Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/341081967

Simulation of Crack Propagation in API 5L X52 Pressurized Pipes Using XFEM-


Based Cohesive Segment Approach

Article in Journal of Pipeline Systems Engineering and Practice · May 2020


DOI: 10.1061/(ASCE)PS.1949-1204.0000444

CITATIONS READS

14 1,128

5 authors, including:

Meng Lin Sylvester Agbo


University of Alberta University of Alberta
16 PUBLICATIONS 135 CITATIONS 16 PUBLICATIONS 97 CITATIONS

SEE PROFILE SEE PROFILE

J.J. Roger Cheng Samer Adeeb


University of Alberta University of Alberta
196 PUBLICATIONS 2,470 CITATIONS 220 PUBLICATIONS 1,865 CITATIONS

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Investigation of pelvic bone fracture mechanism View project

Biomechanical analysis of infant hips View project

All content following this page was uploaded by Meng Lin on 04 May 2020.

The user has requested enhancement of the downloaded file.


Simulation of Crack Propagation in API 5L X52 Pressurized
Pipes Using XFEM-Based Cohesive Segment Approach
Meng Lin 1; Sylvester Agbo 2; Da-Ming Duan 3; J. J. Roger Cheng, M.ASCE 4; and Samer Adeeb 5

Abstract: The cohesive zone model (CZM) is one of the most widely used damage models to describe the fracture processes of brittle and
ductile materials, and has been usually combined with the conventional finite-element method (FEM). CZM in the context of the more
effective extended finite-element method (XFEM) has recently been implemented in many applications, but it has not been widely used
Downloaded from ascelibrary.org by MENG Lin on 02/07/20. Copyright ASCE. For personal use only; all rights reserved.

for crack propagation of pipelines. This paper aims to investigate the capability of the XFEM-based cohesive segment approach implemented
in Abaqus to predict crack propagation of pipelines by calibrating a linearly decreasing traction–separation law with two damage parameters,
the maximum principal stress and the fracture energy. The damage parameters for vintage pipeline steel (API 5L Grade X52) were sys-
tematically calibrated and verified by comparing the numerical results with eight full-scale experiments of pressurized and circumferentially
surface-cracked pipe specimens. A correlation between the damage parameters and material yield strength and fracture toughness is discussed
and an investigation of mesh size sensitivity included. DOI: 10.1061/(ASCE)PS.1949-1204.0000444. © 2020 American Society of Civil
Engineers.
Author keywords: Extended finite-element method; Traction–separation law; Maximum principal stress; Fracture energy.

Introduction resemble the actual conditions of buried pipeline underneath the


ground. For more cost-effective pipeline design, the numerical tech-
Assessment of fracture behavior of steel pipelines has been tradi- niques based on the finite-element method (FEM) or extended
tionally conducted by performing standard fracture toughness tests finite-element method (XFEM) have offered alternative solutions in
on deep-cracked specimens, such as compact tension (CT) speci- numerously simulating the crack initiation and propagation of
mens and single-edge notched bend (SENB) specimens. The valid pipelines subjected to various complex loading conditions.
fracture toughness results require deep-cracked specimens with high In most FEM-based approaches, the cohesive zone model
levels of constraints at the crack tip, which are too conservative es- (CZM) is one of the most widely used numerical techniques for
pecially in the study of commonly used thin-walled steel pipeline crack propagation analysis. The model was originally introduced
with relatively high toughness (Billingham et al. 2003; Wang et al. by Barenblatt (1959, 1962) and Dugdale (1960) in the early 1960s,
2012; Zhang et al. 2010). Zhang et al. (2010) stated that a crack in a who respectively described the near-tip nonlinear processes in
thin-walled pipeline is essentially in a low-constraint configuration quasi-brittle materials and ductile materials with small-scale plas-
even when the pipe is subjected to global bending. In recent years, ticity. The first implementation of CZM in the context of FEM was
shallow-cracked specimens, such as single-edge notched tension proposed by Hillerborg et al. (1976) for brittle fracture in concrete.
(SENT) specimens, which can accurately capture the stress state at In the late 1980s and early 1990s, ductile fracture at the micro-
the crack tip in pipelines, have been highly recommended (Shen scale was studied by Needleman (1987), while ductile fracture at
et al. 2008; Wang et al. 2012). However, there is a high cost asso- the macroscale was studied by Tvergaard and Hutchinson (1992).
ciated with specimen fabrication and testing for small-scale SENT The first implementation of CZM in the context of XFEM was pro-
fracture test specimens. The costs are even higher for physically posed by Wells and Sluys (2001) based on the partition of unity
conducting full-scale experiments on cracked pipes subjected to property of finite elements for a cohesive crack. The CZM model
both internal pressure and external tension and/or bending to considers fracture as a gradual phenomenon and describes a con-
stitutive relation [referred to as the traction–separation law (TSL)]
1 between cohesive traction acting on crack surfaces resisted by crack
Ph.D. Candidate in Structural Engineering, Dept. of Civil and Envir-
propagation and the corresponding separations of surfaces across
onmental Engineering, Univ. of Alberta, Edmonton, AB, Canada T6W 2K6
(corresponding author). ORCID: https://orcid.org/0000-0002-1688-0693. an extended crack tip (cohesive zone) (Jiang 2010; Jin and Sun
Email: lin4@ualberta.ca 2006; Khoei 2015; Zeng 2015). When the separation at the tail
2
Ph.D. Candidate in Structural Engineering, Dept. of Civil and Environ- of the cohesive zone (physical crack tip) reaches a critical value, the
mental Engineering, Univ. of Alberta, Edmonton, AB, Canada T6W 2K6. crack grows, while the cohesive traction vanishes (Jin and Sun
3
Pipeline Integrity Engineer, TransCanada Pipelines Ltd., 450 1st St. 2006). The major advantage of CZM is the removal of crack tip
SW, Calgary, AB, Canada T2P 5H1. stress singularities in classical fracture mechanics. The damage
4
Professor, Dept. of Civil and Environmental Engineering, Univ. of model following a TSL in CZM describes the loss of load or
Alberta, Edmonton, AB, Canada T6W 2K6. deflection-bearing capacity of material as a function of a crack sur-
5
Professor, Dept. of Civil and Environmental Engineering, Univ. of
face separation (i.e., displacement jump), irrespective of the physi-
Alberta, Edmonton, AB, Canada T6W 2K6. Email: adeeb@ualberta.ca
Note. This manuscript was submitted on June 20, 2018; approved on
cal details of damage occurring in an actual material (Schwalbe
August 12, 2019; published online on February 8, 2020. Discussion period et al. 2013; Zeng 2015). Therefore, it can be applied to both brittle
open until July 8, 2020; separate discussions must be submitted for and ductile fracture processes (Schwalbe et al. 2013; Zeng 2015).
individual papers. This paper is part of the Journal of Pipeline Systems The choice of a TSL depends on the brittle or ductile fracture
Engineering and Practice, © ASCE, ISSN 1949-1190. behaviors of materials under consideration. For brittle materials, the

© ASCE 04020009-1 J. Pipeline Syst. Eng. Pract.

J. Pipeline Syst. Eng. Pract., 2020, 11(2): 04020009


Previous Research on Pipelines

Numerical Studies on Modern Pipelines


Recent studies over the past decade using the FEM-based CZM in
simulating fracture behaviors of modern pipelines have achieved
satisfying results. Shim et al. (2011) successfully employed the
method in the simulation of the ductile crack propagation on a cir-
cumferentially cracked pipe by modeling a three-dimensional four-
point bending test on a pipe specimen made of SA-358 Type 304
stainless steel (σy ¼ 220 MPa and σu ¼ 682 MPa). The modeled
crack was a circumferential through-wall crack with a length of
Fig. 1. Typical traction–separation laws: (a) for brittle fracture pro- 37% of pipe circumference. The spiderweb mesh technique was em-
posed by Hillerborg et al. (1976); and (b) for ductile fracture propose ployed in the crack tip region with a mesh size of 2.54 mm. The TSL
by Scheider (2001). (Adapted from Schwalbe et al. 2013.) adopted from Scheider (2001) was used in the work with the dam-
age parameters T 0 ¼ 618 MPa (2.8σy ) and Γ0 ¼ 1,084 N=mm.
Downloaded from ascelibrary.org by MENG Lin on 02/07/20. Copyright ASCE. For personal use only; all rights reserved.

Nonn and Kalwa (2012, 2013) and Scheider et al. (2014) success-
TSL proposed by Hillerborg et al. (1976) [Fig. 1(a)] has been sug- fully calibrated and verified the damage parameters for API 5L pipe-
gested by many experiments, and the influence of TSL shape plays line steel with grade of X65 (σy ¼ 482 MPa and σu ¼ 572 MPa),
a minor role in crack propagation (Scheider and Brocks 2003). X80 (σy ¼ 663 MPa and σu ¼ 713 MPa), and X100 (σy ¼ 756 MPa
However, for ductile materials involving large plastic material de- and σu ¼ 757 MPa). A number of three-dimensional drop weight
formation, there is not a correct shape of TSL determined from tear testing (DWTT) tests and SENT tests were simulated for spec-
experiments and hence it has been usually assumed (Scheider and imens cut from corresponding-grade longitudinal welded pipe. By
Brocks 2003), such as the TSL proposed by Scheider (2001) accurately reproducing the load-deformation and fracture resistance
in modeling ductile fracture [Fig. 1(b)]. The magnitudes of cohesive curves obtained from small-scale tests, the adequate sets of damage
parameters are dependent on the shape of TSL, but all differ- parameters were determined: T 0 ¼ 1,375 MPa (¼2.85σy ) and Γ0 ¼
ent shapes of TSL with their related cohesive parameters are able
900 N=mm for X65 steel, T 0 ¼ 1,600 MPa (¼2.4σy ) and Γ0 ¼
to reproduce experimental results (Scheider and Brocks 2003;
Schwalbe et al. 2013). For a given shape of TSL, the material dam- 900 N=mm for X80 steel, and T 0 ¼ 1,700 MPa (¼2.26σy ) and
age can be only characterized by two damage parameters, cohesive Γ0 ¼ 700 N=mm for X100 steel. The mesh size in cohesive ele-
strength (T 0 ) and cohesive energy (Γ0 ) or complete surface separa- ments was 0.5 × 0.01 × 1.58 mm (¼1=12 of specimen thickness)
for X65 steel, and 0.5 × 0.01 × 1.15 mm (¼1=12 of specimen
tion (δ 0 ), where Γ0 ¼ ∫ δ00 TðδÞdδ is the work needed to crack a unit
thickness) for X80 steel, which are dimensions respectively in the
area of fracture surfaces, equivalent to the area under the TSL curve.
direction of crack propagation, perpendicular to the crack plane
Based on the concept of partition of unity by Melenk and
(width of cohesive element layer), and in the direction of the speci-
Babuška (1996), XFEM extends the conventional FEM by en-
men thickness (Nonn and Kalwa 2013). The TSL used in their mod-
riching additional discontinuous displacement in conjunction with
elings was adapted from Scheider (2001) as shown in Fig. 1(b),
additional degrees of freedom in the elements that can capture the
where the adjusting shape parameters δ 1 ¼ 0.001δ 0 and δ 2 ¼
physical discontinuity, such as cracks (Belytschko and Black 1999;
0.5δ0 resulting in Γ0 ¼ 0.75T 0 δ0 , and it was embedded in the user
Dolbow 1999; Moёs et al. 1999). The XFEM-based CZM (typi-
subroutine UEL of Abaqus/Standard.
cally referred to as the XFEM-based cohesive segment approach)
More recently, XFEM-based CZM has been employed in pipe-
is based on traction–separation cohesive behavior used in conven-
line fracture studies. Liu et al. (2012) successfully employed the
tional FEM. It is based on the intraelement algorithm in which the
method in the simulation of crack behavior of a buried API 5L
discontinuities can be freely laid within elements and are not tied
X65 pipe (σy ¼ 460 MPa and σu ¼ 667 MPa) during the landslide
to element boundaries without the need of remeshing to match the
geometry of the discontinuities (Khoei 2015). Phantom nodes, first process by modeling a three-dimensional pipe in a soil model.
proposed by Hansbo and Hansbo (2004), are introduced in the The damage process was defined by a TSL with the damage ini-
method to represent the discontinuities of the cracked elements. tiation defined by the maximum principal stress σmaxps ¼ 667 MPa
In the commercial finite-element system Abaqus/Standard, the soft- (¼σu ¼ 1.45σy ) taken from the tensile strength and an expo-
ware searches for critical regions of crack initiation in which the nential response for damage evolution with fracture energy GC ¼
stress or strain exceeds a user-defined critical value, after which 180 N=mm obtained from SENB test results. The mesh sensitivity
phantom nodes and their superposed original real nodes move was examined by comparing results of the meshes of 50 × 10 ×
apart. The maximum principal stress (σmaxps ) and the fracture en- 3.1 mm and 25 × 10 × 2.5 mm in the fracture process zone along
ergy (GC ) are two commonly used damage parameters in control- the longitudinal, hoop, and depth direction, respectively, which
ling the crack initiation and propagation. resulted in no significant variation of numerical results. The same
This paper aims to simulate the crack propagation of API 5L values of σmaxps and GC were used by Zhang et al. (2016) in the
Grade X52 steel pipes subjected to the combined effects of internal simulation of the fatigue crack behaviors of X65 pipe with a circum-
pressure and external tension and bending using the XFEM-based ferential elliptical embedded crack under cyclic tensile loadings and
cohesive segment approach implemented in the commercial finite- the classical Paris law–based fatigue model was added to the method.
element system Abaqus/Standard. A suitable set of damage param- The mesh sensitivity was examined by modeling mesh of 0.3, 0.5,
eters, σmaxps and GC , was calibrated and verified using the eight and 0.7 mm in the fracture zone in three directions. Almost the same
full-scale experimental results. The correlation between the damage fatigue crack growth pattern was obtained using mesh sizes of 0.3 and
parameters and the material yield strength and fracture toughness is 0.5 mm, indicating that the element size of 0.5 mm was fine enough
discussed. The effect of mesh sensitivity on the numerical results to produce accurate results. Hojjati-Talemi (2016) and Hojjati-
was examined in one of the models. Talemi et al. (2016) employed the method in the simulation of

© ASCE 04020009-2 J. Pipeline Syst. Eng. Pract.

J. Pipeline Syst. Eng. Pract., 2020, 11(2): 04020009


Fig. 2. (Color) Location of a circumferential crack in (a) a model; and (b) an experiment.
Downloaded from ascelibrary.org by MENG Lin on 02/07/20. Copyright ASCE. For personal use only; all rights reserved.

dynamic brittle fracture of X70 pipe (σy ¼ 520 MPa at 23° and σy ¼
760 MPa at −196°) at low temperature by modeling a two-
dimensional Charpy V-notch (CVN) impact test and a DWTT test.
The damage parameters used were σmaxps ¼ 1,064 MPa (¼1.4σy )
and GIC ¼ K 2IC =ðE=ð1 − ν 2 ÞÞ ¼ 2.7 N=mm, where fracture tough-
pffiffiffiffi
ness K IC ¼ 25 MPa m at −196° derived from measured CVN at
the lower shelf of the ductile-to-brittle-transition curve known as
Barsom-Rolfe correlation (Barsom and Rolfe 1970). The mesh sizes Fig. 3. (Color) Numerical model configuration and locations in which
used were 0.15 × 0.15 mm in the CVN model and 0.5 × 0.5 mm ε0.5L were obtained.
in the DWTT model at their potential crack propagation regions
and the size increased gradually far from the area of interest. In all
the studies mentioned, the damage parameters were not calibrated
Minnesota) machine on the top-side tongue plate when the bottom-
and verified from extensive experiments; on the contrary, they were
side tongue plate was fixed, and consequently the pipe specimen
often estimated, e.g., from tensile strength or Charpy V-notch im-
would experience an eccentric tension resulting in a certain level
pact energy. In fact, the implementation of XFEM-based CZM in
of bending. A digital image correlation (DIC) system was used to
pipeline fractures studies is still in its early stages.
evaluate the tensile stain along the pipe surface on the tension side
aligned with the circumferential crack. The strains were evaluated
Experimental Testing on X52 Pipe by tracking the movements of a speckle pattern produced by black
and white spray paint [Fig. 2(b)] on the pipe surface from a se-
The fracture behavior of vintage pipeline steel is starting to become quence of recorded images. Additionally, DIC was used to measure
the focus of much research in recent years. In this paper, the vintage the growth of the crack mouth opening displacement (CMOD) by
API 5L Grade X52 steel pipes that were experimentally studied tracking the displacements of two selected reference points on each
by Abdulhameed et al. (2016) and Lin (2015) at the University side of the crack. A number of strain gauges were positioned at a
of Alberta are investigated numerically. The experimental work in- quarter of pipe length, 0.5L (L is defined as the half-length of pipe)
cluded eight full-scale burst tests on circumferentially surface- away from the end plate at 90° intervals around the pipe circum-
cracked pipe sections subjected to both internal pressure and eccen- ference to measure the tensile strain (ε0.5L ) and hoop strain at those
tric tension (Abdulhameed et al. 2016), and a number of small-scale locations. Clinometers were attached to both end plates to measure
material tests consisting of 25 tension coupon tests and 24 Charpy the rotation caused by the eccentric loading. A schematic represen-
V-notch impact tests on specimens cut from the same pipe material tation of the test setup and location of the crack and strain gauges is
(Lin 2015). In the full-scale burst tests, each pipe specimen was cut shown in Fig. 3. The experimental results showed that the burst
out of the Enbridge vintage X52 Norman Wells pipeline with an load and the tensile strain capacity of the pipe were affected by the
outer diameter of 324 mm (NPS 12) and a wall thickness of level of internal pressure applied and by the initial crack configu-
6.9 mm (0.27″). The specimens’ length was approximately taken ration. Additionally, the crack depth showed greater influence on
equal to either four or six times the pipe outer diameter, and they the tensile strain capacity than the crack length.
were capped by steel plates at two ends, on which a pair of tongue
plates were welded with an eccentricity of 50 mm to the pipe lon-
gitudinal axis. Each pipe specimen was specially cut with a circum- Numerical Methods and Simulation
ferential surface crack with a depth either 25% or 50% of the pipe
wall thickness and a length of 5% or 15% of the pipe circumfer-
ence, located near the girth weld at the middle length of the pipe XFEM-Based Cohesive Segment Approach
specimen. The loading was applied in two steps. In the first step, an The XFEM-based cohesive segment approach implemented in the
internal pressure causing hoop stress corresponding to approxi- commercial finite-element system Abaqus/Standard was used in the
mately 70% or 30% of the specified minimum yield strength simulation of the crack propagation of API 5L Grade X52 steel pipes
(SMYS) was applied on the pipe specimen by pressurized water subjected to the combined effects of internal pressure and eccentric
through an opening at the bottom end plate. In the second step, tension. This approach was based on traction–separation cohesive
displacement-controlled tensile loading was applied using the behavior used in conventional FEM to simulate a moving crack
material testing system (MTS Systems Corporation, Eden Prairie, along an arbitrary, solution-dependent path in the bulk material

© ASCE 04020009-3 J. Pipeline Syst. Eng. Pract.

J. Pipeline Syst. Eng. Pract., 2020, 11(2): 04020009


Table 1. Basic information of tests and models
Pipe specimen dimensions Crack dimensions Internal pressure level
Test or Outer diametera Pipe length Wall thickness Crack depth Crack lengthb Internal pressure Hoop stress/SMYS
model (mm) (mm) (mm) (mm) (mm) (MPa) (%)
Test 1 324 1,828.8 6.95 1.7 50 11.65 77
Model 1 304.8 1,828.8 6.8 1.7 50 11.65 73
Test 2 324 1,828.8 6.8 1.5 50 3.47 23
Model 2 304.8 1,828.8 6.8 1.5 50 3.5 22
Test 3 324 1,828.8 6.8 3.1 50 11.67 77
Model 3 304.8 1,828.8 6.8 3.1 50 11.65 73
Test 4 324 1,828.8 6.8 3.3 50 4.74 31
Model 4 304.8 1,828.8 6.8 3.3 50 4.65 29
Test 5 324 1,219.2 6.8 1.4 150 11.65 77
Model 5 304.8 1,219.2 6.8 1.4 150 11.65 73
Test 6 324 1,219.2 6.8 1.8 150 4.61 31
Downloaded from ascelibrary.org by MENG Lin on 02/07/20. Copyright ASCE. For personal use only; all rights reserved.

Model 6 304.8 1,219.2 6.8 1.8 150 4.65 29


Test 7 324 1,219.2 6.8 3.5 (3.0)c 150 11.65 77
Model 7 304.8 1,219.2 6.8 3.3 150 11.65 73
Test 8 324 1,219.2 6.8 2.7 (2.8)c 150 4.65 31
Model 8 304.8 1,219.2 6.8 2.7 150 4.65 29
a
The minor discrepancy between the outer diameter dimension of the experiment and the model is due to an initial confusion about the actual outer diameter of
NPS12. It is expected to have minimum effect on the results.
b
Actual initial crack length of each pipe specimen was not measured, thus the target value (5% or 15% pipe circumference) was used (Abdulhameed et al.
2016). Simulated XFEM crack length of each model was half of the value in the table due to the symmetry around the YZ-plane.
c
Actual initial crack depths were measured from two samples cut from pipe Specimens 7 and 8 after pipe failure using fractography (Abdulhameed et al. 2016).
Additional measured values are shown in parentheses.

(Abaqus 2014). The near-tip asymptotic singularities were not full-scale tests, and a corresponding circumferential XFEM crack
considered, and only the displacement jump across a cracked was respectively added at the middle length of each pipe model
element was considered. The fracture process defined by a trac- (Fig. 3). Because of symmetry around the YZ-plane, only half
tion–separation response model consists of a damage initiation of the pipe was modeled. The entire pipe was modeled by combin-
criterion and a damage evolution law. The response is initially ing a solid part (40 mm long) at the center and two shell parts at the
assumed to be linear elastic until a defined damage initiation cri- sides using the shell-solid coupling constraint to reduce the time
terion is satisfied (either stress or strain based), after which the required for the numerical calculations. The end plates and tongues
material damage occurs according to a defined damage evaluation were added at each end of the pipe using the tie constraint and were
law. The damage evolution law is defined either based on energy simulated as rigid bodies and represented by two reference nodes
dissipated due to the fracture energy per unit area or based on the with an eccentricity of 50 mm to the longitudinal axis of the pipe.
effective displacement at the completed surface separation. The The numerical model of the pipe was first loaded with the internal
available damage evaluation law in Abaqus/Standard is either lin- pressure, then subjected to a tensile displacement assigned at the
early decreasing [Fig. 1(a)] or exponentially decreasing. top reference node with the bottom reference node fixed but
In this paper, the crack propagation associated with the pipe allowing rotation about the X-axis. The basic information of each
fracture was assumed mainly under the Mode I fracture type (open- test and model are summarized in Table 1, and the geometry of a
ing mode), where crack opens perpendicular to the crack plane typical numerical model is illustrated in Fig. 3.
caused by tension or bending. The maximum principal stress The two shell parts of the pipe were modeled using a 4-node
(σmaxps ) and fracture energy (GC ) were selected as two key damage linear shell element with reduced integration (S4R) and with a
parameters to characterize the fracture process, which respectively global mesh size of 5 mm. The solid part of the pipe was modeled
control the crack initiation and the resistance against crack propa- using an 8-node linear brick element with reduced integration
gation of Mode I fracture. A higher value of σmaxps resulted in a (C3D8R), and a mesh construction was carefully defined with a
better deflection-bearing ability of pipeline before damage initia- mesh size between 0.5 and 5 mm. In the blocked partitioned region
tion, while a higher value of GC resulted in a better fracture tough- around the crack propagating path, the fine mesh was employed
ness to resist the crack propagation after damage initiation. A linear with the element height lh ¼ 0.5 mm (corresponding to 1/13 of
softening traction–separation law shown in Fig. 1(a) was used for the pipe wall thickness), length ll ¼ 0.5 mm, and thickness lt ¼
damage evolution. This shape of TSL was considered suitable 2 mm [Fig. 2(a)]. Dimensions lh , ll , and lt are in the direction
to simulate the fracture behaviors of the plastic X52 pipe material of the pipe wall thickness or crack propagation (in the radial direc-
because the pipe fractures demonstrated in the experiments by tion), oriented perpendicular to the crack plane (in the longitudinal
Abdulhameed et al. (2016) were considered to be mostly brittle direction), and oriented parallel to the crack plane (in the hoop/
because of the primarily flat fracture surfaces without significant circumferential direction), respectively.
plastic tearing and dimpling.

Material Properties
Setup of Numerical Modeling
The pipe was simulated as an elastoplastic isotropic material. The
A total of eight three-dimensional finite-element models were de- material properties were taken from the average of true stress versus
veloped using Abaqus/Standard to simulate the eight experimental true plastic strain curves measured from six coupons cut in the

© ASCE 04020009-4 J. Pipeline Syst. Eng. Pract.

J. Pipeline Syst. Eng. Pract., 2020, 11(2): 04020009


CMOD, and fracture surface appearances, as shown in Figs. 5–8
and Table 2.
Two definitions were used for the critical tensile strains (tensile
strains at the onset of failure) produced by the experimental com-
bined loadings (ε0.5L and εavg ): ε0.5L is defined as the tensile strain
at the onset of failure measured at a quarter of the pipe length away
from the end plate at 90° intervals around the pipe circumference (at
an angle of 0°, 90°, 270°, or 180° from the crack), and εavg is defined
as the averaged tensile strain at the onset of failure measured from
the outer surface at an angle of 0° from the crack (at the tension side
of pipe) along the pipe length in a range from 10% to 40% of the
pipe length (from 0.2L to 0.8L) on both sides away from crack. For
Tests and Models 1–4, this range is 185–730 mm from the crack.
Fig. 4. True stress versus true plastic strain input in the modeling. For Tests and Models 5–8, this range is 120–490 mm from the
crack. Regarding the CMOD, there were also two critical defini-
Downloaded from ascelibrary.org by MENG Lin on 02/07/20. Copyright ASCE. For personal use only; all rights reserved.

tions, CMODfailure and CMODcritical : CMODfailure is the CMOD at


the point in time when the water starts to seep out of the crack in the
longitudinal direction of the base metal of X52 pipe specimen by test or the crack propagates through the whole pipe wall thickness
Lin (2015). To be more specific, the Young’s modulus was 199 GPa, in the model, and CMODcritical is the CMOD at the point in time
the Poisson’s ratio was 0.3, the 0.2% offset yield strength was when the applied load was almost constant but the CMOD in-
411 MPa, and the ultimate tensile strength was 473 MPa with a true creased sharply, which is simply represented by the time when the
plastic strain of 0.147. The curve of the true stress versus true plastic load reaches 98% of the maximum load in each test or model.
strain is illustrated in Fig. 4. Additional true fracture stress σf with
corresponding true plastic strain εf was particularly added to assist Comparison of Strains Produced by Internal Pressure
with the crack propagation in the modeling. They can be calculated After the first loading step of internal pressure, the numerical lon-
from the reduction of cross-sectional area at fracture (Davis 2004): gitudinal or tensile strains and hoop strains produced by the ap-
σf ¼ ef =ð1 − qÞ ¼ 1,006 MPa and εf ¼ lnð1=ð1 − qÞÞ ¼ 1.125, plied internal pressure were examined by both experimental and
where the engineering fracture stress ef ¼ 325 MPa and reduction theoretical strains. The averaged strains measured from critical lo-
of cross-sectional area q ¼ 67.7% were measured from six fractured cations shown in Fig. 3 were used in the comparison, in which ex-
tension coupons (Lin 2015). perimental strain gauges were installed at a quarter of the pipe
An initial investigation in determining a suitable set of damage length from the end plate at 90° intervals around the pipe circum-
parameters was conducted by Lin et al. (2017) and Agbo et al. ference. The theoretical strains were calculated based on Barlow’s
(2017) for Tests 3, 7, and 8. Lin et al. (2017) compared the numeri- formula. For a long section of a thin-walled pipe with capped ends,
cal developments of CMOD against the applied tension force ob- the resisting hoop and longitudinal stress and strain caused by the
tained from various combinations of σmaxps and GC for Test 7 to internal pressure are (Adeeb 2011) σl ¼ PD=4t, σh ¼ PD=2t,
study the effect of changing one parameter when the other was kept εl ¼ σL =E − νσh =E, and εh ¼ σh =E − νσl =E, where σl , σh , εl ,
constant. Additionally, Agbo et al. (2017) compared the numerical and εh are longitudinal stress, hoop stress, longitudinal strain, and
developments of tensile strain measured at a quarter length of the hoop strain, respectively, P, D, and t are internal pressure, pipe
pipe away from end plate and at an angle of 90° from the crack outer diameter, and pipe wall thickness, respectively, and E and ν
location against the applied tension force obtained from various are Young’s modulus and Poisson’s ratio, respectively.
combinations of σmaxps and GC for Test 8. Both studies concluded For the eight models, the magnitude differences between the
that σmaxps ¼ 700 MPa and GC ¼ 900 N=mm based on a mesh numerical and the theoretical hoop or tensile strain were less than
size of lh × ll × lt ¼ 0.85 × 0.95 × 2 mm applied in the region 0.00003, while those between the numerical and the experimental
hoop or tensile strain were less than 0.00007. To be more specific,
around the crack-propagating path were able to estimate the exper-
the numerical hoop strains were roughly 3% lower to 1% higher
imental burst load by no more than 6% difference. This research
than the theoretical hoop strains, and they were roughly 4% lower
improved the numerical accuracy by using a finer mesh lh × ll ×
to 6% higher than the experimental hoop strains except for Model 6,
lt ¼ 0.5 × 0.5 × 2 mm in the critical region based on a mesh sen-
which were 18% higher. The numerical tensile strains were roughly
sitivity investigation discussed subsequently. After the calibration
10%–13% lower than the theoretical tensile strains, and they were
and verification of the damage parameters using the eight tests,
roughly 15%–25% lower than the experimental tensile strains.
σmaxps ¼ 750 MPa and GC ¼ 900 N=mm were found to be more
suitable in reproducing the overall experimental results, thus this Comparison of Tensile Strains Produced by
set of values was input to the damage modeling. Combined Loadings
At failure, the numerical tensile strains measured from the outer pipe
surface at 0° from the crack were plotted along the pipe length for
Numerical Results and Discussion each model and were respectively compared with their experimental
tensile strain profiles obtained from the DIC technique (Fig. 5). The
numerical models were able to predict the tensile strains in a region
Validation of Damage Parameter
very close to the crack where the DIC technique was not able to
The validation of the damage parameter set of σmaxps ¼ 750 MPa produce meaningful results owing to the lack of speckles painted
and GC ¼ 900 was based on the comparisons between the nu- at those locations (Abdulhameed et al. 2016). The tensile strain de-
merical and experimental results for the eight tests and models creased from the pipe end to the pipe center at which the crack was
in terms of the hoop and tensile strains produced by the internal located. The local compressive strain (negative values) generated
pressure, tensile strains produced by the combined internal pressure around the crack was caused by a large-scale tension generated at
and external tension and bending, end plate rotation, burst loads, the inside (pressurized) surface of the pipe. The averaged tensile

© ASCE 04020009-5 J. Pipeline Syst. Eng. Pract.

J. Pipeline Syst. Eng. Pract., 2020, 11(2): 04020009


Downloaded from ascelibrary.org by MENG Lin on 02/07/20. Copyright ASCE. For personal use only; all rights reserved.

Fig. 5. (Color) Comparison of tensile strains measured at 0° from the crack along the pipe length at failure obtained from (a) Models and Tests 1 and
2; (b) Models and Tests 3 and 4; (c) Models and Tests 5 and 6; and (d) Models and Tests 7 and 8.

Fig. 6. (Color) Comparison of force–rotation curves obtained from (a) Models and Tests 1 and 2; (b) Models and Tests 3 and 4; (c) Models and Tests 5
and 6; and (d) Models and Tests 7 and 8.

© ASCE 04020009-6 J. Pipeline Syst. Eng. Pract.

J. Pipeline Syst. Eng. Pract., 2020, 11(2): 04020009


Downloaded from ascelibrary.org by MENG Lin on 02/07/20. Copyright ASCE. For personal use only; all rights reserved.

Fig. 7. (Color) Comparison of force–CMOD curves obtained from (a) Models and Tests 1 and 2; (b) Models and Tests 3 and 4; (c) Models and Tests 5
and 6; and (d) Models and Tests 7 and 8.

strain εavg and tensile strain at critical locations ε0.5L measured at


pipe failure are listed in Table 2. Among total tests and models,
the difference between experimental and numerical tensile strains
εavg or ε0.5L were less than 30%. The general higher experimental
tensile strain at failure might be caused by the original crack fab-
ricated on pipe specimens being less sharp. Another reason might be
the existence of multiple initial cracks rather than a single crack,
which was observed from fractographic studies of the fracture sur-
faces (Abdulhameed et al. 2016). Both possibilities would require
higher energy to cause pipe fracture, resulting in higher tensile strain
capacity.

Comparison of Burst Loads, End Plate Rotation,


and CMOD
During the second loading step of longitudinal eccentric tensile
loading, the developments of the end plate rotation and CMOD
for each model were respectively plotted against the reaction force
generated at the top reference node and compared with the corre-
sponding experimental results in Figs. 6 and 7. As shown in Table 2,
the numerical burst loads were roughly 13% lower to 2% higher
than the experimental burst loads for all tests and models, resulting
in a maximum difference of 303 kN measured in Test and Model 6.
Tests and Models 1 and 7 fractured at the maximum load, while
Tests and Models 2–6 and 8 fractured slightly after the maximum
load was reached. The numerical end plate rotations at failure were
roughly 20% lower to 12% higher than experimental rotations for
all tests and models, resulting in a maximum difference of 0.85°
Fig. 8. (Color) Comparison of fracture surfaces obtained from
measured in Test and Model 2. The numerical CMODfailure val-
(a) Model 1; (b) Test 1 in the longitudinal direction; and (c) Test 1 in
ues were roughly 32% lower to 22% higher than experimental
the circumferential direction. [Adapted (b and c) from Abdulhameed
CMODfailure , resulting in a maximum difference of 0.77 mm mea-
et al. 2016.]
sured in Test and Model 3. The numerical CMODcritical values were

© ASCE 04020009-7 J. Pipeline Syst. Eng. Pract.

J. Pipeline Syst. Eng. Pract., 2020, 11(2): 04020009


Table 2. Comparison between numerical and experimental results at failure
Tensile strain (%)
Burst load Rotation at CMODfailure Reduction
ε0.5L εavg
(max load) end plate (CMODcritical ) of pipe wall
Test or model (kN) (degrees) (mm) At 0° At 90°/270° In 0.2L–0.8L thickness (%)
Test 1 2,299 5.08 2.11 (1.19) — 1.950 3 32.4
Model 1 2,335 5.67 1.83 (1.31) 2.620 1.340 2.898 9.3
Difference þ2% þ12% −13% (þ10%) — −31% −3% −23%
Test 2 3,100 (3,109) 6.82 2.16 (1.09) — — 8 27.9
Model 2 2,975 (2,977) 7.67 2.16 (1.21) 5.351 4.276 7.405 9.7
Difference −4% þ12% 0% (þ11%) — — −7% −18%
Test 3 1,623 (1,664) 0.98 2.37 (1.18) 0.510 0.328 0.530 19.1
Model 3 1,653 0.78 1.60 (1.31) 0.419 0.239 0.414 9.0
Difference þ2% −20% −32% (þ11%) −18% −27% −22% −10%
Test 4 2,061 (2,075) 1.04 2.05 (1.16) 0.533 0.321 0.553 20.6
Model 4 1,919 0.96 1.55 (0.80) 0.480 0.271 0.492 8.5
Downloaded from ascelibrary.org by MENG Lin on 02/07/20. Copyright ASCE. For personal use only; all rights reserved.

Difference −7% −8% −24% (−31%) −10% −16% −11% −12%


Test 5 1,934 (1,962) 1.86 1.52 (0.77) — 0.900 1.846 26.5
Model 5 1,786 (1,854) 1.86 1.85 (0.83) 1.364 0.700 1.391 10.3
Difference −8% 0% þ22% (þ8%) — −22% −25% −16%
Test 6 2,261 (2,268) 1.40 1.56 (0.90) 1.204 0.647 1.324 27.9
Model 6 1,958 (1,997) 1.28 1.74 (0.90) 0.942 0.488 0.966 9.1
Difference −13% −9% þ12% (0%) −22% −25% −27% −19%
Test 7 1,304 0.26 1.30 (0.90) 0.208 0.177 0.217 19.1
Model 7 1,261 0.25 1.42 (1.00) 0.157 0.151 0.155 8.7
Difference −3% −4% þ9% (þ11%) −25% −15% −29% −10%
Test 8 1,831 (1,844) 0.42 1.27 (0.85) 0.293 0.218 0.305 20.6
Model 8 1,657 (1,669) 0.36 1.50 (0.89) 0.241 0.184 0.237 9.0
Difference −10% −14% þ18% (þ5%) −18% −16% −22% −12%

roughly equal to 10% higher than experimental CMODcritical , ex- the maximum principal stress σmaxps and fracture energy GC can be
pect for Test and Model 4, which were 31% lower. The differ- correlated to the cohesive strength T 0 and cohesive energy Γ0 in the
ences in CMODcritical values between the models and the tests were FEM-based CZM, with the relationship σmaxps ¼ T 0 and GC ¼ Γ0 .
minimal compared to the differences in CMODfailure between the The accurate damage parameters or cohesive parameters should be
models and tests. This is perhaps because CMODcritical does not determined from experiments and optimized from numerical sim-
consider the dynamic effects due to the sudden release of internal ulations. In CZM, the cohesive strength T 0 can be taken from the
pressure during the test. maximum value of stress over a notched tensile bar’s instantaneous
Comparison of Fracture Surfaces cross section at fracture, which is equal to the true tensile strength
The numerical fracture surfaces were compared with corresponding σu for fracture without localized necking, or higher than the true
experimental fracture surfaces analyzed by the fractographic method tensile strength for fracture with localized necking (Cornec et al.
(Abdulhameed et al. 2016). A typical appearance of the experimen- 2003; Schwalbe et al. 2013). The cohesive energy Γ0 can be taken
tal fracture surface (Test 1) emanating from the original machined from linear elastic energy release rate GIc ¼ K 2Ic =E 0 within the
crack (flaw) is shown in Fig. 8(b) viewed in the longitudinal di- framework of linear elastic fracture mechanics, or from the J-integral
rection of the pipe and in Fig. 8(c) viewed in the circumferential at initiation of stable ductile crack extension within the framework of
direction of pipe, where E is the original crack depth, C is the re- elastic-plastic fracture mechanics J i ≈ K 2Ic =E 0 , where E 0 ¼ E for
duced pipe wall thickness due to the plastic necking, and A is the plane stress assumption, while E 0 ¼ E=ð1 − ν 2 Þ for plane strain
original pipe wall thickness. The experimental fracture surface was assumption (Cornec et al. 2003; Schwalbe et al. 2013). For ductile
considered to be mostly brittle because it was primarily flat without structural steels, a rough estimation of cohesive parameters was
significant plastic tearing and dimpling [Fig. 8(b)], although some suggested by Schwalbe et al. (2013) around T 0 ≈ 3σy and
plasticity was observed in the inner pipe surface owing to the inter- Γ0 ≈ K 2Ic =E 0 . Recalling previous numerical studies on modern
nal pressure. Table 2 summarizes the reduction of pipe wall thick- API 5L pipeline steel, Nonn and Kalwa (2013) suggested T 0 ¼
ness for each test and model as well as their 10%–23% difference in 2.85σy and T 0 ¼ 2.4σy for X65 and X80 steels, respectively, and
reduction of pipe wall thickness. The higher percentage of experi-
Γ0 ≈ J i at initiation of ductile crack extension in the J-resistance
mental reduction of pipe wall thickness indicated that the fracture
curve obtained for DWTT tests for both materials. Hojjati-Talemi
experienced more plasticity in tests than in models. This might be
(2016) and Hojjati-Talemi et al. (2016) suggested σmaxps ¼ 1.4σy
caused by the immense impact resulting from the sudden release of
internal pressure when fracture happened during tests, while it was and GIC ≈ K 2IC =ðE=ð1 − ν 2 ÞÞ for dynamic brittle fracture of
not simulated in the numerical modeling. X70 steel at low temperature, where K IC was correlated from
Charpy V-notch impact energy. The difference of values of cohesive
parameters was caused by selecting the different shape of TSL,
Correlation between Damage Parameters and which implied the TSL shape dependency for the calibration of
Yield Strength and Fracture Toughness
cohesive parameters.
Because the XFEM-based cohesive segment method was based on In this paper, the material properties of X52 pipe steel under
traction–separation cohesive behavior used in conventional FEM, consideration measured or empirically estimated by Lin (2015)

© ASCE 04020009-8 J. Pipeline Syst. Eng. Pract.

J. Pipeline Syst. Eng. Pract., 2020, 11(2): 04020009


Downloaded from ascelibrary.org by MENG Lin on 02/07/20. Copyright ASCE. For personal use only; all rights reserved.

Fig. 9. (Color) XFEM crack propagation obtained from Model 3 regarding (a) maximum principal stress; and (b) status of enriched element.

included 0.2% offset yield strength σy ¼ 411 MPa, Young’s modu- initiation. The same behavior was obtained when using σmaxps ¼
lus E ¼ 199 GPa, Poisson’s ratio ν ¼ 0.3, estimated plane strain 800 MPa and GC ¼ 900 N=mm. For Element 1, the energy dissi-
pffiffiffiffi
stress intensity factor K Ic ¼ 207 MPa m based on CVN impact pated accompanying the crack opening can be roughly calculated
energy ¼ 167 J, and estimated plane strain fracture toughness from the area under the resulting trapezoid shape, which was roughly
J IC ¼ GIc ¼ K 2Ic =ðE=ð1 − ν 2 ÞÞ ¼ 196 N=mm. Then the relation- 535 N=mm (¼2.7K 2Ic =ðE=ð1 − ν 2 ÞÞ) for the case of using σmaxps ¼
ship between the damage parameters σmaxps ¼ 750 MPa and 750 MPa and GC ¼ 900 N=mm, and 600 N=mm (¼3K 2Ic =ðE=
GC ¼ 900 N=mm and yield strength and fracture toughness were ð1 − ν 2 ÞÞ) for the case of using σmaxps ¼ 800 MPa and GC ¼
σmaxps ¼ 1.8σy and GC ¼ 4.6K 2Ic =ðE=ð1 − ν 2 ÞÞ. The calibration of 900 N=mm. However, even though the energy dissipated until fail-
damage parameters was based on the agreement of experimental ure did not reach 900 N=mm, a lower input GC is not recommended
results at pipe failure, which is defined in the model as the point because it would result in failure loads that are significantly lower
of time when the crack tip (or element damage) reaches the inner than the experimental results, which was also shown previously by
surface of the numerical pipe. As shown in Fig. 9, the original Lin et al. (2017) and Agbo et al. (2017). Although the enriched el-
XFEM crack represented by the bottom two red layers of elements ements were not completely damaged in the analysis, the numerical
[Fig. 9(b)] started to propagate through the first enriched element results agreed well with the experimental results.
[Element 1 in Fig. 9(a)] in which σmaxps reached 750 MPa. After- Some researchers (Anvari et al. 2006; Scheider 2009; Schwalbe
ward, the crack propagated mainly along the direction of pipe wall et al. 2013; Siegmund and Brocks 2000) have also proposed that
thickness until the crack tip reached the inner surface of the pipe, the shape of TSL and the related cohesive parameters are strongly
where the enriched Element 2 in Fig. 9(a) was damaged (defined as dependent on the stress triaxiality (h), defined by the hydrostatic
failure). The status of the enriched element in Fig. 9(b) showed that stress (mean normal stress) over the von Mises equivalent stress
all elements through the remaining ligament were damaged h ¼ ððσ11 þ σ22 þ σ33 Þ=3Þ=σv . A typical relationship was pro-
(STATUSXFEM < 1). This can also be seen in Fig. 10, where posed by Anvari et al. (2006), who plotted curves of normalized
the shape of the TSL obtained from Element 1 in Model 3 roughly cohesive strength (T 0 =σy ) and normalized cohesive energy (Γ0 =
followed the linearly decreasing pattern, but the final value of σy Dcell ) against stress triaxiality (h) for rate-insensitive material
σmaxps did not decrease to zero when Element 2 is damaged. Also, based on a unit cell simulation using the Gurson-Tvergaard-
the numerically obtained shape was slightly different from the TSL Needleman (GTN) model on a ferritic steel. GTN is another widely
curve based on σmaxps ¼ 750 MPa and GC ¼ 900 N=mm because used damage model in simulating the ductile crack process, but it
of the significant drop in the early stage caused by the crack requires up to nine damage parameters calibrated from a fracture
resistance curve test and results are strongly mesh size dependent
(Schwalbe et al. 2013). As shown in Fig. 11, as h increases from 1
to 5, T 0 =σy increases from 2 to 4, while Γ0 =σy Dcell decreases from
1.5 to 0.2, where Dcell is the element height of the GTN unit cell.
Lower triaxiality resulted in a decrease of cohesive strength but an
increase of cohesive energy. For Model 3, the triaxiality at the crack
tip at failure obtained from Element 2 was 0.677 when using
σmaxps ¼ 750 MPa and GC ¼ 900 N=mm, while it was 0.690
when using σmaxps ¼ 800 MPa and GC ¼ 900 N=mm. Both cases
resulted in a triaxiality value lower than 1, which explained the
relatively lower ratio of σmaxps =σy around 1.8 but higher ratio of
GC =ðK 2Ic =ðE=ð1 − ν 2 ÞÞÞ around 4.6 computed in our models.

Investigation of Mesh Size Sensitivity


Fig. 10. (Color) Comparison of the shape of TSL between the input
A suitable finite-element mesh has a significant effect on numer-
damage parameters and the numerical model obtained from the first
ical stability and computational accuracy, which has been ad-
damaged enriched element (Element 1) on Model 3.
dressed by numerous researchers in the past. As a rule of thumb,

© ASCE 04020009-9 J. Pipeline Syst. Eng. Pract.

J. Pipeline Syst. Eng. Pract., 2020, 11(2): 04020009


Following h ≤ GIc E 0 =ð10σ2u Þ proposed by Carpinteri (1989)
and Carpinteri and Colombo (1989), the critical element size in this
research was calculated to be 32 mm based on GC ¼ 900 N=mm
and σmaxps ¼ 750 MPa replacing σu , which seemed too coarse. As
in this paper, the crack propagated through the pipe wall thickness
direction, the critical element size of 0.085 mm corresponding to
1/80 of the pipe wall thickness suggested by Oliveira (2013)
seemed too fine. On the other hand, the element size corresponding
to 1/12 of the specimen thickness used in Nonn and Kalwa (2012,
2013) and Scheider et al. (2014) was more appropriate for our
analyses. The finally chosen mesh size in the partitioned critical
region around the crack propagating path in our study was
lh × ll × lt ¼ 0.5 × 0.5 × 2 mm, where the element height lh cor-
responded to 1/13 of the pipe wall thickness (Fig. 3). In addition,
this mesh size was examined in Model 3 as the adequate size in
Downloaded from ascelibrary.org by MENG Lin on 02/07/20. Copyright ASCE. For personal use only; all rights reserved.

predicting the pipe failure in terms of burst load, end plate rotation,
CMODfailure , and tensile strains at various critical locations. Two
Fig. 11. Dependence of normalized cohesive strength and energy on sets of damage parameters were included in the investigation to
stress triaxiality for ferritic steel determined from GTN unit cell simu- exclude the effect caused by inputting different material properties.
lations by Anvari et al. (2006). The first set, σmaxps ¼ 750 MPa and GC ¼ 900 N=mm, was
generally suitable for the eight models in producing experimental
results, while the second set, σmaxps ¼ 800 MPa and GC ¼
900 N=mm, was most suitable for Model 3 itself. The effect of
the recommended size of elements in the direction of crack propa- mesh size on the numerical results was examined by changing
gation should be in the range of 0.05 to 0.25 mm for ductile ma- the mesh size in the region around the crack propagating path
terials in FEM (Schwalbe et al. 2013). Carpinteri (1989) and in the lh , ll , and lt directions (Table 3). In the second set, further
Carpinteri and Colombo (1989) examined mesh refinement on reduction of lh , ll , and lt to 0.25, 0.25, and 1 mm, respectively,
the regularity of simulated load-deformation responses of a resulted in less than 10% variation of the numerical results in
three-point bending beam for different brittleness numbers (sE ) all respects, including 0.2%, 0.8%, and 0.1% variation in burst
using the FEM-based linear elastic fracture mechanics (LEFM) ap- load, 2.1%, 8.2%, and 1.0% variation in end plate rotation at fail-
proach. For brittle material with a low value of sE , the cohesive zone ure, 7%, 9.7%, and 1.1% variation in CMOD at failure, and 3%,
is confined to a small crack tip region, thus a refined mesh is re- 5.8%, and 0.2% variation in εavg . The variations obtained from the
quired. They proposed a lower bound of sE ¼ GIC =ðσu bÞ ¼ first set were slightly higher but still acceptable to ease the com-
GIC =ðσu mhÞ ≥ 0.0008=m to obtain reasonable results for a brittle putational burden caused by analyzing considerable numbers of el-
material with an ultimate tensile strain εu ¼ σu =E ¼ 0.000087, ements. The comparisons of tensile strains at the tension side of the
where h and m are finite-element size and numbers, b ¼ mh pipe along the pipe length at failure obtained from two sets are
is the beam depth. The material characteristic length lch ¼ shown in Figs. 12 and 13. The higher variation caused by changing
GIC E 0 =ðσ2u Þ defined by Hillerborg et al. (1976) can be correlated ll indicated it was more critical in predicting fracture behaviors than
to brittleness number by sE ¼ εu ðlch =bÞ ¼ εu ðlch =mhÞ. Then the the dimensions in the other two directions. Future studies may be
lower bound of sE can be rewritten as lch =h ≥ 0.0008=εu ¼ 9.2. conducted in further reducing the element length oriented
Based on a minimum number of 10 elements per characteristic perpendicular to the crack plane to minimize its sensitivity.
length, the critical finite-element size h around the crack tip should
satisfy h ≤ GIc E 0 =ð10σ2u Þ, where E 0 ¼ E for the plane stress
assumption and E 0 ¼ E=ð1 − ν 2 Þ for the plane strain assumption. Conclusions
The mesh sensitivity based on the same specimen configuration
and material properties was further studied by Moës and Belytschko This paper qualified the use of the XFEM-based cohesive segment
(2002) and Mojiri (2010) using the XFEM-based LEFM approach. approach implemented in the commercial finite-element analysis
Moës and Belytschko (2002) examined the independence of XFEM Abaqus/Standard for crack propagation analysis of pipes. It is the
results to the mesh matching or not matching the cohesive crack first to systematically calibrate and verify the damage parameters
path. Mojiri (2010) obtained very similar results as Carpinteri and for vintage pipeline steel (API 5L Grade X52) from eight full-scale
Colombo (1989) but using fewer elements, which demonstrated pressurized and circumferentially surface-cracked pipe models un-
the accuracy and efficiency of XFEM. In addition to the critical mesh til a good agreement is achieved between the numerical and exper-
size, proper meshing technique is often recommended to ensure re- imental results. Based on a linearly deceasing traction–separation
liable results. Oliveira (2013) suggested defining a structured mesh, law, the most suitable set of damage parameters is σmaxps ¼
such as a hexahedral shape in three dimensions, with high density to 750 MPa and GC ¼ 900 N=mm, which can be correlated to the
achieve satisfying results in the XFEM framework. Additionally, for material yield strength and fracture toughness around σmaxps ¼
efficient use of computational resources, proper mesh partitions are 1.8σy and GC ¼ 4.6K 2Ic =ðE=ð1 − ν 2 ÞÞ. An investigation of mesh
recommend allowing a finer mesh around the crack but a coarser sensitivity was conducted to achieve the numerical stability and
mesh for other regions. In the study of stationary cracks in SENT computational accuracy. The mesh size lh × ll × lt ¼ 0.5 × 0.5 ×
and SENB simulations, 80 elements per partition length correspond- 2 mm around the crack-propagating path resulted in 10% varia-
ing to 0.05 mm near the crack was suggested, resulting in less than tion of numerical results, while ll oriented perpendicular to the
3% error of stress intensity factor (K I ). The generated computation crack plane might be further reduced to minimize its sensitivity.
time was significantly less than that using a constant element size of This paper serves as a good basis in effectively predicting crack
0.2 mm for total specimen geometry (Oliveira 2013). propagation of pipelines with various steel grades using XFEM.

© ASCE 04020009-10 J. Pipeline Syst. Eng. Pract.

J. Pipeline Syst. Eng. Pract., 2020, 11(2): 04020009


Table 3. Investigation of mesh sensitivity on fracture behaviors of Model 3
Tensile strain at fracture (%)
Rotation at
ε0.5L
Mesh size in the region Burst load end plate CMODfailure
Model around crack, lh × ll × lt (kN) (degrees) (mm) At 0° At 90°/270° At 180° εavg
Model 3 with 1 × 0.5 × 2 1,664 0.82 1.56 0.440 0.248 0.029 0.437
σmaxps ¼ 750 MPa 0.5 × 0.5 × 2 1,653 0.78 1.60 0.419 0.239 0.031 0.414
and GC ¼ 900 N=mm 0.25 × 0.5 × 2 1,649 0.77 1.69 0.409 0.236 0.031 0.403
0.5 × 1 × 2 1,716 1.09 2.10 0.546 0.298 0.021 0.546
0.5 × 0.25 × 2 1,617 0.66 1.40 0.367 0.216 0.035 0.360
0.5 × 0.5 × 1 1,639 0.73 1.51 0.395 0.229 0.033 0.390
Change of lh
Difference from lh ¼ 1 to lh ¼ 0.5 −0.7% −4.9% þ2.6% −4.8% −3.6% þ6.9% −5.3%
Difference from lh ¼ 0.5 to lh ¼ 0.25 −0.2% −1.3% þ5.6% −2.4% −1.3% 0.0% −2.7%
Change of ll
Difference from ll ¼ 1 to ll ¼ 0.5 −3.7% −28.4% −23.8% −23.3% −19.8% þ47.6% −24.2%
Downloaded from ascelibrary.org by MENG Lin on 02/07/20. Copyright ASCE. For personal use only; all rights reserved.

Difference from ll ¼ 0.5 to ll ¼ 0.25 −2.2% −15.4% −12.5% −12.4% −9.6% þ12.9% −13.0%
Change of lt
Difference from lt ¼ 2 to ll ¼ 1 −0.8% −6.4% −5.6% −5.7% −4.2% þ6.5% −5.8%
Model 3 with 1 × 0.5 × 2 1,707 1.03 1.78 0.531 0.290 0.023 0.534
σmaxps ¼ 800 MPa 0.5 × 0.5 × 2 1,694 0.97 1.86 0.496 0.275 0.024 0.497
and GC ¼ 900 N=mm 0.25 × 0.5 × 2 1,690 0.95 1.99 0.481 0.270 0.025 0.482
0.5 × 1 × 2 1,766 1.39 2.39 0.678 0.360 0.015 0.690
0.5 × 0.25 × 2 1,680 0.89 1.68 0.468 0.262 0.027 0.468
0.5 × 0.5 × 1 1,693 0.96 1.84 0.495 0.275 0.024 0.496
Change of lh
Difference from lh ¼ 1 to lh ¼ 0.5 −0.8% −5.8% þ4.5% −6.6% −5.2% þ4.3% −6.9%
Difference from lh ¼ 0.5 to lh ¼ 0.25 −0.2% −2.1% þ7.0% −3.0% −1.8% þ4.2% −3.0%
Change of ll
Difference from ll ¼ 1 to ll ¼ 0.5 −4.1% −30.2% −22.2% −26.8% −23.6% þ60.0% −28.0%
Difference from ll ¼ 0.5 to ll ¼ 0.25 −0.8% −8.2% −9.7% −5.6% −4.7% þ12.5% −5.8%
Change of lt
Difference from lt ¼ 2 to ll ¼ 1 −0.1% −1.0% −1.1% −0.2% 0.0% 0.0% −0.2%

Fig. 12. (Color) Comparisons of tensile strains at the tension side of pipe at failure obtained from Test 3 and Model 3 using σmaxps ¼ 750 MPa and
GC ¼ 900 N=mm: (a) change of lh ; (b) change of ll ; and (c) change of lt .

© ASCE 04020009-11 J. Pipeline Syst. Eng. Pract.

J. Pipeline Syst. Eng. Pract., 2020, 11(2): 04020009


Downloaded from ascelibrary.org by MENG Lin on 02/07/20. Copyright ASCE. For personal use only; all rights reserved.

Fig. 13. (Color) Comparisons of tensile strains at the tension side of pipe at failure obtained from Test 3 and Model 3 using σmaxps ¼ 800 MPa and
GC ¼ 900 N=mm: (a) change of lh ; (b) change of ll ; and (c) change of lt .

Future work will focus on exploring new damage initiation criteria εavg = averaged tensile strain measured from the outer
that take the constraints around the crack tip into consideration. surface along the pipe length and at an angle of 0°
from the crack in a range from 10% to 40% of pipe
length (0.2L–0.8L) at both sides away from the
Acknowledgments crack (185–730 mm from the crack for Tests and
Models 1–4; 120–490 mm from the crack for Tests
The financial support of the Natural Sciences and Engineering and Models 5–8);
Research Council of Canada (NSERC), Enbridge Pipelines, Inc.,
ε0.5L = tensile strain measured at a quarter of the pipe
and TransCanada Pipelines Limited is thankfully acknowledged.
length (1=4 × 2L ¼ 0.5L) from the end plate at
90° intervals around the pipe circumference; and
Notation σmaxps = maximum principal stress in XFEM-based
cohesive segment approach.
The following symbols are used in this paper:
CMODcritical = crack mouth opening displacement at the point in
time when the applied load is almost constant but References
the CMOD increases sharply, which is simply
represented by the time when the load reaches Abaqus. 2014. Abaqus analysis user’s guide. Vélizy-Villacoublay, France:
Dassault Systèmes.
98% of the maximum applied load in each test and
Abdulhameed, D., C. Cakiroglu, M. Lin, R. Cheng, J. Nychka, M. Sen, and
model;
S. Adeeb. 2016. “The effect of internal pressure on the tensile strain
CMODfailure = crack mouth opening displacement at failure when capacity of X52 pipelines with circumferential flaws.” J. Pressure
the water starts to seep out of the crack in the test Vessel Technol. 138 (6): 061701. https://doi.org/10.1115/1.4033436.
or the crack tip (or element damage) reaches the Adeeb, S. 2011. Introduction to solid mechanics and finite element analysis
inner surface of pipe in the model; using Mathematica. Dubuque, IA: Kendall Hunt.
Gc = fracture energy in XFEM-based cohesive segment Agbo, S., M. Lin, A. Ahmed, J. J. R. Cheng, and S. Adeeb. 2017. “Predic-
approach; tion of burst load in pressurized pipelines using extended finite element
method (XFEM).” In Proc., 6th Int. Conf. on Engineering Mechanics
L = half of the pipe length;
and Materials. Montreal: Canadian Society for Civil Engineering.
T 0 = cohesive strength in FEM-based cohesive zone Anvari, M., I. Scheider, and C. Thaulow. 2006. “Simulation of dynamic
model; ductile crack growth using strain-rate and triaxiality-dependent cohe-
Γ0 = cohesive energy in FEM-based cohesive zone sive elements.” Eng. Fract. Mech. 73 (15): 2210–2228. https://doi.org
model; /10.1016/j.engfracmech.2006.03.016.

© ASCE 04020009-12 J. Pipeline Syst. Eng. Pract.

J. Pipeline Syst. Eng. Pract., 2020, 11(2): 04020009


Barenblatt, G. I. 1959. “The formation of equilibrium cracks during brittle Melenk, J. M., and I. Babuška. 1996. “The partition of unity finite ele-
fracture. General ideas and hypotheses. Axially-symmetric cracks.” ment method: Basic theory and applications.” Comput. Methods Appl.
J. Appl. Math. Mech. 23 (3): 622–636. https://doi.org/10.1016/0021 Mech. Eng. 139 (96): 289–314. https://doi.org/10.1016/S0045-7825
-8928(59)90157-1. (96)01087-0.
Barenblatt, G. I. 1962. “The mathematical theory of equilibrium cracks Moës, N., and T. Belytschko. 2002. “Extended finite element method for
in brittle fracture.” Adv. Appl. Mech. 7 (C): 55–129. https://doi.org/10 cohesive crack growth.” Eng. Fract. Mech. 69 (7): 813–833. https://doi
.1016/S0065-2156(08)70121-2. .org/10.1016/S0013-7944(01)00128-X.
Barsom, J. M., and S. T. Rolfe. 1970. “Correlations between K Ic and Mojiri, S. 2010. “Numerical analysis of cohesive crack growth using ex-
Charpy V-notch test results in the transition-temperature range.” In tended finite element method (X-FEM).” M.Sc. thesis, International
Impact testing of metals, 281–302. West Conshohocken, PA: ASTM. Center for Numerical Methods in Engineering.
Belytschko, T., and T. Black. 1999. “Elastic crack growth in finite elements Moёs, N., J. Dolbow, and T. Belytschko. 1999. “A finite element method
with minimal remeshing.” Int. J. Numer. Methods Eng. 45 (5): 601–620. for crack growth without remeshing.” Int. J. Numer. Methods Eng.
https://doi.org/10.1002/(SICI)1097-0207(19990620)45:5<601::AID 46 (1): 131–150. https://doi.org/10.1002/(SICI)1097-0207(19990910)
-NME598>3.0.CO;2-S. 46:1<131::AID-NME726>3.0.CO;2-J.
Billingham, J., J. V. Sharp, J. Spurrier, and P. J. Kilgallon. 2003. Review of Needleman, A. 1987. “A continuum model for void nucleation by inclusion
the performance of high strength steels used offshore. Cranfield, UK: debonding.” J. Appl. Mech. 54 (3): 525–531. https://doi.org/10.1115/1
Cranfield Univ. .3173064.
Downloaded from ascelibrary.org by MENG Lin on 02/07/20. Copyright ASCE. For personal use only; all rights reserved.

Carpinteri, A. 1989. “Softening and snap-back instability in cohesive Nonn, A., and C. Kalwa. 2012. “Simulation of ductile crack propagation in
solids.” Int. J. Numer. Methods Eng. 28 (7): 1521–1537. https://doi.org high-strength pipeline steel using damage models.” In Vol. 3 of Proc.,
/10.1002/nme.1620280705. ASME 2012 9th Int. Pipeline Conf., 597–603. New York: ASME.
Carpinteri, A., and G. Colombo. 1989. “Numerical analysis of catastrophic Nonn, A., and C. Kalwa. 2013. “Analysis of dynamic ductile fracture
softening behaviour (snap-back instability).” Comput. Struct. 31 (4): propagation in pipeline steels: A damage-mechanics’ approach.”
607–636. https://doi.org/10.1016/0045-7949(89)90337-4. In Proc., 6th Int. Pipeline Technology Conf. Duisburg, Germany:
Cornec, A., I. Scheider, and K. H. Schwalbe. 2003. “On the practical ap- European Pipeline Research Group.
plication of the cohesive model.” Eng. Fract. Mech. 70 (14): 1963–1987. Oliveira, F. 2013. “Crack modelling with the extended finite element
https://doi.org/10.1016/S0013-7944(03)00134-6. method.” M.Sc. thesis, Instituto Superior Técnico.
Davis, J. R. 2004. Tensile testing. 2nd ed. Materials Park, OH: ASM Scheider, I. 2001. “Simulation of cup-cone fracture in round bars using the
International. cohesive zone model.” Comput. Fluid Solid Mech. 460–462. https://doi
Dolbow, J. 1999. “An extended finite element method with discontinuous .org/10.1016/B978-008043944-0/50681-8.
enrichment for applied mechanics.” Ph.D. thesis, Dept. of Mechanical Scheider, I. 2009. “Derivation of separation laws for cohesive models in the
Engineering, Northwestern Univ. course of ductile fracture.” Eng. Fract. Mech. 76 (10): 1450–1459.
Dugdale, D. S. 1960. “Yielding of steel sheets containing slits.” J. Mech. https://doi.org/10.1016/j.engfracmech.2008.12.006.
Phys. Solids 8 (2): 100–104. https://doi.org/10.1016/0022-5096(60) Scheider, I., and W. Brocks. 2003. “The effect of the traction separation law
90013-2. on the results of cohesive zone crack propagation analyses.” Key Eng.
Hansbo, A., and P. Hansbo. 2004. “A finite element method for the sim- Mater. 251–252: 313–318. https://doi.org/10.4028/www.scientific.net
ulation of strong and weak discontinuities in solid mechanics.” Comput. /KEM.251-252.313.
Methods Appl. Mech. Eng. 193 (33–35): 3523–3540. https://doi.org/10 Scheider, I., A. Nonn, A. Völling, A. Mondry, and C. Kalwa. 2014. “A
.1016/j.cma.2003.12.041. damage mechanics based evaluation of dynamic fracture resistance
Hillerborg, A., M. Modéer, and P.-E. Petersson. 1976. “Analysis of crack in gas pipelines.” Procedia Mater. Sci. 3: 1956–1964. https://doi.org/10
formation and crack growth in concrete by means of fracture mechanics .1016/j.mspro.2014.06.315.
and finite elements.” Cem. Concr. Res. 6 (6): 773–781. https://doi.org Schwalbe, K.-H., I. Scheider, and A. Cornec. 2013. Guidelines for applying
/10.1016/0008-8846(76)90007-7. cohesive models to the damage behaviour of engineering materials and
Hojjati-Talemi, R. 2016. “Numerical simulation of dynamic brittle fracture structures. New York: Springer.
of pipeline steel subjected to DWTT using XFEM-based cohesive seg- Shen, G., R. Bouchard, J. A. Gianetto, and W. R. Tyson. 2008. “Fracture
ment technique.” Frattura ed Integritá Strutturale (Fracture Struct. toughness evaluation of high strength steel pipe.” In Vol. 6 of Proc.,
Integrity) 10 (36): 151–159. https://doi.org/10.3221/IGF-ESIS.36.15. PVP2008 ASME 2008 Pressure Vessel and Piping Division Conf.,
Hojjati-Talemi, R., S. Cooreman, and D. Van Hoecke. 2016. “Finite element 1275–1282. New York: ASME.
simulation of dynamic brittle fracture in pipeline steel: A XFEM-based Shim, D., M. Uddin, F. Brust, and G. Wilkowski. 2011. “Cohesive zone
cohesive zone approach.” Proc. Inst. Mech. Eng. Part L: J. Mater. Des. modeling of ductile crack growth in circumferential through-wall
Appl. 232 (5): 1–14. https://doi.org/10.1177/1464420715627379. cracked pipe tests.” In Vol. 3 of Proc., ASME 2011 Pressure Vessels
Jiang, H. 2010. “Cohesive zone model for carbon nanotube adhesive sim- and Piping Conf., 685–690. New York: ASME.
ulation and fracture/fatigue crack growth.” Ph.D. thesis, Dept. of Siegmund, T., and W. Brocks. 2000. “A numerical study on the correlation
Mechanical Engineering, Univ. of Akron. between the work of separation and the dissipation rate in ductile frac-
Jin, Z.-H., and C. T. Sun. 2006. “A comparison of cohesive zone modeling ture.” Eng. Fract. Mech. 67 (2): 139–154. https://doi.org/10.1016
and classical fracture mechanics based on near tip stress field.” Int. J. /S0013-7944(00)00054-0.
Solids Struct. 43 (5): 1047–1060. https://doi.org/10.1016/j.ijsolstr.2005 Tvergaard, V., and J. W. Hutchinson. 1992. “The relation between crack
.06.074. growth resistance and fracture process parameters in elastic-plastic
Khoei, A. R. 2015. Extended finite element method: Theory and applica- solids.” J. Mech. Phys. Solids 40 (6): 1377–1397. https://doi.org/10
tions. Chichester, UK: Wiley. .1016/0022-5096(92)90020-3.
Lin, M. 2015. “Characterization of tensile and fracture properties of X52 Wang, E., W. Zhou, G. Shen, and D. Duan. 2012. “An experimental
steel pipes and their girth weld.” M.Sc. thesis, Dept. of Civil and study on J(CTOD)-R curves of single edge tension specimens for X80
Environmental Engineering, Univ. of Alberta. steel.” In Vol. 4 of Proc., ASME 2012 9th Int. Pipeline Conf., 241–248.
Lin, M., S. Agbo, J. J. R. Cheng, N. Yoosef-Ghodsi, and S. Adeeb. 2017. New York: ASME.
“Application of the extended finite element method (XFEM) to simulate Wells, G. N., and L. J. Sluys. 2001. “A new method for modelling cohesive
crack propagation in pressurized steel pipes.” In Vol. 3B of Proc., cracks using finite elements.” Int. J. Numer. Methods Eng. 50 (12):
ASME 2017 Pressure Vessels and Piping Conf. New York: ASME. 2667–2682. https://doi.org/10.1002/nme.143.
Liu, P. F., B. J. Zhang, and J. Y. Zheng. 2012. “Finite element analysis of Zeng, Y. 2015. “Feasibility study of cohesive zone model on crack
plastic collapse and crack behavior of steel pressure vessels and piping propagation in pipeline steel under monotonic and fatigue loading.”
using XFEM.” J. Fail. Anal. Prev. 12 (6): 707–718. https://doi.org/10 M.Sc. thesis, Faculty of Civil Engineering and Geosciences, Delft Univ.
.1007/s11668-012-9623-8. of Technology.

© ASCE 04020009-13 J. Pipeline Syst. Eng. Pract.

J. Pipeline Syst. Eng. Pract., 2020, 11(2): 04020009


Zhang, Y. M., M. Fan, Z. M. Xiao, and W. G. Zhang. 2016. “Fatigue Zhang, Z., J. Xu, B. Nyhus, and E. Østby. 2010. “SENT (single edge
analysis on offshore pipelines with embedded cracks.” Ocean notch tesion) methodology for pipeline applications.” In Proc.,
Eng. 117 (May): 45–56. https://doi.org/10.1016/j.oceaneng.2016 18th European Conf. on Fracture. Freiburg, Germany: Aedification
.03.038. Publishers.
Downloaded from ascelibrary.org by MENG Lin on 02/07/20. Copyright ASCE. For personal use only; all rights reserved.

© ASCE 04020009-14 J. Pipeline Syst. Eng. Pract.

View publication stats J. Pipeline Syst. Eng. Pract., 2020, 11(2): 04020009

You might also like