13.12.2012 Views

JPE - Sept09 - cover2-4.pmd - Pipes & Pipelines International ...

JPE - Sept09 - cover2-4.pmd - Pipes & Pipelines International ...

JPE - Sept09 - cover2-4.pmd - Pipes & Pipelines International ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

September, 2009 Vol.8, No.3<br />

Scientific<br />

Surveys Ltd, UK<br />

Journal of<br />

Pipeline Engineering<br />

incorporating<br />

The Journal of Pipeline Integrity<br />

Sample issue<br />

Clarion<br />

Technical Publishers, USA


Journal of Pipeline Engineering<br />

Editorial Board - 2009<br />

Obiechina Akpachiogu, Cost Engineering Coordinator, Addax Petroleum<br />

Development Nigeria, Lagos, Nigeria<br />

Mohd Nazmi Ali Napiah, Pipeline Engineer, Petronas Gas, Segamat, Malaysia<br />

Dr Michael Beller, NDT Systems & Services AG, Stutensee, Germany<br />

Jorge Bonnetto, Operations Vice President, TGS, Buenos Aires, Argentina<br />

Mauricio Chequer, Tuboscope Pipeline Services, Mexico City, Mexico<br />

Dr Andrew Cosham, Atkins Boreas, Newcastle upon Tyne, UK<br />

Prof. Rudi Denys, Universiteit Gent – Laboratory Soete, Gent, Belgium<br />

Leigh Fletcher, MIAB Technology Pty Ltd, Bright, Australia<br />

Roger Gomez Boland, Sub-Gerente Control, Transierra SA,<br />

Santa Cruz de la Sierra, Bolivia<br />

Daniel Hamburger, Pipeline Maintenance Manager, El Paso Eastern <strong>Pipelines</strong>,<br />

Birmingham, AL, USA<br />

Prof. Phil Hopkins, Executive Director, Penspen Ltd, Newcastle upon Tyne, UK<br />

Michael Istre, Engineering Supervisor, Project Consulting Services,<br />

Houston, TX, USA<br />

Dr Shawn Kenny, Memorial University of Newfoundland – Faculty of Engineering<br />

and Applied Science, St John’s, Canada<br />

Dr Gerhard Knauf, Mannesmann Forschungsinstitut GmbH, Duisburg, Germany<br />

Lino Moreira, General Manager – Development and Technology Innovation,<br />

Petrobras Transporte SA, Rio de Janeiro, Brazil<br />

Prof. Andrew Palmer, Dept of Civil Engineering – National University of Singapore,<br />

Singapore<br />

Prof. Dimitri Pavlou, Professor of Mechanical Engineering,<br />

Technological Institute of Halkida , Halkida, Greece<br />

Sample issue<br />

Dr Julia Race, School of Marine Sciences – University of Newcastle,<br />

Newcastle upon Tyne, UK<br />

Dr John Smart, John Smart & Associates, Houston, TX, USA<br />

Jan Spiekhout, NV Nederlandse Gasunie, Groningen, Netherlands<br />

Dr Nobuhisa Suzuki, JFE R&D Corporation, Kawasaki, Japan<br />

Prof. Sviatoslav Timashev, Russian Academy of Sciences – Science<br />

& Engineering Centre, Ekaterinburg, Russia<br />

Patrick Vieth, Senior Vice President – Integrity & Materials,<br />

CC Technologies, Dublin, OH, USA<br />

Dr Joe Zhou, Technology Leader, TransCanada PipeLines Ltd, Calgary, Canada<br />

Dr Xian-Kui Zhu, Senior Research Scientist, Battelle Pipeline Technology Center,<br />

Columbus, OH, USA<br />

❖ ❖ ❖


3rd Quarter, 2009 145<br />

The Journal of<br />

Pipeline Engineering<br />

incorporating<br />

The Journal of Pipeline Integrity<br />

Volume 8, No 3 • Third Quarter, 2009<br />

Contents<br />

Dr Mohamad J Cheaitani ....................................................................................................................................... 149<br />

Approaches for determining limit load and reference stress for circumferential embedded flaws in pipe girth welds<br />

Nigel S Kirk and Dipl-Ing Björn Dobberstein...................................................................................................... 167<br />

The Nord Stream Pipeline’s German landfall: the challenges ahead<br />

Dr Kimberly Cameron and Dr Alfred Pettinger .................................................................................................. 175<br />

Assessing pipeline integrity using fracture mechanics and currently available inspection tools<br />

Navid Nazemi, Sara Kenno, and Sreekanta Das ................................................................................................... 183<br />

Behaviour of wrinkled linepipe subjected to internal pressure and eccentric axial compression load<br />

Kenton Pike ............................................................................................................................................................ 191<br />

Advanced numerical modelling tools aid Arctic pipeline design<br />

Sample issue<br />

Etim S Udoetok and Anh N Nguyen .................................................................................................................... 195<br />

A disc pig model for estimating the mixing volumes between product batches in multi-product pipelines<br />

J P Pruiksma, H J Brink, H M G Kruse, and J Spiekhout .................................................................................. 205<br />

Soil reaction force at the head of the pipeline during the pull-back operation of horizontal directional drilling<br />

Stijn Hertelé, Rudi Denys, and Wim De Waele................................................................................................... 213<br />

Full range stress-strain relation modelling of pipeline steels<br />

Correction ............................................................................................................................................................... 223<br />

A practical approach to pipeline corrosion modelling: Part 2 – Short-term integrity forecasting<br />

❖ ❖ ❖<br />

OUR COVER PICTURE shows a typical landfall using a coffer dam to pull-in the pipe from offshore, one of the<br />

techniques being proposed by the designers of the Nord Stream pipeline project. The twin 48-in diameter, 1220-km<br />

long, pipelines will be the world’s longest subsea pipelines once fully operational in 2011. The first of a series of<br />

three papers describing aspects of the project is published on pages 167-173.


146<br />

1. Disclaimer: While every effort is made to check the<br />

accuracy of the contributions published in The Journal of<br />

Pipeline Engineering, Scientific Surveys Ltd and Clarion<br />

Technical Publishers do not accept responsibility for the<br />

views expressed which, although made in good faith, are<br />

those of the authors alone.<br />

2. Copyright and photocopying: © 2009 Scientific Surveys<br />

Ltd and Clarion Technical Publishers. All rights reserved.<br />

No part of this publication may be reproduced, stored or<br />

transmitted in any form or by any means without the prior<br />

permission in writing from the copyright holder.<br />

Authorization to photocopy items for internal and personal<br />

use is granted by the copyright holder for libraries and<br />

other users registered with their local reproduction rights<br />

organization. This consent does not extend to other kinds<br />

of copying such as copying for general distribution, for<br />

advertising and promotional purposes, for creating new<br />

collective works, or for resale. Special requests should be<br />

addressed to Scientific Surveys Ltd, PO Box 21, Beaconsfield<br />

HP9 1NS, UK, email: info@scientificsurveys.com.<br />

3. Information for subscribers: The Journal of Pipeline<br />

Engineering (incorporating the Journal of Pipeline Integrity)<br />

is published four times each year. The subscription price<br />

for 2009 is US$350 per year (inc. airmail postage). Members<br />

of the Professional Institute of Pipeline Engineers can<br />

subscribe for the special rate of US$175/year (inc. airmail<br />

postage). Subscribers receive free on-line access to all issues<br />

of the Journal during the period of their subscription.<br />

The Journal of Pipeline Engineering<br />

THE Journal of Pipeline Engineering (incorporating the Journal of Pipeline Integrity) is an independent, international,<br />

quarterly journal, devoted to the subject of promoting the science of pipeline engineering – and maintaining and<br />

improving pipeline integrity – for oil, gas, and products pipelines. The editorial content is original papers on all aspects<br />

of the subject. Papers sent to the Journal should not be submitted elsewhere while under editorial consideration.<br />

Authors wishing to submit papers should send them to the Editor, The Journal of Pipeline Engineering, PO Box 21,<br />

Beaconsfield, HP9 1NS, UK or to Clarion Technical Publishers, 3401 Louisiana, Suite 255, Houston, TX 77002, USA.<br />

Instructions for authors are available on request: please contact the Editor at the address given below. All contributions<br />

will be reviewed for technical content and general presentation.<br />

The Journal of Pipeline Engineering aims to publish papers of quality within six months of manuscript acceptance.<br />

Notes<br />

4. Back issues: Single issues from current and past volumes<br />

(and recent issues of the Journal of Pipeline Integrity) are<br />

available for US$87.50 per copy.<br />

5. Publisher: The Journal of Pipeline Engineering is<br />

published by Scientific Surveys Ltd (UK) and Clarion<br />

Technical Publishers (USA):<br />

Scientific Surveys Ltd, PO Box 21, Beaconsfield<br />

HP9 1NS, UK<br />

tel: +44 (0)1494 675139<br />

fax: +44 (0)1494 670155<br />

email: info@scientificsurveys.com<br />

web: www.j-pipe-eng.com<br />

www.pipemag.com<br />

Editor and publisher: John Tiratsoo<br />

email: jtiratsoo@j-pipe-eng.com<br />

Sample issue<br />

v v v<br />

Clarion Technical Publishers, 3401 Louisiana,<br />

Suite 255, Houston TX 77002, USA<br />

tel: +1 713 521 5929<br />

fax: +1 713 521 9255<br />

web: www.clarion.org<br />

Associate publisher: BJ Lowe<br />

email: bjlowe@clarion.org<br />

6. ISSN 1753 2116<br />

www.j-pipe-eng.com<br />

is available for subscribers


3rd Quarter, 2009 147<br />

Editorial<br />

Pipeline outlook: it’s not all gloom<br />

FOR A CONSIDERABLE time, headlines around the<br />

world have been focussing on the economic situation<br />

and the immense difficulties the downturn has imposed on<br />

individuals and companies in many industrial sectors.<br />

There is no denying the fact that huge changes are under<br />

way, and the world as a whole is having to readjust to the<br />

new regime that these changes are introducing. Many<br />

might therefore see this as a poor choice of time at which<br />

to launch a new industry publication. The hydrocarbons’<br />

pipeline industry is, however, particularly buoyant currently,<br />

and forecasts for the next five years are tremendously<br />

positive for both on- and offshore pipeline construction.<br />

Fuelled, of course, by the world’s burgeoning need for<br />

energy, gas pipeline projects have never been of greater<br />

significance, and are focussing on transporting reserves<br />

from more technically-challenging areas than ever before.<br />

Oil, too, is in high demand, and requires transport over<br />

longer distances and through terrain of increasing<br />

complexity and environmental sensitivity.<br />

Two recently-published authoritative reports highlight the<br />

strength of the pipeline industry and its forecast growth<br />

over the next few years. Looking offshore, London-based<br />

Infield Energy Analysts’ fourth edition of its Global<br />

perspectives pipelines and control lines update report provides<br />

an in-depth, independent analysis of the global offshore<br />

pipeline and control line market sectors from 2004 to<br />

2013. The report covers pipelines of all lengths and diameters<br />

including SURF flowlines, trunklines, and conventional<br />

pipelines, as well as all control lines, including<br />

communication, power line, seismic cable,<br />

telecommunication, and umbilicals.<br />

With the long-term prospect of increasing global energy<br />

demand, securing future energy supplies has become a<br />

common global issue. As the report points out, for those<br />

countries with dwindling production rates or low<br />

hydrocarbon reserves, the pressure for energy security is set<br />

to increase, while those with abundant reserves will strive<br />

to attract investment to enable adequate development to<br />

meet both domestic and foreign energy demands. As a<br />

consequence of these demands, there has been growth in<br />

the offshore oil and gas industry since 2004. A lower price<br />

outlook and lack of available credit have certainly affected<br />

the future development of reserves, but growth in the<br />

industry is still expected to continue.<br />

Pipeline and control line installation trends have mirrored<br />

those of the wider offshore industry, which is unsurprising,<br />

considering their crucial role within the offshore oil and<br />

gas infrastructure, and Infield forecasts the total pipelines<br />

and control lines capital expenditure to exceed $265bn<br />

over the period 2009-2013. This equates to 103,435km of<br />

lines being installed, of which 81,293km will be pipelines<br />

and 22,142km will be control lines. Combined, these<br />

represent an increase of 68% in installations over that’s of<br />

the previous five years. The forecasted increase will be<br />

dominated by growth in the pipeline market, with a<br />

significantly-slower growth in the control line market. A<br />

considerable percentage of the forecast pipeline expenditure<br />

is related to far-advanced trunklines, many un-connected to<br />

specific field development projects and, as such, key<br />

infrastructure development.<br />

Sample issue<br />

Infield says that the next five years indicate a change in<br />

market demographics, in which all pipeline segments will<br />

hold fairly equal shares of the installation market. This<br />

follows a period in which conventional pipelines dominated<br />

the installation market, highlighting the industry’s<br />

historically-favoured shallow-water developments. However,<br />

as shallow-water production rates fall, the industry has<br />

sought to discover and develop deeper-water reserves. As a<br />

consequence, subsea construction, umbilical, riser, and<br />

flowline (SURF) installations have grown in the previous<br />

five-year period, and are set to continue increasing in the<br />

forecast period. The largest pipeline installation growth is<br />

expected in the trunk/export lines sector, further<br />

characterizing the increasing demand to secure a diversified<br />

mix of future energy supplies.<br />

Whilst growth is still expected in the control line market,<br />

a decline in communication line installations and slower<br />

growth in the power line sector will be seen compared to<br />

the previous five-year period. The overall control line<br />

growth will predominantly be driven by an increase in<br />

umbilical and bundled pipeline installations, both of which


148<br />

imply the continuing trend to replace installation of single<br />

control lines with combined multiple line installations.<br />

Overall however, as the report highlights, the future for the<br />

pipeline and control line industry is expected to be strong<br />

with a variety of water depths, project sizes, and locations<br />

expected over the next five years.<br />

As far as the onshore industry is concerned, Douglas-<br />

Westwood’s report points out that around 157,000km of<br />

pipelines are planned up to 2013, at a cost of over $178<br />

billion, which is a 15% increase in length installed and a<br />

27% increase in investment relative to the previous fiveyear<br />

period. Gas pipelines will make up 95,341km, and oil<br />

pipelines 35,034km, of the total, in which LNG<br />

transportation will also play a significant role. Some specific<br />

projects that will contribute to these totals are featured in<br />

this issue, among which are reviews of various aspects of the<br />

twin 1220-km long, 48-in diameter, Nord Stream pipelines,<br />

which will be the longest subsea pipelines in the world<br />

when commissioned in 2011 and 2012. The issues<br />

surrounding the design and engineering of pipelines in the<br />

Arctic, a region that is becoming of great significance, are<br />

also becoming increasingly high-profile. As a testament to<br />

this, the proposed pipeline to bring Alaskan gas to markets<br />

in the southern United States is expected to cost over $30<br />

billion, and the latest published cost estimate for the<br />

Mackenzie gas pipeline from the Mackenzie Delta area is<br />

$16 billion.<br />

Publishers merge: new industry<br />

magazine launched<br />

TWO OF THE LEADING providers of technical and<br />

business information for the pipeline industry,<br />

Scientific Surveys Ltd and Great Southern Press (GSP),<br />

have merged. The newly formed company has a global<br />

The Journal of Pipeline Engineering<br />

scope, with head offices in the UK and the Asia Pacific, as<br />

well as a strong presence in Houston and contacts<br />

throughout South America, Europe and the Middle East.<br />

The companies will, together, continue to produce their<br />

full range of pipeline products, and have already launched<br />

a new print magazine, <strong>Pipelines</strong> <strong>International</strong>, which is<br />

supported by a comprehensive online presence (see<br />

www.pipelinesinternational.com) and will reflect the<br />

diversity of the pipeline industry world-wide. As part of the<br />

merger, the new business division will – in association with<br />

Clarion of Houston – continue publication of the Journal<br />

of Pipeline Engineering, along with developing the<br />

comprehensive database of technical papers at<br />

www.pipedata.com, and expanding its involvement with<br />

high-quality training courses and events.<br />

Formation of the new division is intended to build-upon<br />

the reputations of UK-based Scientific Surveys and<br />

Australian GSP in providing technical information, and<br />

the strong partnership with Clarion in Houston will be<br />

developed and enhanced. In addition to the world-renowned<br />

Pipeline Pigging & Integrity Management Conference and<br />

Exhibition in Houston each February, new major<br />

conferences and exhibitions, as well as training, will be<br />

planned elsewhere, including in the Asia-Pacific and Middle<br />

Eastern regions. The partnership will also strengthen the<br />

companies’ other existing products, for example by providing<br />

greater resources and technical expertise to The Australian<br />

Pipeliner magazine.<br />

Sample issue


3rd Quarter, 2009 149<br />

Approaches for determining<br />

limit load and reference stress<br />

for circumferential embedded<br />

flaws in pipe girth welds<br />

by Dr Mohamad J Cheaitani<br />

TWI Ltd, Abington, Cambridge, UK<br />

THIS PAPER PROVIDES an evaluation of approaches for determining limit load (and equivalent reference<br />

stress) for use in failure-assessment diagram (FAD)-based fracture assessment of circumferential<br />

embedded flaws in pipe girth welds. Three-dimensional elastic plastic finite-element analyses have been<br />

conducted on pipe models containing typical circumferential embedded flaws and subject to global bending<br />

loads. ‘J-based’ limit loads (and equivalent reference stresses) and global collapse limit loads have been<br />

determined from the finite-element analyses and used to evaluate existing standard flat-plate solutions,<br />

including those in BS 7910 and R6. A general approach for determining the limit load (and the equivalent<br />

reference stress) is presented. This approach is consistent with both the finite-element results and<br />

reference stress J-estimation scheme and, consequently, allows the development of improved assessment<br />

models.<br />

THE USE OF AN engineering critical assessment (ECA)<br />

to derive flaw acceptance criteria for pipeline girth<br />

welds allows the maximum tolerable size of surface and<br />

embedded circumferential planar flaws to be determined<br />

on a fitness-for-purpose basis. A typical ECA involves<br />

assessing the significance of such flaws with regard to<br />

failure mechanisms, including fracture, which the pipeline<br />

may experience during construction, commissioning, and<br />

service. The most commonly used approach to assess the<br />

significance of flaws with regard to fracture and plastic<br />

collapse is the failure-assessment diagram (FAD), which is<br />

based on the reference stress J-estimation scheme [1, 2] . An<br />

FAD-based assessment involves the calculation of a fracture<br />

parameter (K r , d r , or J r ) and a plastic collapse parameter,<br />

This paper was presented at the Pipeline Technology Conference held<br />

in Ostend, Belgium, on 12-14 October, 2009, and organized by the<br />

University of Gent, Belgium, and Technologisch Instituut vzw, Antwerp,<br />

Belgium.<br />

Author’s contact details:<br />

tel: +44 (0)1223 899000<br />

email: mohamad.cheaitani@twi.co.uk<br />

Sample issue<br />

L r , Fig.1. The fracture parameter characterizes the proximity<br />

to fracture under linear elastic conditions. The plasticcollapse<br />

parameter characterizes the proximity to failure by<br />

yielding mechanisms and is defined as either the ratio of<br />

applied load to the limit load or, equivalently, the ratio of<br />

the reference stress to the yield stress. There is a unique<br />

relationship between a reference stress and a limit load<br />

which enables one to be defined if the other is known:<br />

specifically, a limit load is inversely proportional to the<br />

corresponding reference stress as follows:<br />

Applied load Reference stress<br />

=<br />

Limit load Yield stress<br />

The work described in this paper concerns the development<br />

of a reference-stress (or limit-load) model for use in FADbased<br />

fracture assessments of circumferential embedded<br />

girth weld flaws such as that shown in Fig.2. This work is<br />

considered necessary since existing reference-stress solutions<br />

for embedded flaws are not consistent with the referencestress<br />

solutions that are typically used to assess surface<br />

(1)


150<br />

K r (fracture parameter)<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

flaws, because the approaches used in their derivation are<br />

different. Whereas there are several reference-stress models<br />

for surface-breaking flaws, including some intended<br />

specifically for circumferential flaws in pipe sections, there<br />

are fewer reference-stress models for embedded flaws, and<br />

these are intended for flaws in flat plates.<br />

One of the most widely used solutions is that in BS<br />

7910:2005 [1], which assumes that plastic collapse occurs<br />

locally when the net section stress on a small area<br />

surrounding the embedded flaw reaches the yield or flow<br />

strength of the material. It also assumes that tensile loads<br />

are applied through a pin-jointed coupling, i.e. that there<br />

is negligible bending restraint. The use of this solution to<br />

assess embedded flaws could lead to overly conservative<br />

results, which may in some cases be counter-intuitive: for<br />

example, for a given applied loading, the maximum tolerable<br />

length for an embedded flaw (whose ligament is greater<br />

than or equal to the height of one weld pass) is smaller than<br />

that for a surface-breaking flaw of the same height.<br />

The paper focuses on the evaluation of the existing referencestress<br />

(and the equivalent limit-load) solutions for embedded<br />

flaws and the development of improved models. The existing<br />

solutions are evaluated using data generated from elasticplastic<br />

finite-element analyses of pipe models containing<br />

typical circumferential embedded flaws.<br />

Scope of work<br />

Acceptable<br />

Not acceptable<br />

0<br />

0 0.2 0.4 0.6 0.8 1 1.2<br />

L r (yielding or collapse parameter)<br />

The scope of work includes three-dimensional elastic-plastic<br />

finite-element analyses of pipe models containing typical<br />

circumferential embedded flaws. The pipe models were<br />

loaded by pure bending, which is the pre-dominant loading<br />

mode during pipeline installation, and were perfectly aligned<br />

across the section containing the flaw; the same tensile<br />

properties were assigned to the parent and weld metals.<br />

Data from the finite-element analyses were used to review<br />

and evaluate a number of methods for determination of the<br />

reference stress (or limit load) for embedded flaws, including<br />

the conventional methods recommended in BS 7910 [1]<br />

The Journal of Pipeline Engineering<br />

and R6 [2], and novel methods. It is shown that none of the<br />

existing codified reference-stress (or limit-load) solutions<br />

agree well with the results obtained from the finite-element<br />

analyses. Therefore, a new approach using a novel method<br />

for definition of reference stress, which is fully consistent<br />

with the findings from the finite-element analyses, is<br />

proposed.<br />

The remainder of this paper consists of the following<br />

sections: reference stress J-estimation scheme; approaches<br />

for determining reference stresses (and limit loads) in FADbased<br />

assessments; codified reference-stress (and limit-load)<br />

solutions for embedded flaws; approach adopted for<br />

determining J-based limit loads (M J ); scope of (and results<br />

from) finite-element analyses; comments on global collapse<br />

and J-based limit loads; comparison of flat-plate solutions<br />

with J-based solutions; plastic strain in ligaments; and<br />

summary and conclusions.<br />

Sample issue<br />

Reference stress<br />

J-estimation scheme<br />

The evaluation of reference-stress models is performed<br />

within the context of the reference stress J-estimation<br />

scheme [2], which is defined below.<br />

J is estimated from the following expressions, which<br />

correspond to the material-specific FAD (BS 7910 Level<br />

2B/3B and R6 Option 2):<br />

Je<br />

J = 2<br />

fL ( )<br />

(2)<br />

and<br />

r<br />

3<br />

−0.5<br />

⎛Eεref Lrσ<br />

⎞<br />

y ⎟<br />

f( Lr<br />

) =<br />

⎜ ⎟<br />

⎜ + ⎟<br />

⎜⎝Lrσy2Eε ⎟ ref ⎠<br />

Fig.1. A typical failure-assessment<br />

diagram (FAD).<br />

where, for a given applied bending moment, M:<br />

(3)


3rd Quarter, 2009 151<br />

L r s y is the reference stress (denoted as s ref );<br />

e ref is the reference strain corresponding to s ref and<br />

determined from the stress-strain curve of the<br />

material;<br />

s y is the yield or 0.2% proof strength of the material;<br />

and<br />

E is Young’s modulus.<br />

J e is the elastic value of J at an applied moment M,<br />

determined from data obtained at an applied moment M o ,<br />

as follows:<br />

⎛ M ⎞<br />

= ⎜<br />

⎟<br />

⎜⎜⎝ ⎟<br />

⎠⎟<br />

Je Jo Mo<br />

2<br />

or using the equivalent expression<br />

where:<br />

2<br />

⎛ ⎞<br />

M<br />

e = ⎜ ⎟<br />

o⎜⎟<br />

⎜⎜⎝σ ⎟<br />

o ⎠<br />

J J σ<br />

J o is the elastic value of J determined at M o (for example,<br />

from an elastic finite-element analysis); and<br />

s M and s o are the elastic bending stresses on the pipe<br />

OD corresponding to M and M o , respectively,<br />

determined using elastic section properties.<br />

An alternative estimate of the elastic value of J may be<br />

determined as follows:<br />

2<br />

⎛ ⎞<br />

M1<br />

e = ⎜ ⎟<br />

o⎜⎟<br />

⎜⎜⎝ σ<br />

⎟<br />

o ⎠<br />

J J σ<br />

which is similar to Equn 5 but uses the actual elastic-plastic<br />

stress on the pipe OD (denoted s ), rather than s . In this<br />

M1 M<br />

case, J is not proportional to M e 2 .<br />

The parameter L r , which characterizes the proximity to<br />

plastic collapse, can be expressed as follows:<br />

Lr ML<br />

(4)<br />

(5)<br />

(6)<br />

M<br />

= (7)<br />

or alternatively as:<br />

σref<br />

Lr<br />

= (8)<br />

σy<br />

where M is the limit load.<br />

L<br />

The reference stress J-estimation scheme could also be<br />

applied using alternative expressions of J, such as that<br />

which corresponds to simplified FADs in BS 7910 and R6.<br />

There are a number of approaches for determining M L (and<br />

Fig.2. Idealized curved elliptical embedded flaw in a pipe<br />

(located at 12 o’clock position).<br />

the corresponding s ref ), which are described in the following<br />

sections.<br />

Approaches for determining<br />

limit loads for FAD assessments<br />

The limit load (or plastic collapse) required for the<br />

calculation of reference stress and the parameter L r is one<br />

of the most important elements of an FAD-based assessment<br />

since it serves two functions:<br />

• it ensures that the limit load of the component<br />

containing the flaw under consideration is not<br />

exceeded;<br />

• it ensures that the relationship between elasticplastic<br />

driving force and proximity to plastic collapse<br />

is consistent with the relationship implied by the<br />

failure-assessment curve.<br />

Sample issue<br />

Two potential plastic-collapse modes of a flawed component<br />

can be identified:<br />

• Local collapse: corresponds to failure, by yielding<br />

mechanisms, of the ligament adjacent to the flaw.<br />

This is deemed to occur when the stress in the<br />

remaining ligament reaches the yield strength of the<br />

material. With regard to a circumferential partthickness<br />

flaw in a girth weld (surface-breaking or<br />

embedded), the significance of the circumferential<br />

extent of the remaining ligament on either side of<br />

the flaw is not well defined in any of the existing<br />

standards. Another source of uncertainty is whether<br />

bending of the section containing the flaw is<br />

restrained or not. In the absence of bending restraint,<br />

secondary bending stresses arise due to eccentric<br />

loading. This is caused by movement of the neutral


152<br />

axis, due to the existence of the flaw, compared with<br />

the unflawed condition.<br />

• Global collapse: corresponds to failure, by yielding<br />

mechanisms, of the whole section containing the<br />

flaw. This is deemed to occur when the global<br />

deformation, displacement and/or rotation, of the<br />

section become unbounded. Global collapse occurs<br />

at a higher load than that corresponding to local<br />

collapse.<br />

Most codified limit-load or reference-stress solutions for<br />

part-thickness (surface or embedded) flaws are based on the<br />

local-collapse approach. Although some standards, such as<br />

R6 [2] also provide solutions based on the global-collapse<br />

approach, further checks, such as against finite-element<br />

analyses, may be required to verify that such solutions<br />

provide safe assessments.<br />

An alternative method for determining limit loads, which<br />

requires J data from finite-element analyses, consists of<br />

defining the limit load such that it is consistent with the J<br />

data and the reference stress J-estimation scheme represented<br />

by Equn 2. The main benefit of this approach is that it does<br />

not require the analyst to specify in advance whether a localor<br />

global-collapse model is more suitable. The limit load is<br />

found by solving Equns 2 and 3 using J results from an<br />

elastic-plastic finite-element analysis of the flawed<br />

component. A simple version of this approach is<br />

recommended in Section B.6.4.3(e) of API 579 [3] and in<br />

a slightly different form in Section B.1.89 of API-579-1/<br />

ASME-FFS-1 [4] as follows:<br />

L<br />

r<br />

t<br />

Fig.3. Idealized elliptical embedded flaw in a flat plate, used in BS 7910 (BSI, 2005).<br />

P<br />

=<br />

P<br />

ref<br />

where P ref is determined from the following relationship:<br />

(9)<br />

The Journal of Pipeline Engineering<br />

J 0.002E 1 ⎛ 0.002E⎞<br />

⎟<br />

= 1+ +<br />

⎜<br />

⎜1<br />

⎟<br />

Je σy 2<br />

⎜ + ⎟<br />

⎜⎝ σ ⎟ y ⎠<br />

P= Pref<br />

−1<br />

(10)<br />

where J is the total value of J determined from an elasticplastic<br />

analysis of the flawed component; J e is the elastic J<br />

determined from an elastic analysis by, for example, using<br />

Equns 4 or 5; P is a characteristic applied load (or stress)<br />

such as axial force, bending moment, or a combination<br />

thereof; and P ref is the reference load (or stress) defined as<br />

the load at which the ratio J/J e reaches the value defined by<br />

Equn 10.<br />

If P ref is used to construct a BS 7910 Level 3C FAD (with L r<br />

defined according to Equn 9), it will intersect the<br />

corresponding BS 7910 Level 2B/3B material-specific FAD,<br />

and give the same K r value, at L r = 1.0. Thus, the limit load<br />

P ref is defined in a manner which is consistent with the Level<br />

2B/3B material-specific FAD, at least at L r = 1.0. In this<br />

case, the limit load may depend on the strain-hardening<br />

characteristics of the material.<br />

Sample issue<br />

Codified reference-stress<br />

solutions for embedded flaws<br />

General<br />

Whereas there are several well-established reference-stress<br />

solutions for circumferential surface flaws in pipe girth<br />

welds, there are no such solutions for circumferential<br />

embedded flaws. Consequently, most analysts use referencestress<br />

solutions originally derived for flat plates to assess<br />

circumferential embedded flaws in pipe girth welds. The<br />

most widely used of these solutions are the reference-stress<br />

equations given in BS 7910 [1] and R6 [2]. These are given<br />

in terms of the membrane stress, P m , and through-wall


3rd Quarter, 2009 153<br />

Fig.4. Idealized<br />

rectangular embedded<br />

flaw in a flat plate<br />

subjected to tension<br />

and/or bending loading,<br />

used in R6 (BEGL,<br />

2001).<br />

bending stress, P b . However, given that in a thin-walled<br />

pipe loaded by bending P b is very small compared with P m ,<br />

and in a pipe loaded by tension P b is equal to zero, the<br />

reference-stress equations considered below are given<br />

assuming that P b is equal to zero. Note that references to<br />

equations, sections, and figures in codes and standards are<br />

shown in italics.<br />

BS 7910<br />

The embedded flaw reference stress equation in Equation<br />

P4 of BS 7910 [1] is based on local collapse and assumes<br />

that tensile loads are applied through pin-jointed coupling,<br />

i.e. that there is negligible bending restraint. The equation<br />

was derived by Willoughby and Davey [5] assuming that the<br />

load-bearing area (or ligament) extends to the plate surfaces<br />

above and below the embedded flaw and has a length equal<br />

to the flaw length plus one plate thickness on either end of<br />

the flaw. The solution, with P b set at zero, is as follows:<br />

σ<br />

ref<br />

{ }<br />

⎡ 2 2<br />

2 4 α"<br />

p ⎤<br />

Pmα" + ⎢( Pmα" ) + Pm<br />

( 1 − α"<br />

) + ⎥<br />

⎢ t ⎥<br />

= ⎣ ⎦<br />

2 4 α"<br />

p<br />

( 1 − α"<br />

) +<br />

t<br />

0.5<br />

(11)<br />

where p is the ligament (the smallest distance between the<br />

flaw and the surface), t is the wall thickness, and<br />

2a<br />

α " =<br />

⎛ t ⎞⎟<br />

t ⎜<br />

⎜⎜⎝ 1+<br />

⎟<br />

c⎠⎟<br />

(12)<br />

where 2a is the flaw height and 2c is its length – see Fig.3<br />

(note that the wall thickness in BS 7910 is denoted as B).<br />

R6<br />

R6 [2] provides several reference stress solutions for<br />

embedded flaws in flat plates, which are based on local or<br />

global collapse with loads applied through pin-jointed<br />

coupling (i.e. pin loading) or fixed-grip loading conditions.<br />

The approach used to develop these solutions is described<br />

by Lei and Budden [6]. The solutions cater for flaws located<br />

fully or partially in the tensile stress zone (determined by<br />

the location of the neutral axis). Whilst this distinction is<br />

important when assessing flat plates, it is less significant for<br />

circumferential embedded flaws in pipe girth welds, which<br />

are almost always assumed to be in the tensile stress zone.<br />

Consequently, attention is focussed below on the latter<br />

condition. The solutions are expressed using the following<br />

parameters:<br />

N a c<br />

n<br />

Wt t W k<br />

y<br />

L<br />

off c<br />

L = , a= , b = , = , g =<br />

2 s t t<br />

Sample issue<br />

y<br />

where N L is the limit load and other parameters are illustrated<br />

in Fig.4. The limit-load solutions are given below in terms<br />

of the non-dimensional parameter n L .<br />

Global collapse, pin-loaded (IV.1.6.3-1):<br />

c1<br />

nL<br />

=<br />

2<br />

2ab + 4(<br />

ab)<br />

+ c<br />

1<br />

(IV.1.6.1-1 with l = 0) (13)<br />

2<br />

c = 1-8abk-4( ab)<br />

(IV.1.6.1-3) (14)<br />

1<br />

Valid for:<br />

1<br />

1<br />

> k³ 0 anda£ -k<br />

2<br />

2


154<br />

Stress (MPa)<br />

Global collapse, fixed-grip tension (IV.1.6.3-1 with k = 0):<br />

As above (Equns 13 and 14) but with k = 0 (the limit load<br />

does not depend on the crack position in the cross section).<br />

Local collapse, pin-loaded (IV.1.6.3.2), solution (a):<br />

d= t+ c, t1= t and W > d<br />

This solution is approached when yielding takes place<br />

across the whole loading-bearing area 2(t + c) t.<br />

n<br />

L<br />

600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

Stress (MPa)<br />

0<br />

0 1 2 3 4 5 6 7 8 9 10<br />

400<br />

300<br />

200<br />

100<br />

=<br />

⎛ αγ ⎞<br />

2 ⎜<br />

⎟ ⎟+ ⎜⎝1+ γ<br />

⎟ ⎟⎠ c1<br />

2<br />

⎛ αγ ⎞<br />

4 ⎜<br />

⎟ + c<br />

⎜<br />

1<br />

⎝1+ γ⎠<br />

⎟<br />

(IV.1.6.2-1 with l = 0) (15)<br />

2<br />

Strain %<br />

8αγ<br />

k ⎛ αγ⎞<br />

c1<br />

= 1− −⎜<br />

⎜ ⎟<br />

1 γ ⎜1 γ<br />

⎟ (IV.1.6.2-3) (16)<br />

+ ⎜⎝<br />

+ ⎠ ⎟<br />

Valid for:<br />

1<br />

1<br />

g<br />

> k ³ 0 and a£ - k and b<<br />

2<br />

2 1+<br />

g<br />

True stress R-O n=15<br />

Eng stress R-O n=15<br />

0.2% offset<br />

0<br />

0 0.2 0.4 0.6 0.8<br />

Strain %<br />

The Journal of Pipeline Engineering<br />

True stress R-O n=15<br />

Eng stress R-O n=15<br />

0.2% offset Fig.5. Stress-strain curve used:<br />

Local collapse, pin-loaded (IV.1.6.3.2), solution (b):<br />

d= t+ c, t1= t and W > d<br />

Sample issue<br />

This is approached when yielding spreads through the<br />

smallest ligament along the plate thickness. Solution (b) is<br />

always less than or equal to Solution (a).<br />

⎛ 2α<br />

⎞<br />

d= t ⎜<br />

⎜1 ⎟<br />

⎜⎝<br />

− ⎟+<br />

c, t1= t( 1− 2k)<br />

and W > d<br />

1−2k⎠⎟ n<br />

L<br />

⎛ 2α<br />

⎞<br />

⎜<br />

⎜1 ⎟<br />

⎜⎝<br />

− ⎟(<br />

1+<br />

γ)<br />

1−2k⎠⎟ =<br />

⎛ 2α<br />

⎞⎟<br />

⎜<br />

⎜⎜⎝ 1− ⎟+<br />

γ<br />

1−2k⎟⎠ Local collapse, fixed-grip tension:<br />

1+ g( 1-2a) nL =<br />

1+<br />

g<br />

(a - top) curve up to 10% strain;<br />

(b - bottom) curve up to 0.8% strain.<br />

(17)<br />

(18)<br />

(19)


3rd Quarter, 2009 155<br />

Approach adopted for determining<br />

J-based limit loads (M J )<br />

As evident from the above, there are numerous approaches<br />

for the definition of limit load including local collapse,<br />

global collapse, or J-based methods associated with the<br />

reference stress J-estimation scheme. Each of these<br />

approaches can be applied using a number of options. For<br />

example, local collapse can be defined based on a somewhat<br />

arbitrarily postulated load-bearing area, and either fixedended<br />

or pin-ended supports (reflecting whether bending<br />

of the section containing the flaw is restrained or not). In<br />

practice it is difficult to assess the suitability of these<br />

solutions for use in specific applications without additional<br />

information.<br />

The suitability of the above approaches to assess<br />

circumferential embedded flaws in pipe girth welds is<br />

evaluated in the following sections using data from elasticplastic<br />

finite-element analyses of pipe models containing<br />

typical circumferential embedded flaws. A reference stress<br />

(or limit load) is deemed to be adequate if it allows the<br />

effective elastic-plastic crack driving force expressed in<br />

terms of J integral (or J), to be determined using the FADbased<br />

reference stress J-estimation scheme, with a reasonable<br />

degree of accuracy compared to finite-element results.<br />

Therefore, the meaning of the term ‘limit load’ is extended<br />

to include any load that is used to define the reference stress<br />

in the context of a FAD calculation. Within this framework,<br />

the above approaches are assessed against a J-based limit<br />

load, which for a pipe subjected to bending is referred to as<br />

M J . This is determined using an approach somewhat similar<br />

to that of API-579-1/ASME-FFS-1 in that the limit load is<br />

defined to ensure consistency between the Level 3C FAD<br />

and the Level 2B/3B material-specific FAD (Equn 3).<br />

However, unlike the API 579 model which achieves this<br />

consistency at L r = 1.0, M J is determined such that the Level<br />

3C FAD matches the Level 2B/3B FAD at a range of L r and<br />

applied strain values. For each model, M J is determined for<br />

each load increment (in the finite-element analysis) as<br />

follows:<br />

• L r is determined by solving Equns [2] and [3]<br />

• M J is determined as M/L r (according to Equn 7),<br />

where M is the applied moment.<br />

This approach allows M J to be determined for every load<br />

increment and, consequently, allows M J to be plotted as a<br />

function of any loading parameter, including applied<br />

moment, remote applied stress/strain or L r .<br />

In theory, M J may depend on the position along the crack<br />

front. In the present work, attention has been restricted to<br />

solutions in the centre of the crack, which is usually the<br />

position of maximum J.<br />

Finite-element analyses<br />

Three-dimensional elastic-plastic finite-element analyses<br />

were conducted on pipe models containing a range of<br />

embedded flaw geometries. Four pipe geometries with pipe<br />

radius to thickness ratios in the range 5 to 20 were<br />

considered: in this paper, only results from Series E1 (pipe<br />

outside diameter = 400mm, wall thickness = 20mm), which<br />

were conducted using the stress-strain curve shown in Figs<br />

5a and 5b, are reported. The stress-strain curve was<br />

constructed from the following Ramberg-Osgood powerlaw<br />

hardening relationship:<br />

⎛ n<br />

⎛ ⎞ ⎞<br />

Y ⎜ α⎜<br />

⎟ ⎟<br />

⎜ ⎜ ⎟<br />

E σY σ<br />

⎟ ⎟<br />

⎜⎝<br />

⎜ ⎜ ⎜⎝ ⎟<br />

Y ⎠ ⎠ ⎟<br />

σ σ σ<br />

e=<br />

+ ⎟<br />

where:<br />

e = true strain<br />

s = true stress<br />

s Y = 0.2% proof strength (= 400MPa)<br />

E = Young’s modulus (200,000MPa)<br />

n = strain-hardening exponent (= 15)<br />

a = 0.002/e Y (= 1.0)<br />

e Y = s Y /E<br />

(20)<br />

Each finite-element model contained an elliptical embedded<br />

flaw oriented in the circumferential direction and contained<br />

within a plane perpendicular to the pipe axis. The flaws<br />

considered had a height in the range 3 to 9mm, were<br />

located at 1.5 to 14mm from the pipe inside surface, and<br />

had a length, along the ellipse major axis, in the range 25<br />

to 250mm. The flaws were located at the 12 o’clock<br />

position to ensure that they were subjected to the largest<br />

tensile stresses when the model was loaded by bending. The<br />

models were loaded by pure bending, which is the<br />

predominant loading mode during pipeline installation. In<br />

all models, the pipes were perfectly aligned across the<br />

section containing the flaw and the same tensile properties<br />

were assigned to the finite elements representing parent<br />

and weld metals. The dimensions of the pipes and flaws<br />

considered within Series E1 are summarized in Table 1.<br />

Sample issue<br />

All the analyses were performed using ABAQUS Version<br />

6.6 [7] using a small strain formulation, which is considered<br />

to be acceptable at least up to applied strains in the range<br />

0.5 to 1%. The finite-element meshes consist entirely of<br />

type C3D20R elements, which are 20-noded, quadratic,<br />

three-dimensional elements. As the flaw in the pipe is<br />

symmetric about the vertical plane and the applied load is<br />

also symmetric about this plane, it is only necessary to<br />

model half the pipe.<br />

In addition to analyses using the above-mentioned stressstrain<br />

model, analyses were performed on a number of pipe<br />

models using an elastic-perfectly plastic stress-strain model<br />

to determine a global collapse limit load, denoted M FEA .


156<br />

Results<br />

J-based limit loads -<br />

dependence on applied load<br />

Analyses using the above stress-strain model enabled the Jintegral<br />

to be calculated as a function of the applied load<br />

along the crack front. The highest J values, which generally<br />

occurred at the middle of the crack front adjacent to the<br />

smaller of the two ligaments (below and above the flaw),<br />

were adopted.<br />

The Journal of Pipeline Engineering<br />

Pipe dimension<br />

s,<br />

mm<br />

Flaw<br />

dimension<br />

s,<br />

mm<br />

Outside<br />

diameter<br />

Wall<br />

thickne<br />

ss<br />

Height Length<br />

Ligame<br />

nt<br />

to<br />

ID<br />

Ligame<br />

nt<br />

OD<br />

E1BH3L50L1. 5M0<br />

400 20 3 50 1. 5 15.<br />

5<br />

E1BH3L50L3M0400 20 3 50 3 14<br />

E1BH3L50L14M0400 20 3 50 14 3<br />

E1BH3L50L6M0400 20 3 50 6 11<br />

E1BH3L50L9M0400 20 3 50 9 8<br />

E1BH6L50L3M0400 20 6 50 3 11<br />

E1BH6L50L11M0400 20 6 50 11 3<br />

E1BH9L50L3M0400 20 9 50 3 8<br />

E1BH6L50L6M0400 20 6 50 6 8<br />

E1BH6L50L1. 5M0<br />

400 20 6 50 1. 5 12.<br />

5<br />

E1BH3L25L3M0400 20 3 25 3 14<br />

E1BH3L100L3M0400 20 3 100 3 14<br />

E1BH3L200L3M0400 20 3 200 3 14<br />

E1BH3L250L3M0400 20 3 250 3 14<br />

Sample issue<br />

E1BH6L25L3M0400 20 6 25 3 11<br />

E1BH6L100L3M0400 20 6 100 3 11<br />

E1BH3L25L6M0400 20 3 25 6 11<br />

Table 1. Dimensions of pipe and flaw considered in finite-element analysis models (series E1).<br />

The J results were used to estimate the following J-based<br />

limit loads according to the approach described earlier:<br />

• M J2B E – consistent with Equn 2, representing the<br />

material-specific FAD of BS 7910 (Level 2B/3B)<br />

and R6 (Option 2) with J e estimated using the elastic<br />

pipe bending stress according to Equn 5.<br />

• M J2B EP – consistent with Equn 2, representing the<br />

material-specific FAD of BS 7910 (Level 2B/3B)<br />

and R6 (Option 2) with J e estimated using the<br />

elastic-plastic pipe bending stress in Equn 6.<br />

to


3rd Quarter, 2009 157<br />

The above results are shown in Figs 6 and 7 in terms of<br />

2 M /4ts R vs Lr and applied strain, respectively. Here,<br />

J2B EP y m<br />

t is the pipe thickness, R is the pipe mean radius, and s m y<br />

2 is the pipe yield strength. Plots of M /4ts R vs Lr and<br />

J2B E y m<br />

applied strain are not included since they have broadly<br />

similar shapes to those shown in Figs 6 and 7 (but M is J2B E<br />

approximately 4% higher than M ). J2B EP<br />

2 Fig.6. M /4tv R (Je based on elastic-plastic pipe bending stress) vs L .<br />

J2B EP y m<br />

r<br />

Sample issue<br />

2 Fig.7. M /4ts R (Je based on elastic-plastic pipe bending stress) vs remote strain (%).<br />

J2B EP y m<br />

Figures 6 and 7 indicate that M J2B EP increase slightly with L r<br />

(for L r > 1.0) and applied strain (for strains > 0.2%). This<br />

behaviour is believed to be due to a number of factors<br />

including:<br />

• Component loading (it can be shown that the load<br />

dependence of J-based limit load solutions is


158<br />

2 Fig.8. M x (s /s )/4ts R (Je based on elastic-plastic pipe bending stress) vs L .<br />

J2B EP M1 M y m<br />

r<br />

influenced by the loading considered and is more<br />

significant for components loaded by bending than<br />

by tension).<br />

• The fact that J-based limit loads are not true limit<br />

loads: M J2B E is lower than M FEA , which is a true<br />

global collapse limit load, by approximately 9 to<br />

16%; M J2B EP is lower than M FEA by approximately 13<br />

to 19%.<br />

• Approximations within the reference stress J<br />

estimation scheme.<br />

The fact that M (i.e. M and M ) increases with L and<br />

J J2B E J2B EP r<br />

applied strain implies that a single value of M is an optimal<br />

J<br />

solution only for the L or applied strain value at which it<br />

r<br />

is determined. Furthermore, if M is determined at L = 1,<br />

J r<br />

as recommended in API 579, the use of this solution to<br />

assess load conditions where L > 1 can lead to very<br />

r<br />

conservative results. Equally, if M is determined at a<br />

J<br />

relatively high L value, say at L = 1.2, the use of this<br />

r r<br />

solution to assess load conditions where L < 1.2 can lead<br />

r<br />

to non-conservative results.<br />

Such dependence of M J2B E and M J2B EP on L r and applied<br />

strain somewhat complicates their use to assess standard<br />

solutions and/or develop new equations for determining<br />

M J . However, further work has shown that this load<br />

dependence can be largely eliminated by using either of the<br />

following two approaches:<br />

• If the estimated M J for a given load level (for<br />

example, corresponding to a load increment in the<br />

finite-element analysis) is multiplied by s M1 /s M (the<br />

ratio of the elastic-plastic pipe stress to the elastic<br />

Sample issue<br />

The Journal of Pipeline Engineering<br />

pipe stress at the same load level), the product<br />

2 (M x (s /s ) /4ts R ) or (MJ2B x (s /s ) /<br />

J2B E M1 M y m<br />

EP M1 M<br />

2 4ts R ) is largely independent of Lr and applied<br />

y m<br />

strain and is nearly constant for L > 1 and applied<br />

r<br />

strain > 0.5%. The evidence is illustrated in Figs 8<br />

2 and 9, which show (M x (s /s ) /4ts R ) vs<br />

J2B EP M1 M y m<br />

L and applied strain, respectively. For a given<br />

r<br />

applied moment, the factor (s /s ) depends only<br />

M1 M<br />

on the pipe cross section and its stress-strain<br />

properties and, consequently, can be determined<br />

easily.<br />

• If the results are expressed as non-dimensional<br />

reference stress (determined as s ref /s M1 ), it is seen<br />

that this parameter is largely independent of L r and<br />

applied strain, see Fig.10. This is predictable since<br />

s ref /s M1 is inversely proportional to M J x (s M1 /s M ).<br />

Therefore the beneficial effects of multiplying M J by<br />

s M1 /s M (described above) are included within<br />

s ref /s M1 (here M J refers to both M J2B E and M J2B EP ).<br />

An example of the impact of the above approaches is<br />

illustrated in Fig.11 for E1BH3L50L3M0 (2a = 3mm, 2c =<br />

50mm, p = 3mm), which shows that the use of the correction<br />

factor (s M1 /s M ) or expressing the results in terms of s ref /s M1<br />

enables J to be estimated with greater accuracy. It can also<br />

be seen that J estimates obtained using the global collapse<br />

limit load, M FEA , are significantly lower than those from<br />

finite-element analysis.<br />

The above two approaches give comparable results, but the<br />

second approach is generally preferable since it is relatively<br />

simpler to apply, and is potentially easier to develop into a<br />

versatile assessment model (applicable to pipe loaded by<br />

either tension or bending).


3rd Quarter, 2009 159<br />

Comments on global<br />

collapse and J-based limit loads<br />

2 Fig.9. M x (s /s )/4ts R (Je based on elastic-plastic pipe bending stress) vs remote strain (%).<br />

J2B EP M1 M y m<br />

To facilitate evaluation of the limit loads considered in the<br />

previous section, M J2B E at an applied strain of 0.5% (denoted<br />

M J2B E 0.5% ) and M J2B EP at an applied strain of 0.5% (denoted<br />

M J2B EP 0.5% ) were determined. The results are given in nondimensional<br />

form in Table 2, which also includes M FEA (the<br />

global collapse load determined by finite-element analysis<br />

using an elastic perfectly-plastic stress-strain model). The<br />

following observations can be made:<br />

Sample issue<br />

Fig.10. Non-dimensional reference stress (s ref /s M1 ) with J e based on elastic-plastic pipe bending stress) vs remote strain %.<br />

• The M FEA results are insensitive to crack size and<br />

ligament height and are always greater than J-based<br />

limit loads. This implies that limit loads based on<br />

global collapse (such as M FEA ), could lead to nonconservative<br />

estimates of J (see also Fig.11).<br />

• J-based limit loads decrease as the height increases,<br />

the ligament decreases, or as the length increases,<br />

but changes in height appear to have greater<br />

influence than changes in ligament or length. For<br />

example, considering E1BH3L50L3M0 as a base


160<br />

J, N/mm<br />

80<br />

60<br />

40<br />

20<br />

case (height = 3mm, length = 50mm, ligament =<br />

3mm), the following is observed based on results in<br />

Table 2:<br />

• increasing the height from 3 to 6 and 9mm<br />

leads to reductions in M of 3.8 and<br />

J2B E 0.5%<br />

5.5%, respectively;<br />

• increasing the ligament from 3 to 6 and<br />

9mm, leads to increases in M of 1.8<br />

J2B E 0.5%<br />

and 2.3%, respectively;<br />

• increasing the length from 50 to 100 and<br />

200mm leads to reductions in M of<br />

J2B E 0.5%<br />

1.1 and 1.8%, respectively.<br />

• The J-based limit loads for flaws located at a given<br />

ligament from the ID are nearly equal to those for<br />

flaws located at the same ligament from the OD (for<br />

example E1BH3L50L3M0 and E1BH3L50L14M0).<br />

• M J2B E (with J e based on the elastic pipe bending<br />

stress) is higher than M J2B EP (with J e based on the<br />

elastic-plastic pipe-bending stress) by approximately<br />

4%. This is not surprising since in the former case,<br />

for a given load level, the elastic driving force (J e ) and<br />

f(L r ) are higher, L r is lower, and M J is higher, see<br />

Equns 2 and 7.<br />

Comparison of flat-plate solutions<br />

with J-based solutions<br />

To facilitate comparison of the J-based limit loads with the<br />

codified flat-plate solutions reviewed earlier, the M J results<br />

The Journal of Pipeline Engineering<br />

OD=400mm, t=20mm, e=0.0mm. Embedded flaw near ID: 2a=3.0mm, 2c=50.0mm, pi=3mm, 'E1BH3L50L3M0'<br />

J estimate based on σ ref/σ M1 determined at 0.5% strain<br />

J max (FEA)<br />

J estimate based on M J2B EP 0.5% x (σ M1/σ M)<br />

J estimate based on M J2B EP 0.5%<br />

J estimate based on Global collapse,<br />

0<br />

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0<br />

Remote strain %<br />

Fig. 11. J results for model E1BH3L50L3M0 (from FEA and based on estimates of limit moment).<br />

(M J2B E 0.5% and M J2B EP 0.5% ) were expressed in terms of the<br />

non-dimensional parameter s ref /s M1 , where s M1 is the elasticplastic<br />

pipe-bending stress, and compared with s ref /s M1<br />

determined from the flat-plate solutions (as 1/n L ). The<br />

results are given in Table 3. The same results are reproduced<br />

as the ratio of s ref from the flat-plate solutions to s ref from<br />

M J2B EP 0.5% in Table 4. The following conclusions are drawn:<br />

Sample issue<br />

• s ref /s M1 estimates based on flat-plate solutions and<br />

local collapse with pin loading exceed significantly<br />

s ref /s M1 based on M J2B E 0.5% and M J2B EP 0.5% . Therefore,<br />

flat-plate solutions based on local collapse with pin<br />

loading lead to conservative assessments according<br />

to the following ranking (given in order of decreasing<br />

conservatism):<br />

• R6, local collapse, solution (b), pin loading<br />

(Equn 18)<br />

• BS 7910, local collapse, solution (a), pin<br />

loading (Equn 11)<br />

• R6, local collapse, solution (a), pin loading<br />

(Equn 15)<br />

• Table 4 shows that the BS 7910 flat-plate s ref solution<br />

exceeds s ref based on M J2B EP 0.5% by a margin which<br />

varies depending on flaw size and ligament. The<br />

margin is higher for deeper and longer flaws. Except<br />

for the shortest flaw considered (length = 25mm)<br />

and shallow flaws located near the middle of the<br />

thickness (height = 3mm and ligament >= 6mm) the<br />

margin exceeds 5%. These results apply to the<br />

stress-strain curve considered (low work hardening).


3rd Quarter, 2009 161<br />

J e<br />

based on<br />

Equn 4 Equn<br />

5<br />

Source<br />

FEA<br />

global<br />

collapse<br />

M / F EA<br />

2<br />

4tσ R y m<br />

Table 2. Non-dimensional limit-load results from finite-element analysis, and ligament plastic strain.<br />

Note: (1) Maximum plastic strain (%) on surface (ID or OD) nearest to the flaw (remote strain on OD = 1%).<br />

Limited results, not reported in this paper, indicate<br />

that the margin is lower for high work hardening<br />

materials. It should be noted that there are other<br />

sources of conservatism in BS 7910 assessments of<br />

circumferential embedded flaws in pipes, which<br />

include the equations used to estimate the stress<br />

intensity factor (intended for flaws in flat plates).<br />

FEA<br />

@ 0.<br />

5%<br />

strain<br />

M J 2B<br />

E 0.<br />

5%<br />

2<br />

/<br />

4tσ R y m<br />

FEA<br />

@ 0.<br />

5%<br />

strain<br />

M J 2B<br />

EP<br />

0.<br />

5%<br />

2<br />

/<br />

4tσ R y m<br />

Ligament<br />

plastic<br />

strain<br />

( % )<br />

( 1)<br />

E1BH3L50L1. 5M0<br />

0. 999<br />

0. 872<br />

0. 841<br />

2.<br />

8<br />

E1BH3L50L3M00. 999<br />

0. 888<br />

0. 854<br />

1.<br />

9<br />

E1BH3L50L14M00. 999<br />

0. 890<br />

0. 854<br />

2.<br />

3<br />

E1BH3L50L6M00. 999<br />

0. 904<br />

0. 867<br />

1.<br />

1<br />

E1BH3L50L9M00. 999<br />

0. 909<br />

0. 870<br />

1.<br />

1<br />

E1BH6L50L3M00. 998<br />

0. 855<br />

0. 825<br />

6.<br />

0<br />

E1BH6L50L11M00. 999<br />

0. 853<br />

0. 822<br />

7.<br />

6<br />

E1BH9L50L3M00. 995<br />

0. 840<br />

0. 811<br />

11.<br />

5<br />

E1BH6L50L6M0- 0. 872<br />

0. 839<br />

2.<br />

4<br />

E1BH6L50L1. 5M0<br />

- 0. 848<br />

0. 819<br />

8.<br />

6<br />

E1BH3L25L3M01. 000<br />

0. 901<br />

0. 865<br />

1.<br />

6<br />

E1BH3L100L3M00. 999<br />

0. 878<br />

0. 846<br />

2.<br />

2<br />

E1BH3L200L3M0- 0. 872<br />

0. 841<br />

2.<br />

4<br />

Sample issue<br />

E1BH3L250L3M0- 0. 870<br />

0. 839<br />

2.<br />

5<br />

E1BH6L25L3M0- 0. 880<br />

0. 847<br />

3.<br />

6<br />

E1BH6L100L3M0- 0. 832<br />

0. 805<br />

8.<br />

9<br />

E1BH3L25L6M0- 0. 916<br />

0. 876<br />

1.<br />

0<br />

• The majority of s ref /s M1 estimates based on flatplate<br />

solutions and global collapse (with the plate<br />

width assumed equal to half the pipe’s mean<br />

circumference) are lower than s ref /s M1 based on<br />

M J2B E 0.5% and M J2B EP 0.5% . Therefore, flat-plate<br />

solutions based on global collapse can lead to nonconservative<br />

assessments. However, it can be shown


162<br />

Table 3. s ref /s M1 estimates from J-based limit loads and flat-plate solutions.<br />

Notes:<br />

(1) K based on elastic stress, L uses elastic-plastic stress and is J-based at 0.5% strain.<br />

r<br />

(2) K based on elastic-plastic stress, L uses elastic-plastic stress and is J-based at 0.5% strain.<br />

r<br />

(3) Collapse of load bearing area around crack front.<br />

(4) All plate solutions: width = p x mean pipe radius.<br />

that modifying the R6 pin-loading model (Equn 13)<br />

by adjusting the plate width can lead to solutions<br />

that agree well with s ref /s M1 based on M J2B E 0.5% and<br />

M J2B EP 0.5% .<br />

It may be inferred from the above results that the<br />

conventional definition of local collapse solutions for flat<br />

plates, assuming that the load-bearing area extends one<br />

plate thickness at either side of the flaw, leads to<br />

overestimating the reference stress for embedded flaws in<br />

pipes, which may result in overly conservative assessments.<br />

On the other hand, limit loads based on global collapse of<br />

the pipe cross section (M FEA ) or global collapse in a flat-plate<br />

model, with plate width assumed equal to half the pipe<br />

The Journal of Pipeline Engineering<br />

Source Equn 4 Equn 5 BS 7910<br />

R6 R6 R6 R6<br />

Basis of<br />

σ ef<br />

/σ r M1<br />

M J2B<br />

E 0.<br />

5%<br />

FEA<br />

@ 0.<br />

5%<br />

strain<br />

M J2B<br />

EP<br />

0.<br />

5%<br />

FEA<br />

@ 0.<br />

5%<br />

strain<br />

Flat<br />

plate,<br />

local,<br />

pinned<br />

Flat<br />

plate,<br />

global,<br />

pinned<br />

Flat<br />

plate,<br />

global,<br />

fixed<br />

Flat<br />

plate,<br />

local,<br />

pinned<br />

Flat<br />

plate,<br />

local,<br />

fixed<br />

M odel<br />

( 1)<br />

( 2)<br />

( 3)<br />

, ( 4)<br />

( 3)<br />

, ( 4)<br />

( 3)<br />

, ( 4)<br />

( 3)<br />

, ( 4)<br />

( 3)<br />

, ( 4)<br />

E1BH3L50L1_5M01. 077<br />

1. 117<br />

1. 176<br />

1. 022<br />

1. 013<br />

1. 168<br />

1.<br />

091<br />

E1BH3L50L3M01. 057<br />

1. 099<br />

1. 158<br />

1. 020<br />

1. 013<br />

1. 150<br />

1.<br />

091<br />

E1BH3L50L14M01. 052<br />

1. 096<br />

1. 158<br />

1. 020<br />

1. 013<br />

1. 150<br />

1.<br />

091<br />

E1BH3L50L6M01. 038<br />

1. 082<br />

1. 124<br />

1. 016<br />

1. 013<br />

1. 117<br />

1.<br />

091<br />

E1BH3L50L9M01. 031<br />

1. 077<br />

1. 103<br />

1. 013<br />

1. 013<br />

1. 096<br />

1.<br />

091<br />

E1BH6L50L3M01. 098<br />

1. 138<br />

1. 351<br />

1. 037<br />

1. 026<br />

1. 308<br />

1.<br />

200<br />

E1BH6L50L11M01. 097<br />

1. 138<br />

1. 351<br />

1. 037<br />

1. 026<br />

1. 308<br />

1.<br />

200<br />

E1BH9L50L3M01. 117<br />

1. 157<br />

1. 586<br />

1. 049<br />

1. 039<br />

1. 459<br />

1.<br />

333<br />

E1BH6L50L6M01. 075<br />

1. 117<br />

1. 260<br />

1. 028<br />

1. 026<br />

1. 225<br />

1.<br />

200<br />

E1BH6L50L1_5M01. 107<br />

1. 146<br />

1. 404<br />

1. 041<br />

1. 026<br />

1. 357<br />

1.<br />

200<br />

E1BH3L25L3M01. 042<br />

1. 086<br />

1. 106<br />

1. 010<br />

1. 006<br />

1. 099<br />

1.<br />

061<br />

E1BH3L100L3M01. 069<br />

1. 110<br />

1. 209<br />

1. 041<br />

1. 026<br />

1. 202<br />

1.<br />

120<br />

E1BH3L200L3M01. 077<br />

1. 117<br />

1. 249<br />

1. 085<br />

1. 053<br />

1. 244<br />

1.<br />

143<br />

E1BH3L250L3M01. 080<br />

1. 120<br />

1. 260<br />

1. 109<br />

1. 067<br />

1. 255<br />

1.<br />

149<br />

E1BH6L25L3M01. 066<br />

1. 108<br />

1. 227<br />

1. 018<br />

1. 013<br />

1. 194<br />

1.<br />

130<br />

E1BH6L100L3M01. 128<br />

1. 167<br />

1. 480<br />

1. 076<br />

1. 053<br />

1. 437<br />

1.<br />

273<br />

E1BH3L25L6M01. 024<br />

1. 071<br />

1. 084<br />

1. 008<br />

1. 006<br />

1. 078<br />

1.<br />

061<br />

Sample issue<br />

mean circumference, underestimate the reference stress<br />

compared with J-based solutions.<br />

Based on the above, it may be concluded that the way<br />

forward is to develop new pipe-specific J-based solutions,<br />

which represent loading conditions between global collapse<br />

and conventionally-defined local collapse. This implies<br />

that the required J-based solutions correspond to a larger<br />

load-bearing area (defined by the extent of the ligament on<br />

either side of the flaw) than that associated with conventional<br />

local-collapse solutions.<br />

In further work, simple equations to estimate M J2B E 0.5%<br />

have been developed using a semi-analytical approach


3rd Quarter, 2009 163<br />

Basis of<br />

σ from<br />

code<br />

ref<br />

flat<br />

plate<br />

solution<br />

BS<br />

7910:<br />

local,<br />

pinned<br />

( 2)<br />

, ( 3)<br />

σr ef<br />

( BS<br />

7910)<br />

/<br />

σ ( M )<br />

ref J 2B<br />

EP<br />

0.<br />

5%<br />

effects<br />

of<br />

R6:<br />

global,<br />

pinned<br />

( 2)<br />

,<br />

( 3)<br />

σr ef<br />

σref ( R6)<br />

/<br />

( M )<br />

J 2B<br />

EP<br />

0.<br />

5%<br />

R6:<br />

global,<br />

fix<br />

σr ef<br />

σref ligamen<br />

t size<br />

( 2a=<br />

3,<br />

2c=<br />

50,<br />

p=<br />

1.<br />

5 to<br />

9)<br />

ed<br />

( 2)<br />

,<br />

( R6)<br />

/<br />

( M )<br />

J 2B<br />

EP<br />

0.<br />

5%<br />

Table 4. Ratio of s ref from code flat-plate solution to s ref from M J2B EP 0.5% .<br />

Notes:<br />

(1) M J2B EP 0.5% based on elastic-plastic stress and is determined at 0.5% strain.<br />

(2) Collapse of load bearing area around crack front.<br />

(3) All plate solutions: width = p x mean pipe radius<br />

( 3)<br />

R6:<br />

local,<br />

pinned<br />

( 2)<br />

,<br />

( 3)<br />

σr ef<br />

σref ( R6)<br />

/<br />

( M )<br />

J 2B<br />

EP<br />

0.<br />

5%<br />

E1BH3L50L1_5M01. 05<br />

0. 91<br />

0. 91<br />

1.<br />

05<br />

E1BH3L50L3M01. 05<br />

0. 93<br />

0. 92<br />

1.<br />

05<br />

E1BH3L50L6M01. 04<br />

0. 94<br />

0. 94<br />

1.<br />

03<br />

E1BH3L50L9M01. 02<br />

0. 94<br />

0. 94<br />

1.<br />

02<br />

effects<br />

of<br />

ligamen<br />

t size<br />

( 2a=<br />

6,<br />

2c=<br />

50,<br />

p=<br />

1.<br />

5 to<br />

6)<br />

E1BH6L50L1_5M01. 23<br />

0. 91<br />

0. 90<br />

1.<br />

18<br />

E1BH6L50L3M01. 19<br />

0. 91<br />

0. 90<br />

1.<br />

15<br />

E1BH6L50L6M01. 13<br />

0. 92<br />

0. 92<br />

1.<br />

10<br />

effects<br />

of<br />

ligamen<br />

t size<br />

( 2a=<br />

3,<br />

2c=<br />

25,<br />

p=<br />

3 to<br />

6)<br />

E1BH3L25L3M01. 02<br />

0. 93<br />

0. 93<br />

1.<br />

01<br />

E1BH3L25L6M01. 01<br />

0. 94<br />

0. 94<br />

1.<br />

01<br />

effects<br />

of<br />

height<br />

( 2a=<br />

3 to<br />

9,<br />

2c=<br />

50,<br />

p=<br />

3)<br />

E1BH3L50L3M01. 05<br />

0. 93<br />

0. 92<br />

1.<br />

05<br />

E1BH6L50L3M01. 19<br />

0. 91<br />

0. 90<br />

1.<br />

15<br />

E1BH9L50L3M01. 37<br />

0. 91<br />

0. 90<br />

1.<br />

26<br />

effects<br />

of<br />

height<br />

( 2a=<br />

3 to<br />

6,<br />

2c=<br />

25,<br />

p=<br />

3)<br />

E1BH3L25L3M01. 02<br />

0. 93<br />

0. 93<br />

1.<br />

01<br />

E1BH6L25L3M01. 11<br />

0. 92<br />

0. 91<br />

1.<br />

08<br />

Sample issue<br />

effects<br />

of<br />

length<br />

( 2a=<br />

3,<br />

2c=<br />

25<br />

to<br />

250,<br />

p=<br />

3)<br />

E1BH3L25L3M01. 02<br />

0. 93<br />

0. 93<br />

1.<br />

01<br />

E1BH3L50L3M01. 05<br />

0. 93<br />

0. 92<br />

1.<br />

05<br />

E1BH3L100L3M01. 09<br />

0. 94<br />

0. 92<br />

1.<br />

08<br />

E1BH3L200L3M01. 12<br />

0. 97<br />

0. 94<br />

1.<br />

11<br />

E1BH3L250L3M01. 13<br />

0. 99<br />

0. 95<br />

1.<br />

12<br />

effects<br />

of<br />

length<br />

( 2a=<br />

6,<br />

2c=<br />

25<br />

to<br />

100,<br />

p=<br />

3)<br />

E1BH6L25L3M01. 11<br />

0. 92<br />

0. 91<br />

1.<br />

08<br />

E1BH6L50L3M01. 19<br />

0. 91<br />

0. 90<br />

1.<br />

15<br />

E1BH6L100L3M01. 27<br />

0. 92<br />

0. 90<br />

1.<br />

23


164<br />

calibrated by means of the finite-element results. Additional<br />

work is currently being performed to develop equations to<br />

estimate s ref /s M1 . The outcome of these activities, in terms<br />

of equations that can be used to predict M J and/or s ref /s M1<br />

in routine assessments, will be published in 2010.<br />

Plastic strain in ligament<br />

In order to enable an assessment of strains in the ligament<br />

adjacent to embedded flaws, data on strain concentrations<br />

in the smaller of the two ligaments (below and above the<br />

flaw) were obtained for all cases. Figure 12 shows typical<br />

contour plots for the equivalent plastic strain in the ligaments<br />

adjacent to an embedded flaw. The example shown illustrates<br />

contour plots in model E1BH6L50L3M0, which indicate<br />

that strain concentration occurs in two plastic zones<br />

issue<br />

The Journal of Pipeline Engineering<br />

Fig.12. Contour plots for the equivalent plastic strain in the ligaments adjacent to an embedded flaw (model<br />

E1BH6L50L3M0). Sample<br />

extending from the flaw tip towards the surface at 45 o with<br />

respect to the plane of the flaw.<br />

The plastic strain at the surface on either side of the flaw<br />

corresponding to a remote axial strain on the pipe OD of<br />

1% is given in Table 3. It can be seen that such strains,<br />

which increase as the flaw height or length increase or the<br />

ligament decreases, can be as high as 11.5% (model<br />

E1BH9L50L3). It should be noted that these strains were<br />

obtained in models which were perfectly aligned. Had axial<br />

misalignment been included, the strains in the ligament<br />

would have been even higher.<br />

Further research is required to produce guidance on strain<br />

concentrations in the ligament in the presence of axial<br />

misalignment and strength mismatch. Such guidance can


3rd Quarter, 2009 165<br />

be used to assess the possibility of ligament failure due to<br />

excessive straining. This would be a separate failure criterion<br />

to J-based fracture associated with extension of the flaw.<br />

Summary and conclusions<br />

Three-dimensional elastic-plastic finite-element analyses<br />

have been conducted on pipes containing circumferential<br />

embedded flaws. From these analyses, the elastic-plastic<br />

fracture mechanics’ parameter J has been evaluated and<br />

used to determine limit loads consistent with the reference<br />

stress J-estimation scheme. Two J-based limit loads have<br />

been determined: M J2B E and M J2B EP . In addition, global<br />

collapse limit loads were obtained from elastic-perfectly<br />

plastic finite-element analyses (denoted as M FEA ).<br />

Existing standard solutions and methods for determining<br />

limit load (and/or reference stress) estimates for<br />

circumferential embedded flaws in pipes within the context<br />

of FAD-based assessments have been reviewed and evaluated<br />

against results obtained from the finite-element analyses.<br />

The following conclusions are made:<br />

a) J-based limit-load and reference stress<br />

1 M J2B E (J e based on the elastic pipe bending stress) is<br />

higher than M J2B EP (J e based on the elastic-plastic<br />

pipe bending stress) by approximately 4%.<br />

2 Both M and M are found to increase with<br />

J2B E J2B EP<br />

applied load (bending moment), and hence applied<br />

strain (for strains > 0.2%). This behaviour is believed<br />

to be due to the combined effects of the loading<br />

considered, approximations within the referencestress<br />

J-estimation scheme, and the fact that J-based<br />

limit loads are not true limit loads. Also M and<br />

J2B E<br />

M are both found to increase with L (for L ><br />

J2B EP r r<br />

1.0).<br />

3 If M J2B E and M J2B EP are multiplied by the ratio of the<br />

maximum elastic-plastic stress to the elastic stress<br />

(s M1 /s M ) in the pipe remote from the crack at a<br />

given load level, or if the results are expressed in<br />

terms of s ref /s M1 , the resulting parameters become<br />

largely independent of load level (for L r > 1 and<br />

applied strain > 0.5%). This enables J to be estimated<br />

reliably for a wide range of applied loads.<br />

b) Standard solutions and methods<br />

4 Global-collapse limit loads (such as M FEA ) are higher<br />

than J-based limit loads and insensitive to crack size<br />

and ligament height. Their use in FAD-based<br />

assessments could lead to non-conservative estimates<br />

of J.<br />

5 Flat-plate reference-stress solutions based on local<br />

collapse with pin loading overestimate J-based<br />

solutions determined from M J2B E 0.5% and M J2B EP 0.5%<br />

(values of M J2B E and M J2B EP at 0.5% applied strain)<br />

and, consequently, lead to conservative estimates of<br />

J.<br />

6 In most of the cases considered, the flat-plate<br />

reference-stress solutions based on global collapse<br />

(with a plate width equal to half the pipe<br />

circumference) underestimate J-based solutions<br />

determined from M J2B E 0.5% and M J2B EP 0.5% and,<br />

consequently, can potentially lead to nonconservative<br />

assessments. However, it can be shown<br />

that modifying the R6 pin-loading model (Equn 13)<br />

by adjusting the plate width, can lead to solutions<br />

that agree well with s ref /s M1 based on M J2B E 0.5% and<br />

M J2B EP 0.5% .<br />

c) Equations for estimating J-based limit-load and/or<br />

reference stress<br />

A new general equation for estimating M J2B E 0.5% has<br />

been derived from a semi-analytical approach<br />

calibrated by means of the finite element results.<br />

Additional equations are being developed for<br />

estimating s ref /s M1 . When used in conjunction<br />

with BS 7910 (Level 2B/3B) FAD or R6 Option 2<br />

FAD, the new J-based solutions give an improved<br />

estimate of J, and hence flaw assessment, compared<br />

to using standard codified limit-load solutions based<br />

on local or global collapse. The new solutions will<br />

be published in 2010.<br />

d) Plastic strain concentration<br />

Sample issue<br />

Data on plastic strain concentration in the smaller<br />

of the two ligaments adjacent to embedded flaws<br />

have been obtained. For some cases, the plastic<br />

strain at the surface nearest to the flaw exceeds 10<br />

times the nominal remote strain on the pipe OD.<br />

e) Future work<br />

More work is needed to incorporate the effects of<br />

axial misalignment and strength mismatch and<br />

account for discontinuous yielding and other rates<br />

of strain hardening. More work is also required to<br />

produce guidance on strain concentration in the<br />

ligament to enable the assessment of ligament failure<br />

due to excessive straining.<br />

Acknowledgements<br />

The work was funded, as part the Core Research Programme,<br />

by Industrial Members of TWI, whose support is gratefully<br />

acknowledged. The author also acknowledges the efforts of<br />

Dr Martin Goldthorpe, who conducted the finite-element<br />

analyses, and the valuable support provided by John Wintle<br />

and Dr Simon Smith.


166<br />

References<br />

1. BSI, 2005. BS 7910:2005 including Amendment 1, 2007:<br />

Guide to methods for assessing the acceptability of flaws in<br />

metallic structures. British Standards Institution.<br />

2. BEGL, 2001. R6 Revision 4 and amendments: Assessment of<br />

the integrity of structures containing defects. British Energy<br />

Generation Ltd, Gloucester, UK. (Amendments issued in<br />

subsequent years.)<br />

3. API, 2000. API 579: Fitness-for-service, 1 st Edn, American<br />

Petroleum Institute.<br />

4. API/ASME, 2007. API-579-1/ASME-FFS-1-2007: Fitnessfor-service,<br />

API 579 2 nd Edn, American Petroleum Institute<br />

(API)/American Society of Mechanical Engineers (ASME).<br />

The Journal of Pipeline Engineering<br />

5. A.A.Willoughby and T.G.Davey, 1987. Plastic collapse at<br />

part wall flaws in plates, in ‘Fracture mechanics: perspectives<br />

and directions. Proc. 20th national symposium, Bethlehem,<br />

PA, USA. ASTM STP 1020, pp340-409.<br />

6. Y.Lei and P.J.Budden, 2004. Limit load solutions for plates<br />

with embedded cracks under combined tension and bending.<br />

Int. J.Pressure Vessels and Piping, 81, pp589-597.<br />

7. ABAQUS/Standard User’s Manuals, Version 6.6, Hibbitt,<br />

Karlsson and Sorenson Inc.<br />

Sample issue


3rd Quarter, 2009 167<br />

The Nord Stream Pipeline’s<br />

German landfall: the challenges<br />

ahead<br />

by Nigel S Kirk 1 and Dipl-Ing Björn Dobberstein* 2<br />

1 Project Manager Landfall Germany, Nord Stream AG, Zug, Switzerland<br />

2 Project Engineer Landfall Germany, Nord Stream AG, Zug, Switzerland<br />

IN THIS ISSUE the Journal of Pipeline Engineering features the first of three articles on one of the world’s<br />

biggest current pipeline projects, Nord Stream. The series articles will review various aspects of the<br />

project, with specific attention paid to the pipeline landfall in Germany. The first article describes the Nord<br />

Stream project and general technical details, together with a description of the German landfall (preconstruction)<br />

including the environmental and permitting issues, anticipated construction techniques, and<br />

the expected installation schedule.<br />

The second article will provide a general update of the project (during construction) and describe the<br />

installation design together with the actual construction activities and the challenges encountered. The final<br />

article will review the Nord Stream project on completion of the first pipeline and assess the positive and<br />

negative aspects of the German landfall’s construction.<br />

Project overview<br />

The Nord Stream Pipeline Project consists of two 1223-km<br />

long parallel 48-in diameter offshore pipelines laid across<br />

the Baltic Sea, connecting the pig launchers close to the<br />

compressor station at Portovaya Bay, Russia, and the pig<br />

receivers adjacent Greifswald receiving terminal in<br />

Germany. The pipelines landfall at their most northerly<br />

end in Vyborg, NW of St Petersburg, and generally run<br />

westward through the Gulf of Finland for approximately<br />

440km, then turning generally southward and running<br />

east of the Swedish island of Gotland. The pipelines then<br />

turn to the SW and skirt the Danish island of Bornholm,<br />

continuing in an SSW direction and eventually landfalling<br />

close to Lubmin, east of Greifswald in Germany.<br />

The planned route of the Nord<br />

Stream project<br />

Nord Stream AG is an international joint venture: its<br />

shareholders are OAO Gazprom (51%), BASF/Wintershall<br />

*Author’s contact details<br />

tel: +41 41 766 9209<br />

email: bjoern.dobberstein@nord-stream.com<br />

AG and E.ON Ruhrgas AG (each with 20%) and NV<br />

Nederlandse Gasunie (9%). The company was set up in<br />

September, 2005, to plan, construct, and subsequently<br />

operate, the Nord Stream pipeline. The Nord Stream<br />

Project has a budget of approximately $11.1 billion, with<br />

goods and services being supplied to the project on a<br />

worldwide basis from across Europe, the USA, and Russia.<br />

Sample issue<br />

The European Union’s annual demand for natural gas<br />

imports, which was approximately 314 billion cubic meters<br />

(bcm) in 2005, is forecasted to increase to 509 bcm in 2025:<br />

subsequently, the annual import gap is anticipated to reach<br />

almost 200 bcm by 2025. Nord Stream’s goal is contribute<br />

to closing this gap by connecting the largest gas reserves in<br />

the world with the European gas network: the Nord Stream<br />

pipeline will meet about 25% of this additional import<br />

demand by supplying Europe with 55 bcm/yr of natural<br />

gas. In other terms, 55 bcm of natural gas contains enough<br />

energy to meet the annual demands of 13-14 million<br />

people.<br />

The Nord Stream pipeline will be installed by two of the<br />

world’s largest offshore installation contractors for largediameter<br />

pipelines: Saipem SpA of Italy, and Allseas<br />

Deepwater Contractors of Switzerland. The pipe for the<br />

project will be supplied by Europipe GmbH, Germany, and<br />

OMK Steel, Russia; production is well advanced, and


168<br />

Fig.1. The planned route of the Nord Stream project.<br />

approximately 70% of the first pipeline’s pipes have already<br />

been manufactured. Eupec Pipe Coatings of France will<br />

provide the complete pipe logistical and supply services,<br />

including the concrete weight coating of the pipes. Two<br />

brand-new concrete-coating plants have been built to service<br />

the vast quantity of pipe, one at Mukran on the island<br />

Rugen, Germany, and one at Kotka in the Gulf of Finland.<br />

The Baltic Sea is a highly-sensitive ecological region and, as<br />

a result, Nord Stream AG has carried out extensive and<br />

detailed environmental impact studies and environmental<br />

planning to ensure that the design, installation, and<br />

operation of the pipeline will be environmentally sound.<br />

Construction of pipeline 1 (the North West line) is planned<br />

to start in April, 2010, with first gas expected by September,<br />

2011. The full transport capacity of approximately 55 bcm/<br />

yr will be available on completion of pipeline 2 (the South<br />

East line) in November, 2012.<br />

Technical aspects/data<br />

System design<br />

The pipeline has been designed in accordance with DNV<br />

OS-F101 2000, Rules for submarine pipeline systems, including<br />

the January, 2003, Amendment and corrections. Additionally,<br />

in Germany the DIN - EN 14161 Petroleum and natural gas<br />

industries: pipeline transportation systems (ISO 13623:2000<br />

modified) code has also been satisfied due to authority<br />

requirements. However, should DIN EN 14161 and the<br />

DNV F101 code be in contradiction with each other, then<br />

the DNV code takes priority.<br />

Each of the pipelines will have a transport capacity of<br />

approximately 27.5 bcm/yr of natural gas at reference<br />

conditions of 20°C and 1atm, and the system’s design life<br />

is 50 years.<br />

The Journal of Pipeline Engineering<br />

The pipeline system’s limits are defined as between the pig<br />

launcher and the pig receiver at the Russian and German<br />

landfalls, respectively.<br />

The fundamental design of the Nord Stream pipelines was<br />

based on several factors, including the steady-state and<br />

transient flow operating conditions and overall<br />

environmental, economic, and commercial optimization.<br />

This resulted in dividing the pipeline’s overall length of<br />

1,223km into three MAOP (maximum allowable operating<br />

pressure) sections, as shown in Table 1.<br />

Pipe data<br />

Sample issue<br />

• nominal size = DN 48in (DN1200)<br />

• constant internal diameter ID = 1,153mm<br />

• pipes with longitudinally-welded seams (submergedarc<br />

welding) with an individual pipe length of<br />

approximately 12.2m<br />

• pipe material SAWL 485 I DF according to DNV<br />

standard OS-F101 with the following characteristic<br />

values:<br />

minimum yield strength = 485N/mm²<br />

modulus of elasticity = 2.07 x 10 5 N/mm²<br />

transverse contraction number = 0.3<br />

thermal expansion coefficient = 1.16 x 10 -5 /°C<br />

density = 7,850 kg/m³<br />

External coating<br />

The three-layer anti-corrosion coating was designed in<br />

accordance with ISO 21809-1 External coatings for buried or<br />

submerged pipelines used in pipeline transportation systems.<br />

Designed to a minimum thickness of 4.2mm, it consists of:<br />

• first layer: approximately 0.15mm of FBE (fusionbonded<br />

epoxy)


3rd Quarter, 2009 169<br />

Table 1. Pipeline design data.<br />

• second layer: approximately 0.25mm of adhesive<br />

PE coating<br />

• third layer: approximately 3.8mm of PE coating<br />

Internal coating<br />

The internal coating was designed in compliance with API<br />

RP5L2 Recommended practice for the internal coating of line pipe<br />

for non corrosive gas transmission with an epoxy coating film<br />

thickness of approximately 90mmm and an internal<br />

roughness specified as Rz = 5mmm.<br />

Concrete weight coating<br />

The designed thickness of the concrete coating depends on<br />

a variety of factors including, environmental influences<br />

(water depth, current flow, waves), the pipe properties<br />

(diameter and wall thickness), and the concrete density. In<br />

accordance with DNV OS-F101, the selected concrete<br />

density of 3,040 kg/m³ is achieved by mixing an additional<br />

70% of iron ore to the concrete. The concrete thickness<br />

varies between 60mm and 110mm over the whole pipeline<br />

route.<br />

Anodes<br />

Locat ion<br />

Kilomet<br />

re<br />

post<br />

Russia<br />

section<br />

- dry<br />

Offshore<br />

-<br />

Segment<br />

1<br />

Offshore<br />

-<br />

Segment<br />

2<br />

Offshore<br />

-<br />

Segment<br />

3<br />

Germany<br />

- dry<br />

section<br />

Galvanic anodes, commonly known as sacrificial anodes,<br />

provide a permanent current flow through the pipe,<br />

commonly known as cathodic protection. Anodes are<br />

fitted to the pipes as part of the concrete weight coating<br />

process and are directly electrically connected to the steel<br />

pipe to protect the pipeline from corrosion in those areas<br />

where the basic coating may be defective. The anodes are<br />

designed in accordance with two primary specifications,<br />

namely DNV RP-F103 Cathodic protection of submarine<br />

pipelines by galavanic anodes and ISO 15589-2 Cathodic<br />

protection of pipeline transportation systems. Along the pipeline<br />

route the anode spacing varies between 7 and 12 pipe<br />

lengths.<br />

Depending on the average salinity of the surrounding<br />

seawater either aluminium or zinc anodes will be used on<br />

0 - 0.<br />

5<br />

0.<br />

5<br />

- 300<br />

Design<br />

temper<br />

atures<br />

( oC)<br />

60<br />

max<br />

-38<br />

min<br />

MAOP<br />

( desig<br />

n<br />

pressure<br />

- bar)<br />

Wall<br />

thickne<br />

ss<br />

( mm)<br />

22041. 0<br />

22034. 6<br />

300-675 40<br />

max<br />

-10<br />

min<br />

200 30.<br />

9<br />

675-1223170 26.<br />

8<br />

500m<br />

seawards<br />

up<br />

to<br />

pig<br />

trap<br />

60<br />

max<br />

-25min170 30.<br />

9 ( buried<br />

section)<br />

34.<br />

6 ( above<br />

ground<br />

section)<br />

the Nord Stream pipeline. The weight of each aluminium<br />

anode is between 200kg and 460kg, whereas the zinc<br />

anodes each weigh between 530kg and 1,200kg. The weight<br />

of the anodes is dependent on their density, individual<br />

length, outer pipe diameter, and the concrete coating<br />

thickness. A total of 4,800t of zinc and 5,200t of aluminium<br />

anodes will be attached to the pipelines.<br />

Environment and permitting in<br />

the German EEZ and elsewhere<br />

Within the German Exclusive Economic Zone (EEZ) border<br />

and territorial waters, the pipeline route circumvents and<br />

bisects several national and international designated nature<br />

protection areas.<br />

The most significant zones include the following flora,<br />

fauna habitat (FFH) areas, such as the Greifswalder Bodden<br />

and Parts of Stralsund and Usedom North Head (DE 1747-<br />

301 SCI) and the Greifswalder Boddenrandschwelle and<br />

Parts of the Pomeranian Bight (DE 1749-302 SCI). This<br />

area of the Baltic Sea is designated as the Greifswalder<br />

Bodden and Stralsund national wetland, and is one of the<br />

most important stopover sites during migration, or as a<br />

wintering or moulting area; the following areas have<br />

therefore been dedicated as EU Bird Sanctuary Areas, and<br />

include the Greifswalder Bodden and Southern Stralsund<br />

(DE 1747-402 SPA) and (DE 1747-401 SPA), the<br />

Pomeranian Bight (DE 1552-401 SPA), and the Western<br />

Pomeranian Bight (DE 1649-401 SPA).<br />

Sample issue<br />

At Lubmin, the route runs through the planned<br />

Peenemünder Haken, Struck and Ruden nature reserve.<br />

The landfall dry section design has been altered significantly<br />

here to satisfy the conditions of the Habitats Directive<br />

2130 Fixed dunes with herbaceous vegetation (grey dunes).<br />

During the Great Nordic War (1700 - 1721) 20 ships were<br />

sunk on the Boddenrandschwelle sandbar by the Swedish<br />

army to block access to the Greifswalder Bodden area. This


170<br />

Sample issue<br />

The Journal of Pipeline Engineering<br />

Fig.2. General<br />

layout of the<br />

Nord Stream<br />

pipeline’s<br />

German landfall.


3rd Quarter, 2009 171<br />

barrier of ships, known as ‘Schiffssperre’, is a designated<br />

historic monument. One of the shipwrecks, identified<br />

simply as No 67, has recently been successfully recovered<br />

from the sea bed and placed in wet storage to allow access<br />

for the laybarge and dredging vessels.<br />

The unusual S-shape of the route within the Greifswalder<br />

Bodden has been designed so that the pipelines can be<br />

installed within a defined planning corridor designated as<br />

a marine priority area within the planning policy of the<br />

regional development of Mecklenburg-Western Pomerania.<br />

The environmental and ecological restrictions within the<br />

protected areas have essentially determined the construction<br />

schedule and have had such a significant and direct effect<br />

on the project’s technical aspects that numerous installation<br />

methods have had to be designed, proposed, modified, and<br />

ultimately accepted to minimize the environmental impact<br />

and maintain the construction operations within the<br />

available time period.<br />

Onshore, several environmental-mitigation measures will<br />

have to be undertaken prior to the start of construction at<br />

Lubmin. These include the construction of habitats for<br />

lizards, amphibian pathways, solid partition fencing/<br />

screening, and transplantation of small trees and bushes.<br />

Additional permit restrictions, including dredging and<br />

cofferdam installations, are not allowed to commence<br />

before 15 May, 2010, due to the herring-spawning season<br />

in the Greifswalder Bodden, and all offshore construction<br />

work must be completed in one construction season, i.e. by<br />

31 December, 2010, within the FFH areas in the area of the<br />

German landfall.<br />

Noise-restriction guidelines have been set, at the nearest<br />

residential towns adjacent to the pipeline route, where the<br />

levels shall not exceed 50dB(A) during the daytime and<br />

35dB(A) at night-time; additionally noise levels at the<br />

adjacent marina shall be monitored and shall not exceed<br />

65dB(A) during the daytime and 50dB(A) during nighttime.<br />

Geological formation<br />

The offshore area crossed by the Nord Stream pipeline<br />

within the German sector can be divided into four different<br />

geological sectors: the Greifswalder Bodden, the<br />

Boddenrandschwelle, the Oder subsea valley, and the Oder<br />

Bank. Each area was formed by the glacial processes in the<br />

last ice age and the subsequent marine progression of the<br />

Baltic Sea and can generally be distinguished by their<br />

different water depths.<br />

The Greifswalder Bodden is almost a land-enclosed basin<br />

with a water depth up to 10m. The seabed of this basin is<br />

characterized by the variation of a large number of soil<br />

types, such as sand and gravel, peat, clay, or alluvial mud,<br />

which is typical for lagoon-like areas.<br />

At its eastern edge the basin is restricted towards the Baltic<br />

Sea by a submarine barrier with water depth of less than<br />

5m. This barrier is called the Boddenrandschwelle, and is<br />

formed of glacial till (a cohesive mixture of clay, sand, and<br />

gravel) with loose residual sediments (sand, gravel, cobbles)<br />

generally at the uppermost (surface) layers.<br />

The Oder valley formation, an ancient course of the river<br />

Oder during the end of the last ice age, runs parallel to the<br />

Boddenrandschwelle. The water depth in this subsea valley<br />

ranges to 20m and above. The sea bottom is composed of<br />

coarse-grained fluvial sediments which have been covered<br />

by recent alluvial mud deposits.<br />

The major part of the pipeline route within the German<br />

sector crosses the Pomeranian Bay east of the Oder valley<br />

along the northern foothill of the Oder Bank. The sea<br />

bottom is characterized by marine sediments (largely sand),<br />

with water depths varying from 15m to more than 20m.<br />

German landfall<br />

At the German landfall the Nord Stream pipeline is divided<br />

into three separate sections: the offshore section, the ‘pullin’<br />

section, and the dry section.<br />

The German landfall offshore section commences<br />

approximately 1km seaward of the highly-protected<br />

Greifswalder Boddenrandschwelle and Parts of the<br />

Pomeranian Bight (DE 1749-302) FFH-area, and extends<br />

in a south-westerly direction for 26km to approximately<br />

1,100m from the shoreline at Lubmin. The pull-in section<br />

then commences and ends approximately 220m landward<br />

of the shoreline. The remaining 300m consists of the dry<br />

section up to the pig receivers.<br />

Sample issue<br />

Dry section<br />

The dry section generally comprises an ‘Omega’-shaped<br />

spool arrangement and a combination of valves (emergency<br />

shutdown and gate) together with isolation joints and pig<br />

receivers. The 48-in pipework is connected to the Greifswald<br />

receiving terminal (GRT) by way of twin 38-in diameter<br />

pipelines and supplementary 16-in by-pass lines.<br />

The dry section works were originally designed with a ‘dogleg’<br />

arrangement; however, subsequent to a periodic<br />

environmental survey, it was discovered that the Grey<br />

Dune vegetation had migrated onto the pipeline route,<br />

thereby necessitating a realignment of the pipeline route<br />

and a redesign of the onshore pipework<br />

The dry-section pipework and permanent-work items will<br />

be supported by over 100 reinforced concrete bases of<br />

varying sizes.


172<br />

All the permanent-work materials will be delivered to site<br />

by road including the 105-t valves and 75-t pig receivers. A<br />

standard fabrication process is anticipated, notwithstanding<br />

the size and scale of the valves and pipework. All the welds<br />

will be non-destructively examined by automatic ultrasonic<br />

techniques with the entire permanent works being painted<br />

subsequent to the completion of fabrication.<br />

The dry section will be hydrostatically tested as part of the<br />

project’s pre-commissioning philosophy with the final<br />

‘golden’ welds being carried out at the connection to the<br />

pull-in section. The final reinstatement of the area postconstruction<br />

will necessitate the construction of a small<br />

artificial dune to ensure that the minimum cover levels for<br />

the pipelines are achieved.<br />

Pull-in<br />

The pipelines will be pulled from Saipem’s pipelay barge<br />

Castoro Dieci moored approximately 1100m away from the<br />

shoreline in a water depth of approximately 4.5-5m. The<br />

pipelines will be pulled individually into a pre dredged<br />

trench using a 4-in diameter steel wire and 500-t winch and<br />

piled back-anchor arrangement. At approximately 550m<br />

from the shoreline, the pre-dredged trench is replaced by a<br />

pre-installed cofferdam of approximately 9.5m width; the<br />

cofferdam continues for a further 150m onshore.<br />

As the pipelines are pulled ashore, buoyancy tanks are<br />

fitted to the pipelines at the laybarge to ensure that the pull<br />

loads are not exceeded. At the shoreline, the pipe level<br />

begins to increase with the pipes eventually being pulled<br />

into a lazy-S profile. The pipeline will be pulled to<br />

approximately 220m onshore using a combination of dry<br />

pull and temporary support rollers.<br />

There are currently two options under consideration for<br />

the construction of the cofferdam:<br />

• Option 1: three-wall cofferdam<br />

The seaward cofferdam consists of three sheet pile<br />

walls running parallel to each other and forming<br />

two channels that will each be 9.5m wide. Material<br />

will be dredged from one of the channels, and the<br />

other will be used for storage of the dredged material,<br />

resulting in a total width of approximately 19m.<br />

The channel for the storage of the dredged material<br />

will be additionally supported by piles providing a<br />

support for a steel framework that serves as a platform<br />

for the pile-driving and dredging equipment.<br />

• Option 2: two-wall cofferdam with Bailey bridge<br />

The offshore cofferdam will consist of two parallel<br />

sheet pile walls forming a trench which will be<br />

approximately 9.5m wide, and will be constructed<br />

from a pre-installed Bailey bridge running parallel<br />

to the cofferdam. The Bailey bridge will be supported<br />

The Journal of Pipeline Engineering<br />

by steel piles with a diameter of approximately 1m;<br />

as it is a modular steel structure, it will be quick and<br />

easy to construct. The dredging of the cofferdam<br />

will be undertaken from the bridge, with the dredged<br />

material being stored in the shallow water adjacent<br />

to the bridge. Silt screens will be installed in the<br />

water to contain the sediments released from the<br />

dredged material.<br />

Typical cofferdam and pull-in arrangement<br />

Offshore section<br />

The offshore pipelay will be undertaken by Saipem’s Castoro<br />

10 laybarge using the traditional S-lay method, commencing<br />

immediately after the pull-in operations. The two pipelines<br />

will be individually pulled-in toward the shoreline and laid<br />

in a single trench with a bottom width of 9.5m in the<br />

straight sections and 10.5m in the curved sections (a radius<br />

of 2500m approx.). The bottom trench widths were set at<br />

9.5m and 10.5m in order that the affected areas due to<br />

dredging would be minimized.<br />

The pipelines will be laid sequentially, with the pipeline on<br />

the inner curve laid first. This will prevent the second<br />

pipeline being laid on top of the first in the unlikely event<br />

of an emergency abandonment on the second pipelay.<br />

Additionally, the sequenced lay has been developed to<br />

enable the backfilling operation to commence as early as<br />

possible and to reduce the amount of time the dredged<br />

trench remains open, thus satisfying permit and<br />

environmental requirements. The expected pipeline lay<br />

rate is between 350m/day and 550m/day.<br />

The minimum separation between the pipelines in the<br />

dredged trench will be 2.5m, with a nominal pipeline<br />

centre-to-centre distance of 6m.<br />

Sample issue<br />

The total length of the pre-dredged trench will be<br />

approximately 26km with about 1,800,000m 3 of material<br />

being excavated, temporarily stored, and backfilled.<br />

The dredging and backfilling works within the German<br />

landfall have been subcontracted to a joint venture of<br />

Boskalis Offshore bv of the Netherlands and Rohde Nielsen<br />

A/S of Denmark. The trench will be excavated using<br />

trailing suction-hopper dredgers, bucket-ladder dredgers,<br />

and backhoe dredgers. The dredging operations will<br />

commence with a comprehensive pre-dredge survey of the<br />

anticipated trench and material-storage areas. The pipeline<br />

trench will be excavated well in advance of the laybarge<br />

mobilization in order that the scheduling of dredging,<br />

pipelay, and backfilling is sequenced precisely, and the<br />

works are completed in an efficient manner.<br />

The dredged material will predominantly be transported to<br />

a temporary offshore storage area unless the excavation and<br />

backfilling works are sequenced in parallel.


3rd Quarter, 2009 173<br />

Fig.3. Typical cofferdam and pull-in<br />

arrangement.<br />

The pipeline profile within the landfall area has been<br />

designed to satisfy a variety of criteria in respect of the<br />

burial depth. The cover to the pipeline varies from 1m to<br />

4.5m depending on pipeline stability, pipeline protection,<br />

coastal erosion, and local shipping authority requirements.<br />

The Nord Stream pipeline crosses a sandbar known as the<br />

Boddenrandschwelle, an area where the water depth is<br />

relatively shallow, varying between 2.5m and 4.5m deep.<br />

The pipe trench has to be widened over a length of about<br />

1,100m to ensure a minimum trench width of approximately<br />

50m to allow access for the laybarge.<br />

Several environmental restrictions have been applied to<br />

the dredging process, and include the following;<br />

• Several different types of topsoil (minimum dredged<br />

thickness of 0.3m) have to be dredged, temporarily<br />

stored, and backfilled separately to increase the<br />

possibility of a shorter regeneration period.<br />

• At all the excavation and backfilling locations the<br />

turbidity within the water is limited to 50mg/l<br />

(peak values 100mg/l) above the natural background<br />

levels within a distance of 500m around the dredging<br />

and transport equipment. In areas with fine-grained<br />

material such as clay or silt, these values may not be<br />

achieved during extreme adverse weather conditions,<br />

and therefore the deployment of silt screens and/or<br />

temporary contingency measures may become<br />

necessary.<br />

• Some of the dredged material cannot be stored at<br />

sea because of its high organic content and must be<br />

transported to an onshore location and either<br />

permanently disposed of (at spoil grounds, for<br />

example) or recycled (by soil separation).<br />

• Boulders of a certain size are designated as reefs, and<br />

therefore necessitate specific protection. All the<br />

boulders identified along the German landfall<br />

pipeline route with a size of more than 0.6m in one<br />

dimension will be stored separately and placed back<br />

as near to their original location as practicably<br />

possible after the reinstatement of the topsoil.<br />

To minimize the environmental impact within the German<br />

landfall the dredging tolerances for all dredging works have<br />

been limited to:<br />

• horizontal: +1.0 m/-0.0m<br />

• vertical: +0.0m/-0.3 m<br />

After installation of the pipelines the dredged material<br />

contained within the offshore storage area will be redredged<br />

and backfilled into the trench. Selected coarsegrained<br />

material will be placed directly around the pipeline<br />

to ensure that liquefaction of the material does not occur<br />

and induce buoyancy of the pipeline. Silty and cohesive<br />

soil-type materials will be placed above the coarse material<br />

with topsoil finishing-off the layered backfilling. Any soils<br />

unsuitable for backfilling will remain permanently at the<br />

offshore storage area. Should there be a deficit in respect of<br />

backfill material due to soil quantities being stored<br />

permanently either offshore or onshore, then suitable<br />

backfill material will be imported. The backfilled trench<br />

will be accepted on completion of an approved bathymetric<br />

survey.<br />

Sample issue<br />

After completion of the backfilling and reinstatement the<br />

offshore and pull-in sections of the German landfall will be<br />

hydrostatically tested in combination with the remainder<br />

of the Nord Stream pipeline.<br />

The German landfall section of the Nord Stream project is<br />

highly demanding in respect of environmental restrictions<br />

and technical challenges to this end Nord Stream are<br />

working closely with the various contractors to ensure the<br />

project is delivered on time and within budget.


174<br />

Sample issue<br />

The Journal of Pipeline Engineering


3rd Quarter, 2009 175<br />

Assessing pipeline integrity using<br />

fracture mechanics and<br />

currently available inspection<br />

tools<br />

by Dr Kimberly Cameron* and Dr Alfred Pettinger<br />

Exponent Failure Analysis, Menlo Park, CA, USA<br />

PIPELINE SYSTEMS ARE DESIGNED to comply with the regulatory requirements of each country and<br />

their applicable engineering standards. In the USA, Title 49 of the Code of Federal Regulations (CFR)<br />

establishes the mandatory minimum federal safety standards of pipelines for the transportation of natural<br />

gas (Part 192) and hazardous liquids (Part 195). These mandatory regulatory requirements typically cite<br />

consensus standards promulgated by the American Society of Mechanical Engineers (ASME), the American<br />

Pipeline Institute (API) and ASTM <strong>International</strong>1 . Specific performance criteria for pipeline systems suitable<br />

for the transportation of gas and hazardous liquids are established in ASME B31.8 [Gas transmission and<br />

distribution piping systems] and ASME B31.4 [Pipeline transportation systems for liquid hydrocarbons and other<br />

liquids] and frequently quoted in the construction specification of pipeline systems throughout the world.<br />

Code compliance is established if the designer demonstrates that all specific code requirements and all<br />

reasonably foreseeable load conditions are addressed by the design. The load condition seen by all pipelines<br />

is the load resulting from internal pressure. Because the hoop stress resulting from the internal pressure<br />

in the pipeline is at least twice the axial stress, typically longitudinal cracks and welds are the most susceptible<br />

and a substantial volume of literature addresses longitudinal cracking in pipes. Several pipeline systems,<br />

however, are subjected not only to internal pressure but also to significant external loads, which need to<br />

be evaluated using the code’s so-called occasional load condition. For example, pipeline systems buried in<br />

regions of active landslides, expansive soils, steep topography, and poor foundation conditions can be<br />

subjected to substantial external forces, which produce axial loads in the pipe. These loads can well exceed<br />

the axial pressure load and present a much greater risk for joints like circumferential welds. Guidance on<br />

how to implement some of these geotechnical considerations and how to estimate these external loads are<br />

described in more detail in Refs 1, 2, and 3.<br />

As our analysis will show, circumferential growth of cracks has the potential of causing severe consequences,<br />

typically leading to the rupture of the pipeline with the potential of a full-bore pipe failure. This observation<br />

is also reflected in the spill incident data of the 6th EGIG report [4], where the leading cause of spill incidents<br />

is external interference at 49.7%, followed by construction defects/material failure at 16.7%, corrosion at<br />

15.1%, and ground movement at 7.1%, with ground movement having the largest proportion of ruptures<br />

and landslides, causing more than half of the ground-movement-related spill incidents. Incident-spill data<br />

have been further segregated to only include pipelines in mountain areas [3]. The authors report an incident<br />

rate of 0.32 to 0.8 spill incidents per 1000 km years for mountainous areas in Europe and the USA and,<br />

depending on the sophistication of the geotechnical engineering, a rate of 0.33 to 2.8 spills per 1000 km year<br />

in the Andean Mountains. This rate is slightly larger than the most recent spill incident rates for pipelines<br />

at large, which are typically 0.2 spill incidents per 1000 km year [4]. However, this incident data [3] does<br />

not include the spill incident data from the most recently constructed pipeline system crossing the Andean<br />

*Author’s contact details:<br />

tel: +1 832 325 5700<br />

email: kcameron@exponent.com<br />

Sample issue<br />

continued overleaf<br />

1. Formerly known as the American Society for Testing and Materials.


176<br />

Background on case study<br />

The Camisea system consists of a buried natural gas (NG)<br />

pipeline and a buried natural gas liquid (NGL) pipeline.<br />

The NGL pipeline transports liquid condensates from<br />

Malvinas in the Peruvian Amazon to a fractionation plant<br />

near Pisco, on the coast of Peru south of Lima (see Fig.1).<br />

The pipeline starts in the jungle (“selva”) and climbs up the<br />

east slopes of the Andes Mountains (“sierra”) to a height of<br />

approximately 4,800m, from where it drops steeply towards<br />

the coastal (“costa”) city of Pisco. The NGL pipeline is<br />

approximately 561km long, and telescopes from a nominal<br />

pipe diameter of 14 to 10.75in. The wall thickness of the<br />

NGL pipeline ranges between 0.219 and 0.469in. The 734km<br />

long and larger-diameter NG pipeline shares the same<br />

right-of-way (RoW) along its initial 550km until it follows<br />

the Pacific coast towards Lima. Both pipelines are<br />

constructed of tubular high-strength steel (X70) in<br />

conformance with the American Petroleum Institute (API)<br />

5L standard, welded and inspected per API 1104, and<br />

The Journal of Pipeline Engineering<br />

Mountains, the Camisea transportation system, which is buried in a region where landslides and other<br />

geological hazards are common.<br />

In this paper an elastic plastic fracture mechanics analysis of a pipeline is presented that ruptured due to<br />

external soil loading, to evaluate possible loading conditions and correlate the observed crack propagation<br />

with possible external loading conditions. Next a fracture mechanics based performance criterion is derived<br />

for the most commonly used in-line inspection (ILI) methods, to detect these circumferential cracks; i.e.<br />

the magnetic flux leakage (MFL) tool.<br />

Fig.1. Alignment of the NGL and NG pipeline of the Camisea system in Peru [5].<br />

protected by a triple layer of polyethylene. All girth welds<br />

were x-rayed 24hrs after welding and evaluated per API<br />

1104 [5, 6].<br />

Since the Camisea Transportation System was brought<br />

into service in August, 2004, the NGL pipeline has<br />

experienced six spill incidents involving a release of NGL;<br />

however, no incident has been reported for the NG pipeline<br />

(see Fig.1). Three of these failures occurred at girth welds<br />

and were determined to result from soil loading due to<br />

ground movement [5]. Overall the Camisea pipeline system<br />

has experienced a spill incident rate of approximately 1.1<br />

spill incidents 2 per 1000 km year, which is slightly larger<br />

than the spill incident rate of contemporary South American<br />

pipelines through similar regions that were constructed<br />

with the newest geotechnical means. However, the spill<br />

incident rate of the Camisea system should improve because<br />

2 Using the combined length of the NG and NGL pipeline.<br />

Sample issue


3rd Quarter, 2009 177<br />

Fig.2. Failed pipe segment of fifth spill incident: tearing of the pipe occurred in the heat-influenced zone of the weld.<br />

many new geotechnical stabilization measures have been<br />

constructed along the ROW since 2006 and active<br />

monitoring of the ROW is ongoing [5].<br />

Fracture-mechanics analysis<br />

Pipeline rules for mechanical design intend to ensure<br />

integrity by requiring a set of minimum material and<br />

fabrication quality requirements and seeking to ensure that<br />

the design is such that it can reliably withstand the specified<br />

design loads (internal pressure and external). The latter is<br />

based on the strength of the material of the pipeline and<br />

there is no specific consideration given to the presence of<br />

defects such as weldment flaws, etc.<br />

When it comes to determining the fitness-for-service of a<br />

given operating pipeline, however, due consideration needs<br />

to be given to the behaviour of defects that may be indicated<br />

Fig.3. Low amplification of the pipe<br />

cross-section of the fifth spill<br />

incident.<br />

in an inspection or may be assumed with reasonable<br />

conservatism based on experience, including prior failures.<br />

The API RP 579 Recommended practice for fitness for service,<br />

for example, provides a means for evaluating the acceptability<br />

of a given crack-like flaw in a given pipeline using a failureassessment<br />

diagram that defines a region of acceptability on<br />

a fracture stress intensity factor-based parameter (Y-axis) vs<br />

strength-based parameter (X-axis). The intent is to, via<br />

conservative calculations of each parameter, establish<br />

whether the pipeline can fail by unstable crack propagation<br />

(Y-axis determined) or by plastic collapse (X-axis determined).<br />

The crack propagation portion of the estimation is based<br />

on the science of fracture mechanics that provides a means<br />

of predicting flaw tolerance or the capacity of a structure to<br />

resist the propagation of a given crack for a given set of<br />

loading conditions.<br />

The authors now discuss in detail the application of fracture<br />

mechanics to predicting the behaviour of cracks in the<br />

Sample issue


178<br />

Camisea system, including the subcritical, stable behaviour<br />

prior to unstable fracture or plastic collapse.<br />

In order to quantify the impact of ground movement,<br />

fracture mechanics was used to relate an assumed defect<br />

size to the failure load and to gain some insight into how<br />

quickly tearing can occur. To assess the soil loading and<br />

observed crack growth behaviour, an elastic-plastic fracture<br />

analysis was conducted for the tearing process that led to<br />

the fifth spill incident at KP 125+950, where a<br />

circumferential crack initiated on the outside of the pipe<br />

and grew by progressive tearing to a through-wall 275-mm<br />

long circumferential crack; Fig.2 shows the failed pipe<br />

segment after complete rupture of this crack. The fracture<br />

surface of the crack has three distinct stages of crack<br />

propagation, which are shown in Fig.3; in the figure, the<br />

top is the outside wall of the pipe and the blue arrows<br />

marking the transition between stages.<br />

In this spill incident, the mechanism of both crack initiation<br />

and propagation was ductile tearing caused by soil<br />

movement. It should be noted that under high axial loads,<br />

the biaxial stress state of the pipe could effectively increase<br />

the axial load needed to cause general yielding and allow<br />

tearing of the weld material prior to the general yielding of<br />

the pipe. After this ductile tearing through the majority of<br />

the pipe wall, the final slant fracture occurred by plastic<br />

collapse of the remaining ligament. The difference in the<br />

fracture surface in the three stages of crack growth shown<br />

in Fig.3 may be due to the differences in the rate of ductile<br />

tearing.<br />

In ductile materials, plastic tearing ahead of a crack initiates<br />

when the driving force of the crack, J, reaches J IC . However,<br />

once tearing begins there can be stable crack advance due<br />

to the increasing resistance of the material to crack advance,<br />

The Journal of Pipeline Engineering<br />

which is typically summarized with a J-R curve. A J-R curve<br />

for API 5L X70 steel is shown in Fig.4 [7].<br />

As can be seen in Fig.4, as the length of the crack, Da,<br />

increases, the driving force required for further crack<br />

advance increases. Unstable fracture will occur when the<br />

rate of change of the crack driving force with crack length<br />

becomes greater than the rate of change of the resistance<br />

curve with crack length for a given loading or the remaining<br />

ligament fails by plastic collapse: for API 5L X70, J is IC<br />

reported to be approx. 400 kJ/m2 [7]. In the initial stages of<br />

crack growth, corresponding to the ductile tearing in stage<br />

one in Fig.3, the resistance curve is relatively steep. As the<br />

crack grows, however, the slope of the resistance curve<br />

decreases and it is possible that a transition to a more rapid<br />

ductile tearing begins when the crack reaches a length of<br />

1.4 mm, which corresponds to the blue arrow indicating<br />

the end of the first stage in Fig.3. This transition may<br />

correspond to this change in slope of the resistance curve<br />

seen in Fig.4. This more rapid ductile tearing continues<br />

until the rate of change of the crack driving force with crack<br />

length becomes greater than the rate of change of the<br />

resistance curve with crack length for a given loading or the<br />

remaining ligament fails by plastic collapse<br />

Sample issue<br />

Fig.4. J-R resistance curve for<br />

API 5L X70.<br />

The load necessary to reach a value of J IC = 400kJ/m 2 was<br />

computed using a SENT (single edge notch tension) with a<br />

crack of length a = 1.4mm. The stress in the wall was found<br />

to be approx. 95ksi, greater than the measured yield stress<br />

(82ksi) of the pipe material. This would be consistent with<br />

the observed plastic deformation and lateral contraction of<br />

the pipe. This wall stress correlates to a 1,205kip load in<br />

addition to the typical operating pressure of 2,320psi. It<br />

should be noted that the calculation of the load based on<br />

J is only approximate because J is very sensitive to the stress<br />

near J IC and tearing would probably occur before the


3rd Quarter, 2009 179<br />

Fig.5. EPRI-J based failure<br />

assessment diagram for a centrecracked<br />

panel.<br />

remaining ligament reaches the flow stress (at an additional<br />

load of 716kips).<br />

In order to gain some insight in to how quickly the tearing<br />

may have occurred, the load increase was calculated for a<br />

crack advance Da = 0.25mm. That load increase was found<br />

to be only a 2% increase of the load that had already been<br />

applied to initiate tearing, which means that once the crack<br />

begins to tear through the wall the progressive growth can<br />

happen over a short period of time. Hence, pipelines that<br />

are subjected to continuous ground movement are at an<br />

elevated risk since progressive tearing only requires a small<br />

increase in external loading.<br />

The EPRI-J based failure assessment diagram for centrecracked<br />

panel was used to assess the stability of the cracked<br />

pipe before failure occurred (see Fig.5). The centre-cracked<br />

panel is a close approximation to the SENT solution used<br />

for the J analysis. The failure assessment diagram allows a<br />

body with a crack to be assessed for failure from both crack<br />

growth and plastic collapse. When the point is inside the<br />

curve there will be no plastic collapse. Using the loads<br />

calculated from the J analysis, J r and S r were calculated<br />

to be 0.2 and 1.38 respectively for a 1.4mm deep crack and<br />

n = 10 [8]. Although there is no curve available for<br />

a/w = 0.15 it can be seen that the pipe wall would be on the<br />

verge of plastic collapse. The final slant fracture occurred by<br />

plastic collapse of the remaining ligament after the ductile<br />

tearing. It should be noted that the measured yield strength<br />

is higher than the specified minimum yield strength of<br />

70ksi, so that if the material yield point were lower plastic<br />

collapse would have occurred earlier.<br />

The elastic-plastic analysis shows that an incremental tearing<br />

of 0.25mm can be caused by as little as a 2% increase in the<br />

total applied load, and this indicates that cracks can<br />

propagate easily once they have begun ductile tearing.<br />

Therefore pre-existing circumferential cracks should be<br />

detected before reaching a depth of 1.4mm in cases where<br />

external loading could be significant.<br />

Crack mouth opening<br />

displacement and ILI of<br />

circumferential cracks<br />

1.4 mm deep<br />

crack at J 1C<br />

Following the above analysis, an ILI tool would need to be<br />

able to reliably detect circumferential cracks with a minimum<br />

depth of 1.4mm. Several API 1163-compliant ILI solutions<br />

are available that could theoretically detect circumferential<br />

crack-like features: one is a high-resolution MFL tool; the<br />

other one could be a slightly-modified ultrasound ILI tool.<br />

The MFL tool has traditionally been used as an ILI tool to<br />

detect material loss and characterize corrosion damage.<br />

Most inspection companies advertise their high-resolution<br />

MFL tools to be capable of detecting crack-like features<br />

with a minimum face opening of 0.1mm. Tuboscope (TPS)<br />

developed a detection and accuracy specification for<br />

circumferential cracks for its high-resolution MFL tool.<br />

This specification states that for a pipe thinner than 0.344in<br />

(8.74mm), cracks of a length of 25mm with a depth of more<br />

than 25% can be found at a probability of detection (PoD)<br />

of 90% only when the minimum crack opening is at least<br />

0.1mm. Similarly, for a thicker pipe, the 25-mm long crack<br />

needs to be 30% deep to be detected at a PoD of 90%.<br />

Sample issue<br />

Ultrasound ILI tools are currently the most reliable tools to<br />

detect tight axial cracks with no minimum crack opening<br />

requirement. However, some technical issues will need to<br />

be addressed in order to run an ultrasound ILI tool to<br />

detect circumferential rather than axial cracks; these<br />

technical changes to the UT tool appear to be surmountable.<br />

In order to evaluate the capability of the MFL detection<br />

technique to detect the above-discussed circumferential


180<br />

Crack Mouth Opening Displacement (inches)<br />

0.004<br />

0.003<br />

0.002<br />

0.001<br />

δ<br />

0.1 mm<br />

a<br />

σ<br />

t<br />

σ<br />

0.000<br />

0 500 1000 1500 2000 2500<br />

cracks under either the MAOP or an external soil load that<br />

will yield the entire pipe, an analysis must be performed to<br />

evaluate the crack opening for such cracks under such<br />

loads.<br />

To put an upper bound on the crack mouth opening<br />

displacement (CMOD), the crack geometry is taken as the<br />

crack shown in Fig.6 with the pipe cracked all the way<br />

around the circumference of the pipe. The solutions were<br />

performed for a pipe with R/t = 18.2, a hardening exponent<br />

of n = 10, and different depths into the wall (see Fig.6). This<br />

solution should slightly overestimate the crack opening of<br />

a finite length crack in the actual pipe since the crack is not<br />

all the way around the circumference of the pipe.<br />

Figure 6 shows that under the design operating pressure of<br />

2,700psi (72% SMYS), even a crack that has propagated<br />

halfway through the wall is not reliably detectable by the<br />

MFL technique, since the crack mouth opening is still less<br />

than 0.1mm. At an assumed operating pressure of<br />

approximately 2,320psi a circumferential crack that is 15%<br />

of the way through the wall would have a crack mouth<br />

opening that is an order of magnitude less than the detection<br />

limit of the MFL technique.<br />

In this context it is also useful to consider what the crack<br />

mouth opening displacement would be under both the<br />

operating pressure and external loads such as soil loads.<br />

The CMOD versus applied bending moment and versus<br />

applied tension with the same crack configurations and<br />

pipe geometry as given in Fig.6 are plotted in Figs 7 and 8,<br />

respectively. In both cases the applied load is in addition to<br />

the operating pressure of 2,320psi (86% of the MAOP).<br />

a<br />

Internal Pressure (psi)<br />

a=0.15t<br />

a=0.25t<br />

a=0.5t<br />

Cracked Region<br />

t<br />

Operating<br />

Pressure<br />

The Journal of Pipeline Engineering<br />

The loads at which the cracks could fail by plastic collapse<br />

of the remaining ligament are circled in red in Figs 7 and<br />

8: when the crack is 15% of the wall thickness, an external<br />

soil load of 720kips would cause the remaining ligament to<br />

reach the flow stress (76,000psi) of the material. At this<br />

point the crack opening is only 0.03mm and below the<br />

detection capability of the MFL technique. When the crack<br />

is 25% of the wall thickness, an external soil movement<br />

inducing a tensile load of 595kips would cause the remaining<br />

ligament to reach the flow stress (76,000psi) of the material.<br />

At this point the crack opening is only 0.06mm, which is<br />

still below the detection capability of the MFL technique.<br />

Therefore the MFL tool is most likely not capable of reliably<br />

detecting this size of circumferential cracks prior to them<br />

being susceptible to progressive tearing by small increases<br />

in external loadings.<br />

Sample issue<br />

Fig.6. Plastic CMOD of an outside<br />

circumferential for three different<br />

crack depth versus internal pressure<br />

for a pipe with dimensions R/t =<br />

18.2 and t = 0.375in.<br />

The repeat inspection interval required to mitigate this risk<br />

is determined by computing the difference between the<br />

time to detection, T det , i.e., the amount of time needed<br />

under the loading conditions for the crack to grow to a<br />

detectable size at high probability and confidence for the<br />

chosen method, and T c , the time needed under the loading<br />

conditions for the crack to grow to a critical size. This gives<br />

rise to the computation of the safe inspection interval T s ,<br />

which is simply the difference T c - T det . The repeat inspection<br />

interval is then typically chosen to be a fraction of this the<br />

safe inspection. If the length of time is short between when<br />

the crack can be detected and when the crack is critical, a<br />

short repeat inspection interval is obtained that would be<br />

economically and logistically not viable.<br />

Following the above, it is imperative that pipelines in


3rd Quarter, 2009 181<br />

Fig.7. Plastic CMOD versus<br />

bending moment and internal<br />

pressure of 2,320psi for a pipe<br />

with dimensions R/t = 20 and t =<br />

0.375in.<br />

Fig.8. Plastic CMOD versus tensile<br />

load and internal pressure of<br />

2,320psi for a pipe with<br />

dimensions R/t = 18.2 and t =<br />

0.375in.<br />

Crack Mouth Opening Displacement (inches)<br />

Crack Mouth Opening Displacement (inches)<br />

0.005<br />

0.004<br />

0.003<br />

0.002<br />

0.001<br />

adverse environments benefit from a detailed geological<br />

and geotechnical evaluation, where all potential geotechnical<br />

hazards are properly identified and the pipeline engineer<br />

consults with competent geotechnical engineers during the<br />

design on the need for geotechnical stabilization measures.<br />

During operation, a careful monitoring programme needs<br />

to be implemented to identify any potential geotechnical<br />

hazards along the RoW. This hazard identification should<br />

Internal Pressure = 2,320 psi<br />

a=0.15t<br />

a=0.25t<br />

δ<br />

a<br />

σ<br />

t<br />

σ<br />

0.1 mm<br />

Cracked Region<br />

0.000<br />

0 50 100 150 200<br />

0.010<br />

0.009<br />

0.008<br />

0.007<br />

0.006<br />

0.005<br />

0.004<br />

0.003<br />

0.002<br />

0.001<br />

a<br />

Additional Bending Moment (kips*ft)<br />

a<br />

δ<br />

0.000<br />

0 200 400 600 800<br />

t<br />

Outer Circumferential Crack a=0.15t<br />

Outer Circumferential Crack a=0.25t<br />

Internal Pressure = 2,320 psi<br />

Cracked Region<br />

σ<br />

Sample issue<br />

t<br />

a<br />

Tensile Load (kips)<br />

t<br />

σ<br />

0.1 mm<br />

be integrated into the operator’s overall risk-assessment<br />

methodology and preferably follow guidelines like API<br />

1160, which allow for a rationale evaluation of geotechnical<br />

risks. Exponent’s engineers and engineering geologists<br />

have recently developed a geotechnical risk assessment<br />

method that has now been used to identify and evaluate the<br />

risk along the RoW of the Camisea pipeline in the jungle<br />

and mountains [5].


182<br />

Conclusion<br />

Our presented non-linear elastic-plastic fracture mechanics<br />

analysis of the fifth spill incident indicates that the high<br />

toughness of both the pipe material and the weld material<br />

necessitated the combination of extremely high axial loads<br />

(which exceeded the uniaxial yield strength of the material)<br />

and a biaxial stress state to propagate defects at the<br />

circumferential girth welds. The biaxial stress state, caused<br />

by the internal pressure, essentially increases the load<br />

needed to cause yielding in the pipe and can allow ductile<br />

tearing to occur in the girth welds before the load reaches<br />

this higher yield point for biaxial stress. A substantial axial<br />

load can cause ductile tearing of defects that are considered<br />

allowable by API 1104, since girth welds may be code<br />

compliant, but may still contain allowable deviations from<br />

the ideal condition allowed in the standard. These minor<br />

defects will ultimately provide the stress risers to initiate<br />

cracks as high loads arise. Although the exact propagation<br />

rate was not determined, it seems likely that this tearing<br />

could occur over a short time: this limits the ability of (ILI)<br />

tools in identifying these defects in a timely fashion.<br />

Currently only an ultrasound ILI with a modified sensor<br />

arrangement would reliably detect circumferential cracks<br />

within an actionable timeframe to safeguard the pipeline<br />

against potential geotechnical hazards. However, at present,<br />

insufficient evaluation has been conducted to establish the<br />

relationship between the onset of tearing and the shape,<br />

size, and orientation of welding defects. The timescales of<br />

the ductile tearing have also not been established sufficiently<br />

to rely upon periodic inspection. Hence, detailed<br />

geotechnical studies prior to construction, careful<br />

monitoring of the pipeline’s RoW, and geotechnical risk<br />

assessments appear currently to be the most comprehensive<br />

way to safeguard the integrity of the pipeline. Constructing<br />

the necessary geotechnical stabilization measures can<br />

typically mitigate geotechnical hazards.<br />

When designing pipeline systems, particularly for use in<br />

susceptible geological environments, it is important to<br />

recognize that axial loads generated from soil movement<br />

The Journal of Pipeline Engineering<br />

can be high enough to propagate relatively-small<br />

circumferential flaws and cause failure. Currently-available<br />

commercial pipeline inspection methods do not appear to<br />

provide a suitable means of detecting such flaws and<br />

guarding against failure due to propagation of these<br />

relatively-small flaws. Therefore, in susceptible geologic<br />

environments, extra attention must be given during the<br />

design phase to specifically include suitable and conservative<br />

assumptions of the external loading.<br />

References<br />

1. EGIG, 2005. Report 1970-2004. Gas pipeline incidents.<br />

6th Report of the European Gas Pipeline Incident Data<br />

Group, Doc. Number EGIG 05.R.0002, December.<br />

2. PRCI, 2004. Guidelines for the seismic design and assessment<br />

of natural gas and liquid hydrocarbon pipelines. Catalog No.<br />

L51927, October.<br />

3. Exponent Failure Analysis, 2007. Report: Integrity analysis<br />

of the Camisea transportation system, Peru. Submitted to<br />

the Inter-American Development Bank, June.<br />

4. A.P.S. Selvadurai, J. J. Lee, R.A.A. Todeschini, and H.F.<br />

Somes, 1983. Lateral soil resistance in soil-pipe interaction.<br />

Proc. Conf. <strong>Pipelines</strong> in adverse environments 11, San<br />

Diego, California, November.<br />

5. M. Sweeney, A. H. Gasca, M. G. Lopez, and A. C. Palmer.<br />

<strong>Pipelines</strong> and landslides in rugged terrain: a database, historic<br />

risks and pipeline vulnerability.<br />

6. Germanischer Lloyd, 2007. Report: Auditoria integral de<br />

los sistemas de transporte de gas natural y liquidos de gas<br />

natural del proyecto Camisea. Final audit report of the<br />

Ministerio de Energia y Minas del Peru, No. GLP/GLM/<br />

MEMP/726-07, October.<br />

7. Mauricio Carvalho Silva, Eduardo Hippert Jr., and Claudio<br />

Ruggieri, 2005. Experimental investigation of ductile tearing<br />

properties for API X70 and X80 pipeline steels. Proc. PVP2005<br />

2005 ASME Pressure vessels and piping division conference,<br />

July 17-21, Denver, CO, USA.<br />

8. C. Ruggieri and E. Hippert Jr., 2002. Cell model predictions<br />

of ductile fracture in damaged pipelines. Fatigue and Fracture<br />

Mechanics: 33rd Volume, ASTM STP 1417, Walter G. Reuter<br />

and Robert S. Piascik, Eds, ASTM <strong>International</strong>.<br />

Sample issue


3rd Quarter, 2009 183<br />

Behaviour of wrinkled linepipe<br />

subjected to internal pressure<br />

and eccentric axial compression<br />

load<br />

by Navid Nazemi, Sara Kenno*, and Sreekanta Das<br />

Department of Civil and Environmental Engineering, University of Windsor, Windsor, ON,<br />

Canada<br />

AN NPS10 field linepipe failed due to rupture in the wrinkle. The segment of the field linepipe where<br />

the wrinkle formed and rupture occurred was situated in an unstable ground slope. The rupture<br />

occurred when the linepipe was being brought back to service after its regular shutdown. Post-failure<br />

observation indicated that the shape of the wrinkle was not symmetric and the rupture occurred at the foot<br />

of the wrinkle. The pipeline operator wanted to understand the load condition that can produce this type<br />

of asymmetric wrinkle and rupture. Therefore, two laboratory tests on NPS6 pipe specimens were<br />

conducted to investigate possible load conditions that may have created such a wrinkle and rupture in the<br />

pipe wall. The pipe specimens were tested under eccentric axial load with or without internal pressure.<br />

This paper discusses the test procedure used and results obtained from these two tests. The test results<br />

show that the load combinations applied to these two specimens were able to produce a wrinkle and a<br />

rupture that looked very similar to that of the field linepipe, and these load combinations impose a great<br />

threat to the structural integrity and safety of a field linepipe passing through an unstable slope.<br />

NATURAL GAS AND OIL are being explored in the<br />

far arctic and sub-arctic zones of Canada, USA, and<br />

other countries because of their high demand. As a result,<br />

pipelines are being laid in severe weather zones where<br />

temperature variation between summer and winter can be<br />

40º C or even higher. As a result, it is not uncommon for<br />

geotechnical movements and thermal variation to impose<br />

large forces and displacements on buried pipelines resulting<br />

in local and/or global buckling in these linepipes. Local<br />

buckling in a pipe wall – which is commonly known as a<br />

‘wrinkle’ – forms in the pipe wall if the localized compressive<br />

strain in the pipe wall is higher than yield strain of the pipe<br />

material. Material strain can localize and exceed the yield<br />

strain value if the local displacement in a linepipe is large.<br />

Such displacement may be associated with river crossings,<br />

unstable slopes, thermal load, regions of discontinuous<br />

*Author’s contact details:<br />

tel: +1 519 253 3000 x 2549<br />

e-mail: kenno2@uwindsor.ca<br />

Sample issue<br />

permafrost, and freeze-thaw of the ground. The wrinkle<br />

may grow rapidly if the linepipe is subjected to sustained<br />

deformation and/or loads. The loads and associated<br />

deformations that develop in a field buried linepipe in the<br />

arctic and sub-arctic regions can occur under various loading<br />

conditions that may be idealized as combinations of variable<br />

internal pressure, compressive axial load, lateral load, and<br />

moment.<br />

Recently, an NPS10 linepipe (linepipe with nominal<br />

diameter of 10in, or 254mm) in Canada’s sub-arctic zone<br />

failed in rupture that occurred in the wrinkle region. The<br />

segment of the pipe where failure occurred was situated in<br />

an unstable ground slope and was a buried linepipe. The<br />

segment of failed linepipe is shown in Fig.1: the rupture in<br />

the field-wrinkled NPS10 linepipe was detected when the<br />

pipeline was being brought back to normal service after<br />

regular shutdown. The pipeline operator wanted to<br />

understand why the wrinkle of that unusual shape formed<br />

and a rupture occurred and this was the motivation of the<br />

current study. From physical inspection of the ruptured


184<br />

Rupture<br />

NPS10 field linepipe, it was felt that the wrinkle formed<br />

and subsequently the rupture occurred in the pipe wall due<br />

to application of axial load and associated deformation not<br />

aligned with the axis of linepipe. However, the actual load<br />

history was not known. Therefore, the current research<br />

programme was designed and undertaken at the University<br />

of Windsor for understanding the load conditions that are<br />

able to produce a wrinkle and a rupture in the wrinkle<br />

region similar to the field NPS10 linepipe (Fig.1). Under<br />

the scope of this research programme, two laboratory tests<br />

on NPS6 (pipe with nominal diameter of 6in, or 152mm)<br />

X60 grade (API 2008) pipe specimens subjected eccentric<br />

axial load and deformation with or without internal pressure<br />

were successfully completed and the results are discussed in<br />

this paper.<br />

A literature review on wrinkle formation, growth of wrinkle,<br />

and formation of rupture in the wrinkle revealed that<br />

Top foot of wrinkle<br />

Crest of wrinkle<br />

The Journal of Pipeline Engineering<br />

Fig.1. Rupture in wrinkle of field<br />

NPS10 linepipe.<br />

several studies on pipeline local buckling (wrinkling) and<br />

failures have been undertaken over the last three to four<br />

decades (for example, Gresnigt 1986; Ju and Kyriakides<br />

1988; Yoosef-Ghodsi et al. 1995; Dorey et al. 1999; Das et<br />

al. 2000; Song et al. 2003; Einsfeld et al. 2003; Dorey et al.<br />

2006; Das et al. 2007; Zhang and Das 2008). The majority<br />

of previous research on local (wrinkling) failure of linepipe<br />

was undertaken primarily to understand when and how<br />

wrinkles forms in the field linepipes when subjected to axisymmetric<br />

axial deformation or bending deformation. The<br />

study by Das et al. (2000) shows that X52 grade wrinkled<br />

pipe subjected to monotonically increasing axi-symmetric<br />

axial deformation and internal pressure may not rupture<br />

and multiple wrinkles may form that looks like an ‘accordion’<br />

shape. The other studies by Das et al. (2001 and 2007) show<br />

that the same wrinkled pipe may, however, lose its structural<br />

integrity because of rupture formation in the wrinkle<br />

region if it is subjected to elastic-plastic strain reversals due<br />

Sample issue<br />

Fig.2. Schematic of test setup.


3rd Quarter, 2009 185<br />

Table 1. Test matrix.<br />

Specim<br />

en<br />

to cyclic deformations that develop due to temperature<br />

variations, pressure fluctuations, and freeze-thaw cycles of<br />

the ground. No studies on the investigation of initiation<br />

and growth of a wrinkle that look similar to the field NPS10<br />

linepipe (Fig.1) were found in the open literature.<br />

Test specimens and test setup<br />

The test parameters were chosen with the intention to<br />

simulate wrinkle shape and rupture that look similar to the<br />

field NPS10 buried pipeline (Fig.1). The test specimens<br />

were subjected to eccentric axial compression deformation<br />

with or without internal pressure. Axial compression load<br />

and deformation in the pipelines can occur due to several<br />

factors. The effect of temperature difference between the<br />

tie-in and operating conditions of pipeline is very important.<br />

Compressive forces may also be imposed on pipelines that<br />

are placed in unstable sloping ground because of the earth<br />

movements along the length of the pipe. However, if the<br />

movement of soil is not aligned perfectly in the longitudinal<br />

axis of the pipe an eccentric axial force develops in the pipe<br />

wall.<br />

Two test specimens were made from 152-mm (6-in) nominal<br />

diameter pipe of API X60 grade steel (API 2008). Actual<br />

outside diameter and wall thickness of the pipe were<br />

measured as 168mm and 7.3mm, respectively. Actual yield<br />

strength (SIGMAs y ) for this pipe material at 0.5% strain<br />

was obtained as 422MPa (61.2ksi) from coupon test. The<br />

modulus of elasticity of pipe steel (E) was 201GPa. Both test<br />

specimens were 800mm long and had no girth weld (plain<br />

pipe).<br />

The test setup is shown schematically in Fig.2. The ends of<br />

each pipe specimen were welded to 300-mm long, 300-mm<br />

wide, and 50-mm thick steel plates to contain the water and<br />

internal pressure. The welding was completed by a certified<br />

welder to ensure no leaks occur in the welded locations due<br />

to application of axial load and internal pressure. One set<br />

of steel collars were mounted at the ends of the pipe to<br />

avoid buckling near the connection between the pipe and<br />

the steel end plates. A swivel head was mounted at the top<br />

end of the pipe specimen to allow the specimen to rotate<br />

about the top end. The bottom end was clamped to a steel<br />

base which was clamped to the strong floor. This restricted<br />

any translation and rotation at the bottom end of the pipe<br />

specimen.<br />

Eccen<br />

tricit<br />

y,<br />

e<br />

( mm)<br />

Pressur e test<br />

Axia<br />

l Load<br />

test<br />

Constant<br />

pressure<br />

( MPa)<br />

1<br />

20<br />

9.<br />

6<br />

( 0.<br />

25py<br />

)<br />

2 0.<br />

0MPa<br />

End<br />

conditions<br />

pin-fixed<br />

As shown in Fig.2, an eccentric axial compression load (P a )<br />

was applied to the specimen through a universal hydraulic<br />

loading actuator with an eccentricity (e) of 20mm. The<br />

required internal pressure was applied by filling the pipe<br />

with water and then pressurizing it using an air-driven<br />

water pump.<br />

Instrumentation<br />

The vertical load was controlled and measured through a<br />

3000-kN capacity load cell. Two vertical linear-variable<br />

differential transformers (LVDT) were mounted between<br />

the top swivel head support and the solid steel base, as<br />

shown in Fig.2, to measure the vertical deformation of the<br />

pipe. One pressure transducer was used to control and<br />

acquire the internal pressure. A digital inclinometer was<br />

installed on the top end plate to acquire the rotations of the<br />

top end of the pipe.<br />

Electrical resistance strain gauges of 5-mm gauge length<br />

were installed to measure localized strains in the longitudinal<br />

direction; the gauges were installed before application of<br />

load and pressure, and therefore measured the local strains<br />

from the beginning of the test. Post-buckling strain-gauge<br />

readings on the wrinkle vary rapidly from one point to<br />

another depending on their positions relative to the crest<br />

or foot of the wrinkle. Since the location of the wrinkle is<br />

unknown at the beginning of the test, the strain gauges<br />

were spaced closely in the 300-mm long space at the midheight<br />

of the pipe specimen to measure local strains over<br />

the entire wrinkle region.<br />

Sample issue<br />

Test procedure<br />

and test behaviour<br />

Maximum<br />

pressure<br />

( MPa)<br />

5.<br />

4<br />

( 0.<br />

14py<br />

)<br />

7.<br />

7<br />

( 0.<br />

20py<br />

)<br />

End<br />

conditions<br />

Free<br />

at<br />

top<br />

Each specimen was loaded in two test steps: (i) Axial load<br />

test, when an eccentric axial loading with constant internal<br />

pressure was applied and; (ii) Pressure load test, when a<br />

monotonically increasing pressure load was applied (Table<br />

1).<br />

The internal pressure during the axial load test was chosen<br />

as 0.25p y (9.6MPa) and 0.0p y (zero pressure) for Specimens<br />

1 and 2, respectively, as shown in Table 1. The internal<br />

pressure required to yield the pipe material in the


186<br />

Y<br />

Rupture<br />

U<br />

circumferential direction is referred to as p y and the value<br />

of p y for this pipe material was found to be 38.4MPa<br />

(5560psi).<br />

First, the axial load test was undertaken on each specimen.<br />

The purpose was to simulate the initiation and growth of<br />

wrinkle for two scenarios: (i) assuming the first possibility<br />

that the wrinkle in the field linepipe may have formed and<br />

grew when the linepipe was in operation and then; (ii)<br />

assuming the other possibility, that is, the wrinkle may have<br />

initiated and grew when the linepipe was in shutdown<br />

condition. Accordingly, the internal pressure for Specimen<br />

1 was 0.25p y during the axial load test when the wrinkle<br />

formed and grew and no internal pressure was applied<br />

during the axial load test for Specimen 2 (Table 1).<br />

C<br />

Rupture<br />

Sample issue<br />

Specimen 2 Specimen 1 Field Specimen<br />

E<br />

The Journal of Pipeline Engineering<br />

Fig.3. Load-deformation behaviour.<br />

Fig.4. Final deformed shape of<br />

specimens.<br />

The load-deformation behaviours of both specimens are<br />

shown in Fig.3. The axial compression load was applied<br />

with small incremental loads at an eccentricity of 20mm to<br />

apply a moment along with the axial compression load on<br />

the specimens. As the axial load was increasing, the pipe<br />

specimens yielded (point Y in Fig.3) and then the axial load<br />

reached the maximum load carrying capacity of the pipe, as<br />

shown by point U in Fig.3. At this point, a wrinkle initiated<br />

at about the mid-height of the specimens. Application of<br />

axial load was then controlled by the displacement rather<br />

than the load. With the application of furt her axial<br />

deformation, the load-carrying capacity decreased as the<br />

wrinkle began to grow and became more asymmetric. The<br />

wrinkle at point C in Fig.3, closed from inside pipe wall and<br />

as a result, the load carrying capacity started to increase, as


3rd Quarter, 2009 187<br />

Fig.5. Strain-load plot for Specimen 1.<br />

shown by the path C-E. The axial load test was then<br />

discontinued when the axial deformation was 75mm and<br />

77mm for Specimens 1 and 2, respectively. Small surface<br />

cracks were observed at the top foot and also at the crest of<br />

wrinkle of both specimens. However, both pipe specimens<br />

were able to maintain structural integrity and no rupture or<br />

leak occurred. It is worth mentioning that the internal<br />

pressure was maintained constant during the application<br />

of axial load and deformation.<br />

The maximum loads for Specimens 1 and 2 were obtained<br />

as 1151kN and 1176kN, respectively (Table 2). Since<br />

Specimen 2 had no internal pressure, the load carrying<br />

capacity was slightly higher than Specimen 1; however, the<br />

post-wrinkle-initiation behaviour for Specimen 1 was more<br />

stable than Specimen 2. The pipe specimens at this stage<br />

developed wrinkles that look very similar to the field<br />

NPS10 linepipe. Though both specimens maintained<br />

structural integrity when the axial load test was discontinued,<br />

Fig.6. Strain-load plot for<br />

Specimen 1.<br />

Y<br />

U<br />

-<br />

the linepipe with a wrinkle like that may not be suitable for<br />

inline inspection tool to pass through and thus, it can be<br />

considered as a deformation failure.<br />

In the next test step, that is in the pressure test, the top end<br />

of the pipe specimens were made free from rotations and<br />

translations. The internal pressure was monotonically<br />

increased and as a result, rupture occurred in the wrinkle<br />

region. The rupture occurred at the top foot of the wrinkle<br />

for Specimen 1 when the internal pressure was 5.4MPa<br />

(780psi) and at the crest of Specimen 2 when the pressure<br />

was 7.7MPa (1100psi) (see Fig.4 and Table 1). Therefore, it<br />

can be concluded that the field NPS10 linepipe was subjected<br />

to an eccentric axial load similar to the one applied in the<br />

axial load test step for the laboratory specimen. The axial<br />

load developed due to a landslide in the unstable slope<br />

where the field linepipe was situated. This might have<br />

happened when the field linepipe was in operation. Then,<br />

as the pressure was being brought back to resume the<br />

Sample issue<br />

C


188<br />

Specim<br />

en<br />

Maxim um<br />

strain<br />

( % ) Maxim<br />

um<br />

rotati<br />

on<br />

Foot Crest<br />

( dreg<br />

rees)<br />

pipeline operation after its regular shutdown, a rupture<br />

occurred at the foot of the wrinkle similar to the way the<br />

rupture occurred in Specimen 1.<br />

Strains and rotations<br />

As mentioned earlier, a large number of strain gauges were<br />

installed on the middle 300mm long segment of each pipe<br />

specimen to make an effort in acquiring the strain histories<br />

at critical locations of the wrinkle as wrinkle initiates and<br />

grows. The strain history for Specimen 1 is shown in Fig.5.<br />

From this figure, it can be observed that the strain at both<br />

feet of the wrinkle increased monotonically as the axial<br />

loading continued. The strain stabilized at both feet when<br />

the wrinkle came in contact from inside face of the pipe<br />

wall (Figs 3 and 5). However, for the crest, the strain<br />

increased only until when the maximum load capacity<br />

reached (point U in Figs 3 and 5). Then, the strain at the<br />

crest reduced slightly since the crest area relaxed as tension<br />

developed locally in the crest. The maximum compressive<br />

strains obtained from this specimen at the foot and at the<br />

crest of wrinkle are 13.7% and 3.6%, respectively (Table 2).<br />

It is important to note that these maximum compressive<br />

strain values were obtained from the axial load test and all<br />

the strain gauges failed and became non-functional when<br />

the wrinkle came in contact (point C in Fig.3).<br />

Inclinometers were installed on the top end plate of pipe<br />

specimens to measure the rotation that occurred during<br />

formation and growth of wrinkle. Figure 6 shows the<br />

rotation history of Specimen 1. It can be seen that the<br />

increase in rotation until initiation of wrinkle (point U)<br />

was small and it was only 0.37º. The rotation began to<br />

increase rapidly as soon as the wrinkle started to grow and<br />

the total rotation at top end of Specimen 1 reached to 3.3º<br />

until the wrinkle closed from inside pipe wall (point C in<br />

Fig.3). The change in rotation was very small once the<br />

wrinkle closed from inside pipe wall and the maximum<br />

rotation for Specimen 1 was 3.6º (Table 2).<br />

Similar behaviour in strain and rotation were also observed<br />

for Specimen 2. However, the maximum compressive<br />

strains at crest and foot of the wrinkle in Specimen 2 were<br />

4.9% and 13%, respectively. The maximum rotation at the<br />

top end of Specimen 2 was recorded at 2.8º.<br />

Maxim<br />

um<br />

load<br />

( kN)<br />

Ruptu<br />

re<br />

locati<br />

on<br />

1 13. 7<br />

3. 6<br />

3. 6<br />

1151 foot<br />

2 13. 0<br />

4. 9<br />

2. 8<br />

1176 crest<br />

Conclusions<br />

The Journal of Pipeline Engineering<br />

The following conclusions are made based on the<br />

experimental test data obtained from two tests conducted<br />

under the scope of this study. Therefore, these conclusions<br />

are limited to the pipe specimen and loading history that<br />

were applied to these two specimens.<br />

Sample issue<br />

1. Both test specimens produced wrinkle shape that<br />

look similar to the one observed in the field linepipe.<br />

Therefore, it can be concluded that the field NPS10<br />

linepipe was subjected to an eccentric axial load.<br />

The axial load might have developed due to<br />

movement of soil in the unstable slope.<br />

2. The shape of wrinkle and location of rupture of the<br />

field linepipe correlates better with those of<br />

Specimen 1. An axial load test on this specimen was<br />

undertaken in presence of internal pressure. Thus,<br />

it can be concluded that the wrinkle in field linepipe<br />

initiated and grew when it was in operation. The<br />

rupture occurred when the pipeline was being<br />

brought back to its normal operation after its<br />

scheduled shutdown.<br />

Acknowledgements<br />

This work was completed with financial assistance from the<br />

Natural Science and Engineering Research Council of<br />

Canada.<br />

References<br />

Table 2. Strains and rotations.<br />

American Standard (API), 2008. Specifications for Linepipe:<br />

API 5L. American Petroleum Institute, Washington, DC,<br />

USA.<br />

S.Das, J.R.Cheng, and D.W.Murray, 2007. Behavior of wrinkled<br />

steel pipelines subjected to cyclic axial loading. Canadian J.<br />

Civil Engineering, 34, pp 598-607.<br />

S.Das, J.J.R.Cheng, D.W.Murray, S.A.Wilkie, and Z.J.Zhou,<br />

2001. Wrinkle behavior under cyclic strain reversals in<br />

NPS12 pipe. Proc. 20th <strong>International</strong> Conference on Offshore<br />

Mechanics and Arctic Engineering, ASME, Rio de Janeiro,<br />

Brazil, pp129-138.<br />

S.Das, J.J.R.Cheng, D.W.Murray, S.A.Wilkie, and Z.J.Zhou,


3rd Quarter, 2009 189<br />

2000. Laboratory study of wrinkle development and strains<br />

for NPS12 linepipe. Proc, 3rd <strong>International</strong> Pipeline Conf,<br />

Calgary, Canada, pp909-915.<br />

A.B.Dorey, D.W.Murray, and J.J.R.Cheng, 2006. Critical buckling<br />

strain equations for energy pipelines – a parametric study.<br />

Journal of Offshore Mechanics and Arctic Engineering, 128, 3,<br />

pp248-255.<br />

A.B.Dorey, D.W.Murray, J.J.R.Cheng, G.Y.Grondin, and<br />

Z.I.Zhou, 1999. Testing and experimental results for NPS30<br />

pipe under combined loads. Proc. 18th <strong>International</strong><br />

Conference on Offshore Mechanics and Arctic Engineering,<br />

ASME, St. John’s, Canada, Paper No. OMAE99/PIPE-<br />

5022.<br />

R.A.Einsfeld, D.W.Murray, and N.Yoosef-Ghodsi, 2003. Buckling<br />

analysis of high-temperature pressurized pipelines with soilstructure<br />

interaction. Journal of Brazilian Society of Mechanical<br />

Science and Engineering, 25, 2, pp1-14.<br />

A.M.Gresnigt, 1986. Plastic design of buried steel pipelines in<br />

settlement areas. Heron, 31, 4, pp40-51.<br />

G.T.Ju and S.Kyriakides, 1988. Thermal buckling of offshore<br />

pipelines. Journal of Offshore Mechanics and Arctic Engineering,<br />

110, pp355-364.<br />

X.Song, D.W.Murray, and J.J.R.Cheng, 2003. Numerical solutions<br />

for pipeline wrinkling. Engineering Report AAT MQ82346,<br />

Department of Civil & Environmental Engineering,<br />

University of Alberta, Edmonton, Canada.<br />

N.Yoosef-Ghodsi, G.L.Kulak, and D.W.Murray, 1995. Some<br />

test results for wrinkling of girth-welded linepipe. Proc. 14th<br />

<strong>International</strong> Conference on Offshore Mechanics and Arctic<br />

Engineering, ASME, Copenhagen, Denmark, pp379-388.<br />

Y.Zhang, and S.Das, 2008. Failure of X52 wrinkled pipelines<br />

subjected to monotonic axial deformation. Journal of Pressure<br />

Vessel Technology, 130, 2, pp130-136.<br />

Sample issue


190<br />

Sample issue<br />

The Journal of Pipeline Engineering


3rd Quarter, 2009 191<br />

Advanced numerical modelling<br />

tools aid Arctic pipeline design<br />

by Kenton Pike<br />

JP Kenny, Houston, TX, USA<br />

A<br />

THREE-dimensional (3D) finite-element (FE) simulator tools has been developed to deal with common<br />

challenges in Arctic regions: ice gouging and permafrost. The Arctic oil and gas market has garnered<br />

much renewed attention as of late, and the team at J P Kenny is developing tools to help ensure pipeline<br />

designs for future Arctic developments are as safe and economical as possible.<br />

REMOTE LOCATIONS, extreme cold and harsh<br />

weather conditions, lack of infrastructure, difficult<br />

transportation of materials, goods and services, sensitive<br />

environments, and limited construction windows are merely<br />

some of the challenges faced when designing, constructing,<br />

and operating in Arctic regions. As a result of the high costs<br />

and long cycle times associated with developing oil and gas<br />

in the Arctic, break-even oil prices can be as high as $61/<br />

barrel. Safety of course takes precedence; however,<br />

unnecessary conservatism should be avoided to reduce<br />

expenses when technically and practically feasible.<br />

Ice gouging<br />

Ice gouging (or ice scouring) occurs when environmental<br />

forces drive ice features (icebergs or ice-ridges) that extend<br />

deeper than the water depth through the seabed soil. Ice<br />

gouging occurs offshore in Arctic and sub-Arctic regions,<br />

such as in the shallow Beaufort Sea and offshore<br />

Newfoundland. With the majority of estimated Arctic oil<br />

and gas reserves being held offshore, ice gouging could<br />

potentially govern the design of pipelines and subsea<br />

architecture for many future field developments.<br />

Current practice in pipeline design to mitigate the ice<br />

gouging hazard is to bury the pipeline deep enough to reach<br />

Author’s contact details:<br />

tel: +1 281 675 1045<br />

e-mail: kenton.pike@jpkenny.com<br />

a safety target against pipeline system failure. Contact with<br />

the keel of the ice feature is avoided; however, pipelines are<br />

installed in a region where some soil displacement can be<br />

transferred to the pipeline. It is therefore critical to<br />

accurately predict sub-gouge soil displacement.<br />

J P Kenny utilizes the Coupled Eulerian Lagrangian (CEL) FE<br />

method, available in ABAQUS/Explicit, to model the ice<br />

gouge process and has carried out extensive validation<br />

work to ensure its models behave accurately. The major<br />

advantage realized by this modelling technique is that it<br />

overcomes mesh distortion and convergence issues<br />

experienced by other methods. In the CEL FE formulation,<br />

the seabed soil is modelled using an Eulerian material that<br />

is allowed to freely flow throughout a fixed mesh. Because<br />

the mesh does not distort, very large deformations<br />

experienced during the ice gouge process can be realistically<br />

simulated.<br />

Sample issue<br />

The model (Fig.1), consisting of a rigid ice keel, Eulerian<br />

seabed, and Lagrangian pipeline, provides a fully-coupled<br />

numerical solution for ice-soil-pipeline interaction events.<br />

In running the model, the first step of the analysis allows<br />

the soil to reach an in-situ initial stress state; during the<br />

second step, the ice keel is translated through the seabed<br />

causing soil failure and displacement. The soil forms a<br />

frontal mound, displaces to the side, creating berms, and<br />

also displaces below the gouge, imposing strains on the<br />

buried pipeline. As pipeline strain demand and response<br />

are determined explicitly, the results of the model can be<br />

used to optimize pipeline burial depth requirements based<br />

on limit state-based design criteria. J P Kenny has performed


192<br />

The Journal of Pipeline Engineering<br />

Fig.1. The CEL FE ice-gouge model. Fig.2. Simplified CEL FE ice-gouge model.<br />

Fig.3. CEL FE model vs centrifuge data.<br />

validation, sensitivity and hypothetical design studies that<br />

have demonstrated the reliability and applicability of the<br />

tool.<br />

Ice-gouge model validation<br />

Modelling of the ice-gouge phenomenon has been<br />

approached through small- and medium-scale experimental<br />

testing, analytical and empirical formulations, simplified<br />

structural analyses, and advanced numerical techniques.<br />

However, uncertainty remains on the magnitude and extent<br />

of subgouge soil deformations, giving rise to uncertainty in<br />

pipeline burial depth requirements.<br />

The results obtained using the model were compared with<br />

existing experimental centrifuge data. In order to match<br />

the centrifuge testing conditions, the trench geometry,<br />

trench backfill material, and the pipeline were removed<br />

from the model, as shown in Fig.2.<br />

Sample issue<br />

The centreline horizontal subgouge displacement profile<br />

resulting from a simulation with clay soil compared to<br />

centrifuge data is shown in Fig.3. The gouge depth and<br />

width were 1.2m and 10m, respectively. The accuracy at<br />

one gouge depth is approximately 85% with good correlation<br />

extending downward.<br />

Ice-gouge model application<br />

Using the developed model, a hypothetical study was<br />

carried out assuming a 305-mm (12-in) diameter pipeline<br />

with varying wall thicknesses. The pipe was placed in a 3.61m<br />

deep trench, 1m wide at the bottom and with side slopes<br />

2V:3H. The burial depth was 3m measured from the top of<br />

the pipe. The trench soil was modelled using a softer<br />

material relative to the surrounding native seabed. The<br />

model was run a number of times for a range of D/t values<br />

and depths.


3rd Quarter, 2009 193<br />

Fig.4. A hypothetical design application for<br />

the ice-gouge model.<br />

In order to satisfy the buckling ultimate limit state, the ratio<br />

of the design compressive strain to the design resistance<br />

strain (e Sd /e Rd ) must be less than unity. Fig.4 presents this<br />

ratio versus the ratio of the clearance depth/gouge depth<br />

for the hypothetical study. The clearance depth is defined<br />

as the available soil cover below the ice keel base measured<br />

from the top of the pipe.<br />

Using this methodology, governing failure mechanisms<br />

can be investigated and required burial depths can be<br />

determined. The results of the hypothetical study indicated<br />

that lowering the value of D/t may result in a shallower<br />

burial requirement for maintenance of pipeline integrity.<br />

If the required burial depth is technically feasible, one<br />

would have to weigh trenching cost versus material cost.<br />

Figure 5 shows an isometric view of a typical FE model<br />

output, illustrating pipeline deformation and soil mound<br />

and berm formation; Fig.6 shows the same in plan view.<br />

Determining the pipeline burial depth in order to protect<br />

against encroaching ice keels has traditionally been a<br />

conservative aspect of Arctic pipeline design due to analysis<br />

procedures and a lack of explicit criteria. Traditional<br />

approaches using simple structural models (special beam<br />

elements and soil springs) do not truly represent the icegouge<br />

process, and are inherently conservative in predicting<br />

sub-gouge displacements. Realistic 3-D simulation can help<br />

reduce the unknowns associated with ice-soil-pipeline<br />

interaction, safely reduce unnecessary conservatism, and<br />

ultimately, save on trenching costs.<br />

Permafrost<br />

Permafrost is permanently frozen soil, covering about half<br />

of Canada and Russia and 85% of Alaska. The existence of<br />

permafrost presents a significant challenge to the design,<br />

construction, and operation of pipelines on Arctic terrain.<br />

<strong>Pipelines</strong> transporting warm hydrocarbons can transfer<br />

heat to the surrounding soil, causing the ground to thaw<br />

Fig.5. Isometric view of the ice-soil-pipeline interaction.<br />

Sample issue<br />

Fig.6. A plan view of the ice-soil-pipeline interaction.<br />

Fig.7. 3-D permafrost thawsettlement-pipeline<br />

interaction model.<br />

Fig.8. One-dimensional whiplash curve solution vs the FE<br />

model predictions (pure conduction).


194<br />

Fig.9. One-dimensional Neumann solution vs FE model<br />

predictions (latent heat effect).<br />

during the years of operation, and lose load-carrying<br />

capacity. Differential ground settlement is likely to occur,<br />

overstressing pipelines and inducing bending strains.<br />

JP Kenny has developed a 3-D FE model for investigating<br />

the effects of permafrost on Arctic pipelines. The model<br />

predicts unsteady-state heat transfer, thaw settlement, and<br />

global deformation processes of a pipeline buried in<br />

permafrost soil.<br />

The pipeline-permafrost soil interaction process is simulated<br />

in two decoupled steps:<br />

• unsteady-state, two-dimensional (2-D), FE simulation<br />

of heat transfer for a specific time period;<br />

• 3-D FE prediction of long-term soil settlement due<br />

to consolidation and pipeline deformation based<br />

on the mapped temperature distribution.<br />

The influence of the soil settlement on the heat transfer<br />

process is considered a second order effect.<br />

Thermal model validation<br />

The Journal of Pipeline Engineering<br />

The capabilities of the thermal model to predict pure<br />

conduction and to simulate the latent heat effect were<br />

validated by comparison with theoretical solutions. The<br />

pure conduction solution, also known as the “whiplash<br />

curve” solution, considers only the thermal diffusion in the<br />

permafrost soil with surface temperature fluctuation and<br />

geothermal gradient boundaries. Simulated temperature<br />

gradients in spring and summer in comparison to the<br />

theoretical solution are presented in Fig.8; the theoretical<br />

upper- and lower-bound temperatures due to fluctuation<br />

are also shown in the figure.<br />

So as to verify the capability of the developed model to<br />

simulate the latent heat effect, simulation results were<br />

compared with the Neumann solution, which is a onedimensional<br />

theoretical solution of simple heat conduction<br />

involving the latent heat effect. As shown in Fig.9, the<br />

results of the simulation exhibit excellent correspondence<br />

with the Neumann solution after one- and ten-month thaw<br />

periods.<br />

The model is designed to give accurate predictions of the<br />

amount of thawed ground around a pipe, and the<br />

corresponding strain on the pipeline. The thaw settlement<br />

of the pipeline is assessed based on the actual growing size<br />

of the thaw ‘bulb’, instead of the commonly-used total<br />

thickness of the thaw-unstable permafrost layers. The<br />

predictions of the current model provide clear scenarios of<br />

a pipeline’s thaw-settlement evolution and corresponding<br />

bending strains over its lifetime, and help avoid overconservative<br />

designs. Similar to the ice-gouge model, this<br />

model improves upon traditional methods that employ<br />

structural modelling.<br />

The 3-D model that has been developed can be used during<br />

the initial pipeline design phase to improve safety and cost<br />

effectiveness; the evaluation can aid the selection of proper<br />

construction and implementation methods used to<br />

minimize the impact on the permafrost and the surrounding<br />

environment. In addition, the model can be used to assess<br />

existing pipelines embedded in permafrost soils from both<br />

thermal and mechanical perspectives.<br />

Sample issue


3rd Quarter, 2009 195<br />

A disc pig model for estimating<br />

the mixing volumes between<br />

product batches in multiproduct<br />

pipelines<br />

by Etim S Udoetok* and Anh N Nguyen<br />

Colonial Pipeline Co, Alpharetta, GA, USA<br />

IN MULTI-PRODUCT PIPELINES, mixing occurs between the product batches. A new model for<br />

estimating the mixing volume, developed by modelling the leading interface as a concentric disc similar<br />

to the action of a separation pig, is presented in this paper. Comparison with field data and other models<br />

shows that the proposed new model is reliable over a wider range of flow conditions.<br />

DUE TO THE IMPORTANCE of product pipelines in<br />

the petroleum industry, there is a need for better<br />

understanding of the relationship between theory and field<br />

results for the mixing that occurs when products are<br />

transported in batches [1, 2, 3]. The problem of estimating<br />

the mixing interface is common in multi-product pipelines<br />

where the batching involves dispatching different products<br />

through a pipeline in continuous succession with no<br />

medium employed to separate the different products [4].<br />

Under these conditions there is always mixing at the<br />

boundary of two adjacent product streams, so a volume of<br />

contaminate product is formed between the two products<br />

[4]. The accurate estimation of the mixed volume affects the<br />

pipeline operation economically, because underestimating<br />

will result in contamination of product batches and possibly<br />

lead to failure to meet optimum product quality, while<br />

overestimating will result in directing valuable product<br />

into transmix tank for shipping back to refinery for<br />

reprocessing at an additional cost.<br />

A number of models have been developed for predicting<br />

the volume of the mixture, using experimental and/or<br />

theoretical approaches. However, practical field<br />

measurements over a wide range of conditions have showed<br />

that these models are not reliable in most cases. In this<br />

paper, a new model for estimating the mixing volume is<br />

*Author’s contact details:<br />

tel: +1 678 762 2467<br />

email: eudoetok@colpipe.com<br />

described based on modelling the leading interface as a<br />

concentric disc, similar to the action of a separation pig.<br />

Method<br />

The turbulent flow velocity profile gives a near-straight<br />

velocity profile in the central region of a pipe, as shown in<br />

Fig.1. The profile is described be the power-law as [5]:<br />

Sample issue<br />

max 0<br />

1 n<br />

⎛r0−r⎞ ⎜ ⎟<br />

u<br />

= ⎜<br />

u ⎜⎜⎝ ⎟<br />

r ⎟⎠<br />

and the average velocity is:<br />

2<br />

2n<br />

V = u<br />

( n+ 1)( 2n+ 1)<br />

max<br />

1<br />

where n = , r is the pipe radius, and f is friction<br />

f 0<br />

coefficient (the Appendix shows methods for calculating f).<br />

In order to estimate the mixing volume, it is assumed that<br />

the turbulence creates an imaginary and concentric disc pig<br />

at the leading interface and this disc pig prevents mixing<br />

and gives the characteristic near-straight velocity profile in<br />

the pipe centre (Fig.2). The radius, r’, of the imaginary disc<br />

pig depends on the flow conditions: it is large for high<br />

Reynolds’ number (Re) and small for low Re.<br />

(1)<br />

(2)


196<br />

Since r’ < r , the slow leading product close to the wall lags<br />

0<br />

behind the disc pig in a turbulent mixing action. The radius<br />

of the imaginary disc pig is found by equating the velocity<br />

to a fraction, e, of the average velocity:<br />

ur (') = eV<br />

Equations 1 and 2 can be combined into Equn 3 to give the<br />

disc pig radius as:<br />

⎛ n<br />

2 ⎞<br />

⎜ ⎛ 2n<br />

⎞<br />

n ⎟<br />

r′ = r ⎜<br />

0 1<br />

⎜ ⎟<br />

e ⎟<br />

⎜ −⎜<br />

⎟ ⎟<br />

⎜ ⎜<br />

⎟<br />

( n 1)( 2n 1)<br />

⎟ ⎟<br />

⎜⎝<br />

⎜ ⎝ + + ⎠⎟ ⎠<br />

⎟<br />

The fractional area not swept by the disc is given as:<br />

(3)<br />

(4)<br />

n<br />

2<br />

2 2 2<br />

A′ ⎛ ⎞<br />

πr0−πr′ ⎜ ⎛ 2n<br />

⎞<br />

n ⎟<br />

= = 1−⎜1 ⎜ ⎟<br />

e ⎟<br />

2 ⎜ −⎜<br />

⎟ ⎟<br />

A πr<br />

⎜ ⎜<br />

⎟<br />

0<br />

( n 1)( 2n 1)<br />

⎟ ⎟<br />

⎜⎝<br />

⎜ ⎝ + + ⎠⎟ ⎠<br />

⎟<br />

(5)<br />

Assuming the constant of proportionality is absorbed into<br />

e, the volume of the mixture is given as:<br />

The Journal of Pipeline Engineering<br />

Fig.1. Turbulent velocity profile. Fig.2. Disc pig separation generated by turbulence.<br />

Fig.3. Plot of volume of mixture (kerosene + gasoline) for a 12-in diameter, 100-mile long, pipeline.<br />

⎛ n<br />

2⎞<br />

⎜ ⎛ 2 ⎞<br />

⎜ ⎜ ⎛ 2n<br />

⎞<br />

n ⎟ ⎟<br />

V = ⎜1−⎜1 ⎜ ⎟<br />

e ⎟<br />

⎜ −⎜ ⎟ ⎟ ⎟ V<br />

⎜ ⎜ ⎜ ⎟<br />

( n 1)( 2n 1)<br />

⎟ ⎟<br />

⎝ + + ⎠ ⎟<br />

⎜⎜⎝ ⎜⎝<br />

⎟⎠<br />

⎟⎠<br />

Sample issue<br />

m pipe<br />

where e is constant and the only unknown to be found by<br />

experiment. In this model, the mixed volume is a function<br />

of Re, pipe roughness, and pipe length. The Reynolds’<br />

number is evaluated using the 50/50 mixture properties<br />

(see Appendix B for the viscosity equation for liquid<br />

mixtures).<br />

(6)<br />

Comparison with other models<br />

In order to check the performance of the proposed model,<br />

Equn 6, field data provided by Colonial Pipeline Co was<br />

used to find e as 0.585. For the determination of n, a<br />

roughness value of 0.000025m was used for the pipes.<br />

The models used in the comparison included Smith and<br />

Schulze [2, 3], Birge [6], Taylor [7], Sjenitzer [8], Hull and<br />

Kent [9], Jablonski [10], Austin and Palfrey [4], Levenspiel<br />

[1] (see Aappendix C for the model equations.)


3rd Quarter, 2009 197<br />

Fig.4. Plot of volume of mixture (kerosene + gasoline) for a 30-in diameter, 100-mile long, pipeline.<br />

Fig.5. Plot of volume of mixture (kerosene + gasoline) for a 30-in diameter, 350-mile long, pipeline.<br />

When the estimated volume of mixture was plotted against<br />

the Reynolds’ number (see Figs 3-5), it was observed that<br />

the proposed model compared favourably with most of the<br />

existing models. As the diameter of the line was changed<br />

(Figs 3 and 4), the models nearing the proposed models<br />

changed. Similarly, as the length of the line was changed<br />

(Figs 4 and 5), the models nearing the proposed model<br />

changed. The deviation of other models from the proposed<br />

model was high for lower Reynolds’ numbers.<br />

Sample issue<br />

For constant Reynolds’ number and varying pipe length<br />

(see and compare Tables 1-4), it was observed that some of<br />

the existing models overestimated the mixed volume for<br />

short pipe lengths (say L 400 miles) most of the models<br />

underestimated the mixing, but deviation from proposed<br />

model was low. Holding the flow rate constant and varying<br />

the diameter of the pipe also caused deviations of some<br />

existing models from the proposed model (see Tables 5-7).


198<br />

Re = 2691025<br />

Products:<br />

kerosene<br />

+ gasoline<br />

3 Model Mix volum<br />

e ( m ) Mix<br />

volum<br />

e ( brl)<br />

Smith&Schulze177. 49<br />

1.<br />

12<br />

Barge126. 70<br />

0.<br />

80<br />

Taylor169. 12<br />

1.<br />

06<br />

Sjenitzer46. 39<br />

0.<br />

29<br />

Hull&Kent149. 10<br />

0.<br />

94<br />

Jablonski320. 39<br />

2.<br />

02<br />

Austin&Palfrey172. 81<br />

1.<br />

09<br />

LevenspielN/ A<br />

N/<br />

A<br />

Proposed 72. 02<br />

0.<br />

45<br />

Re = 2691025<br />

Products:<br />

kerosene<br />

+ gasoline<br />

3 Model Mix volum<br />

e ( m ) Mix<br />

volum<br />

e ( brl)<br />

Smith&Schulze272. 79<br />

1.<br />

72<br />

Barge182. 82<br />

1.<br />

15<br />

Taylor239. 17<br />

1.<br />

5<br />

Sjenitzer68. 87<br />

0.<br />

43<br />

Hull&Kent210. 86<br />

1.<br />

33<br />

Jablonski485. 62<br />

3.<br />

05<br />

Austin&Palfrey244. 38<br />

1.<br />

54<br />

LevenspielN/ A<br />

N/<br />

A<br />

Sample issue<br />

Proposed 144. 05<br />

0.<br />

91<br />

Re = 2691025<br />

Products:<br />

kerosene<br />

+ gasoline<br />

3 Model Mix volum<br />

e ( m ) Mix<br />

volum<br />

e ( brl)<br />

Smith&Schulze419. 24<br />

2.<br />

64<br />

Barge263. 80<br />

1.<br />

66<br />

Taylor338. 23<br />

2.<br />

13<br />

Sjenitzer102. 23<br />

0.<br />

64<br />

Hull&Kent298. 20<br />

1.<br />

88<br />

Jablonski736. 06<br />

4.<br />

63<br />

Austin&Palfrey345. 61<br />

2.<br />

17<br />

LevenspielN/ A<br />

N/<br />

A<br />

Proposed 288. 10<br />

1.<br />

81<br />

The Journal of Pipeline Engineering<br />

Table 1. Length<br />

50 miles,<br />

diameter 24in.<br />

Table 2. Length<br />

100 miles,<br />

diameter 24in.<br />

Table 3. Length<br />

200 miles,<br />

diameter 24in.


3rd Quarter, 2009 199<br />

Table 4. Length<br />

400 miles,<br />

diameter 24in.<br />

Table 5. Length<br />

500 miles,<br />

diameter 16in.<br />

Table 6. Length<br />

500 miles,<br />

diameter 20in.<br />

Re = 2691025<br />

Products:<br />

kerosene<br />

+ gasoline<br />

3 Model Mix volum<br />

e ( m ) Mix<br />

volum<br />

e ( brl)<br />

Smith&Schulze644. 32<br />

4.<br />

05<br />

Barge380. 64<br />

2.<br />

39<br />

Taylor478. 33<br />

3.<br />

01<br />

Sjenitzer151. 77<br />

0.<br />

95<br />

Hull&Kent421. 72<br />

2.<br />

65<br />

Jablonski1115. 66<br />

7.<br />

02<br />

Austin&Palfrey488. 77<br />

3.<br />

07<br />

LevenspielN/ A<br />

N/<br />

A<br />

Proposed 576. 20<br />

3.<br />

62<br />

Re = 4036538<br />

Products:<br />

kerosene<br />

+ gasoline<br />

3 Model Mix volum<br />

e ( m ) Mix<br />

volum<br />

e ( brl)<br />

Smith&Schulze328. 37<br />

2.<br />

07<br />

Barge190. 37<br />

1.<br />

20<br />

Taylor189. 21<br />

1.<br />

19<br />

Sjenitzer53. 61<br />

0.<br />

34<br />

Hull&Kent155. 79<br />

0.<br />

98<br />

Jablonski481. 22<br />

3.<br />

03<br />

Austin&Palfrey190. 42<br />

1.<br />

20<br />

LevenspielN/ A<br />

N/<br />

A<br />

Sample issue<br />

Proposed 338. 87<br />

2.<br />

13<br />

Re = 3229231<br />

Products:<br />

kerosene<br />

+ gasoline<br />

3 Model Mix volum<br />

e ( m ) Mix<br />

volum<br />

e ( brl)<br />

Smith&Schulze513. 46<br />

3.<br />

23<br />

Barge297. 45<br />

1.<br />

87<br />

Taylor335. 18<br />

2.<br />

11<br />

Sjenitzer101. 95<br />

0.<br />

64<br />

Hull&Kent286. 55<br />

1.<br />

80<br />

Jablonski822. 78<br />

5.<br />

18<br />

Austin&Palfrey340. 16<br />

2.<br />

14<br />

LevenspielN/ A<br />

N/<br />

A<br />

Proposed 508. 75<br />

3.<br />

20


200<br />

Re = 1614615<br />

Products:<br />

kerosene<br />

+ gasoline<br />

3 Model Mix volum<br />

e ( m ) Mix<br />

volum<br />

e ( brl)<br />

Smith&Schulze2061. 02<br />

12.<br />

96<br />

Barge1189. 81<br />

7.<br />

48<br />

Taylor1980. 03<br />

12.<br />

45<br />

Sjenitzer750. 49<br />

4.<br />

72<br />

Hull&Kent1903. 85<br />

11.<br />

97<br />

Jablonski4360. 58<br />

27.<br />

43<br />

Austin&Palfrey2062. 35<br />

12.<br />

97<br />

LevenspielN/ A<br />

N/<br />

A<br />

Proposed 2062. 53<br />

12.<br />

97<br />

Other variations of flow conditions were made, and it was<br />

observed in all cases that there were always some models<br />

close to the proposed model. The models whose prediction<br />

approximated the proposed model more often include<br />

those by Austin and Palfrey, Hull and Kent, Taylor, and<br />

Smith and Schulze.<br />

Discussion<br />

The proposed model encompasses variables used by existing<br />

models and in addition includes the effect of the pipe<br />

roughness. The model predicts that mixing volume increases<br />

with pipe roughness. The proposed model has been<br />

developed by a simplified theoretical approach involving<br />

the finding of only one constant, and it accurately predicts<br />

field data over a wider range of flow conditions compared<br />

to existing models. Unlike some models, the dispersion<br />

coefficient or Schmidt’s number were not used to formulate<br />

the proposed new model. The new model is directly<br />

proportional to the pipe length unlike other models, which<br />

are mostly proportional to the square root of the pipe<br />

length; therefore, applying the proposed model to a pipeline<br />

with sections of varying diameter gives a better estimate of<br />

the flow mixing. The new model is based on turbulent flows<br />

and works for all turbulent flow conditions. The proposed<br />

new model agrees with most models that maintaining the<br />

velocity as high as possible minimizes mixing. The new<br />

model also implies that if the Reynolds’ number tends to<br />

infinity, the pipe is perfectly smooth, and f tends to zero,<br />

the turbulent velocity profile is perfectly straight and there<br />

will be no interfacial mixing.<br />

However, the proposed model and other models may<br />

deviate from practical data due to factors such as pump<br />

start-ups and shutdowns, varying flow rates, valve switching<br />

time between product batches, and complicated pipe<br />

networks at the depots and pump stations.<br />

Conclusions<br />

The Journal of Pipeline Engineering<br />

A new model for predicting the mixing volume between<br />

product batches has been developed by modelling the<br />

leading interface as a concentric disc, similar to the action<br />

of a separation pig. The separation disc is assumed to have<br />

its outer radius as the radius where the pipe velocity is equal<br />

to 58.5% of the average velocity. The model uses the<br />

Reynolds’ number, pipe roughness, and pipe length to<br />

accurately predict the mixed volume over a wider range of<br />

conditions than existing models.<br />

Symbols used<br />

Sample issue<br />

A = area (m2 )<br />

D = diameter of pipe (m)<br />

E = constant (d” 1)<br />

F = friction coefficient<br />

K = dispersion coefficient<br />

L = pipe length (m)<br />

n = power-law constant<br />

Re = Reynolds’ number<br />

r, z = cylindrical coordinates<br />

r = pipe radius (in, m)<br />

0<br />

u = velocity at r (m/s)<br />

u = maximum velocity usually at r = 0 (m/s)<br />

max<br />

V = average velocity (m/s)<br />

V = volume (m3 , Mbbl)<br />

e = pipe roughness (m)<br />

n = kinematic viscosity<br />

References<br />

Table 7. Length<br />

500 miles,<br />

diameter 40in.<br />

1. O.Levenspiel, 1958. How much mixing occurs between<br />

batches? Pipe Line Industry, 5, pp51-54.<br />

2. S.S.Smith and R.K.Schulze, 1948. Interfacial mixing


3rd Quarter, 2009 201<br />

characteristics of products in products pipeline – Part 1. The<br />

Petroleum Engineer, 20, 8, pp94-104.<br />

3. Idem, 1948, 20, 9, pp7-12.<br />

4. J.E.Austin and J.R.Palfrey, 1964. Mixing of miscible but<br />

dissimilar liquids in a serial flow in a pipeline. Proc. Inst.<br />

MechE, 178, 1, 15, pp377-395.<br />

5. M.C.Potter and D.C.Wiggert, 2002. Mechanics of fluids,<br />

3rd Edn. Brooks/Cole California, pp298-305.<br />

6. E. A.Birge, 1947. Contamination control in products<br />

pipelines. Oil and Gas Journal, 46, p176.<br />

7. G.I.Taylor, 1954. The dispersion of matter in turbulent flow.<br />

Proc. R. Soc., 223, p446.<br />

8. F.Sjenitzer, 1958. How much do products mix in a pipeline?<br />

The Pipeline Engineer, D31-D34.<br />

9. D.E.Hull and J.W.Kent, 1952. Radioactive tracers to mark<br />

interfaces and measure intermixing in pipelines. Ind. Eng.<br />

Chem., 44, 11, p2745.<br />

10. V.S.Jablonski, 1946. Neftjanoje Chozjajstvo, 2, p56.<br />

11. T.K.Serghides, 1984. Estimate friction Factor accurately.<br />

Chemical Engineering, 91, 5, pp63-64.<br />

12. S.E.Haaland, 1983. Simple and explicit formulas for the<br />

friction factor in turbulent flow. Trans ASIVIE, J. of Fluids<br />

Engineering, 103, 5, pp89-90.<br />

13. W.R.Gambill, 1959. How to estimate mixtures viscosities.<br />

Chemical Engineering, 66, pp151-152.<br />

The paper continues with the Appendix, on pages 202-204<br />

Sample issue


202<br />

Appendix<br />

Sample issue<br />

The Journal of Pipeline Engineering


3rd Quarter, 2009 203<br />

Appendix (continued)<br />

Sample issue


204<br />

Appendix (continued)<br />

Sample issue<br />

The Journal of Pipeline Engineering


3rd Quarter, 2009 205<br />

Soil reaction force at the head of<br />

the pipeline during the pull-back<br />

operation of horizontal<br />

directional drilling<br />

by J P Pruiksma 1 , H J Brink 2 , H M G Kruse 1 , and J Spiekhout* 2<br />

1 Deltares, National Institute Unit Geo-Engineering, Delft, Netherlands<br />

2 NV Nederlandse Gasunie, Groningen, Netherlands<br />

THE SUCCESS OF A horizontal-directional drilling (HDD) project is largely dependent on the success<br />

of the pull-back operation, when the product pipe is installed in the created borehole. For global design<br />

purposes and global engineering practice, a calculation method is available in the Dutch standard. This is a<br />

quick and relatively simple method for the calculation of the distribution of normal forces between the<br />

pipeline and the borehole wall, and gives a reasonable estimate of the maximum pull-back force. The reason<br />

for possible pulling problems, however, cannot be explained with this method. Therefore knowledge of the<br />

behaviour of the head of the pipeline at the connection with the pull-back equipment is required.<br />

This behaviour was investigated using a pull-back model based on the finite-element code ABAQUS. Since<br />

the model calculations consume a considerable amount of time, the results are compared with analytical<br />

solutions. It is observed for which situations analytical solutions can be used to calculate the soil reaction<br />

force relatively quickly. Forces on the coating and borehole wall penetrations during the pull-back operation<br />

can therefore be assessed relatively easy.<br />

Further research into this subject from members of the same team of authors, plus others, and titled<br />

Horizontal directional drilling – the influence of uplift and downlift during the pull-back operation of the steel pipeline<br />

string, will be published in the Fourth Quarter, 2009, issue of the Journal of Pipeline Engineering.<br />

THE SUCCESS OF A horizontal-directional drilling<br />

(HDD) project is largely dependent on the success of<br />

the pull-back operation, when the product pipe is installed<br />

in the created borehole. The cost of damaged pipelines and<br />

the costs of additional measures during and after the pullback<br />

operation can be considerable [1]. Recently, in the<br />

Netherlands, problems occurred during the pull-back<br />

operation at some locations where relatively large diameter<br />

pipelines where installed: the problems varied from high<br />

pulling forces to uncompleted pull-back operations due to<br />

a jammed pipeline. The nature of the pull-back problems is<br />

related to the pipeline-soil interaction during the pull-back<br />

operation.<br />

*Author’s contact details:<br />

tel: +31 5 0521 2190<br />

email: j.spiekhout@gasunie.nl<br />

Sample issue<br />

The current Dutch method for calculating the pull-back<br />

force on the product pipe is based on the soil-pipeline<br />

interaction, and considers the distribution of the normal<br />

forces between the pipeline and the wall of the pre-reamed<br />

borehole. The method was developed more then ten years<br />

ago [2]. For global design purposes and global engineering<br />

practice, it is a quick and relatively simple method for the<br />

calculation of the distribution of normal forces between<br />

the pipeline and the borehole wall, and gives a reasonable<br />

estimate of the maximum pull-back force. The reason for<br />

the pulling problems, however, cannot be explained with<br />

this method. Recent research explained that the behaviour<br />

of the head of the pipeline is of major importance in the<br />

pull-back operation [5], and therefore the behaviour of the<br />

head of the pipeline at its connection with the pull-back<br />

equipment in the curved sections of a horizontal-directional<br />

drilling has been investigated in more detail.


206<br />

Engineering practice<br />

The pull-back force calculation used in engineering practice<br />

is partly based on friction between the pipeline and the<br />

borehole wall [2], and is described in the Dutch Standard<br />

[4]. The magnitude of the friction is caused by the normal<br />

force which the pipeline exerts on the borehole wall. In<br />

curved sections of the borehole, the pipe is subjected to<br />

elastic bending. From elementary beam theory it is known<br />

that if a beam is bent into a perfect circle the bending<br />

moment is:<br />

EI<br />

M = (1)<br />

R<br />

where EI is the bending stiffness of the beam (pipeline) and<br />

R the circle radius. EI can be calculated as follows:<br />

π<br />

EI E D D<br />

64<br />

4 4<br />

= ( 0 − i )<br />

(2)<br />

in which E is the Young’s modulus of the pipeline material,<br />

D 0 the outer diameter, and D i the inner diameter of the<br />

pipeline.<br />

The bending moment can only exist if the pipeline is able<br />

to mobilize reaction forces, and the forces of the moment<br />

The Journal of Pipeline Engineering<br />

must be provided by the soil. The Dutch Standard uses<br />

Hetényi’s theory [3a] to calculate the soil reaction forces.<br />

This is done by applying a moment on an infinite beam on<br />

elastic foundation (see also Fig.1).<br />

Hetényi’s solution can be written as:<br />

Sample issue<br />

Fig.1. Soil reaction at the end of an<br />

elastic bend in an infinite pipeline.<br />

Fig.2. Soil reaction when the head of<br />

the pipeline is located in a bend.<br />

2<br />

EIλ<br />

−λ<br />

x<br />

Qrbend , = kvy = e sin λx<br />

(3)<br />

DR<br />

0<br />

kD<br />

4 v o<br />

`λ<br />

= (4)<br />

4EI<br />

where:<br />

x = position, moment is applied at x = 0<br />

Q r, bend = maximum soil reaction near the end of the<br />

bend (N/mm 2 )<br />

k v = vertical modulus of subgrade reaction (N/mm 3 )<br />

y = displacement (mm)<br />

EI = bending stiffness of the pipe (Nmm 2 )<br />

R = radius of the bend (mm)<br />

It can be shown that the equation has a maximum for<br />

lx = p/4. This gives the maximum reaction stress in the<br />

subsequent equation.<br />

Equation 5 is used in the Dutch Standard [4] for calculating<br />

the maximum soil reaction force. This equation gives the


3rd Quarter, 2009 207<br />

Fig.3. Pipeline and drill pipe<br />

(consisting of a number of<br />

elements) are bent into the shape<br />

of the bore path.<br />

Fig.4. One element of a pipeline or<br />

drill pipe and its interaction with soil<br />

and drilling fluid.<br />

spring force F(u n )<br />

-b<br />

maximum soil reaction at the crossover point from straight<br />

bore path section to circular bend, or the crossover point<br />

from circular bend to straight section, when the head of the<br />

pipeline is located beyond the curved section (bend).<br />

2<br />

0.322λ<br />

EI<br />

Qrbend , = kvy= (5)<br />

DR<br />

o<br />

This calculation method is valid when the pipeline is<br />

pulled-in entirely in the curved section (bend) of the<br />

borehole. When only the head of the pipeline is located in<br />

a curved section or close to a curved section, the distribution<br />

of the soil reaction forces is different.<br />

b<br />

plastic<br />

k(u n )<br />

1<br />

pipe line<br />

friction force f w<br />

u n<br />

u t<br />

nel_pipe<br />

net force,<br />

submerged weight<br />

wall of bore hole<br />

pipe line wall<br />

u n<br />

spring k(u n )<br />

Soil reaction at the head of the<br />

pipeline inside a bend<br />

To calculate the maximum soil reaction at the head of the<br />

pipeline when it is in a circular bend, Hetényi’s theory [3b]<br />

is used. Hetényi gives a solution for a semi-infinite beam on<br />

elastic foundation with a moment applied at the end<br />

(Fig.2). The solution for the soil reaction stress can be<br />

written as:<br />

2<br />

2EIλ<br />

−λ<br />

x<br />

Qrhead , = kvy = e (cosλx−sin λx)<br />

DR<br />

0<br />

u n<br />

nodes on centre line<br />

of bore path<br />

Sample issue<br />

drilling<br />

pipe<br />

nel<br />

(6)<br />

Δu<br />

nodes on beam elements<br />

b= gap between pipe and wall<br />

centre of bore hole<br />

Fig.5. Spring stiffness of tube support element between beam and borehole.


208<br />

The Journal of Pipeline Engineering<br />

Fig.6. Simulation results of the soil reaction forces for a pipeline lying in the upward circular bend of a bore path and at<br />

certain distances beyond the bend.<br />

Sample issue<br />

Fig.7. Borehole wall penetration as a results of soil-reaction forces.


3rd Quarter, 2009 209<br />

Fig.8. Simulation results of the soil-reaction forces for a pipeline lying in the upward circular bend of a bore path and at<br />

certain distances beyond the bend. A gap of 100mm around the pipeline represents the borehole.<br />

Sample issue<br />

Fig.9. Displacement normal to bore path for simulations with 100-mm gap. The borehole wall penetration occurs when<br />

displacements are larger than +100mm or smaller than -100mm.


210<br />

if the moment is applied at x = 0 and the pipeline extends<br />

from x = 0 to infinity. It can be shown that Q r has a<br />

maximum for x = 0, which is given by the formula:<br />

2<br />

λ EI<br />

Qrhead , = kvy= 2<br />

(7)<br />

DR<br />

o<br />

Compared to the maximum soil reaction stress that occurs<br />

when the pipe is pulled-in entirely in the curved section, a<br />

much higher soil reaction stress is calculated when the head<br />

of the pipeline is located in the curved section:<br />

2<br />

λ EI<br />

Qrhead , = kvy= 2<br />

(8)<br />

DR<br />

o<br />

This, depending on the radius of the borehole, can result<br />

in an unexpectedly higher pulling force than calculated<br />

using the method commonly used in engineering practice.<br />

Finite-element model<br />

description<br />

A recently developed pull-back model was used to investigate<br />

the distribution of the soil-reaction forces in curved sections<br />

of the bore path. The pull-back model uses the ABAQUS<br />

finite-element code, and the MATLAB package was used<br />

for the input for the model and to control the calculations<br />

[7]. The model considers the pipeline and the drill pipes,<br />

which are both modelled using beam elements. For a<br />

certain length of the pipeline and drill pipe, both are<br />

divided into beam elements with a bending and pulling<br />

stiffness.<br />

In the model, the pipeline and drill pipe are initially<br />

straight, and they are bend into the shape of the bore path<br />

in the first simulation step, as shown in Fig.3. The model<br />

simulates the pull-back operation by a series of a prescribed<br />

displacements (Du). To describe the penetration behaviour<br />

of the pipeline and the drill pipes into the borehole wall<br />

during pulling, the interactions between the soil and the<br />

pipeline, and between the soil and the drill pipe, are<br />

described in the model, using spring elements at the end<br />

nodes of a two-noded beam element. For each node on the<br />

beam which undergoes a displacement normal to the bore<br />

path, a user-defined spring stiffness k(u ) describes the<br />

n<br />

interaction with the soil and drilling fluid in the borehole.<br />

The value of k(u n ) can be defined arbitrarily using piecewise<br />

linear segments. In the current model, two segments were<br />

used, a linear spring stiffness for borehole wall penetration<br />

and a linear stiffness in the gap between the pipe (or drill<br />

pipe) and the borehole wall, which represents the drilling<br />

fluid.<br />

By defining the spring function k(u n ) the force normal to<br />

The Journal of Pipeline Engineering<br />

the borehole wall, F n , is calculated. A friction force is then<br />

calculated according to the formula F w = c + mF n in which<br />

F w is the total friction force at a node, F n the normal force,<br />

m the friction coefficient for the pipeline-borehole wall<br />

contact, and c the cohesion, a measure for the friction<br />

between the pipeline and the drilling fluid, which is<br />

independent of the normal force.<br />

Model simulations<br />

Simulations were performed in order to investigate the<br />

behaviour of the head of the pipeline during the pull-back<br />

operation in the upward circular bend of a horizontal<br />

directional drill. To compare with analytical results, the net<br />

weight of the pipeline is set to zero; the pipeline diameter<br />

(D 0 ) was 1.21m, the wall thickness 22.7mm, and the pipe<br />

material was steel with EI=3.1343e 9 Nm 2 . The soil was<br />

chosen to be very soft, with a spring stiffness k = k v D 0 =<br />

130kN/m 2 . Half of the total bore path was modelled as a<br />

straight section of 200m, and a curved section with a<br />

bending radius R = 1000D 0 . The last part of the bore path<br />

beyond the curved section (the bend) is an inclined straight<br />

section to the exit point with a length varying from 0 to<br />

100m.<br />

In the simulations, the pipeline is considered to be located<br />

along the bore path without pulling in order to observe the<br />

soil reaction as the head of the pipeline is moves to different<br />

positions beyond the bend. A first set of simulations was<br />

carried out using no gap between the pipeline and borehole<br />

wall (i.e. no borehole): the resulting soil-reaction stress is<br />

presented in Fig.6. It can be seen that the analytical<br />

solution of the Dutch standard (a moment applied on an<br />

infinite beam on an elastic foundation) gives similar results<br />

to the Abaqus model when the head of the pipeline is<br />

beyond the bend. The analytical solution for a moment<br />

applied on a semi-infinite beam on elastic foundation also<br />

gives similar results to the Abaqus model when the head of<br />

the pipeline is located in the bend. The slight difference<br />

between FEM and analytical solution is due to the fact that<br />

the FEM simulation is geometrically nonlinear and that the<br />

pipeline in the simulations is finite in length.<br />

Sample issue<br />

As the head of the pipeline moves further beyond the end<br />

of the bend into the next straight section, the FEM solution<br />

more and more resembles the analytical solution of the<br />

moment applied on an infinite beam.<br />

From these simulations it can be seen that the maximum<br />

soil-reaction stress as calculated in the Dutch standard<br />

underestimates the soil reaction by a factor of 6.2, as<br />

calculated above, when the head of the pipeline is inside<br />

the bend. Of course the higher soil reaction force leads to<br />

a significant penetration of the pipeline into the borehole<br />

wall, as presented in Fig.7, where the maximum penetration<br />

of the head of the pipeline inside the bend is 126.5mm. The<br />

penetration is 20.8mm when the head of the pipeline is<br />

located beyond the curved section.


3rd Quarter, 2009 211<br />

A second set of simulations were performed with a 100-mm<br />

gap between the pipeline and borehole wall to simulate the<br />

borehole, the results for which are shown in Figs 8 and 9.<br />

The pull-back operation<br />

It is interesting to observe that the borehole-wall penetration<br />

and soil-reaction stress differ by a factor of 6.2, depending<br />

on the location of the head of the pipeline, compared to the<br />

Dutch standard. This results in:<br />

• Primarily:<br />

higher pulling forces due to higher frictional<br />

forces<br />

higher normal forces on the coating of the<br />

pipeline<br />

• Secondarily:<br />

higher pulling forces due to borehole-wall<br />

penetration<br />

higher forces on the pull-back equipment<br />

(connection pipeline and drill pipes)<br />

Firstly, the soil-reaction force leads to a higher normal force<br />

between the pipeline and the borehole wall. By integrating<br />

the Hétenyi formulae [3a] and [3b], the contribution of the<br />

friction to the pulling force can be determined. For the<br />

situation that the head of the pipeline is located in the<br />

curved section, the friction contribution is a factor 1.26<br />

higher than when the head of pipeline is located beyond<br />

the curved section. For the entire pipeline in a borehole<br />

with two bends, this factor for the friction contribution<br />

amounts to 4.26/4 = 1.07.<br />

Secondly, the pulling force is influenced by the effects of<br />

the borehole-wall penetration. The repeated penetration<br />

(every time a drill pipe is disconnected) leads to a ‘bulldozer’<br />

effect during the pull-back operation. In the case of a large<br />

repeated penetration, the pipeline will follow a path below<br />

the created borehole, so that the forces on the connections<br />

between the pull-back equipment and the drill pipes will<br />

considerably increase [5].<br />

Conclusions<br />

A model for the pull-back of pipelines has been created, and<br />

simulations have been performed to study the behaviour of<br />

the pipeline in the borehole during the pull-back operation<br />

in a horizontal directional drilling project. The model can<br />

describe the complex set of interactions between the<br />

pipeline, the drilling pipe, the drilling fluid, and the soil in<br />

the borehole.<br />

From the simulations and analytical solutions, it can be<br />

concluded that the soil-reaction forces are much higher<br />

when the head of the pipeline is located in the bend<br />

compared to when the head of the pipeline has passed<br />

through the bend.<br />

Depending on the ground conditions and the bending<br />

radius, the high soil reaction stress in the curved section<br />

may lead to damage to the pipeline coating, and may lead<br />

to penetration of the borehole wall, which in turn leads to<br />

high pulling forces and may lead to a stuck pipeline or to<br />

damaged pull-back equipment.<br />

References<br />

1. H.M.G.Kruse and H.J.Brink, 2007. Soil related risks during<br />

the pull back operation of horizontal directional drilling. Int.<br />

No-Dig conf., Rome.<br />

2. P.P.T.Litjens and H.J.A.M.Hergarden, 2001. A calculation<br />

method to determine pulling forces in a pipeline during<br />

installation with horizontal directional drilling. Von der<br />

production zur service Schrift (Schriftenreihe aus dem institut<br />

for Rohrleitungsbau Oldenburg).<br />

3a. M.Hetényi, 1946. Beams on elastic foundations. University<br />

of Michigan, equation 6a, page 14.<br />

3b. Idem, equation 20a, page 25.<br />

4. Requirements for pipeline installation, 2003. Dutch Standard<br />

ICS 23.040.10, NEN Delft.<br />

5. H.J.Brink, H.M.G.Kruse, H.Luebbers, H.J.A.M.Hergarden,<br />

and J.Spiekhout. Design guidelines for the bending radius<br />

for large diameter HDD. Journal of Pipeline Engineering, 6, 4.<br />

6. J.P.Pruiksma and H.M.G.Kruse, 2008. Soil pipeline<br />

interaction during the pull back operation of horizontal<br />

directional drilling. Int. No-Dig conf., Moscow.<br />

7. Idem, 2009. Modelling the soil pipeline interaction during<br />

the pull back operation of horizontal directional drilling.<br />

Publication CT-221, Centre for Underground Construction,<br />

Gouda, Netherlands<br />

Sample issue


212<br />

Sample issue<br />

The Journal of Pipeline Engineering


3rd Quarter, 2009 213<br />

Full range stress-strain relation<br />

modelling of pipeline steels<br />

by Stijn Hertelé* 1 , Rudi Denys 2 , and Wim De Waele 2<br />

1 FWO Flanders aspirant, Laboratory Soete, Ghent University, Belgium<br />

2 Laboratory Soete, Ghent University, Belgium<br />

IT IS STANDARD PRACTICE to model the post-yield behaviour of pipeline steels by means of the<br />

Ramberg-Osgood (RO) equation. However, errors can be made when the strain-hardening exponent<br />

or the slope of the stress-strain curve in the post-yield loading range varies with increasing strain. A new<br />

‘UGent’ stress-strain model, outlined in this paper, has been developed to address this problem. It<br />

successfully describes the ‘double-n’ strain hardening seen in contemporary high-strength TMCP pipeline<br />

steels.<br />

THE POST-YIELD BEHAVIOUR of pipeline steels is<br />

often approximated by a constitutive equation which<br />

enables parametric studies in finite-element modelling<br />

applications and provides necessary information to perform<br />

some higher-level variants of ECA analysis. To this purpose,<br />

the Ramberg-Osgood (RO) model [1] is attractive because<br />

of its simplicity. It is usually defined in engineering stress<br />

and strain terms (s and e, respectively) using the 0.2% proof<br />

stress in a form proposed by Hill [2]:<br />

s ⎛ s ⎞<br />

⎟<br />

e = + 0.002<br />

⎜ ⎟<br />

E<br />

⎜ ⎟<br />

R ⎟<br />

n<br />

⎜⎝ p0.2<br />

⎠<br />

In this equation, E is the elasticity modulus, R p0.2 the 0.2%<br />

proof stress, and n a so-called strain-hardening exponent.<br />

The RO equation is almost exclusively applied in pipeline<br />

research, and is recommended in some standards such as<br />

CSA Z662 (App. K) [3] and API 1104 (App. A) [4]. There,<br />

This paper was presented at the Pipeline Technology Conference held<br />

in Ostend, Belgium, on 12-14 October, 2009, and organized by the<br />

University of Gent, Belgium, and Technologisch Instituut vzw, Antwerp,<br />

Belgium.<br />

* Author’s contact details:<br />

email: Stijn.hertele@ugent.be<br />

(1)<br />

it is presented under a slightly modified form, which<br />

enables the use of the more common yield stress at 0.5%<br />

total strain:<br />

s ⎛ s ⎞<br />

e = + ⎜0.005 ⎟ ⎜<br />

⎜ − ⋅⎜ ⎟<br />

E ⎜<br />

⎟<br />

⎝ E ⎠⎟ ⎜ ⎜R<br />

⎟<br />

⎝ ⎟⎠<br />

n<br />

⎛ R ⎞ t0.5<br />

⎟<br />

Sample issue<br />

t0.5<br />

Figure 1 shows the model and its three parameters E, Rt0.5<br />

and n. Note that the power law, which uses n as an<br />

exponent, applies for the full range of the post-yield<br />

behaviour (small-scale yielding as well as extensive yielding).<br />

The strain-hardening exponent n is an important parameter<br />

because it describes the entire post-yield stress-strain<br />

behaviour. Since this exponent is difficult to visually identify,<br />

some relations have been developed that express n as a<br />

function of easily measureable tensile characteristics. Firstly,<br />

if the emphasis is put on an accurate description of the<br />

onset of yielding, n is often estimated using R p0.01 and R p0.2 ,<br />

the 0.01% and 0.2% proof stresses:<br />

⎛ 0.2 ⎞<br />

ln ⎜<br />

⎟<br />

⎜⎝<br />

⎟<br />

0.01⎠⎟<br />

n =<br />

⎛R⎞⎟ ln<br />

⎜ p0.2<br />

⎜ ⎟<br />

⎜ ⎟<br />

⎜⎝R⎟ p0.01<br />

⎠<br />

(2)<br />

(3)


214<br />

Fig.1. The Ramberg-Osgood model as put forward in CSA Z662 and API 1104.<br />

True stress, σ (MPa)<br />

750<br />

730<br />

710<br />

690<br />

670<br />

650<br />

630<br />

610<br />

590<br />

570<br />

550<br />

Sample issue<br />

Experimental curve<br />

Ramberg-Osgood model, using Eq. (3)<br />

Ramberg-Osgood model, using Eq. (4)<br />

The Journal of Pipeline Engineering<br />

0 1 2 3 4 5 6 7 8<br />

True strain, ε (%)<br />

Fig.2. Illustration of the limitations of the Ramberg-Osgood model using Equns 3 and 4.


3rd Quarter, 2009 215<br />

Secondly, if a good global description of post-yield behaviour<br />

is desired, the following rule is frequently applied (R m is the<br />

ultimate tensile stress, uELpl the plastic component of the<br />

uniform elongation, expressed in %):<br />

⎛uEL ⎞ pl<br />

ln ⎜<br />

⎟<br />

⎜⎝ 0.2 ⎟ ⎟⎠<br />

n =<br />

⎛ R ⎞⎟<br />

ln<br />

⎜ ⎟<br />

m<br />

⎜⎜⎝R ⎟ p0.2<br />

⎠<br />

Both Equns 3 and 4 are analytically derived from the<br />

Ramberg-Osgood equation, so their accuracy depends on<br />

the extent to which the Ramberg-Osgood model provides<br />

a good approximation of reality.<br />

Limitations of the<br />

Ramberg-Osgood model<br />

Despite the popularity of the RO model, previous and<br />

ongoing research [5, 6] shows that contemporary highstrength<br />

pipeline steels exhibit a more complex post-yield<br />

behaviour. In particular, modern TMCP pipeline steels<br />

exhibit the so-called ‘double-n’ behaviour, which can be<br />

described by two different characteristic strain-hardening<br />

exponents: the one accounting for small-scale yielding, the<br />

other for extensive yielding. Thus, depending of the<br />

procedure used in establishing n, the single-exponent RO<br />

equation is either capable of well representing the smallscale<br />

yielding stage, or the extensive yielding stage. To<br />

(4)<br />

illustrate the restrictions of Equns 3 and 4 in describing<br />

TMCP pipeline steels, both equations were applied to an<br />

experimental curve of a Grade X70 steel (Fig.2: true stress<br />

and strain). It is clear that both approaches can produce<br />

results that vary from very poor to unacceptable.<br />

Nevertheless, an accurate full-range stress-strain description<br />

may be desired in a range of applications. Important<br />

examples which can be referred to are higher-level FAD<br />

analyses, such as the SINTAP Level 3 FAD [7], and strainbased<br />

design [8-10]. This issue indicates the need for a<br />

better approximation of the post-yield stress-strain relation<br />

of high-strength TMCP pipeline steels.<br />

Sample issue<br />

UGent model<br />

Fig.3. The UGent model and its parameters.<br />

Considering the limitations of the RO equation, a new true<br />

stress-true strain (UGent) model has been developed and is<br />

illustrated, with its parameters, in Fig.3. Note that the<br />

occurrence of a Lüders plateau is not incorporated in the<br />

current study. In essence, the model consists of a<br />

combination of two RO power laws, with a smooth transition<br />

in between. In mathematical terms, the UGent model is<br />

defined by three equations:<br />

⎧ ⎪<br />

RO1(<br />

σ) σ≤σ1 ε= ⎪<br />

⎨RO<br />

→ ( σ) σ < σ≤σ ⎪ ⎪⎩ RO2(<br />

σ) σ2 < σ<br />

1 2 1 2<br />

(5a)<br />

(5b)<br />

(5c)<br />

First, Equn 5a represents the ‘small-scale’ yielding portion<br />

of the stress-strain curve. This portion is modelled by a RO


216<br />

True stress, σ (MPa)<br />

True stress, σσ (MPa)<br />

True stress, σσ (MPa)<br />

700<br />

650<br />

600<br />

550<br />

500<br />

450<br />

400<br />

0 1 2 3 4 5 6 7 8 9 10 11 12 13<br />

650<br />

600<br />

550<br />

500<br />

450<br />

True strain, ε (%)<br />

Test 1<br />

Y/T = 0.73<br />

Ramberg-Osgood model<br />

UGent model<br />

experimental curve<br />

400<br />

0 1 2 3 4 5 6 7 8 9 10 11 12 13<br />

700<br />

650<br />

600<br />

550<br />

500<br />

Ramberg-Osgood model<br />

=<br />

poor representation of<br />

post-yield response<br />

transition stage<br />

small-scale yielding<br />

True strain, ε (%)<br />

Test 2<br />

Y/T = 0.78<br />

Ramberg-Osgood model<br />

UGent model<br />

experimental curve<br />

Sample issue<br />

450<br />

0 1 2 3 4 5 6 7 8 9<br />

True strain, ε (%)<br />

extensive yielding<br />

Test 3<br />

Y/T = 0.82<br />

Ramberg-Osgood model<br />

UGent model<br />

experimental curve<br />

The Journal of Pipeline Engineering<br />

Fig.4a-c (top-bottom). Representative<br />

stress-strain curves from pipeline<br />

TMCP steel and their curve-fits using<br />

the Ramberg-Osgood model and the<br />

UGent model.


3rd Quarter, 2009 217<br />

Fig.4d-f (top-bottom). Representative<br />

stress-strain curves from pipeline<br />

TMCP steel and their curve-fits using<br />

the Ramberg-Osgood model and the<br />

UGent model.<br />

True stress, σ (MPa)<br />

True stress, σ (MPa)<br />

700<br />

650<br />

600<br />

550<br />

500<br />

450<br />

0 1 2 3 4 5 6 7 8 9<br />

950<br />

900<br />

850<br />

800<br />

750<br />

True strain, ε (%)<br />

Test 4<br />

Y/T = 0.86<br />

Ramberg-Osgood model<br />

UGent model<br />

experimental curve<br />

700<br />

0 1 2 3 4 5 6 7<br />

True strain, ε (%)<br />

Test 5<br />

Y/T = 0.87<br />

Sample issue<br />

True stress, σ (MPa)<br />

800<br />

750<br />

700<br />

650<br />

Ramberg-Osgood model<br />

UGent model<br />

experimental curve<br />

600<br />

0 1 2 3 4 5 6 7<br />

True strain, ε (%)<br />

Test 6<br />

Y/T = 0.88<br />

Ramberg-Osgood model<br />

UGent model<br />

experimental curve


218<br />

True stress, σ (MPa)<br />

True stress, σ (MPa)<br />

750<br />

700<br />

650<br />

600<br />

550<br />

700<br />

650<br />

600<br />

550<br />

Test 7<br />

Y/T = 0.90<br />

Ramberg-Osgood model<br />

UGent model<br />

experimental curve<br />

0 1 2 3 4 5 6 7 8<br />

True strain, ε (%)<br />

500<br />

0 1 2 3 4 5 6<br />

True strain, ε (%)<br />

Test 8<br />

Y/T = 0.94<br />

Ramberg-Osgood model<br />

UGent model<br />

power-law relation with a strain hardening exponent n 1 :<br />

n1<br />

σ ⎛ σ ⎞<br />

RO1 ( σ)<br />

= + 0.002 ⎜<br />

⎟<br />

E ⎜⎜⎝σ ⎟<br />

⎠ ⎟<br />

(5a)<br />

0.2<br />

experimental curve<br />

Next, Equn 5b (shown below) models a smooth curve shape<br />

transition.<br />

For the sake of completeness, it is noteworthy that Equn 5b<br />

was obtained from a differential equation with boundary<br />

conditions defined by Equns 5a and 5c, in order to assure<br />

a smooth full-range curve [6].<br />

Finally, Equn 5c represents the RO equation, with strain-<br />

RO<br />

Sample issue<br />

The Journal of Pipeline Engineering<br />

Fig.4g-h (top-bottom). Representative<br />

stress-strain curves from pipeline<br />

TMCP steel and their curve-fits using<br />

the Ramberg-Osgood model and the<br />

UGent model.<br />

hardening exponent n 2 , of the ‘extensive’ yielding portion<br />

of the stress-strain curve (translated over a term –De):<br />

n2<br />

σ ⎛ σ ⎞<br />

RO2 ( σ) = + 0.002 ⎜<br />

⎟ −Δε<br />

E ⎜⎜⎝σ ⎟ ⎟⎠<br />

0.2<br />

(5c)<br />

The strain-translation term De in Equn 5c – note the<br />

minus sign – ensures a continuous curve at s 2 , and is given<br />

by:<br />

2 1 2 1 1 1 1 1<br />

0.002 ⎡ n + n + n + n +<br />

σ2 −σ1 σ2 −σ<br />

⎤<br />

1<br />

Δ ε = ⋅⎢ −<br />

⎥<br />

n2 n1<br />

σ2− σ ⎢<br />

1 ( n2 + 1) ⋅ σ0.2 ( n1+<br />

1)<br />

⋅σ<br />

⎥ (6)<br />

⎢⎣ 0.2 ⎥⎦<br />

⎡ ⎤ ⎡ ⎤<br />

= ⎢ ⎥ ⎢ ⎥<br />

⎣ ⎦ ⎢ ⎥<br />

⎣ ⎦<br />

n<br />

1 n 1 1 1 1<br />

2<br />

n<br />

n<br />

2<br />

+ n<br />

2<br />

+ n<br />

1<br />

+ n<br />

1<br />

+<br />

σ σ<br />

σ−σ 1 σ σ 1 0.002 σ −σ ( ) 0.002<br />

1<br />

σ −σ<br />

1<br />

1→2 σ + ( ) + 0.002 ⋅ − − ⋅ −<br />

E σ σ ( ) ( ) 0.2 2<br />

−σ 1 σ σ<br />

0.2<br />

σ<br />

0.2 2<br />

−σ<br />

1 n<br />

( 1) 2<br />

n<br />

n ( 1)<br />

1<br />

2<br />

+ ⋅ σ n<br />

0.2 1<br />

+ ⋅σ<br />

0.2<br />

(5b)


3rd Quarter, 2009 219<br />

n, n 1, n 2 (-)<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

n (UGent model)<br />

1<br />

n (UGent model)<br />

2<br />

n (Ramberg-Osgood model)<br />

Fig. 4(a)<br />

Fig. 4(b)<br />

Fig. 4(c)<br />

0.70 0.75 0.80 0.85 0.90 0.95 1.00<br />

Y/T (-)<br />

trendline for n 1 (UGent model)<br />

Fig. 4(d)<br />

Fig. 4(e)<br />

Fig. 4(f)<br />

Fig. 4(g)<br />

Fig. 4(h)<br />

increasing tendency<br />

towards<br />

double-n behavior<br />

trendline for n 2 (UGent model)<br />

and n (Ramberg-Osgood model)<br />

Fig.5. Strain-hardening exponents obtained through curve-fitting as a function of Y/T-ratio. ‘Double-n’-behaviour is more<br />

pronounced for higher Y/T-ratios.<br />

Table 1. Tensile<br />

characteristics and model<br />

parameters for the eight<br />

characteristic tensile tests.<br />

Test<br />

no.<br />

Tensile<br />

charac<br />

teristi<br />

cs<br />

Y/<br />

T<br />

Rp 0.<br />

2<br />

( = Y)<br />

MPaksi The model has six independent parameters: E, s 0.2 , n 1 , n 2 ,<br />

s 1 , and s 2 . Of these, E and s 0.2 represent the model’s<br />

elasticity modulus and 0.2% proof stress, respectively. The<br />

strain-hardening exponents n 1 and n 2 characterize the<br />

‘double-n’ behaviour. Note that:<br />

• n 1 , which describes the small-scale yielding area, is of<br />

Rambe<br />

rg-<br />

Osgood<br />

σ / 0 . 2<br />

Rp0. 2<br />

n<br />

Curv<br />

e-fitted<br />

model<br />

parame<br />

ters<br />

σ / 0 . 2<br />

Rp0. 2<br />

UGent<br />

model<br />

n n σ /σ 1 2 1 0.<br />

2<br />

σ /σ 2 0.<br />

2<br />

1 0. 73<br />

432 62. 3 0. 983<br />

9. 0 1. 011<br />

10. 5 9. 5 1. 19<br />

1.<br />

37<br />

Sample issue<br />

2 0. 78<br />

424 61. 5 0. 963<br />

10. 2 1. 013<br />

13. 6 11. 0 1. 13<br />

1.<br />

35<br />

3 0. 82<br />

515 74. 7 0. 967<br />

12. 0 1. 008<br />

19. 1 13. 3 1. 08<br />

1.<br />

15<br />

4 0. 86<br />

521 75. 6 0. 958<br />

13. 4 0. 999<br />

23. 5 15. 1 1. 04<br />

1.<br />

14<br />

5 0. 87<br />

736 106. 7 0. 988<br />

15. 7 1. 003<br />

24. 9 16. 5 0. 90<br />

1.<br />

13<br />

6 0. 88<br />

649 94. 1 1. 004<br />

17. 8 1. 018<br />

20. 6 18. 3 1. 11<br />

1.<br />

15<br />

7 0. 90<br />

605 87. 7 0. 960<br />

16. 3 0. 995<br />

30. 0 18. 3 1. 02<br />

1.<br />

12<br />

8 0. 94<br />

584 84. 7 0. 975<br />

24. 1 1. 006<br />

52. 1 27. 1 1. 03<br />

1.<br />

06<br />

particular importance for flaw integrity. Indeed, the<br />

crack driving force, for instance expressed in terms<br />

of CMOD, is highly influenced by the initial strainhardening<br />

behaviour [11].<br />

• n 2 , which describes the extended yielding area,<br />

might be related to uniform elongation capacity.


220<br />

n, n 1, n 2 (-)<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

n (UGent model)<br />

This statement follows from Equn 5c and the onset<br />

of necking criterion for true stress and strain,<br />

ds/de = s. Further validation is, however, needed<br />

to confirm and identify this supposed relation.<br />

The stress values s 1 and s 2 define the intervals of application<br />

of the different sub-equations. Generally, they define the<br />

end of small-scale yielding and the initiation of extensive<br />

yielding.<br />

Experimental validation<br />

The UGent model has been validated using 146 stressstrain<br />

curves of Grade X60 to X100 pipeline steels. Defined<br />

as the ratio between 0.2% proof stress and ultimate tensile<br />

stress, the yield-to-tensile ratio (Y/T) varied from 0.68 to<br />

0.94. This paper elaborates eight representative results,<br />

which are summarized in Table 1. The findings from this<br />

selection are confirmed in a more in-depth validation on<br />

the entire set of 146 curves, which will be reported in the<br />

near future.<br />

The experimental stress-strain curves were transformed to<br />

true stress – true strain up to the point of necking, using the<br />

following well-known conversion relations:<br />

ε = ln( 1+ e)<br />

(7)<br />

σ = s⋅ ( 1+<br />

e)<br />

(8)<br />

For the ease of data manipulation, all curves were next<br />

reduced to a set of 100 points: 80 of these points were taken<br />

1<br />

n 2(UGent model)<br />

no trend observed<br />

Fig. 4(b)<br />

Fig. 4(a)<br />

n (Ramberg-Osgood model)<br />

Fig. 4(c)<br />

Fig. 4(d)<br />

Fig. 4(h)<br />

The Journal of Pipeline Engineering<br />

400 450 500 550 600<br />

Rp0.2 (MPa)<br />

650 700 750 800<br />

Fig.6. Strain-hardening exponents obtained through curve-fitting as a function of R p0.2 . No trend is visible.<br />

Fig. 4(g)<br />

Fig. 4(f)<br />

Fig. 4(e)<br />

in the strain interval [0% – 1%], since that area shows the<br />

elastic-to-plastic transition, a zone which is generally of<br />

particular importance in plastic analysis. The resulting data<br />

sets were least-squares’ curve-fitted with the Ramberg-<br />

Osgood model and the UGent model, using the Levenberg-<br />

Marquardt algorithm [12].<br />

Figure 4 shows the eight resulting experimental curves,<br />

along with their RO and UGent curve fits. The parameters<br />

obtained are given in a dimensionless form in Table 1. It<br />

can be seen that the UGent model produces an almostperfect<br />

approximation of the experimental curves, whereas<br />

the RO model shows considerable errors for some cases<br />

(tests 3, 4, 7, and 8). The deviation is especially significant<br />

at small-scale yielding. To illustrate the different stages of<br />

strain hardening, some additional explanation is given on<br />

a clear example in Fig.4c (test 3). As discussed, the poor<br />

performance of the RO-model can be attributed to the<br />

‘double-n’ behaviour of the steel.<br />

Sample issue<br />

Figures 5 and 6 depict the variation of the strain-hardening<br />

exponents n 1 and n 2 (UGent model) and n (RO model) as<br />

a function of the Y/T-ratio and Rp0.2, respectively. Two<br />

important observations can be made:<br />

• the difference between n 1 and n 2 increases with Y/<br />

T ratio (Fig.5). In other words, pipeline TMCP<br />

steels with a high Y/T-ratio display ‘double-n’<br />

behaviour. A similar trend could not be observed<br />

using R p0.2 (Fig.6).<br />

• the exponent n (RO-model) correlates with n 2<br />

(UGent model). Thus, in the case where there is a


3rd Quarter, 2009 221<br />

‘double-n’ behaviour, the RO curve-fit merely<br />

describes the extensive yielding stage. In contrast,<br />

the UGent model manages to describe the smallscale<br />

yielding stage as well, through n 1 . It must be<br />

mentioned that, for n 1 to be a representative<br />

parameter for small-scale yielding, SIGMA 1 must<br />

exceed SIGMA 0.2 . This was not the case in test 5.<br />

Summary and conclusion<br />

It is standard practice to model the post-yield behaviour of<br />

pipeline steels by means of the Ramberg-Osgood (RO)<br />

equation. However, errors can be made when the strainhardening<br />

exponent, or the slope of the stress-strain curve<br />

in the post-yield loading range, varies with increasing<br />

strain. The ‘UGent’ stress-strain model outlined in this<br />

paper describes the ‘double-n’ strain hardening seen in<br />

contemporary high-strength pipeline TMCP steels. It may<br />

therefore contribute towards a better numerical description<br />

of mechanical pipeline-related problems.<br />

Acknowledgments<br />

The authors would like to acknowledge the FWO (Fund for<br />

Scientific Research), Flanders, for its financial support.<br />

References<br />

1. W.Ramberg and W.R.Osgood, 1943. Description of stressstrain<br />

curves by three parameters. National Advisory<br />

Committee for Aeronautics, Technical Note 902.<br />

2. H.N.Hill, 1944. Determination of stress-strain relations from<br />

the offset yield strength values. National Advisory Committee<br />

for Aeronautics, Technical Note 927.<br />

3. CSA Z662, 2007. Oil and gas pipeline systems.<br />

4. API 1104, 2007. Welding of pipelines and related facilities.<br />

5. R.Denys, P.De Baets, A.Lefevre, and W.De Waele, 2002.<br />

Material tensile properties in relation to the failure behaviour<br />

of girth welds subject to plastic longitudinal strains. Proc.<br />

Conf. Application & Evaluation of High-Grade Linepipes in<br />

Hostile Environments, Yokohama, Japan, November 7-8,<br />

pp159-172.<br />

6. S.Hertelé, W.De Waele, and R.Denys. To be published.<br />

7. S.Webster and A.Bannister, 2000. Structural integrity<br />

assessment procedure for Europe - of the SINTAP programme<br />

overview. Engineering Fracture Mechanics, 67, 6, pp481-514.<br />

8. W.De Waele, 2004. Effect of material properties on the<br />

plastic straining capacity of defective welds. PRICM 5: the 5 th<br />

Pacific Rim <strong>International</strong> Conference on Advanced Materials<br />

and Processing, Beijing, China, November 2-5, pp2659-<br />

2662.<br />

9. R. Denys, P.De Baets, and W.De Waele, 2003. Weld metal<br />

test performance requirements - a critical appraisal of future<br />

needs. Thermec’ 2003, Pts 1-5, Vols 426-430, pp4153-4158.<br />

10. R. Denys, 2007. Interaction between material properties,<br />

inspection accuracy and defect acceptance levels in strain<br />

based pipeline design. Proc. NATO Advanced Research<br />

Workshop on Safety, Reliability and Risks Associated with<br />

Water, Oil and Gas <strong>Pipelines</strong>, Alexandria, Egypt, February 4-<br />

8, pp45-64.<br />

11. R. Denys, 2008. Weld metal strength mismatch: past, present<br />

and future. Proc. <strong>International</strong> Symposium to Celebrate<br />

Prof. Masao Toyoda’s Retirement from Osaka University,<br />

Osaka, Japan, pp115-148.<br />

12. D.W.Marquardt, 1963. An algorithm for least-squares<br />

estimation of nonlinear parameters. J. Soc. for Industrial and<br />

Applied Mathematics, 11, 2, pp431-441.<br />

Sample issue


222<br />

Sample issue<br />

The Journal of Pipeline Engineering


3rd Quarter, 2009 223<br />

Correction<br />

A practical approach to pipeline corrosion modelling: Part 2 – Shortterm<br />

integrity forecasting<br />

cv<br />

R<br />

R<br />

by Dr Érika S M Nicoletti, Ricardo Dias de Souza,<br />

and Dr Sérgio da Cunha Barros<br />

WE REGRET that some of the equations and figures published in this paper in our 2nd Quarter issue,<br />

2009 (Vol 8 No 2) were reproduced incorrectly. Readers are asked to note the corrected versions<br />

publsihed below, and are advised that a fully-corrected version of the paper, and of the issue itelf, are<br />

available on the Journal website www.j-pipe-eng,com.<br />

It goes without saying that we are very sorry for the problems that this may cause, and apologize both to<br />

the authors and to readers.<br />

The following are the correct verions of the equations and figures involved:<br />

σ<br />

R<br />

r = (2)<br />

Li<br />

Li<br />

j= n+ i<br />

∑<br />

dj<br />

j=− i n<br />

=<br />

(2n+ 1). j= nΔ+ its<br />

∑ dj<br />

j=− i n<br />

=<br />

(2n+ 1). Δt −Δt<br />

s c<br />

(3aa)<br />

(3ab)<br />

σ = R . cv<br />

(3b)<br />

Li<br />

Li<br />

Fig.2. Logic flowchart for metal-loss<br />

growth under general hot-spot<br />

conditions.<br />

>N<br />

d = d + R . Δ t<br />

(4a)<br />

fi i Li f<br />

Pif = f( df, li, wi)<br />

(5)<br />

R<br />

rc<br />

d −d<br />

=<br />

Δt<br />

2 1<br />

i<br />

(8a)<br />

σ = R . cv<br />

(8b)<br />

Sample issue<br />

rc<br />

d<br />

σ<br />

Li<br />

Li<br />

=<br />

=<br />

rc<br />

dINSP<br />

d i>=perc 0.8dLi. Y dLi = di<br />

Loop i<br />

∑ + = j n i<br />

j=<br />

i−n<br />

d Li<br />

N<br />

d j<br />

2n<br />

+ 1<br />

. cv


224<br />

>N1<br />

INSP1B<br />

d1B<br />

f(x)<br />

0,25<br />

0,20<br />

0,15<br />

0,10<br />

0,05<br />

0,00<br />

INSP1<br />

Segmetation<br />

Criterion<br />

Loop j<br />

N2<br />


Sample issue


Sample issue

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!