding of large
electron beam welding (EBW) defect level realizes the ISO13919–1: thickness structure with the advantages of small post-welding defor­
2019-B quality standard. The VV has a complex hyperboloid structure mation, narrow heat affected zone (HAZ) and large aspect ratio. How­
with the cross section of double-layer D-shaped, and contains an upper ever, the vacuum chamber and welding motion system of electron beam
window, a middle window and a lower window for diagnostic systems, welding machine is unable to weld the large window collar parts

* Corresponding author.
E-mail address: zhliu@ipp.ac.cn (Z. Liu).

https://doi.org/10.1016/j.fusengdes.2023.113973
Received 10 July 2023; Received in revised form 23 August 2023; Accepted 23 August 2023
Available online 30 August 2023
0920-3796/© 2023 Elsevier B.V. All rights reserved.
J. Ma et al. Fusion Engineering and Design 195 (2023) 113973

Welding tests were carried out by the ZD150–60C–CV66M vacuum


electron beam welding machine with working vacuum of about 1.7 ×
10− 4 mbar, accelerating voltage of 150 kV and the detailed welding
parameters were listed in Table 1. To remove the surface oxide, oil and
water stains and reduce the production of welding defects, the 316 L
plates were subjected to mechanical polishing and alcohol washing
before welding. The arc ignition plate and arc retract plate with the size
of 60×60×160 mm were added in the front and end of the welding
plates for cutting the crater easily and the 360×20×15 mm size of pallet
was added to ensure full penetration. The scanning deflection with
Fig. 1. Vacuum vessel window collar welding tool.
rectangle shape was added during the EBW process, which was shown in
Fig. 2.
directly, and the collar can only move in plane but not rotate. In order to
The hardness measurement, tensile tests and microstructure obser­
satisfy the full penetration of the flanging part, the back of the weld must
vation were carried out from five sampling positions: top, upper, middle,
be added with a pallet, and the maximum penetration depth is about
lower and root, which were marked as T, U, M, L and R regions, as shown
160 mm, as shown in Fig. 1(b).
in Fig. 3. The distribution of microhardness in the weld cross section was
The EBW of large-thick plates is a deep penetration welding mode
measured by DHV-1000 Z micro Vickers hardness tester, with a load of
along with keyhole effects. Microstructure inhomogeneity which affects
4.9 N and a hosting time of 15 s. Tensile tests at room temperature were
the mechanical properties often occurs along the thickness direction of
carried out on MTS 809 universal tensile testing machine at a strain rate
joint. There are few reports and studies on the microstructure and
of 2 × 10− 3 s− 1, and the size of specimen was shown in Fig. 4. The
properties inhomogeneity of EBW joint. Bhanu et al. [4] prepared 8 mm
electron backscattered diffraction (EBSD) specimens were finely ground
thick EBW dissimilar joints between P91 ferritic / martensitic steel and
to 5000 # sandpaper, and 0.03 μm silica suspension was used for me­
Incoloy 800HT nickel-based alloy and characterized the microstructure
chanical polishing. The HAZ and weld were scanned at a 3 μm step size
of weld involving nickel-based alloy. The results showed that the
using an OXFORS C–NANO electron backscatter diffraction system and
microstructure of weld center exhibited a mixed solidification mode
a SIGMA300 field emission scanning electron microscope after slight
with equiaxed dendrites and cellular structures. Maurya et al. [5]
corrosion by aqua regia solution (75% HCl + 25% HNO3).
studied the microstructure, mechanical properties and corrosion
behavior of dissimilar welded joints. There were significant differences
3. Results and discussion
in the microstructure of the dissimilar joints at the weld and interface,
the microstructure inhomogeneity had a significant effect on the me­
3.1. Microstructure of weld
chanical properties of the welded joints, including microhardness, ten­
sile and impact strength. Wang et al. [6] carried out the EBW of 40 mm
3.1.1. Macroscopic morphology characteristics
thick 316 L and analyzed the microstructure and properties after
Fig. 3 showed the macroscopic morphology in the weld section of the
welding. The study found that the weld microstructures were mainly
columnar crystal, fine dendrite and equiaxed crystal. The hardness in the
middle of the weld was the highest, the strength of the weld was obvi­
ously better than that of the base metal. Chen et al. [7] welded 50 mm
thick 304 stainless steel plate at one time via EBW. The solidification
mode varied with the solidification rates which were diverse in various
depths of the weld, and the grains gradually decreased from top to root.
Elmer et al. [8] studied the effects of different alloy compositions and
cooling rates on the solidification mode and microstructure of the weld.
The alloy only solidified in single-phase ferrite or single-phase austenite
at a high cooling rate, the microstructure changed with the partial solid
transformation of ferrite to austenite. Through previous work, it was
found that the thickness of materials they studied was generally no more
than 100 mm, which was because too large thickness of the plate
increased the difficulty of welding and the forming quality was difficult
to guarantee. However, two 160 mm thick 316 L stainless steel plates
were well welded via vacuum electron beam welding in this paper. The
weld was sampled from top to root for observing the macro/micro
morphology and testing the performances. The change law of micro­
structure and properties was analyzed, which provided theoretical
guidance for the welding of VV window collar applied to CFETR.

2. Experimental procedure

Two 316 L stainless steel plates (ASTM A240/A240M-16a) with the


size of 300×25×160 mm were selected as the experimental material. Fig. 2. Electron beam welding process.

Table 1
Electron beam welding process parameters.
Voltage Ua/ Beam current Ib/ Focus current If/ Velocity v/(mm•s −
Working distance L/ Scanning Amplitude A/ Oscillation frequency fp/
1
kV mA mA ) mm shape mm Hz

150 390 2300 5 400 Rectangle 3 100

2
J. Ma et al. Fusion Engineering and Design 195 (2023) 113973

Table 3
Solidification mode of austenitic stainless steel.
Solidification mode Solidification structure Creq / Nieq

A L→L+γ→γ <1.25
AF L→L+γ→L+γ+δ→γ+δ 1.25~1.48
FA L→L+δ→L+γ+δ→δ+γ 1.48~1.95
F L→L+δ→δ→δ+γ >1.95

Fig. 3. The sampling position of weld for tests.

Fig. 4. The size of the tensile samples.

160 mm thick 316 L electron-beam-welded joint, which obviously


clarified the large weld depth-width ratio of about 16:1 and the weld
reinforcement of about 3 mm. The weld width varied greatly along the
thickness direction from top to root, where the maximum was about 10
mm (T region) and minimum was about 3 mm (R region). There were no
cracks, nail tips and other defects found by phased array ultrasonic
testing (UT) after welding, which proved the welded joint matched the
ISO13919–1: 2019 B quality requirements.
Fig. 5. Pseudo-binary phase diagram of Fe-Cr-Ni.
3.1.2. Solidification modes of weld along the thickness direction
The chemical composition of 316 L measured by direct reading austenite precipitation in the liquid phase. The remaining liquid phase
spectrometer was shown in Table 2, which was directly related to the performed a complete transformation into γ austenite, either directly or
phase transformation of austenitic steel during solidification process. the through peritectic reaction with δ-ferrite [11]. Subsequently, a solid
solidification mode could be divided into four types i.e. austenite (A), phase transition took place, where δ-ferrite transformed into γ austenite,
austenite-ferrite (AF), ferrite-austenite (FA) and ferrite (F) according to and only a small amount of δ-ferrite remained in the austenite matrix.
the range of Creq/Nieq as shown in Table 3 [9]. Different solidification This phenomenon resulted in the segregation of rich Cr and poor Ni in
modes was corresponding to different solidification structures. The the solidified structure.
calculation formula of chromium-nickel equivalent proposed by Ham­ Fig. 6 illustrated the microstructure evolution of the weld along the
mar-Svensson’s is more accurate to predict the solidification mode of thickness direction, with observations taken from five different regions.
austenitic stainless steels which contain N and Cu [10]: The microstructure of the T region at the top of weld was shown in Fig. 6
(a), which consists of coarse austenite and reticular or acicular δ-ferrite
Creq = Cr + 1.37Mo + 1.5Si + 2Nb + 3Ti (1)
in the matrix. The formation of the ferrite-austenite microstructure may
be attributed to the non-equilibrium solidification mode during the
Nieq = Ni + 0.31Mn + 22C + 14.2N + Cu (2)
electron beam welding process, resulting in incomplete ferrite-austenite
The Creq/Nieq ratio of the experimental 316 L stainless steels was transformation. The U region was mainly composed of austenite and
determined to be 1.68 by calculation. It was evident that the ferrite in skeleton-like ferrite, as shown in Fig. 6(b). The M region structure was
weld was precipitated as initial phase under equilibrium conditions with composed of columnar austenite and skeletal/dendritic ferrite as shown
FA solidification mode according to the pseudo-binary phase diagram of in Fig. 6(c). Fig. 6(d) displayed the microstructure of L region, where
Fe-Cr-Ni (Fig. 5). However, the solidification mode may change with different ferritic morphologies (mesh-like, dendrite-like) and austenitic
cooling rate because of the rapid heating and cooling involved in elec­ structures coexisted, indicating a significant change in the solidification
tron beam welding process. The root of weld experienced the fastest rate of L region. The dendrite spacing of ferrite decreased significantly
cooling rate, leading to the direct precipitation of metastable austenite with the change of solidification mode. It was reported that the dendrite
as the primary phase, replacing stable ferrite [11]. arm spacing (DAS) in the electron-beam-welded joint was related to the
The δ-ferrite precipitated first during the FA mode, which led to the cooling rate (CR) [12,13].
enrichment and consumption of ferritic forming elements, such as Cr
DAS = 80(CR)− 0.33
(3)
and Mo, thus promoted the nucleation and growth of ferrite. Simulta­
neously, austenitic forming elements (such as Ni and Mn) were excluded The DAS decreased with the weld depth increased since the lowest
into the liquid phase, creating favorable conditions for the formation of CR in R region and the highest in T region based on the above equation,

Table 2
Composition of 316L(wt.%).
C Si Mn P S Ni Cr Mo Cu N

0.022 0.50 1.36 0.032 0.004 10.12 16.31 2.04 0.3 0.037

3
J. Ma et al. Fusion Engineering and Design 195 (2023) 113973

Fig. 6. Microstructure of weld center in different depth: (a) T region; (b) U region; (c) M region; (d) L region; (e) R region; (f) Pore in L region.

which explained that the R region comprised primarily of columnar and


fine equiaxed crystal structures. Differences in grain morphology of
different regions primarily was because of the alloy component of weld
segregation during solidification process, which resulted in variations in
composition overcooling along the depth direction. The criterion of
composition overcooling can be expressed by the following formula
[14].
GL mC0 1 − k0
= (4)
R DL k0

where GL-liquid temperature gradient, R-crystal growth rate, mL-liquid


line slope, C0-original composition concentration, DL-solute diffusion
coefficient in liquid phase, K0-solute equilibrium distribution coeffi­
cient. For alloy systems with defined composition, the composition
overcooling is mainly influenced by the temperature gradient (G) and
solidification rate at the liquid-solid interface front during solidification.
Smaller G/R value is corresponding to greater degrees of composition
overcooling, the grain growth morphology transitions from planar to
Fig. 7. Keyhole effect.
equiaxed as the overcooling tendency increases, successively manifest­
ing as planar, cellular, columnar, dendrite, and equiaxed gradually. The
highest peak heating temperature occurred in T region, which yielded the weld upon solidification. The primary bubbles of pores with large
the largest G in the liquid phase of the liquid-solid interface front and a size and uniform shape exhibited an equilibrium state in the gas-liquid
quit narrow overcooling zone. The increasing composition overcooling phase of the weld pool that satisfied both thermal and mechanical
introduced changes in the ferrite morphology, progressing towards equilibrium. During the vacuum electron beam welding process, the
dendritic and equiaxed crystals. A solid phase transformation occurred heat source of electron beam acted on the keyhole directly, generating
when the temperature reduced below the solidus line, which separated high wall surface superheat. The keyhole size was related to the gas
austenite forming elements towards the austenite side and ferrite density, surface tension, and superheat of the gas phase in the primary
forming elements towards the ferritic side. The dendrite center richen in bubble [15].
Cr remained untransformed into austenite and formed skeleton-like
ferrite. With a further reduction of G, the weld center composition un­ 3.1.3. Microstructural analysis of weld along the thickness direction
derwent significant overcooling, which facilitated the formation of new Fig. 8 illustrated the grain orientation distribution at various depths
crystal nucleus. Consequently, the microstructure exhibited fine equi­ within the cross-section of the welded joint. The Inverse Pole
axed ferrite, and the solidification mode changed from FA to AF. Figure (IPF) described the changes in grain size, orientation, and growth
A porosity with a diameter of approximately 48 μm in L region was morphology along the width and thickness direction of weld. The weld
evidently observed from Fig. 6(d) and (f), which may arise from the microstructure was primarily composed of axial dendrites in the center
keyhole effect caused by the rapid evaporation of low melting point and gradient-distributed columnar crystals. The columnar grains grew
elements in 316 L under the impact of high-density electron beam, as from both sides of the weld pool perpendicular to the weld boundary,
demonstrated in Fig. 7. The metal surrounding the beam in the molten and axial grains grew along the depth of the weld. The orientation of
pool was in superheated state through the action of forced convection. grains in the weld was primarily dominated by [001] and [100] di­
The metal vapor couldn’t escape through buoyancy and formed pores in rections. R region exhibited relatively uniform orientation with mostly

4
J. Ma et al. Fusion Engineering and Design 195 (2023) 113973

Fig. 8. IPF of different sampling regions: (a)T-region; (b)U-region; (c)M-region; (d)L-region; (e)R-region.

[100] orientation of equiaxed crystal in the center, [001] orientation of with a coarse grain size of up to 200 μm displayed cellular and columnar
columnar crystal on both sides, and the dendritic ferrite dispersed in the morphology as shown in Fig. 8(b) and (c). It was clarified that the grain
austenite matrix. The weld microstructure demonstrated substantial size of T and U regions was significantly larger than that of L and R
inhomogeneity from T region to R region. The top microstructure of region, which suggested that coarse grain size mainly arose due to the
weld displayed a composition of axial coarse grain and columnar grain, high energy density, low cooling rate of electron beam welding and long
while the root grain structure transformed into columnar and equiaxed residence time at high temperature. The coarsening of microstructure
fine crystal. It was believed that the faster cooling rate contributed to resulted in a sharp compromise of performances.
form the fine equiaxed crystal structure. The HAZ near the fusion line The size of grains mainly depends on the ratio of nucleation rate N to

5
J. Ma et al. Fusion Engineering and Design 195 (2023) 113973

growth rate G. It was observed that the grain size decreased with the
ratio of N/G increased as shown in Fig. 9, both N and G were closely
related to the overcooling degree under unbalanced solidification con­
ditions [16]. However, N and G decreased when the overcooling degree
increased to a certain value, increasing the degree of overcooling
moderately can refine the grain. Scanning deflection is often added to
produce stirring effect in industry, so as to increase the degree of over­
cooling, improve the weld morphology and enhance the mechanical
properties of electron-beam-welded joints.
Affected by the welding thermal cycle and thermal exchange, the
energy received by the HAZ from the near base metal side to the near
weld side enhanced gradually, and the microstructure showed the
characteristics of continuous transition. The HAZ showed a high auste­
nitizing temperature and large grain sizes near the weld, while the HAZ
near the base metal underwent grain refinement after recrystallization.
Therefore, there were significant differences in grain orientation and
morphology between both sides near the fusion line where represented
the ideal location for forming a continuous liquid film with low melting
point impurities and brittle phases which usually caused mechanical
property degradation. In addition, HAZ also supply a favorable site for Fig. 10. Hardness distribution along thickness direction.
residual stress concentration and crack propagation. Hence, the HAZ is
deemed as the weakest area of weld [17]. microstructure. From T region to M region of the weld, grains changed
from coarse dendrites to fine dendrites, the grain size became smaller
and the hardness increased. From M region to R region, the change of
3.2. Mechanical properties of weld along the thickness direction solidification mode led to the decrease of ferrite quantity, and the
refinement of grains led to the increase of grain boundary quantity,
3.2.1. Microhardness analysis which resulted in the increase of hardness.
Microhardness is a critical parameter for evaluating the quality of Fig. 11 depicted the microhardness variations taken from R region
welded joints, which depends on not only the material but also the along the width direction of the weld. Hardness levels displayed a
microstructure and welding defects. The DHV-1000 Vickers hardness decreasing trend from the weld center to base metal. The microhardness
tester was subjected to measure the cross-sectional hardness of the of the weld center was 15–20 HV higher than that of the base metal. The
welded joint along the thickness and width directions. Fig. 10 presented equiaxed dendrite distribution range of the weld center was believed to
the outcomes of the weld center in T, U, M, L, and R regions. The be the highest hardness area of the welded joint. The main reason why
hardness exhibited an increasing trend from the top to the root, which the hardness near the fusion line not much different from that of the base
ranged between 150 and 160HV in T region, 200–220HV in U and M metal was that the grains near the fusion line were slender columnar
regions, and 220–230HV in L region. The largest hardness of R region crystals, which were coarser than the weld center, which reduced the
increased to 242.5HV, which exhibited 56% higher than that of T region hardness.
(156.5HV). Grains in the top of weld were coarse because of the most
significant welding heat input. The coarse network ferrite precipitated 3.2.2. Tensile properties analysis
during the solidification process, leading to a notable hardness reduction Tensile strength is another essential parameter to evaluate the me­
in T region when compared to base metal (~217HV). From T region to M chanical properties of electron-beam-weld joints. The test results of the
region, the decreasing temperature gradient (G) resulted in the trans­ tensile specimens in different regions showed that the tensile strength
formation of coarse reticular crystals to fine dendrites and equiaxed from T region to R region along the thickness direction of the weld was
crystals. The solidification mode changed from FA to AF along the highly inhomogeneous as shown in Fig. 12. The tensile properties along
thickness direction with the decreasing ferrite content. The composition
overcooling inflicted changes from dendritic crystal to equiaxed crystal,
resulting in further grain refinement and increased hardness. In sum­
mary, the variation of hardness was closely related to the

Fig. 9. Relationships between N, G and △T. Fig. 11. Hardness distribution along width direction (R region).

6
J. Ma et al. Fusion Engineering and Design 195 (2023) 113973

Fig. 12. Engineering stress-axial displacement curve of different regions.

with fracture location were listed in Table 4. It was observed that the
fracture position of all specimens was located at base metal as shown in
Fig. 13, which indicated that the weld strength was higher than that of
the base metal.
The ultimate tensile strength (UTS) from the T region to the R region
specimens along the thickness direction of the weld exhibited a high
inhomogeneity. The T region and R region specimens showed high
strength with an average UTS exceeded 700 MPa and the U region and L
region was about 688 MPa, while the M region exhibited the lowest. This
distribution law was mainly related to the microstructure and weld
surface characteristics. The weld grains in the T region were mostly
equiaxed fine dendrites, which were greatly affected by fine grain
strengthening and thus kept high strength. The grains in T region were
coarse columnar crystals which resulted in a low hardness. However,
due to the characteristics of electron beam deep penetration welding,
the top of weld was subjected to a high heat input which caused the large
weld width, and the protective effect of weld on the joint was the
strongest [18]. The tensile properties reflected the average mechanical
properties of a larger area, while the microhardness represented the
performance changes in the small local area. Hence, the T region
exhibited the high strength-low hardness feature, and the high strength
of the U region also came from the protection of the weld. Compared
with the L region, the difference in performance between the M region
and L region mainly came from the transformation of the microstruc­
ture. It can be observed from the previous content that the trans­
formation of the microstructure was mainly related to the composition
overcooling determined by the cooling rate (R) and the temperature

Table 4
Tensile properties along thickness direction.
Region UTS / MPa Fracture location

T 708.62 Base metal Fig. 13. Tensile specimen after fracturing.


U 688.64 Base metal
M 677.49 Base metal
L 687.05 Base metal
R 715.22 Base metal

7
J. Ma et al. Fusion Engineering and Design 195 (2023) 113973

gradient (G). Fusion Technology Program of China [grant number 2018-000052-73-


01-001228], the Collaborative Innovation Program of Hefei Science
4. Conclusion Center, CAS [grant number 2022HSC-CIP025], Youth Innovation Pro­
motion Association CAS [grant number 2019433] and the National
In this paper, the microstructure and mechanical properties of 160 Nature Science Foundation of China [grant number 12105185]. This
mm thick 316 L electron-beam-welded joint were investigated in detail. work was also supported by the Experimental Advanced Super­
The following conclusions can be drawn: conducting Tokamak device and Anhui Province Key Laboratory of
Special Welding Technology.
1 The optimized welding process parameters were selected to achieve
welding through two 160 mm thick 316 L stainless steel plates, References
resulting in a weld with large depth-to-width ratio and no discernible
surface cracks and nail tip defects. [1] Y. Song, S. Wu, J. Li, et al., Concept design of CFETR tokamak machine, IEEE Trans.
2 The temperature gradient and cooling rate along the depth direction Plasma Sci. 42 (2014) 503–509. https://ieeexplore.ieee.org/document/6716984.
[2] H. Ji, J. Wu, H. Wu, et al., Analysis of welding deformation on CFETR 1/32 vacuum
of weld were different, morphing the solidification mode from FA to vessel mockup, Fusion Eng. Des. 144 (2019) 160–163. https://www.sciencedirect.
AF mode eventually. The weld structure transformed from coarse com/science/article/pii/S0920379619306520?via%3Dihub.
mesh-like ferrite in T region to dendrites in M and L regions and [3] L. Xiu, J. Wu, Z. Liu, et al., Welding distortion control technology in CFETR vacuum
vessel, Fusion Eng. Des. 160 (2020), 111851. https://www.sciencedirect.com/scie
finally culminated in fine equiaxed crystals in R region with uniform nce/article/pii/S0920379620303999?via%3Dihub.
grain orientation. [4] V. Bhanu, A. Malakar, A. Gupta, et al., Electron beam welding of P91 steel and
3 The mechanical properties of the weld cross-section displayed in­ incoloy 800HT and their microstructural studies for advanced ultra super critical
(AUSC) power plants, Inter J. Press Vessel Piping 205 (2003), 105010, https://doi.
homogeneity along the thickness direction, which mainly depended org/10.1016/j.ijpvp.2023.105010.
on the changing microstructure. The shape of grains was regulated [5] A.K. Maurya, S.M. Pandey, R. Chhibber, et al., Structure–property relationships
by the temperature gradient and grain growth rate. As the cooling and corrosion behavior of laser-welded X-70/UNS S32750 dissimilar joint, Archiv.
Civ. Mech. Eng 23 (2023) 81, https://doi.org/10.1007/s43452-023-00627-5.
speed accelerated along the thickness direction, the overcooling
[6] Z. Wang, D. Wang, Z. Wen, et al. Study on microstructure and mechanical property
degree increased and led to continuous refinement of the grain and of electron beam welding joint of 40mm-thickness stainless steel. Dong. Turb.,
enhanced hardness. The tensile test indicated that the strength of the 2020, (01): 22–27. https://kns.cnki.net/kcms2/article/abstract?v=3uoqIhG8
C44YLTlOAiTRKibYlV5Vjs7i8oRR1PAr7RxjuAJk4dHXotVREiQgwjkaAF084H33f
weld surpassed that of the base metal. Moreover, the upper of the
q-YM70zBle3NjDCIafreahL&uniplatform=NZKPT.
weld featured a large melting width which offered a robust protec­ [7] Q. Chen, D. Li, C. He, et al., Microstructure difference analysis of large thickness
tive effect on the joint and thus increased the strength. welded joint with EBW, Trans. Chi Weld Inst. 36 (09) (2015) 79–82. +117, https:
//kns.cnki.net/kcms2/article/abstract?v=3uoqIhG8C44YLTlOAiTRKibYlV5Vjs7ir
5D84hng_y4D11vwp0rrtQ7GwkwasGWUd_Uqj6t9DnpAwjCh2zjUh
In this paper, EBW was used to realize the one-time penetration of nE-V_IpHUEu&uniplatform=NZKPT.
160 mm ultra-thick plate and good forming quality. The microstructure [8] J. Elmer, S. Allen, T. Eagar, Microstructural development during solidification of
of the weld was analyzed by means like OM, EBSD and other material stainless-steels alloys, Metal. Trans. 20 (10) (1989) 2117–2131. https://link.
springer.com/article/10.1007/BF02650298.
characterization methods, and the relationship between microstructure [9] J. Ma, Y. Yang, W. Tong, et al., Microstructural evolution in AISI 304 stainless steel
and mechanical properties was deeply explored. The inhomogeneity of during directional solidification and subsequent solid-state transformation, Mater.
microstructure and properties along the thickness and width direction Sci. Eng. A 444 (1–2) (2007) 64–68. https://linkinghub.elsevier.com/retrieve/pii
/S0921509306017904.
were analyzed, which was a major feature of this paper. The main [10] A. Schino, M. Mecozzi, M. Barteri, et al., Solidification mode and residual ferrite in
contents of this research can provide very important reference value for low-Ni austenitic stainless steels, J. Mater. Sci. 35 (2) (2000) 375–380. https://lin
the studies of electron beam welding on ultra-thick stainless steels, it k.springer.com/article/10.1023/A:1004774130483.
[11] L. Zhu, X. Liang, Effect of cooling rate on microstructure and solidification mode in
also provides theoretical guidance for the application of electron beam
Cr15Mn9Cu2Ni1N austenitic stainless steel, Foun Tech. 30 (07) (2009) 864–867.
welding in nuclear fusion and other fields of large structural parts. https://kns.cnki.net/kcms2/article/abstract?v=3uoqIhG8C44YLTlOAiTRKgch
rJ08w1e75TZJapvoLK2GcYYMAEceN-fecpNpAZGGvana8_OzM0gUfmo1xCHhwN
d_uDW_FPWh&uniplatform=NZKPT.
CRediT authorship contribution statement
[12] J. Kar, S. Roy, G. Roy, Effect of beam oscillation on microstructure and mechanical
properties of AISI 316L electron beam welds, Metal Mater Trans A 48 (4) (2017)
Jianguo Ma: Methodology, Investigation, Writing – original draft, 1759–1770. https://link.springer.com/article/10.1007/s11661-017-3976-2.
Conceptualization. Zhiyong Wang: Methodology, Investigation, [13] C. Hsieh, X. Guo, C. Chang, et al., Dendrite evolution of delta (δ) ferrite and
precipitation behavior of sigma (σ) phase during multipass dissimilar stainless
Writing – review & editing. Zhihong Liu: Writing – review & editing, steels welding, Metals Mater. Inter. 16 (3) (2010) 349–356. https://link.springer.
Supervision. Haibiao Ji: Data curation, Writing – review & editing. com/article/10.1007/s12540-010-0602-x.
Xiaowei Xia: Writing – review & editing, Supervision. Chengwen Li: [14] Z. Wang, J. Wu, Z. Liu, et al., Mechanical characterization and performance of
reduced scale mock-up for the CFETR first wall manufactured by SLM process,
Writing – review & editing. Zhenfei Liu: Investigation. Zhongtao Fusion Eng. Des. 182 (2022), 113235. https://linkinghub.elsevier.com/retrieve/pii
Zhang: Investigation, Data curation. Jiefeng Wu: Writing – review & /S0920379622002319.
editing, Supervision. [15] Y. Luo, J. Han, L. Zhu, et al., Study on inducement and equilibrium mechanism of
pore defects in vacuum electron beam welding, Trans. Chi. Weld. Inst. 38 (8)
(2017) 107–110. https://kns.cnki.net/kcms2/article/abstract?v=3uoqIh
Declaration of Competing Interest G8C44YLTlOAiTRKibYlV5Vjs7iAEhECQAQ9aTiC5BjCgn0RqbG-vH_BL0Rk3k1z0T
_m4TyfsxGUgExkJ996_Y6cQ9r&uniplatform=NZKPT.
[16] H. Li, Research on the relation of heat-treatment temperature with the quantity of
The authors declare that they have no known competing financial separating perovskite crystal and its grain size in blast furnace slag containing
interests or personal relationships that could have appeared to influence titanium, Energy Metal. Indus. 19 (4) (2000) 25–27. https://kns.cnki.net/kcms2
the work reported in this paper. /article/abstract?v=3uoqIhG8C44YLTlOAiTRKgchrJ08w1e79zTD32bjb4wBAii
s5n6JUJOi9UAWIdNe5_IlKN3saW0cTyiMPObKR0GoYvWmUbfw&uniplatform=N
ZKPT.
Data availability [17] M. Alali, I. Todd, B. Wynne, Through-thickness microstructure and mechanical
properties of electron beam welded 20mm thick AISI 316L austenitic stainless
steel, Mater Des. 130 (2017) 488–500. https://linkinghub.elsevier.com/retrieve
The authors do not have permission to share data.
/pii/S0264127517306998.
[18] Q. Lin. Research on fracture mechanical properties of electron beam welding joint
for thick stainless steel. Master’s thesis. SUES, 2015. https://kns.cnki.net/reade
Acknowledgement r/review?invoice=SC1HITUccNvAWNAAjzYypoR4DJi8WQ5a%2F3T
QP3iEvlu7eHwxCHBLYDLqw%2BthB3ApLnixsZxTysJSE9SeqHePbmH1pUEt8ybn
KQO4bF3pn2l4R1aTva4UhJie41l4iAeqD4yNZPTtJhExPXs4gWM3XVMKm7aNe
This research was funded by Comprehensive Research Facility for OgrdFlC%2FUm0asE%3D&platform=NZKPT&product=CMFD&filename

8
J. Ma et al. Fusion Engineering and Design 195 (2023) 113973

=1016159515.nh&tablename=cmfd201602&type=DISSERTATION&scope=tria l&cflag=overlay&dflag=&pages=&language=chs&trial=&nonce=DC25D0F64
E724967AD47FC519D31F7D1.

You might also like