Next Article in Journal
Computing the Fatigue Life of Cold Spray Repairs to Simulated Corrosion Damage
Next Article in Special Issue
Effect of Yield Strength Distribution Welded Joint on Crack Propagation Path and Crack Mechanical Tip Field
Previous Article in Journal
Boosting Piezo/Photo-Induced Charge Transfer of CNT/Bi4O5I2 Catalyst for Efficient Ultrasound-Assisted Degradation of Rhodamine B
Previous Article in Special Issue
Characterization of Mechanical Heterogeneity in Dissimilar Metal Welded Joints
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Mechanical Heterogeneity on Strain and Stress Fields at Crack Tips of SCC in Dissimilar Metal Welded Joints

1
School of Mechanical Engineering, Xi’an University of Science and Technology, Xi’an 710054, China
2
School of Science, Xi’an University of Science and Technology, Xi’an 710054, China
*
Author to whom correspondence should be addressed.
Submission received: 7 July 2021 / Revised: 31 July 2021 / Accepted: 5 August 2021 / Published: 9 August 2021
(This article belongs to the Special Issue Extreme Mechanics in Multiscale Analyses of Materials)

Abstract

:
The crack tip strain and stress condition are one of the main factors affecting stress corrosion cracking (SCC) behaviors in the dissimilar metal welded joint of the primary circuit in the pressurized water reactor. The mechanical property mismatch of base metal and weld metal can significantly affect the stress and strain condition around the crack tip. To understand the effect of different weld metals on strain and stress fields at SCC crack tips, the effects of strength mismatch, work hardening mismatch, and their synergy on the strain and stress field of SCC in the bi-material interface, including plastic zone, stress state, and corresponding J-integral, are investigated in small-scale yielding using the finite element method. The results show a significant effect of the strength mismatch and work hardening mismatch on the plastic zone and stress state in the weld metal and a negligible effect in the base metal. J-integral decreases with the single increase in either strength mismatch or work hardening mismatch. Either the increase in strength mismatch or work hardening mismatch will inhibit the other’s effect on the J-integral, and a synthetic mismatch factor can express this synergistic effect.

1. Introduction

As the operating time of nuclear power plants increases, stress corrosion cracking (SCC) has become a widespread issue during the last few decades [1]. SCC is a slow crack growth process under the combination of environment, material, and load stress in the crack tip area [2]. Because of the heterogeneous microstructure and mechanical properties in the bi-material interface regions, the dissimilar metal weld joint (DMWJ) is more susceptible to SCC and becomes the weakest link of structural materials in the pressurized water reactor (PWR) [3,4]. The stress and strain condition around the crack tip is a critical factor affecting the SCC behaviors [1]. Therefore, it is essential to calculate strain and stress fields at the crack tip of SCC for predicting the crack growth direction and rate of SCC quantitatively. The mechanical property mismatch is an important factor to be considered because it can significantly affect the stress and strain condition around the crack tip.
In general, the mechanical property mismatch is defined by the strength mismatch ratio of the yield strength of the weld metal and base metal. Many efforts were made regarding the effect of strength mismatch on the crack-tip stress and strain field [5] and corresponding crack driving force [6,7], crack-tip constraint [8,9]. In most of these studies, the default weld metal and base metal have the same hardening properties, i.e., the true stress–strain curves of two metals after yielding are parallel. Nevertheless, a few studies showed that the work hardening mismatch could significantly affect the crack-tip stress and strain fields and crack driving force with the same yield strength [10,11,12]. In fact, the yield strength and work hardening of the base metal and weld metal are mismatched simultaneously in the DMWJ. Due to the use of different base metals and weld metals, the mismatch states are different, and the degradation of materials in the welded joint will occur and change the macro mechanical properties due to the influence of extreme environments such as irradiation and high temperature [13]. Moreover, the “local mismatch” was proposed meaning the difference of yield strength and work hardening near the crack [14,15]. Therefore, the effect of mechanical property mismatch on the crack-tip stress and strain field of SCC needs to be studied in more detail.
However, when it comes to the effects of strength mismatch and work hardening mismatch on the crack-tip stress and strain field and crack driving force, it can be found that previous studies focused on one of them with the other unchanged or ignored. The synergy (inhibition or promotion) of these two mismatches has not been studied. In the present study, the effects of strength mismatch, work hardening mismatch, and their synergy on the strain and stress field of interface SCC crack tip are investigated in small-scale yielding by the finite element method (FEM) based on elastic–plastic fracture mechanics (EPFM) approach. The plastic zone and stress level near the crack tip and corresponding J-integral are also discussed. The results presented in this work can provide more detailed recommendations for selecting welding materials and evaluating the structural integrity of the DMWJ from the perspective of the SCC crack driving force.

2. Materials and Methods

2.1. Materials and Specimen Geometry

The Ramberg-Osgood equation can be used to express the true stress–strain (σ-ε) curves of both the weld metal and base metal in the DMWJ, which is written as
ε ε 0 = σ σ 0 + α σ σ 0 n ,
where σ0 is the yield strength, ε0 is the yield strain, α is the Ramberg–Osgood coefficient, and n is the work hardening exponent. The safe-end DMWJ is utilized to join the ferric steel pipe-nozzle of the pressure vessel with the austenitic stainless steel safe-end. Ferritic low alloy steel A508, usually used for the pipe-nozzle of the nuclear pressure vessel, is selected as the base metal. The material mechanical parameters of A508 are Young’s modulus E = 183,000 MPa, Poisson’s ratio ν = 0.3, yield stress σ0 = 410 MPa, strain hardening exponent n = 5.41, and the Ramberg–Osgood coefficient α = 5.11 at the operating temperature of about 340 °C [16,17]. At present, the world’s uses of safe-end weld metal are 309L/308L, Inconel 82, and Inconel 152 in the United States; Soudonel 690 in Belgium; Thermanit 690 in Germany; Nic 703D in Japan; and Sanicro 71 in Sweden. They all have different yield strengths and work hardening coefficients. For example, the yield strength of alloy 82 is 315 MPa, and the work hardening coefficient is 7.01 at 320 °C [18]; the yield strength of alloy 52 is 380 MPa, and the work hardening coefficient is 3.29 at 340 °C [19]; the yield strength of 309L/308L is 270 MPa, and the work hardening coefficient is 3.74 at 295 °C [20]. Therefore, a wide range of the strength and work hardening mismatch ratio should be selected in this study for almost covering the yield strength and work hardening coefficient of all weld metal.
To investigate effects of strength mismatch, work hardening mismatch, and their synergy on the strain and stress field, it is assumed that Young’s modulus E, Poisson’s ratio ν, and the Ramberg–Osgood coefficient α in the weld metal are consistent with base metal, but the yield strength and work hardening exponent vary in the weld metal. The strength mismatch ratio is usually defined as
M S = σ W σ B ,
where σW and σB are the yielding strength of the weld metal and base metal. MS varies from 0.7 to 1.3, which is considered to be a typical range [21]. The MS is 0.7, 0.85, 1, 1.15, and 1.3, and the corresponding yield strength of the weld metal is 287, 348.5, 410, 471, and 533 MPa in this study. Similar to the strength mismatch, the work hardening mismatch ratio is defined as
M N = n W n B ,
where nW and nB are the work hardening exponent of the weld metal and base metal, respectively. The range of MN from 0.5 to 1.8 is considered, and the value interval is 0.1. Therefore, the work hardening exponent of the weld metal varies from 2.71 to 9.74, which is approximately coincident with the variation range of work hardening exponent in other studies [22,23].
As shown in Figure 1, according to ASTM E399-17 [24], 1T-CT welded joint specimen (W = 50 mm, a = 0.5 W) was used in finite element analysis, which is often used in SCC experiments [25]. The crack is located at the interface of the weld metal and base metal.

2.2. FEM Model

The two-dimensional finite element model with 10738 8-node biquadrate plane strain quadrilateral (CPE8) elements is shown in Figure 2. The meshes around the crack tip were refined with 1728 CPE8 elements, as shown in Figure 2b. The crack growth direction is the X direction in the coordinate system. The finite element model was calculated by ABAQUS code using the incremental theory of plasticity. The stress intensity factor (K) is usually used as a parameter to measure the external load of the CT specimen in SCC experiments. To investigate the synergistic effect of two mismatches on the strain and stress field under the same load, K is set as constant 30 MPa·m1/2, which is usually used in SCC experiments [26,27]. The specimen was loaded by concentrated force in the center point of two loaded holes in the vertical direction, and all other motions of the center point were restrained except the vertical direction. The coupling constraint was defined between the center point and the loaded hole.

3. Results and Discussion

3.1. Plastic Zone

The local stress and strain distribution near the crack tip can significantly affect the SCC growth rate and direction. Therefore, they are usually used as mechanical parameters to predict the SCC behavior quantitatively [28,29]. The plastic zone around the crack tip is investigated in this section, and the stress distribution will be investigated next section. To investigate the plastic zone on both sides of the SCC crack, as shown in Figure 3, the plastic zone area (surrounded by 0.2% equivalent plastic strain isoline) of the weld metal and base metal are defined and expressed by ApW and ApB, respectively.
The plastic zone area of the weld metal and base metal are shown in Figure 4, where ApW and ApB are normalized by the plastic zone area of any side homogeneous base metal (MS = MN = 1) Apref. It can be seen from Figure 4a that when work hardening is evenly matched (MN = 1), the plastic zone area of the weld metal decreases with increasing strength mismatch ratio, consisting with the study of Lee and Kim [30]. When yield strength is evenly matched (MS = 1), the plastic zone area of the weld metal also decreases with increasing work hardening mismatch ratio. Meanwhile, with increasing MS from 0.7 to 1.3, the change degree of ApW with MN is reduced. Additionally, the change degree of ApW with MS is reduced with increasing MN from 0.5 to 1.8. It indicates that when the yield strength and work hardening exponent are mismatched simultaneously, the increase in either inhibits the effect of the other on the plastic zone area of the weld metal. Figure 4b shows that when work hardening is evenly matched (MN = 1), the plastic zone area of the base metal increases with increasing MS. When yield strength is evenly matched (MS = 1), the plastic zone area of the base metal also increases with increasing MN, which is opposite to the trend of change in the weld metal. This opposite phenomenon can be explained as the increase in the plastic zone in the weld metal releases more high stress at the crack tip. Therefore, the high stress to be released in the base metal decreases, the plastic zone of the base metal decreases correspondingly, and vice versa. Those results are consistent with the other studies [29,30]. However, when comparing Figure 4b with Figure 4a, the variation in the plastic zone area in the base metal (from 0.62 to 1.33) is much smaller than weld metal (from 0.17 to 13.06). Thus the effects of strength mismatch and work hardening mismatch on the plastic zone area are negligible in the base metal but significant in the weld metal.
It is assumed that the plastic zone area is regarded as the crack driving force in addition to the applied loading [15]. Therefore, the ratio of ApW and ApB is greater than 1, indicating that the crack driving force in the weld metal is greater than the base metal, and the crack might grow into the weld metal and vice versa. Figure 5 shows the ratio between the plastic zone area of the weld metal and base metal. The ratio of ApW and ApB is dominated by the ApW as the variation in ApW is much greater than ApB. The results in Figure 5 further indicate that the crack growth direction depends on both strength mismatch and work hardening mismatch. For example, the crack might grow into the base metal when MS > 1, without considering the work hardening mismatch (MN = 1), but into the weld metal when MN = 0.7. Therefore, it is not enough to consider one mismatch only for estimating the crack growth direction, and the synergistic effect of strength mismatch and work hardening mismatch must be considered.

3.2. Stress Distribution

For convenience, the “point match” method, i.e., selecting one or several points near the crack tip to characterize the stress level [31], is used to study the stress field at the crack tip. Additionally, the 20 μm can be used as the distance from the characteristic point to the crack tip of the SCC [2]. Therefore, to investigate stress level on both sides of the SCC crack, two points in r = 0.02 mm, θ = ±90°, are selected in this study, where r and θ are the polar coordinates centered at the crack tip with θ = 0° corresponding to the interface. The stress level in the base metal and weld metal is represented by the Von Mises stress at these two points and is expressed by σVB and σVW.
Figure 6 shows the Von Mises stress in the weld metal and base metal for different strength mismatches and work hardening mismatches, where σVB and σVW are normalized by the Von Mises stress of the homogeneous base metal (MS = MN = 1) σV0 in r= 0.02 mm, θ = 90°. The results in Figure 6a suggest that when work hardening is evenly matched (MN = 1), the Von Mises stress in the weld metal increases with increasing strength mismatch ratio MS, consistent with other studies [5,29]. When yield strength is evenly matched (MS = 1), the Von Mises stress in the weld metal decreases with increasing work hardening mismatch ratio MN. Meanwhile, the change degree of σVW with MN decreases with increasing MS (from 0.7 to 1.3), and the change degree of σVW with MS increases with increasing MN (from 0.5 to 1.8). It indicates that when the yield strength and work hardening exponent are mismatched simultaneously, the increase in MS inhibits the effect of MN on the stress level in the weld metal. The increase in MN promotes the effect of MS on the stress level in the weld metal. Figure 6b shows that the tread of σVB is similar to that of σVW, which is because the stresses in the weld metal and base metal are similar to the relationship between force and reaction force. However, the maximum variation in σVB, which occurs in MS = 0.7 and MN = 1.8, is less than 8% compared with the homogeneous base metal. It implies that the effects of strength mismatch and work hardening mismatch on stress in the base metal can be neglected compared with variation in stress level in the weld metal (33%).
To compare the Von Mises stress in the weld metal and base metal, Figure 7 illustrates the ratio between them. The variation in σVW/σVB is consistent with the variation in σVW (Figure 6a) because the ratio of σVB and σVW is dominated by σVW. As MS changes, the σVW/σVB curve is not a simple translation. It demonstrates that the synergistic effect of strength mismatch and work hardening mismatch on stress distribution at the crack tip is significant.

3.3. J-Integral

The stress and strain fields at the crack tip are not in accordance with the HRR (Hutchinson–Rice–Rosengren) field distribution due to the mechanical property mismatch of the base metal and weld metal. However, it is notable that the J-integral is still an effective parameter to measure the scale of stress and strain region at the SCC crack tip [32]. When the stress intensity factor (K) is constant, the mechanical property mismatch affects the J-integral.
Figure 8 shows the J-integral with different strength mismatches and work hardening mismatches. The J-integral for homogeneous base metal (MS = MN = 1) is 4.651, and its value is 4.602, calculated by the EPRI (Electric Power Research Institute) approach. The relative error is 1.07%, which demonstrates that the FEM analyses are reliable in this study. The J-integral decreases with increasing strength mismatch ratio MS when work hardening is evenly matched (MN = 1), which is consistent with Lee’s research [6]. When yield strength is evenly matched (MS = 1), the J-integral also decreases with increasing work hardening mismatch ratio MN, and the variation range less than 1% when MN > 1. In addition, with increasing MS from 0.7 to 1.3, the J-integral curve to MN does not simply translate, and the change in J-integral reduces from 29.6% to 11.7%. With increasing MN from 0.5 to 1.8, the change in J-integral with MS reduces from 17.8% to 1.5%. Therefore, when the yield strength and work hardening exponent are mismatched simultaneously, the increase in either will inhibit the effect of the other on J-integral rather than superimposed on each other simply. The results in Figure 8 further indicate that higher yield strength and work hardening coefficient of weld metal will lead to a smaller crack driving force (J-integral) for interface SCC crack.
Further, in order to characterize the synergistic effect of strength mismatch and work hardening mismatch on J-integral, the mismatch factor is attempted to define as
M 2 = M N M S ,
Figure 9 shows a fitting empirical equation curve of J-integral about M2 based on Figure 8, where the J-integral is normalized by the J-integral of homogeneous base metal Jref for engineering applications. There is an excellent fitting effect between J/Jref and M2, and the equation is expressed as
J / J r e f = a M 2 b + c ,
where a = 0.01709, b = −3.358, and c = 0.9872, the coefficient of correlation R-square (R2) is 0.988. Figure 9 further indicates that the J-integral decreases as a power function as the mismatch factor M2 increases, and the variation (less than 1.6%) in J-integral can be neglected when M2 > 1.
Figure 10 illustrates the comparison between the FEA results and predicted J-integral values with MS = 0.8, 0.9, 1.1, and 1.2. Good agreements are achieved, and the maximum relative error is less than 1%. Therefore, M2 quantifies the interdependencies of effects among strength mismatch and work hardening mismatch on J-integral and might be a suitable strategy to estimate the J-integral of interface SCC crack for DMWJs. For example, after J-integral of homogeneous base metal Jref is obtained by the Engineering Treatment Model (ETM) or EPRI, the J-integral of interface crack for DMWJs can be estimated according to the mismatch factor M2 between the weld metal and base metal. However, it should be noted that as a practical attempt to characterize the synergistic effect of strength mismatch and work hardening mismatch on J-integral, M2 and corresponding Equation (5) are applicable in CT specimen with small scale yielding in this research. The effects of constraints by different geometries and loads on M2 need to be further investigated.
The J-integral, with respect to strength mismatch application, was extended based on the EPRI approach [8,33] and used in the structural integrity assessment, such as R6 or BS 7910. However, this study demonstrates that the synergistic effect of strength mismatch and work hardening mismatch on the J-integral is significant and must be considered in small-scale yielding. To consider only one effect would result in a conservative or non-conservative result. Therefore, it is recommended to calculate the crack driving force of SCC related to both strength and hardening mismatch in the integrity design and evaluation of the DMWJ in nuclear power plants.

4. Conclusions

In conclusion, the synergistic effect of strength mismatch and work hardening mismatch on the strain and stress fields of SCC crack tip, including plastic zone, stress state, and corresponding J-integral, were investigated in small-scale yielding by FEM numerical analyses based on the EPFM approach. The following main conclusions were drawn from the results obtained:
  • The effect of strength mismatch and work hardening mismatch on the plastic zone area is negligible in the base metal but significant in the weld metal. The increase in either will inhibit the effect of the other on the plastic zone area of the weld metal. The crack growth direction depends on the synergistic effect of two mismatches;
  • The effect of strength mismatch and work hardening mismatch on the stress distribution is negligible in the base metal but significant in the weld metal. The increase in strength mismatch ratio will inhibit the effect of the work hardening mismatch ratio on the stress level of the weld metal. The increase in work hardening mismatch ratio will promote the effect of the strength mismatch ratio on the stress level of the weld metal;
  • J-integral decreases with the single increase in either strength mismatch or work hardening mismatch. Either the increase in strength mismatch or work hardening mismatch will inhibit the other’s effect on the J-integral, and this synergistic effect may be expressed by a synthetic mismatch factor M2. Further study of the mismatch factor M2 and its use are currently being carried out;
  • Higher yield strength and work hardening coefficient of weld metal will reduce the crack driving force (including plastic zone and J-integral) for interface SCC crack. It is recommended to calculate the strain and stress field and crack driving force of SCC for both strength and hardening mismatch in structural integrity assessment.

Author Contributions

Conceptualization, S.Z. and H.X.; methodology, S.Z. and S.W.; software, S.Z., Y.Z. and Y.S.; validation, S.W., Y.S., F.Y. and Y.Z.; writing—original draft preparation, S.Z.; writing—review and editing, H.X., S.W. and F.Y.; supervision, H.X. and F.Y.; project administration, H.X.; funding acquisition, H.X. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China, grant number 52075434; Natural Science Foundation of Shaanxi Provincial Department of Education, grant number 2021KW-36.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Andresen, P.L. A Brief History of Environmental Cracking in Hot Water. Corrosion 2018, 75, 240–253. [Google Scholar] [CrossRef]
  2. Xue, H.; Li, Z.; Lu, Z.; Shoji, T. The effect of a single tensile overload on stress corrosion cracking growth of stainless steel in a light water reactor environment. Nucl. Eng. Des. 2011, 241, 731–738. [Google Scholar] [CrossRef]
  3. Liu, X.; Hwang, W.; Park, J.; Van, D.; Chang, Y.; Lee, S.H.; Kim, S.-Y.; Han, S.; Lee, B. Toward the multiscale nature of stress corrosion cracking. Nucl. Eng. Technol. 2018, 50, 1–17. [Google Scholar] [CrossRef]
  4. Dong, L.; Peng, Q.; Xue, H.; Han, E.-H.; Ke, W.; Wang, L. Correlation of microstructure and stress corrosion cracking initiation behaviour of the fusion boundary region in a SA508 Cl. 3-Alloy 52M dissimilar weld joint in primary pressurized water reactor environment. Corros. Sci. 2018, 132, 9–20. [Google Scholar] [CrossRef]
  5. Zhang, Z.L.; Hauge, M.; Thaulow, C. Two-parameter characterization of the near-tip stress fields for a bi-material elastic-plastic interface crack. Int. J. Fract. 1996, 79, 65–83. [Google Scholar] [CrossRef]
  6. Lee, H. Estimation of crack driving forces on strength-mismatched bimaterial interfaces. Nucl. Eng. Des. 2002, 212, 155–164. [Google Scholar] [CrossRef]
  7. Castelluccio, G.M.; Cravero, S.; Bravo, R.; Ernst, H. Mismatch effect on fracture driving force in mismatched girth welded pipes. In Proceedings of the ASME 2011 Pressure Vessels and Piping Conference, Baltimore, MA, USA, 17–21 July 2011. [Google Scholar] [CrossRef]
  8. Koo, J.-M.; Huh, Y.; Seok, C.-S. Plastic η factor considering strength mismatch and crack location in narrow gap weldments. Nucl. Eng. Des. 2012, 247, 34–41. [Google Scholar] [CrossRef]
  9. Su, L.; Xu, J.; Song, W.; Chu, L.; Gao, H.; Li, P.; Berto, F. Numerical Investigation of Strength Mismatch Effect on Ductile Crack Growth Resistance in Welding Pipe. Appl. Sci. 2020, 10, 1374. [Google Scholar] [CrossRef] [Green Version]
  10. Min, Z.; Shi, Y.; Zhang, X. Influence of strength mis-matching on crack driving force and failure assessment curve of weldment. Int. J. Press. Vessels Pip. 1997, 70, 33–41. [Google Scholar] [CrossRef]
  11. Østby, E.; Zhang, Z.; Thaulow, C. Constraint effect on the near tip stress fields due to difference in plastic work hardening for bi-material interface cracks in small scale yielding. Int. J. Fract. 2001, 111, 87–103. [Google Scholar] [CrossRef]
  12. Lee, H.; Kim, Y.-J. Interfacial crack-tip constraints and J-integral for bi-materials with plastic hardening mismatch. Int. J. Fract. 2007, 143, 231–243. [Google Scholar] [CrossRef]
  13. Herold, H.; Zinke, M.; Hübner, A. Investigations on the use of Nitrogen Shielding Gas in Welding and its Influence on the Hot Crack Behaviour of High-Temperature Resistant Fully Austenitic Ni- and Fe-Base Alloys. Weld. World 2005, 49, 50–63. [Google Scholar] [CrossRef]
  14. Wang, H.; Wang, G.; Xuan, F.; Tu, S. An experimental investigation of local fracture resistance and crack growth paths in a dissimilar metal welded joint. Mater. Des. 2013, 44, 179–189. [Google Scholar] [CrossRef]
  15. Zerbst, U.; Ainsworth, R.; Beier, H.; Pisarski, H.; Zhang, Z.; Nikbin, K.; Nitschke-Pagel, T.; Münstermann, S.; Kucharczyk, P.; Klingbeil, D. Review on fracture and crack propagation in weldments—A fracture mechanics perspective. Eng. Fract. Mech. 2014, 132, 200–276. [Google Scholar] [CrossRef]
  16. Tanguy, B.; Besson, J.; Piques, R.; Pineau, A. Ductile to brittle transition of an A508 steel characterized by Charpy impact test: Part I: Experimental results. Eng. Fract. Mech. 2005, 72, 49–72. [Google Scholar] [CrossRef] [Green Version]
  17. Rudland, D.; Brust, F.; Shim, D.J.; Zhang, T. Further welding residual stress and flaw tolerance assessment of dissimilar metal welds with alloy 52 inlays. In Proceedings of the ASME 2010 Pressure Vessels and Piping Division/K-PVP Conference, Bellevue, WA, USA, 18–22 July 2010. [Google Scholar] [CrossRef]
  18. Kim, J.W.; Lee, K.; Kim, J.S.; Byun, T.S. Local mechanical properties of Alloy 82/182 dissimilar weld joint between SA508 Gr.1a and F316 SS at RT and 320 °C. J. Nucl. Mater. 2009, 384, 212–221. [Google Scholar] [CrossRef]
  19. Wang, H.T.; Wang, G.Z.; Xuan, F.Z.; Liu, C.J.; Tu, S.T. Local Mechanical Properties and Microstructures of Alloy52M Dissimilar Metal Welded Joint between A508 Ferritic Steel and 316L Stainless Steel. Adv. Mater. Res. 2012, 509, 103–110. [Google Scholar] [CrossRef]
  20. Horsten, M.G.; Belcher, W.P. Fracture toughness and tensile properties of irradiated reactor pressure vessel cladding material. In Proceedings of the Radiation on Materials: 20th International Symposium, West Conshohocken, PY, USA, 12–16 January 2001. [Google Scholar]
  21. Duan, C.; Zhang, S. Two-parameter J-A estimation for weld centerline cracks of welded SE(T) specimen under tensile loading. Theor. Appl. Fract. Mech. 2020, 107, 102435. [Google Scholar] [CrossRef]
  22. Fan, K.; Wang, G.; Xuan, F.; Tu, S. Effects of work hardening mismatch on fracture resistance behavior of bi-material interface regions. Mater. Des. 2015, 68, 186–194. [Google Scholar] [CrossRef]
  23. Ding, P.; Wang, X. Solutions of the second elastic–plastic fracture mechanics parameter in test specimens under biaxial loading. Int. J. Press. Vessels Pip. 2013, 111–112, 279–294. [Google Scholar] [CrossRef]
  24. ASTM Standard E399-17. Standard Test Method for Linear-Elastic Plane Strain Fracture Toughness of Metallic Materials; Annual Book of ASTM Standards; ASTM International: West Conshohocken, PY, USA, 2017. [Google Scholar]
  25. Zhu, R.; Wang, J.; Zhang, Z.; Han, E.-H. Stress corrosion cracking of fusion boundary for 316L/52M dissimilar metal weld joints in borated and lithiated high temperature water. Corros. Sci. 2017, 120, 219–230. [Google Scholar] [CrossRef]
  26. Andresen, P.L.; Morra, M.M. Stress Corrosion Cracking of Stainless Steels and Nickel Alloys in High-Temperature Water. Corrosion 2008, 64, 15–29. [Google Scholar] [CrossRef]
  27. Jenks, A.R.; White, G.A.; Crooker, P. Crack growth rates for evaluating PWSCC of thick-wall alloy 690 material and alloy 52, 152, and variant welds. In Proceedings of the Pressure Vessels and Piping Conference, Waikoloa, HI, USA, 16–20 July 2017. [Google Scholar] [CrossRef]
  28. Lu, Z.; Shoji, T. Unified Interpretation of Crack Growth Rates of Ni-base Alloys in LWR Environments. J. Press. Vessel Technol. 2005, 128, 318–327. [Google Scholar] [CrossRef]
  29. Xue, H.; Shoji, T. Quantitative Prediction of EAC Crack Growth Rate of Sensitized Type 304 Stainless Steel in Boiling Water Reactor Environments Based on EPFEM. J. Press. Vessel Technol. 2006, 129, 460–467. [Google Scholar] [CrossRef]
  30. Lee, H.; Kim, Y.-J. Interfacial crack-tip constraints and J-integrals in plastically mismatched bi-materials. Eng. Fract. Mech. 2001, 68, 1013–1031. [Google Scholar] [CrossRef]
  31. Yang, S.; Chao, Y.; Sutton, M. Higher order asymptotic crack tip fields in a power-law hardening material. Eng. Fract. Mech. 1993, 45, 1–20. [Google Scholar] [CrossRef]
  32. Tomoyuki, F.; Keiichiro, T.; Naohiro, I.; Yoshinobu, S.; Gen, N. Derivation of J integral for evaluation of stress corrosion cracking behavior in a plastic deformation field. Corros. Eng. 2012, 61, 30. [Google Scholar] [CrossRef]
  33. Souza, R.F.; Ruggieri, C. Fracture assessments of clad pipe girth welds incorporating improved crack driving force solutions. Eng. Fract. Mech. 2015, 148, 383–405. [Google Scholar] [CrossRef]
Figure 1. Geometry of 1T-CT specimen.
Figure 1. Geometry of 1T-CT specimen.
Materials 14 04450 g001
Figure 2. Finite element model of CT specimen. (a) The whole model and (b) meshes around the crack tip.
Figure 2. Finite element model of CT specimen. (a) The whole model and (b) meshes around the crack tip.
Materials 14 04450 g002
Figure 3. Diagram of plastic zone area of the base metal and weld metal.
Figure 3. Diagram of plastic zone area of the base metal and weld metal.
Materials 14 04450 g003
Figure 4. Plastic zone area. (a) Weld metal and (b) base metal.
Figure 4. Plastic zone area. (a) Weld metal and (b) base metal.
Materials 14 04450 g004
Figure 5. The ratio between plastic zone area of the weld metal and base metal.
Figure 5. The ratio between plastic zone area of the weld metal and base metal.
Materials 14 04450 g005
Figure 6. Von Mises stress. (a) Weld metal and (b) base metal.
Figure 6. Von Mises stress. (a) Weld metal and (b) base metal.
Materials 14 04450 g006
Figure 7. The ratio between Von Mises stress in the weld metal and base metal.
Figure 7. The ratio between Von Mises stress in the weld metal and base metal.
Materials 14 04450 g007
Figure 8. J-integral with different strength and strain hardening mismatches.
Figure 8. J-integral with different strength and strain hardening mismatches.
Materials 14 04450 g008
Figure 9. Fitting function of J-integral with different mismatch factors.
Figure 9. Fitting function of J-integral with different mismatch factors.
Materials 14 04450 g009
Figure 10. Comparison between the FEA results and predicted J-integral.
Figure 10. Comparison between the FEA results and predicted J-integral.
Materials 14 04450 g010
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhang, S.; Xue, H.; Wang, S.; Sun, Y.; Yang, F.; Zhang, Y. Effect of Mechanical Heterogeneity on Strain and Stress Fields at Crack Tips of SCC in Dissimilar Metal Welded Joints. Materials 2021, 14, 4450. https://0-doi-org.brum.beds.ac.uk/10.3390/ma14164450

AMA Style

Zhang S, Xue H, Wang S, Sun Y, Yang F, Zhang Y. Effect of Mechanical Heterogeneity on Strain and Stress Fields at Crack Tips of SCC in Dissimilar Metal Welded Joints. Materials. 2021; 14(16):4450. https://0-doi-org.brum.beds.ac.uk/10.3390/ma14164450

Chicago/Turabian Style

Zhang, Shun, He Xue, Shuai Wang, Yuman Sun, Fuqiang Yang, and Yubiao Zhang. 2021. "Effect of Mechanical Heterogeneity on Strain and Stress Fields at Crack Tips of SCC in Dissimilar Metal Welded Joints" Materials 14, no. 16: 4450. https://0-doi-org.brum.beds.ac.uk/10.3390/ma14164450

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop