You are on page 1of 7

Proceedings of the 2012 9th International Pipeline Conference IPC2012 September 24-28, 2012, Calgary, Alberta, Canada

IPC2012-90653

SIMULATION OF DUCTILE CRACK PROPAGATION IN HIGH-STRENGTH PIPELINE STEEL USING DAMAGE MODELS
Aida Nonn Salzgitter Mannesmann Forschung GmbH Duisburg, Germany Christoph Kalwa EUROPIPE GmbH Mlheim an der Ruhr, Germany

ABSTRACT The performance of engineering design of high-strength steel pipelines has revealed the necessity to revise current design procedures. Therefore, an improved and detailed comprehension of fracture mechanisms and development of failure prediction tools are required in order to derive new design criteria. In last decades the most successful failure prediction tools for steel structures subjected to various type of loading can be encountered in the field of damage mechanics. This paper aims to describe ductile fracture behavior of highstrength steel pipelines by applying three different damage models, Gurson-Tvergaard-Needelman (GTN), Fracture Locus Curve (FLC) and Cohesive Zone (CZ). These models are evaluated regarding their capability to estimate ductile crack propagation in laboratory specimens and linepipe components without adjusting the calibrated parameters. It can be shown that appropriate parameter sets can be identified to reproduce load-deformation and fracture resistance curves accurately. The strain rate effect on the fracture behavior is examined by dynamic tests on the BDWT specimens. Finally, the shortcomings of the applied models are pointed out with the reference to possible extensions and modifications. INTRODUCTION Recently, the development of new high-strength steel grades has been promoted within many pipeline research projects with the objective to cover the requirements resulting from long-distance transport of oil and gas. The safe and at the same time cost-effective use of these steel grades can be only ensured, if it is possible to fully characterize the mechanical properties and understand the fracture performance of the pipelines under given loading conditions. In this context, the influence of different material (e.g. strain hardening, strain rate sensitivity, etc.) and geometry (e.g. diameter to wall thickness

ratio, etc.) factors on the fracture behavior should be quantitatively described for consideration in pipeline design standards. Furthermore, the outcome of full-scale tests [1] and [2] has shown that the current pipeline standard rules for prediction of crack arrest are no longer applicable for new grades since they are based on semi-empirical correlations, [3][5]. Hence, an improved and detailed comprehension of dynamic fracture mechanisms and development of failure prediction tools are required in order to modify the pipeline design procedures. Accompanying the development of suitable experimental methods, damage mechanics models as a part of numerical approach have successfully demonstrated their capability to capture damage mechanisms at the local scale. The main objective of this paper is to provide a comparative study with respect to the description of ductile fracture for X100 material between three different widely used damage models, Gurson-Tvergaard-Needleman (GTN), Fracture Locus Curve (FLC) and Cohesive Zone (CZ). The application of these models requires the determination of corresponding parameters. In case of GTN model, the parameters have already been identified and verified by metallographic analyses and tests on notched tensile, fracture mechanics (SENB and SENT type) and Battelle-Drop-WeightTear (BDWT) specimens in [37] and [38]. The existing experimental data from tensile tests on notched specimens with three different radii is used to derive the Fracture Locus Curve. The CZ model parameters are adjusted by means of test results on the deep notched SENT specimens until an agreement between numerical and experimental data points is achieved. Finally, all three damage models are scrutinized regarding the accuracy to reproduce shallow notched SENT and BDWT test results. Based on these findings possible modifications are proposed for the improvement of the model prediction capability.

Copyright 2012 by ASME

DUCTILE DAMAGE MECHANICS MODELS Gurson-Tvergaard-Needelman (GTN) Model The micromechanics-based GTN model describes the ductile failure process consisting of nucleation, growth and coalescence of micro-voids at second phases and inclusion sites inside the material. The major disadvantages of the original GTN model [6]-[9] include strong mesh size sensitivity and incapability to capture slant fracture appearance under shear-dominated stress state typical for low-constraint geometries, such as thin-walled pipelines. One possibility to overcome this obstacle in simulation of long-distance cracks lies in frequent automatic remeshing by using a fine grid mesh in the fracture zones only at the time when fracture occurs [10] and [11]. Another possibility refers to usage of non-local damage model with microstructural length directly implemented in the material constitutive law [12]. Regarding simulation of shear fracture, the original GTN model has to be modified to describe shear softening due to void distortion and shear localization under low triaxiality. In [13] and [14] the void shape is introduced as an additional parameter to capture the influence of sheardominated stress on damage behavior. In order to address the issue of damage induced softening in shear-dominated deformation, the third invariant of stress with Lode parameter has been used allowing to distinguish between axisymmetric and shear-dominated stress state, [15]-[20]. Further modifications of the original GTN model are related to the plastic anisotropy especially pronounced for high strength pipeline steels [21]. Fracture Locus Curve (FLC) Model Generally, ductile crack initiation in metallic materials is controlled by stress triaxiality and equivalent plastic strain [22][25]. The latter decreases with increasing level of stress triaxiality at initiation. The relation between these two variables at the onset of stable crack growth can be expressed by the failure locus curve, which represents an empirically derived exponential limit curve. In analogy to modification of GTN model, the extension [26]-[28] of the original failure locus curve [24] has been proposed to account for damage accumulation in combined shear and void growth modes. The derivation of the fracture locus curve requires a testing program to cover the range of different stress states and parallel numerical analyses for the evaluation of local stress variables (hybrid experimental-numerical analysis). For the stress triaxiality levels >1.0, tensile tests on the round notched bar specimens have proven to be the most convenient, since they allow for simple setting of different stress states by varying the notch geometry. Cohesive Zone (CZ) Model The widely used CZ model originally introduced in [29] and [30] can be used for the simulation of brittle but also ductile fracture [31]-[34] in the metallic materials. According to this

model the fracture process occurs in the cohesive zone by overcoming the existing cohesive strength between two adjacent virtual surfaces. The numerical realization of the CZ model has been achieved using cohesive elements, the interface elements defined between the continuum elements. The underlying constitutive law between the traction and displacement provides the phenomenological description of the fracture process. The major advantage of this model lies in its robustness and the fact that only three parameters, cohesive strength T 0, separation 0 and the cohesive law are required. Identification of the damage parameters The GTN damage parameters already determined from previous analyses [38] are summarized in Table 1. The crack growth with the GTN model is simulated with model parameter f*. The crack extends over one element when material losses its stress carrying capacity at f*=fu*=1/q1. Table 1
f0

GTN model parameters


fN eN sN fC k q1 q2 ly [mm] 0.2

1.5x10-4 0.005 0.3 0.1 0.02 4.0 1.5 1.0

In order to derive the FLC, the evaluation of the local stress/strain field values is required in terms of stress triaxiality h and equivalent plastic strain evpl. Therefore, elastic-plastic calculations are performed with Abaqus/Standard without including the damage development. The local field values are determined at the location, at which the experimental crack initiation occurs. Depending on the notch geometry, the fracture onset takes place either on the notch root (for notch radius =0.2) or in the center of the specimen (for notch radii =0.5 and 1.0mm). The resulting 3 h/evpl data pairs for all 3 notch geometries are used for the construction of the FLC, which can be described by an exponential function, see Figure 1.
1.0

equivalent plastic strain e vpl

0.8 0.6
0.4 0.2

r=0.2mm r=0.5mm r=1.0mm Fracture Curve Fit

evpl =3.99e-1.57h

0.0
0.0 1.0 2.0 stress triaxiality h 3.0

Figure 1

Derivation of FLC

Besides the definition of failure curve, FLC model also requires the specification of the post-initiation softening for the crack propagation. After the initiation in a material point within

Copyright 2012 by ASME

an element, additional work is required for the crack to propagate through the element. This work can be quantified as fracture energy, which is considered in the exponential damage evolution law. In this paper, GTN and FLC models already implemented in Abaqus/Explicit are used for the simulation of ductile fracture. The parameters of CZ model are calibrated by means of fracture mechanics tests on the deep notched SENT specimen with a/W=0.52. The resulting parameters are given by T0=1300MPa and 0=90N/mm. Here, the cohesive law [35] developed by Helmholtz Zentrum Geesthacht (HZG) is applied for the simulation of ductile crack propagation. This law has been embedded in the user subroutine UEL for Abaqus/Standard [36]. In the simulation with CZ model, the interface elements open during loading and finally lose their stiffness leading to the disconnection of continuum elements. Hence, the crack propagation is only possible along the element boundaries. As shown in [37], the cohesive elements provided by Abaqus are not suitable to describe ductile fracture in metallic materials without performing additional modifications. MATERIAL PROPERTIES The X100 material for the investigations is extracted from the same section and location (approx. 6 oclock) of the large diameter pipe section with outer diameter OD=48 and wall thickness wt=18.4mm. The longitudinally welded pipe has been manufactured by EUROPIPE GmbH in the UOE process from the thermo-mechanically rolled plate. The mechanical properties in terms of yield and tensile strengths have been obtained by tensile tests on the smooth round bar specimens. Figure 2 shows the resulting flow curve extended to 100% true plastic strain with Hollomon potential function.
1000

process after the pipe production, it is assumed that thermal ageing effects are manifested in pronounced yield strength and significant Lders plateau compared to untreated pipes. The 0.2% yield strength Rp0.2=756MPa and tensile strength Rm=757MPa lead to very high Y/T ratio of about 100% and thus to low strain hardening exponent. The metallographic analyses have been performed with objective to identify and quantify the relative microstructure entities. Light optical microscopy (LOM) on the polished and HNO3 etched sample shows ferritic-bainitic microstructure for X100 steel characterized by ferrite grains aligned in the rolling direction and bainite bands, see Figure 3. By applying LOM to the polished unetched sample, large inclusions with diameter d>1m have been made visible. Inclusions of the type calcium sulphide (CaS), titanium nitride (TiN) and aluminium oxides (Al2O3) are expected due to the chemical composition of the material. In contrast to above-mentioned inclusions, the detachment of the particles with diameters d<1m leading to formation of secondary voids should occur at larger strains. There are two major types of particles contributing to the nucleation of secondary voids, bainite packages, which consists of finest precipitated iron carbides (Fe3C cementite), and M-Aconstituents (combination of martensite + retained austenite). These two microstructural entities can be observed by SEM of HNO3 etched surfaces at high magnifications, see Figure 3. In numerical simulation with GTN model, the volume fraction of secondary voids has been considered with the parameter fN.
300nm

30m

1:500

1:40000

Flow curve

900

True stress [MPa]

iron carbides ferrite retained austenite

martensite

Figure 3
800

Microstructure of X100 from LOM (left) and SEM (right) analyses

Hollomon potential function:


700
n Ke p

K=898MPa, n=0.032
600

0.0

0.2

0.4 0.6 0.8 True plastic strain [-]

1.0

Figure 2

True stress-strain curve for X100

The geometry of the tested specimen is defined by diameter D=8mm and gauge length of L=40mm. The course of the engineering stress-strain curves for X100 exhibits pronounced yield strength ReH=801MPa and wide range of Lders strain. Since the pipeline material was subjected to a simulated coating

FRACTURE MECHANICS TESTS The results from fracture mechanics tests on pin-loaded SENT specimens with initial crack ratio a/W=0.26 is used to validate the proposed damage parameters. Two clip gauges located 1.5mm and 16.9mm away from the specimen surface serve to determine crack opening displacements COD1 and COD2 and fracture toughness in terms of CTOD and J-integral values. The fracture resistance at different crack extensions is obtained by multi-specimen technique with 8 SENT specimens. A 3D FE model is created and calculated using all three different models. Special attention is given to the modeling of the contact between the pin and the welded grips in order to describe the specimen rotation accurately. The numerical model

Copyright 2012 by ASME

Opening displ. COD1 [mm]

contains solid 8-node elements (C3D8R) with reduced integration in the damaged area. Additionally, elastic-plastic FE calculations are performed without taking into account damage process. The experimental load-deformation curve with the load drop due to the damage induced softening can be well reproduced by all three damage models. The deviation from the elastic-plastic curve indicates the onset of the failure, see Figure 4.
120 100 80
Load [kN]

2.4 2.0 1.6


SENT 13x13x200, a/W=0.26 COD1=1.6374(a)0.5397

1.2 0.8 0.4 0.0 0.0 0.2 0.4 0.6 0.8 1.0 Crack growth a [mm]

Exp. Curve Fit FEM-GTN FEM-FLC FEM-CZM

1.2

1.4

60 40 20
SENT 13x13x200, a/W=0.26 Exp. FEM-ELPL FEM-GTN FEM-FLC FEM-CZM

Figure 6

Fracture resistance COD1 vs. a

0
0.0 0.5 1.0 1.5 2.0 2.5 Opening displ. COD1 [mm]

Figure 4

Load-displacement curves for SENT specimens

The profiles of the ductile tearing along the thickness are presented in Figure 5. Although none of the models in their original versions is capable to simulate slant fracture at the specimen surface, the CZ model provides best approximation of the experimental crack front with crack slightly preceding in the mid-plane. In contrast, distinct difference in the crack extensions between mid-plane and surface is obtained with GTN and FLC models.

INSTRUMENTED BDWT TESTS The purpose of the BDWT (Batelle-Drop-Weight-Tear) test in the pipeline industry is to avoid brittle fracture based on the evaluated fracture surface appearance and transition temperature. Within this study BDWT test results are used to scrutinize the suitability of the used models to describe the fracture behavior subjected to dynamic loading. By applying the same parameters as in the quasi-static case, all three models are able to describe correctly the load-time curve with slight underestimation of the load drop onset, see Figure 7. The strain rate dependence of the material is accounted for in the plasticity model by defining the yield strength values as a function of different strain rate levels. The difference in the accuracy of numerical predictions becomes evident with respect to the slope of the declining curve in the steady-state region. The delayed crack growth with FLC model leads to the slower descent in comparison to the experimental one. The best agreement to experimental measurements is obtained with CZ model.
500

GTN
mid-plane surface

initial crack tip

FLC

CZM

400
Load F [kN]

Figure 5

Simulated crack fronts (thick red line)

300

Regarding fracture resistance in terms of COD1 vs. a, all numerical curves correspond well to the experimental one as shown in Figure 6. The best estimation of the fracture resistance can be achieved with GTN model during entire crack propagation. Both FLC and CZ models slightly overestimate the experimental curve. While the level of the COD1 at crack initiation is negligibly higher than the testing value, the deviation of the CZ model predicted curve increases continuously with growing crack. The opposite is the case for FLC model with initiation COD1 value distinctly higher when compared to experimental one, while the slope of the curve is predicted correctly with increasing crack propagation.

200

100

Exp.1 Exp.2 FEM-GTN FEM-FLC FEM-CZM 1.0E-03 2.0E-03 3.0E-03 Time [s] 4.0E-03 5.0E-03

0 0.0E+00

Figure 7

Load-time curves for BDWT specimens

CRACK PROPAGATION IN THE LINEPIPE The following section shows the first results achieved with GTN model to simulate the crack propagation in the linepipe.

Copyright 2012 by ASME

Development of the model A 3D FE model in Abaqus/Explicit is created to simulate the dynamic ductile fracture in the pipe without considering the backfill effects. The total pipe length is set to 6.2m (ca. 5xOD). Furthermore, only a quarter pipe model is analyzed by taking into account symmetry conditions in length and circumferential directions. The initial internal pressure of 163bar is equal to 72% of the X100 yield strength. The equivalent axial tension of 270MPa resulting from the internal pressure has been employed at the one end of the pipe. The selected 0.2mm thickness of the predefined damage elements layer for X100 corresponds to the damage parameter, element height, determined from the tests under quasi-static loading. The influence of the mesh on the fracture behavior is examined by choosing two different element lengths (0.61mm for fine mesh (FM) and 5mm for the coarse mesh (CM)) in axial direction. Like in case of SENT and BDWT specimens the whole FE mesh consists of linear 8-node brick elements (C3D8R) with reduced integration. The total number of the elements in FM and CM is 250,460 (Model 1) and 1,801,356, (Model 2) respectively. It is obvious that the simulation of the dynamic propagation in the model with the FM requires a significant increase of the CPU time in comparison to the coarser meshed model. The element size in thickness direction is selected to 1.53mm for X100 material resulting from 12 element layers over 18.4mm thickness. The mesh becomes coarser with the increasing distance from the damage element region. Besides the description of the dynamic fracture resistance with suitable ductile damage models, the biggest challenge in the simulation lies in the implementation of the complex loading scenario. In the first step, the pragmatic approach has been chosen by substituting the fluid dynamic loading with simplified loading boundary conditions. This approach has been realized in Abaqus/Explicit with the user subroutine VDLOAD, by which two loading regions have been defined, one with flap opening occurrence behind the moving crack tip and the other in front of it. As already demonstrated in [39] based on the experimental findings [40], the pressure decay behind the crack tip can be estimated with an exponential function. Figure 8 shows the equivalent plastic strains of the deformed pipe after the crack propagation of ca. 6m. The results have been obtained from the simulation with Model 1 mesh and GTN model. The deformed pipe profile also displays the typical wave pattern (wrinkling), which according to [41] originate from the plastic stretching and subsequent formation of the bulge in direct vicinity of the crack tip. The fracture process zone is limited to the transition region between the crack tip and the flattened necking zone.

typical wave pattern (wrinkling)

Figure 8

Deformed pipe due to crack growth with GTN model

The comparison between two models (Model 1 and Model 2) for X100 with respect to crack tip location as a function of time is presented in Figure 9. The calculations are conducted with GTN model using the same parameters as for the simulation of SENT and BDWT tests. As evident in Figure 9, there is a pronounced mesh size effect in the direction of the crack propagation. The reduction of the element size in the axial direction by factor 8.3 leads to the decrease of the crack propagation velocity from 330m/s to 288m/s. In order to reduce the computing time, the local remeshing procedure should be reconsidered or the application of non-local GTN model. In addition to the GTN model, the dynamic crack propagation will be also reproduced with the FLC and CZ models.
6.0
Crack tip distance from the pipe center, m
5.0

350 crack velocity crack tip distance


Crack velocity, m/s

300
250

4.0 3.0 2.0 1.0


0.0

200
150

Model 1
Model 2

100
50

8 10 12 Time, ms

14

16

18

Figure 9

Crack propagation vs. time, comparison between Model 1 and 2

CONCLUSIONS The major conclusions from the conducted comparative study are: Ductile damage behavior of pipeline steel material can be successfully characterized using three different damage mechanics approaches By applying damage parameters calibrated on quasistatic tests, the maximum load level including the

Copyright 2012 by ASME

beginning of the load drop can be quantitatively well described for BDWT test Although the slant fracture cannot be reproduced, the result indicates that all models might be suitable for modeling of dynamic fracture propagation in the linepipe Further analyses are required to address the issues such as mesh sensitivity, influence of the single parameters and the reduction of the model size and subsequently computational time

11

12

13

ACKNOWLEDGMENTS The authors would like to thank EUROPIPE GmbH and German Academic Exchange Service for the generous financial support. REFERENCES 1 Vogt, G., Re, G. and Demofonti, G., EPRG recommendation for crack arrest toughness for high strength line pipe steels, 9th biennial joint technical meeting on line pipe research, Houston: PRCI/EPRG, 1993, pp. 25.1-25.12 2 Demofonti, G., Mannucci, G., Hillenbrand, H.-G. and Harris, D., Suitability evaluation of X100 steel pipes for high pressure gas transportation pipelines by full scale tests, 14th joint technical meeting on pipeline research, EPRG-PRCI-APIA, 2003 3 Maxey, W., Keifner, J. F., and Eiber, R. J., Ductile Fracture Arrest in Gas Pipelines, A.G.A. Catalogue number L32176, May 1976 4 Maxey, W., Dynamic crack propagation in line pipe, Sih G., Mirabile M., editors, Analytical and experimental fracture mechanics, Rome, 1981, pp. 109-123 5 Wiedenhoff, W., Vogt, G. and Peters, P., Toughness requirements of large-diameter line pipe in gas transmission pipelines, International seminar on fracture in gas pipelines, Moscow, 1984, pp. 95-117 6 Gurson, A., Continuum theory of ductile rupture by void nucleation and growth: Part I-yield criteria and flow rules for porous ductile media, J. Eng. Mater. Technology, 1977, 99, pp. 2-15 7 Gurson, A., Continuum theory of ductile rupture by void nucleation and growth: Part I-yield criteria and flow rules for porous ductile media, J. Eng. Mater. Technology, 1977, 99, pp. 2-15 8 Tvergaard, V.: Influence of voids on shear band instability under plane strain conditions, Int. J. Fract., 1981, 17, pp. 389-407 9 Tvergaard, V. and Needleman, A., Analysis of the cup-cone fracture in a round tensile bar, Acta metal., 1984, 32(1), pp. 157-169 10 Perrin, G., Martinez, M., Odru, P., Luu, T., Pineau, A., Tanguy, B. and Besson, J., Towards X100 gas pipes

14

15

16

17

18

19

20

21

22

23

24

assessment against propagating axial crack hazard, ISOPE 2005, Seoul Wawrzynek, P.A. and Ingraffea, A.R., An interactive approach to local remeshing around a propagating crack, Finite Elem. Anal. Des., 1989, 5(1), pp. 87-96 Reusch, F., Svendsen, B. and Klingbeil, D., A nonlocal extension of Gurson-based ductile damage modeling, Comput. Mater. Sci., 2003, 26, pp. 219229 Gologanu, M., Leblond, J.B., Perrin, G. and Devaux, J., Recent extensions of Gursons model for porous ductile metals, Suquet, P. (Ed.), Continuum Micromechanics, Springer Verlag, 1995 Hutchinson, J.W. and Pardoen T., An extended model for void growth and coalescence, J. Mech. Phys. Solids, 2000, 48, pp. 2467-2512 Zhang, K.S., Bai, B.S. and Francois, D., Numerical analysis of the influence of the Lode parameter on void growth, Int. J. Solids Struct., 2001, 38(32-33), pp. 5847-5856 Nahshon, K. and Hutchinson, J.W., Modification of the Gurson Model for shear failure, European Journal of Mechanics A/Solids, 2008, 27, pp. 1-17 Nahshon, K. and Xue, Z., A modified Gurson model and its application to punch-out experiments, Engng. Fract. Mech., 2009, 76(8), pp. 997-1009 Jackiewicz, J., "Use of a modified Gurson model approach for the simulation of ductile fracture by growth and coalescence of microvoids under low, medium and high stress triaxiality loadings," Engng. Fract. Mech., 2011, 78-3, pp. 487-502 Nielsen, K.L. and Tvergaard, V., "Failure by void coalescence in metallic materials containing primary and secondary voids subject to intense shearing," International Journal of Solids and Structures, 2011, 489, pp. 1255-1267 Butcher, C., Chen, Z.T.: A void coalescence model for combined tension and shear, Modelling and Simul. Mater. Sci. Eng., 2009, 17(2), pp. 025007 Tanguy, B., Luu, T.T., Perrin, G., Pineau, A., Besson, J., Plastic and damage behaviour of a high strength X100 pipeline steel: Experiments and modeling, Int. J. Press. Vess. Piping, 2008, 85, pp. 322-335 McClintock, F.A, A criterion of ductile fracture by the growth of holes, J. Appl. Mech., 1968, 35, pp. 363371 Rice, J.R. and Tracey, D.M., On the ductile enlargement of voids in triaxial stress fields, J. Mech. Phys. Solids, 1969, 17, pp. 201-217 Hancock, J.W. and Mackenzie, A.C., On the mechanisms of ductile failure in high-strength steels subjected to multi-axial stress states, J. Mech. Phys. Solids, 1976, 24, pp. 147-169

Copyright 2012 by ASME

25 Schlter, N., Grimpe, F., Bleck, W. and Dahl, W., "Modelling of the damage in ductile steels," Computational Materials Science, 1996, 7, pp. 27-33 26 Bao, Y. and Wierzbicki, T., On fracture locus in the equivalent strain and stress triaxiality space, Int. J. Mech. Sci., 2004, 46, pp. 81-98 27 Bai, Y. and Wierzbicki, T., A new model of metal plasticity and fracture with pressure and Lode dependence, Int. J. Plast., 2008, 24, pp. 1071-1096 28 Bai, Y. and Wierzbicki, T., Application of the extended Coulomb-Mohr model to ductile fracture, Int. J. Fract., 2010, Vol. 161, pp.1-20 29 Dugdale, D.S., Yielding of steel sheets containing slits, J. Mech. Phys. Solids 8, 1960, pp. 100-104 30 Barenblatt, G., The mathematical theory of equilibrium cracks in brittle fracture, Advances in Applied Mechanics 7, 1962, pp. 55-129 31 Negre, P., Steglich, D. and Brocks, W., Crack extension at an interface: prediction of fracture toughness and simulation of crack path deviation, Int. J. Fract., 134, 2005 3-4, pp. 209 229 32 Roy, Y.A. and Dodds, R.H., Simulation of Ductile Crack Growth in Thin Aluminum Panels Using 3-D Surface Cohesive elements, Int. J. Fract., 110, 2001, pp. 21-45 33 Scheider, I., Schdel, M., Brocks, W. and Schnfeld, W., Crack Propagation Analyses with CTOA and Cohesive Model: Comparison and Experimental Validation, Engng. Fract. Mech, 73, 2006, pp. 252263 34 Anvari, M., Scheider, I. and Thaulow, C., Simulation of dynamic ductile crack growth using strain-rate and

35

36

37

38

39

40

41

triaxiality-dependent cohesive elements, Engng. Fract. Mech., 2006, 73, pp. 2210-2228 Scheider, I. and Brocks, W, Simulation of cup-cone fracture using the cohesive model. Engng. Fract. Mech., 2003, 70, pp. 943961 Scheider, I., The Cohesive Model Foundations and Implementation, Helmholtz Zentrum Geesthacht (HZG), 2nd Edition (11/2011) Nonn, A. and Kalwa, C., "Anwendung numerischer Modelle zur Simulation des duktilen Bruchverhaltnes in den hochfesten Pipelinesthlen," Tagungsband der DVM Tagung Werkstoffprfung, Berlin, Germany, pp. 219224, 2011 Nonn, A. and Kalwa, C., "Modeling of damage behavior of high strength pipeline steel," Proc. of ECF18, 363_proceeding.pdf, 18th European Conference on Fracture, Dresden, Germany, 2010 Shim, D.J., Wilkowski, G., Rudland, D., Rothwell, B., Merrit, J., Numerical simulation of dynamic ductile fracture propagation using cohesive zone modeling, Proceedings of the 7th International Pipeline Conference IPC2008, Calgary, Canada, 2008 Berardo, G., Salvini, P., Mannucci, G. and Demofonti, G., On Longitudinal Propagation of a Ductile Fracture in a Buried Gas Pipeline: Numerical and Experimental Analysis, Proceedings of the International Pipeline Conference, Calgary, Canada, Vol. 1, pp. 287-294, 2000 Schindler, H.J., Prediction of ductile crack extension and arrest by an analytical approach, Proceedings of the 8th International Pipeline Conference IPC2010, Calgary, Canada, 2010

Copyright 2012 by ASME

You might also like