147
University of Calgary PRISM: University of Calgary's Digital Repository Graduate Studies The Vault: Electronic Theses and Dissertations 2019-01-11 Finite Element Modeling of Buried Longitudinally Welded Large-Diameter Oil Pipelines Subject to Fatigue Anisimov, Evgeny Anisimov, E. (2019). Finite Element Modeling of Buried Longitudinally Welded Large-Diameter Oil Pipelines Subject to Fatigue (Unpublished master's thesis). University of Calgary, Calgary, AB. http://hdl.handle.net/1880/109466 master thesis University of Calgary graduate students retain copyright ownership and moral rights for their thesis. You may use this material in any way that is permitted by the Copyright Act or through licensing that has been assigned to the document. For uses that are not allowable under copyright legislation or licensing, you are required to seek permission. Downloaded from PRISM: https://prism.ucalgary.ca

Finite Element Modeling of Buried Longitudinally Welded

  • Upload
    others

  • View
    5

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Finite Element Modeling of Buried Longitudinally Welded

University of Calgary

PRISM: University of Calgary's Digital Repository

Graduate Studies The Vault: Electronic Theses and Dissertations

2019-01-11

Finite Element Modeling of Buried Longitudinally

Welded Large-Diameter Oil Pipelines Subject to

Fatigue

Anisimov, Evgeny

Anisimov, E. (2019). Finite Element Modeling of Buried Longitudinally Welded Large-Diameter Oil

Pipelines Subject to Fatigue (Unpublished master's thesis). University of Calgary, Calgary, AB.

http://hdl.handle.net/1880/109466

master thesis

University of Calgary graduate students retain copyright ownership and moral rights for their

thesis. You may use this material in any way that is permitted by the Copyright Act or through

licensing that has been assigned to the document. For uses that are not allowable under

copyright legislation or licensing, you are required to seek permission.

Downloaded from PRISM: https://prism.ucalgary.ca

Page 2: Finite Element Modeling of Buried Longitudinally Welded

UNIVERSITY OF CALGARY

Finite Element Modeling of Buried Longitudinally Welded Large-Diameter Oil Pipelines Subject

to Fatigue

by

Evgeny Anisimov

A THESIS

SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

IN PARTIAL FULFILMENT OF THE REQUIREMENTS FOR THE

DEGREE OF MASTER OF SCIENCE

GRADUATE PROGRAM IN MECHANICAL ENGINEERING

CALGARY, ALBERTA

JANUARY, 2019

© Evgeny Anisimov 2019

Page 3: Finite Element Modeling of Buried Longitudinally Welded

ii

Abstract

The design and construction of large diameter buried pipelines primarily for crude oil

transportation is governed in Canada by CSA Z662, ASME B31.4, and ASME BPVC Section VIII.

Although these codes provide general guidelines on pipeline design, many aspects of modelling

the pipeline are not given in detail, and the results can vary significantly based on how these details

are modelled. Engineers often adopt a very conservative approach and this results in pipelines that

are overdesigned and therefore unnecessarily costly. Following the design code, this thesis

provides a detailed fatigue analysis (FA) of a large diameter buried liquid pipeline and incorporates

the effects of the stress concentrations associated with manufacturing defects and tolerances. A

stress analysis of the pipe is first performed using the finite element method (FEM), and results

obtained are used in conjunction with both elastic and elastic-plastic FA life assessment models to

predict fatigue damage (FD). The results of a FEM and FA performed on four standard pipeline

OD’s show that a 20% increase in the outside diameter (OD) to wall thickness (WT) ratio can be

achieved when plasticity is considered. This is equivalent to one to two increments of standard

WT or the percent reduction of a pipeline construction cost. In the analyses process, where the

code leaves significant room for interpretation, this thesis provides clarity on appropriate

procedures to follow. Examples include how to accurately model the weld profile, and the

misalignments due to the manufacturing process. Furthermore, a simple calculation tool is

developed that can be used to approximate hot-spot elastic stresses.

Keywords: Large Diameter Pipeline Fatigue, Fatigue of Welded Connections, Elastic-Plastic

Fatigue Analysis, Fatigue Damage.

Page 4: Finite Element Modeling of Buried Longitudinally Welded

iii

Preface

The thesis focuses on the design codes related to fatigue analysis of pipelines and discusses

the challenges associated with implementation of the codes, analyzing the procedures of North

American codes in more detail. Special attention is paid to manufacturing misalignments of the

pipes’ weld region and analysis of stresses in that region. The magnification of stresses due to

misalignments is further discussed from the standpoint of fatigue analysis. Addressed the global

aim of the research – development of the simple and easy to use model that can help engineers

with reliable assessment of design stresses in the pipe and its fatigue-safe design.

Chapter 1 provides the background information on the pipe manufacturing processes,

materials used to build the pipelines and their properties related to fatigue degradation. Further

discussion focuses on the elastic and the elastic-plastic models of materials’ behavior, including

the von Mises and the Tresca yielding criteria. Finally, the basis of the three main fatigue

assessment procedures are discussed, including the stress life and the strain life approaches dealing

with non-planar defects (such as pores and others metallurgical defects), as well as the crack

growth approach used for the assessment of structures with planar defects (such as cracks and weld

undercuts).

Chapter 2 is dedicated to manufacturing tolerances used in pipeline design and

manufacturing quality control. The various types of misalignments (manufacturing defects) are

discussed, including weld discontinuity, offset of pipe plate at the weld region, peaking of the weld

region, and ovality of the pipes’ body. As a summary, the research problem is formulated, and the

pipe design parameters selected for the model development, including complete pipe geometry

with manufacturing defects, materials, and pipe loading.

Page 5: Finite Element Modeling of Buried Longitudinally Welded

iv

Chapter 3 shows the steps taken toward the development of both the mathematical (Hand

calculations) and the final element (ABAQUS calculations) models used for calculation of stress-

strain states of the pipeline due to various loadings. The computing was focused on obtaining of

the stresses and strains at the most critical location in the pipe – structural hot-spot (at the weld

toe). The models capture the effects of various types of misalignment, internal pressure, soil

pressure, and temperature, on the stress rise at the hot-spot, including bending stress developed.

Chapter 5 provides concluding remarks and discusses major results of the research work

discussed in this thesis. The significant increase of structural stresses due to misalignment was

demonstrated with the help of elastic and elastic-plastic analyses. The stress rise resulted in

dramatic increase of the fatigue damage due to cyclic pressurizing of the pipeline. Another

important observation is the conservatism involved in the elastic fatigue analysis. Elastic-plastic

fatigue analysis suggested the possible reduction of the pipe wall thickness without compromising

the fatigue performance of the modeled pipeline. Newly implemented accounting for the weld

profile, not observed in the standards before, can provide the

Chapter 6 outlines the future work and recommends the areas for improvement. Accounting

for the residual stresses due to manufacturing and during cyclic loading in the model can be very

important for more detailed analysis of stress-strain states at the critical locations and can be

extremely useful when accompanied by the more advanced fatigue assessment methodologies

based on fracture mechanics principles. The crack growth approach in fatigue analysis would

utilize the accurate stress-strain data at the crack tip to yield more accurate predictions of fatigue

damage.

Page 6: Finite Element Modeling of Buried Longitudinally Welded

v

Acknowledgements

I would like to acknowledge Dr. Meera Singh and Dr. Les Sudak for their professional

guidance and valuable critical discussions related to the research work discussed in this thesis.

I am grateful for the opportunity to participate in the project that addressed some real-life

challenges that engineers have in the pipeline industry; I was able to research the problems

associated with structural integrity and safety. The research project benefited me professionally

and personally, I was interacting with professionals from the industry, learned many new things

and furthered my knowledge during my studies.

I would also like to thank Darryl Stoyko and Robert Thom from Stress Engineering

Services, Inc. for critical discussions and reviews.

Page 7: Finite Element Modeling of Buried Longitudinally Welded

vi

Dedication

I would like to dedicate the thesis to my family, especially to my wife Natalia, to my sons,

Maxim and Denis, and thank them for all the support and understanding provided during my

studies. I also dedicate the thesis to my grandfather Vladislav Anisimov a civil engineer who

sparked my interest to the field of engineering.

Page 8: Finite Element Modeling of Buried Longitudinally Welded

vii

Table of Contents

Abstract .............................................................................................................................. ii Preface ............................................................................................................................... iii Acknowledgements ............................................................................................................v Dedication ......................................................................................................................... vi Table of Contents ............................................................................................................ vii

List of Tables .................................................................................................................... ix List of Figures and Illustrations .......................................................................................x List of Symbols, Abbreviations and Nomenclature .................................................... xiii

Epigraph ......................................................................................................................... xvi

Chapter 1 INTRODUCTION ...................................................................................1 1.1. Motivation ...............................................................................................................1 1.2. Background ............................................................................................................2

1.3. Objective .................................................................................................................4 1.4. Thesis Outline .........................................................................................................5

Chapter 2 LITERATURE REVIEW.......................................................................7 2.1 Early FE Analyses ..................................................................................................8 2.2 Recent FE Analyses..............................................................................................12

2.3.1. Hot-Spot Stress ...............................................................................................12

2.3.2. Weld Distortion and Wall Thickness ...........................................................15 2.3 Governing Codes ..................................................................................................19 2.4 Conclusions of Literature Review ......................................................................20

Chapter 3 STANDARD PROCEDURES ..............................................................21 3.1. Pipeline Codes ......................................................................................................21

3.2. Pipeline Geometry ................................................................................................23 3.2.1. Pipe ODs and WTs .........................................................................................24 3.2.2. Weld Misalignments ......................................................................................26

3.2.2.1. Radial Misalignment .............................................................................28

3.2.2.2. Angular Misalignment ..........................................................................30

3.2.2.3. Ovality Misalignment ...........................................................................30 3.2.3. Welding Defects ..............................................................................................31 3.2.3.1. Weld Reinforcement .............................................................................31

3.2.3.2. Welding Cracks .....................................................................................33 3.3. Pipe-Soil Interaction ............................................................................................34 3.4. Pipeline Materials ................................................................................................35 3.5. Loading .................................................................................................................42 3.6. Linearization of Stresses......................................................................................44

3.7. Analytical Model ..................................................................................................45 3.7.1. Stress due to Misalignment ...........................................................................45

3.7.2. Stress due to Soil ............................................................................................47 3.8. Fatigue Assessment ..............................................................................................49 3.8.1. Stress-Life Curves ..........................................................................................50 3.8.2. Elastic Fatigue Analysis .................................................................................52

Page 9: Finite Element Modeling of Buried Longitudinally Welded

viii

3.8.3. Modified Elastic Fatigue Analysis ................................................................53 3.8.4. Elastic-Plastic Fatigue Analysis ....................................................................56

3.8.5. Elastic Fatigue Analysis of Welds .................................................................57 3.9. Summary and Problem Definition .....................................................................61

Chapter 4 MODEL DEVELOPMENT .................................................................63 4.1. Static Finite Element Model ................................................................................64 4.1.1. Geometry of Model ........................................................................................64

4.1.2. Material Model ...............................................................................................69 4.1.3. Boundary Conditions .....................................................................................73 4.1.4. Model Meshing and Convergence ................................................................74

4.1.5. Data Extraction ..............................................................................................77

Chapter 5 RESULTS AND DISCUSSION ...........................................................79 5.1. Finite Element Model ..........................................................................................79 5.1.1. Validation of FEM .........................................................................................83

5.2. Fatigue Analysis ...................................................................................................97

Chapter 6 CONCLUSIONS AND FUTURE WORK ........................................105

6.1. Conclusions .........................................................................................................105 6.2. Future Work .......................................................................................................107

APPENDIX A – MATLAB Numerical Solution .........................................................108

APPENDIX B – MATLAB Cycle-Counting ................................................................110

APPENDIX C – ABAQUS Input File ..........................................................................112

APPENDIX D – ABAQUS Report Example ...............................................................118

APPENDIX E – MATLAB Code for the ABAQUS Data ..........................................121

References .......................................................................................................................124

Page 10: Finite Element Modeling of Buried Longitudinally Welded

ix

List of Tables

Table 1 Nominal WTs for different ODs ...................................................................................... 24

Table 2 Permissible Specified ODs and WTs ............................................................................... 25

Table 3 Example DFs and WTs used in Keystone pipeline .......................................................... 25

Table 4 Permissible Variation of WT ........................................................................................... 26

Table 5 Permissible radial misalignment for different pipe thicknesses (North America) ........... 29

Table 6 Permissible radial misalignment (BS PD 5500) .............................................................. 29

Table 7 Permissible angular misalignment (BS PD 5500) ........................................................... 30

Table 8 Permissible ovality misalignment (API 5L) .................................................................... 31

Table 9 Permissible weld reinforcement, inch (mm) .................................................................... 32

Table 10 Soil properties ................................................................................................................ 34

Table 11 Constants for a polynomial fit of experimental data in the calculation of number of

cycles to failure ..................................................................................................................... 55

Table 12 Pipeline parameters [in or (mm)] considered in this research ....................................... 61

Table 13 Standard pipeline WTs for selected ODs and steel material at an internal pressure of

10 MPa .................................................................................................................................. 65

Table 14 Pipeline defects .............................................................................................................. 67

Table 15 Weld bead dimensions obtained in this study ................................................................ 68

Table 16 Pipeline steel material .................................................................................................... 69

Table 17 Displacement constraints ............................................................................................... 73

Table 18 Parameters used in calculation of SCFs due to misalignments ..................................... 84

Table 19 Stress magnification at different weld locations ............................................................ 91

Table 20 Results of analysis of the design hoop stresses Sh [MPa] for a pipe of OD 914 mm

and WT 17.5 mm .................................................................................................................. 92

Table 21 Results of fatigue analysis obtained at accumulated fatigue damage of 0.5................ 100

Table 22 Results of fatigue analysis obtained at accumulated fatigue damage of 1.0................ 103

Page 11: Finite Element Modeling of Buried Longitudinally Welded

x

Table 23 Construction cost savings associated with WT reduction on a 4700 km pipeline ....... 104

List of Figures and Illustrations

Figure 1 Misalignments of a welded pipe ....................................................................................... 3

Figure 2 Schematic of the DSAW weld profile .............................................................................. 3

Figure 3 Structural stress concept ................................................................................................. 13

Figure 4 Welded connections in: (a) seamless pipeline, (b) pipeline with longitudinal seam,

and (c) pipeline with helical seam; the hatched area shows the plane of connection of

two pipes ............................................................................................................................... 27

Figure 5 Common types of weld misalignment in longitudinally welded pipe: (a) radial, (b)

angular, and (c) ovality ......................................................................................................... 28

Figure 6 Detailed schematic of radial misalignment .................................................................... 29

Figure 7 Schematic of abutting plates before welding .................................................................. 33

Figure 8 Weld widths for different wall thicknesses of plates tapered at 60° .............................. 33

Figure 9 Mechanical properties of pipeline steels ........................................................................ 36

Figure 10 Typical engineering stress-strain tensile curves for some X steels as per API 5L [3] . 36

Figure 11 Typical engineering and true stress-strain tensile curves for X42 steel ....................... 38

Figure 12 Difference between experimental stress-strain tensile curve for X42 steel and

Ramberg-Osgood fit near the yield strength ......................................................................... 39

Figure 13 Tresca and von Mises yield criteria in (a) hydrostatic and (b) plane stresses .............. 40

Figure 14 Stress tensor at the longitudinal weld in pipeline ......................................................... 41

Figure 15 Fatigue loading showing (a) spectrum loading and (b) constant amplitude loading .... 42

Figure 16 In-service pressure history diagram.............................................................................. 43

Figure 17 Cycle-counted in-service pressure history showing Pmin, Pmax, and nk ......................... 44

Figure 18 Through-wall bending stress and ovality of pipe cross-section due to transmitted

pressure ................................................................................................................................. 48

Figure 19 Schematic classification of fatigue life approaches ...................................................... 49

Page 12: Finite Element Modeling of Buried Longitudinally Welded

xi

Figure 20 Standardized tensile specimen (a) versus real component (b)...................................... 50

Figure 21 Stress life (S-N) curve .................................................................................................. 51

Figure 22 Neuber’s relationship between linear and non-linear stresses and strains [83] [84]

[85] ........................................................................................................................................ 59

Figure 23 Flow chart showing process of model refinement ........................................................ 63

Figure 24 Geometry of model ....................................................................................................... 65

Figure 25 Schematic of the SAW-processed pipe region ............................................................. 66

Figure 26 Traces of weld profiles used to generate an average weld profile ............................... 67

Figure 27 Geometry of the weld bead profile showing (solid dots) experimental data and

(solid line) 4th-order polynomial approximation ................................................................... 67

Figure 28 Geometry of the weld region including (bold white line) radial and (bold black

line) angular misalignments .................................................................................................. 69

Figure 29 Construction of a tangent to Ramberg-Osgood’s curve from yield point on Hooke’s

curve ...................................................................................................................................... 70

Figure 30 Example of a numerical solution for the tangent point on a Ramberg-Osgood curve . 71

Figure 31 True Stress-Strain curves for pipe steel material .......................................................... 72

Figure 32 Pipe (highlighted by circles) surrounded by a soil box with constraints ...................... 73

Figure 33 Meshing of a pipe showing detailed meshing at the hot-spot (black line indicates

the path used for an SCL) ..................................................................................................... 74

Figure 34 Meshing of soil box around pipe .................................................................................. 75

Figure 35 Refinement of global (away from discontinuity) and local (at the weld toe) meshes

showing von Mises Stress/Strain – Mesh Element Size relationship ................................... 76

Figure 36 Hoop stress distribution maps for a misaligned pipe during elastic loading ................ 80

Figure 37 Hoop stress distribution maps for a misaligned pipe during elastic-plastic loading .... 80

Figure 38 Through-thickness (curved) actual and (linear) linearized stress distributions

obtained for a pipe of 914 mm OD and 14.3 mm WT from an SCL positioned at the hot-

spot (at 0 mm WT coordinate) normal to the pipe wall with no misalignment by using

(a) elastic and (b) elastic-plastic analysis, and with misalignment by using (c) elastic and

(d) elastic-plastic analysis ..................................................................................................... 81

Page 13: Finite Element Modeling of Buried Longitudinally Welded

xii

Figure 39 Schematic of (a) the SAW butt joint, represented in the form of (b) fillet in stepped

flat bar, showing (c) equivalent load at the base of reinforcement and (d) real shear

stress diagram with its (dashed line) approximation ............................................................. 85

Figure 40 The (dots) SCFs for different (connected dots) transition radiuses Wr (a)

Wr =0.25WT, (b) Wr =7.145 mm, (c) Wr=3-7/t, and (d) Wr=5 mm ...................................... 89

Figure 41 Secondary bending (curved arrows) due to: (a) axial, (b) angular, and (c) ovality

misalignments, and (d) due to soil; the red dashed line indicates the plane of a

hypothetical crack or SCL, and 1 through 4 are the hot-spot locations ................................ 90

Figure 42 Hoop stress calculated with mathematical model, power-law-fitted, and

extrapolated until solutions of (dashed line) non-misaligned and (solid line) misaligned

conditions intersect (power-law-fitted) ................................................................................. 93

Figure 43 Solutions for Hoop stress (linear fit) in (non)misaligned pipe of OD 914 mm (a)

without and (b) with km.weld accounted .................................................................................. 94

Figure 44 Solutions for Hoop stress (power law fit) in pipe of OD 914 mm (a) without km.weld

and (b) with km.weld ................................................................................................................. 96

Figure 45 Accumulated fatigue damage plots for pipe diameters (a) 610 mm, (b) 864 mm, (c)

914 mm, and (d) 1219 mm, calculated with (solid lines) misalignment and with (contour

lines) no misalignment .......................................................................................................... 99

Figure 46 Relationship between OD and WT at accumulated fatigue damage of 0.5 for (blue)

BS elastic, (red) ASME elastic, and (grey) ASME elastic-plastic analyses ....................... 102

Page 14: Finite Element Modeling of Buried Longitudinally Welded

xiii

List of Symbols, Abbreviations and Nomenclature

Symbol Unit Definition

𝐴0, 𝐵0 Geometric constants for a fillet weld

BD [𝑚𝑚] Burial depth

CAE Computer aided engineering

𝐶𝑢𝑠 [0 ÷ 1] Conversion factor

𝐷 [𝑚𝑚] Outside diameter

𝐷𝑓 [0 ÷ 1] Accumulated FD

𝐷𝑙 [1.0 ÷ 1.3] Deflection lag factor

𝑑 Factor for nominal probability of failure

𝐸 [𝑀𝑃𝑎] Young’s modulus

𝐸′ [𝑀𝑃𝑎] Modulus of soil reaction (≈ 0 ÷ 20 for loose to compact soil)

𝐸𝑇,𝑘 [𝑀𝑃𝑎] Young’s modulus at assessed temperature

𝐸𝐹𝐶 [𝑀𝑃𝑎] Young’s modulus of material used to obtain experimental S-N curve

FD Fatigue damage

FEM Finite element method

𝐻 [𝑚𝑚] Pipe BD

𝐾 [𝑀𝑃𝑎] Strength coefficient

𝐾𝑏 [≈ 0.1] Soil bedding constant

𝐾𝑒,𝑘 Fatigue penalty factor

𝑘 Coefficient of deformation of the weld joint

𝑘𝑚 Stress magnification factor

𝑙1,2 [𝑚𝑚] Weld section lengths supporting the shear distributed load

𝑚 Slope of S-N curve

𝑁𝑘 Number of cycles to failure

𝑛𝑘 Number of assessed cycles

𝑛 [0 ÷ 1] Strain hardening exponent

OD [𝑚𝑚] Outside diameter

𝑃𝑖 [𝑀𝑃𝑎] Internal pressure

𝑃𝑠 [𝑀𝑃𝑎] Soil pressure on the pipe above the water table

Page 15: Finite Element Modeling of Buried Longitudinally Welded

xiv

𝑅 Stress ratio

SAW Submerged arc-welded

𝑆𝑎 [𝑀𝑃𝑎] Stress amplitude

𝑆𝑎𝑙𝑡,𝑘 [𝑀𝑃𝑎] Effective alternating equivalent stress

𝑆𝑒 [𝑀𝑃𝑎] Fatigue endurance limit

SCF Stress concentration factor

SD Standard deviation

𝑆𝐿 [𝑀𝑃𝑎] Stress in pipe’s longitudinal direction

𝑆𝐻 [𝑀𝑃𝑎] Stress in pipe’s hoop direction

𝑆𝑟 [𝑀𝑃𝑎] Stress range in most critical direction

𝑇 [℃] Assessment temperature

𝑡 [𝑚𝑚] Wall thickness

UOE U-to-O shaped and expanded

𝑈𝑖 [𝑚𝑚] Displacement components

𝑊𝑇 [𝑚𝑚] Wall thickness

𝑊𝑤 [𝑚𝑚] Weld bead width

𝑊ℎ [𝑚𝑚] Weld bead height (reinforcement)

𝑊𝑟 [𝑚𝑚] Weld toe radius

𝑊𝛼 [°] Weld reinforcement angle

𝛼 [℃−1] Coefficient of linear thermal expansion

𝛼𝑓 [°] Friction angle

𝛼𝑑 [°] Dilation angle

𝛿𝑜 [𝑚𝑚] Offset misalignment

𝛿𝑝 [𝑚𝑚] Peaking misalignment

휀 [𝑚𝑚 𝑚𝑚⁄ ] Strain

휀𝑒𝑛𝑔 [𝑚𝑚 𝑚𝑚⁄ ] Engineering strain

휀𝑝 [𝑚𝑚 𝑚𝑚⁄ ] Plastic strain

휀𝑡𝑟𝑢𝑒 [𝑚𝑚 𝑚𝑚⁄ ] True strain

휀𝑦 [𝑚𝑚 𝑚𝑚⁄ ] Yield strain

𝜈 [0 ÷ 0.5] Poisson’s ratio

Page 16: Finite Element Modeling of Buried Longitudinally Welded

xv

𝜌 [𝑘𝑔 𝑚3⁄ ] Mass density (unit weight of soil fill)

𝜌𝑚 Material structural characteristic constant

𝜎 [𝑀𝑃𝑎] Normal stress

𝜎𝑏 [𝑀𝑃𝑎] Bending stress (through-wall)

𝜎𝑒𝑛𝑔 [𝑀𝑃𝑎] Engineering stress

𝜎𝑚 [𝑀𝑃𝑎] Membrane stress

𝜎𝑚𝑖𝑗,𝑘 [𝑀𝑃𝑎] Component stresses at the end of cycle

𝜎𝑛𝑖𝑗,𝑘 [𝑀𝑃𝑎] Component stresses at the start of cycle

𝜎𝑢 [𝑀𝑃𝑎] Ultimate tensile stress

𝜎𝑡𝑟𝑢𝑒 [𝑀𝑃𝑎] True stress

𝜎𝑦 [𝑀𝑃𝑎] Yield stress at room temperature

𝜎𝑦𝑇 [𝑀𝑃𝑎] Yield stress at assessed temperature

𝜎𝑦𝐶 [𝑀𝑃𝑎] Cohesion yield stress

𝜏𝑚 [𝑀𝑃𝑎] Shear stress at the fillet weld section

𝜏𝑚′ [𝑀𝑃𝑎] Maximum value of the shear stress at the rectangular reinforcement

∆𝑝𝑖𝑗,𝑘 [𝑚𝑚 𝑚𝑚⁄ ] Change in plastic strain range components for the 𝑘𝑡ℎ loading cycle

∆𝑆𝑃,𝑘 [𝑀𝑃𝑎] Effective equivalent stress range

(∆𝑦

𝐷) Deflection (or ovality) due to soil

∆𝑦 [𝑚𝑚] Vertical deflection of pipe due to soil

∆휀𝑒𝑓𝑓,𝑘 [𝑚𝑚 𝑚𝑚⁄ ] Effective strain range

∆휀𝑒𝑙,𝑘 [𝑚𝑚 𝑚𝑚⁄ ] Elastic strain range

∆휀𝑝𝑒𝑞,𝑘[𝑚𝑚 𝑚𝑚⁄ ] Plastic equivalent stain range

∆𝜎𝑖𝑗,𝑘 [𝑀𝑃𝑎] Range of normal component stresses

∆𝜎 [𝑀𝑃𝑎] Normal stress range

∆𝜎′ [MPa] Additional local stress at upper fillet weld in stepped flat bar

∆𝜎′′ [MPa] Additional local stress at lower fillet weld in stepped flat bar

∆𝜎′′′ [MPa] Additional local stress at both fillet welds in stepped flat bar

∆𝜏 [𝑀𝑃𝑎] Shear stress range

∆𝜃 Ovality misalignment

Page 17: Finite Element Modeling of Buried Longitudinally Welded

xvi

Epigraph

In the behavior of structures, truth can usually be found only by testing and observing

genuine structural members built with the materials as they are, with imperfections which cannot

be avoided.

Fritz Leonhardt, 1976

~ Gold Medalist, The Institution of Structural Engineers

Page 18: Finite Element Modeling of Buried Longitudinally Welded

1

Chapter 1 INTRODUCTION

1.1. Motivation

According to Natural Resources Canada, Canada is the fourth-largest producer and

exporter of crude oil in the world and owns 10% of the world’s proven oil reserves (as of December

2017) [1]. Approximately 5.1 million barrels of crude oil per day [MMb/d] are transported within

Canada via 840,000 km of pipeline infrastructure, of which 117,000 km is pipeline with a diameter

of up to 48 inches. Currently, 62.4% of large-diameter transmission pipelines in Canada are

federally regulated [1]-[2]. Multiple reports of pipeline rupture events have been accumulated over

40 years by the National Energy Board (NEB) [2], showing that during the last decade alone,

nearly 50% of the reported incidents related to crude oil transportation were caused by cracking

due to fatigue. These fatigue related incidents resulted in more than 54,000 barrels of crude oil

spilled, which is more than 60% of the total amount spilled during the time period.

It is very important to account for all possible factors that may negatively impact the

integrity and reduce the fatigue life of a pipeline. The pipeline fatigue design must incorporate the

manufacturing defects and in-service conditions, both of which introduce stresses/strains, which

for safe operation should not exceed critical values at critical locations. These design

considerations are governed by design codes. In order to avoid failures while performing stress

analyses and fatigue life predictions for large-diameter pipes, engineers generally attempt to follow

the codes and to incorporate a conservative approach in their designs. However, this often results

in a greater wall thickness (WT) than required. This consequently contributes to a considerable

increase in a pipeline’s overall cost. Therefore, optimizing a pipe WT can result in significant

savings when multi-kilometer lines are being designed.

Page 19: Finite Element Modeling of Buried Longitudinally Welded

2

1.2. Background

Larger pipes normally have outer diameters (OD’s) ranging between 457.0 and

2,174.0 mm, wall thicknesses (WT’s) ranging between 7.1 and 52.0 mm, and come in standard

lengths between 6.0 and 24.0 m, corresponding to standards API 5L [3] and CSA Z245.1 [4]. In

Canada, such pipes are usually buried 0.6 to 1.2 m below the ground, as per CSA Z662 [5]. They

are normally manufactured by the cold-forming of flat steel plates using the U-ing O-ing and

expanding (UOE) process. During this process, the plate is first formed into a U-shape and then

pressed into an O-shape between two semicircular dies [6]. Subsequently, the longitudinal seam is

welded using the Double Submerged Arc Welding (DSAW) process [7] to connect the abutting

edges of the deformed plate and complete an O-shape. Finally, the pipe is expanded using an

internal mandrel to improve its roundness [8].

Pipelines, particularly liquid-carrying ones, are subjected to repeated thermal and pressure

loads while in service. This cyclic loading can result in fatigue failures at loads much lower than

those observed in static failures. Geometric discontinuities within the loaded pipe act as stress

risers that result in magnification of a local stress [9]. These stress risers promote fatigue failure,

as they act as perfect sites for cracks to initiate. Several such stress risers in large-diameter pipe

can originate from the UOE manufacturing process. Specifically, they can occur due to peaking

and radial misalignment, as shown in Figure 1, in the vicinity of the SAW seam. Furthermore, the

double-SAW welding process used to create the longitudinal seam of a potentially misaligned pipe

can introduce a large stress concentration due to the weld profile, as shown in Figure 2.

Page 20: Finite Element Modeling of Buried Longitudinally Welded

3

Figure 1 Misalignments of a welded pipe Figure 2 Schematic of the DSAW weld profile

In order to predict the fatigue life of a large-diameter pipe, all stress risers must be

accounted for. As an example, it was found that cyclic tensile stresses can increase by 50% by

increasing the angle between the weld reinforcement and the base plate, 𝑊𝛼, from 0° to 60° [10]

[11]. A similar strong influence on stress development and fatigue strength has been reported to

be caused by the reinforcement, 𝑊ℎ, the weld radius, 𝑊𝑟, [11] [12], and the axial (offset), 𝛿𝑜, or

angular (peaking), 𝛿𝑝, weld misalignment [12] [13]. Although this topic is being actively

researched, there is a lack of data in the available literature on the analysis of welded pipes with

combined misalignments.

The UOE and SAW processes produce residual stresses in pipe, the most detrimental of

which from a fatigue perspective are the residual tensile stresses observed after welding. Although

the stress distribution after welding [14] can become more uniform after expansion [8], the

uncertainties in the experimentally measured data discussed in [15] [16] and [17] indicate

significant variation in residual stresses that may ambiguously influence fatigue tests on welded

structures [18] or may have a barely noticeable impact on fatigue strength [19]. However, narrower

weld beads were reported to result in slightly lower residual stresses [20]. Another stress

magnification factor is the over-pressurization overloads frequently observed in the engineering

practice, which have been found to accelerate the failure of a component due to fatigue [18].

Page 21: Finite Element Modeling of Buried Longitudinally Welded

4

In Canada, pipelines are generally designed with reference to the codes such as API 5L [3],

API 579 [21], ASME BPVC Section VIII Part 2 [22], ASME B31.4 [23], CSA Z245.1 [4],

CSA Z622 [5], BS 7608 [24], and BS 7910 [25]. Although the codes are critical to assuring the

safe operation of pipelines, the application of code requirements is often not straightforward. For

example, the assessment methodology in the British standards BS 7608 [24] and BS 7910 [25],

used to account for the stresses due to weld misalignments, is limited to hand calculations, and its

implementation in FEM analysis is not discussed. A similar situation exists with North American

standards API 579 [21] and ASME BPVC Section VIII Part 2 [22], although FEM is among the

preferred methods for stress-strain analyses. Furthermore, the geometry of weld profiles seems

also not to be well defined in the North American standards. As a result of these uncertainities in

the codes, engineers are often left to provide their best guess on how to move forward in their

analysis. This can lead to underdesign or overdesign of the pipelines analyzed.

1.3. Objective

The overall objective of this project is to provide a systematic assessment methodology for

fatigue life that can be used to design and analyze pipelines so that they are safe from a fatigue

perspective and are also cost-efficient. In this work, some difficulties associated with the

interpretation of codes, modeling misalignments due to the UOE process, and modeling weld seam

profiles are highlighted. The work discusses these challenges in the context of a stress analysis

performed on a standard large-diameter pipe using FEM, and includes the effects of manufacturing

tolerances, misalignments and their combinations, and typical thermo-mechanical loading. The

FEM results are subsequently used to make fatigue life predictions and to guide best practices in

large-diameter pipe design.

Page 22: Finite Element Modeling of Buried Longitudinally Welded

5

Therefore, to achieve the objective of the project, the research focuses on the calculation

of acceptable pipe WT, including combinations of commonly observed manufacturing

misalignments from the perspective of fatigue-safe design, including axial and angular

misalignments applied to a range of commonly used pipe diameters. It also uses the Level 2 Fatigue

Assessment Methods from ASME BPVC Section VIII Part 2 [22] and discusses the conservatism

involved in elasticity-based methods when selecting the WT for a given pipe design.

1.4. Thesis Outline

Chapter 2 overviews the published literature associated with the stress and fatigue analyses

of welds used in large-diameter pipelines as well as the pipe manufacturing tolerances governed

by the design codes. Emphasis is placed on allowable pipe weld geometry and weld misalignment

defects.

Chapter 3 provides a discussion on the development of the analytical and FE models used

in the stress analysis of buried oil-carrying pipes, and includes the geometry of the model, the

material model, boundary conditions, and the model meshing. The calculation procedures

described in Chapter 3 closely follow the guidance from pipe design codes, and discussion is

provided regarding the limitations of the codes as they relate to the modeling of weld profiles and

combinations of weld misalignments. Finally, the chapter introduces the elastic and the elastic-

plastic fatigue assessment methodologies used in predictions of fatigue damage, including a

discussion on the rainflow cycle-counting of load history and the linearization of stresses obtained

from FEM.

Page 23: Finite Element Modeling of Buried Longitudinally Welded

6

Chapter 4 discusses the results of the analyses described in the previous chapter and

provides a detailed assessment of allowable design wall thicknesses for the studied pipe

geometries. This chapter highlights the significance of combined misalignment with regard to the

fatigue damage of pipes, and discusses the conservatism involved with the elastic fatigue

assessment methods.

Chapter 5 summarizes the results and presents the conclusions of the research work

accomplished in this thesis. It further provides the basis for future work and specific

recommendations to further the understanding of the mechanical behavior of misaligned in-service

large-diameter pipes.

Page 24: Finite Element Modeling of Buried Longitudinally Welded

7

Chapter 2 LITERATURE REVIEW

The first oil-carrying pipelines in North America were built in the second half of the 19th

century, shortly after oil was discovered. Pipe manufacturers turned relatively quickly to steel,

which offered higher strength to weight ratios than cast or wrought iron. They also enabled lower

processing temperatures that favored the fabrication of longitudinally welded larger pipes by

means of the lap welding technique [26]. The growing need for oil transportation capacity led to

the development of better-quality steels with fewer defects and to advances in welding processes.

For example, the (double) submerged-arc welding technique was introduced after World War II

and proved to be more reliable than lap joining in the furnace [26]. The design and manufacturing

of pressure vessels and pipes in North America have been governed by ASME VIII code since

1925, by API 5L since 1928, and by ASME B31 since 1935 [27].

Pipeline design has greatly benefited from improvements in quality-assurance methods and

inspection tools. These include hydrostatic testing standards, the introduction of radiographic

inspection in 1948 and in-line inspection in the 1960s [26]. Studies of pipeline failures have

recognized the need for addressing the issues related to fatigue cracking [26]. Miner’s work on

cumulative fatigue damage (1945) and Langer’s design fatigue curves for pressure vessels (1961)

have greatly contributed to the pipeline design capabilities offered by the design codes in the 2000s

[27].

The following sections will discuss in detail the difficulties associated with the FE

modeling of the welded connections when following the governing codes. Specifically, literature

pertaining to FE modeling of weld profiles and weld misalignments in pipelines and their effects

on the stresses and fatigue damage are reviewed.

Page 25: Finite Element Modeling of Buried Longitudinally Welded

8

2.1 Early FE Analyses

Pollard and Cover (1972) [10], both American metallurgists, concluded in their literature

review on the fatigue in steel weldments by stating “…fatigue strength is determined primarily by

the geometry of the weldment and the soundness of the weld metal, and so even if the weld

reinforcement is ground off, the fatigue strength may still be influenced by the soundness of the

weld metal”.

Since this thesis focuses on the fatigue performance of UOE pipes influenced by

geometrical tolerances that are acceptable by North American standards, the literature reviewed in

this thesis is concerned with the FE modeling of pipes’ weld profiles and related challenges. FE

modeling has improved greatly over recent years mainly due to advances in computing

capabilities. Therefore, the widely utilized analytical solutions could further benefit from and be

extended by numerical FEM-aided solutions. However, the design codes used for piping do not

provide detailed guidance on FE modeling.

The effects of various geometrical parameters of a full-penetration butt weld on fatigue

strength were studied by Berge and Myhre at Det Norske Veritas in 1977 [28]. Using experimental

and FE analysis the authors showed that misalignment may seriously deteriorate the fatigue

strength of such welds. In Canada, early FE analyses of misaligned line pipes were conducted by

Worswick and Pick (1985) at the University of Waterloo [29]. Their study showed a significant

reduction in limit load (transition from elastic to plastic behaviour) associated with increases of

local bending stress due to weld misalignment, signifying the need to account for these effects in

pipeline design. Later, in 1991, Ferreira and Branco [30] discussed the importance of strength

analysis of misaligned welded joints as opposed to the “good-practical-experience” exercised in

the governing codes at that time when specifying permissible levels of misalignment, which were

Page 26: Finite Element Modeling of Buried Longitudinally Welded

9

found to reduce the fatigue strength of welded components markedly. In all three mentioned FE

studies, the mesh size used in the FE models was rather coarse with sharp transitions at the weld

discontinuity, and no convergence study had been performed due to the expenses associated with

computing capabilities at that time [29]. Nonetheless, the study [30] demonstrated a strong

influence of both axial and angular distortions on the local stress at the weld toe in tension (loaded

normal to the plane of a hypothetical crack) and therefore on the fatigue strength of longitudinally

welded tubular structures.

Berge and Myhre [28] used 2D quadrilateral and triangular elements that appear to be first-

order and of an average size of 1.25-2.22 mm for a global mesh, and a size of 0.1 mm was used

for the elements at the stress concentration. The mesh size was refined to a depth of only 1.6 mm

at a total through-thickness size of 20 mm. The authors in [28] concluded that the large scatter in

the published S-N data and derived design curves, which were most probably influenced by

accidental misalignments (𝛿0 𝑡⁄ ≈ 0.125), prohibited close comparison to the FEM results. This

highlighted that the optimization of welds must consider the ratio between the bending and axial

(membrane) stresses 𝜎𝑏 𝜎𝑚⁄ , which depend on the geometry in the vicinity of the weld and

boundary conditions 𝜆, when allowable misalignment is assessed. The study also identified one

main problem with the assessment of 𝜎𝑏 𝜎𝑚⁄ for each joint in question, which was impractical and

led to the selection of conservative values of 𝜆. Nowadays, this may pose much less of a problem

because of the much more advanced computational tools available, which allow for relatively

quick FEM analyses even for a highly detailed 3D model.

Another important observation made by Berge and Myhre [28] was related to the increase

of a toe angle (flank angle), 𝑊𝛼, with increased (axial) misalignment, 𝛿0, that contributes to local

bending stresses and thus impairs the fatigue strength. The bending stresses were found to decrease

Page 27: Finite Element Modeling of Buried Longitudinally Welded

10

rapidly with plate thickness. Although, a complete FEM analysis was not done to quantify the

effect of toe angle alone, the combination of 𝛿0 and 𝑊𝛼 indicated an almost 40 % decrease in

fatigue strength at eccentricity 𝛿0 = 𝑡 2⁄ , which resulted in 𝑊𝛼 = 26.6°. It may be obvious to

assume that the combination of axial and angular weld distortions would increase the toe angle,

and therefore the bending stresses at the toe even more.

Worswick and Pick [29] used quadrilateral, i.e., quadratic isoparametric (second-order),

elements that included 8-node quadrilaterals for 2D and 20-node brick elements for a 3D model

for both elastic and the elastic-plastic analyses. Although the authors in [29] used more advanced

mesh elements to address the issue with a singularity at a critical location, no mesh convergence

study was published. The stress-strain curve which was obtained through the uniaxial tensile

testing in the FEM model was approximated by a series of linear segments. The axial misalignment

of 2.286 mm, modeled for a pipe of 11.2 mm WT and 914 mm OD, resulted in approximately a

15% decrease in limit load. The authors in [29] assumed that the FEM approach might be an

alternative to potentially difficult analytical solutions in fracture mechanics techniques.

Ferreira and Branco [30] used isoparametric 8-node elements in 2D model as well. The

authors used strain-life (e-N) approach to calculate crack initiation life and linear fracture

mechanics approach to calculate crack propagation (a-N) life and, therefore, have predicted total

(S–N) life, which was verified experimentally. Aside from the general observation of reduced

fatigue strength with mainly angular (peaking, 𝛿𝑝) misalignment, and that the governing codes

need to shift from good-practical-experience to more detailed strength analysis, the authors showed

that the inner weld toe is a more significant contributor to fatigue strength reduction than the outer

weld toe. Furthermore, the work of Ong and Hoon [31] clearly indicated the need for more strict

regulation of the angular distortion, 𝛿𝑝, and the authors proposed eliminating or moderating it, or

Page 28: Finite Element Modeling of Buried Longitudinally Welded

11

otherwise the 1% of diametral deviation (ovality or out-of-roundness) allowable in ASME BPVC

(1989) may yield a non-conservative design. Interestingly, the same allowable ovality can still be

found in recent North American codes, including ASME BPVC (2015).

Andrews [32] modeled distorted welds with cubic isoparametric (third-order) elements and

performed an elastic FE analysis, and showed a reduction of fatigue strength due to axial

misalignment similar to that published by Berg and Myhre [28]. A good correlation between the

experimental results and the results of FEM was found when the stress ranges were measured at a

distance of several millimeters away from the weld toe [32]. Interestingly, stress extrapolation

techniques based on a similar idea are commonly used in modern practice (BS 7608 [24]). It would

be worth adding here the efforts of Berg and Myhre [28] to separately analyze the bending stress,

𝜎𝑏, and membrane stress, 𝜎𝑚, at the critical location (hot-spot). Nowadays, this is implemented

via the stress linearization techniques used in some modern codes (API 579 [21] and ASME BPVC

Section VIII [22]).

Analytical solutions to the problem of weld misalignments were also extensively studied

and supported by FE analysis. The solutions for welded plates developed by Maddox (1985) [33],

whose research contributed greatly to the development of British standards BS 7910 [25] and BS

7608 [24] as well as others, were furthered by Andrews in 1996 [32]. Discrepancies between the

theoretical work of Zeman (1994) [34] [35] and FEM results for longitudinally welded pipes were

critically addressed and refined by Ong and Hoon in 1996 [31].

During the second half of the 20th century, the research on weld misalignments involving

FEM analysis mainly focused on the axial and the angular weld distortions separately. It is worth

noting the observations of Berge and Myhre [28] regarding the effects of weld misalignment on

the weld profile. There are complex synergetic effects due to actual welding processes that

Page 29: Finite Element Modeling of Buried Longitudinally Welded

12

influence the geometry of the welded region, and these combinations need to be carefully analyzed

to more accurately predict the fatigue properties of a welded structure.

2.2 Recent FE Analyses

The previous research involving FEM suggested the need for the standardization of some

principles and procedures to bring more agreement in the FEM community.

Since 1998, the ASME BPVC committee, led by Osage, started completely rewriting

Section VIII Division 2, and in 2007 published the modernized version of the code with the critical

changes discussed in [36], and updated it in 2015 [37]. One of the significant changes introduced

in the revised code was the estimation of the minimum WT by using design-by-analysis (DBA)

requirements in Part 5 of the code, which included the (mesh-insensitive) structural stress concept

for computation of the membrane, 𝜎𝑚, and bending, 𝜎𝑏, stresses at the critical location using FEA

as well as the procedures for elastic and elastic-plastic FEM analysis and models for the stress-

strain curves. Recommendations for the linearization of stresses at the critical location were

obtained by using FEA. In the design-by-rule (DBR) requirements, weld efficiency factors were

introduced in order to account for the types of welded connections and the explicit design rules for

combined loadings.

2.3.1. Hot-Spot Stress

The accuracy of FEM predictions of stresses at the critical locations (hot-spots), i.e., the

weld toe in the case of longitudinally welded UOE pipes, subsequently influences the estimation

of fatigue life, and is strongly dependent on the weld details and dimensionality (2D verse 3D) as

well as on the material model used in FEM (elastic verse elastic-plastic) and on the methodology

used to extract these stresses (i.e., extrapolation verse linearization).

Page 30: Finite Element Modeling of Buried Longitudinally Welded

13

In 2001 Dong published research [38] on the definition and procedure for obtaining

structural stress, 𝜎𝑠 = 𝜎𝑚 + 𝜎𝑏, which has been claimed to yield FEM mesh-size-insensitive

results even for a high-stress singularity. Assuming equilibrium between the axial, 𝜎, and shear, 𝜏,

stresses, this procedure transforms a complex stress state at a critical location (weld toe) of a

structure into an equivalent simple structural stress state (Figure 3), as opposed to the stress

extrapolation techniques for reference points away from the location of concern (i.e. the weld toe).

Stress extrapolation approximates the local stress at the critical location and cannot correctly

capture the stress concentration effects in locations with apparent weld joint discontinuities, and

thus may not be reliable according to Dong [38]. One of the codes that uses various stress

extrapolation techniques is BS 7608 [24]. The structural stress concept is today used in ASME

BPVC VIII [22].

Figure 3 Structural stress concept

Doerk et al. (2003) [39] systematically examined various methods for calculating the

structural stresses in welded structures by using FEM and discussed the validity of claims made

by Dong [38]. Shell elements (8-node quadratic) and solid elements (20-node isoparametric,

reduced-integration) were used in [39] to model 2D and 3D welded joints, and the use of a finer

mesh of at least 0.4𝑡 was suggested when using higher-order elements. However, one element in

the thickness direction is sufficient. While the weld is frequently omitted from the FEA based on

Page 31: Finite Element Modeling of Buried Longitudinally Welded

14

shell elements, or weld geometry is significantly simplified [39], solid elements allow for the weld

to be modeled in greater detail. In contrast to 2D models, the calculation of structural stress was

reported to show mesh-size sensitivity in 3D models, especially when a stress concentration

becomes more localized due to geometry (at the weld toe). Doerk concluded the research in [39]

with the statement “…the structural hot-spot stress approach remains to be relatively coarse,

however, very practical…”. Although the 3D models allow for complex shapes to be analyzed by

FEM and potentially lead to more accurate results, care should be exercised when ensuring their

validity.

Rohart et al. (2015) [40] addressed the conservatism involved in the elastically calculated

design limits of pressure vessels. The modeling of more complex geometries and loading scenarios

leads to more conservative designs compared to results based on a more realistic elastic-plastic

representation of a material’s behaviour and accounts for local effects more accurately. This

observation was supported by Möller et al. (2017) [41], who reported lower scatter in results when

plasticity is considered in the FE model. Therefore, it is not only the geometry problem when

modeling structures with discontinuities, but also in modeling the material’s behaviour.

Goyal, El-Zein, and Glinka (2016) [42] proposed a method for the calculation of mesh-

independent peak stress at the hot-spot via extraction of the through-thickness membrane and

bending components of true non-linear stress from a coarse-mesh FE model by the stress

linearization technique. In [42], cruciform weld joints were investigated, and while the through-

thickness variation of the membrane stress was shown to be mesh-insensitive, the bending

component showed mesh-insensitivity in the middle part of the plate section, 0.75𝑡 ≥ 𝑥 ≥ 0.25𝑡,

and increased at the weld toe with the mesh element size. This was used in [42] as the basis for

that proposed method, which uses linearly distributed bending stresses to calculate the peak stress

Page 32: Finite Element Modeling of Buried Longitudinally Welded

15

at the hot-spot. It can be true that a linear distribution of bending stress may exist in the case of a

symmetric geometry of a welded region. However, this may not be true in the case of non-

symmetric welded regions such as the ones found in UOE pipes, especially when they are

misaligned. Moreover, Goyal, El-Zein, and Glinka (2016) [42] reported up to a 10% difference in

calculated peak stress compared to the fine-mesh FE model. The finest mesh elements were

modeled at the weld toe, which had a size of about 0.2 mm, and 6 of these elements comprised the

weld toe radius at 𝑊𝑇 𝑊𝑟⁄ = 2.9. Furthermore, the differences between elastic and elastic-plastic

material models, which are both frequently used in FEM, may further increase the discrepancies.

Therefore, the simplified method for calculation of the peak stress at the hot-spot proposed in [42]

may lead to ambiguous results in the case of welds with combined misalignment.

2.3.2. Weld Distortion and Wall Thickness

Lillemäe et al. (2012) [43] investigated the effect of combined misalignment on the fatigue

strength of thin-wall butt-welded specimens by using both experimental testing and FE modeling,

which included detailed weld profiles. Tensile load was applied to specimens and S–N curves

plotted using notch stress range. The modeled weld was represented by fine mesh elements for

plane strain elements at the weld toes with sizes ranging from 0.025 mm to 0.1 mm, and other

locations were modeled with plane stress elements of 0.1-5.0 mm size. The structural hot-spot

stress was obtained by linear surface stress extrapolation. It was shown that thinner sections

undergo greater straightening of the welded region in tension and thus promote greater bending

compared to thicker sections. The corrected analytical solution for angularly distorted welds

presented by Kuriyama et al. [44] in 1971 accounted for the straightening and yielded a solution

for the stress concentration factor (SCF) that is within the range obtained by the FEM.

Interestingly, a similar formulation is now adopted in BS 7910 [25].

Page 33: Finite Element Modeling of Buried Longitudinally Welded

16

Another interesting observation by Lillemäe et al. [43] was that the structural stress could

be under- or over-estimated depending on the weld toe location, and their results showed that the

SCF rapidly increases with angular misalignment, especially at the weld toe located inside the

angled weld profile. This would correspond to the weld toe located inside of an outward peaking

pipe, and structural stresses can increase even more when axial misalignment is also considered,

in which case there will be a need to distinguish between two inside weld toe locations because

one of them would have a greater flank angle, 𝑊𝛼, and thus experience greater bending. The

opposite was reported for weld toes located outside the angled weld profile. Indeed, analytical

solutions available to date seem to be incapable of distinguishing between different weld toe

locations, and the calculated SFC and hot-spot stress is therefore averaged. This may lead to non-

conservative estimates of a design limit. Overall, the study in [43] represents a detailed FE

modeling example that discusses many geometrical parameters of a misaligned weld. However,

the 2D linear-elastic model used in the study may not adequately represent the local behavior of a

real structure, and the 2D model may not adequately capture the stress states of complex geometry.

There are no details on the convergence study of developed FE model. Furthermore, [43] focused

on the behaviour of a weld joint and neglected any surrounding structures.

Nykänen and Björk (2015) [45] analyzed the effects of weld geometry of as-welded butt

joints on stress concentrations at the weld toe using data available in the literature. Constant

amplitude fatigue tensile test results were analysed using the design S–N curve [24] based on

nominal stress. The authors discussed the stress concentrations in these welds due to angular

(0.2-0.3°) and axial (0.1-0.2 mm) misalignments, and mentioned possible variations of the SCF

depending on the weld toe location. The authors attempted to assess the effects of combined

misalignments using conservative analytical solutions when the misalignments were known, or by

Page 34: Finite Element Modeling of Buried Longitudinally Welded

17

using the FE-based solutions developed by Anthes et al. [46] in 1993 when only WT, 𝑡 (3-40 mm),

and flank angle, 𝑊𝛼 (14-43°), were known, and the weld toe radius, 𝑊𝑟, was fixed to 1 mm for all

the analyzed cases. The non-linear effects were disregarded because the models used in the

calculations were based on linear material behaviour. This resulted in lower errors for thicker

sections and larger errors for thinner sections. Furthermore, the misalignments assessed in the

study [45] were rather small, i.e., a tenth of a millimeter or a degree, and were not readily available

for many of the sampled data, and may not correctly represent the longitudinal welds in UOE

pipes. Additionally, the detailed analysis of the welded region could benefit from other weld shape-

specific parameters such as weld width and height, and curvature due to pipe diameter.

Lillemäe et al. (2016) [11] observed a noticeable (30%) increase in the fatigue strength of

high-quality welds measures using hot-spot stress in S–N approach. This increase was attributed

to lower stress concentrations at the weld toe due to lower welding distortion such as undercut

(< 0.05 mm), weld height (< 1 mm), flank angle (< 30 °), and higher transition radius (> 0.5 mm).

The linear-elastic 2D FE model was built with detailed weld regions based on the actual geometry

of fatigue test specimens using plane stress elements of 0.2 mm. This model was validated with

the solid 3D model with a 2% difference, and the hot-spot stress was determined using surface

linear extrapolation.

Möller et al. (2017) [41] modeled angularly distorted butt welds with detailed weld profiles

obtained from image analysis, although neglecting the axial misalignment. In contrast to Lillemäe

et al. [43], Möller et al. (2017) [41] considered plasticity in the FE model, which showed lower

scatter in the results. The nominal stresses as well as notch stresses and strains were used to

determine the fatigue resistance from S–N curves. The fatigue life of developed weld profiles in

lower-quality butt joints was reduced, and more significantly in high-cycle fatigue.

Page 35: Finite Element Modeling of Buried Longitudinally Welded

18

Pachoud, Manso, and Schleiss (2017) [47] studied the influence of the weld profile

parameters on SCF values, and both the weld height, 𝑊ℎ, and the flank angle, 𝑊𝛼, in axially

misaligned welds were reported to significantly contribute to stress concentrations at the weld toe,

which was calculated using a 2D linear-elastic FE model. The structural stress at the weld toe was

obtained by using both surface extrapolation and through-thickness linearization. For example, in

a 30 mm-thick plate with a butt weld profile of 𝑊ℎ = 0.75 𝑚𝑚, 𝑊𝑤 = 22.3 𝑚𝑚, 𝑊𝛼 = 10 °,

𝑊𝑟 = 1 𝑚𝑚, and misalignment of 𝛿𝑜 = 2.1 𝑚𝑚, the SCF increased to 1.86. The SCF was equal

to 2.02, with an additional change in 𝑊ℎ = 0.75 → 2.1 𝑚𝑚, and the SCF reached 2.52 after an

additional change in 𝑊𝛼 = 10 → 25 °. The authors compared their FE solutions for the SCFs with

analytical solutions from the literature, and aside from the scatter within analytical solutions, they

were observed to generally overestimate the SCFs compared to the FEM solutions, and this trend

increased with WT. Apparently, this difference was due to the fact that the available analytical

solutions did not explicitly account for the weld reinforcement height, 𝑊ℎ, and width, 𝑊𝑤. The

strong influence of 𝑊ℎ and 𝑊𝛼 on the fatigue strength of butt welds has also been reported for

thinner sections [11].

Shiozaki et al. (2018) [19] conducted the S–N fatigue tests and FEM analyses on lap joints

under bending using a 2D elastic model build with plane strain elements of 0.2 mm size, reduced

to approximately 0.04-0.10 mm at the weld toe. The results of this study may be particularly useful

for the analysis of the effects of toe radius in longitudinally welded tubular structures because of

some similarities in the loading and the boundary conditions. Specifically, a simply supported

beam of a lap joint that is allowed to rotate at the supports while under pulsating bending, which

is imposed at the free ends extended outside the supports, would represent the loading condition

of an axially misaligned welded region of a pipe experiencing similar bending during pulsations

Page 36: Finite Element Modeling of Buried Longitudinally Welded

19

of the pipe’s internal pressure. Certainly, the pipe’s longitudinal weld would also experience a

tensile component of stress acting in the pipe’s circumferential direction. However, the bending

component of the total stress would contribute to the opening and closing of a hypothetical crack

at the weld toe of an axially misaligned pipe weld, similarly to a lap weld.

Therefore, qualitative analysis of the results published by Shiozaki et al. [19] could

contribute to the selection of weld toe radii for the FEM modeling of misaligned pipes. Although

a lap weld would rather exaggerate the axial weld misalignment in pipes, 𝛿0 = 𝑡, it was also

studied in early research by Berge [28] and Andrews [32]. Reduction in the 𝑊𝑇 𝑊𝑟⁄ ratio from

14.50, 5.8, 2.90, to 1.93 yielded dramatic reduction in the slopes of the S-N curve from 33 × 10−5,

25 × 10−5, 3 × 10−5, to 1 × 10−5 respectively, and showed corresponding increase in the

allowable nominal stress range from 280 MPa, 475 MPa, 620 MPa, to 650 MPa at 106 cycles. The

fatigue strength was improved with increased 𝑊𝑟, and more significantly at 𝑊𝑇 𝑊𝑟⁄ = 5.8 ÷ 2.90.

The authors in [19] also analyzed the residual stresses of welded regions in the as-welded condition

and after heat treatment, and reported insignificant impacts, if any, on fatigue strength.

2.3 Governing Codes

The pipeline governing codes guide the process of selection of WT using well established

rules and analytical solutions (ASME B31.4 [23], ASME BPVC VIII [22], BS 7608 [24], and

BS 7910 [25]). The pipeline design based on these rules (design-by-rule (BDR)) considers the

influence of manufacturing processes and manufacturing tolerances in the form of factors applied

to calculated stresses and WT. Pipeline engineer may utilize additional degree of freedom offered

by the design-by-analysis (DBA) methodology (ASME BPVC VIII [22] or API 579-1/ASME FFS-

1 Fitness-For-Service [21]) when designing a pipeline with specific detailed geometry (i.e.

including weld and misalignments). The DBA analysis is usually performed using FE modeling.

Page 37: Finite Element Modeling of Buried Longitudinally Welded

20

However, the governing codes do not provide any guidance on how to approach such modeling.

Therefore, design engineers perform FE modeling based on their experience.

2.4 Conclusions of Literature Review

The discussion of the available literature to date on the problem of misaligned pipe welds

shows a great deal of progress in both physical testing and FEM analysis aimed at assessing

stresses and fatigue strength due to critical details of the welded region. However, the studies

reported in the reviewed literature are mainly limited to the weld part loaded uniaxially and did

not include complete pipe section and effects of typical loading. Therefore, the question remains

open as to how the combination of axial and radial misalignments in large-diameter pipes could

be modeled using FEM based on manufacturing tolerances, and how it would contribute to the

development of local stresses and fatigue behaviour.

Page 38: Finite Element Modeling of Buried Longitudinally Welded

21

Chapter 3 STANDARD PROCEDURES

The modern codes used in pipeline design provide the engineer with flexible assessment

methodologies and enable the fatigue design of components with specific geometries by using FE

modeling. However, few details are provided in codes on how to model manufacturing defects and

the combination of misalignments that occur after UOE manufacturing and SAW welding of

pipelines. This Chapter focuses on code procedures and methodologies used in fatigue analysis,

including the definition of permissible pipeline geometry, pipeline materials, and loading. The

stress concentration factor (SCF) was obtained for the SAW weld studied in this research. The

SCF can be equal to one in the case of a smooth transition between the base plate and the weld

bead; however, the real SAW profile of a UOE-processed pipe is rather sharp and magnifies the

applied loads. Moreover, as was discussed earlier, governing standards make use of DFs to account

for a specific welding process and do not take into account the variability of the weld profile

parameters.

3.1. Pipeline Codes

North American codes present the basis for pipeline designs in many countries. The ASME

code for pressure piping, B31, developed in the US, has part B31.4 [23] dedicated to buried

pipeline transportation systems for liquid hydrocarbons and other liquids. The ASME B31 code

was used as a basis for the Canadian code CSA Z662 [5], similarly to CSA 245.1 [4], which is

based on API 5L [3]. Many advanced design rules reference the more complete ASME Boiler &

Pressure Vessel Code (BPVC) Section VIII [22], which is the most comprehensive standard guide

for designing efficient pressurized equipment. The ASME BPVC Section VIII [22] shares similar

design principles with API 579-1/ASME FFS-1 (Fitness-For-Service) [21].

Page 39: Finite Element Modeling of Buried Longitudinally Welded

22

A buried piping design is usually based on the minimum specified yield strength (SMYS

or 𝜎𝑦𝑚𝑖𝑛) of selected construction materials and service temperatures below 120°C (ASME B31.4

[23], ASME BPVC VIII [22]). The codes take into account manufacturing defects in the form of

design factors and safety factors. These factors are used to modify 𝜎𝑦𝑚𝑖𝑛 and calculate the allowable

design stress, 𝑆, and obtain the nominal WT, 𝑡 = 𝑃𝐷 2𝑆⁄ , for a pipeline to ensure pressure

containment [48], where 𝑆 = 𝐹 ∙ 𝐸 ∙ 𝜎𝑦𝑚𝑖𝑛, 𝐹 is a safety factor, and 𝐸 is a weld efficiency factor.

The stress acting in the hoop direction, 𝑆𝐻, of a pipe section is usually the largest and, therefore,

should be limited to 𝑆𝐻 ≤ 𝑆. The stress due to thermal gradients and the flexibility stress acting in

the longitudinal direction, 𝑆𝐿, are combined with the hoop stress, 𝑆𝐻, to further adjust the design-

allowable stress. The WT is normally adjusted to a standard value as prescribed in B31.10 [49].

This design method is named as design-by-rule (DBR), which requirements apply to commonly

used pressure vessel shapes and pressure loading.

When the tolerances provided in the code for DBR are exceeded, the engineer may follow

the design-by-analysis (DBA) procedures (from ASME BPVC VIII code [22] or API 579-1/ASME

FFS-1 Fitness-For-Service [21]) to qualify the design. The pipeline design should be evaluated for

each applicable failure mode (plastic collapse, local failure, collapse from buckling, failure from

cyclic loading) to establish the design WT. A fatigue analysis shall be performed on a piping

system to check its suitability for cyclic operating conditions according to B31.3 [50], which

references ASME BPVC VIII code [22].

The DBA methodology utilizes results from a detailed FEM stress analysis to determine

the suitability of a component. However, the DBA methodology does not provide any

recommendations on the stress analysis method. The same issue exists with European codes

Page 40: Finite Element Modeling of Buried Longitudinally Welded

23

BS 7608 [24] and BS 7910 [25]. It is expected that the design engineer will provide an accurate

validated analysis, which implies that such analysis is to be performed by a professional with

special expertise. Nevertheless, the code provides the user with all necessary pre-processing

(physical properties, strength parameters, monotonic and cyclic stress-strain curves) and post-

processing (failure modes models) tools to support the FE analysis.

3.2. Pipeline Geometry

Although it has been shown by multiple researchers that the geometrical defects of welded

pipes can greatly influence the service life of pipelines, it is difficult to eliminate the pipeline

defects associated with manufacturing. Therefore, the tolerances prescribed by the pipeline design

governing codes limit the occurrence, size, and location of possible defects. The following parts

of this chapter overview the possible defects associated with the manufacturing processes for larger

pipelines. Extensive knowledge about the existing defects is extremely important in fatigue

assessment of the component.

The tolerances used in the design of a pipeline include those for OD, WT, mechanical

properties of construction materials, operation loading, pipeline location, and weld quality/type.

The following sections focus on the geometrical pipeline tolerances in the following most recent

standards: API 5L [51], ASME BPVC Section VIII [22], CSA Z662 [5], CSA Z245.1 [4],

BS PD 5500 [52], BS 7608 [24], and BS 7910 [25].

Page 41: Finite Element Modeling of Buried Longitudinally Welded

24

3.2.1. Pipe ODs and WTs

In [5], the nominal pipe WTs for different pipe ODs are divided into two categories, as

shown in Table 1, depending on how close the pipe section is to the compressor to properly account

for a pressure drop between pumping stations.

Table 1 Nominal WTs for different ODs

OD Close to a compressor Away from a compressor

16-20 0.252 0.189

22-36 0.2-0.3 0.220

38-54 0.311 0.252

It is a common practice to design a pipeline with a larger WT close to a compressor station

and gradually reduce the design WT for pipeline sections farther away from the compressor. This

reduces the cost of pipeline manufacturing. Notably, the values for WTs can be set independently

of distance from a compressor station for ODs in the range of 22-36 in.

The nominal values of ODs and WTs are used in pipeline design as the baseline, which is

usually extended/modified according to the specific design requirements/parameters, including but

not limited to the construction material, manufacturing method, defects, operating conditions,

operating environment, and pipeline location. In general, the design parameters specify the total

critical design stress the pipeline structure must withstand during its service life with no incidents

or hazardous consequences. After all the design factors are accounted for, the pipe dimensions can

be specified as summarized in Table 2 [51]. Similar values can be found in [4], with ODs of 356-

508, 559-914, and 965-1372 mm and WTs of 4.8-7.1, 5.6-7.1, and 6.4-7.1 mm.

Page 42: Finite Element Modeling of Buried Longitudinally Welded

25

Table 2 Permissible Specified ODs and WTs

OD, inch (mm) WT, inch (mm)

Light sizes Regular sizes

14-18 (356-457) 0.177-0.281 (4.5-7.1) 0.281-1.771 (7.1-45)

18-22 (457-559) 0.188-0.281 (4.8-7.1) 0.281-1.771 (7.1-45)

22-28 (559-711) 0.219-0.281 (5.6-7.1) 0.281-1.771 (7.1-45)

28-34 (711-864) 0.219-0.281 (5.6-7.1) 0.281-2.050 (7.1-52)

34-38 (864-965) - 0.219-2.050 (5.6-52)

38-56 (965-1422) - 0.250-2.050 (6.4-52)

56-72 (1422-1829) - 0.375-2.050 (9.5-52)

72-84 (1829-2134) - 0.406-2.050 (10.3-52)

An example of utilizing the design factors in the determination of a WT of an oil

transporting pipeline with an OD of 36” made from API 5L X70/X80 steel material, operated at

10 MPa internal pressure, for the TransCanada Keystone XL pipeline design, depending on

transportation, handling, bending, welding, and installation/back-fill, is represented in Table 3

[53]. Notably, the design WT is smaller when higher-grade steels are considered. This can be seen

when comparing the WTs for API 5L X70 and API 5L X80 in Table 3. Furthermore, the total

length of the combined Keystone pipelines is designed to be around 4700 km, which represents a

significant amount of steel to be used in manufacturing and installation processes, and is associated

with great financial expenditure.

Table 3 Example DFs and WTs used in Keystone pipeline

Design Factor DF 0.80 0.72 0.60 0.50

API 5L X70: WT, mm 11.811 13.081 15.697 18.999

API 5L X80: WT, mm 10.312 11.506 13.792 16.510

Page 43: Finite Element Modeling of Buried Longitudinally Welded

26

The variations in pipe WT for welded pipes with OD > 457 mm, according to [4], are

shown in Table 4. The variations for WTs of ≤ 5, 5÷15, and ≥ 15 mm, according to [51], should

be limited to ±0.5, ±0.1t, and ±1.5 mm respectively.

Table 4 Permissible Variation of WT

WT, inch (mm) %

< 0.375 (9.5) +17, -8

0.375-0.496 (9.6-12.6) +15, -8

> 0.5 (12.7) +12, -8

Finally, common nominal lengths of a pipe are 20, 30, 40, 50, 60, and 80 ft or 6, 9, 12, 15,

18, and 24 m [51] [4].

3.2.2. Weld Misalignments

The standards governing a pipeline design limit the occurrences and sizes of manufacturing

defects within the tolerances, and prescribe design safety factors for other design parameters. The

manufacturing tolerances and safety factors have been developed through extensive experience in

pipeline operations as well as experimental research.

The plane (location) and direction of the critical stress assessed in the design depend on

the specific pipeline manufacturing process, and are often related to the location of a major

structural defect. In the case of welded structures, the most critical factor would be the plane of a

hypothetical crack at the weld root. The direction of stress depends on the actual loading, and in

most cases is normal to the plane of a hypothetical crack, and referred to as normal stress, 𝜎. Shear

stress, 𝜏, tangential to the critical plane (in-plane stress), can also develop. For instance, in a

seamless straight pipeline, the critical plane is located at the weld root between two abutting

welded pipes parallel to the pipe section, as depicted in Figure 4(a). In a longitudinally welded

Page 44: Finite Element Modeling of Buried Longitudinally Welded

27

pipe, such a plane is perpendicular to both the pipe wall and the pipe section (Figure 4(b)), and the

critical plane is located at the weld root almost perpendicular to the pipe wall. Although the

position of a critical plane in a pipe with a helical seam is variable with pipe length, it remains

tangential to the seam and perpendicular to the pipe wall (Figure 4(c)). The critical stress usually

acts normal to the critical plane. In the case of seamless pipe, this stress acts in the pipe’s

longitudinal direction and is referred to as longitudinal stress, 𝜎𝐿. It can be due to pipe-soil

interaction. In longitudinally welded pipe, the critical stress acts in the pipe’s hoop direction

(horizontally and transverse to the pipe) and is referred to as hoop stress, 𝜎ℎ, and can be due to

pipe internal pressure. In a helical seam, the stress is oriented at an angle constant to the pipe’s

longitudinal direction and also tangential to the pipe wall, and can be due to both the pipe-soil

interaction and the internal pressure. In a more conservative estimate, the effects of the welding

process, including the orientation of a weld seam, are accounted for in the form of DFs.

(a) (b) ©

Figure 4 Welded connections in: (a) seamless pipeline, (b) pipeline with longitudinal seam,

and (c) pipeline with helical seam; the hatched area shows the plane of connection of two

pipes

Three major misalignments of a longitudinal weld seam can be observed in SAW-UOE

manufactured pipe, as shown in Figure 5 [24] [52] [25]. It is obvious that the defects associated

Tv L

Th

Page 45: Finite Element Modeling of Buried Longitudinally Welded

28

with the weld region would have a tremendous impact on the stress rise and therefore should be

limited, as discussed in the next sections.

(a) (b) (c)

Figure 5 Common types of weld misalignment in longitudinally welded pipe: (a) radial, (b)

angular, and (c) ovality

3.2.2.1. Radial Misalignment

Radial misalignment is also called axial or linear misalignment, or radial offset that arises

due to axial or plate thickness mismatch at the weld region, 𝛿𝑜 (Figure 5(a)). The radial offset of

the abutting edges of the same nominal WT, measured either at the outer or inner surfaces of the

pipe (high-low radial offset) (Figure 6) must not exceed the values presented in Table 5 (CSA and

API), including 10% of WT or 0.8 mm at the pipe ends, and 10% of WT or 1.5 mm away from the

pipe ends, as prescribed in CSA Z245.1 [4] and API 5L [51]. This value shall be limited to 0.063”

(1.6 mm) according to CSA Z662 [5]. Maximum misalignment of the weld beads, 𝑑, measured

between the centers of the weld beads 𝑀1 and 𝑀2 along the middle line of plates, must not exceed

0.1” (3 mm) for WT ≤ 0.8” (20 mm), or 0.16” (4 mm) for greater WT (Figure 6).

Weld

Dmax

Dmin

θ°

δp

2l

δo

Page 46: Finite Element Modeling of Buried Longitudinally Welded

29

Figure 6 Detailed schematic of radial misalignment

The offset within the allowable tolerance provided in Table 5 (ASME) must be faired at a

3-to-1 taper transition over the width of the finished weld, and a taper of 3×offset for differences

more than ¼ of the thinnest WT or 0.125” (3 mm) (ASME BPVC Section VIII Part 2 [22]).

Table 5 Permissible radial misalignment for different pipe thicknesses (North America)

CSA Z245.1 and API 5L ASME BPVC Section VIII Part 2

𝑡, inch (mm) ≤ 0.590 (15) 0.590-0.984 (15-25) > 0.984 (25) ≤ 0.5 (13) > 0.5 (13)

𝛿𝑜, inch (mm) 0.060 (1.5) 0.1t 0.098 (2.5) 0.2t 0.094 (2.5)

According to BS PD 5500 [52], the middle line of plates and the surfaces of the plates must

be aligned within the tolerances shown in Table 6. The circumferential tolerance for OD ≤ 25”

(650 mm) is ± 0.197” (5 mm), and 0.25% of circumference for greater ODs.

Table 6 Permissible radial misalignment (BS PD 5500)

Middle line of plates Surfaces of plates

𝑡, inch (mm) 𝛿𝑜𝑚𝑖𝑑𝑑𝑙𝑒, inch (mm) 𝑡, inch (mm) 𝛿𝑜ℎ𝑖𝑔ℎ−𝑙𝑜𝑤

, inch (mm)

≤ 0.394 (10) 0.039 (1) ≤ 0.472 (12) 0.25WT

0.394-1.968 (10-50) 0.1WT or 0.118 (3) 0.472-1.968 (12-50) 0.118 (3)

It can be seen that the different standards prescribe similar tolerances for radial

misalignment. However, the maximum difference can be as large as 1.5 mm. In this research, the

𝛿𝑜high

𝛿𝑜low

d

M1

M2

𝛿𝑜middle

Page 47: Finite Element Modeling of Buried Longitudinally Welded

30

value of 1.5 mm for 𝛿𝑜 will be used, corresponding to the portions of pipeline away from the

connection between pipe segments, i.e., pipe ends, according to CSA Z245.1 [4] and API 5L [51].

3.2.2.2. Angular Misalignment

Angular misalignment is also called outward peaking or inward peaking, 𝛿𝑝 (Figure 5(b)).

Peaking shall not exceed 0.125” (3.2 mm) according to API 5L [51], and shall not exceed 3 mm

within 200 mm of each pipe end according to CSA Z245.1 [4]. The maximum permitted values

for peaking (inward or outward), in excess of those shown in Table 7 (BS PD 5500 [52]), are only

permitted when supported by special fatigue analysis but must not exceed 0.197” (5 mm) for

t/D ≤ 0.025, and 0.394” (10 mm) for larger WT/OD ratios (or must not exceed WT).

Table 7 Permissible angular misalignment (BS PD 5500)

𝑡, inch (mm) 𝛿𝑝, inch (mm)

WT < 0.118 (3) 0.059 (1.5)

0.118 (3) ≤ WT < 0.236 (6) 0.094 (2.5)

0.236 (6) ≤ WT < 0.354 (9) 0.118 (3)

0.354 (9) ≤ WT t/3

Although there is insignificant difference between North American standards, European

standards tolerate much larger angular misalignments for vessels of larger WT.

3.2.2.3. Ovality Misalignment

Ovality or ovalization is also called inconsistency of diameter or out-of-roundness

(O-of-R), ∆𝜃 (Figure 5(c)), and shall be limited to the difference between the 𝐷𝑚𝑎𝑥 and 𝐷𝑚𝑖𝑛, not

exceeding 5% [5] or 1% [22]. According to [52], ovality shall not exceed (1

2+

625

𝑂𝐷 𝑚𝑚)% or 1%,

whichever is smaller. The ovality limit due to bending shall be limited to the range ∆𝜃≤ ∆𝜃𝑐𝑟𝑖𝑡=

0.03 ÷ 0.06, where ∆𝜃= 2 (𝐷𝑚𝑎𝑥−𝐷𝑚𝑖𝑛

𝐷𝑚𝑎𝑥+𝐷𝑚𝑖𝑛) is calculated with the OD [5]. Detailed allowance for

ovality misalignment is given in Table 8 [51]. For pipe of 𝐷 𝑡 ≥ 75⁄ the 𝐷𝑚𝑎𝑥 (𝐷𝑚𝑖𝑛) shall be

Page 48: Finite Element Modeling of Buried Longitudinally Welded

31

<1% larger (smaller) than the specified 𝐷, and at 𝐷 𝑡 ≤ 75⁄ the maximum differential between the

𝐷𝑚𝑎𝑥 or 𝐷𝑚𝑖𝑛 shall not exceed 12.7 mm for 𝐷 ≤ 1067 mm and 15.9 mm for 𝐷 ≥ 1067 mm [4].

Ovality can be measured using the same method used for measuring peaking [52].

Table 8 Permissible ovality misalignment (API 5L)

𝐷, inch (mm) 𝐷 tolerance, inch (mm) O-of-R tolerance, inch (mm)

Pipe body Pipe end Pipe body Pipe end

6.625-24 (168-610) ± 0.0075 𝐷

max. ± 0.125 (3.2)

± 0.005 𝐷

max. ± 0.063 (1.6)

0.02 𝐷 0.015 𝐷

24-56 (610-1422)

± 0.005 𝐷

max. ± 0.160 (4.0)

max. ± 0.063 (1.6)

0.015 𝐷

max. 0.6 (15)

for 𝐷 𝑡⁄ ≤ 75

0.01 𝐷

max. 0.5 (13)

for 𝐷 𝑡⁄ ≤ 75

Other pipe geometry defects, such as dents, discussed in [51] and [4], are not considered

in this research.

3.2.3. Welding Defects

Welding produces a variety of defects, including metallurgical and geometrical; since the

former is discussed in the codes from the perspective of residual stresses and is mainly considered

in crack propagation studies, this section will be focused on geometrical defects of welds, and

these will be considered in the research.

3.2.3.1. Weld Reinforcement

As-deposited inside and outside weld bead surfaces shall not extend above the adjacent

original parent metal surface by more than the values presented in Table 9, in accordance with

ASME BPVC Section VIII Part 2 [22], API 5L [51], CSA Z662 [5], CSA Z245.1 [4], BS 5500

[52], and shall not be below the prolongation of the applicable adjacent original parent metal

Page 49: Finite Element Modeling of Buried Longitudinally Welded

32

surface. Weld beads must be ground down to 0.02” (0.5 mm) from each pipe end to the distance

of 4” (100 mm) [51].

Table 9 Permissible weld reinforcement, inch (mm)

ASME BPVC API 5L CSA Z662 CSA Z245.1 BS 5500

𝑡 𝑊ℎ 𝑡 𝑊ℎ 𝑡 𝑊ℎ 𝑊ℎ 𝑊ℎ

Int. Ext. Ext.

< 0.094 0.031 ≤ 0.512 0.138 0.138 < 0.394 0.098 < 0.157 0.0394 + 0.1𝑊𝑤

< (2.4) (0.8) ≤ (13) (3.5) (3.5) < (10) (2.5) < (4) or (1) + 0.1𝑊𝑤

0.094-0.187 0.062 > 0.512 0.138 0.177 > 0.394 0.138 max. 0.197 (5)

(2.4-4.8) (1.5) > (13) (3.5) (4.5) > (10) (3.5) Int.

0.187-1.0 0.094 Penetration

(4.8-25) (2.5) 0.0394 + 0.3𝑊𝑤

1.0-2.0 0.125 or (1) + 0.3𝑊𝑤

(25-51) (3.0) max. 0.118 (3)

The welding process can negatively impact the geometry of the parent material adjacent to

the weld bead, and this is related to the reduction in WT 𝑡. The reduction of 𝑡 due to welding shall

not exceed 1/32” (1 mm) or 0.1𝑡 (ASME BPVC Section VIII Part 2 [22]). In North American

standards, this defect is referred to as undercut, which shall be not deeper than 0.5 mm (CSA

Z245.1 [4]), or 0.06𝑡𝑛𝑜𝑚 and 0.039” (1 mm) (CSA Z662 [5]). Undercut must be ≤ 0.016” (0.4 mm)

deep, or ≤ 0.031” (0.8 mm) deep when its length is ≤ 0.5𝑡, in any given 12” (300 mm) length of

weld (API 5L [51]). In BS 5500 [52], the reduction of 𝑡 is also referred to undercut, and is limited

to ≤ 0.039” (1 mm) or ≤ 0.025𝑡 in the case of a double-sided full-penetration butt weld between

co-planar plates (type 5.2D in BS 7608 [24]); weld root concavity or shrinkage groove must not

exceed 0.059” (1.5 mm). Since the weld root of the Double SAW method is located at the middle

line of the welded plates, the occurrence of an external shrinkage groove is unlikely.

Page 50: Finite Element Modeling of Buried Longitudinally Welded

33

Although the tolerances of weld height 𝑊ℎ can be found in most standards, the weld bead

width calculation, 𝑊𝑤 = 𝑡 2⁄ , as well as the weld root/toe radius definition, 𝑊𝑟 ≥ 0.25𝑡, are found

only in BS 7608 [24]. However, it can be approximately calculated from the geometry of the

abutting plates before welding (Figure 7). For example, the root gap can be set to 0.051-0.075”

(1.3-1.9 mm) for mechanized or automatic welding (CSA Z662 [5]), and the taper angle can be set

to 60° (ASME B31.4 [23]). Therefore, the 𝑊𝑤 for different standard 𝑡 is normally as shown in

Figure 8.

Figure 7 Schematic of abutting plates before welding

Figure 8 Weld widths for different wall thicknesses of plates tapered at 60°

3.2.3.2. Welding Cracks

Planar defects such as cracks, lack of penetration, and incomplete fusion shall be

unacceptable regardless of location [4] [52]. However, surface defects that have a depth ≤ 0.05𝑡

and do not encroach on 𝑡𝑚𝑖𝑛, such as undercuts, can be acceptable [51]. For gas or liquid service,

the maximum imperfection depth may exceed 0.5𝑡 or 0.25𝑊ℎ, provided that an analysis to

determine fatigue crack growth is carried out [5]. The maximum permissible size of any indication

10.00

15.00

20.00

25.00

30.00

35.00

9 14 19 24 29

Wel

d W

idth

, mm

Wall Thickness, mm

60°

Root gap

Page 51: Finite Element Modeling of Buried Longitudinally Welded

34

shall be limited to 𝑡 4⁄ or 4 mm, and indications separated by more than 25 mm may have a size

of 𝑡 3⁄ or 6 mm, whichever is less [22]. In any 300 mm-length of weld, elongated slag inclusions

shall not exceed 1.5 × 50 mm, and isolated slag inclusions shall not exceed 2.5 mm or 𝑡 3⁄ ×10

mm [5] [4]. Spherical porosity, circular slag inclusion, or gas pockets shall not exceed 𝑡 4⁄ or

3 mm, and the projected area shall be limited to 3, 4, and 5% at 𝑡 < 14 mm, 𝑡 = 14-18 mm, and

𝑡 > 18 mm respectively in any 150 mm length of weld [5] [4] [52]. Wormholes of length ≤ 6 mm

× width ≤ 1.5 mm and inclusions of length = 𝑡 ≤ 100 mm × width = 𝑡 10⁄ ≤ 4 mm are prescribed

in [52]. Similar tolerances on inclusions are given in [51].

3.3. Pipe-Soil Interaction

Oil pipelines located in non-developed areas in Canada are buried 2 ft (0.6 m) below the

ground (CSA Z622 [5]). The ASME B31.4 [23] code, which is used to design the liquid lines,

references the procedures for the design of restrained underground piping as prescribed in ASME

B31.1 [54] and in Guidelines for the Design of Buried Steel Pipe (American Lifelines Alliance,

2001) [55]. The procedures referenced provide guidance on modeling pipe-soil interaction. For

example, the Winkler model is widely used due to its simplicity in modeling pipe-soil interaction

with soil springs to represent soil forces in three principal directions [55].

The pipe burial conditions and the soil properties specific to the pipeline installation site

should be used in pipeline design to account for the effects of soil on a buried pipe. In this study,

a soil represented by mainly clay with the properties summarized in Table 10 will be considered.

Table 10 Soil properties

Soil 𝐸

[MPa]

𝜈 𝜌

𝑔 𝑚𝑚3⁄

𝛼

[mm/K]

𝜎𝑦𝐶

[MPa]

𝛼𝑓

[°]

𝛼𝑑

[°]

휀𝑝

10 0.3 0.0015 1×10-4 0.03 0 0 0

Page 52: Finite Element Modeling of Buried Longitudinally Welded

35

A relatively high coefficient of friction, 𝜇 = 0.3, between pipeline and soil can be selected

in case the pipeline is not provided with the protective coating. The soil friction angle of 𝛼𝑓 = 0°

can be selected (saturated soil showing undrained shear strength) to simplify the analysis [56], in

which case the Mohr-Coulomb criterion reduces to Tresca criterion due to only cohesive behavior

of soil (clay) [57]. Since the study focuses mainly on stresses on the pipe, this would be sufficient

to provide the necessary static pressure of a sustained load on pipeline due to soil and avoid

potential issues with convergence in the FE model. A soil box would provide a more convenient

and more realistic way of modeling the pipe-soil interaction [58].

3.4. Pipeline Materials

Steels manufactured according to the API 5L [51] are widely used in pipeline engineering.

In particular, low-carbon ferrite-perlite grades X42 through X80 are commonly used. Higher-grade

steels (e.g., X90, X100, and X120) are also used, benefiting from the higher strength and toughness

of the bainitic microstructure. While there is a significant difference between maximum yield

strength (𝜎𝑦𝑚𝑎𝑥) and maximum tensile strength (𝜎𝑡

𝑚𝑎𝑥) of lower-grade steels, this difference

reduces dramatically with the steel grade, as shown in Figure 9. This figure, constructed based on

data from [3] per API 5L [51], gives rise to the 𝜎𝑦/𝜎𝑡 ratio from 0.65 to 0.91, resulting in reduction

of elongation at fracture from around 25% to 5%. Materials with a 𝜎𝑦/𝜎𝑡 ratio close to unity may

be brittle because of limited plasticity at 𝜎𝑦 ≈ 𝜎𝑡.

Higher-strength materials contain relatively higher volume fractions of secondary particles

such as carbides 𝐹𝑒3𝐶 on which voids will nucleate and as a consequence, relatively lower fracture

resistances can be observed [59]. This implies that smaller grains and/or no secondary particles is

beneficial for toughness and crack resistance [60] [61]. When a crack is initiated, it can propagate,

Page 53: Finite Element Modeling of Buried Longitudinally Welded

36

and both higher-strength and lower-strength steels may be susceptible to sudden changes in the

crack propagation modes. For example, ductile to ductile-brittle mixed-mode types of failure have

been observed on the fracture surfaces [62]. It was shown in [63] that both low-strength and high-

strength steels (Figure 9) can have very similar resistances to fatigue damage.

Figure 9 Mechanical properties of pipeline steels

The increased grades of the X series of steels manufactured as per the API 5L standard [51]

shows an increased strength [64] [65], as shown in Figure 10, reduction of elongation at fracture,

휀, and decreases in both hardening capacity and fracture resistance [59].

Figure 10 Typical engineering stress-strain tensile curves for some X steels as per API 5L

[3]

200

300

400

500

600

700

800

900

1000

1100

1200

40 50 60 70 80 90 100 110 120

Stre

ss, M

Pa

Steel Grade, X

𝜎𝑡𝑚𝑎𝑥

𝜎𝑦𝑚𝑎𝑥

𝜎𝑡𝑚𝑖𝑛

𝜎𝑦𝑚𝑖𝑛

0 0.02 0.04 0.06 0.08 0.10 0.12 0.14 0.16 0.18 0.20

σ,

MP

a

1000

800

600

400

200

0

ε, mm/mm

X70 X80

X100

X52

𝑡𝑎𝑛(𝛼) = 𝐸 α

yield strength

Page 54: Finite Element Modeling of Buried Longitudinally Welded

37

The linear portion of each stress-strain curve in Figure 10 is referred to as the elastic portion

of the tensile curve. It has the same slope, since the Young’s modulus of elasticity, 𝐸, is the

material property and is almost the same for all X steels, 𝐸 ≈ 207 𝐺𝑃𝑎. This is due to their very

similar chemical compositions regardless of differences in the thermal-mechanical treatment.

Therefore, values of 𝐸 are not prescribed in API 5L [51]. The possible variability in actual 𝐸

reported in the available literature is most likely due to variations in the testing conditions such as

temperature. Hence, all X steels have same Hooke’s relationship for the elastic region,

𝜎 = 휀𝐸, (1)

until the yield strength, 𝜎𝑦, which is unique for each steel grade and signifies the transition to the

elastic-plastic region (the curvilinear portion of each tensile curve), as shown in Figure 10.

When the stress is above the yield strength, the material undergoes plastic deformation,

and the elastic-plastic region of the tensile curve can be described using

𝜎 = 𝐾휀𝑛, (2)

where 𝐾 is the strength coefficient and 𝑛 is the strain hardening exponent. Both can be obtained

experimentally by following the standard-practice ASTM E646 [66].

True stress-strain tensile experimental data can be fitted with the widely used Ramberg-

Osgood material model [67],

휀𝑡𝑟𝑢𝑒 =𝜎𝑡𝑟𝑢𝑒

𝐸+ 𝐾 (

𝜎𝑡𝑟𝑢𝑒

𝜎𝑦)

𝑛

(3)

or

휀𝑡𝑟𝑢𝑒 =𝜎𝑡𝑟𝑢𝑒

𝐸+ 𝛼

𝜎𝑡𝑟𝑢𝑒

𝐸(

𝜎𝑡𝑟𝑢𝑒

𝜎𝑦)

𝑛−1

, (4)

where coefficient 𝛼 can be computed as

Page 55: Finite Element Modeling of Buried Longitudinally Welded

38

𝛼 = 𝐾 (𝜎𝑦

𝐸)

𝑛−1

, (5)

and 𝛼𝜎𝑡𝑟𝑢𝑒

𝐸|

𝜎𝑡𝑟𝑢𝑒→𝜎𝑦

= 휀𝑦 is the yield offset (0.2% or 0.002 mm/mm),

true stress,

𝜎𝑡𝑟𝑢𝑒 = (1 + 휀𝑒𝑛𝑔)𝜎𝑒𝑛𝑔, (6)

and true strain,

휀𝑡𝑟𝑢𝑒 = 𝑙𝑛(1 + 휀𝑒𝑛𝑔). (7)

It follows from Eq. (3) that the total strain 휀𝑡𝑟𝑢𝑒 is the sum of the elastic part 𝜎𝑡𝑟𝑢𝑒 𝐸⁄ and

the plastic part 𝐾(𝜎𝑡𝑟𝑢𝑒 𝜎𝑦⁄ )𝑛

, which is present even at lower stresses within the proportionality

region, Eq. (1). For example, this can be seen from the X42 steel tensile data in Figure 11 [68],

cut-off at strain equal to 0.05 for clarity.

Figure 11 Typical engineering and true stress-strain tensile curves for X42 steel

Although the elastic-plastic region seems to have an almost exact fit, the curvature of the

yielding region is clearly omitted from the Ramberg-Osgood fit. A significant deviation from the

linear portion of the elastic region of both the engineering and the true stress-strain curves can be

0 0.01 0.02 0.03 0.04 0.05

σ,

MP

a

500

400

300

200

100

0

ε, mm/mm

true stress-strain engineering stress-strain Ramberg-Osgood fit

Page 56: Finite Element Modeling of Buried Longitudinally Welded

39

seen from just above 150 MPa to around 350 MPa, at the intersection of all three curves. This

inaccuracy may be insignificant in the case of softer materials with high strain-hardening capacity

and smooth elastic-plastic transition. However, the inaccuracy needs to be addressed in the case of

higher-strength materials such as X42 and higher-grade steels that show more pronounced yielding

plateaus (see Figure 12) in order to avoid possible errors associated with further use of these data.

Figure 12 Difference between experimental stress-strain tensile curve for X42 steel and

Ramberg-Osgood fit near the yield strength

One of the solutions would be connecting the yield strength point with a line tangent to the

Ramberg-Osgood fit, which can subsequently be utilized in other calculations such as FEM.

The results of uniaxial tensile test of material can be used to predict yielding under multi-

axial loading conditions by calculating an equivalent tensile stress developed by Ludwig Heinrich

Edler von Mises, and, thus, also known as equivalent von Mises stress, 𝜎𝑣, or von Mises yield

criterion,

𝜎𝑣 = √[(𝜎11 − 𝜎22)2 + (𝜎22 − 𝜎33)2 + (𝜎33 − 𝜎11)2 + 6(𝜏122 + 𝜏23

2 + 𝜏312 )] 2⁄ , (8)

which is a circle of a cylinder with axis [1 1 1] and radius √2 3⁄ 𝜎𝑦 in case of hydrostatic stress,

(𝜎11 − 𝜎22)2 + (𝜎22 − 𝜎33)2 + (𝜎33 − 𝜎11)2 = 𝑐𝑜𝑛𝑠𝑡 at 𝜎11 = 𝜎22 = 𝜎33, Figure 13(a), or has

true stress-strain engineering stress-strain Ramberg-Osgood fit

σ,

MP

a

400

300

200

100

0

ε, mm/mm

0 0.0025 0.005 0.0075 0.01

𝜎𝑦𝑡𝑟𝑢𝑒

𝜎𝑦𝑓𝑖𝑡

휀𝑒𝑡𝑟𝑢𝑒

휀𝑒𝑓𝑖𝑡

휀𝑝

∆휀

∆𝜎

Page 57: Finite Element Modeling of Buried Longitudinally Welded

40

elliptical shape in plane stress, 𝑐𝑜𝑛𝑠𝑡 = 2𝜎112 |𝜎11→𝜎𝑦

at 𝜎3 = 0, because 𝜎1 − 𝜎2 plane (1 1 0),

would cut the yield surface of a cylinder at 45° angle, resulting in the ellipse projection on the

𝜎1 − 𝜎2 plane, Figure 13(b). The von Mises yield criterion curve would intersect the principal

stress axes, normalized by 𝜎𝑦, at unity when stress reaches yield stress of material.

Figure 13 Tresca and von Mises yield criteria in (a) hydrostatic and (b) plane stresses

The plane stress condition is practical for ductile materials as they are compressible, show

plasticity and shearing at stresses higher than the yield limit, 𝜎𝑦. Although von Mises criterion

provides better agreement with experimental data, the Tresca criterion is still used because of its

simplicity,

𝑚𝑎𝑥 (|𝜎11 − 𝜎22|, |𝜎22 − 𝜎33|, |𝜎33 − 𝜎11|) 2⁄ = 𝜏𝑦 = 𝜎𝑦 2⁄ . (9)

The Tresca yield criterion for plane stress condition is derived from the yield surface of a

hexagonal prism, Figure 13, similarly to von Mises criterion.

The equivalent stress should be limited to some design value, which is usually related to

material’s yield limit, 𝜎𝑦, and is calculated based on actual loading of a component. In

longitudinally welded straight portion of a pipeline, the weld seam would normally experience

𝜎1 𝜎2

𝜎3

𝜎1 = 𝜎2 = 𝜎3

von Mises Tresca

ඨ2

3𝜎𝑦

(a)

-1

𝜎1

𝜎𝑦

𝜎2

𝜎𝑦

-1

1 1

0

Tresca

von Mises

(b)

ඨ2

3𝜎𝑦

√2𝜎𝑦

Page 58: Finite Element Modeling of Buried Longitudinally Welded

41

stresses as shown in Figure 14. Both the normal (out-of-plane) and the shear (in-plane) components

of the stress tensor are depicted in Figure 14. The stress tensor shows geometrical relations between

the stress components in a continuous material and directions of acting stresses, wherein 1 = 𝑋,

2 = 𝑌, and 3 = 𝑍.

Figure 14 Stress tensor at the longitudinal weld in pipeline

The stress in hoop direction, 𝜎𝑦𝑦, is usually the largest in longitudinally welded pipes

because it acts normal to the structural discontinuity (weld seam, the red line in Figure 14) and

tend to open the hypothetical crack; longitudinal stress (𝜎𝑥𝑥 in Figure 14) is usually the second

largest stress, and the stress in radial direction (𝜎𝑧𝑧 in Figure 14) is the smallest. For the pipeline

to provide the safe pressure containment, all stresses should be determined, and the equivalent

stress (or stress in most critical direction/plane) should be calculated and limited to the design

value based on mechanical properties of pipeline material used (API 5L [51] and ASME BPVC

Section VIII [22]). The pipeline loading should be properly accounted in the calculation of stresses.

𝜎𝑧𝑧

𝜏𝑧𝑦 𝜏𝑧𝑥

𝜎𝑦𝑦

𝜏𝑦𝑧

𝜏𝑦𝑥

𝜎𝑥𝑥

𝜏𝑥𝑧

𝜏𝑥𝑦

X Y

Z

σLong.= σxx

σRadial= σzz

σHoop= σyy

Z

Y X

σHoop

Page 59: Finite Element Modeling of Buried Longitudinally Welded

42

3.5. Loading

A pipeline can be exposed to extreme loading conditions, including aggressive

environments, high temperatures, multiaxial loading states, etc. A pipeline may experience internal

pressure fluctuations and temperature gradients associated with oil and gas transportation and

seasonal weather changes, and location-specific conditions may include winds acting on above-

ground pipes, currents that strain sub-sea pipes, soil pressure in the case of underground pipes,

landslides, dents, etc. All applicable loads and their combinations shall be considered, and a

loading histogram shall be developed. The actual random amplitude loading spectra (Figure 15(a))

can be very complex and are usually analyzed and converted to blocks of constant amplitude

loading (Figure 15(b)) by a standard algorithm, such as rain-flow cycle counting, in order to

simplify further use in fatigue life assessment [21] [22] [24]. The in-service conditions of the

component must be appropriately considered to correctly identify the failure criteria at the design

stage for timely repair/replacement during service.

(a) (b) Figure 15 Fatigue loading showing (a) spectrum loading and (b) constant amplitude

loading

The pressure history of the liquid pipeline, as shown in Figure 16, was obtained from the

literature [69], and was cycle-counted as per ASME BPVC Section VIII Part 2 [22] using the

rainflow counting algorithm from [70], and is shown in Figure 17. This is an example of a typical

time

load

reversal – valley

reversal – peak

mean

range

block 1 block 2 block 3

time

load

max

mean

range

am

pli

tud

e

min

Page 60: Finite Element Modeling of Buried Longitudinally Welded

43

loading observed in liquid lines between pumping stations during transportation of oil [69]. Similar

loading spectrum can be found in [71]. Notably, the large variations in pressure are typical. The

pipeline sections located away from pump stations normally experience smaller ranges of pressure

per cycle due to pressure drop, and, therefore, would be considered non-fatigue-prone areas.

However, these sections are usually designed with lower WT and may result in the stress cycles

comparable to those observed in pipeline sections immediately downstream the pump station [71].

Therefore, the actual pressure readings at the location of interest should be used in fatigue design.

Figure 16 In-service pressure history diagram

The selected pressure history contains cycles that exceed the design limit pressure, 𝑃𝑖, of

10 MPa, which can represent a potential worst case loading scenario, as pipelines that transport

liquids, which are generally incompressible, tend to experience transient pressure fluctuations

during service for different reasons [72] [73] [74] [75] [76]. Unfortunately, some of these cycles

can lead to disastrous consequences due to pipeline failure. The overpressure cycles presented in

Figure 17 are expected to accelerate fatigue damage due to pulsating bending of a misaligned weld,

especially for loads with stress ratio 𝑅 ≥ 0, as analyzed in [18].

0

2

4

6

8

10

0 50 100 150 200 250

Pre

ssu

reP

i, M

Pa

Days

Page 61: Finite Element Modeling of Buried Longitudinally Welded

44

Figure 17 Cycle-counted in-service pressure history showing Pmin, Pmax, and nk

The cycle-counting process of the data obtained from some real spectra, such as the one

presented in Figure 16, can be controlled as described in APPENDIX B – MATLAB Cycle-

Counting. A complete input file, *.inp, with all programmed features and properties of the elastic-

plastic model is given in APPENDIX C – ABAQUS Input File. The elastic model can be obtained

by deleting the elastic-plastic properties from the elastic-plastic model.

3.6. Linearization of Stresses

Through-thickness stress linearization of the actual non-linear stress distribution is

prescribed by the ASME standard to obtain different components of stress at the hot-spot,

including the membrane stress 𝜎𝑚 and bending stress 𝜎𝑏. After the analysis in ABAQUS is

finished, the calculated stresses for each loading cycle are linearized at the hot-spot along the Stress

Classification Line (SCL). The approximate path of an SCL is used for stress linearization as

required by ASME BPVC Section VIII Part 2 [22].

To calculate the membrane and bending stresses, ABAQUS [77] utilizes the integration of

stresses through the section as:

𝜎𝑚 =1

𝑡∫ 𝜎 ∙ 𝑑𝑥

𝑡

2

−𝑡

2

and 𝜎𝑏 =6

𝑡2 ∫ 𝜎 ∙ 𝑥 ∙ 𝑑𝑥

𝑡

2

−𝑡

2

, (10)

ASME BPVC Section VIII Part 2 [22] utilizes the integration of stresses through the section using

0

20

40

60

0

2

4

6

8

10

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 Nu

mb

er o

f C

ycle

s n

k

Inte

rnal

Pre

ssu

re

Pm

in, P

ma

x,M

Pa

Cycle

Pmin

Pmax

Cycles

Page 62: Finite Element Modeling of Buried Longitudinally Welded

45

𝜎𝑚 =1

𝑡∫ 𝜎 ∙ 𝑑𝑥

𝑡

0 and 𝜎𝑏 =

6

𝑡2 ∫ 𝜎 ∙ (𝑡

2− 𝑥) ∙ 𝑑𝑥

𝑡

0, (11)

where 𝑡 is the WT and 𝜎 is the stress at a coordinate 𝑥 along the SCL path. It should be noted that

both integrations yield the same results.

Assuming that the Level 1 fatigue analysis screening criteria are satisfied (ASME BPVC

Section VIII [22]), and the stresses and strains are obtained using FEA, the fatigue analysis can be

performed by using one or several of the available fatigue assessment methods. These are based

on an elastic stress analysis and equivalent stresses, or based on an elastic-plastic stress analysis

and equivalent strains for smooth bar fatigue curves. They can also be based on the fatigue curves

for welded joints when performing fatigue assessment of welds based on elastic analysis and

structural stress. Each of the fatigue assessment methods available in ASME BPVC Section VIII

[22] follows step-by-step procedures discussed in the next section.

3.7. Analytical Model

The analytical model primarily focuses on the development of stress at the structural hot-

spot of a buried pipe in the hoop direction due to the weld, its misalignment, and soil.

3.7.1. Stress due to Misalignment

Critical stress for a longitudinal weld is the one perpendicular to the longitudinal direction

of a weld seam (⊥ to a defect or discontinuity) acting in the axial direction of the welded steel

plates, i.e., the hoop stress 𝑆ℎ,

𝑆ℎ =𝑃∙𝐷

2∙𝑡, (12)

calculated based on 𝑃 (the internal pressure of a pipe), its 𝐷 (the outside diameter), and 𝑡 (the wall

thickness). For example, 𝑆ℎ = 261 𝑀𝑃𝑎 when 𝑃 = 10 𝑀𝑃𝑎, 𝐷 = 914 𝑚𝑚, and 𝑡 = 17.5 𝑚𝑚.

This stress is a nominal membrane stress that occurs in the pipe wall away from the weld. To

Page 63: Finite Element Modeling of Buried Longitudinally Welded

46

obtain the design stress, the nominal stress needs to be multiplied by a design stress magnification

factor, 𝑘𝑚, which should include the effects from all possible stress magnifiers such as the effects

of weld geometry and misalignments, and stress magnification due to soil.

Special attention was paid to the modeling of a weld profile and misalignment to capture

their effects on the hot-spot’s stresses/strains.

The critical spot in the weld is the weld toe, which for a DSAW seam is located on both

the outer and inner surfaces of a pipe at the transition between the plate and weld

crown/reinforcement. The weld toe is a stress concentrator which magnifies the 𝑆ℎ by a factor 𝑘𝑚

(𝑘𝑡) – stress magnification (or stress concentration) factor (SCF), BS 7608 [24],

𝑆ℎ.𝑑𝑒𝑠𝑖𝑔𝑛 = 𝑆ℎ ∙ (1 + ∑ 𝑘𝑚.𝑖𝑛𝑖=1 ). (13)

The contributions to the stress rise due to various types of weld misalignments, discussed

in 3.2.2, will be further addressed. The 𝑛 types of different contributors to the total SCF need to

be considered in design, where each magnification factor comes with a sign depending on the

location of the critical spot considered and represents the contribution to the total stress from the

secondary bending due to a particular effect BS 7910 [25].

Therefore, having abutting plates with equal thicknesses, the SCFs can be calculated for

the axial misalignment (BS 7910 [25]), 𝑘𝑚.𝑎𝑥𝑖𝑎𝑙, as

𝑘𝑚.𝑎𝑥𝑖𝑎𝑙 =𝜎𝑏.𝑎𝑥𝑖𝑎𝑙

𝑃𝑚= 3

𝛿𝑜

𝑡=

6𝛿𝑜

𝑡1(1−𝜈2)(

1

1+(𝑡2 𝑡1⁄ )0.6), (14)

for the angular misalignment (BS 7910 [25]), 𝑘𝑚.𝑎𝑛𝑔𝑢𝑙𝑎𝑟, as

𝑘𝑚.𝑎𝑛𝑔𝑢𝑙𝑎𝑟 =𝜎𝑏.𝑎𝑛𝑔𝑢𝑙𝑎𝑟

𝑃𝑚=

3𝛿𝑝

𝑡(1−𝜈2){

𝑡𝑎𝑛ℎ(𝛽 2⁄ )

𝛽 2⁄}, (15)

and for the ovality misalignment (BS 7910 [25]), 𝑘𝑚.𝑜𝑣𝑎𝑙𝑖𝑡𝑦, as

Page 64: Finite Element Modeling of Buried Longitudinally Welded

47

𝑘𝑚.𝑜𝑣𝑎𝑙𝑖𝑡𝑦 =𝜎𝑏.𝑜𝑣𝑎𝑙𝑖𝑡𝑦

𝑃𝑚=

1.5(𝐷𝑚𝑎𝑥−𝐷𝑚𝑖𝑛)𝑐𝑜𝑠2𝜃

𝑡{1+𝑝(1−𝜈2)

2𝐸(

𝐷𝑚𝑡

)3

}, (16)

where parameter 𝛽 is calculated from

𝛽 =2𝑙

𝑡{

3(1−𝜈2)𝜎𝑚𝑚𝑎𝑥

𝐸}

0.5

, (17)

𝛿𝑜 – axial mismatch, 𝑡 – plate (pipe wall) thickness, 𝜎𝑏 – bending stress at the weld toe due

to misalignment, 𝑃𝑚 – applied membrane stress, 𝛿𝑝 – peaking, ν – Poisson’s ratio, 𝑃 – max.

operating pressure, mean diameter 𝐷𝑚 = 𝐷 − 𝑡, E – modulus of elasticity, 𝜎𝑚𝑚𝑎𝑥 – max. membrane

stress (𝑆ℎ), 𝑙 – half of a distance between the points when the peaking departures from the perfect

circle (found geometrically using SolidWorks between tangent points after 𝛿𝑜 and 𝛿𝑝 applied), 𝜃

– angle at which the weld is located w.r.t. the ovalized section (0° and 90° are considered).

3.7.2. Stress due to Soil

It is important to account for the stresses coming from the bending of a pipe wall due to

soil pressure (external pressure). Following the previous methodology, we introduce 𝑘𝑚.𝑠𝑜𝑖𝑙

(Guidelines for the Design of Buried Steel Pipe [55]),

𝑘𝑚.𝑠𝑜𝑖𝑙 =𝜎𝑏.𝑠𝑜𝑖𝑙

𝑃𝑚, (18)

where 𝜎𝑏.𝑠𝑜𝑖𝑙 is a through-wall bending due to soil,

𝜎𝑏.𝑠𝑜𝑖𝑙 = 4𝐸 (∆𝑦

𝐷) (

𝑡

𝐷), (19)

(∆𝑦

𝐷) is a deflection (or ovality) due to soil,

(∆𝑦

𝐷) =

𝐷𝑙𝐾𝑏𝑃𝑠(𝐸𝐼)𝑒𝑞

(𝐷𝑚

2)

3+0.061𝐸′,

(20)

and 𝑃𝑠 is the soil pressure on the pipe above the water table,

Page 65: Finite Element Modeling of Buried Longitudinally Welded

48

𝑃𝑠 = 𝜌𝐻, (21)

(𝐸𝐼)𝑒𝑞 – equivalent pipe wall stiffness per unit length of pipe (may include pipe lining and pipe

coating), ∆𝑦 – vertical deflection of pipe, 𝐷𝑚 – pipe mean diameter, 𝐷𝑙 – deflection lag factor

(≈ 1.0-1.5), 𝐾𝑏 – bedding constant (≈ 0.1), 𝐼 is moment of inertia of wall cross section per unit

length of a plane pipe,

𝐼 =𝑡3

12, (22)

𝐸′ – modulus of soil reaction (≈ 0-20 MPa, for loose to compact soil), 𝜌 – unit weight of soil fill

((0 ÷ 125) 𝑙𝑏 𝑓𝑡3⁄ , 120 𝑙𝑏 𝑓𝑡3⁄ = 1.96 × 10−5 𝑁 𝑚𝑚3⁄ , an 𝐻 – pipe burial depth [55] [78], as

shown in Figure 18.

Figure 18 Through-wall bending stress and ovality of pipe cross-section due to transmitted

pressure

H

∆y

P

σb

WT

Page 66: Finite Element Modeling of Buried Longitudinally Welded

49

3.8. Fatigue Assessment

The fatigue life assessment of such industrial components as pipelines is critical, as it may

involve substantial financial losses or even human lives in case of a catastrophic failure, as well as

pollution and other environmental impacts [72] [73] [74] [75] [76]. This phenomenon requires

special attention and accompanies developments in different fields of engineering. Different

fatigue assessment methods have been developed and used in engineering practice (Figure 19).

The use of each of the methods summarized in Figure 19 results in the estimation of number of

cycles to failure, 𝑁𝑓, at certain magnitudes of load, stress amplitudes, 𝑆𝑎, or strain amplitudes, 휀𝑎.

With the increasing complexity of components, it is often not sufficient to use the

generalized assessment-specific criteria, which need to be specified more accurately by the user to

reflect the properties of a real component as closely as possible (Figure 20).

Figure 19 Schematic classification of fatigue life approaches

Standardized specimens (Figure 20(a)) and procedures are used to simulate the behavior of

a full-scale component/product (Figure 20(b)). Preferably, the component itself is tested under

service conditions; however, obtaining the experimental data may be very expensive.

Nf (log scale)

Sa (

log

sca

le)

Crack

Initiation Period

Strain Life Approach

(ε-N)

Crack

Growth Period

Fracture

Mechanics (a-N)

Total Life Stress Life Approach

(S-N)

Se

Page 67: Finite Element Modeling of Buried Longitudinally Welded

50

(a) (b) Figure 20 Standardized tensile specimen (a) versus real component (b)

This section focuses on design aspects related to fatigue and discusses various S–N

approaches used to determine the fatigue damage/life of a component. These include fatigue

analysis based on elastic stress analysis and equivalent stress ranges (in this work, this is referenced

as Elastic Fatigue Analysis, BS 7608 [24]), fatigue analysis based on elastic stress analysis and

equivalent stresses (in this work, this is referenced as Modified Elastic Fatigue Analysis, ASME

BPVC Section VIII Part 2 [22], Level 2, Method A), and fatigue analysis based on elastic-plastic

stress analysis and equivalent strain (in this work, this is referenced as Elastic-Plastic Fatigue

Analysis, ASME BPVC Section VIII Part 2 [22], Level 2, Method B).

3.8.1. Stress-Life Curves

Methods of fatigue assessment based on stress-life (S-N) fatigue curves (BS 7608 [24],

ASME BPVC VIII [22] or API 579-1/ASME FFS-1 Fitness-For-Service [21]) are also known as

Level 2 methods. The first systematic investigation of S-N curves was carried out in the late 19th

century by the German railway engineer August Wӧhler, whose work was focused on the causes

of fractures in railroad axles. The Wӧhler curves show the number of cycles to failure, 𝑁𝑓, at a

certain stress amplitude, 𝑆𝑎. One of interesting properties of these curves is the stress level called

fatigue strength, also known as fatigue limit or endurance limit, 𝑆𝑒, below which the life of a

component under cyclic loading, is theoretically infinite and where fatigue failure should not

occur.

Page 68: Finite Element Modeling of Buried Longitudinally Welded

51

The S-N assessment deals with non-planar flaws, which would apply to the weld profile

and associated geometrical defects. The basic S-N curve is shown in Figure 21 (BS 7608 [24]),

and can be obtained through the constant-amplitude force-controlled fatigue test (ASTM E466

[79]), and it shows the effects of variable amplitude (spectrum loading) and the aggressiveness of

the environment on an SAW type of weld (type 5.2D in BS 7608 [24]).

The basic S-N curves relate predominantly to membrane stress, 𝜎𝑚; however, the fatigue

strength of real structures depends on the degree of through-thickness bending, and thus bending

stress, 𝜎𝑏, BS 7608 [24]. Therefore, the basic S-N curve needs to be modified with all possible

deviations from the idealized scenario to obtain a design S-N curve that will reflect the behavior

of a real component, including in-service loading, surface finishing, notches, welding,

misalignments, environmental effects, etc.

Figure 21 Stress life (S-N) curve

The mentioned effects can be included in the basic S-N curve through modification of the

membrane, 𝜎𝑚, or the nominal, 𝜎𝑛𝑜𝑚, stress in the form of stress amplification factors (SAF),

𝜎𝑡𝑜𝑡𝑎𝑙 = 𝜎𝑛𝑜𝑚(1 + ∑ 𝑘𝑖𝑛𝑖 ). (23)

The next step in S-N fatigue life assessment is calculating the number of cycles to failure

(NCF), 𝑁𝑓. Both European and North American standards use equations derived from

static

limitations

s

Nf (log scale)

Sa (

log

sca

le),

MP

a

Sa.c=53

Sa.v=31

1 m

1 m+2

constant amplitude

5 Nav=

10

7

Nac=

10

7

Sa=2σy

Page 69: Finite Element Modeling of Buried Longitudinally Welded

52

experimentally obtained best-fitted data. In BS 7608 [24] this is done based on parameters specific

for different weld connections, and the NCF is computed for a range of cyclic stress, which does

not account for the effects of mean stress. In ASME BPVC [22], the effect of mean stress is

accounted for in the CNF equation, whose parameters are based on a specific group of materials

used to obtain the fatigue curve.

Finally, in the case of constant amplitude loading, blocks of 𝑛 number of cycles can be

used to compute the fatigue damage (FD) of a component using Miner’s rule,

𝐷𝑓 =𝑛

𝑁𝑓, (24)

which can include the FDs caused by each loading block, 𝐷𝑓 = ∑𝑛

𝑁𝑓.

3.8.2. Elastic Fatigue Analysis

The elastic fatigue analysis was adopted from BS 7608 [24] and BS 7910 [25], which are

well established and widely used European standard practices for fatigue analysis of welded

structures. This analysis can be used as a reference for comparison to similar yet more detailed

analyses based on North American standards, ASME BPVC Section VIII Part 2 [22] and API 579-

1 [21]. The type of weld corresponding to a longitudinal SAW seam weld of a large-diameter pipe

is a double-sided full-penetration butt weld between co-planar plates. This corresponds to weld

design class 5.2D according to BS 7608 [24]. The number of cycles to failure, 𝑁𝑘, due to cyclic

loading can be calculated by using

𝑁 = 10𝑙𝑜𝑔(𝐶0)−𝑑∙𝑆𝐷−𝑚∙𝑙𝑜𝑔(𝑆𝑟), (25)

where for a selected weld design class, 𝐶0 = 3.988 × 1012 is a constant, 𝑆𝐷 = 0.2095 is a

standard deviation, 𝑑 = 3 is a factor for a nominal probability of failure of 0.14% equal to three

standard deviations of 𝑙𝑜𝑔(𝑁𝑘) below the mean, 𝑚 = 3 is a slope of the 𝑆 − 𝑁 curve, and 𝑆𝑟 is

Page 70: Finite Element Modeling of Buried Longitudinally Welded

53

the applied stress range, i.e., algebraic difference between the two extremes (reversals) of a given

cycle and is equal to ∆𝜎𝑚 + ∆𝜎𝑏 + ∆𝜎𝑝 at the hot-spot, and is obtained through stress linearization.

Although the BS standard states that the use of a stress component, 𝜎𝑖𝑗, acting normal to the weld

toe is sufficient, the use of all stress components is preferred.

The BS code advises that the maximum accumulated FD be limited to 0.5 for critical cases.

Therefore, the calculated values of FD can be normalized to 1.0 to align it with the maximum

allowable FD, as per the ASTM standard.

The elastic fatigue analysis in ASME BPVC [22] and API 579-1/ASME FFS-1 [21] utilizes

similar methodology in Level 2 – Method A, or plasticity corrected Method B.

3.8.3. Modified Elastic Fatigue Analysis

The elastic fatigue analysis can be performed in accordance with API 579-1 [21] and

ASME BPVC Section VIII Part 2 [22], Level 2, Method A. Normally, the effective strain (or the

von Mises equivalent total strain [80]) range, ∆휀𝑒𝑓𝑓,𝑘, one of the main components of the driving

force for fatigue damage, is approximated from an elastic analysis and incorporated in calculation

of fatigue penalty factor (plasticity correction factor), 𝐾𝑒,𝑘. However, a considerable variability in

analysis results can be observed since the stress results are sensitive to the mesh density of the

FEM [37] and the 𝐾𝑒,𝑘 factor can yield inaccurate results for larger stress ranges [81]. Therefore,

the optional modified steps related to the calculation of 𝐾𝑒,𝑘 provided in the ASME standard can

be implemented in order to improve the computational results. Thus, the fatigue analysis in this

section is based on the elastic effective equivalent stress range, ∆𝑆𝑃,𝑘, and the plastic equivalent

strain range, ∆휀𝑝𝑒𝑞,𝑘.

Page 71: Finite Element Modeling of Buried Longitudinally Welded

54

The driving force for a fatigue damage – effective alternating equivalent stress (stress

amplitude) for the 𝑘𝑡ℎ cycle, 𝑆𝑎𝑙𝑡,𝑘, which is needed to calculate the number of cycles to failure,

𝑁𝑘, and the total FD, 𝐷𝑓, caused by a fatigue cycle, can be computed using

𝑆𝑎𝑙𝑡,𝑘 =𝐾𝑓∙𝐾𝑒,𝑘∙∆𝑆𝑃,𝑘

2, (26)

where ∆𝑆𝑃,𝑘 is the effective equivalent stress range,

∆𝑆𝑃,𝑘 = √(∆𝜎11,𝑘−∆𝜎22,𝑘)2

+(∆𝜎11,𝑘−∆𝜎33,𝑘)2

+(∆𝜎22,𝑘−∆𝜎33,𝑘)2

+6(∆𝜎12,𝑘2 +∆𝜎13,𝑘

2 +∆𝜎23,𝑘2 )

2

2, (27)

which is obtained using the range of elastic component stresses, ∆𝜎𝑖𝑗,𝑘 = 𝜎𝑖𝑗,𝑘𝑒𝑛𝑑 − 𝜎𝑖𝑗,𝑘

𝑠𝑡𝑎𝑟𝑡, between

the stress tensors of the start 𝜎𝑖𝑗,𝑘𝑠𝑡𝑎𝑟𝑡 and end 𝜎𝑖𝑗,𝑘

𝑒𝑛𝑑 points of a cycle at a hot-spot, including ∆𝜎𝑚,

∆𝜎𝑏, and ∆𝜎𝑝 components, obtained after stress linearization; 𝐾𝑓 is a fatigue strength reduction

factor that accounts for the local weld notch and can be kept equal to unity in the case the FE model

already comprises the weld discontinuity; 𝐾𝑒,𝑘 is a fatigue penalty factor. 𝐾𝑒,𝑘 can be calculated

from the relationship between the plastic and elastic equivalent total strain ranges as

𝐾𝑒,𝑘 =∆ 𝑒𝑓𝑓,𝑘

∆ 𝑒𝑙,𝑘, (28)

where the effective strain range, ∆휀𝑒𝑓𝑓,𝑘, a von Mises-based measure of the strain state [81] for

which a fatigue crack is open during a cycle [82], is

∆휀𝑒𝑓𝑓,𝑘 = ∆휀𝑒𝑙,𝑘 + ∆휀𝑝𝑒𝑞,𝑘 (29)

and the elastic strain range, ∆휀𝑒𝑙,𝑘, is

∆휀𝑒𝑙,𝑘 =∆𝑆𝑃,𝑘

𝐸𝑇,𝑘, (30)

∆휀𝑝𝑒𝑞,𝑘 is a plastic equivalent strain range obtained from ABAQUS for each pressure increment,

Page 72: Finite Element Modeling of Buried Longitudinally Welded

55

∆휀𝑝𝑒𝑞,𝑘 =√2[(∆𝑝11,𝑘−∆𝑝22,𝑘)

2+(∆𝑝11,𝑘−∆𝑝33,𝑘)

2+(∆𝑝22,𝑘−∆𝑝33,𝑘)

2+6(∆𝑝12,𝑘

2 +∆𝑝13,𝑘2 +∆𝑝23,𝑘

2 )2

]

3, (31)

where ∆𝑝𝑖𝑗,𝑘 is the change in plastic strain range components at the point under evaluation for the

𝑘𝑡ℎ loading condition or cycle, as per ASME BPVC Section VIII Part 2 [22].

The value of 𝐾𝑒,𝑘 increases with larger plastic strains and becomes larger than unity, and

thus increasing the value of 𝑆𝑎𝑙𝑡,𝑘, used in the calculation of the number of cycles to failure,

𝑁𝑘 = 10𝐶1+𝐶3𝑌+𝐶5𝑌2+𝐶7𝑌3𝐶9𝑌4𝐶11𝑌5

1+𝐶2𝑌+𝐶4𝑌2+𝐶6𝑌3𝐶8𝑌4𝐶10𝑌5 , (32)

based on a smooth bar fatigue curve, where 𝑌 is a coefficient,

𝑌 = (𝑆𝑎𝑙𝑡,𝑘

𝐶𝑢𝑠) (

𝐸𝐹𝐶

𝐸𝑇), (33)

𝐶1 through 𝐶11 are the constants that depend on 𝑆𝑎𝑙𝑡,𝑘 (Table 11), 𝐶𝑢𝑠= 6.894757 is the conversion

factor for stresses in MPa (ASME BPVC Section VIII Part 2 [22]), 𝐸𝐹𝐶=195 GPa for the material

used to obtain the constants 𝐶1 through 𝐶11 (ASME BPVC Section VIII Part 2 [22]), and 𝐸𝑇 is the

Young’s modulus at service temperature.

Table 11 Constants for a polynomial fit of experimental data in the calculation of number

of cycles to failure

Constant 𝑆𝑎𝑙𝑡,𝑘 ≤ 45 𝑀𝑃𝑎 45 𝑀𝑃𝑎 < 𝑆𝑎𝑙𝑡,𝑘 ≤ 214 𝑀𝑃𝑎 214 𝑀𝑃𝑎 < 𝑆𝑎𝑙𝑡,𝑘 ≤ 3999 𝑀𝑃𝑎

𝐶1 0 +2.25451 +7.999502 𝐶2 0 −4.642236 × 10−1 +5.832491 × 10−2 𝐶3 0 −8.312745 × 10−1 +1.500851 × 10−1 𝐶4 0 +8.634660 × 10−2 +1.273659 × 10−4 𝐶5 0 +2.020834 × 10−1 −5.263661 × 10−5 𝐶6 0 −6.940535 × 10−3 0 𝐶7 0 −2.079726 × 10−2 0 𝐶8 0 +2.010235 × 10−4 0 𝐶9 0 +7.137717 × 10−4 0 𝐶10 0 0 0 𝐶11 0 0 0

Page 73: Finite Element Modeling of Buried Longitudinally Welded

56

The procedure is repeated for each cycle in the loading history. Miner’s rule is used to

compute the accumulated FD,

𝐷𝑓 = ∑𝑛𝑘

𝑁𝑘≤ 1𝑀

𝑘=1 . (34)

It should be noted that Miner’s rule was derived for elastic analysis and does not account

for local sequencing effects associated with plasticity.

3.8.4. Elastic-Plastic Fatigue Analysis

The elastic-plastic fatigue analysis can also be performed in accordance with API 579-1

[21] and ASME BPVC Section VIII Part 2 [22], Level 2, Method B (the strain-based version of

Method A [37]).

The calculation of FD is based on an approach similar to that described in the Modified

Elastic Fatigue Analysis. However, the range of elastic-plastic component stresses, ∆𝜎𝑖𝑗,𝑘 =

𝜎𝑖𝑗,𝑘𝑒𝑛𝑑 − 𝜎𝑖𝑗,𝑘

𝑠𝑡𝑎𝑟𝑡, shall be used in the calculation of ∆𝑆𝑃,𝑘, Eq. (27). Subsequently, ∆𝑆𝑃,𝑘 from Eq.

(27) and ∆휀𝑝𝑒𝑞,𝑘 from Eq. (31) are substituted in Eq. (29) for the calculation of the effective strain

range, ∆휀𝑒𝑓𝑓,𝑘. Thus, the fatigue analysis in this section is based on the elastic-plastic effective

equivalent stress range, ∆𝑆𝑃,𝑘, and the plastic equivalent stain range, ∆휀𝑝𝑒𝑞,𝑘.

Next, the effective alternating equivalent stress, 𝑆𝑎𝑙𝑡,𝑘, is computed according to

𝑆𝑎𝑙𝑡,𝑘 =𝐸𝑇,𝑘∙∆ 𝑒𝑓𝑓,𝑘

2, (35)

and is used to calculate the number of cycles to failure, as shown in Eq. (32) and Eq. (33).

Miner’s rule, Eq. (34), is used to compute the accumulated FD.

Page 74: Finite Element Modeling of Buried Longitudinally Welded

57

3.8.5. Elastic Fatigue Analysis of Welds

Fatigue assessment of welds can also be performed according to ASME BPVC Section

VIII Part 2 [22] and API 579-1 [21] (Level 2 – Method C) and can be used as a reference to the

data obtained by the previously discussed methods. This analysis utilizes the linear elastic stress

analysis of structural stresses and Neuber’s method for local plasticity correction. This method is

recommended for seam weld joints that have not been machined and is based on the structural

stress (comprised of membrane, 𝜎𝑚, and bending, 𝜎𝑏, stresses) normal to the hypothetical crack

plane, i.e., the components of stress in the hoop direction of a pipe.

First, the stress ranges due to pressure cycles can be calculated for structural membrane

stress range, ∆𝜎𝑚,𝑘𝑒 , as:

∆𝜎𝑚,𝑘𝑒 = 𝜎𝑚

𝑚,𝑘𝑒 − 𝜎𝑛

𝑚,𝑘𝑒 , (36)

and for the structural bending stress range, ∆𝜎𝑏,𝑘𝑒 , as:

∆𝜎𝑏,𝑘𝑒 = 𝜎𝑚

𝑏,𝑘𝑒 − 𝜎𝑛

𝑏,𝑘𝑒 , (37)

at the point under evaluation (see structural hot-spot in Figure 33). Subsequently, the elastically

calculated structural stress range, ∆𝜎𝑘𝑒, can be determined as:

∆𝜎𝑘𝑒 = ∆𝜎𝑚,𝑘

𝑒 + ∆𝜎𝑏,𝑘𝑒 , (38)

and the structural strain range, ∆휀𝑘𝑒, as:

∆휀𝑘𝑒 =

∆𝜎𝑘𝑒

𝐸𝑇𝑚𝑒𝑎𝑛

. (39)

These values are further used in the determination of the corresponding ranges of local nonlinear

structural stress, ∆𝜎𝑘, and strain, ∆휀𝑘, by simultaneously solving Neuber’s Rule,

∆𝜎𝑘 ∙ ∆휀𝑘 = ∆𝜎𝑘𝑒 ∙ ∆휀𝑘

𝑒, (40)

Page 75: Finite Element Modeling of Buried Longitudinally Welded

58

and modification of Eq. (56) in the form of a model for the material hysteresis loop stress-strain

curve,

∆휀𝑘 =∆𝜎𝑘

𝐸𝑇𝑚𝑒𝑎𝑛

+ 2 (∆𝜎𝑘

2𝐾𝑐𝑠𝑠)

1

𝑛𝑐𝑠𝑠. (41)

The solution for the obtained relationship,

∆𝜎𝑘𝑒∙∆ 𝑘

𝑒

∆𝜎𝑘=

∆𝜎𝑘

𝐸𝑇𝑚𝑒𝑎𝑛

+ 2 (∆𝜎𝑘

2𝐾𝑐𝑠𝑠)

1

𝑛𝑐𝑠𝑠, (42)

i.e., ∆𝜎𝑘 and the corresponding ∆휀𝑘, can be found by the numerical method used in 1.4.1.2 for the

calculation of a tangent point, and modified for low-cycle fatigue by using

∆𝜎𝑘 = (𝐸𝑇𝑚𝑒𝑎𝑛

1−𝜈2 ) ∆휀𝑘, (43)

and implemented similarly to the merged stress-strain curve in APPENDIX A – MATLAB (see

Neuber’s rule).

Level 2 – Method C takes into account local plasticity using Neuber’s rule. A notch, i.e., a

local sharp change (discontinuity) of a component’s geometry, is one of the major contributors to

stress rise. Heinz Neuber developed the model, which relates local linear stress-strain conditions

at the notch root to non-linear stress-strain conditions above the proportionality limit (Figure 22)

[83] [84] [85]. This model has been proven to yield estimates very close to experimental data [83].

Neuber’s rule calculates the local notch-root stress-strain history using the total strain energy

density, 휀 ∙ 𝜎, [84] [85].

Page 76: Finite Element Modeling of Buried Longitudinally Welded

59

Figure 22 Neuber’s relationship between linear and non-linear stresses and strains [83] [84]

[85]

Second, the mean stress, 𝜎𝑚𝑒𝑎𝑛,𝑘, and the stress ratio, 𝑅𝑘, for each cycle can be calculated

from

𝜎𝑚𝑒𝑎𝑛,𝑘 =𝜎𝑚𝑎𝑥,𝑘 + 𝜎𝑚𝑖𝑛,𝑘

2 (44)

and

𝑅𝑘 =𝜎𝑚𝑖𝑛,𝑘

𝜎𝑚𝑎𝑥,𝑘 (45)

respectively, where the maximum stress in a cycle was obtained as 𝜎𝑚𝑎𝑥,𝑘 from

𝜎𝑚𝑎𝑥,𝑘 = 𝑚𝑎𝑥[( 𝜎𝑚𝑚,𝑘𝑒 + 𝜎𝑚

𝑏,𝑘𝑒 ), ( 𝜎𝑛

𝑚,𝑘𝑒 + 𝜎𝑛

𝑏,𝑘𝑒 )]. (46)

Similar procedure is used to obtain minimum stress, 𝜎𝑚𝑖𝑛,𝑘 from

𝜎𝑚𝑖𝑛,𝑘 = 𝑚𝑖𝑛[( 𝜎𝑚𝑚,𝑘𝑒 + 𝜎𝑚

𝑏,𝑘𝑒 ), ( 𝜎𝑛

𝑚,𝑘𝑒 + 𝜎𝑛

𝑏,𝑘𝑒 )], (47)

𝜎𝑚𝑒𝑎𝑛,𝑘 and 𝑅𝑘, including the structural stress exponent, 𝑚𝑠𝑠 = 3.6, are subsequently used in the

calculation of the mean stress correction factor, 𝑓𝑀,𝑘,

if 𝜎𝑚𝑒𝑎𝑛,𝑘 ≥ 0.5 ∙ 𝑌𝑆𝑇𝑚𝑎𝑥 && 𝑅𝑘 > 0 && 𝑎𝑏𝑠(∆𝜎𝑚,𝑘

𝑒 + ∆𝜎𝑏,𝑘𝑒 ) ≤ 2 ∙ 𝑌𝑆𝑇𝑚𝑒𝑎𝑛

;

𝑓𝑀,𝑘 = (1 − 𝑅𝑘)^(1/𝑚𝑠𝑠);

else 𝑓𝑀,𝑘 = 1; end

(48)

σ,

MP

a

ε, mm/mm

𝜎𝑛𝑜𝑡𝑐ℎ = 𝑘𝑡𝜎𝑛𝑜𝑚 (𝜎휀)𝑛𝑜𝑡𝑐ℎ = (𝑘𝑡𝜎𝑛𝑜𝑚)2 𝐸⁄

𝜎 = 𝐸휀

𝜎 = 𝐾휀𝑛 𝜎𝑛𝑜𝑡𝑐ℎ = 𝑘𝑡𝜎𝜎𝑛𝑜𝑚

휀 𝑛𝑜

𝑡𝑐ℎ

=𝑘

𝑡휀 𝑛

𝑜𝑚

휀 𝑛𝑜

𝑡𝑐ℎ

=𝑘

𝑡휀 𝑛

𝑜𝑚

휀𝑛𝑜𝑚

𝜎𝑛𝑜𝑚

𝜎𝑦

휀𝑦

Neuber hyperbola Linear stress-strain Non-linear stress-strain Proportionality limit

Page 77: Finite Element Modeling of Buried Longitudinally Welded

60

The reader is referred to Eq. (57) used to calculate 𝜎𝑇.

Third, the equivalent structural stress range parameter, ∆𝑆𝑒𝑠𝑠,𝑘, is calculated from

∆𝑆𝑒𝑠𝑠,𝑘 =∆𝜎𝑘

𝑡𝑒𝑠𝑠(

2−𝑚𝑠𝑠2∙𝑚𝑠𝑠

)∙𝐼

1𝑚𝑠𝑠 ∙𝑓𝑀,𝑘

, (49)

where 𝐼 is the correction factor,

𝐼1

𝑚𝑠𝑠 =1.23−0.364𝑅𝑏,𝑘−0.17𝑅𝑏,𝑘

2

1.007−0.306𝑅𝑏,𝑘−0.178𝑅𝑏,𝑘2 , (50)

𝑅𝑏,𝑘 is the ratio,

𝑅𝑏,𝑘 =|∆𝜎𝑏,𝑘

𝑒 |

|∆𝜎𝑏,𝑘𝑒 |+|∆𝜎𝑚,𝑘

𝑒 |, (51)

and the structural stress effective TW, 𝑡𝑒𝑠𝑠, can be defined using the actual WT, 𝑡, using the

statement

if 𝑡 ≥ 16 𝑚𝑚; 𝑡𝑒𝑠𝑠 = 𝑡; else 𝑡𝑒𝑠𝑠 = 16 𝑚𝑚 end. (52)

Fourth, the number of cycles to failure, 𝑁, is calculated from

𝑁 =𝑓𝐼

𝑓𝐸(

𝑓𝑀𝑇

∆𝑆𝑒𝑠𝑠,𝑘)

1

ℎ, (53)

where the fatigue improvement factor, 𝑓𝐼, can be kept as equal to unity in the case no fatigue

improvement method is considered, the environmental modification factor, 𝑓𝐸 , can be selected to

be equal to 4, 𝑓𝑀𝑇 is the temperature adjustment factor defined as

𝑓𝑀𝑇 =𝐸𝑇𝑚𝑒𝑎𝑛

𝐸𝑅𝑇, (54)

and 𝐶 = 11577.9 and ℎ = 0.3195 are the coefficients for the welded joint fatigue curve of ferritic

and stainless steels for lower prediction interval (99%, −3𝜎). Similar to the fatigue assessment

methods previously discussed (see Eq. (32)), no allowance for corrosion is included in Eq. (53).

Page 78: Finite Element Modeling of Buried Longitudinally Welded

61

Fifth, the fatigue damage (FD) is calculated similarly to the previous sections, by using

Miner’s rule, Eq. (34).

3.9. Summary and Problem Definition

The research work will focus on elastic (3.8.2 Elastic Fatigue Analysis and 3.8.3 Modified

Elastic Fatigue Analysis) and elastic-plastic (3.8.4 Elastic-Plastic Fatigue Analysis) fatigue

assessment methods (ASME BPVC Section VIII Part 2 [22]), of submerged-arc welded (SAW)

large-diameter oil pipelines with standard outside diameters (ODs) of 610-1219 mm and standard

wall thicknesses (WTs, 𝑡) of 9.53-23.83 mm produced using the UOE process from X56 grade

steel and operated onshore 1000 mm below the ground (BD) at a permanent internal pressure (IP)

of 10 MPa and service temperature of 80 °C. There parameters have been selected with the

reference to the Keystone pipeline design. This will include modeling of longitudinal seam-welded

regions with associated misalignments and defects within the manufacturing tolerances of radial

offset 𝛿𝑜 of 1.5 mm, angular peaking 𝛿𝑝 of 3.2 mm, weld reinforcement 𝑊ℎ of 2 mm, and

reinforcement angle 𝜃 of 27°, with the loading cycles obtained from a common in-service loading

spectrum for a period of 50 years of design fatigue life to provide more accurate service-specific

predictions, as shown in Table 12. An elastic-plastic model (3.8.4) will be included in the analysis

to account for hot-spot plasticity developed during cycling.

Table 12 Pipeline parameters [in or (mm)] considered in this research

Pipeline Geometry Pipeline Defects In-Service Loading

𝐷 𝑡 𝛿𝑜 𝛿𝑝 𝑊ℎ BD 𝑃𝑖 Cycle

16-56 0.188-2.050 ≤ t/10 ≤ 0.125 ≤ 0.157 24 1400* Spectrum

(457-1422) (7.1-52) (1.5) (3.2) (4) (≥ 600) 10** Spectrum

* in [psig]; ** in [MPa]

Page 79: Finite Element Modeling of Buried Longitudinally Welded

62

The thesis will review the applicability of the available standards to large-diameter

pipelines and will develop a fatigue assessment methodology that will address the range of

parameters related to pipeline manufacturing and service conditions in order to meet industrial

needs in non-conservative and cost-effective pipeline design, and will provide an optimization

scheme for existing standards. The work will focus on the analysis of mechanical stresses and

strains in a pipeline, with special attention to the effects of misalignment.

The thesis will focus on the calculation of acceptable pipe WTs with combinations of

commonly observed manufacturing misalignments, including axial and angular misalignments,

from the perspective of fatigue-safe designs applied to the range of commonly used pipe diameters

and discusses the conservatism involved in elasticity-based methods when selecting WTs for pipe

design.

Page 80: Finite Element Modeling of Buried Longitudinally Welded

63

Chapter 4 MODEL DEVELOPMENT

In Chapter 4 the analytical and FE methods will be used to determine stresses at critical

locations of pipeline and to predict the fatigue life of pipeline based on its geometry tolerable by

design codes and typical loading expected in service. The wall thickness (WT) of a pipeline in the

initially conservative FE model (INPUT) will be reduced gradually until the cumulated fatigue

damage (FD) reaches unity value (GOAL), as shown in Figure 23. This process will be repeated

for elastic and elastic-plastic fatigue assessment methods (ASME BPVC Section VIII Part 2 [22]).

The results of different fatigue methods will be analyzed and compared in the next Chapter.

Figure 23 Flow chart showing process of model refinement

The design of the SAW-UOE manufactured pipeline started first with the models available

in the literature, including the effects of pipe geometry, weld profile, weld misalignments, internal

loading due to the transported medium, and external loading due to soil. Stress concentration

factors due to each effect were calculated and their sum was used as a multiplication factor for the

stress in the hoop direction to calculate the design stresses developed at the structural hot-spot.

This initial step was performed to obtain a mathematical model for a specific case discussed in this

study. Next, the fatigue life predictions were based on the verified FE model (verified hot-spot

stresses) for more detailed study of the component stresses across the WT of a pipe, including

Conservative Model

Reduce WT and Update

the FE Model

Obtain Stresses/Strains

Predict Fatigue Life

YES

NO Cumulated

FD < 1

Non-Conservative Model

GOAL INPUT

Page 81: Finite Element Modeling of Buried Longitudinally Welded

64

membrane stress, bending stress, and peak stress components, which are needed for accurate

predictions of FD as prescribed by governing design codes.

The calculations were performed in two steps: first, stresses and strains were obtained at a

structural hot-spot after applying a histogram of service loading onto the FEM modeled pipe; the

second, the FEM results were used in the fatigue analysis, including stress linearization,

calculation of equivalent stresses and strains, numbers of cycles to failure and fatigue damage.

Multiple software packages were used for the calculations, including Computer Aided Engineering

(CAE) software ABAQUS by Dassault Systèmes [77] [86] used in the FEM analysis, and

MATLAB by MathWorks, Inc. [87], which was used in other calculations and analyses.

4.1. Static Finite Element Model

The ABAQUS model comprises a pipe of commonly used standard OD and WT, weld

profile, and soil, as well as thermal loading and pressure history to accurately capture pipeline and

service conditions.

4.1.1. Geometry of Model

Four widely used pipeline ODs were utilized, and the pipeline WT was calculated

according to ASME B31.4 [23] and ASME B36.10 [49] using an internal pressure, 𝑃𝑖, of 10 MPa,

and temperatures of 0-80 °C, normally used in oil transportation, and adjusted to the nearest

standard WT as prescribed in ASME B36.10 [49], Table 13. The WTs were incrementally changed

using standard WTs from ASME B36.10 [49] to study the development of fatigue damage due to

service loading.

Page 82: Finite Element Modeling of Buried Longitudinally Welded

65

Table 13 Standard pipeline WTs for selected ODs and steel material at an internal pressure

of 10 MPa

𝐷 [mm] 610 864 914 1219

𝑡 [mm] 11.91 17.48 19.05 23.83

A soil box (Figure 24) was selected over the approximation of interaction between pipe

and soil with springs. The soil box is more convenient due to the weld profile included in the model

and is expected to provide accurate representation of sustained load due to gravity [58].

Figure 24 Geometry of model

The dimensions of the soil box were adjusted to simulate a pipeline trench with the burial

depth (BD) defined by the pipeline construction standard CSA Z662 [5]. In [5], a minimum BD of

0.6 m is required for oil pipelines located in non-developed areas in Canada.

One of the most critical features of pipeline is the welded region. The weld zone is a

complex and heterogeneous region that contains different microstructural features of different

morphologies that have different fatigue properties [88] [89] which have been found to change

with temperature [90]. Furthermore, the welded region of a pipe is highly susceptible to corrosion

and hydrogen embrittlement, especially at the heat-affected zone (HAZ) [91]. A typical double

SAW weld profile from the UOE process is shown in Figure 25. Three distinctly different regions

can be found within the welded region, including base plate, HAZ, and weldment. These regions

normally have different microstructures and mechanical properties.

Page 83: Finite Element Modeling of Buried Longitudinally Welded

66

Figure 25 Schematic of the SAW-processed pipe region

It has been shown by other researchers that weld bead geometry has a great impact on

stresses developed at the weld toe, which has been found to be extremely sensitive to weld

reinforcement, its angle to the welded plate, and a weld toe radius, as discussed earlier; bending

due to weld misalignment is also known as a great contributor to stress rise.

Special attention was paid to the modeling of the weld profile to capture its effects on the

hot-spot’s stresses. Although some of the mentioned critical weld parameters are discussed in

standards, some of them are only presented schematically. For instance, the weld bead width

calculation, (𝑊𝑤 = 0.5𝑊𝑇), and the weld root/toe radius definition, (𝑊𝑟 ≥ 0.25𝑊𝑇), are specified

only in BS 7608 [24], and none of the standards explicitly discusses or references a way to combine

radial 𝛿𝑜 and angular 𝛿𝑝 (Figure 1) misalignments. Weld profile curvature is also omitted from the

standards. The standards do however distinguish between weld types by introducing either the

weld joint efficiency factors used as a multiplier of the appropriate allowable stress for a given

material (ASME BPVC Section VIII Part 2 [22]) or weld quality specifications (weld classes) used

to determine the coefficients for fatigue curves (BS 7608 [24]). It seems that there is no unified

approach available in standards to be used as an input for FEM. The authors of this report believe

that there is a need for more accurate and statistically supported definitions of weld geometry and

misalignments that can be used in FEM for elastic and elastic-plastic fatigue analyses. Therefore,

the allowable weld misalignments were included in the model based on tolerances found in the

Base Plate

External

Weldment

Internal

Weldment

HAZ

HAZ

HAZ

HAZ

Base Plate

Page 84: Finite Element Modeling of Buried Longitudinally Welded

67

ASME, API, and BS codes, as well as the weld toe radius to more accurately account for the UOE

manufacturing defects (Table 14).

Table 14 Pipeline defects

𝛿𝑜 [mm] 𝛿𝑝 [mm] 𝑊ℎ [mm] 𝑊𝑟 [mm]

≤ 1.5 ≤ 3.2 ≤ 4 5

An average weld profile was obtained from available literature such as [92] [93] by

normalizing the traced SAW weld profiles of various transmission pipelines with ODs ranging

from 864 mm to 1219 mm ODs with WTs between 8 mm and 22 mm. The variations in 30 SAW

weld profiles were studied (Figure 26).

Figure 26 Traces of weld profiles used to generate an average weld profile

The resultant average weld profile is shown in Figure 27.

Figure 27 Geometry of the weld bead profile showing (solid dots) experimental data and

(solid line) 4th-order polynomial approximation

0

1

2

3

0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19Wel

d H

eigh

tW

h,

mm

Weld Width Ww, mm

Page 85: Finite Element Modeling of Buried Longitudinally Welded

68

An average weld bead geometry, obtained through this analysis in the form of a 4th order

polynomial, i.e.,

𝑊ℎ = −9.215 × 10−5𝑊𝑤4 + 3.485 × 10−3𝑊𝑤

3 − 6.291 × 10−3𝑊𝑤2 + 0.5671𝑊𝑤

+ 0.03417,

(55)

was used to obtain 𝑊ℎ [𝑚𝑚] and build the outside weld bead part. The inside weld bead part was

scaled down by a factor of 1.098 to match an average measured weld width 𝑊𝑤 and weld height

𝑊ℎ of a weld reinforcement, as shown in Table 15. The values of 𝑊𝑤 and 𝑊ℎ were found to vary

insignificantly with OD and WT used to obtain Eq. (40). The values for 𝑊ℎ from Table 15 are

well below those prescribed in API 5L [51] and CSA Z662 [4], [5]. The Inside and outside weld

beads were modeled within 0-1 mm of misalignment, which is less than that prescribed in API 5L

[51] for the WTs studied, i.e., 3-4 mm.

Table 15 Weld bead dimensions obtained in this study

Weld Bead Location 𝑊𝑤 [mm] 𝑊ℎ [mm]

Outside 18.98 ±4.71 2.02 ±0.40

Inside 17.29 ±4.58 1.84 ±0.44

The weld profile and misalignments were carefully modelled in order to capture their

effects on the hot-spot’s stresses/strains (Figure 28).

Page 86: Finite Element Modeling of Buried Longitudinally Welded

69

Figure 28 Geometry of the weld region including (bold white line) radial and (bold black

line) angular misalignments

The reinforcement angle 𝑊𝛼 was measured to be 27.3° at a hot-spot (circled area in Figure

28) for a non-misaligned weld and 33.5° for a misaligned weld observed due to the combined

effect of axial misalignment (𝛿𝑜 = 1.5 𝑚𝑚) and angular misalignment (𝛿𝑝 = 3.2 𝑚𝑚) specified

according to CSA Z245.1 [4] and API 5 L [51], as shown in Figure 28. The weld toe radius 𝑊𝑟

was set to 5 mm, and assuming there is no weld undercut.

4.1.2. Material Model

The material model for FEA was developed based on the approaches published in the

literature. The calculations were based on the material properties of pipeline steel extracted from

API 5L [51], Table 16. Although the material properties of weld and parent material may differ

and result in different fatigue strengths, the present research focuses on stress concentration due to

geometry of a welded region. Therefore, same properties have been used for all parts of a modelled

pipeline, Table 16.

Table 16 Pipeline steel material

Steel 𝐸

[MPa]

𝜈 𝜌 𝛼

[mm/K]

𝜎𝑦

[MPa]

𝜎𝑢

[MPa]

𝑛 𝐾

[MPa]

X56 207000 0.3 0.0078 1×10-5 358 455 -0.24 684

Page 87: Finite Element Modeling of Buried Longitudinally Welded

70

The data from Table 16 were used to generate the stress-strain curves for installation and

service temperatures of 0 °C and 80 °C respectively.

The true stress-strain response is modeled using the Ramberg-Osgood equation,

휀 =𝜎

𝐸+ (

𝜎

𝐾)

1

𝑛, (56)

as dictated by ASME BPVC Section VIII Part 2 [22]. The coordinates of the yield point at the

assessed temperatures 𝑇 were calculated from the yield stress and defined as

𝜎𝑦𝑇= 𝜎𝑦 ∙ 𝑒(𝐶0+𝐶1𝑇+𝐶2𝑇2+𝐶3𝑇3+𝐶4𝑇4+𝐶5𝑇5) (57)

and from the Young’s modulus,

𝐸𝑇 = (7 ∙ 10−17𝑇6 + 10−12𝑇5 − 10−9𝑇4 + 6 ∙ 10−5𝑇2 − 0.0646 ∙ 𝑇 + 203.66) × 103. (58)

In [22], the constants for Eq. (57) were taken as: 𝐶0 = 3.38037095 × 10−2, 𝐶1 =

−1.7355438 × 10−3, 𝐶2 = 8.32638097 × 10−6, 𝐶3 = −2.11471664 × 10−8, 𝐶4 =

3.29874954 × 10−11, and 𝐶5 = −2.69329508 × 10−14 (ASME BPVC Section VIII Part 2 [22]).

One of the connecting points, with coordinates [yield stress, yield strain] on Hooke’s curve,

defined in Eq. (1), lies outside the Ramberg-Osgood curve and can be connected to it by the

tangent line, as shown in Figure 29.

Figure 29 Construction of a tangent to Ramberg-Osgood’s curve from yield point on

Hooke’s curve

Page 88: Finite Element Modeling of Buried Longitudinally Welded

71

Therefore, derivation of a function for the tangent line has the initial form of

𝑥1 − 𝑥0 = 𝑚(𝑦1 − 𝑦0), (59)

where [𝑥0, 𝑦0] is the coordinate for the yield point of a material,

[𝑥0, 𝑦0] = [𝜎𝑦

𝐸, 𝜎𝑦],

(60)

[𝑦1, 𝑥1] is the coordinate for point of tangent line point of a tangent line on a Ramberg-Osgood

curve,

[𝑥1, 𝑦1] = [𝜎

𝐸+ (

𝜎

𝐾)

1

𝑛, 𝜎],

(61)

and 𝑚 = 𝑓′(𝜎) is the derivative of Eq. (56) w.r.t. 𝜎,

𝑚 = 𝑓′(𝜎) =𝐾

−1𝑛∙𝜎

1𝑛

−1

𝑛+

1

𝐸.

(62)

This results in

[𝜎

𝐸+ (

𝜎

𝐾)

1

𝑛] −

𝜎𝑦

𝐸= [

𝐾−

1𝑛∙𝜎

1𝑛

−1

𝑛+

1

𝐸] (𝜎 − 𝜎𝑦),

(63)

which can be solved for 𝜎 numerically by finding the lowest difference between the solutions of

the two curves on the left-hand side and right-hand side of the equation, as shown in Figure 30.

Figure 30 Example of a numerical solution for the tangent point on a Ramberg-Osgood

curve

300

350

400

450

500

-0.005 0.005 0.015 0.025 0.035 0.045 0.055 0.065 0.075 0.085 0.095

Stre

ss σ

, MP

a

Strain ε, mm/mm

Solution for left-hand side of Eq. (9) at pipe installation temperatureSolution for right-hand side of Eq. (9) at pipe installation temperatureSolution of Eq. (9) at pipe installation temperatureSolution for left-hand side of Eq. (9) at pipe service temperatureSolution for right-hand side of Eq. (9) at pipe service temperature

Page 89: Finite Element Modeling of Buried Longitudinally Welded

72

The equation for the tangent line from the yield point to the Ramberg-Osgood curve can

be found from Eq. (63), which yields as solution for the strain coordinate, 휀 = 𝜎𝑦 𝐸⁄ . The strain

hardening exponent 𝑛 and strength coefficient 𝐾 were adjusted to the installation and service

temperatures using linear interpolation of the experimental data for a low-alloy carbon steel found

in ASME BPVC Section VIII Part 2 [22]. The resultant stress-strain curves are shown in Figure

31. This approach was implemented to incorporate the elastic region into the Ramberg-Osgood

formulation and to avoid errors related to apparent plasticity at stresses lower than the yield stress.

The generated stress-strain curves were implemented in the FEM analysis.

Notably, the obtained cyclic stress-strain curve is very similar to that of the Prager

formulation found in [22] until the tangent point at strain 휀, and after that point the Prager solution

for the stress-strain curve diverges toward lower stresses.

Figure 31 True Stress-Strain curves for pipe steel material

A complete MATLAB implementation of the numerical solution for the merged stress-

strain curve is given in APPENDIX A – MATLAB .

The soil box around the pipe was modelled using Mohr-Coulomb plasticity as clay with

the properties summarized in Table 10 and was added to the model to account for the effects of

soil on buried pipe.

0

100

200

300

400

500

0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05 0.055

Stre

ss σ

, MP

a

Strain ε, mm/mm

Installation Temperature

Service Temperature

Page 90: Finite Element Modeling of Buried Longitudinally Welded

73

4.1.3. Boundary Conditions

The boundary conditions applied to the model are summarized below in Table 17. The pipe

was allowed to move freely in the 𝑋𝑌 plane and to expand radially within the soil box [58] (but

not longitudinally) to simulate the soil-pipe interaction and underground restraint of a pipe [94],

as shown in Figure 32. Gravity was added to the model to account for stresses due to soil pressure

on the pipe. The displacement components 𝑈1, 𝑈2, and 𝑈3 in Table 17 act in the 𝑋, 𝑌, and 𝑍

directions respectively, as shown in Figure 32.

Table 17 Displacement constraints

Entity 𝑈1 𝑈2 𝑈3 Comment

Pipe Ends • Longitudinal displacement restrained

Soil Fronts • Frontal displacement restrained

Soil Sides • • Only vertical displacement allowed

Soil Bottom • • • No displacement allowed

Figure 32 Pipe (highlighted by circles) surrounded by a soil box with constraints

Page 91: Finite Element Modeling of Buried Longitudinally Welded

74

A static pressure of 10 MPa was evenly distributed over the inside surface of the pipe in

order to simulate the hydrostatic condition. A loading history from Figure 17 was applied to the

FE model to study the fatigue damage.

Thermal loading due to in-service oil transmission was included to account for stresses due

to differences between the installation and service conditions of 0 °C and 80 °C respectively. While

a temperature of 0 °C was kept constant for the soil, the temperature was increased to 80 °C for

the pipeline before application of the pressure cycles.

4.1.4. Model Meshing and Convergence

The pipe was partitioned to allow for a gradual mesh refinement of the pipe and meshed

with the 20-node quadratic brick elements with reduced integration (C3D20R, hex-dominated

elements [86]). The parts of the pipe away from the weld were meshed with a single through-

thickness 10 mm long elements, reduced in size to 1 mm around the weld region and further

reduced to 0.2 mm at the weld toe (hot-spot), as shown in Figure 33.

Figure 33 Meshing of a pipe showing detailed meshing at the hot-spot (black line indicates

the path used for an SCL)

Page 92: Finite Element Modeling of Buried Longitudinally Welded

75

The soil box was meshed with 15-node quadratic triangular prism wedge elements (C3D15,

[86]), which allowed for a sufficiently accurate modeling of surface-to-surface interactions in a

soil-pipe system, as shown in Figure 34. The element size of 50 mm near the external pipe surface

was implemented and refined to 5 mm elements at the pipe weld, and the edges of the soil box

were meshed with 500 mm elements, which was also the size for the elements in the longitudinal

direction throughout the FEM model.

Figure 34 Meshing of soil box around pipe

A convergence test was done on the model to confirm its computational validity. First, the

FEM mesh was refined globally to stabilize the solution for the nominal stresses and strains away

from the structural discontinuity. Second, the FEM mesh in the vicinity of the hot-spot was refined

to accurately capture the stresses and strains at the discontinuity.

The results presented in Figure 35 show a relatively quick convergence of a solution, both

globally and locally.

Page 93: Finite Element Modeling of Buried Longitudinally Welded

76

Figure 35 Refinement of global (away from discontinuity) and local (at the weld toe)

meshes showing von Mises Stress/Strain – Mesh Element Size relationship

The convergence of the global solution was found to be strongly related to a surface-to-

surface interaction between pipe and soil. Coarser element sizes in the pipe mesh and an overly

refined mesh in the soil box led to inaccurate solution that is represented by penetration of the

modeled bodies in a final deformed state, specifically in the weld region. Notably, it seems that

there is no similar thorough study in the literature on the modeling of pipe-soil interactions

involving s weld profile. The misaligned weld region is a geometrical irregularity and hence

required special care in the selection of the finite elements. The penetration of the modeled bodies

was dramatically reduced and localized to some pipe-soil contact points when the element size of

280.667

285.241285.129284.844284.969285.009285.040

374.347

355.267354.214352.433352.372352.282

0.001910

0.0019370.0019370.0019350.0019360.0019360.001936

0.003228

0.0049070.004851

0.0047910.0047940.004786

0.0015

0.0025

0.0035

0.0045

0.0055

280

300

320

340

360

380

0.01 0.1 1 10 100

Stra

in ε

, mm

/mm

von

Mis

es S

tres

s σ

, MP

a

Mesh Element Size, mm

Global Mesh - Mises Stress Local Mesh - Mises StressGlobal Mesh - Strain Local Mesh - Strain

Page 94: Finite Element Modeling of Buried Longitudinally Welded

77

a pipe mesh reduced to 50 mm. This computational inaccuracy is eliminated at mesh element sizes

below 40 mm, and the solution is stable when elements finer than 20 mm are used.

Another indication of the model’s convergence is that the solution for a nominal hoop

stress, found by subtraction of hoop stresses due to soil and heat from the final state, which also

includes pressure, is in good agreement with a closed form solution for hoop stress,

𝑆𝐻 =𝑃𝐷

2𝑡, (64)

showing only a 0.42% difference. It is important that the solution for the thermal load is also in

reasonable agreement with the procedure in ASME B31.4 [23], where the longitudinal stress due

to heat in an anchored and fully restrained portion of a buried pipe is

𝑆𝐿 = −𝐸𝛼∆𝑇 + 𝑆𝐻𝜈, (65)

showing only a 1.47% difference.

4.1.5. Data Extraction

The results of the simulation in ABAQUS were exported in *.rpt format, and contained

linearized stresses for each cycle, including the start and end points, and converted to *.xlsx format

for convenient use in MATLAB; individual component stresses were extracted by using a

following example command, which helps to separate numerical data from alphabetical.

[v,T,vT]=xlsread(FileName,2,CellName);

a=regexp(vT,'\s+','split');

n=numel(a{1});

m=numel(a); ComponentStress=transpose(reshape(str2double([a{:}]),n,m));

CellName=['A' num2str(Start/End) ':' 'A' num2str(Start/End)];

In (FileName,2,CellName), 2 is the page number in *.xlsx file.

The Start and End coordinates for a cycle within the *.xlsx file were calculated based on

the location of the extracted data. An example of linearized stress at a single cycle point (OD

Page 95: Finite Element Modeling of Buried Longitudinally Welded

78

914 mm WT 17.5 mm at a pressure of 2.8 MPa) is presented in APPENDIX D – ABAQUS Report

Example. The number of intervals for the SCL, which is defined in Figure 33, is a variable

parameter selected by the user in ABAQUS during export of the linearized stresses and is set as

constant throughout the analysis, 40, for a convenient subsequent reading of those data in

MATLAB. The “for loop” with the coded data extraction algorithm was used to retrieve the data

points for all pressure cycles sequentially; each sequence calculated the FD for one cycle and added

it to the total value. The number of extracted points is the double of number of pressure cycles, 60

in this research, and the maximum was observed to be more than 170 (106 lines per each cycle in

the report file) depending on the computational iterations during the ABAQUS calculations. The

number of iterations was observed to increase with plasticity due to high-pressure cycles.

ABAQUS produces multiple data points when solving for a cycle due to adjustments to the internal

time increments used in the calculations. The total number of lines can be more than 18,000.

Therefore, a separate subroutine was coded in MATLAB in order to eliminate the intermediate

points between the Start and End cycle points.

An example of conditioning of the ABAQUS report file (*.rpt or *.xlsx) and extraction of

the membrane component of the linearized stress (bending and peak components of stress can be

extracted similarly) is presented in APPENDIX E – MATLAB Code for the ABAQUS Data.

After extraction the data (hot-spot stresses and strains) were used in fatigue analyses

presented in Chapter 3: Elastic Fatigue Analysis (3.8.2), Modified Elastic Fatigue Analysis (3.8.3),

and Elastic-Plastic Fatigue Analysis (3.8.4).

Page 96: Finite Element Modeling of Buried Longitudinally Welded

79

Chapter 5 RESULTS AND DISCUSSION

This Chapter summarizes the results of the FEM and fatigue analyses. The effects of the

misalignment on the local stresses and the use of different fatigue assessment methods in the

calculation of allowable WT are discussed. The results of FE model are compared to the results of

an approximate model developed to simply calculate local stresses. Specifically, the hoop stress

and the longitudinal stress obtained from the FE model are verified far from the weld.

Subsequently, a detailed analysis of SCFs due to each loading mode (internal pressure, soil

pressure, temperature), associated with varying degrees of misalignment, is given. Finally,

validation of newly proposed analytical method of obtaining the stresses at four weld toes is

detailed.

5.1. Finite Element Model

Since this study focuses primarily on elastic and elastic-plastic fatigue analyses, it is

important to show the differences in the ABAQUS simulation results between the two models. A

pipe modeled with an OD of 914 mm, WT of 14.3 mm, and a maximum tolerable misalignment

was selected as an example to show the stress distribution plots in the through-thickness direction

of a pipe due to expected in-service loading. Figure 36 and Figure 37 show the hoop stress maps

with the misaligned pipe calculated using the elastic and elastic-plastic models respectively. The

stress in the hoop direction was selected for comparison, as it is of greater importance for a

longitudinally welded pipe, in which the plane of the hypothetical crack is situated perpendicular

to the hoop direction and opens under fluctuations of the hoop stress. The analyses do not show

any difference in the hoop stresses due to soil, as shown in Figure 36(a) and Figure 37(a), or due

to heat, Figure 36(b) and Figure 37(b). However, a significant difference in the stress distributions

can be observed between the models when the design pressure of 10 MPa is applied.

Page 97: Finite Element Modeling of Buried Longitudinally Welded

80

(a) Soil/Gravity (b) Soil/Gravity + Heat

(c) Soil/Gravity + Heat + Pressure

Figure 36 Hoop stress distribution maps for a misaligned pipe during elastic loading

(a) Soil/Gravity (b) Soil/Gravity + Heat

(c) Soil/Gravity + Heat + Pressure

Figure 37 Hoop stress distribution maps for a misaligned pipe during elastic-plastic loading

Page 98: Finite Element Modeling of Buried Longitudinally Welded

81

Stre

ss σ

, MP

a

(a) (b)

Stre

ss σ

, MP

a

(c) (d)

Wall Thickness t, mm Wall Thickness t, mm Figure 38 Through-thickness (curved) actual and (linear) linearized stress distributions

obtained for a pipe of 914 mm OD and 14.3 mm WT from an SCL positioned at the hot-

spot (at 0 mm WT coordinate) normal to the pipe wall with no misalignment by using (a)

elastic and (b) elastic-plastic analysis, and with misalignment by using (c) elastic and (d)

elastic-plastic analysis

-200

-100

0

100

200

300

400

500

600

700

800

900

0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15

-200

-100

0

100

200

300

400

500

600

700

800

900

0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15

S11S11mS11mbS22S22mS22mbS33S33mS33mb

-200

-100

0

100

200

300

400

500

600

700

800

900

0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15

-200

-100

0

100

200

300

400

500

600

700

800

900

0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15

Stress along Y axis of model – transverse direction

Linearized stress along Y axis – membrane

Linearized stress along Y axis – membrane + bending

Stress along X axis of model – hoop direction

Linearized stress along X axis – membrane

Linearized stress along X axis – membrane + bending

Stress along Z axis of model – longitudinal direction

Linearized stress along Z axis – membrane

Linearized stress along Z axis – membrane + bending

Page 99: Finite Element Modeling of Buried Longitudinally Welded

82

Although the stress distribution patterns for both models seem to be similar, if not the same,

the stress is almost exclusively concentrated at the inside right-hand weld toe when the elastic

model considered, as shown in Figure 36(c); the elastic-plastic behavior of the structure resulted

in dissipation of stress over a much larger volume of material adjacent to the mentioned weld toe

part, Figure 37(c), which experienced larger bending due to misalignment.

This difference between the stress distributions shown in Figure 36 and Figure 37 is due to

local plasticity developed in the elastic-plastic analysis, resulting in a lower through-thickness

stress gradient, which can be much steeper when a sharper weld toe radius or no radius is used.

The stress distributions shown in Figure 36 and Figure 37 are constant in longitudinal direction.

Notably, the through-thickness stress linearization is prescribed by the ASME standard to

obtain different components of stress at the hot-spot, including membrane stress 𝜎𝑚 and bending

stress 𝜎𝑏, as visualized in Figure 38 (see Hoop stress, 𝑆22). Stress linearization clearly eliminates

the discussed difference, e.g., compare Figure 38(a) to Figure 38(b) or Figure 38(c) to Figure 38(d),

with the elastic result showing only 0.18% and 5.44% larger values for 𝜎𝑚 and 𝜎𝑏 respectively.

This fact indirectly signifies the importance of a good convergence required from the FE model.

Comparing Figure 38(a) to Figure 38(c) or Figure 38(b) to Figure 38(d), misalignment was

found to have a significant impact on the stress rise at the hot-spot. While there is no change in the

membrane component of stress, 𝜎𝑚, the dramatic increase in hot-spot stress shown in Figure 38 is

mainly due to the contribution of its bending component, 𝜎𝑏, when weld misalignment is

considered, resulting in a 42.5% increase in total linearized stress, as shown in Figure 38(c-d).

Moreover, the elastic model showed 0.16% and 11.99% larger values for 𝜎𝑚 and 𝜎𝑏 respectively.

Increasing the misalignment results in a larger observed discrepancy between the stresses,

including linearized stresses, obtained through the elastic and elastic-plastic analyses.

Page 100: Finite Element Modeling of Buried Longitudinally Welded

83

Therefore, weld misalignment is expected to seriously compromise the fatigue

performance of a UOE-manufactured pipeline and is also expected to cause larger discrepancy

between solutions for FD based on the elastic and elastic-plastic analyses.

Although misalignment was found to significantly increase normal stresses, the shear

stresses of the assessed pipeline were calculated to be below 100 MPa, which is much lower than

normal stresses (> 500 MPa), and the relationship ∆𝜏 ≤ ∆𝜎/3 from ASME BPVC Section VIII

Part 2 [22] was satisfied. Since the structural shear stress range is negligible, multiaxial Elastic

Fatigue Analysis of Welds was not required and was not performed in this study. Furthermore, the

strain-life Level 3 Fatigue Assessment, ASME BPVC Section VIII Part 2 [22], based on the critical

plane approach was not performed either since the FEM results showed that the plane normal to

the hoop stress is the only critical plane.

5.1.1. Validation of FEM

This section is dedicated to the analytical calculation of a SCF that would reflect the actual

weld profile due to the SAW-UOE manufacturing process.

The SCFs due to maximum tolerable misalignments, listed with other parameters in Table

18, are as follows: 𝑘𝑚.𝑎𝑥𝑖𝑎𝑙 = 0.2571 (Eq. (14)), 𝑘𝑚.𝑎𝑛𝑔𝑢𝑙𝑎𝑟 = 0.5981 (Eq. (15)) (𝛽 = 0.3085,

Eq. (17)), 𝑘𝑚.𝑜𝑣𝑎𝑙𝑖𝑡𝑦 = 0.00001 (Eq. (16)). The initial ovality of a pipe does not modify the hoop

stress 𝑆ℎ much (𝑘𝑚.𝑜𝑣𝑎𝑙𝑖𝑡𝑦 = 0.00001 at 𝜃 = 90°), but axial and angular misalignments contribute

the most to the overall stress at the weld toe, resulting in maximum design stress 𝑆ℎ.𝑑𝑒𝑠𝑖𝑔𝑛 =

261 × 1.856 = 484 𝑀𝑃𝑎, which is the total stress at the pipe weld toe due to maximum allowable

misalignments and internal pressure inside the pipe.

Page 101: Finite Element Modeling of Buried Longitudinally Welded

84

Table 18 Parameters used in calculation of SCFs due to misalignments

𝛿𝑜

[mm]

𝑡

[mm]

2𝑙

[mm]

𝜈 𝜎𝑚𝑚𝑎𝑥

[MPa]

𝐸

[GPa]

𝛿𝑝

[mm]

𝜃

[°]

𝑝

[MPa]

1.5 17.5 92 0.3 261 207 3.2 90 10

In the present work, for the pipe without a coating, the bending stress at the weld toe (hot-

spot) due to soil is 𝜎𝑏.𝑠𝑜𝑖𝑙 = 16 𝑀𝑃𝑎 (Eq. (19)). Therefore, 𝑘𝑚.𝑠𝑜𝑖𝑙 = 0.0620 (Eq. (18)), which

magnifies the hoop stress to the value of 𝑆ℎ.𝑑𝑒𝑠𝑖𝑔𝑛 = 261 × (1 + 0.0620) = 277 𝑀𝑃𝑎.

Since the governing codes do not account for the variations in weld profile geometry, the

following part will be dedicated to a method that can be used to calculate SCF due to specific weld

profile. The SCF solution for SAW type of weld connection is not explicitly addressed in the

available literature [95] [96]. The SAW profile was treated as a fillet on stepped flat bar, which

has some features as similar to SAW. However, there are difficulties in obtaining the SCFs due to

weld width, 𝑊𝑤, which range, covered in [95] and [96], is different from that of SAW weld.

Additionally, for a completely automated model, there is a need for the set of equations to obtain

the SCF. Therefore, the following shows the development of the mathematical model for the stress

concentration for a fillet on stepped flat bar at the hot-spot adopted from [84].

The geometry of the SAW profile, shown in Figure 39(a), can be further developed into a

fillet on stepped flat bar, Figure 39(b), which has an approximated weld profile with maintained

weld height (crown), 𝑊ℎ, length of the weld attachment, 𝑊𝑤, weld radius, 𝑊𝑟, and wall thickness,

𝑡 [84]. This approximation has been found to yield an accurate SCF for SAW weld profile [84].

The approximated distributed shear load profiles for the fillet weld are shown schematically in

Figure 39(c) [84].

Page 102: Finite Element Modeling of Buried Longitudinally Welded

85

Figure 39 Schematic of (a) the SAW butt joint, represented in the form of (b) fillet in

stepped flat bar, showing (c) equivalent load at the base of reinforcement and (d) real shear

stress diagram with its (dashed line) approximation

The additional local stress on the upper fillet, ∆𝜎′, lower fillet, ∆𝜎′′, and on both fillets,

∆𝜎′′′, due to distributed shear loads on the surface of the reinforcement base, Figure 39(c), can be

obtained from

∆𝜎′ = 𝜏𝑚 (0.212 − 0.25𝑊𝑟

1

(𝑡 2⁄ )1 + 0.093𝑊𝑟

2

(𝑡 2⁄ )2), (66)

∆𝜎′′ = 𝜏𝑚 (−0.215𝑊𝑟

1

(𝑡 2⁄ )1 + 0.123𝑊𝑟

2

(𝑡 2⁄ )2), (67)

and

(a) (b)

(c) (d)

R l1 l

Wr

l l

Ww

t W

h

Ww

l

l1

τ'm

R

τxy

A

l2 x

τ m

a

b

c d

Page 103: Finite Element Modeling of Buried Longitudinally Welded

86

∆𝜎′′′ = 𝜏𝑚 (0.212 − 0.125𝑊𝑟

1

(𝑡 2⁄ )1+ 0.03

𝑊𝑟2

(𝑡 2⁄ )2) (68)

respectively, developed in [84], where 𝜏𝑚𝑙1= 𝜏𝑥 and 𝜏𝑚𝑙2

= 𝜏𝑚𝑙1

𝑥3

𝑊𝑟2 are the shear stresses, shown

in Figure 39(d), integrated in a range from 𝑥 = 0 to 𝑥 = 𝑊𝑟, and 𝑙1 and 𝑙2 are the section lengths

supporting the shear distributed load [84].

It follows that the equality of the areas of the rectilinear (𝑎𝑏𝑐) and curvilinear (𝑑𝑏𝑐) shear

stress diagrams in Figure 39(d) results in

𝜏𝑚′ 𝑙1

2= ∫ 𝜏𝑥𝑑𝑥

𝑙

0=

𝑊ℎ∙𝑡∙𝜎

𝑡+2𝑊ℎ, (69)

which yields

𝜏𝑚′ = 𝜎 ∙

2𝑊ℎ∙𝑡

𝑙1(𝑡+2𝑊ℎ). (70)

Similarly, can be found

(𝜏𝑚′ −𝜏𝑚)𝑙1

2=

𝜏𝑚

𝑊𝑟3 ∫ 𝑥3𝑑𝑥

𝑊𝑟

0=

𝜏𝑚𝑊𝑟

4, (71)

which yields

𝜏𝑚 = 𝜏𝑚′ 2𝑙1

2𝑙1+𝑊𝑟. (72)

Therefore, substitution of Eq. (70) into Eq. (72) yields

𝜏𝑚 = 𝜎 ∙2𝑊ℎ∙𝑡

𝑙1(𝑡+2𝑊ℎ)∙

2𝑙1

(2𝑙1+𝑊𝑟), (73)

where 𝜎 is an arbitrary value of the hoop stress in this case.

Page 104: Finite Element Modeling of Buried Longitudinally Welded

87

Finally, the SCF can be calculated from

𝑘𝑚.𝑤𝑒𝑙𝑑 =𝜎𝑏.𝑤𝑒𝑙𝑑

𝑃𝑚=

∆𝜎′′′

𝜎=

2𝑊ℎ∙𝑡

𝑙1(𝑡+2𝑊ℎ)×

2𝑙1

(2𝑙1+𝑊𝑟)(0.212 − 0.125

𝑊𝑟1

(𝑡 2⁄ )1 + 0.03𝑊𝑟

2

(𝑡 2⁄ )2), (74)

where 𝑙1 is calculated from

𝑙1 =𝜎

𝜏𝑚′

∙2𝑊ℎ ∙ 𝑡

(𝑡 + 2𝑊ℎ) (75)

and 𝜏𝑚′ , the maximum value of the shear stress in the section under the rectangular reinforcement,

is found from

𝜏𝑚′ = 𝜏𝑥|

𝑥=𝐿

2

𝑚𝑎𝑥 =𝜎

𝐴0𝑠ℎ (𝐵0

𝑊𝑤

2). (76)

The geometrical constant 𝐴0 is obrained as:

𝐴0 =𝑐ℎ(𝐵0

𝑊𝑤2

)−1

𝐵0𝑊ℎ(

𝑡+2𝑊ℎ

𝑡), (77)

and 𝐵0

𝐵0 =2

𝑡√

2𝑡+𝑊ℎ

𝑊ℎ𝑘, (78)

where 𝑘 is the coefficient of deformation of the weld joint

𝑘 = 0.9 (𝑡

𝑡+𝑊ℎ)

2

. (79)

The Fourier hyperbolic functions used in Eq. (76) and Eq. (77) can be calculated from

𝑠ℎ (𝐵0𝑊𝑤

2) =

𝑒𝐵0

𝑊𝑤2 −𝑒

−𝐵0𝑊𝑤

2

2 and 𝑐ℎ (𝐵0

𝑊𝑤

2) =

𝑒𝐵0

𝑊𝑤2 +𝑒

−𝐵0𝑊𝑤

2

2 [84]. (80)

The value of 𝑙1 is independent of the applied stress, as it has a purely geometrical meaning.

In the present research, the values of 𝑊𝑤 and 𝑊ℎ are kept constant, as they have been found

to vary insignificantly with WT during the calculation of an average weld profile. As the focus of

Page 105: Finite Element Modeling of Buried Longitudinally Welded

88

the research is on the potential reduction of WT for a pipe with a particular OD, the only variable

is WT for the ODs studied. Different relationships between SCFs, WT, and 𝑊𝑟 are represented in

Figure 40. The SCF decreases with WT when 𝑊𝑟 is computed according to [25] as 𝑊𝑟 = 0.25𝑡, as

shown in Figure 40(a), due to the increase of calculated 𝑊𝑟 with WT. This relationship is used

when the weld undercut is removed and 𝑅 is introduced to improve fatigue properties of the weld

joint when fracture mechanics is considered. The SCF has a parabolic distribution when the 𝑅 is

kept constant, as shown in Figure 40(b), and is lowest for lower WTs.

It would be ideal to have a set of 𝑊𝑟 that allows for the lowest gradient of SCF across the

studied WTs, and the only solution found in this study is shown in Figure 40(c), when 𝑊𝑟 = 3 −7

𝑡.

However, the calculated SCFs have high magnitudes, which would require larger design

parameters for a pipeline, such as OD and/or WT. The value of 𝑊𝑟 = 5 𝑚𝑚 is closer to the

untreated SAW joints and shows a 25 % lower gradient of SCF for WTs of 8.74 mm through

28.58 mm compared to previously computed SCFs. Furthermore, there is no additional

information on the value of 𝑊𝑟 in North American standards [21] [22]. Therefore, for convenience,

the radius 𝑊𝑟 will be kept as equal to 5 mm for all further calculations and simulations in the

present work.

The weld due to SAW processing of the UOE pipe magnifies the hoop stress, resulting in

𝑘𝑚.𝑤𝑒𝑙𝑑 = 0.1524 (Eq. (74)) and the design value 𝑆ℎ.𝑑𝑒𝑠𝑖𝑔𝑛 = 261 × (1 + 0.1524) = 301 𝑀𝑃𝑎.

Page 106: Finite Element Modeling of Buried Longitudinally Welded

89

Str

ess

Conce

ntr

atio

n F

acto

r k m

.wel

d

Wel

d R

oot

Rad

ius

Wr,

mm

(a) (b)

(c) (d) Wall Thickness t, mm

Figure 40 The (dots) SCFs for different (connected dots) transition radiuses Wr (a)

Wr =0.25WT, (b) Wr =7.145 mm, (c) Wr=3-7/t, and (d) Wr=5 mm

By considering the signs of each modification factor with respect to the location of a critical

spot, as shown in Figure 41, the stresses that may be observed due to different combinations of

discussed effects are presented in Table 19.

2

3

4

5

6

7

8

1.00

1.02

1.04

1.06

1.08

1.10

1.12

1.14

1.16

1.18

1.20

1.22

8 10 12 14 16 18 20 22 24 26 28 30

2

3

4

5

6

7

8

1.00

1.02

1.04

1.06

1.08

1.10

1.12

1.14

1.16

1.18

1.20

1.22

8 10 12 14 16 18 20 22 24 26 28 30

2

3

4

5

6

7

8

1.00

1.02

1.04

1.06

1.08

1.10

1.12

1.14

1.16

1.18

1.20

1.22

8 10 12 14 16 18 20 22 24 26 28 30

2

3

4

5

6

7

8

1.00

1.02

1.04

1.06

1.08

1.10

1.12

1.14

1.16

1.18

1.20

1.22

8 10 12 14 16 18 20 22 24 26 28 30

Page 107: Finite Element Modeling of Buried Longitudinally Welded

90

(a) (b)

(c) (d)

Figure 41 Secondary bending (curved arrows) due to: (a) axial, (b) angular, and (c) ovality

misalignments, and (d) due to soil; the red dashed line indicates the plane of a hypothetical

crack or SCL, and 1 through 4 are the hot-spot locations

For example, the angle (𝑎𝑏𝑐) tend to increase to (𝑎′𝑏′𝑐′) and to open the hypothetical crack

at hot-spot location 3 when load 𝑃 is applied (due to 𝑆ℎ.𝑛𝑜𝑚𝑖𝑛𝑎𝑙), as shown in Figure 41(a); a similar

situation can be observed at location 1; however, hypothetical crack tend to close at locations 2

and 4, resulting in subtraction of the magnification factor 𝑘𝑚.𝑎𝑥𝑖𝑎𝑙 from the total 𝑘𝑚, while at

locations 1 and 3 the factor 𝑘𝑚.𝑎𝑥𝑖𝑎𝑙 is added to 𝑘𝑚. The same assessment was applied to other

cases, as shown in Figure 41(b) through Figure 41(d).

In Figure 41(b) the load 𝑃 is applied in the direction of the welded plates at the junction

between the true pipe circle (dashed curved line) and the misaligned part (solid line), also

represented by points 𝑎 and 𝑑; with secondary bending being experienced at all points, (𝑎𝑏𝑐𝑑).

Therefore, hypothetical cracks open at hot-spot locations 1 and 4 and close at locations 2 and 3

when internal pressure is applied. There are two locations of the weld assessed in Figure 41(c) and

a b

c d a' b'

c' d'

1

2

3

4

P

P

P P

P

P

Page 108: Finite Element Modeling of Buried Longitudinally Welded

91

Figure 41(d), i.e., the upper location and the side location. The analysis of possible scenarios for

calculation of 𝑘𝑚 is summarized in Table 19 and indicates that the part of the weld located at the

inner surface of a pipe at location 1 experiences the largest stresses, while the lowest stress was

obtained at location 2 for the upper weld location (see Figure 41(a) for spot locations/numbers).

Table 19 Stress magnification at different weld locations

Spot 𝑆ℎ.𝑛𝑜𝑚

[MPa]

𝑘𝑚.𝑎𝑥

0.2571

𝑘𝑚.𝑎𝑛

0.5981

𝑘𝑚.𝑜𝑣𝑎𝑙

0.0000

𝑘𝑚.𝑠𝑜𝑖𝑙

0.0620

1+∑𝑘𝑚

1.9172

𝑆ℎ.𝑑𝑒𝑠𝑖𝑔𝑛

[MPa]

𝑘𝑚.𝑤𝑒𝑙𝑑

0.0689

𝑆ℎ.𝑑𝑒𝑠𝑖𝑔𝑛𝑚.𝑤𝑒𝑙𝑑

[MPa]

Upper weld location ( and )

1 261 + + - + 1.9172 501 + 519

2 261 - - + - 0.0828 22 + 40

3 261 + - + - 0.5970 156 + 174

4 261 - + - + 1.4030 366 + 384

Side weld location ( )) and )) )

)) 1 261 + + + - 1.7932 468 + 486

)) 2 261 - - - + 0.2068 54 + 72

)) 3 261 + - - + 0.7210 188 + 206

)) 4 261 - + + - 1.2790 334 + 352

The stress developed at location 1 in a non-misaligned pipe is calculated to be equal to

322 MPa. An upper weld location was selected for development of an ABAQUS-based FEM with

a soil box, including axial and angular misalignments.

The Analytical (Hand) prediction of stresses at the hot-spot based on the analytical model

described in this section shows design hoop stresses, 𝑆ℎ.𝑑𝑒𝑠𝑖𝑔𝑛, similar to those obtained by using

Page 109: Finite Element Modeling of Buried Longitudinally Welded

92

an FEM approach for both misaligned and the non-misaligned pipeline conditions Table 20; the

result of Hand calculations of stresses seem to be conservative, showing (2 ÷ 20)% larger stresses

than those predicted in ABAQUS. The result of Hand calculations, shown in Table 20, would

represent an actual conservatism that can be involved in design-by-rule governed by the design

codes when compared to the results of a more detail FE modeling. Notably, the stresses calculated

using elastic and elastic-plastic FE analysis, shown in Table 20, are almost the same since the

linearized stresses have been used in both cases.

It can be concluded that the Hand calculations discussed in this work provide pipeline

design engineers with a comprehensive tool for detailed analyses of critical hot-spot stresses

observed at the plate-weld transition due to manufacturing defects as well as in-service conditions,

such as pressure from the transported medium and soil. The advantage of using the FEM approach

is in obtaining the complete set of stress components, including 𝜎𝑚, 𝜎𝑏, and 𝜎𝑝, needed for

advanced fatigue assessment.

Table 20 Results of analysis of the design hoop stresses Sh [MPa] for a pipe of OD 914 mm

and WT 17.5 mm

Misalignment Hot-spot

Location

Method

Hand:

Upper Weld

Hand:

Side Weld

FEM:

Elastic

FEM:

Elastic-Plastic

NO 1 261 261 250 250

YES 1 501 468 431 432

YES 2 22 54 69 69

YES 3 156 188 153 155

YES 4 366 334 345 345

Page 110: Finite Element Modeling of Buried Longitudinally Welded

93

The discrepancy between the Hand and ABAQUS calculations is likely due to differences

between the exact geometry of the pipe section modeled with FEM and the geometry used to derive

the closed-form solutions for each individual stress magnification factor, 𝑘𝑚 (Section 3.7). The

Hand calculations are summarized in Figure 42.

Stre

ss S

h.d

esig

n, M

Pa

Wall Thickness t, mm

Figure 42 Hoop stress calculated with mathematical model, power-law-fitted, and

extrapolated until solutions of (dashed line) non-misaligned and (solid line) misaligned

conditions intersect (power-law-fitted)

68, 46

66, 7165, 7763, 109

322

15.20 24.060

100

200

300

400

500

600

700

800

900

1,000

1,100

1,200

1,300

1,400

1,500

1,600

1,700

9 13 17 21 25 29 33 37 41 45 49 53 57 61 65 69

OD 610 mm

OD 610 mm - Misaligned

OD 864 mm

OD 864 mm - Misaligned

OD 914 mm

OD 914 mm - Misaligned

OD 1219 mm

OD 1219 mm - Misaligned

Intersection

Page 111: Finite Element Modeling of Buried Longitudinally Welded

94

The solutions for the 𝑆ℎ.𝑑𝑒𝑠𝑖𝑔𝑛 of misaligned pipes of selected ODs (Figure 42) significantly

diverge from the non-misaligned solutions with decreasing WT, 𝑡, because the through-thickness

bending, 𝜎𝑏, becomes the dominant component of total stress 𝑆ℎ.𝑑𝑒𝑠𝑖𝑔𝑛. The two conditions for

each pipe have a common solution at certain value of 𝑡 and corresponding 𝑆ℎ.𝑑𝑒𝑠𝑖𝑔𝑛, and drift

toward lower 𝑡 and larger 𝑆ℎ.𝑑𝑒𝑠𝑖𝑔𝑛 with increasing OD exponentially (Figure 42). A significant

increase in WT of a pipe of OD = 914 mm due to misalignment can be observed (from 15.20 mm

to 24.06 mm) at the cut-off value of allowable stress for the selected material 𝑆 = 0.9𝜎𝑦 = 0.9 ∙

359[𝑀𝑃𝑎] = 322 𝑀𝑃𝑎. The slope of the solution for a misaligned pipe is larger than that for a

pipe without misalignment, as shown in Figure 42. The slope increases with pipe OD. The slopes

were studied in detail for pipe of OD 914 mm and compared to FE results, Figure 43.

Stre

ss S

h.d

esig

n, M

Pa

Wall Thickness t, mm

(a) (b) Figure 43 Solutions for Hoop stress (linear fit) in (non)misaligned pipe of OD 914 mm (a)

without and (b) with km.weld accounted

24.98, 148.00

25.24, 141.00

322

15.07 21.37

14.10

20.470

100

200

300

400

500

600

700

800

14 18 22 26

In-Hand - not misaligned

In-Hand - misaligned

Abaqus - not misaligned

Abaqus - misaligned

24.98, 169.00

25.24, 141.00

322

17.37 21.9714.10

20.47

0

100

200

300

400

500

600

700

800

14 18 22 26

Page 112: Finite Element Modeling of Buried Longitudinally Welded

95

Two versions of the calculation of the 𝑆ℎ.𝑑𝑒𝑠𝑖𝑔𝑛 are presented in Figure 43; the one with

𝑘𝑚.𝑤𝑒𝑙𝑑 not included in total 𝑘𝑚 (the same as in the standards), Figure 43(a), and the other, with

𝑘𝑚.𝑤𝑒𝑙𝑑 included in total 𝑘𝑚 (as shown in this research), Figure 43(b) (see 𝑆ℎ.𝑑𝑒𝑠𝑖𝑔𝑛𝑚.𝑤𝑒𝑙𝑑 in Table 19).

There is no obvious benefit of accounting for the stress magnification due to weld profile, 𝑘𝑚.𝑤𝑒𝑙𝑑,

in the Hand calculations, as the total stress is simply increases by the 𝑘𝑚.𝑤𝑒𝑙𝑑 factor and the

predictions just become more conservative.

However, when the power law is used to fit the data (i.e., the coefficient of determination

becomes 𝑅2 = 1, resulting in a 100% fit), the use of the 𝑘𝑚.𝑤𝑒𝑙𝑑 factor may be justified, as shown

in Figure 44. The difference in the estimated 𝑡 between the non-misaligned and misaligned

geometries has a factor of 2.52 = (23.72 − 21.68) (15.04 − 14.23)⁄ when 𝑘𝑚.𝑤𝑒𝑙𝑑 is not

considered, as shown in Figure 44(a); however, this factor is more than two times lower, at only

1.18 = (25.22 − 21.68) (17.22 − 14.23)⁄ , when 𝑘𝑚.𝑤𝑒𝑙𝑑 is considered, as shown in Figure

44(b). Therefore, a common factor (or increment of 𝑡) for both the non-misaligned and the

misaligned geometries of a pipe with a particular OD can be used to make a correlation (between

Hand and FEM solutions) based on either method, whether elastic or elastic-plastic stress-strain

analysis. This means that both the non-misaligned and the misaligned geometries can be adjusted

with almost same increment of 𝑡, with discrepancy of only 0.55 𝑚𝑚 = (25.22 − 21.68) −

(17.22 − 14.23), as shown in Figure 44(b), compared to more than double of it, 1.23 =

(23.72 − 21.68) − (15.04 − 14.23), as shown in Figure 44(a), when 𝑘𝑚.𝑤𝑒𝑙𝑑 is not considered.

Moreover, dissonance was found to increase with reduced WT and increased OD, as shown in

Figure 42, which means that the proposed addition to the total 𝑘𝑚 in the form of the 𝑘𝑚.𝑤𝑒𝑙𝑑 factor

may benefit the large-diameter pipelines even more (than small-diameter pipelines).

Page 113: Finite Element Modeling of Buried Longitudinally Welded

96

Stre

ss S

h.d

esig

n, M

Pa

(a)

(b)

Wall Thickness t, mm Figure 44 Solutions for Hoop stress (power law fit) in pipe of OD 914 mm (a) without km.weld

and (b) with km.weld

Therefore, the use of the 𝑘𝑚.𝑤𝑒𝑙𝑑 factor can save on computational effort and the associated

costs when a more precise data-to-curve fit is used (Figure 44). It enables the engineer to estimate

the conservatism of the analytical model in WT-equivalent almost independent of misalignment.

62.09, 79.50

60.96, 80.20

322

15.04

23.72

14.23

21.68

25.14, 144.75

0

100

200

300

400

500

600

700

800

14 18 22 26 30 34 38 42 46 50 54 58 62

Hand - not misaligned

Hand - misaligned

Abaqus - not misaligned

Abaqus - misaligned

60.25, 93.50

60.96, 80.20

322

17.22

25.22

14.23 21.68

24.89, 153.98

0

100

200

300

400

500

600

700

800

14 18 22 26 30 34 38 42 46 50 54 58 62

Page 114: Finite Element Modeling of Buried Longitudinally Welded

97

Thus, only one series of FE analyses would be needed, for either non-misaligned or misaligned

pipeline model for the range of WTs, to determine the design WT.

5.2. Fatigue Analysis

The results of the fatigue analysis represent the relationships between WTs and FDs

computed for four different ODs as depicted in Figure 45. An example of the FD plots for the pipe

with a non-misaligned weld also included for comparison in Figure 45(c). The obtained data points

collected through the analysis are included in the data-to-curve-fit based on the two constants

power law, which was found to approximate the obtained data accurately, as shown in Table 21(I).

The coefficient of determination, 𝑅2 (Table 21(I)), was calculated to be in the range between 0.97

and 1.00, indicating a 97 to 100 % goodness of data fit.

Although the elastic solutions for the FD obtained in accordance with BS and ASME codes

are very close, as shown in Figure 45(a-d), a significant difference in allowable WT was observed

between the elastic and elastic-plastic solutions, signifying a lower fatigue damage of assessed

pipe design when elastic-plastic analysis considered. The fatigue analysis solutions become

unstable at FDs larger than unity, since decreasing the WT results in an increase of nominal stresses

and through-thickness plasticity can be developed, the stress-strain state known as general

yielding. Therefore, further analysis proceeded with a cut-off value for the maximum acceptable

accumulated FD of 0.5 in order to avoid the influence of excessive plasticity.

Interestingly, the standard design WT obtained from the analysis based on the elastic data,

(Table 21(I)) was found to be exactly the same as the ones predicted using [97] and [94] according

to ASME 31.4 [23]; i.e., WT of 11.9 mm for OD 610 mm, WT of 17.5 mm for OD 864 mm, WT

of 19.1 mm for OD 914 mm, and WT of 23.8 mm for OD 1219 mm. It is worth noting that the

Page 115: Finite Element Modeling of Buried Longitudinally Welded

98

actual values may differ depending on actual pressure history and weld misalignment; however,

the trend should not change.

Another important observation noted during analysis of the linearized stresses in Figure 38

was that the fatigue life of a pipeline is significantly reduced with misalignment due to secondary

bending developed at hot-spot. For instance, depending on the assessment method, a pipeline with

an OD of 914 mm can be designed with 23.30% to 44.41% lower actual WT (or with 25.00% to

45.88% lower standard WT) than that of a misaligned pipeline, as shown in the data represented

by full squares in Figure 45(c) and calculations in Table 21(I). Therefore, for the fixed WT and

pressure history, a non-misaligned pipeline can be designed with significantly lower OD

(approximately 610 mm), instead of 914 mm when the pipeline is misaligned. It can be seen from

Figure 45(a) and Figure 45(c). However, an approximately 3-6% reduction in 𝜎𝑦 can be expected

for UOE-manufactured pipes with larger OD/WT ratios, i.e., with lower WTs, based on the

experimental and predicted data discussed in [17]; this would require recalculation of cyclic stress-

strain curves used in ABAQUS simulations for more accurate estimates of FD.

As far as a misaligned pipeline is concerned, a significant WT reduction without

compromising the fatigue performance of a pipeline (fatigue damage value remains 0.5) can be

achieved when the elastic-plastic analysis performed, i.e., between 7.59% and 16.46% when actual

WT considered, or between 7.57% and 19.98% when standard WT considered (Table 21(I)). The

area of envelope between elastic and elastic-plastic solutions in Figure 45(c) increases significantly

with misalignment; this indicates an increase in design conservatism with increased misalignment

when elastic analysis performed. Moreover, conservatism increases for every OD studied at FDs

lower than 0.5, i.e. at larger WTs.

Page 116: Finite Element Modeling of Buried Longitudinally Welded

99

Acc

um

ula

ted

Fat

igu

e D

amag

e D

f

(a)

(b)

(c)

(d)

Wall Thickness t, mm

Figure 45 Accumulated fatigue damage plots for pipe diameters (a) 610 mm, (b) 864 mm,

(c) 914 mm, and (d) 1219 mm, calculated with (solid lines) misalignment and with (contour

lines) no misalignment

0.0

0.5

1.0

5 10 15 20 25 30

0.0

0.5

1.0

5 10 15 20 25 30

0.0

0.5

1.0

5 10 15 20 25 30

0.0

0.5

1.0

5 10 15 20 25 30

BS-e (elastic analysis according to [17])

▪ ASME-e/p (elastic-plastic analysis according to [16])

ASME-e (elastic analysis according to [16])

Page 117: Finite Element Modeling of Buried Longitudinally Welded

100

Table 21 Results of fatigue analysis obtained at accumulated fatigue damage of 0.5

Method

Data-to-Curve Fit Actual Standard

𝑅2

Adj. 𝑅2

1st const. a

Power b

2nd const. c

WT* [mm]

ΔWT [mm]

ΔWT [%]

WT** [mm]

ΔWT [mm]

ΔWT [%]

I. Power Law Data-to-Curve Fit, 𝑭𝑫 = 𝒂 × 𝑾𝑻𝒃 + 𝒄

OD 610 mm

A (A – B) 0.9903 0.9838 198.0 × 101 -4.601 0.103 10.50 0.41 3.86 11.91 1.60 13.43

B (A – C) 0.9962 0.9937 484.6 × 107 -10.190 0.215 10.09 1.21 11.56 10.31 2.38 19.98

C (B – C) 0.9940 0.9900 264.3 × 1010 -13.270 0.119 9.28 0.81 8.01 9.53 0.78 7.57

OD 864 mm

A (A – B) 0.9999 0.9998 338.3 × 101 -3.066 -0.057 17.14 0.40 2.33 17.48 0.00 0.00

B (A – C) 0.9974 0.9957 325.9 × 106 -7.412 0.223 16.74 2.42 14.11 17.48 1.60 9.15

C (B – C) 0.9987 0.9979 188.3 × 1010 -10.910 0.159 14.72 2.02 12.06 15.88 1.60 9.15

OD 914 mm

A (A – B) 1.0000 1.0000 825.4 × 101 -3.377 -0.019 17.54 0.58 3.30 19.05 1.57 8.24

B (A – C) 0.9985 0.9977 105.9 × 104 -5.216 0.091 16.96 2.89 16.46 17.48 3.17 16.64

C (B – C) 0.9965 0.9947 797.3 × 104 -6.244 0.082 14.65 2.31 13.61 15.88 1.60 9.15

OD 914 mm – No Misalignment

A (A – B) 1.0000 1.0000 361.0 × 10−1 -1.804 -0.093 9.75 -1.74 -15.12 10.31 -1.60 -13.43

B (A – C) 0.9908 0.9814 565.8 × 102 -4.747 -0.026 11.48 -0.85 -7.99 11.91 -1.60 -13.43

C (B – C) 0.9931 0.9862 393.0 × 105 -7.718 0.017 10.59 0.89 7.75 11.91 0.00 0.00

OD 1219 mm

A (A – B) 0.9972 0.9991 104.0 × 1012 -10.780 0.221 22.48 0.13 0.59 23.83 0.00 0.00

B (A – C) 0.9987 0.9996 312.3 × 1012 -11.130 0.199 22.34 1.83 8.13 23.83 3.21 13.47

C (B – C) 1.0000 1.0000 921.6 × 108 -8.651 0.112 20.65 1.70 7.59 20.62 3.21 13.47

Page 118: Finite Element Modeling of Buried Longitudinally Welded

101

II. Linear Data-to-Curve Fit, 𝑾𝑻 = 𝒂 × 𝑶𝑫 + 𝒄

OD 610 mm

A (A – B) 0.9726 0.9588 0.01941 -0.5919 11.24 0.48 4.26 11.91 1.60 13.43

B (A – C) 0.9787 0.9681 0.01986 -1.3740 10.76 1.82 16.17 10.31 2.38 19.98

C (B – C) 0.9917 0.9876 0.01850 -1.8600 9.42 1.34 12.44 9.53 0.78 7.57

OD 864 mm

A (A – B) 0.9726 0.9588 0.01941 -0.5919 16.17 0.35 2.18 17.48 0.00 0.00

B (A – C) 0.9787 0.9681 0.01986 -1.3740 15.82 2.05 12.65 17.48 1.60 9.15

C (B – C) 0.9917 0.9876 0.01850 -1.8600 14.12 1.69 10.71 15.88 1.60 9.15

OD 914 mm

A (A – B) 0.9726 0.9588 0.01941 -0.5919 17.14 0.33 1.91 17.48 0.00 0.00

B (A – C) 0.9787 0.9681 0.01986 -1.3740 16.81 2.09 12.20 17.48 1.60 9.15

C (B – C) 0.9917 0.9876 0.01850 -1.8600 15.05 1.76 10.49 15.88 1.60 9.15

OD 1219 mm

A (A – B) 0.9726 0.9588 0.01941 -0.5919 23.06 0.17 0.76 23.83 0.00 0.00

B (A – C) 0.9787 0.9681 0.01986 -1.3740 22.88 2.37 10.26 23.83 3.21 13.47

C (B – C) 0.9917 0.9876 0.01850 -1.8600 20.69 2.19 9.58 20.62 3.21 13.47

A – BS elastic analysis; B – ASME elastic analysis; C – ASME elastic-plastic analysis; A – B – result of B subtracted from A; A – C – result of C subtracted from A; B – C – result of C subtracted from B. * calculated at the accumulated fatigue damage value of 0.5. ** next standard value according to ASME B31.10 larger than that obtained as *.

The experimental data were further refined by linearizing the relationship between the ODs

and WTs at a fatigue damage of 0.5, as shown in Figure 46. Linearization of the OD–WT

relationship yielded the equations for the individual fatigue assessment methods, see Table 21(II),

which closely follow the general solution for WT in the from of Eq. (64), with only 5%

Page 119: Finite Element Modeling of Buried Longitudinally Welded

102

discrepancy observed on the studied interval of ODs, as compared to the BS elastic method [24].

One of the contributions of the present study, however, is in discovering the actual difference in

calculated design WT between elastic and elastic-plastic analyses. This difference is at least one

increment of a standard WT. The results of the fatigue analysis suggest that the design of large-

diameter pipelines may benefit from reductions in WT, contributing to significant budget savings

on material when multi-kilometer transmission lines are designed.

Figure 46 Relationship between OD and WT at accumulated fatigue damage of 0.5 for

(blue) BS elastic, (red) ASME elastic, and (grey) ASME elastic-plastic analyses

The WTs were also calculated at FD as equal 1.0, as shown in Table 22, using equations

from Table 21(I). The computed WTs were plotted as OD–WT relationship and linearized in order

to refine the calculations for WT. The results show that savings on WT increased with OD, between

1.06% and 8.57% when the actual WT considered, and between 0.00% and 10.14% when the

standard WT was considered, as presented in Table 22. The trend was similar to that of the data in

Table 21(II) and Table 22 respectively (see bolted data points). In other words, while the ΔWT

increases with OD at FD equal 1.0, at FD equal 0.5 the ΔWT is almost same throughout the ODs.

This demonstrates an increase in the WT savings for larger ODs at FD equal 1.0, while at FD equal

0.5 the WT savings remain relatively unchanged throughout ODs.

9

12

15

18

21

24

600 650 700 750 800 850 900 950 1000 1050 1100 1150 1200 1250

Wal

l Th

ickn

ess

t, m

m

Outside Diameter D, mm

BS-e at 0.5DASME-e at 0.5DASME-e/p at 0.5DLinear (BS-e at 0.5D)Linear (ASME-e at 0.5D)Linear (ASME-e/p at 0.5D)

Page 120: Finite Element Modeling of Buried Longitudinally Welded

103

Table 22 Results of fatigue analysis obtained at accumulated fatigue damage of 1.0

Method

Data-to-Curve Fit Actual Standard

𝑅2

Adj. 𝑅2

1st const. a

Power b

2nd const. c

WT* [mm]

ΔWT [mm]

ΔWT [%]

WT** [mm]

ΔWT [mm]

ΔWT [%]

Linear Data-to-Curve Fit, 𝑾𝑻 = 𝒂 × 𝑶𝑫 + 𝒄

OD 610 mm

A (A – B) 0.9982 0.9973 0.01905 -2.802 8.82 -0.48 -5.17 9.53 0.00 0.00

B (A – C) 0.9923 0.9885 0.01844 -1.949 9.30 0.09 1.06 9.53 0.79 8.29

C (B – C) 0.9791 0.9687 0.01633 -1.236 8.73 0.57 6.17 8.74 0.79 8.29

OD 864 mm

A (A – B) 0.9982 0.9973 0.01905 -2.802 13.66 -0.33 -2.33 14.27 0.00 0.00

B (A – C) 0.9923 0.9885 0.01844 -1.949 13.98 0.78 5.74 14.27 0.00 0.00

C (B – C) 0.9791 0.9687 0.01633 -1.236 12.87 1.11 7.94 14.27 0.00 0.00

OD 914 mm

A (A – B) 0.9982 0.9973 0.01905 -2.802 14.61 -0.30 -1.98 15.88 0.00 0.00

B (A – C) 0.9923 0.9885 0.01844 -1.949 14.91 0.92 6.30 15.88 1.61 10.14

C (B – C) 0.9791 0.9687 0.01633 -1.236 13.69 1.22 8.16 14.27 1.61 10.14

OD 1219 mm

A (A – B) 0.9982 0.9973 0.01905 -2.802 20.42 -0.11 -0.53 20.62 0.00 0.00

B (A – C) 0.9923 0.9885 0.01844 -1.949 20.53 1.75 8.57 20.62 1.57 7.61

C (B – C) 0.9791 0.9687 0.01633 -1.236 18.67 1.86 9.06 19.05 1.57 7.61

A – BS elastic analysis; B – ASME elastic analysis; C – ASME elastic-plastic analysis; A – B – result of B subtracted from A; A – C – result of C subtracted from A; B – C – result of C subtracted from B. * calculated at the accumulated fatigue damage value of 0.5. ** next standard value according to ASME B31.10 larger than that obtained as *.

Page 121: Finite Element Modeling of Buried Longitudinally Welded

104

It can be seen from Figure 45 that the solutions for WT are almost the same for all three

methods at FD equal to 1.0; the elastic solutions differ more from elastic-plastic solution with

increasing OD. Although this trend is similar in the case of WTs [mm] obtained at an FD of 0.5,

the ΔWT [%] decreases due to changes in the slope ΔWT/OD from

(2.37 − 1.82) (1219 − 610)⁄ = 9.0 × 10−4 to (1.75 − 0.09) (1219 − 610)⁄ = 2.7 × 10−3, as

can be seen from Table 21(II) and Table 22.

Additionally, when the number of cycles to failure in the Elastic-Plastic Fatigue Analysis

method (3.8.4) is calculated according to the equation used in the Elastic Fatigue Analysis of

Welds (3.8.5), Eq. (53), ASME BPVC Section VIII Part 2 [22], the solutions from both methods

were found to be very close, signifying a good agreement with the Neuber approximation.

In summary, larger pipeline material savings can be expected for lines designed with larger

ODs at accumulated FD close to 100%, at accumulated FD close to 50% pipeline designs with

lower ODs can also benefit from the material savings. For example, if the cost of a UOE

manufactured pipeline is in the range of $ (400 ÷ 1000) 𝑡𝑜𝑛⁄ , considering an approximately

4,700 km long pipeline, as in the case of the Keystone project [53], a 10 % reduction in WT

according to Table 22 would result in $(0.07 ÷ 0.17)𝑚𝑙𝑟𝑑 of total savings, as shown in Table 23.

This is a significant amount at the total estimated pipeline cost of $(5 ÷ 8) 𝑚𝑙𝑟𝑑.

Table 23 Construction cost savings associated with WT reduction on a 4700 km pipeline

Method used Standard pipe dimensions Steel

Density

[𝑔 𝑐𝑚3⁄ ]

Pipeline

Weight

[ton]

Pipeline

Cost

[$ 𝑡𝑜𝑛⁄ ]

Budget

Savings

[𝑚𝑙𝑟𝑑$]

OD

[mm]

WT at FD=1

[mm]

Elastic 914 15.88 7.85 1488173 1000

Elastic-Plastic 914 14.27 7.85 1653112 1000 0.17

Page 122: Finite Element Modeling of Buried Longitudinally Welded

105

Chapter 6 CONCLUSIONS AND FUTURE WORK

The present work discusses standard fatigue analyses as well as difficulties associated with

the development of the FE model based on standard procedures used in pipeline design. The weld

misalignment and the weld profile both significantly influenced the stresses. Methods of

accounting for those effects analytically and using the FEM are proposed. This included revision

of analytical calculation for the hot-spot stresses and optimization of the FE model due to a

complex interaction between soil, pipe, and profile of the misaligned weld.

6.1. Conclusions

A mathematical representation of an average SAW weld profile was obtained based on 30

SAW welds from the literature and used in the FE model for a more realistic representation of

pipeline geometry.

Analytical models available in standard procedures were reviewed. Proposed the detailed

yet simple methodology for calculating the stresses at all four weld toes of UOE-manufactured

pipe based on analysis of bending stresses at each weld toe. The results of this method were close

to the results of the FE analysis and helped explain the variations in calculated hot-spot stresses.

A discussion on modification of the analytical model for the calculation of SCFs based on

actual weld profile is provided. The use of actual weld profile results in a quick determination of

WT design when a more precise data-to-curve fit is considered; the set of FE models with different

WTs may be built for either non-misaligned or misaligned pipe.

The approximation of the tensile curves used in the elastic-plastic analysis in ABAQUS

was addressed. Specifically, the discrepancy between the real and approximated tensile-curves

near the yield point of higher-strength steels was reduced by modifying the existing standardized

Page 123: Finite Element Modeling of Buried Longitudinally Welded

106

model for a true stress-strain curve, which generation via numerical solution was automated in

MATLAB.

A methodology for modeling (combined) weld misalignments in UOE pipes has been

proposed. A detailed pipe model included weld profile, a combination of the axial and angular

weld misalignments, surrounding soil, thermal load and internal pressure. An extensive

convergence study is presented.

It has been shown that the weld misalignment of a longitudinally welded pipeline due to

UOE manufacturing results in significant stress rise at the structural hot-spot. Consequently, it is

detrimental to in-service fatigue performance. For example, at FD equal 0.5, the non-misaligned

pipeline of OD 914 mm witnessed a 23.30% increase of actual WT due to maximum allowable

misalignment when ASME elastic-plastic fatigue analysis considered. A 27.70% increase of WT

was observed when ASME elastic fatigue analysis considered. After the actual WT was adjusted

to the next nearest standard WT, the increase in WT was found to be 25.00% or 31.86% based on

ASME elastic-plastic or ASME elastic fatigue analysis respectively. A similar trend can be

observed at FD equal to 1.0.

The difference was discovered between elastic and elastic-plastic solutions also for FD of

a pipeline due to in-service pressure fluctuations. In the range of studied ODs, the percent savings

on a WT lies in the range between 9.58% and 12.44% for actual WT and between 7.57% and

13.47% for standard WT, at FD equal to 0.5; at FD equal to 1.0, the percent savings on a WT was

between 6.17% and 9.06% for actual WT and between 0.00% and 10.14% for standard WT.

Reported percent values are proportional to the range of 0 to 2 increments of standard WTs and

directly proportional to the weight reduction of a pipe (i.e. associated cost savings).

Page 124: Finite Element Modeling of Buried Longitudinally Welded

107

The result of fatigue analysis suggests that the design of large-diameter pipelines may

benefit from reduction of WT and make significant budget savings on material when the multi-

kilometre transmission lines designed for a 100% FD; at lower values of accumulated FD, 50%,

the pipeline designs with lower ODs can also benefit from the material savings.

The results of this study highlighted the WT reduction capabilities of the elastic-plastic

fatigue analysis compared to conservative estimates based on elastic fatigue analysis done on

large-diameter oil pipelines.

6.2. Future Work

Although, the results of this work have been obtained with the use of construction material

free from metallurgical defects, welding is also known to produce residual stresses and

microstructural defects. The weld microstructure is usually heterogeneous and may contain micro-

cracks or voids at the weld toe surface (undercut) or under the surface. Furthermore, welding and

other UOE pipe manufacturing processes produce residual stresses, some of which, specifically

tensile residual stresses, are known as significant contributors to fatigue crack propagation.

Additionally, pipe expanding and hydrotest homogenize the residual stress distribution and

introduce compressive tensile stresses. Therefore, fatigue life of a welded pipeline should be

further analyzed by using a fracture mechanics approach ASME BPVC Section VIII Part 2 [22],

[37], BS 7608 [24], and BS 7910 [25] to support the WT reduction capabilities when the more

advanced methods of fatigue assessment are used.

Page 125: Finite Element Modeling of Buried Longitudinally Welded

108

APPENDIX A – MATLAB Numerical Solution

Merged Stress-Strain Curve

% Level of refinement for the solution of a tangent line refine=1000; % Generation of the Stress coordinates varSTmax=zeros(refine+1,1); for i=1:refine+1; S1=(varSTmax(i,1)+(SysTmax-

50)+((Sutst-(SysTmax-50))/refine)*(i-1)); varSTmax(i)=S1; end % varE1=S/Ey+(S/Kcss)^(1/ncss)-Sys/Ey; varE1Tmax=zeros(refine+1,1); for i=1:refine+1;

e1=(varSTmax(i,1)/EyTmax)+(varSTmax(i,1)/KcssBaseTmax)^(1/ncssBaseTmax)-

SysTmax/EyTmax; varE1Tmax(i)=e1; end % varE2=((((Kcss^(-1/ncss))*S^(1/ncss-1))/ncss)+1/Ey)*(S-Sys); varE2Tmax=zeros(refine+1,1); for i=1:refine+1; e2=((((KcssBaseTmax^(-

1/ncssBaseTmax))*varSTmax(i,1)^(1/ncssBaseTmax-

1))/ncssBaseTmax)+1/EyTmax)*(varSTmax(i,1)-SysTmax); varE2Tmax(i)=e2; end % Minimum difference between solutions deltaEzeroTmax=abs(varE2Tmax-varE1Tmax); % Index for the point corresponding to minimum difference MinErrorIndexTmax=find(deltaEzeroTmax == min(deltaEzeroTmax(:))); % Coordinates for the tangent point of Ramberg-Osgood curve from index StressMEITmax=varSTmax(MinErrorIndexTmax,1); StrainMETmax=varE1Tmax(MinErrorIndexTmax,1); % Ramberg-Osgood portion of a merged curve CCStep=20; varScyclicTmax=zeros(CCStep+1,1); for i=1:CCStep+1;

S1cyclic=StressMEITmax+(varScyclicTmax(i,1)+((Sutst-

StressMEITmax)/CCStep)*(i-1)); varScyclicTmax(i)=S1cyclic; end varEcyclicTmax=zeros(CCStep+1,1); for i=1:CCStep+1;

e2cyclic=(varScyclicTmax(i,1)/EyTmax)+(varScyclicTmax(i,1)/KcssBaseTmax)^(1/n

cssBaseTmax); varEcyclicTmax(i)=e2cyclic; end % Elastic portion of a merged curve StressSCurveElasticTmax=transpose([0 SysTmax]); SStrainCurveElasticTmax=transpose([0 SysTmax/EyTmax]); % Tangent portion of a merged curve varStangentTmax=zeros(CCStep+1,1); for i=1:CCStep+1;

S1tangent=SysTmax+(varStangentTmax(i,1)+((Sutst-SysTmax)/CCStep)*(i-1));

varStangentTmax(i)=S1tangent; end varEtangentTmax=zeros(CCStep+1,1); for i=1:CCStep+1;

e2tangent=(StressMEITmax/EyTmax+(StressMEITmax/KcssBaseTmax)^(1/ncssBaseTmax)

)+(((KcssBaseTmax^(-1/ncssBaseTmax))*((StressMEITmax)^(1/ncssBaseTmax-

1)))/ncssBaseTmax+1/EyTmax)*(varStangentTmax(i,1)-StressMEITmax);

varEtangentTmax(i)=e2tangent; end Tangent=[varStangentTmax,varEtangentTmax]; % Merged Curve at Service Temperature StressSCurveBaseTmax=[StressSCurveElasticTmax;varScyclicTmax]; SStrainCurveBaseTmax=[SStrainCurveElasticTmax;varEcyclicTmax]; MergedBaseTmax=[StressSCurveBaseTmax,SStrainCurveBaseTmax]; xlswrite('BaseTangentSolutionTserv.xlsx',varSTmax,1); xlswrite('BaseTangentSolutionTserv.xlsx',varE1Tmax,2); xlswrite('BaseTangentSolutionTserv.xlsx',varE2Tmax,3); xlswrite('BaseTangentSolutionTserv.xlsx',StressMEITmax,4); xlswrite('BaseTangentSolutionTserv.xlsx',StrainMETmax,5);

Page 126: Finite Element Modeling of Buried Longitudinally Welded

109

Neuber’s Rule

% Level of refinement for the solution refine=100000; % Generation of the Stress coordinates varSTmax=zeros(refine+1,1); for i=1:refine+1;

S1=(varSTmax(i,1)+((Sutst+100)/refine)*(i-1)); varSTmax(i)=S1; end % varE1=S/Ey+(S/Kcss)^(1/ncss)-Sys/Ey; varEk1Tmax=zeros(refine+1,1); for i=1:refine+1;

e1=deltaSek*deltaEek/(varSTmax(i,1)); varEk1Tmax(i)=e1; end % varE2=((((Kcss^(-1/ncss))*S^(1/ncss-1))/ncss)+1/Ey)*(S-Sys); varEk2Tmax=zeros(refine+1,1); for i=1:refine+1;

e2=(varSTmax(i,1)/EyTmax)+2*(varSTmax(i,1)/(2*KcssBaseTmax))^(1/ncssBaseTmax)

; varEk2Tmax(i)=e2; end % Minimum difference between solutions deltaEzeroTmax=abs(varEk2Tmax-varEk1Tmax); % Index for the point corresponding to minimum difference MinErrorIndexTmax=find(deltaEzeroTmax == min(deltaEzeroTmax(:))); % Coordinates for the tangent point of Ramberg-Osgood curve from index StressMEITmax=varSTmax(MinErrorIndexTmax,1); StrainMETmax=varEk1Tmax(MinErrorIndexTmax,1); figure; plot(varEk1Tmax,varSTmax); hold on; plot(varEk2Tmax,varSTmax); hold on; plot(StrainMETmax,StressMEITmax,'b*');% hold off; suptitle('Solution for the Neubers Rule'); ylabel('Stress S, MPa'); xlabel('Strain e, mm/mm'); legend('Left-hand equation (Neubers Rule)','Right-hand equation (Hysteresis

Curve)','Solution (Intersection)'); set(gcf,'PaperPositionMode','auto','PaperPosition',[0 0 20 10]); deltaSk = StressMEITmax; deltaEk = StrainMETmax;

Page 127: Finite Element Modeling of Buried Longitudinally Welded

110

APPENDIX B – MATLAB Cycle-Counting

clear; clc Input = inputdlg({'Agressive Loading History','Average Loading

History','Accuracy - Bin Size'},'Input Data for Rainflow Cycle Counting',[1

40; 1 40; 1

40],{'RainflowAgressiveLoading.xlsx';'RainflowAverageLoading.xlsx';'0.5'}); FileNameAgr = char([Input(1,1)]); FileNameAve = char([Input(2,1)]); AccBinSize=str2double(Input(3,1)); %% Cycles Counted for Min/Max of Agressive Loading History (Larger Bin) s = xlsread(FileNameAgr); tp=sig2ext(s); rf=rainflow(tp); t = transpose(rf); Pamp=t(1:end,1); Pmean=t(1:end,2); Count=t(1:end,3); Min=Pmean-Pamp; Max=Pmean+Pamp; Delta=Max-Min; Cycle=[Min Max]; % Rounding to desired accuracy acc = AccBinSize; Cycle=round(Cycle/acc)*acc; Cycles=[Cycle Count];

% Construction of a plot Cycles=[Min Max Count]; Transp = transpose(Cycles); rfm=rfmatrix(Transp,20,20); % Surface Bar Plot with Color Bar figure suptitle('Rainflow Cycle Counting - Min/Max') subplot(2,2,1); b = bar3(rfm); title('Agressive Loading History (Larger Bin)') zlabel('Number of Cycles') ylabel('Cycle Max') xlabel('Cycle Min') colorbar for k = 1:length(b); zdata = b(k).ZData; b(k).CData = zdata; end % Remove Zeros for i = 1:numel(b) index = logical(kron(rfm(:, i) == 0, ones(6, 1))); zData = get(b(i), 'ZData'); zData(index, :) = nan; set(b(i), 'ZData', zData); end view(45,30)

set(gcf,'PaperPositionMode','auto','PaperPosition',[0 0 20 10]); print(gcf,'Rainflow Cycle Counting Min-Max.png','-dpng','-r300');

Page 128: Finite Element Modeling of Buried Longitudinally Welded

111

% Finding the unique combinations of Min&Max U = unique(Cycles(:,[1 2 3]), 'rows'); % Extracting the unique combinations of Min&Max Size=size(U,1); NofT=zeros(Size,1); for i=1:Size; R = [U(i,1) U(i,2) U(i,3)]; % Counting the number of occurences of a unique combination if U(i,3)<=0.5; C = (ismember(Cycles,R,'rows')')/2; else C = ismember(Cycles,R,'rows')'; end S = sum(C); NofT(i)=S; end Table=[U NofT]; Table(:,3) = []; %Delete 3rd column SortedCycles=sortrows(Table,[2 1]); ZeroRow = zeros(1,3); SortedCycles = [SortedCycles ; ZeroRow]; % Delete Cycles with Start=End giving up the number of Cycles to previous

Cycle for i=1:Size; if (SortedCycles(i+1,2)-SortedCycles(i+1,1))<=0; SortedCycles(i,3)=SortedCycles(i,3)+SortedCycles(i+1,3); SortedCycles(i+1,:)=0; end end SortedCycles; SortedCycles = SortedCycles(any(SortedCycles,2),:); Sum = sum(SortedCycles); Size=size(SortedCycles,1) xlswrite('Cycles Counted for Min-Max of Agressive Loading History (Larger

Bin).xlsx',SortedCycles); MinMaxTable = zeros(1,1); for i=1:Size; row1 = SortedCycles(i,1); MinMaxTable = [MinMaxTable ; row1]; row2 = SortedCycles(i,2); MinMaxTable = [MinMaxTable ; row2]; end if MinMaxTable(2,1) - MinMaxTable(3,1) <=0; MinMaxTable(1:3) = []; else MinMaxTable(1) = []; end MinMaxTable xlswrite('Cycles History for Min-Max of Agressive Loading History (Larger

Bin).xlsx',MinMaxTable);

Page 129: Finite Element Modeling of Buried Longitudinally Welded

112

APPENDIX C – ABAQUS Input File

**

** PARTS

**

*Part, name=CutSoil

*End Part

**

*Part, name=Pipe

*End Part

**

**

** ASSEMBLY

**

*Assembly, name=Assembly

**

*Instance, name=Pipe-1, part=Pipe

*Element, type=C3D20R

*Element, type=C3D15

*Nset, nset=Pipeline, generate

*Elset, elset=Pipeline, generate

** Section: Steel

*Solid Section, elset=Pipeline, material=X56

,

*End Instance

**

*Instance, name=CutSoil-1, part=CutSoil

*Element, type=C3D15

*Nset, nset=SoilBox, generate

*Elset, elset=SoilBox, generate

** Section: SoilBox

*Solid Section, elset=SoilBox, material=Soil

,

*End Instance

**

*Nset, nset=PipeEnds, instance=Pipe-1

*Elset, elset=PipeEnds, instance=Pipe-1, generate

*Nset, nset=SoilBottom, instance=CutSoil-1

*Elset, elset=SoilBottom, instance=CutSoil-1

*Nset, nset=SoilFronts, instance=CutSoil-1

*Elset, elset=SoilFronts, instance=CutSoil-1, generate

*Nset, nset=SoilSides, instance=CutSoil-1

*Elset, elset=SoilSides, instance=CutSoil-1

*Nset, nset=Tpipe, instance=Pipe-1, generate

*Elset, elset=Tpipe, instance=Pipe-1, generate

Page 130: Finite Element Modeling of Buried Longitudinally Welded

113

*Nset, nset=Tsoil, instance=CutSoil-1, generate

*Elset, elset=Tsoil, instance=CutSoil-1, generate

*Surface, type=ELEMENT, name=PipeExtSurf

*Surface, type=ELEMENT, name=PipeIntSurf

*Surface, type=ELEMENT, name=SoilCavitySurf

*End Assembly

*Amplitude, name=Pcycle

1., 2.75, 2., 2.91, 3., 2.69, 4., 3.17

5., 2.64, 6., 3.43, 7., 2.58, 8., 3.69

9., 2.52, 10., 3.94, 11., 2.47, 12., 4.2

13., 2.41, 14., 4.46, 15., 2.35, 16., 4.72

17., 2.3, 18., 4.98, 19., 2.24, 20., 5.24

21., 2.18, 22., 5.5, 23., 2.13, 24., 5.76

25., 2.07, 26., 6.02, 27., 2.01, 28., 6.28

29., 1.96, 30., 6.54, 31., 1.9, 32., 6.8

33., 1.85, 34., 7.06, 35., 1.79, 36., 7.31

37., 1.73, 38., 7.57, 39., 1.68, 40., 7.83

41., 1.62, 42., 8.09, 43., 1.56, 44., 8.35

45., 1.51, 46., 8.61, 47., 1.45, 48., 8.87

49., 1.39, 50., 9.13, 51., 1.34, 52., 9.39

53., 1.28, 54., 9.65, 55., 1.22, 56., 9.91

57., 1.17, 58., 10.17, 59., 1.11, 60., 10.43

**

** MATERIALS

**

*Material, name=Soil

*Density

0.0015,

*Elastic

10., 0.3

*Expansion

0.0001,

*Mohr Coulomb

0.,0.

*Mohr Coulomb Hardening

0.03,0.

*Material, name=X56

*Density

0.0078,

*Elastic

207000., 0.3

*Expansion

1e-05,

*Plastic

370., 0., 0.

Page 131: Finite Element Modeling of Buried Longitudinally Welded

114

423.972, 0.0122871, 0.

428.222, 0.0131397, 0.

432.472, 0.0140509, 0.

436.722, 0.0150241, 0.

440.971, 0.016063, 0.

445.221, 0.0171714, 0.

449.471, 0.0183534, 0.

453.721, 0.0196131, 0.

457.971, 0.020955, 0.

462.22, 0.0223836, 0.

466.47, 0.0239038, 0.

470.72, 0.0255207, 0.

474.97, 0.0272394, 0.

479.22, 0.0290656, 0.

483.469, 0.0310049, 0.

487.719, 0.0330634, 0.

491.969, 0.0352473, 0.

496.219, 0.0375631, 0.

500.468, 0.0400178, 0.

504.718, 0.0426183, 0.

508.968, 0.0453721, 0.

337., 0., 80.

387.223, 0.00825483, 80.

393.31, 0.00909148, 80.

399.398, 0.0100163, 80.

405.485, 0.0110375, 80.

411.572, 0.0121636, 80.

417.659, 0.013404, 80.

423.747, 0.0147686, 80.

429.834, 0.0162683, 80.

435.921, 0.0179144, 80.

442.008, 0.0197192, 80.

448.096, 0.0216957, 80.

454.183, 0.0238579, 80.

460.27, 0.0262207, 80.

466.357, 0.0287999, 80.

472.445, 0.0316124, 80.

478.532, 0.0346761, 80.

484.619, 0.03801, 80.

490.706, 0.0416344, 80.

496.794, 0.0455707, 80.

502.881, 0.0498417, 80.

508.968, 0.0544715, 80.

**

** INTERACTION PROPERTIES

Page 132: Finite Element Modeling of Buried Longitudinally Welded

115

**

*Surface Interaction, name=SS

1.,

*Friction, slip tolerance=0.005

0.3,

*Surface Behavior, pressure-overclosure=HARD

**

** PREDEFINED FIELDS

**

** Name: Tpipe Type: Temperature

*Initial Conditions, type=TEMPERATURE

Tpipe, 15.

** Name: Tsoil Type: Temperature

*Initial Conditions, type=TEMPERATURE

Tsoil, 15.

**

** INTERACTIONS

**

** Interaction: SS

*Contact Pair, interaction=SS, type=SURFACE TO SURFACE, adjust=0.0, tied

SoilCavitySurf, PipeExtSurf

** ----------------------------------------------------------------

**

** STEP: Gravity

**

*Step, name=Gravity, nlgeom=NO, inc=1000

*Static

1., 1., 1e-40, 1.

**

** BOUNDARY CONDITIONS

**

** Name: PipeEnds Type: Displacement/Rotation

*Boundary

PipeEnds, 3, 3

** Name: SoilBottom Type: Displacement/Rotation

*Boundary

SoilBottom, 1, 1

SoilBottom, 2, 2

SoilBottom, 3, 3

** Name: SoilFronts Type: Displacement/Rotation

*Boundary

SoilFronts, 3, 3

** Name: SoilSides Type: Displacement/Rotation

*Boundary

SoilSides, 1, 1

Page 133: Finite Element Modeling of Buried Longitudinally Welded

116

SoilSides, 3, 3

**

** LOADS

**

** Name: Gravity Type: Gravity

*Dload

, GRAV, 0.0098, 0., -1., 0.

**

** OUTPUT REQUESTS

**

*Restart, write, frequency=0

**

** FIELD OUTPUT: F-Output-1

**

*Output, field, variable=PRESELECT

**

** HISTORY OUTPUT: H-Output-1

**

*Output, history, variable=PRESELECT

*End Step

** ----------------------------------------------------------------

**

** STEP: Temperature

**

*Step, name=Temperature, nlgeom=NO, inc=1000

*Static

1., 1., 1e-40, 1.

**

** PREDEFINED FIELDS

**

** Name: Tpipe Type: Temperature

*Temperature

Tpipe, 80.

**

** OUTPUT REQUESTS

**

*Restart, write, frequency=0

**

** FIELD OUTPUT: F-Output-1

**

*Output, field, variable=PRESELECT

**

** HISTORY OUTPUT: H-Output-1

**

*Output, history, variable=PRESELECT

Page 134: Finite Element Modeling of Buried Longitudinally Welded

117

*End Step

** ----------------------------------------------------------------

**

** STEP: Pressure

**

*Step, name=Pressure, nlgeom=NO, inc=1000

*Static

1., 1., 1e-40, 1.

**

** LOADS

**

** Name: P Type: Pressure

*Dsload

PipeIntSurf, P, 10.

**

** OUTPUT REQUESTS

**

*Restart, write, frequency=0

**

** FIELD OUTPUT: F-Output-1

**

*Output, field, variable=PRESELECT

**

** HISTORY OUTPUT: H-Output-1

**

*Output, history, variable=PRESELECT

*End Step

Page 135: Finite Element Modeling of Buried Longitudinally Welded

118

APPENDIX D – ABAQUS Report Example

********************************************************************************

Statically Equivalent Linear Stress Distribution across a Section,

written on Date Indicated

Source

-------

ODB: C:/*.odb

Step: Pressure

Frame: Increment 1: Step Time = 1.000

Linearized Stresses for stress line 'SCL'

Start point, Point 1 - (9.90777969360352, 438.650207519531, 1000)

End point, Point 2 - (10.3930807113647, 456.264678955078, 1000)

Number of intervals - 40

------------------------------------- COMPONENT RESULTS -------------------------------------

S11 S22 S33 S12 S13 S23

0 1.50064 189.975 -77.1071 -26.9611 0.000914721 2.49023e-05

0.440529 10.7086 157.468 -84.0969 -17.6226 -0.000104373 -3.28035e-06

0.881058 14.3533 136.637 -89.2531 -12.6624 7.09119e-06 -3.40971e-07

1.32159 15.1787 123.072 -93.0747 -9.49782 1.68619e-06 -6.67067e-07

1.76212 14.8283 113.689 -95.9948 -7.21984 -1.1096e-06 -8.46526e-07

2.20264 14.0649 106.76 -98.3025 -5.49221 9.55155e-08 -9.00495e-07

2.64317 13.0834 101.161 -100.277 -4.11358 -9.94643e-08 -9.69991e-07

3.0837 12.0422 96.4699 -101.996 -3.00724 -2.70706e-07 -9.96492e-07

3.52423 11.0827 92.4651 -103.486 -2.1285 -7.29473e-08 -1.14596e-06

3.96476 10.1398 88.9386 -104.827 -1.40159 -1.90878e-08 -9.49357e-07

4.40529 9.28576 85.7652 -106.035 -0.804824 -9.98379e-08 -1.09816e-06

4.84582 8.51961 82.8544 -107.138 -0.316621 -4.35051e-07 -1.00983e-06

5.28635 7.81651 80.1152 -108.17 0.0859991 -6.35731e-07 -1.1078e-06

5.72688 7.11305 77.5458 -109.152 0.429193 -2.72144e-07 -1.74487e-06

6.1674 6.47442 75.1609 -110.059 0.700721 -2.87416e-07 -1.14502e-06

6.60793 5.89226 72.8494 -110.928 0.932453 -3.56561e-07 -1.34228e-06

7.04846 5.36137 70.604 -111.76 1.13076 -3.24729e-07 -1.46821e-06

7.48899 4.87179 68.4115 -112.565 1.30073 4.05415e-08 -1.61171e-06

Page 136: Finite Element Modeling of Buried Longitudinally Welded

119

7.92952 4.44361 66.2878 -113.331 1.43976 -1.33458e-06 -1.73046e-06

8.37005 4.03886 64.1871 -114.082 1.56454 -1.87036e-06 -1.85905e-06

8.81058 3.6815 62.1241 -114.808 1.67181 5.07757e-06 -2.30485e-06

9.25111 3.3524 60.0883 -115.518 1.76732 6.20876e-08 -2.30742e-06

9.69164 3.05445 58.0568 -116.217 1.85668 -2.61125e-05 -1.48921e-06

10.1322 2.79171 56.0627 -116.894 1.93788 2.15678e-06 -2.85157e-06

10.5727 2.55446 54.0459 -117.57 2.01811 9.4648e-05 -6.83548e-06

11.0132 2.35023 52.0747 -118.223 2.09557 -8.69029e-06 -3.05306e-06

11.4538 2.172 50.0695 -118.878 2.17593 -0.000362451 1.1048e-05

11.8943 2.02049 48.1103 -119.511 2.25759 -6.92792e-05 -1.22991e-06

12.3348 1.90113 46.1086 -120.147 2.34632 0.00139442 -6.3865e-05

12.7753 1.80143 44.1468 -120.766 2.43878 0.000754958 -3.79915e-05

13.2159 1.73949 42.1653 -121.379 2.54468 -0.00573958 0.00024583

13.6564 1.69255 40.2103 -121.979 2.65446 -0.00544583 0.000235095

14.0969 1.68092 38.2634 -122.567 2.7894 0.0224755 -0.0010039

14.5375 1.68282 36.3355 -123.145 2.92736 0.034734 -0.00155395

14.978 1.72309 34.4469 -123.699 3.09257 -0.0744971 0.00329452

15.4185 1.78724 32.5311 -124.254 3.26926 -0.190682 0.0084657

15.859 1.79555 31.0792 -124.688 3.4821 0.185249 -0.00804487

16.2996 1.88391 29.3371 -125.184 3.81863 0.583377 -0.023698

16.7401 1.66834 28.2721 -125.568 3.91926 -0.0998983 0.00343847

17.1806 1.1641 27.7303 -125.882 3.7096 -1.37179 0.0571848

17.6212 0.552789 27.249 -126.209 3.31325 -2.62428 0.11596

Membrane

(Average) Stress 5.57059 68.5078 -112.326 -0.393342 -0.0558054 0.00241053

Bending

Stress, Point 1 6.8275 54.8041 18.4895 -7.22837 0.166929 -0.00720545

Membrane plus

Bending, Point 1 12.3981 123.312 -93.837 -7.62171 0.111124 -0.00479492

Bending

Stress, Point 2 -6.8275 -54.8041 -18.4895 7.22837 -0.166929 0.00720545

Membrane plus

Bending, Point 2 -1.25691 13.7037 -130.816 6.83502 -0.222735 0.00961598

Peak Stress,

Point 1 -10.8974 66.6635 16.7299 -19.3393 -0.110209 0.00481983

Page 137: Finite Element Modeling of Buried Longitudinally Welded

120

Peak Stress,

Point 2 1.8097 13.5453 4.60649 -3.52178 -2.40155 0.106344

------------------------------------- INVARIANT RESULTS -------------------------------------

Bending components in equation for computing

membrane plus bending stress invariants are: S11, S22, S33, S12, S13, S23

Max. Mid. Min. Tresca Mises

Prin. Prin. Prin. Stress Stress

Membrane

(Average) Stress 68.5103 5.56816 -112.327 180.837 159.001

Membrane plus

Bending, Point 1 123.833 11.8769 -93.8371 217.67 188.534

Membrane plus

Bending, Point 2 16.3562 -3.909 -130.816 147.173 138.159

Peak Stress,

Point 1 71.2182 16.7302 -15.4525 86.6707 75.883

Peak Stress,

Point 2 14.5787 5.63313 -0.250339 14.8291 12.9333

Page 138: Finite Element Modeling of Buried Longitudinally Welded

121

APPENDIX E – MATLAB Code for the ABAQUS Data

clear; clc %% ABAQUS Input Input = inputdlg({'File Name - Elastic-Plastic Stress','File Name - Elastic-

Plastic Strain','File Name - Elastic Stress','File Name - Rainflow Counted

Cycles','Service Life - Years'},'Input Raw Data',[1 70; 1 70; 1 70; 1 70; 1

70],{'ElasticPlasticStress.xlsx';'ElasticPlasticStrain.xlsx';'ElasticStress.x

lsx';'Cycles Counted for Min-Max of Agressive Loading History (Larger

Bin).xlsx';'50'}); FileNameElasticPlasticStress = char([Input(1,1)]); FileNameElasticPlasticStrain = char([Input(2,1)]); FileNameElastic = char([Input(3,1)]); RainflowCountedCycles = char([Input(4,1)]); ServiceLife = str2double(Input(5,1)); %% Wall Thickness Meaasurement [v,T,vT]=xlsread(FileNameElastic,'A62:A62'); a=regexp(vT,'\s+','split');

n=numel(a{1}); m=numel(a); thickness=transpose(reshape(str2double([a{:}]),n,m)); t=thickness(1,2); xlswrite(['RefinedCell' FileNameElasticPlasticStress],t,1,'D1');

%% Extraction of Elastic Stress Data Points t = readtable(FileNameElastic); [rows, columns] = size(t); Size = round(rows/106); CycleN=Size; % Zeroing the Table of Results Stresses=zeros(1,10); for j=1:CycleN; CycleNumber=j; % Identification of Cycles from the ABAQUS Report file CycleStartCell = CycleNumber; if CycleNumber<=1; LinearStressStart=71; else LinearStressStart=71+106*(CycleStartCell-1); end LinearizedStart=['A' num2str(LinearStressStart) ':' 'A'

num2str(LinearStressStart)]; % Read Loading History from Raw Data [v,T,vT]=xlsread(FileNameElastic,LinearizedStart);

a=regexp(vT,'\s+','split'); n=numel(a{1}); m=numel(a); ElasticStart=transpose(reshape(str2double([a{:}]),n,m)); Stresses = [Stresses ; ElasticStart]; end %xlswrite('ExtractedStresses.xlsx',Stresses); %% Cleaning of Elastic Stress Data (Removing the Intermidiate Points) t = Stresses; [rows, columns] = size(t); % Creation of Min values tmin = zeros(rows+2,columns); for i=1:rows-2; trow1 = t(i+1,1:end); trow2 = t(i+2,1:end); if trow1(1,5) < trow2(1,5); tmin(i+1,1:end) = trow1; end end % Creation of Max values tmin(rows+1,:) = []; tmin(rows+1,:) = [];

Page 139: Finite Element Modeling of Buried Longitudinally Welded

122

tmax = t-tmin; Zeros = zeros(2,columns); tmax = [tmax ; Zeros]; for i=1:rows-2; trow1 = t(i+1,1:end); trow2 = t(i+2,1:end); if trow1(1,5) > trow2(1,5); tmax(i+2,1:end) = [0]; end end tmax(rows+1,:) = []; tmax(rows+1,:) = []; tmax(1,1:end) = [0]; % Creation of Min values (Cont.) for i=1:rows-2; trow1 = t(i+1,1:end); trow2 = t(i+2,1:end); if trow1(1,5) < trow2(1,5); tmin(i+2,1:end) = [0]; end end % Combining the Cleaned Min/Max Loading History tMinMax = tmin+tmax; % Deleting Zero Rows tMinMax = tMinMax(any(tMinMax,2),:); [rows, columns] = size(tMinMax); % Restoring Numbering of Rows for i=1:rows; tMinMax(i,1) = i; end tMinMax; xlswrite(['Refined' FileNameElastic],tMinMax);

tMinMax(1:end,2) = [0]; tMinMax(1:end,10) = [0]; [datarow, datacolumn] = size(tMinMax); B = zeros(1,1); for i=1:datarow; A = {[num2str(tMinMax(i,1)) ' ' num2str(tMinMax(i,2)) ' '

num2str(tMinMax(i,3)) ' ' num2str(tMinMax(i,4)) ' ' num2str(tMinMax(i,5)) ' '

num2str(tMinMax(i,6)) ' ' num2str(tMinMax(i,7)) ' ' num2str(tMinMax(i,8)) ' '

num2str(tMinMax(i,9)) ' ' num2str(tMinMax(i,10))]}; B = [B ; A]; end B(1) = []; xlswrite(['RefinedCell' FileNameElastic],B);

%% Extraction of Elastic Stress Data Points (Sm) t = readtable(FileNameElastic); [rows, columns] = size(t); Size = round(rows/106); CycleN=Size; % Zeroing the Table of Results Stresses=zeros(1,9); for j=1:CycleN; CycleNumber=j; % Identification of Cycles from the ABAQUS Report file CycleStartCell = CycleNumber; if CycleNumber<=1; LinearStressStart=65; else LinearStressStart=65+106*(CycleStartCell-1); end LinearizedStart=['A' num2str(LinearStressStart) ':' 'A'

num2str(LinearStressStart)]; % Read Loading History from Raw Data [v,T,vT]=xlsread(FileNameElastic,LinearizedStart);

a=regexp(vT,'\s+','split'); n=numel(a{1}); m=numel(a); ElasticStart=transpose(reshape(str2double([a{:}]),n,m)); Stresses = [Stresses ; ElasticStart]; end

Page 140: Finite Element Modeling of Buried Longitudinally Welded

123

%% Cleaning of Elastic Stress Data (Removing the Intermidiate Points) t = Stresses; [rows, columns] = size(t); % Creation of Min values tmin = zeros(rows+2,columns); for i=1:rows-2; trow1 = t(i+1,1:end); trow2 = t(i+2,1:end); if trow1(1,5) < trow2(1,5); tmin(i+1,1:end) = trow1; end end % Creation of Max values tmin(rows+1,:) = []; tmin(rows+1,:) = []; tmax = t-tmin; Zeros = zeros(2,columns); tmax = [tmax ; Zeros]; for i=1:rows-2; trow1 = t(i+1,1:end); trow2 = t(i+2,1:end); if trow1(1,5) > trow2(1,5); tmax(i+2,1:end) = [0]; end end tmax(rows+1,:) = []; tmax(rows+1,:) = []; tmax(1,1:end) = [0]; % Creation of Min values (Cont.) for i=1:rows-2; trow1 = t(i+1,1:end); trow2 = t(i+2,1:end); if trow1(1,5) < trow2(1,5); tmin(i+2,1:end) = [0]; end end % Combining the Cleaned Min/Max Loading History tMinMax = tmin+tmax; % Deleting Zero Rows tMinMax = tMinMax(any(tMinMax,2),:); [rows, columns] = size(tMinMax); % Restoring Numbering of Rows for i=1:rows; tMinMax(i,1) = i; end tMinMax; xlswrite(['Refined' FileNameElastic],tMinMax);

tMinMax(:,2) = []; tMinMax(:,8) = []; [datarow, datacolumn] = size(tMinMax); B = zeros(1,1); for i=1:datarow; A = {[num2str(tMinMax(i,1)) ' ' num2str(tMinMax(i,2)) ' '

num2str(tMinMax(i,3)) ' ' num2str(tMinMax(i,4)) ' ' num2str(tMinMax(i,5)) ' '

num2str(tMinMax(i,6)) ' ' num2str(tMinMax(i,7))]}; B = [B ; A]; end B(1) = []; xlswrite(['RefinedCell' FileNameElastic],B,2);

Page 141: Finite Element Modeling of Buried Longitudinally Welded

124

References

[1] "Crude oil facts," Government of Canada: Natural Resources Canada, 13 September 2018. [Online].

Available: https://www.nrcan.gc.ca/energy/facts/crude-oil/20064. [Accessed 19 September 2018].

[2] "Pipeline Ruptures," National Energy Board, 21 June 2017. [Online]. Available: https://www.neb-

one.gc.ca/sftnvrnmnt/sft/pplnrptr/pplnrptr-eng.xls. [Accessed 19 September 2018].

[3] "API 5L PSL2," Trio Steel and Engineering LLP, [Online]. Available: http://www.triosteel.com.

[4] CSA Z245.1-14: Steel pipe, Toronto: CSA Group, 2014.

[5] CSA Z622-15: Oil and gas pipeline systems, Toronto: CSA Group, 2016.

[6] A. Takekazu, N. Kimihiro, Y. Koji and S. Nobuhisa, "Development of High Performance UOE Pipe for

Linepipe," JFE Technical Report, vol. 18, no. 3, pp. 23-35, 2013.

[7] K. Shohei, H. Naoya and O. Kenji, "Multiple-Electrode Submerged Arc Welding Process with Low Heat

Input," JFE Technical Report, vol. 20, no. 1, pp. 106-111, 2015.

[8] Q. Ren, T. Zou, D. Li, D. Tang and Y. Peng, "Numerical study on the X80 UOE pipe forming process,"

Journal of Materials Processing Technology, vol. 215, pp. 264-277, 2015.

[9] E. Anisimov, M. Singh and L. Sudak, "Fatigue Analysis of The Longitudinally Welded Large-Diameter

Oil Pipelines," in 26th Canadian Congress of Applied Mechanics (CANCAM 2017), Victoria, 2017.

[10] B. Pollard and R. Cover, "Fatigue of Steel Weldments," Welding Research Supplement, no. 11, pp.

544-554, 1972.

[11] I. Lillemäe, H. Remes, S. Liinalampi and A. Itävuo, "Influence of weld quality on the fatigue strength

of thin normal and high strength steel butt joints," Weld World, vol. 60, pp. 731-740, 2016.

[12] H. Remes and P. Varsta, "Statistics of Weld Geometry for Laser-Hybrid Welded Joints and its

Application Within Notch Stress Approach," Welding in the World, vol. 54, no. 7, pp. 189-207, 2010.

[13] T. N. Nguyen and M. Wahab, "The effect of weld geometry and residual stresses on the fatigue of

welded joints under combined loading," Journal of Materials Processing Technology, vol. 77, pp.

201-208, 1998.

[14] Z. Barsoum and I. Barsoum, "Residual stress effects on fatigue life of welded structures using LEFM,"

Engineering Failure Analysis, vol. 16, pp. 449-467, 2009.

[15] L. Rodrigues, J. Freire and R. Vieira, "Measurement of residual stresses in UOE-SAW line pipes,"

Experimental Techniques, no. 2, pp. 58-62, 2008.

Page 142: Finite Element Modeling of Buried Longitudinally Welded

125

[16] L. Lothhammer, M. Viotti, C. Veiga and A. A. Jr, "Experimental Evaluation of Residual Stresses in Pipes

Manufactured by UOE and ERW Processes," Experimental Mechanics, vol. 57, p. 287–296, 2017.

[17] T. Zou, D. Li, GuanghanWu and Y. Peng, "Yield strength development from high strength steel plate

to UOE pipe," Materials and Design, vol. 89, pp. 1107-1122, 2016.

[18] C. Sonsino, "Effect of residual stresses on the fatigue behaviour of welded joints depending on

loading conditions and weld geometry," International Journal of Fatigue, vol. 31, pp. 88-101, 2009.

[19] T. Shiozaki, N. Yamaguchi, Y. Tamai, J. Hiramoto and K. Ogawa, "Effect of weld toe geometry on

fatigue life of lap fillet welded ultra-high strength steel joints," International Journal of Fatigue, vol.

116, pp. 409-420, 2018.

[20] M. Roy, J. Walsh and P. J. Withers, "The effect of the weld fusion zone shape on residual stress in

submerged arc welding," International Journal of Advanced Manufacturing Technology, pp. 1-14,

2016.

[21] API 579-1/ASME FFS-1 2007 Fitness-For-Service, New York: American Petroleum Institute, 2007.

[22] ASME BPVC Section VIII Part 2: Rules for construction of pressure vessels, New York: The American

Society of Mechanical Engineers, 2015.

[23] ASME B31.4-2012: Pipeline Transportation Systems for Liquid Hydrocarbons and Other Liquids, New

York: The American Society of Mechanical Engineers, 2012.

[24] BS 7608:2014+A1:2015: Guide to fatigue design and assessment of steel products, The British

Standards Institution, 2015.

[25] BS 7910:2013+A1:2015: Guide to methods for assessing the acceptability of flaws in metallic

structures, The British Standards Institution, 2015.

[26] J. F. Kiefner, "Oil Pipeline Characteristics and Risk Factors: Illustrations from the Decade of

Construction," American Petroleum Institute, New York, 2001.

[27] G. A. Antaki, Piping and Pipeline Engineering: Design, Construction, Maintenance, Integrity, and

Repair, New York: Marcel Dekker, 2003.

[28] S. Berge and H. Myhre, "Fatigue Strength of Misaligned Cruciform and Butt Joints," Norwegian

Maritime Research, no. 1, pp. 29-39, 1977.

[29] M. J. Worswick and R. J. Pick, "Influence of Weld Misalignment and Weld Metal Overmatch on the

Prediction of Fracture in Pipeline Girth Welds," International Journal of Pressure Vessels, vol. 21, pp.

209-234, 1985.

Page 143: Finite Element Modeling of Buried Longitudinally Welded

126

[30] J. Ferreira and C. Branco, "Influence of misalignment on the fatigue strength of butt welds,"

International Journal of Fatigue, vol. 13, no. 5, pp. 405-409, 1991.

[31] L. Ong and K. Hoon, "Bending Stresses at Longitudinal Weld Joints of Pressurized Cylindrical Shells

due to Angular Distortion," Journal of Pressure Vessel Technology, vol. 118, pp. 369-373, 1996.

[32] R. Andrews, "The Effect of Misalignment on the Fatigue Strength of Welded Cruciform Joints,"

Fatigue & Fracture of Engineering Materials & Structures, vol. 19, no. 6, pp. 755-768, 1996.

[33] S. Maddox, "Report 279: Fitness-for-purpose assessment of misalignment in transverse butt welds

subject to fatigue loading," The Welding Institute, 1985.

[34] J. Zeman, "On the problem of angular misalignment at longitudinal weld joints of cylindrical shells,"

International Journal of Pressure Vessels & Piping, vol. 58, pp. 179-184, 1994.

[35] N. Bock and J. L. Zeman, "On bending stresses at longitudinal weld joints of cylindrical shells due to

peaking," International Journal of Pressure Vessels & Piping, vol. 60, pp. 103-106, 1994.

[36] T. Pastor and D. Osage, "Modernization of Pressure Vessel Design Codes ASME Section VIII, Division

2, 2007 Edition," Journal of Pressure Vessel Technology, vol. 129, no. 11, pp. 754-758, 2007.

[37] D. A. Osage, "Fatigue Assessment for In-Service Components – A New Part for API 579-1/ASME FFS-

1 Fitness-For-Service," Procedia Engineering, vol. 133, pp. 320-347, 2015.

[38] P. Dong, "A structural stress definition and numerical implementation for fatigue analysis of welded

joints," International Journal of Fatigue, vol. 23, pp. 865-876, 2001.

[39] O. Doerk, W. Fricke and C. Weissenborn, "Comparison of different calculation methods for structural

stresses at welded joints," International Journal of Fatigue, vol. 25, pp. 359-369, 2003.

[40] P. Rohart, S. Panier, S. Hariri and Y. Simonet, "Contribution to pressure vessels design of innovative

methods and structure - Part I," International Journal of Pressure Vessels and Piping, vol. 134, pp.

25-32, 2015.

[41] B. Möller, J. Baumgartner, R. Wagener, H. Kaufmann and T. Melz, "Low cycle fatigue life assessment

of welded high-strength structural steels based on nominal and local design concepts," International

Journal of Fatigue, vol. 101, pp. 192-208, 2017.

[42] R. Goyal, M. El-Zein and G. Glinka, "A robust stress analysis method for fatigue life prediction of

welded structures," Weld World, vol. 60, pp. 299-314, 2016.

[43] I. Lillemäe, H. Lammi, L. Molter and H. Remes, "Fatigue strength of welded butt joints in thin and

slender specimens," International Journal of Fatigue, vol. 44, pp. 98-106, 2012.

Page 144: Finite Element Modeling of Buried Longitudinally Welded

127

[44] K. Y, S. Y, K. T and O. T., "Low-cycle fatigue strength of butt welded joints with angular distortion,"

Iinternational Institute of Welding, 1971.

[45] T. Nykanen and T. Bjork, "Assessment of fatigue strength of steel butt-welded joints in as-welded

condition - Alternative approaches for curve fitting and mean stress effect analysis," Marine

Structures, vol. 44, pp. 288-310, 2015.

[46] R. Anthes, V. Kottgen and T. Seeger, "Notch shape factors of butt joints and double-T butt joints,"

Schweissen & Schneiden, vol. 45, no. 12, pp. 221-228, 1993.

[47] A. Pachoud, P. Manso and A. Schleiss, "New parametric equations to estimate notch stress

concentration factors at butt welded joints modeling the weld profile with splines," Engineering

Failure Analysis, vol. 72, pp. 11-24, 2017.

[48] J. Ellenberger, Piping Systems & Pipelines: ASME B31 Code Simplified, New York: McGrew-Hill, 2005.

[49] ASME B36.10M-2004: Welded and Seamless Wrought Steel Pipe, New York: The American Society

of Mechanical Engineers, 2004.

[50] "ASME B31.3-2014: Process Piping," The American Society of Mechanical Engineers, New York, 2014.

[51] API 5L 2012: Specification for Line Pipe, Washington: American Petroleum Institute, 2012.

[52] BS PD 5500:2015+A1:2015: Specification for unfired fusion welded pressure vessels, The British

Standards Institution, 2015.

[53] "Keystone XL Pipeline Project," [Online]. Available: https://keystonepipeline-xl.state.gov/.

[Accessed 12 June 2018].

[54] ASME B31.1: Power Piping, New York: The American Society of Mechanical Engineers, 2007.

[55] Guidlines for the Design of Buried Steel Pipe, American Society of Civil Engineers, 2005.

[56] N. Stromblad, "Modeling of Soil and Structure Interaction Subsea," Chalmers University of

Technology, Department of Applied Mechanics, Goteborg, 2014.

[57] B. F. Hikooei, "Numerical Modeling of Pipe-Soil Interaction Under Transverse Direction," University

of Calgary, Department of Civil Engineering, Calgary, 2013.

[58] H. Lee, "Finite element analysis of a buried pipeline," University of Manchester, Engineering and

Physical Science, Manchester, 2010.

[59] I. Scheidera, A. Nonn, A. Völling, A. Mondry and C. Kalwa, "A damage mechanics based evaluation

of dynamic fracture resistance in gas pipelines," Procedia Materials Science, vol. 3, p. 1956 – 1964,

2014.

Page 145: Finite Element Modeling of Buried Longitudinally Welded

128

[60] S. Y. Shin, K. J. Woo, B. Hwang and S. L. Sangho Kim, "Fracture-Toughness Analysis in Transition-

Temperature Region of Three American Petroleum Institute X70 and X80 Pipeline Steels,"

Metallurgical and Materials Transactions, vol. 40A, pp. 867-876, 2009.

[61] Y. Murakami, Metal Fatigue: Effects of Small Defects and Nonmetallic Inclusions, Kidlington: Elsevier

Science Ltd, 2002.

[62] H. Chunyong, L. He, Z. Weiwei, Y. Kun, C. Qiang and M. Qiurong, "Crack control technology for the

X80 large OD 1422 mm line pipes," Natural Gas Industry, vol. 3, pp. 479-484, 2016.

[63] A. Bussiba, P. Darcis, J. McColskey, C. McCowan, T. Siewert, G. Kohn, R. Smith and J. Merritt, "Fatigue

crack growth rates in six pipeline steels," in 6th International Pipeline Conference, Calgary, 2006.

[64] J. Capelle and G. Pluvinage, "Modification of failure risk by the use of high strength steels in

pipelines," Transilvania University of Brasov, Brasov, 2012.

[65] D. B. Rosado, W. D. Waele, D. Vanderschueren and S. Hertelé, "Latest developments in mechanical

properties and metallurgical features of high strength line pipe steels," in Sustainable Construction

and Design, 2013.

[66] "E646: Standard Test Method for Tensile Strain-Hardening Exponents (n-Values) of Metallic Sheet

Materials," ASTM International, 2015.

[67] W. Ramberg and W. R. Osgood, "Technical Note No. 902: Description of stress–strain curves by three

parameters," National Advisory Committee For Aeronautics, Washington, 1943.

[68] D. S. Cronin and R. J. Pick, "Prediction of the failure pressure for complex corrosion defects,"

International Journal of Pressure Vessels and Piping, vol. 79, pp. 279-287, 2002.

[69] M. J. Rosenfeld and J. F. Kiefner, "Basics of Metal Fatigue in Natural Gas Pipeline Systems — A Primer

for Gas Pipeline Operators," Technology for Energy Pipelines, Houston, 2006.

[70] A. Nieslony, "Rainflow Counting Algorithm," 4 April 2010. [Online]. Available:

https://www.mathworks.com/matlabcentral/fileexchange/3026-rainflow-counting-algorithm.

[Accessed 25 May 2017].

[71] J. Kiefner and W. Maxey, "Model helps prevent failure," Oil & Gas Journal, vol. 8, pp. 54-58, 2000.

[72] Pipetech International, [Online]. Available: http://pipepressure.com/library/petroleum-pipeline-

case-studies/2/. [Accessed 27 April 2017].

[73] L. Wang, L. A. Dykhno, G. J. Zabaras, Z. Huo, C. R. Holloway, N. B. Dee and W. N. Jr., "Study examines

causes of slugging in subsea flowline," Offshore, [Online]. Available: https://www.offshore-

mag.com/articles/print/volume-74/issue-6/pipelines-flowlines/study-examines-causes-of-slugging-

in-subsea-flowline-p1.html. [Accessed 27 April 2017].

Page 146: Finite Element Modeling of Buried Longitudinally Welded

129

[74] "Pipeline Accident Report NTSB/PAR-02/02: Pipeline Rupture and Subsequent Fire in Bellingham,

Washington, June 10, 1999," National Transportation Safety Board, Washington, 2002.

[75] "Pipeline Investigation Report P11H0011," Transportation Safety Board of Canada, Beardmore,

2011.

[76] "Pipeline Investigation Report P12H0105," Transportation Safety Board of Canada, Buick, 2012.

[77] Abaqus 6.12, Abaqus/CAE User’s Manual, Providence: Dassault Systèmes, 2012.

[78] R. K. Watkins and L. R. Anderson, Structural Mechanics of Buried Pipes, Boca Raton: CRC Press LLC,

1999.

[79] "E466: Standard Practice for Conducting Force Controlled Constant Amplitude Axial Fatigue Tests of

Metallic Materials," ASTM International, 2015.

[80] ASM Handbook: Volume 19 - Fatigue and Fracture, ASM International, 1997.

[81] J. Emslie, C. Watson and K. Wright, "ASME III Fatigue assessment plasticity correction factors for

austenitic stainless steels," in Pressure Vessels & Piping Conference, Anaheim, 2014.

[82] T. Topper and T. Lam, "Effective strain–fatigue life data for variable amplitude fatigue," International

Journal of Fatigue, vol. 19, no. 1, pp. 137-143, 1998.

[83] J. John H. Crews, "TN D-5253: Elastoplastic stress-strain behavior at notch roots in sheet specimens

under constant-amplitude loading," NASA, Washington, 1969.

[84] D. Navrocky, Calculations for welded joints in view of stress concentration, Leningrad: Mechanical

Engineering, 1968.

[85] P. P. Milella, Fatigue and Corrosion in Metals, Springer Science & Business Media, 2012.

[86] Abaqus 6.11: Analysis User’s Manual - Elements, vol. IV, Providence: Dassault Systèmes, 2011.

[87] MATLAB® Primer, Natick: The MathWorks, Inc., 2018.

[88] Z.-P. Zhao, G.-Y. Qiao, G.-P. Li, W.-W. Yang, B. Liao and F.-R. Xiao, "Fatigue properties of

ferrite/bainite dual-phase X80 pipeline steel welded joints," Science and Technology of Welding and

Joining, vol. 22, no. 3, pp. 217-226, 2017.

[89] Y. Kim, C. Kim, W. Kim and K. Shin, "Fatigue Crack Growth Behavior for Welded Joint of X80 Pipeline

Steel," Journal of the Korean Society for Nondestructive Testing, vol. 29, no. 1, pp. 43-47, 2009.

[90] D.-H. Jung, J.-K. Kwon, N.-S. Woo, Y.-J. Kim, M. Goto and S. Kim, "S–N Fatigue and Fatigue Crack

Propagation Behaviors of X80 Steel at Room and Low Temperatures," Metallurgical and Material

Transactions, vol. 45, pp. 654-662, 2014.

Page 147: Finite Element Modeling of Buried Longitudinally Welded

130

[91] S. K. Sharma and S. Maheshwari, "A review on welding of high strength oil and gas pipeline steels,"

Journal of Natural Gas Science and Engineering, vol. 38, pp. 203-217, 2017.

[92] X. Chen, H. Lu, G. Chen and X. Wang, "A comparison between fracture toughness at different

locations of longitudinal submerged arc welded and spiral submerged arc welded joints of API X80

pipeline steels," Engineering Fracture Mechanics, vol. 148, no. 8, pp. 110-121, 2015.

[93] R. Li, X. Zuo, Y. Hu, Z. Wang and D. Hu, "Microstructure and properties of pipeline steel with a

ferrite/martensite dual-phase microstructure," Materials Characterization, vol. 62, no. 8, p. 801–

806, 2011.

[94] L.-C. Peng, "Stress Analysis Methods for Underground Pipe Lines: Part 2 - Soil-Pipe Interaction," Pipe

Line Industry, no. 5, pp. 68-72, 1978.

[95] W. D. Pilkey and D. F. Pilkey, Peterson's Stress Concentration Factors, 3rd Edition, New Jersey: Wiley,

2008.

[96] W. C. Young, R. G. Budynas and A. M. Sadegh, Roark's Formulas for Stress and Strain, 8th Edition,

McGraw-Hill Education, 2011.

[97] L.-C. Peng, "Stress Analysis Methods for Underground Pipe Lines: Part 1 - Basic Calculations," Pipe

Line Industry, no. 4, pp. 67-71, 1978.