Sie sind auf Seite 1von 10

Engineering Failure Analysis 57 (2015) 413–422

Contents lists available at ScienceDirect

Engineering Failure Analysis


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

Failure analysis on welded joints of 347H austenitic boiler tubes


Han-sang Lee ⁎, Jine-sung Jung, Doo-soo Kim, Keun-bong Yoo
Power Generation Laboratory, KEPCO Research Institute, 105 Munji-ro, Yuseong-Gu, Daejeon 305–760, Republic of Korea

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

Article history: Microstructural and thermo-mechanical analyses using the finite element method (FEM) were
Received 6 March 2015 performed on the welded joint of 347H boiler tubes in a coal-fired power plant after 3600 h of op-
Received in revised form 18 August 2015 eration. The cracks in the failed tube started on the inner surface of the heat-affected zone (HAZ).
Accepted 18 August 2015
This area had a high hardness value and volume fraction of strain-induced martensite. The plastic
Available online 22 August 2015
deformation around the crack was concentrated near the grain boundaries. This failure occurred
as the stabilizing effect disappeared due to carbide dissolution in the heat-affected zone, and plas-
Keywords: tic deformation and tensile residual stress were formed at the inner side due to the solidification
Reheat cracking
contraction of the outer bead.
Intergranular cracking
© 2015 Elsevier Ltd. All rights reserved.
Stabilized steel
Electron backscattered diffraction analysis

1. Introduction

There are continuing researches on the improvement of the efficiency of power plants due to economic and environmental
needs. In the case of coal-fired power plants, much of the continued effort is focused on enlarging the boilers and increasing the
steam temperature. These are possible only if the material's reliability in terms of the creep, steam oxidation, hot corrosion,
fatigue resistance, and weldability of its main components, such as its boiler tubes, is confirmed. Many tubes are arranged inside
the coal-fired boiler and connected through welding. Various types of alloys are applied according to each location's steam
temperature and pressure. Ferrite alloys are used in most parts. However, austenitic alloys which are expensive but have
excellent high temperature properties are applied to high temperature parts. With the increase in the steam temperature, the
use of austenitic stainless steel is being continuously expanded. In particular, austenitic stainless steels such as 347H,
347HFG, Super304H, and HR3C are mainly used on superheater and reheater tubes that are exposed to the highest temperatures
inside boilers [1].
Many failures at the welded joint of 347H reheater tubes have recently occurred during the short-term (one- to two-year) oper-
ation of the tubes, for which failure analysis has been required. 347 and 321 are steels stabilized by adding the strong carbide-forming
elements of Nb and Ti to form Nb and Ti carbides. This not only increases the resistance to intergranular corrosion in the heat-affected
zone at 450–800 °C, but also improves the creep strength [2–4]. In addition, the characteristics of the welded joint are affected by var-
ious factors such as the material, the welding process, the design, and the operational environments. Therefore, these factors must be
validated for accurate failure analysis. In this study, the microstructure of the welded joint of 347H, where cracks occurred after 3600 h
of operation, was analyzed. Moreover, an attempt was made to identify the cause of the failure of the welded joint by comparing the
joint with an as-welded specimen with the finite element method.

⁎ Corresponding author.

http://dx.doi.org/10.1016/j.engfailanal.2015.08.024
1350-6307/© 2015 Elsevier Ltd. All rights reserved.
414 H. Lee et al. / Engineering Failure Analysis 57 (2015) 413–422

Table 1
Chemical composition of the 347H tube (wt.%).

C Si Mn P S Ni Cr Nb

Specification 0.04–0.10 0.75 max 2.0 max 0.045 max 0.03 max 9.0–13.0 17.0–19.0 8 × C min 11 × C max
Measured value 0.072 0.32 1.99 0.03 0.012 10.1 18.02 0.97

2. Experiment method

The failed tube was located at the final reheater, where the designed steam temperature of 590 °C was highest in the coal-fired
boiler. The material was 347H austenitic stainless steel, the chemical composition of which is shown in Table 1. The failure shown
in Fig. 1 (a and b) occurred about 3600 h after the operation. Cracks were detected in the lower heat-affected zone of the welded
joint through non-destructive testing. To compare with the failed tube, as-welded specimens were made according to the welding
procedure specifications. The outer diameter and thickness of the tube were 63.5 and 4.3 mm, respectively, and two-layer, two-
pass welding was conducted via gas tungsten arc welding (GTAW). The detailed welding conditions are shown in Table 2. In the
first pass, the welding was performed at a lower current, voltage, and torch movement speed than in the second pass, to form a
back bead. The heat inputs of the first and second passes–1419 and 1477 J/s, respectively–did not significantly differ.
Grinding and polishing were performed on each test sample. The microstructure was analyzed using an optical microscope and a
scanning electron microscope (JSM-7001F, JEOL) after etching with a solution mixture of 10 ml nitric acid, 20 ml hydrochloric acid,
and 30 ml pure water. Electron backscattered diffraction analysis (EBSD, Oxford Channel 5) was performed on the heat-affected
zones. The hardness of the base material, the heat-affected zone, and the fusion zone was measured at an interval of 0.5 mm in a
load condition of 200 gf using a Vickers hardness tester.
As for the computational simulations based on the finite element method, thermal and mechanical analyses were conducted using
Sysweld. In such analyses, the welded joint had 33,520 3D finite elements. To simulate the arc heat source, Goldak's double ellipsoidal
heat source model was assumed. The welded joint was clamped in a way that rigid body motion is not possible. The air temperature
was set at 20 °C, and the heat exchange coefficient for free air cooling was used as a surface boundary condition. The thermal and me-
chanical analyses were performed from 0 to 18,000 s. In addition, the temperature was measured at a distance of 5 mm from the
welded joint with a K-type thermocouple to derive the parameters for the thermal analysis.

3. Results and discussion

3.1. Microstructure and hardness

The crack of the failed tube started on the inner surface of the heat-affected zone at a distance of 0.3 mm from the boundary of the
fusion zone and the heat-affected zone, and grew toward the outer side of the tube [Fig. 2(a)]. Here, the direction of the crack growth

Fig. 1. Crack of the 347H welded joint detected via non-destructive testing after 3600 h of operation: (a) outer side and (b) inner side.
H. Lee et al. / Engineering Failure Analysis 57 (2015) 413–422 415

Table 2
Welding process conditions of the as-welded specimen.

Voltage (V) Current (A) Velocity (mm/s) Heat input (J/s)

First pass 12 90 0.67 1419


Second pass 14 100 0.89 1477

can be confirmed from the thickness of the oxide layer around the crack. In Fig. 2(b), a 120 μm-thick oxide layer was formed at the
inner side, where the crack started. However, at the outer side, the oxide layer was only 40 μm thick. The crack grew along the
grain boundary; and while it branched out in many directions, the main growth direction was toward the outer side of the tube. In
addition, as in the rectangular dotted region in Fig. 2(a), multiple dark band layers were formed at the heat-affected zones
[Fig. 2(c)]. These could be partly seen on the base material, but their density was especially high around the heat-affected zone.
In most cases, cracks occurred in the heat-affected zones within 2 mm from the boundary between the fusion zone and the heat-
affected zone. Therefore, microstructural analysis in this region was essential. First, the microstructure was observed in areas 0.1, 1.0,
and 5.0 mm away from the boundary of the fusion zone and the heat-affected zone of the failed tube [Fig. 3(a)–(c)]. Many Cr carbides
were formed along the grain boundaries at the area 0.1 mm away from the boundary, but no precipitate was found inside the grain
[Fig. 3(a)]. Relatively small amounts of Cr carbides were formed in the grain boundaries at the area 1 mm away from the boundary.
Nb carbides were partly observed in the grain [Fig. 3(b)]. These Cr carbides in the grain boundaries and Nb carbides in the grain
had the chemical compositions of 51.5Cr–42Fe–6Ni–0.5Si and 75Nb–16Fe–6Cr–3Ni (wt.%). Moreover, in the area 5 mm away, the
number of Cr carbides at the grain boundary decreased even further [Fig. 3(c)]. The same locations were observed for the as-
welded specimen [Fig. 3(d)–(f)]. As in Fig. 3(d), in the area 0.1 mm away from the boundary, most precipitates in the grain interior
and boundary were dissolved while fine carbides at the grain interior were observed in the area 1 mm away [Fig. 3(e)]. Moreover,
in the area 5 mm away, not only fine Nb carbides at the grain interior but also over 1 μm coarse Nb carbides were observed [Fig. 3(f)].
During the welding process, the heat-affected zone is exposed to the solidus temperature of stainless steel for a short time. Here,
the Nb carbides at the heat-affected zone near the fusion boundary can be dissolved. Only a small part of the dissolved carbon is
precipitated again as Nb carbides after the welding due to the rapid cooling rate, and the stabilizing effect equal to the amount of
the dissolved carbon disappears. According to many studies on Nb-stabilizing austenitic stainless steel, 70–80% of the NbC was
dissolved when heat treatment was applied for 30 min to 1 h at 1300 °C [5]. Foklhard mentioned that the dissolution rate of Nb
carbide increased rapidly when the weld metal AWS E347 was maintained at above 1300 °C for 5 s [6]. Therefore, it can be concluded

Fig. 2. Optical image of the cross-sectional area for the welded joint where the crack occurred: (a) low magnification, (b) tube inner side, and (c) higher magnification in
the rectangular dotted region of (a).
416 H. Lee et al. / Engineering Failure Analysis 57 (2015) 413–422

Fig. 3. Microstructure according to the distance from the boundary between the fusion zone and the heat-affected zone: (a) 0.1 mm, (b) 1 mm, (c) 5 mm of the failed
tube, (d) 0.1 mm, (e) 1 mm, and (f) 5 mm of the as-welded specimen.

that Cr carbides were formed instead of Nb carbides at areas 0.1 mm and 1 mm from the boundary between the fusion zone and the
heat-affected zone in Fig. 3 (a and b) because the Nb carbides were dissolved during the welding process and did not precipitate again
at a high cooling rate [Fig. 3 (d and e)].
The degrees of hardness of the failed and as-welded specimens were measured at a 1 mm interval from the upper part through the
fusion zone to the lower part in the direction of (1) in Fig. 4(a). The hardness of the base material was 150–160 HV in the case of the as-
welded specimen, and significantly increased to 200–220 HV after the operation for 3600 h. This hardness increase can be attributed
to the precipitation of the carbides according to the increase in the operating time at a high temperature. Sourmail et al. explained that
the hardness of solution-annealed stainless steels is a function of time and temperature [7]. Carbide and an intermetallic compound, if
exposed to a high temperature for a long time, are precipitated on austenitic stainless steel. As the precipitated carbides inhibit the
movement of dislocations, the creep property is improved and the hardness increases.
H. Lee et al. / Engineering Failure Analysis 57 (2015) 413–422 417

Fig. 4. Results of the Vickers hardness on the cracked tube and the as-welded specimen: (a) measurement direction, (b) HAZ, and (c) fusion zone.

Also, the hardness was examined at a 0.5 mm interval from the outer side to the inner side of the tube at the heat-affected zone, as
in directions (2) and (3) in Fig. 4(a). The results are shown in Fig. 4 (b and c). In the failed tube, the hardness in the heat-affected zone
was 200–220 HV, similar to that of the base material, and lower than 210–230 HV in the fusion zone. As for the as-welded specimen,

Fig. 5. SEM micrograph and EBSD analysis results of the crack tip area of the heat-affected zones: (a) low magnification, (b) higher magnification of the rectangular
dotted region and the crystal structure, and (c) orientation imaging microscopy map.
418 H. Lee et al. / Engineering Failure Analysis 57 (2015) 413–422

Table 3
Thermal analysis results according to the heat input efficiency.

Heat input efficiency Peak temperature (°C) Bead width (mm)

First pass Second pass Outer surface Inner surface

Measured – 688 788 7.7 5.8


Simulated 0.575 720 800 7.2 5.3
0.600 742 822 7.5 5.7
0.625 764 844 7.8 6.1
0.650 784 865 8.1 6.5

its hardness in the heat-affected zone was 140 HV from the outer surface to a 2 mm depth, and lower than the 160–170 HV in the
fusion zone and the 150–160 HV of the base material. However, as the distance from the outer surface increased from 2 mm to
3.5 mm, the hardness of the specimen in both the heat-affected zone and the fusion zone increased rapidly to 200 HV. Such a hardness
increase did not occur at the base material, which was distant from the welded joint, so it can be thought to occur during the welding
procedure.
Irvine, who was the first to mention strain-induced precipitation hardening (SIPH), evaluated the creep characteristics of 18%Cr–
12%Ni–1%Nb steel, which are similar to 347H in the 1960s. He concluded that strain-induced precipitation significantly reduces elon-
gation [5]. In the results of his study, the amount and size of the NbC increased fast at the grain interior with the increase in the plastic
deformation. This distribution of the carbides at the grain interior relatively increased the resistance to deformation compared with
that at the grain boundary. Therefore, it caused grain boundary sliding and intergranular cracking in a creep environment. In addition,
Irvine pointed out that such strain-induced precipitation hardening can occur in the heat-affected zone of the welded joint for a high-
constraint condition. In the as-welded specimen in this study, the hardness increase at the inner side of the heat-affected zone could
have been caused by the plastic deformation in the high-constraint condition.
EBSD analysis was performed on the rectangular dotted region of the failed tube [Fig. 5(a)]. The dark bands formed in the heat-
affected zone in Fig. 5(b) were not γ austenite of fcc (face-centered cubic) but α' martensite of bct (body-centered tetragonal), and
its volume fraction was 32%. This was a marked increase compared with the 2% increase at the base material. The volume fraction
of α' martensite increased at the inner side of the heat-affected zone, and it can be thought that dark bands composed of α' martensite
were formed during the welding process. The γ austenite of stainless steel is a metastable phase formed by the addition of Cr and Ni. α'
martensite is formed if large amounts of processing and hardening occur during deformation [8,9]. This is called “strain-induced

Fig. 6. Parameter changes of the heat source model according to the welding torch angle: (a) backhand, (b) perpendicular, and (c) forehand techniques.
H. Lee et al. / Engineering Failure Analysis 57 (2015) 413–422 419

Table 4
Thermal analysis results according to the length ratio and the power ratio.

Welding technique Length ratio (Lf:Lr) Power ratio (Pf:Pr) Peak temperature (°C) Bead width (mm)

First pass Second pass Outer surface Inner surface

(a) Backhand 1:2 2:1 737 809 7.3 5.7


(b) Perpendicular 1:1 1:1 720 793 7.4 5.8
(c) Forehand 2:1 1:2 704 776 7.7 5.8

martensite,” the volume fraction of which is related to the temperature, deformation level, and rate [10]. Okayasu et al. studied the
strain-induced martensitic transformation that occurs on 304 stainless steels during tensile testing. They reported that the formation
of strain-induced martensite started above the yield strength, and that the volume fraction increased with the strength and the elon-
gation [11]. Therefore, the bands shown in Fig. 5(b) can be assumed to be the strain-induced martensites formed by thermal stress
during the welding process, and they also affected the hardness increase in the heat-affected zone in Fig. 4(b). Fig. 5(c) shows the ori-
entation imaging microscopy map from the EBSD analysis. The crystal orientation was measured at a 0.8 μm interval. In the case of the
fcc crystal structure, the crystal orientation difference from the adjacent measurement point was expressed as 2–5°, 5–10°, and over
10°. Through this, the site of the plastic deformation might be verified. While the twin boundary and the grain boundary formed a high
misorientation angle of over 10°, the surroundings of the grain boundary, where the crack occurred, had a misorientation angle of 2–
10°. This means the plastic deformation occurs not at the grain interior but the grain boundary, and the formation and propagation of
the crack follows the grain boundary. In many studies [13–15], including that of Van Wortel [12], the failure mechanism of relaxation
cracking for austenitic stainless steel is explained as follows. If exposed to high temperatures, fine carbides are formed at dislocations
inside the grain. As these carbides inhibit the movement of the dislocations, there is no deformation inside the grain, and the stress is
concentrated on the grain boundary. Thus, grain boundary deformation occurs at the creep environment. In addition, Ben-Haroe et al.
performed cold rolling on 347 stainless steels and observed long-term microstructural changes at 600 °C and 750 °C [16]. While no
change was detected for up to 500 h at 600 °C, not only the growth of the carbides but also their recovery and recrystallization oc-
curred at 750 °C. Based on this, it can be assumed that the cracks occurred at the grain boundary of the heat-affected zone in the failed
tube because the hardened state of the grain interior due to the precipitation of the fine carbides was maintained without recovery
and recrystallization for up to 3600 h at 590 °C.

3.2. Thermal and mechanical analyses

Several models have been proposed to simulate the moving heat source in FEM analysis [17–20]. In this study, the commercial soft-
ware package Sysweld was used as the FEM analysis tool, and the analysis was performed using Goldak's double ellipsoidal heat
source model. The bead width of each pass and the temperature change at the outer surface of the tube 5 mm away from the welded
joint were measured, and the heat source parameters were repeatedly corrected to make them consistent with the experiment
results.
The parameters of the heat source model are divided into the heat input efficiency, length ratio, and power ratio. The heat input
efficiency is how much of the heat energy generated from the arc is used to melt the filler metal and the base material. It is influenced
by various factors such as the welding method, the protective gas, the arc length, etc. Table 3 shows the results of the analysis of the
peak temperature and the bead width of the first and second passes with changing heat input efficiency. When the heat input

Fig. 7. Comparison of the temperature changes between the thermal analysis and the experimental measurement during the welding.
420 H. Lee et al. / Engineering Failure Analysis 57 (2015) 413–422

Fig. 8. Distribution of the plastic deformation and the residual stress in the axial direction through mechanical analysis [plastic deformation after the (a) first pass and
(b) second pass, and residual stress after the (c) first pass and (d) second pass].

efficiency increased from 0.575 to 0.650, the peak temperature and the bead width at the inner and outer sides of the tube also in-
creased. In the case of the measured values, the peak temperatures of the first and second passes were 688 °C and 788 °C, respectively,
and the outer and inner bead widths were 7.7 mm and 5.8 mm, respectively. The heat input efficiency should be lower than 0.575,
considering the measured peak temperature. However, as the variation in the bead width is smaller than in the peak temperature
when changing the parameters of the heat source model, the 0.6 heat input efficiency was selected.
The other parameters were the length ratio (Lf:Lr) and the heat input ratio (Pf:Pr) between the front ellipsoid and the rear ellipsoid
in the double ellipsoidal heat source model in Fig. 6. The factors that mainly affected the length ratio and the power ratio could be the
angle of the welding torch and the used techniques, such as the forehand, perpendicular, and backhand techniques [21]. Therefore, the
analysis was carried out for the following three categories: the backhand technique, in which the heat input is larger in front of the
heat source model, as shown in Fig. 6(a); the perpendicular technique, in which the heat input is larger at the center of the heat source
model, as in Fig. 6(b); and the forehand technique, wherein the heat input is larger at the back of the heat source model in Fig. 6(c).
The comparison of the analysis results in Table 4 showed that the backhand technique has a higher peak temperature and a smaller

Fig. 9. Formation mechanism of the plastic deformation and the residual stress at the inner side of the tube during the welding.
H. Lee et al. / Engineering Failure Analysis 57 (2015) 413–422 421

bead width in the first and second passes. In the forehand technique, the peak temperature was lower and the bead width larger for
the first and second passes. Among these three cases, the bead width of the forehand technique was consistent with the measured
value, as was the peak temperature. The forehand technique generally has low heat input efficiency, a shallow depth, and a large
bead width. These features are consistent with the results of the analysis on the forehand technique in Table 4. J. Chen et al. analyzed
the heat distribution of the GTAW heat source according to the torch angle, and compared the results with the experiment results [22].
It was referred that the decrease in the torch angle in the forehand technique caused the wider distribution area of the heat sources,
and not only the peak temperature but also the heating rate decreased. The analysis results in Table 4 show a lower peak temperature
in the forehand technique, but no peak temperature reduction in the heating rate was observed because the total length of the heat
source was fixed.
The parameters of the heat source model in the forehand method were further optimized, and the temperature changes for the
length and power ratios of 1.8:1 and 1:2.5 are shown in Fig. 7. The bead widths of the outer and inner sides were 7.7 mm and
5.8 mm, respectively, the same as the measured values. In the first pass in Fig. 7, a difference of 20 °C at the peak temperature occurred
between the analyzed and measured values. This could have been caused by the non-reflection of the Ar purging to prevent the ox-
idation of the back beads, and by changing the heat input to form back beads in the computational simulations.
The distributions of the residual stress in the axial direction through the mechanical analysis are shown in Fig. 8 (a and b). The ten-
sile stress was formed at the inner side of the heat-affected zone after the first pass [Fig. 8(a)]. After the second pass, the tensile stress
in the same area significantly increased to 380 MPa, and the compressive stress formed in the heat-affected zone of the outer side of
the tube [Fig. 8(b)]. While there was less than 2% plastic deformation after the first pass [Fig. 8(c)], the over 16% plastic deformation
was concentrated in the heat-affected zone after the second pass [Fig. 8(d)]. The analysis results showed that the region with high
plastic deformation corresponded to the inner side of the tube, close to the boundary between the fusion zone and the heat-
affected zone, where the crack occurred, as shown in Fig. 2 (a and b). This is also identical to that in the area where the hardness in-
creased significantly at the inner side of the heat-affected zone in the as-welded specimen [Fig. 4(b)].
As a result, the formation of the tensile residual stress and the plastic deformation at the inner side of the heat-affected zone can be
explained as shown in Fig. 9. When the second pass is carried out after completing the first pass, the fusion zone of the second pass
starts to solidify and contract mostly in the axial direction [Fig. 9(a)]. Here, as the rectangular dotted region constrained by the
base material is relatively thin, as shown in Fig. 9(b), elongation and plastic deformation occur through the contraction of the outer
bead. Also, during the long-term operation at around 600 °C, the precipitation of the fine Nb carbide caused the hardening of the
grain interior in the heat-affected zone, and of coarse Cr23C6 at the grain boundary. In these conditions, the plastic deformation con-
centrates on the relatively weak grain boundary and results in an intergranular crack.
To prevent SIPH failure, the ASME Boiler and Pressure Vessel Code Sec. 1 provides guidelines for heat treatment after cold working.
On 347H, post-welding heat treatment at over 1095 °C is recommended if the cold deformation is over 15% for an operating temper-
ature range of 540–675 °C. Today, however, there are no guidelines for welded joints through the ASME code and for post-welding
heat treatment (PWHT) at 347H. Based on the results of this study, post-welding heat treatment can be considered a possible solution
because the plastic deformation is over 15% at the inner side of the heat-affected zone. Alternatives for improving the welding process
could be preheating and increasing the interpass temperature to minimize the residual stress from the thermal stress.

4. Conclusion

– Based on the following results, the intergranular crack of the 347H welded joint was found to have been caused by reheat cracking
or strain-induced precipitation hardening.
• Formation of coarse Cr carbides in the heat-affected zone of the failed tube
• The high volume fraction of the strain-induced martensite on the inner side of the HAZ and the concentration of the plastic de-
formation around the grain boundary
• Dissolution of NbC at the HAZ and increase in the hardness at the inner side of the as-welded specimen.
– The results of the thermal and mechanical analyses showed that the tensile residual stress and the plastic deformation were
highest on the inner surface of the HAZ, where the crack started. These increased greatly during the second pass.
– During the solidification of the second pass, rapid volume contraction occurred in the axial direction. The plastic deformation and
the tensile residual stress were concentrated at the relatively thin inner side of the HAZ.

Acknowledgment

This study was supported by the Korean Institute of Energy Technology Evaluation and Planning (code #: 20131020102200) of the
Republic of Korea.

References

[1] State of Knowledge for Advanced Austenitics, ERRI, 2009 (Report No. 1020241).
[2] Fujio Abe, Creep-resistant Steels, Woodhead Publishing, 2008.
[3] A. John Sedriks, Corrosion of Stainless Steels, 2nd ed. John Wiley & Sons, 1996.
[4] H.S. Khatak, Corrosion of Austenitic Stainless Steels, Narosa Publishing House, 2002.
[5] K.J. Irvine, The effect of heat treatment and microstructure on the high-temperature ductility of 18%Cr–12%Ni–1%Nb steels, J. Iron Steel Inst. (1960) 166–179.
[6] Erich Folkhard, Welding of Stainless Steels, Springer-Verlag/Wien, 1988.
422 H. Lee et al. / Engineering Failure Analysis 57 (2015) 413–422

[7] T. Sourmail, Precipitation in creep-resistant austenitic stainless steels, Mater. Sci. Technol. 17 (2001) 1.
[8] W.J. Dan, A model for strain-induced martensitic transformation of TRIP steel with the strain rate, Comput. Mater. Sci. 40 (2007) 101.
[9] I. Tamura, Deformation-induced martensitic transformation and transformation-induced plasticity in steels, Met. Sci. 16 (1982) 245.
[10] X. Chen, Dynamic behavior of SUS304 stainless steel at elevated temperatures, J. Mater. Sci. 39 (2004) 4869.
[11] M. Okayasu, Strain-induced martensite formation in austenitic stainless steel, J. Mater. Sci. 48 (2013) 6157.
[12] Hans van Wortel, Control of Relaxation Cracking in Austenitic High-temperature Components. Corrosion 2007 Conference & Expo. NACE International: Paper No.
07423, 2007.
[13] Q. Auzoux, Effect of pre-strain on the creep of three AISI 316 austenitic stainless steels in relation to the reheat cracking of weld-affected zones, J. Nucl. Mater. 400
(2010) 127.
[14] S. Hong, Unraveling the Origin of the strain-induced precipitation of M23C6 in a plastically deformed 347 austenite stainless steel, Mater. Charact. 94 (2014) 7.
[15] J. Swaminathan, Sensitization-induced stress corrosion failure of AISI 347 stainless steel fractionator furnace tubes, Eng. Fail. Anal. 18 (2011) 2211.
[16] I. Ben-Haroe, Evolution of the microstructure of AISI 347 stainless steel during heat treatment, Mater. Sci. Technol. 9 (1993) 620.
[17] S.H. Ko, Effects of surface depression on pool convection and geometry in stationary GTAW, Weld. Res. Suppl. 80 (2001) 39.
[18] D.W. Cho, A study on V-groove GMAW for various welding positions, J. Mater. Process. Technol. 213 (2013) 1640.
[19] J. Goldak, A new finite element model for welding heat sources, Metall. Mater. Trans. B 15 (1984) 299.
[20] H.Y. Fang, New general double ellipsoid heat source model, Sci. Technol. Weld. Join. 10 (2005) 361.
[21] Larry Jeffus, Welding Skills, Processes, and Practices for Entry-level Welders, 1st ed. Delmar Cengage Learning, 2010.
[22] J. Chen, Improvement of welding heat source models for the TIG-MIG hybrid welding process, J. Manuf. Process. 16 (2014) 485.

Das könnte Ihnen auch gefallen