Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

Marine Structures 71 (2020) 102733

Contents lists available at ScienceDirect

Marine Structures
journal homepage: http://www.elsevier.com/locate/marstruc

Accurate evaluation of fracture parameters for a surface-cracked


tubular T-joint taking welding residual stress into account
Ramy Gadallah a, *, Seiichiro Tsutsumi a, Satoyuki Tanaka b, Naoki Osawa c
a
Joining and Welding Research Institute, Osaka University, Ibaraki, Osaka, Japan
b
Graduate School of Engineering, Hiroshima University, Higashi-Hiroshima, Hiroshima, Japan
c
Graduate School of Engineering, Osaka University, Suita, Osaka, Japan

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

Keywords: An efficient numerical analysis procedure based on FEM was implemented for evaluating the
Tubular joint effect of welding residual stress (WRS) on mixed-mode stress intensity factors (MM-SIFs) for a
Welding simulation non-planar surface-cracked tubular T-joint with a rounded weld toe. A multi-pass welding
Welding residual stress
simulation was carried out using real welding conditions by means of an effective in-house
Mixed-mode stress intensity factors
Surface crack
welding simulation code. A fully compatible 3D finite element mesh generation system was
Crack face tractions implemented to model non-planar perpendicular surface-cracked tubular T-joints with rounded
weld toes. The adequacy of the mesh generation system was verified with the reference solutions.
A fully automated mapping code was developed to assign the six components of the simulated
WRS to each element face over the crack surface. A developed fracture approach was employed to
accurately compute MM-SIFs resulting from WRS in which normal and shear components of WRS
were considered. MM-SIFs were calculated carefully for several surface cracks with different
shapes under different loading conditions. The behavior of MM-SIF solutions resulting from the
combination of external load and WRS was influenced significantly by the behavior of those
produced due to WRS only. In addition, the value of the computed SIFs resulting from the
combination of external load and WRS was almost doubled for mode-I SIF compared to those
obtained by external load only and WRS only especially near the crack end locations. The
behavior of the calculated fracture parameters was also discussed.

1. Introduction

Steel tubular joints are commonly used in offshore and onshore structural members such as tubular trusses, bridges and
communication towers. These tubular structural members are joined by the welding process, which results in very complicated ge­
ometries at the connecting locations. In addition, complex stress distributions are generated around the welded joints due to the
induced welding residual stress (WRS) and applied external loads. Surface defects, e.g. cracks, are often found around welded joints in
tubular structures, which can propagate and lead to premature failure if the structure is repeatedly subjected to external loads.
Tubular T-joints are critical spots in structures, especially around welded joints. An unavoidable WRS is produced due to welding,
which its distribution is very complicated as it is influenced by the joint geometry and applied welding conditions. The presence of a
crack in a WRS field may have a detrimental effect on structural integrity where WRS is one of the major factors that accelerates the

* Corresponding author.
E-mail address: ramy@jwri.osaka-u.ac.jp (R. Gadallah).

https://doi.org/10.1016/j.marstruc.2020.102733
Received 25 June 2019; Received in revised form 22 October 2019; Accepted 30 January 2020
Available online 10 February 2020
0951-8339/© 2020 Elsevier Ltd. All rights reserved.
R. Gadallah et al. Marine Structures 71 (2020) 102733

Nomenclature

a crack depth
ASFs arbitrary stress fields
BC boundary condition
c crack half-length
d brace diameter
D chord diameter
KI, KII, KIII stress intensity factors for mode-I, mode-II and mode-III, respectively
L chord length
t brace wall thickness
T chord wall thickness
x, y, z Cartesian coordinates
YI , YII , YIII normalized stress intensity factors for mode-I, mode-II and mode-III, respectively
θ the angle between the brace and chord axes
ρ’ weld toe radius
σ nom nominal stress
ϕ crack eccentric angle
ψ flank angle
CFT crack face traction
DIM domain integral method
FE finite element
FEA finite element analysis
FEM finite element method
i-ISM inherent strain-based iterative substructure method
IIM interaction integral method
ISM iterative substructure method
MM-SIFs mixed-mode stress intensity factors
SIFs stress intensity factors
TEP thermo-elastic-plastic
WPS welding procedure specification
WRS welding residual stress

crack propagation rate and concerns with fatigue strength and buckling strength [1–4]. It is, therefore, necessary to evaluate WRS
distributions rigorously for tubular T-joints. Welding is considered a complex process, so efficient methods are needed to predict WRS
precisely. The finite element method (FEM) has considered a robust tool for predicting WRS accurately. A large number of welding
simulations have been performed by means of FEM, until now, for different welded joints to predict WRS [5–10].
Cracks frequently initiate at stress concentration spots such as weld toes. For welded joints, the situation becomes more compli­
cated when cracks initiate at high-stress concentration regions in addition to their presence in large WRS fields. Therefore, stress
intensity factors (SIFs) should be calculated rigorously to obtain a reliable fatigue crack propagation rate. The accuracy of SIFs can be
improved by including the effect of WRS in SIF calculation methods [11–14]. To estimate SIFs accurately, the superposition method is
considered an efficient tool for cracks in arbitrary stress fields (ASFs). Therefore, to improve the accuracy of SIFs taking WRS into
account, the applied numerical integration method should include the crack face traction (CFT) integral [11,12]. There are different
methods are available to calculate SIFs including the effect of WRS based on the FEM or the weight function method [15,16]. Other
promising methods such as the extended FEM [17], the meshfree method [18–20] and the wavelet Galerkin method [21,22] are also
used to estimate SIFs. Several methods are used to calculate SIFs by means of FEM such as the domain integral method (DIM) [23], the
virtual crack closure method [24] and the interaction integral method (IIM) [25].
The IIM has emerged an efficient method and the most accurate and readily technique to extract mixed-mode SIFs (MM-SIFs) [25].
The IIM was derived by means of the DIM by combining two states, namely, actual fields and known auxiliary fields [26,27]. The IIM
was then developed to extract MM-SIFs along planar interface cracks and curved crack tips in 3Ds [23,28]. Walter et al. have developed
interaction integral procedures to calculate SIFs for 3D curved cracks in stress fields where CFT is taken into account [25].
To date, very limited research works have investigated MM-SIFs resulting from WRS for surface-cracked tubular joints by means of
the fracture mechanics approach. Most of the research works that performed on tubular joints have examined either tubular joints
subjected to external loads without taking WRS into account [29–37] or WRS determination (i.e. simulation or measurement) of
tubular joints [38–40]. Recently, Chang et al. [41] have performed fatigue finite element analysis (FEA) for welded tubular joints by
means of a continuum damage mechanics approach. However, in their work, they did not discuss the fracture behavior of the examined
tubular joints. For complicated welded joints with complex curved cracks subjected to arbitrary external loads (e.g. tubular joints)
normal and shear components of WRS may have a considerable effect on fracture behavior and consequently on fatigue crack
propagation rate. It is, therefore, necessary to include the six components of WRS in the applied numerical integration method in order

2
R. Gadallah et al. Marine Structures 71 (2020) 102733

Fig. 1. Schematic overview of the tubular T-joint [34].

to improve the accuracy of MM-SIF solutions for such complicated geometries. For this reason, it is needed to implement a numerical
analysis procedure to accurately evaluate the effect of WRS on fracture parameters (i.e. MM-SIFs).
This paper aims to implement a numerical analysis procedure based on FEM to evaluate the effect of WRS on fracture parameters
for a non-planar surface-cracked tubular T-joint with a rounded weld toe. Firstly, a multi-pass thermo-elastic-plastic (TEP) FEA is
performed for uncracked tubular T-joint based on real welding conditions by means of an effective welding simulation in-house code.
The induced WRS distributions are then verified with reference solutions taken from the literature. A fully compatible 3D finite
element (FE) mesh generation system is implemented to generate surface-cracked tubular T-joints with rounded weld toes. MM-SIF
solutions resulting from WRS in addition to external load are calculated by means of an efficient fracture approach [11] modified
by the authors. WARP3D [42], an open multifunctional code, is used to calculate MM-SIFs. The IIM implemented in WARP3D
(WARP3D-IIM) is adopted to calculate MM-SIFs. Using the modified WARP3D-IIM, the six components of WRS can be considered in the
computed MM-SIFs. A fully automated mapping code is developed to assign the six components of the simulated WRS to the corre­
sponding elements over the crack surface based on the superposition method. Before examining the effect of WRS on the behavior of
MM-SIFs, the implemented mesh generation system is verified with the reference solutions. MM-SIF solutions are then calculated
under different loading conditions including external load only, WRS only and a combination of external load and WRS for several
surface cracks with different shapes. Fracture behavior resulting from WRS is also discussed.

2. Prediction of WRS

2.1. Welding simulation

Welding is considered a complex process, so an effective welding simulation tool is needed to predict WRS accurately. So that,
JWRIAN (Joining and Welding Research Institute ANalysis) was adopted to perform welding simulation. JWRIAN is a robust in-house
code for welding simulation based on FEM in which it provides a consecutive coupled thermal and mechanical FEA. Algorithms are
incorporated in this code to conduct consecutive analyses for computing temperature profile, WRS and welding deformation in welded
structures [43]. The Iterative Substructure Method (ISM) [44] is implemented in this code to reduce the computation time of welding
simulation. The ISM accelerates the computation time by partitioning the FE model into two regions, namely, a nonlinear region that is
exposed to the heat source and linear region, which symbolizes the remaining part of the FE model. In addition, an iterative method is
implemented in JWRIAN to guarantee the continuity of tractions between the two regions. The ISM is efficient to solve large welding
problems with a number of elements less than one million. To solve large-scale welding problems (i.e. more than one million elements),
another approach called the inherent strain-based Iterative Substructure Method (i-ISM) was implemented in this code [43]. Using the
i-ISM, the induced inherent strain varies only in the surrounding region to the heat source in which it is not influenced by the defined
constraints away from the heat source. To reduce the computation time when solving large-scale welding problems based on the i-ISM,
the size of the region close to that exposed to the heat source should be small. This is carried out by relaxing the constraints given to the
region away from the heated region. For more details on the theories implemented in JWRIAN refer to Refs. [43–45]. JWRIAN
conducts TEP FEA using consecutively coupled formulations where thermal analysis is performed firstly to predict temperature his­
tories. Mechanical analysis is then conducted in which the computed temperature histories are used to predict displacement, strain,
and stress. JWRIAN has been widely used for predicting WRS and welding deformation accurately for a large number of engineering
problems [5,11,12,43–49].

2.2. Geometry and TEP FE model

A TEP FEA was performed for a steel tubular T-joint based on real welding conditions. The geometry of the applied tubular T-joint is

3
R. Gadallah et al. Marine Structures 71 (2020) 102733

Fig. 2. TEP FE model of a tubular T-joint. (a) typical one-half symmetric FE model, (b) close-up view of the saddle location, (c) x-z cross-section of
the saddle location, (d) mesh of weld passes at the crown location.

Table 1
Applied welding conditions based on welding specifications given in the WPS.
Pass No. Welding current (A) Arc voltage (V) Travel speed (mm/min) Heat input (kJ/mm)

1 166.5 24 300.0 0.6


2 193.5 24 225.5 1.0
3 188.5 24 200.5 1.1
4 194.5 24 230.5 1.0
5 197.5 24 325.5 0.7
6 184.5 24 349.5 0.6
7 167.5 24 301.5 0.6
8 170.0 24 376.5 0.5
9 167.5 24 261.5 0.7

presented in Fig. 1. The tubular T-joint is composed of a brace connected to a chord by a weld. The chord weld toe has a curvature
radius of ρ’ ¼ 3 mm [34]. The brace has an outer diameter of 139.7 mm with a wall thickness of 5 mm, while the chord has an outer
diameter of 406.4 mm with a wall thickness of 12.7 mm. The chemical composition and mechanical properties of the applied steel are
reported in Ref. [34]. The brace was connected with the chord using the flux-cored arc welding process. A welding wire with a
classification of AWS A5.20 (E71T-1M H8) [50] was adopted in which a shielding gas of Argon þ 20–25% CO2 was used. Two different
FE models were applied in this study. The first model was used in the TEP FEA to compute WRS, which is called TEP FE model while the
other model was employed in the fracture analysis, which is called cracked FE model.
A typical one-half TEP FE model was generated using 8-node hexahedral elements and symmetric boundary conditions (BCs) were
applied. The TEP FE model is illustrated in Fig. 2(a). To prevent the rigid body motion during TEP FEA a node at one end of the chord
member was fixed in the z-direction (i.e. Uz ¼ 0) and another node at the other end of the chord member was fixed in y- and z-directions
(i.e. Uy ¼ Uz ¼ 0). The TEP FE model does not include crack (i.e. uncracked FE model). The region from the chord weld toe to 1.75 mm
in the chord thickness was modeled with a very fine mesh; see Fig. 2(c). This is to overcome the singularity when mapping the
simulated WRS to the cracked FE model [12]. A fine mesh was used for the weld mesh especially around the saddle location in order to
improve the accuracy of simulated WRS (see Fig. 2(b)). The TEP FE model consists of 165,765 nodes and 152,134 elements. The mesh
of weld passes is shown in Fig. 2(d). The longitudinal chord and brace directions are represented by the y- and z-directions,
respectively.
A nine-pass welding simulation was carried out around the brace-chord intersection. Although the mesh of the nine weld passes was
generated before carrying out TEP FEA, the weld passes were treated as dummy passes. When the first pass, for example, receives its
heat input the other passes are treated as dummy passes, etc. The applied welding conditions of each pass are given in Table 1 in which
an arc efficiency of 80% was used. The applied welding conditions are based on welding specifications given in the welding procedure
specification (WPS). An inter-pass temperature of 200 � C was considered in the thermal analysis as recommended in the WPS. A
tubular heat source was used in welding simulation to ensure that each pass receives the applied heat input. Since the surface crack was

4
R. Gadallah et al. Marine Structures 71 (2020) 102733

Fig. 3. Applied temperature-dependent material properties [5,51]. (a) temperature-dependent thermo-physical properties, (b)
temperature-dependent mechanical properties.

Fig. 4. Axial WRS distributions at the outer and inner surfaces of the chord member. (a) current welding simulation, (b) reference [38].

initiated at the saddle location [34], the heat source should not start or end at this location to avoid the resulting problems in the
induced WRS. For this reason, the heat source started from the negative y-direction and ended at the positive y-direction passing by the
saddle location (see Fig. 1). The applied nine-passes have the same welding sequence as it is illustrated in Fig. 1. The performed TEP
FEA represents ‘as-weld’ condition where there was no heat treatment performed after welding in Ref. [34]. The
temperature-dependent material properties used in TEP FEA are shown in Fig. 3. The applied material properties in Fig. 3 represent
properties for steel materials that are close to that used in Ref. [34].

2.3. Welding simulation results

Due to the lack of measured and simulated WRS for the employed tubular T-joint, the tendency of the simulated WRS predicted by
the current welding simulation was verified with those obtained by similar target tubular T-joints. To verify the accuracy of the current
welding simulation, WRS distributions at the outer and inner surfaces of the chord and brace members are compared with those
predicted by numerical analysis reported in Ref. [38]. For the sake of verification, two examples are presented in Figs. 4 and 5. Fig. 4
illustrates the axial WRS distributions of the chord member obtained by the current welding simulation and those given by Ref. [38].
While Fig. 5 shows the radial WRS distributions of the brace member predicted by the current welding simulation and those obtained
by Ref. [38]. From Fig. 4(a) it is noticed that, at the outer surface, a large tensile WRS was produced at the chord weld toe followed by a
compressive one in the region near the chord weld toe while a negligible WRS was observed at the remaining region. On the other
hand, WRS induced at the inner surface has a contradictory behavior compared to that obtained at the outer surface especially in the
region near the chord weld toe; see Fig. 4(a). It is also noticed that some differences are apparent in the magnitudes of WRS at the weld

5
R. Gadallah et al. Marine Structures 71 (2020) 102733

Fig. 5. Radial WRS distributions at the outer and inner surfaces of the brace member. (a) current welding simulation, (b) reference [38].

Fig. 6. Verification of the tendency of the simulated WRS with experimental measurements.

toe location; this may be due to the difference of the strength grade of the applied steel in this study and that used in Ref. [38]. Further,
a small radial WRS was obtained near the brace weld toe at the outer and inner surfaces for a short distance from the brace weld toe
followed by a negligible WRS as shown in Fig. 5(a). Although the difference of the weld geometry, the number of weld passes,
employed temperature-dependent material properties and applied heat input that used in the current welding simulation and those
adopted in Ref. [38], the comparisons illustrated in Figs. 4 and 5 indicate that the general characteristics of the simulated WRS in this
study are in good agreement with those reported in Ref. [38].
Moreover, the tendency of the simulated WRS was verified with experimental measurements reported in Ref. [39]. For the sake of
verification, two examples are illustrated in Fig. 6. WRS results, shown in Fig. 6, are taken at particular locations near the weld (i.e.
point 3 and point 6) according to Ref. [39]. The results represent WRS distribution into depth for the axial direction on the chord
member and the circumferential direction in the brace member. The simulated WRS has a good tendency with experiments. Some
differences were observed between the simulated WRS and experiments as the depth increases where this may be due to the difference
of the geometry of the applied tubular T-joint, the employed steel grade, and welding conditions. In general, the accuracy of the
simulated WRS predicted by the current welding simulation was confirmed with those obtained numerically (i.e. Figs. 4 and 5) and
experimentally (i.e. Fig. 6) for similar target tubular T-joints.
Fig. 7 illustrates the maximum temperature distribution generated after the last weld pass. A uniform temperature distribution, as
well as full penetration welding, were obtained due to the applied heat input. Further, 2D cross-sectional maps of the induced WRS are
presented in Fig. 8. WRS distributions illustrated in Fig. 8 are taken at the chord weld toe where this is the location of surface cracks.
Surface cracks are placed at the saddle location through the chord thickness as it is reported in Ref. [34]. It is, therefore, necessary to
examine the distribution of the induced WRS through the chord thickness. The results presented in Fig. 8 demonstrate that typical WRS

6
R. Gadallah et al. Marine Structures 71 (2020) 102733

Fig. 7. Maximum temperature distribution after the last weld pass.

Fig. 8. 2D maps of typical WRS distributions. (a) von Mises stress, (b) σxx, (c) σyy, (d) σzz.

distributions were obtained under the applied welding conditions. It is noticed that a large tensile WRS was induced near the chord
weld toe especially around the saddle location (see Fig. 8). Fig. 8(a) reveals that the maximum von Mises stress was produced at the
weld zone. It is also observed that the produced WRS presented in Fig. 8(b)–(d) gives a typical WRS distribution in which a tensile WRS
is induced at the chord upper surface (i.e. chord weld toe) while compressive one is produced at the chord lower surface. The variation
of WRS distribution through the chord thickness may have a considerable influence on the behavior of fracture parameters, as it will be
discussed later. Due to the applied multi-passes, a maximum deformation of 3.59 mm was obtained along the longitudinal brace
direction.

7
R. Gadallah et al. Marine Structures 71 (2020) 102733

Fig. 9. Procedure to generate a surface-cracked tubular T-joint with a rounded weld toe. (a) surface-cracked flat plate mesh generated by FEAcrack
[61], (b) surface-cracked T-butt mesh with a rounded weld toe, (c) mapping-1 to generate the brace member, (d) building a partial mesh using MSC
Patran [62], (e) mapping-2 to generate the chord member, (f) building a partial mesh using MSC Patran [62].

3. Fracture analysis

3.1. WARP3D-IIM

WARP3D is an open-source code used for nonlinear analysis of structures and solids under static, dynamic and thermal loadings
[42]. WARP3D code has been employed in several research works [11–14,25,37,52–54] to estimate fracture parameters for different
engineering problems. The modified WARP3D-IIM [11] which considers the effect of the six components of an ASF (i.e. WRS) was used
to calculate MM-SIFs. The IIM provides actual fields such as displacement, strain, and stress as well as post-processes auxiliary fields (e.
g. MM-SIFs and T-stresses [55,56]). The domain of interaction integral that contains the actual and auxiliary terms, IðsÞ, is defined as
[25]:
Z � � Z h � � i Z
IðsÞ ¼ σij uaux
j;1 þ σ aux
ij uj;1 σ jk ε aux
jk δ1i q;i dV þ σ ij uaux
j;1i ε aux
ij;1 þ σ aux
ij;1 uj;1 qdV tj uaux
j;1 qdS; (1)
V V Sþ þS

where the superscript ‘aux’ denotes the auxiliary field components, uj represents displacement components, δij is the Kronecker delta, V
and S are finite volume and finite surface respectively, q is weight function, and tj represents CFT components. Sþ and S are the upper
R
and lower crack surfaces, respectively. The CFT-integral is represented by the third integral, tj uaux
j;1 qdS, in Eq. (1). The efficiency of
Sþ þS
this integral in improving the accuracy of SIFs was demonstrated in Refs. [11,13,14,25].
R
Using the value of IðsÞ calculated in Eq. (1), the interaction integral, IðsÞ, is given as; IðsÞ ¼ IðsÞ= qðsÞds [25]. Where LC is the
LC
segment length of the crack front. MM-SIF solutions (KI, KII and KIII) are then calculated using the value of IðsÞ as [25,42]; KI ¼ E* IðsÞ=
2, KII ¼ E* IðsÞ=2 and KIII ¼ E* IðsÞ=2ð1 þ νÞ. The definition of E* is given as, E* ¼ E for plane stress condition whereas E* ¼ E= ð1 ν2 Þ
for plane strain condition. Where E is Young’s modulus and ν represents Poisson’s ratio.
To estimate MM-SIF solutions accurately for cracks in ASFs, the CFT-integral should be considered [11,13,14,25]. To consider the
effect of an ASF in MM-SIF solutions stress field can be applied as tractions over the crack surface. Thus for linear elastic problems, the
superposition method has emerged a powerful approach for cracks in ASFs. This means that if the CFT-integral is included the accuracy
of the computed MM-SIFs can be improved considerably.
On the other hand, most FE commercial fracture programs do not consider the CFT-integral. Further, WARP3D provides this in­
tegral in the implemented IIM. However, this integral is only available for a uniform stress field applied to the crack surface. Therefore,
the CFT-integral provided by WARP3D-IIM in its current release cannot be used for cracks in ASFs (e.g. WRS field). Moreover, for
cracks in complicated welded joints as well as non-planar crack geometries the possibility of obtaining MM-SIFs is high. For this reason,

8
R. Gadallah et al. Marine Structures 71 (2020) 102733

Table 2
Details of the employed cracked FE tubular T-joints [63].
Geometry Tubular joint geometry Weld geometry Crack size

α β γ τ θ [ o] ρ0 ψ [o] a/T a/c

T1 14 0.5 18 0.6 90 0.0 43 0.5 0.4


T4 14 0.5 8 0.6 90 0.0 43 0.5 0.4

Fig. 10. Typical one-quarter cracked FE model of a tubular T-joint with loading and BCs.

Gadallah et al. [11] have modified the CFT-integral available in WARP3D-IIM in which stress components are applied as traction force
vectors to each element face center over the crack surface. Therefore, the modified WARP3D-IIM considers the effect of the six
components of an ASF in the calculated MM-SIF solutions. Where the normal and shear components of an ASF are considered in the
CFT-integral, which contribute to improving the accuracy of IðsÞ defined in Eq. (1). The efficiency of the modified WARP3D-IIM was
demonstrated by different numerical examples [11].

3.2. Mesh generation system of a surface-cracked tubular T-joint with a rounded weld toe

Full 3D FE meshing techniques for cracks at weld toes are very few where some research works [1,57,58] have reported simplified
approaches. Bowness and Lee [59] have reported a detailed procedure to model cracked tubular joints in which cracks are placed at
weld toes using 20-node hexahedral elements. In addition, Cao et al. [60] have proposed a similar approach to generate a cracked
tubular Y-joint with a surface crack placed at the weld toe. However, in the aforementioned research works, a sharp weld toe (i.e.
unrounded weld toe) was modeled.
In this study, a fully compatible 3D FE mesh generation system was implemented to model a non-planar perpendicular surface crack
with a rounded weld toe using 20-node hexahedral elements. The mesh generation system generally follows the procedure reported by
Bowness and Lee [59]. The implemented system is efficient to generate surface-cracked tubular T-joints with sharp and rounded weld
toes. The implemented system generates surface-cracked tubular T-joints with rounded weld toes based on a surface-cracked flat plate.
The basic cracked mesh which represents a surface-cracked flat plate is generated by FEAcrack software [61] (see Fig. 9(a)). A T-butt
mesh with a rounded weld toe is then generated by reshaping the surface-cracked flat plate mesh to generate a fillet weld with rounded
weld toe. The attachment of the T-butt mesh is built using MSC Patran [62] (see Fig. 9(b)). To construct the tubular T-joint, two
mapping techniques, namely, mapping-1 and mapping-2 were developed and were applied in the implemented mesh generation
system. Mapping-1 is used to transform the T-butt mesh into a brace member (see Fig. 9(c)). A partial mesh is added to the brace
member using MSC Patran [62] as a preparation step to generate the chord member (see Fig. 9(d)). Mapping-2 is then applied to
generate the chord member (see Fig. 9(e)). A partial mesh is added to the chord member using MSC Patran [62] to construct the target
surface-cracked tubular T-joint (see Fig. 9(f)).

9
R. Gadallah et al. Marine Structures 71 (2020) 102733

Fig. 11. The von Mises stress distribution of the deformed cracks. (a) geometry T1, (b) geometry T4.

Fig. 12. Verification of the implemented mesh generation system. (a) geometry T1, (b) geometry T4.

3.3. Verification numerical examples

3.3.1. FE model
To demonstrate the efficiency of the implemented mesh generation system two numerical examples, taken from Ref. [63], were
investigated. The details of the applied tubular joints are given in Table 2. Where α ¼ 2L=D, β ¼ d=D, γ ¼ D=2T, τ ¼ t= T. Due to the
symmetries of the tubular joint and loading condition, a typical one-quarter cracked FE model was used. The cracked FE model and
applied loading and BCs are presented in Fig. 10. The chord end was fixed as reported in Ref. [34]. Semi-elliptical surface-cracked FE
tubular T-joints were generated with 20-node hexahedral elements. A very fine mesh along the crack front was used. An axial tensile
load was applied to the top surface of the brace member.
Due to the lack of reference solutions that investigated surface-cracked tubular T-joints with rounded weld toes, cracked FE tubular
T-joints with sharp weld toes were used in this section to be compatible with the selected reference solutions. The linear elastic analysis
was applied to calculate MM-SIFs for the employed cracked FE tubular T-joints. WARP3D-IIM was used to calculate MM-SIFs in which
the problem was treated as a remote loading problem (i.e. no traction forces were applied over the crack surface). Young’s modulus of
E ¼ 203 GPa and Poisson’s ratio of ν ¼ 0.3 were used where they represent typical properties of steel materials. The material properties

10
R. Gadallah et al. Marine Structures 71 (2020) 102733

Table 3
Details of the employed tubular T-joint.
Model Tubular joint geometry Weld geometry

α β γ τ θ [o] L/T ρ’ [mm] ψ [o]


T-1–T-4 6.8 0.3 16 0.4 90 1.18 3.0 45

Table 4
Details of the applied cracked FE models.
Model Crack shape Crack size

a/T a/c

T-1 shallow crack 0.25 0.30


T-2 elongated-shallow crack 0.39 0.25
T-3 half-thickness crack 0.50 0.49
T-4 deep crack 0.88 0.69

Fig. 13. Surface-cracked FE model of a tubular T-joint with a rounded weld toe. (a) typical one-half symmetric FE model with applied loading and
BCs, (b) x-z cross-section of the saddle location, (c) FE mesh of the applied cracks.

adopted in the fracture analysis emulate those used in the welding simulation at room temperature. The material properties used in this
section were also adopted in the following fracture analyses.

3.3.2. MM-SIFs results


Fig. 11(a) and (b) presents the von Mises stress distribution of the deformed cracks for geometries T1 and T4, respectively. Due to
the difference in the applied tubular T-joint geometry (i.e. γ-value), a pronounced difference was observed in the induced von Mises
stress as well as produced deformation for geometries T1 and T4. As geometry T1 has a larger γ-value compared to that of geometry T4
the induced von Mises stress and produced deformation of geometry T1 are clearly larger than that obtained by geometry T4. It is also
noticed that the induced stress at the crack end location is higher than that generated at the crack deepest point where this may in­
fluence the behavior of SIF solutions at these locations. The calculated SIF solutions by means of WARP3D-IIM were validated with two
well-established reference solutions. The first reference solution [63] represents a numerical solution while the other one [64] is an
analytical solution which developed to determine SIF solutions of welded tubular joints with surface cracks placed at any location
pffiffiffiffiffi pffiffiffiffiffi pffiffiffiffiffi
around the weld. The calculated MM-SIFs are normalized as; YI ¼ KI =σnom πa, YII ¼ KII =σnom πa and YIII ¼ KIII =σnom πa. Where σnom
represents applied nominal stress. The behavior of MM-SIF solutions calculated by WARP3D-IIM of the two numerical examples is
illustrated in Fig. 12. The crack end and crack deepest point locations are given at 2ϕ/π ¼ 0 and 2ϕ/π ¼ 1, respectively. The MM-SIF

11
R. Gadallah et al. Marine Structures 71 (2020) 102733

Fig. 14. Distribution of MM-SIFs along the crack front for a surface-cracked tubular T-joint FE model subjected to external tensile load. (a) model T-
1, (b) model T-2, (c) model T-3, (d) model T-4.

solutions were treated as plane stress condition at the crack end location while as plane strain condition at the element corner nodes
along the crack front except the crack end location. It is noticed that the value of YI at the crack end location is larger than that obtained
at the crack deepest point where this is due to the induced stress field at these locations (see Fig. 11). Since the reference solutions
provide only the mode-I solution (i.e. KI value), so the value of YI calculated by WARP3D-IIM was verified with those given by the
reference solutions (see Fig. 12). The results illustrated in Fig. 12 reveal good agreement between the value of YI obtained by
WARP3D-IIM and those estimated by the reference solutions, especially at the crack deepest point location. The reason why
WARP3D-IIM gave a larger value of YI at the crack end location compared to those obtained by the reference solutions is because of the
IIM that implements a simplified assumption that does not consider the stress singularity at notches [13]. The restraining of the
attachment has also effect in which it reduces the singularity from r0.5 to about r0.4 as reported in Ref. [65]. Therefore, stress sin­
gularities generated due to the previously mentioned reasons were superimposed and gave this difference at the crack end location. In
general, the calculated SIFs by WARP3D-IIM using the implemented mesh generation system give a good matching with those obtained
by the reference solutions, which demonstrate the adequacy of the implemented mesh generation system.

4. Evaluation of MM-SIFs for surface cracks in WRS fields

4.1. Applied cracked FE models

Since non-uniform WRS fields were induced around the saddle location and through the chord thickness a one-half symmetric
cracked FE model with a rounded weld toe was employed to evaluate the effect of WRS on MM-SIFs. The same geometry of the tubular
T-joint shown in Fig. 1 was utilized in which semi-elliptical surface cracks were placed at the saddle weld toe location. Generally,

12
R. Gadallah et al. Marine Structures 71 (2020) 102733

Fig. 15. Flowchart of the developed automated mapping code.

fatigue cracks in ASFs (e.g. WRS field) grow with an irregular shape [13]. For the sake of simplicity, in this study, semi-elliptical surface
cracks were used. Four-cracked FE models, namely, T-1, T-2, T-3 and T-4 with different crack shapes were investigated. The details of
the employed tubular T-joint are given in Table 3. The details of the applied cracked FE models are given in Table 4. A typical one-half
cracked tubular T-joint FE model was generated and symmetric BCs were applied. The two ends of the chord member were fixed. The
cracked FE model and applied loading and BCs are illustrated in Fig. 13(a) and (b). The FE mesh of the applied cracks is shown in
Fig. 13(c). A very fine mesh was generated along the crack front. The cracked FE model was generated with 20-node hexahedral
elements. An axial tensile load of 161.5 kN was applied to the top surface of the brace member. The same material properties used in
Section 3.3.1 were adopted in this section, as well.

4.2. Calculation of MM-SIFs under external tensile load

Before examining the effect of the simulated WRS on the behavior of MM-SIFs, MM-SIF solutions were evaluated firstly under
external mechanical load. Fig. 14 illustrates the distribution of MM-SIF solutions of the applied cracked FE models. The crack end
locations are given at 2ϕ/π ¼ 0 and 2ϕ/π ¼ 2 while the crack deepest point location is located at 2ϕ/π ¼ 1. Due to the complicated
welded geometry as well as the applied non-planar cracks, MM-SIFs were produced. However, it is noticed that mode-I SIF (i.e. KI
value) is the dominant mode. It is also observed that as the crack depth to chord thickness ratio (a/T) increases the behavior of KI
decreases by approaching the crack deepest point location while the behavior of KII changes from a plateau shape to a convex shape
(see Fig. 14(a)–(d)). As for KIII, there is no considerable influence was observed except that the value of KIII near the crack end locations
increases as a/T increases. For the shallow crack and elongated-shallow crack (i.e. models T-1 and T-2), a plateau KI behavior was

13
R. Gadallah et al. Marine Structures 71 (2020) 102733

Fig. 16. 2D map of WRS distribution in x-direction including the applied surface cracks (T-1–T-4) placed at the saddle weld toe.

Fig. 17. Mapping of the simulated WRS to a crack surface. (a) TEP FE mesh at the saddle location, (b) crack FE mesh of model T-4, (c) WRS
distribution in x-direction through the chord thickness, (d) mapped WRS over the crack surface.

obtained especially for model T-2 which emulates the real fatigue crack shape, see Fig. 14(a) and (b). Further, it is noticed that the
value of KI at the crack end locations is higher than that obtained at the crack deepest point location, as it is illustrated in Fig. 14(a)–(d).
This means that, under the employed loading and BCs, the surface crack placed at the saddle weld toe will propagate faster from the
crack end locations compared to the crack deepest location.

4.3. Mapping of the simulated WRS to a crack surface

As mentioned before two FE models were employed where the first model was used for performing the TEP FEA while the other
model was used for conducting the fracture analysis. To evaluate the effect of the simulated WRS on the behavior of MM-SIFs, the six
components of the simulated WRS should be included in the fracture analysis. In the modified WARP3D-IIM [11], the six components
of an ASF can be applied as traction forces over the crack surface. Due to the complexity of the crack geometry (i.e. non-planar crack), a
fully automated mapping code was developed to assign the six components of the simulated WRS to each element face center over the
crack surface based on the superposition method. With the information of the TEP FE model and cracked FE model as well as simulated
WRS components, each element face center over the crack surface can be assigned six components of the simulated WRS in a short
time. A flowchart of the developed mapping code is illustrated in Fig. 15.
Fig. 16 shows the distribution of WRS in x-direction taken at the chord weld toe. The employed surface cracks are placed at the
saddle location. The applied cracks mentioned in Table 4 are plotted over the WRS 2D map with the actual size where it reveals the
variation of the WRS field over each crack surface. From Fig. 16, it is clear that a non-uniform WRS field is applied to each crack
surface. Fig. 17 illustrates an example of WRS mapping from the TEP FE mesh to the corresponding crack FE mesh. The crack FE mesh

14
R. Gadallah et al. Marine Structures 71 (2020) 102733

Fig. 18. 2D maps of the mapped WRS over the crack surface of model T-1. (a) normal stress components, (b) shear stress components.

Fig. 19. 2D maps of the mapped WRS over the crack surface of model T-2. (a) normal stress components, (b) shear stress components.

of model T-4 was selected to show the variation of the mapped WRS over the crack surface. The mesh of the TEP FE model, as well as
crack FE mesh of model T-4, are shown in Fig. 17(a) and (b), respectively. Fig. 17(c) presents WRS distribution in x-direction through
the chord thickness. The mapped WRS over the crack surface of model T-4 is illustrated in Fig. 17(d). From Fig. 17, it is clear that the
mapped WRS on the crack surface is close to that predicted by TEP FEA. The developed mapping code was used to map the normal and
shear components of the simulated WRS for the applied cracks in which their effect was considered in the calculated MM-SIFs.

4.4. Influence of WRS on the behavior of MM-SIFs

The adequacy of the modified WARP3D-IIM was verified including WRS induced by different in-plane weld lines placed on flat
plates by means of JWRIAN [11]. However, the modified WARP3D-IIM was not applied yet for practical engineering problems (i.e.
complicated welded joints). Cracked tubular T-joints represent a good example to demonstrate the efficiency of the modified
WARP3D-IIM. Therefore, an efficient numerical analysis procedure was implemented to evaluate the effect of WRS on MM-SIFs for
surface-cracked tubular T-joints with rounded weld toes.
Using the developed mapping code, the simulated WRS was mapped from the TEP FE model to the cracked FE model. Figs. 18–21
show 2D maps of the simulated WRS components to the corresponding crack surfaces. The mapped WRS distributions presented in
Figs. 18–21 are based on the global coordinate system. It is noticed that the normal components of WRS are more dominate than shear

15
R. Gadallah et al. Marine Structures 71 (2020) 102733

Fig. 20. 2D maps of the mapped WRS over the crack surface of model T-3. (a) normal stress components, (b) shear stress components.

Fig. 21. 2D maps of the mapped WRS over the crack surface of model T-4. (a) normal stress components, (b) shear stress components.

components. However, the values of the shear components are not small. It is also observed that the surface cracks of models T-1–T-3
are placed in tensile normal stress fields, see Figs. 18(a)-20(a). As for the deep crack (i.e. model T-4), it is noticed that the crack deepest
point exists in compressive normal stress fields (see Fig. 21(a)). However, the compressive stress field σxx, which surrounds the deepest
point, is more dominant than σyy and σzz.
Figs. 22–25 illustrate the distributions of MM-SIFs along the crack front including the effect of WRS. The crack end locations are
placed at 2ϕ/π ¼ 0 and 2ϕ/π ¼ 2 while the crack deepest point is given at 2ϕ/π ¼ 1. The results presented in Figs. 22(a)-25(a) depict the
behavior of MM-SIFs resulting from WRS only. The behavior of MM-SIFs resulting from the combination of external tensile load and
WRS is shown in Figs. 22(b)-25(b). It is noticed that when WRS is considered the behavior of MM-SIFs is influenced significantly, see
Figs. 22(a)-25(a).
Due to the consideration of WRS, non-uniform distributions of MM-SIFs were obtained. From Figs 22(a)-25(a), it is observed that KI
is the dominant mode compared to KII and KIII. However, the behavior of KII and KIII is not trivial. For models T-1–T-3, since the cracks
are placed in a tensile normal WRS field (see Figs. 18(a)-20(a)) a positive KI value along the crack front was obtained; see Figs. 22(a)-24
(a). Whereas for the deep crack (model T-4) a negative KI value was observed in the region around the crack deepest point (see Fig. 25

16
R. Gadallah et al. Marine Structures 71 (2020) 102733

Fig. 22. Distribution of MM-SIFs along the crack front for model T-1 under (a) WRS only, (b) a combination of external load and WRS.

Fig. 23. Distribution of MM-SIFs along the crack front for model T-2 under (a) WRS only, (b) a combination of external load and WRS.

(a)). This is due to the existence of a part of the deep crack (i.e. the region around the crack deepest point) in a compressive normal
WRS field; see Fig. 21(a). This demonstrates that the implemented numerical analysis procedure is valid and efficient for any crack
shape in a WRS field in which MM-SIFs for cracks in compressive WRS fields can be calculated efficiently.
On the other hand, the results illustrated in Figs. 22(b)-25(b) present the behavior of MM-SIFs resulting from the external load with
consideration of the effect of WRS. It is observed that the behavior of MM-SIF solutions shown in Figs. 22(b)-25(b) is influenced
considerably by the behavior of MM-SIFs resulting from WRS only (see Figs. 22(a)-25(a)). In addition, the value of the computed SIFs
produced due to the combination of external load and WRS presented in Figs. 22(b)-25(b)) is almost doubled for KI compared to those
obtained in Fig. 14 and Figs. 22(a)-25(a), especially near the crack end locations. Based on the behavior of MM-SIF solutions illustrated
in Figs. 22(b)-25(b), it is clear that the cracks will propagate with irregular shapes.

5. Conclusions

An efficient numerical analysis procedure based on FEM was implemented for evaluating the effect of WRS on fracture parameters
(MM-SIFs) for a non-planar surface-cracked tubular T-joint with a rounded weld toe. A multi-pass welding simulation was carried out
using real welding conditions by means of an effective in-house welding simulation code. A cracked mesh generation system was
implemented to generate surface-cracked tubular T-joints with rounded weld toes. The modified WARP3D-IIM was employed to
compute MM-SIF solutions for several cracks under different loading conditions including external load only, WRS only and combi­
nation of external load and WRS. Based on the results discussed in this study, the following conclusions can be drawn:

17
R. Gadallah et al. Marine Structures 71 (2020) 102733

Fig. 24. Distribution of MM-SIFs along the crack front for model T-3 under (a) WRS only, (b) a combination of external load and WRS.

Fig. 25. Distribution of MM-SIFs along the crack front for model T-4 under (a) WRS only, (b) a combination of external load and WRS.

1. Full WRS distributions were obtained for a tubular T-joint with a rounded weld toe resulting from a multi-pass welding simulation
based on real welding conditions. Where the accuracy of the simulated WRS was verified with numerical and experimental data
taken from the literature.
2. A fully compatible 3D FE mesh generation system was implemented to model non-planar perpendicular surface-cracked tubular T-
joints with rounded weld toes in which the efficiency of the implemented system was verified with well-established reference
solutions.
3. A fully automated mapping code was developed to assign the six components of an ASF to each element face center over the crack
surface based on the superposition method.
4. MM-SIFs resulting from the simulated WRS was calculated efficiently for several surface cracks with different shapes by means of
the modified WARP3D-IIM as well as the developed mapping code.
5. The behavior of MM-SIF solutions resulting from the combination of the external load and WRS is influenced significantly by the
behavior of those produced due to WRS only. In addition, the value of the computed SIFs produced due to the combination of
external load and WRS is almost doubled for KI especially near the crack end locations.
6. The implemented numerical analysis procedure is valid and efficient for any crack shape in an ASF in which MM-SIFs for cracks in
compressive WRS fields can be calculated efficiently.

18
R. Gadallah et al. Marine Structures 71 (2020) 102733

Acknowledgment

The first author gratefully acknowledges Prof. Hidekazu Murakawa (Osaka University) and Prof. Sherif Rashed (Osaka University)
for the fruitful discussions.

References

[1] Lie ST, Li G, Cen Z. Analysis of cracked tubular joints using coupled finite and boundary element methods. Eng Struct 2000;22:272–83. https://doi.org/
10.1016/S0141-0296(98)00094-7.
[2] Soh CK, Fung TC, Qin F, Gho WM. Behavior of completely overlapped tubular joints under cyclic loading. J Struct Eng 2001;127:122–8. https://doi.org/
10.1061/(ASCE)0733-9445(2001)127:2(122).
[3] Lee MM. Strength, stress and fracture analyses of offshore tubular joints using finite elements. J Constr Steel Res 1999;51:265–86. https://doi.org/10.1016/
S0143-974X(99)00025-5.
[4] Madi Y, Matheron P, Recho N, Mongabure P. Low cycle fatigue of welded joints: new experimental approach. Nucl Eng Des 2004;228:161–77. https://doi.org/
10.1016/j.nucengdes.2003.06.016.
[5] Gadallah R, Tsutsumi S, Hiraoka K, Murakawa H. Prediction of residual stresses induced by low transformation temperature weld wires and its validation using
the contour method. Mar Struct 2015;44:232–53. https://doi.org/10.1016/j.marstruc.2015.10.002.
[6] Lee C-H, Chang K-H, Van Do VN. Modeling the high cycle fatigue behavior of T-joint fillet welds considering weld-induced residual stresses based on continuum
damage mechanics. Eng Struct 2016;125:205–16. https://doi.org/10.1016/j.engstruct.2016.07.002.
[7] Rettenmeier P, Roos E, Weihe S. Fatigue analysis of multiaxially loaded crane runway structures including welding residual stress effects. Int J Fatig 2016;82:
179–87. https://doi.org/10.1016/j.ijfatigue.2015.04.009.
[8] Tchoffo Ngoula D, Beier HT, Vormwald M. Fatigue crack growth in cruciform welded joints: influence of residual stresses and of the weld toe geometry. Int J
Fatig 2017;101:253–62. https://doi.org/10.1016/j.ijfatigue.2016.09.020.
[9] Hemmesi K, Farajian M, Fatemi A. Application of the critical plane approach to the torsional fatigue assessment of welds considering the effect of residual
stresses. Int J Fatig 2017;101:271–81. https://doi.org/10.1016/j.ijfatigue.2017.01.023.
[10] Lee C-H, Chang K-H. Finite element computation of fatigue growth rates for mode I cracks subjected to welding residual stresses. Eng Fract Mech 2011;78:
2505–20. https://doi.org/10.1016/j.engfracmech.2011.06.006.
[11] Gadallah R, Osawa N, Tanaka S, Tsutsumi S. A novel approach to evaluate mixed-mode SIFs for a through-thickness crack in a welding residual stress field using
an effective welding simulation method. Eng Fract Mech 2018;197:48–65. https://doi.org/10.1016/j.engfracmech.2018.04.040.
[12] Gadallah R, Osawa N, Tanaka S, Tsutsumi S. Critical investigation on the influence of welding heat input and welding residual stress on stress intensity factor
and fatigue crack propagation. Eng Fail Anal 2018;89:200–21. https://doi.org/10.1016/j.engfailanal.2018.02.028.
[13] Gadallah R, Osawa N, Tanaka S. Evaluation of stress intensity factor for a surface cracked butt welded joint based on real welding residual stress. Ocean Eng
2017;138:123–39. https://doi.org/10.1016/j.oceaneng.2017.04.034.
[14] Gadallah R, Osawa N, Tanaka S. Numerical estimation on stress intensity factors for surface cracks in a welding residual stress field. In: ASME 35th Int. Conf.
Ocean. Offshore Arct. Eng., ASME; 2016. https://doi.org/10.1115/OMAE2016-54495. V003T02A026.
[15] Bao R, Zhang X, Yahaya NA. Evaluating stress intensity factors due to weld residual stresses by the weight function and finite element methods. Eng Fract Mech
2010;77:2550–66. https://doi.org/10.1016/j.engfracmech.2010.06.002.
[16] Pouget G, Reynolds AP. Residual stress and microstructure effects on fatigue crack growth in AA2050 friction stir welds. Int J Fatig 2008;30:463–72. https://doi.
org/10.1016/j.ijfatigue.2007.04.016.
[17] Nagashima T, Miura N. Crack analysis in residual stress field by X-FEM. J Comput Sci Technol 2009;3:136–47. https://doi.org/10.1299/jcst.3.136.
[18] Tanaka S, Suzuki H, Sadamoto S, Okazawa S, Yu TT, Bui TQ. Accurate evaluation of mixed-mode intensity factors of cracked shear-deformable plates by an
enriched meshfree Galerkin formulation. Arch Appl Mech 2017;87:279–98. https://doi.org/10.1007/s00419-016-1193-x.
[19] Tanaka S, Suzuki H, Sadamoto S, Sannomaru S, Yu T, Bui TQ. J-integral evaluation for 2D mixed-mode crack problems employing a meshfree stabilized
conforming nodal integration method. Comput Mech 2016;58:185–98. https://doi.org/10.1007/s00466-016-1288-9.
[20] Tanaka S, Suzuki H, Sadamoto S, Imachi M, Bui TQ. Analysis of cracked shear deformable plates by an effective meshfree plate formulation. Eng Fract Mech
2015;144:142–57. https://doi.org/10.1016/j.engfracmech.2015.06.084.
[21] Tanaka S, Okada H, Okazawa S, Fujikubo M. Fracture mechanics analysis using the wavelet Galerkin method and extended finite element method. Int J Numer
Methods Eng 2013;93:1082–108. https://doi.org/10.1002/nme.4433.
[22] Tanaka S, Suzuki H, Ueda S, Sannomaru S. An extended wavelet Galerkin method with a high-order B-spline for 2D crack problems. Acta Mech 2015;226:
2159–75. https://doi.org/10.1007/s00707-015-1306-6.
[23] Gosz M, Moran B. An interaction energy integral method for computation of mixed-mode stress intensity factors along non-planar crack fronts in three
dimensions. Eng Fract Mech 2002;69:299–319. https://doi.org/10.1016/S0013-7944(01)00080-7.
[24] Nose M, Amano H, Okada H, Yusa Y, Maekawa A, Kamaya M, et al. Computational crack propagation analysis with consideration of weld residual stresses. Eng
Fract Mech 2017;182:708–31. https://doi.org/10.1016/j.engfracmech.2017.06.022.
[25] Walters MC, Paulino GH, Dodds RH. Interaction integral procedures for 3-D curved cracks including surface tractions. Eng Fract Mech 2005;72:1635–63.
https://doi.org/10.1016/j.engfracmech.2005.01.002.
[26] Stern M, Becker EB, Dunham RS. A contour integral computation of mixed-mode stress intensity factors. Int J Fract 1976;12:359–68. https://doi.org/10.1007/
BF00032831.
[27] Yau JF, Wang SS, Corten HT. A mixed-mode crack analysis of isotropic solids using conservation laws of elasticity. J Appl Mech 1980;47:335. https://doi.org/
10.1115/1.3153665.
[28] Gosz M, Dolbow J, Moran B. Domain integral formulation for stress intensity factor computation along curved three-dimensional interface cracks. Int J Solid
Struct 1998;35:1763–83. https://doi.org/10.1016/S0020-7683(97)00132-7.
[29] Gao F, Xiao Z, Guan X, Zhu H, Du G. Dynamic behavior of CHS-SHS tubular T-joints subjected to low-velocity impact loading. Eng Struct 2019;183:720–40.
https://doi.org/10.1016/j.engstruct.2019.01.027.
[30] Wang Z, Zhang Y, Wang Y, Du X, Yuan H. Numerical study on fatigue behavior of tubular joints for signal support structures. J Constr Steel Res 2018;143:1–10.
https://doi.org/10.1016/j.jcsr.2017.12.016.
[31] Zhao X, Liu J, Xu X, Sivakumaran KS, Chen Y. Hysteretic behaviour of overlapped tubular k-joints under cyclic loading. J Constr Steel Res 2018;145:397–413.
https://doi.org/10.1016/j.jcsr.2018.02.035.
[32] Wei X, Wen Z, Xiao L, Wu C. Review of fatigue assessment approaches for tubular joints in CFST trusses. Int J Fatig 2018;113:43–53. https://doi.org/10.1016/j.
ijfatigue.2018.04.007.
[33] Qu H, Li A, Huo J, Liu Y. Dynamic performance of collar plate reinforced tubular T-joint with precompression chord. Eng Struct 2017;141:555–70. https://doi.
org/10.1016/j.engstruct.2017.03.037.
[34] Yagi K, Tanaka S, Kawahara T, Nihei K, Okada H, Osawa N. Evaluation of crack propagation behaviors in a T-shaped tubular joint employing tetrahedral FE
modeling. Int J Fatig 2017;96:270–82. https://doi.org/10.1016/j.ijfatigue.2016.11.028.
[35] Th�evenet D, Ghanameh MF, Zeghloul A. Fatigue strength assessment of tubular welded joints by an alternative structural stress approach. Int J Fatig 2013;51:
74–82. https://doi.org/10.1016/j.ijfatigue.2013.02.003.
[36] Shen W, Choo YS. Stress intensity factor for a tubular T-joint with grouted chord. Eng Struct 2012;35:37–47. https://doi.org/10.1016/j.engstruct.2011.10.014.

19
R. Gadallah et al. Marine Structures 71 (2020) 102733

[37] Qian X, Dodds RH, Choo YS. Mode mixity for tubular K-joints with weld toe cracks. Eng Fract Mech 2006;73:1321–42. https://doi.org/10.1016/j.
engfracmech.2006.01.014.
[38] Jang GC, Chang KH, Lee CH. Characteristics of the residual stress distribution in welded tubular T-joints. J Mech Sci Technol 2007;21:1714–9. https://doi.org/
10.1007/BF03177399.
[39] Puymbroeck E Van, Nagy W, Fang H, De Backer H. Determination of residual weld stresses with the incremental hole-drilling method in tubular steel bridge
joints. Procedia Eng 2018;213:651–61. https://doi.org/10.1016/j.proeng.2018.02.061.
[40] Cao Y, Meng Z, Zhang S, Tian H. FEM study on the stress concentration factors of K-joints with welding residual stress. Appl Ocean Res 2013;43:195–205.
https://doi.org/10.1016/j.apor.2013.09.006.
[41] Chang K-H, Kang S-U, Wang Z-M, Muzaffer S, Hirohata M. Fatigue finite element analysis on the effect of welding joint type on fatigue life and crack location of
a tubular member. Arch Appl Mech 2019. https://doi.org/10.1007/s00419-019-01513-4.
[42] Healy B, Gullerud A, Koppenhoefer K, Roy A, RoyChowdhury S, Petti J, et al. WARP3D-Release 17.7.1. Report No. UILU-ENG-95-2012, civil engineering. IL
61801, USA: University of Illinois; 2016.
[43] Murakawa H, Ma N, Huang H. Iterative substructure method employing concept of inherent strain for large-scale welding problems. Weld World 2015;59:
53–63. https://doi.org/10.1007/s40194-014-0178-z.
[44] Nishikawa H, Serizawa H, Murakawa H. Actual application of FEM to analysis of large scale mechanical problems in welding. Sci Technol Weld Join 2007;12:
147–52. https://doi.org/10.1179/174329307X164274.
[45] Wang J, Ma N, Murakawa H. An efficient FE computation for predicting welding induced buckling in production of ship panel structure. Mar Struct 2015;41:
20–52. https://doi.org/10.1016/j.marstruc.2014.12.007.
[46] Ma N, Huang H, Murakawa H. Effect of jig constraint position and pitch on welding deformation. J Mater Process Technol 2015;221:154–62. https://doi.org/
10.1016/j.jmatprotec.2015.02.022.
[47] Wang J, Rashed S, Murakawa H. Mechanism investigation of welding induced buckling using inherent deformation method. Thin-Walled Struct 2014;80:
103–19. https://doi.org/10.1016/j.tws.2014.03.003.
[48] Deng D, Murakawa H. Influence of transformation induced plasticity on simulated results of welding residual stress in low temperature transformation steel.
Comput Mater Sci 2013;78:55–62. https://doi.org/10.1016/j.commatsci.2013.05.023.
[49] Wang J, Yin X, Murakawa H. Experimental and computational analysis of residual buckling distortion of bead-on-plate welded joint. J Mater Process Technol
2013;213:1447–58. https://doi.org/10.1016/j.jmatprotec.2013.02.009.
[50] American Welding Society (AWSA5.20/A5.20M). Specification for carbon steel electrodes for flux cored arc welding. 2005.
[51] Bhatti AA, Barsoum Z, Murakawa H, Barsoum I. Influence of thermo-mechanical material properties of different steel grades on welding residual stresses and
angular distortion. Mater Des 2015;65:878–89. https://doi.org/10.1016/j.matdes.2014.10.019.
[52] Matos CG, Dodds RH. Probabilistic modeling of weld fracture in steel frame connections part I: quasi-static loading. Eng Struct 2001;23:1011–30. https://doi.
org/10.1016/S0141-0296(00)00107-3.
[53] Matos CG, Dodds RH. Probabilistic modeling of weld fracture in steel frame connections Part II: seismic loading. Eng Struct 2002;24:687–705. https://doi.org/
10.1016/S0141-0296(01)00133-X.
[54] Matos CG, Dodds RH. Modeling the effects of residual stresses on defects in welds of steel frame connections. Eng Struct 2000;22:1103–20. https://doi.org/
10.1016/S0141-0296(99)00055-3.
[55] Paulino GH, Kim J-H. A new approach to compute T-stress in functionally graded materials by means of the interaction integral method. Eng Fract Mech 2004;
71:1907–50. https://doi.org/10.1016/j.engfracmech.2003.11.005.
[56] Yu H, Wu L, Guo L, Li H, Du S. T-stress evaluations for nonhomogeneous materials using an interaction integral method. Int J Numer Methods Eng 2012;90:
1393–413. https://doi.org/10.1002/nme.4263.
[57] Qian X, Dodds RH, Choo YS. Elastic–plastic crack driving force for tubular X-joints with mismatched welds. Eng Struct 2005;27:1419–34. https://doi.org/
10.1016/j.engstruct.2005.03.013.
[58] Qian X, Dodds RH, Choo YS. Elastic–plastic crack driving force for tubular K-joints with mismatched welds. Eng Struct 2007;29:865–79. https://doi.org/
10.1016/j.engstruct.2006.06.024.
[59] Bowness D, Lee MMK. Fracture mechanics assessment of fatigue cracks in offshore tubular structures. OTO Report 2000/077, Health and Safety Executive. 2002.
http://www.hse.gov.uk/research/otopdf/2000/oto00077.pdf.
[60] Cao JJ, Yang GJ, Packer JA, Burdekin FM. Crack modeling in FE analysis of circular tubular joints. Eng Fract Mech 1998;61:537–53. https://doi.org/10.1016/
S0013-7944(98)00091-5.
[61] Integrity Quest. 3D finite element software for cracks, User’s manual ver. 3.2.034. 2018.
[62] MSC patran 2017. User’s guide n.d.
[63] Lee MMK, Bowness D. Estimation of stress intensity factor solutions for weld toe cracks in offshore tubular joints. Int J Fatig 2002;24:861–75. https://doi.org/
10.1016/S0142-1123(01)00209-2.
[64] Chiew SP, Lie ST, Lee CK, Huang ZW. Parametric equations for stress intensity factors of cracked tubular T&Y-joints. In: 13th int. Offshore polar eng. Conf. 25-30
may. Honolulu, Hawaii, USA: International Society of Offshore and Polar Engineers; 2003. p. 255–62.
[65] Bowness D, Lee MMK. Prediction of weld toe magnification factors for semi-elliptical cracks in T–butt joints. Int J Fatig 2000;22:369–87. https://doi.org/
10.1016/S0142-1123(00)00012-8.

20

You might also like