September, 2013 Vol.12, No.3 Sa m pl e co py -n ot f or d is t rib ut incorporating The Journal of Pipeline Integrity io n Journal of Pipeline Engineering Great Southern Press Clarion Technical Publishers Journal of Pipeline Engineering Editorial Board - 2013 Sa m pl e co py -n ot f or d is t rib ut io n Dr Husain Al-Muslim, Pipeline Engineer, Consulting Services Department, Saudi Aramco, Dhahran, Saudi Arabia Mohd Nazmi Ali Napiah, Pipeline Engineer, Petronas Gas, Segamat, Malaysia Dr-Ing Michael Beller, Rosen Engineering, Karlsruhe, Germany Jorge Bonnetto, Operations Director TGS (retired), TGS, Buenos Aires, Argentina Dr Andrew Cosham, Atkins, Newcastle upon Tyne, UK Dr Sreekanta Das, Associate Professor, Department of Civil and Environmental Engineering, University of Windsor, ON, Canada Leigh Fletcher, Welding and Pipeline Integrity, Bright, Australia Daniel Hamburger, Pipeline Maintenance Manager, Kinder Morgan, Birmingham, AL, USA Dr Stijn Hertele, Universiteit Gent – Laboratory Soete, Gent, Belgium Prof. Phil Hopkins, Executive Director, Penspen Ltd, Newcastle upon Tyne, UK Michael Istre, Chief Engineer, Project Consulting Services, Houston, TX, USA Dr Shawn Kenny, Memorial University of Newfoundland – Faculty of Engineering and Applied Science, St John’s, Canada Dr Gerhard Knauf, Salzgitter Mannesmann Forschung GmbH, Duisburg, Germany Prof. Andrew Palmer, Dept of Civil Engineering – National University of Singapore, Singapore Prof. Dimitri Pavlou, Professor of Mechanical Engineering, Technological Institute of Halkida , Halkida, Greece Dr Julia Race, School of Marine Sciences – University of Newcastle, Newcastle upon Tyne, UK Dr John Smart, John Smart & Associates, Houston, TX, USA Jan Spiekhout, DNV Kema, Groningen, Netherlands Prof. Sviatoslav Timashev, Russian Academy of Sciences – Science & Engineering Centre, Ekaterinburg, Russia Patrick Vieth, President, Dynamic Risk, The Woodlands, TX, USA Dr Joe Zhou, Technology Leader, TransCanada PipeLines Ltd, Calgary, Canada Dr Xian-Kui Zhu, Principal Engineer, Edison Welding Institute, Columbus, OH, USA ��� 3rd Quarter, 2013 141 The Journal of Pipeline Engineering incorporating io n The Journal of Pipeline Integrity Volume 12, No 3 • Third Quarter, 2013 rib ut Contents ❖❖❖ Sa m pl e co py -n ot f or d is t Guest Editorial..............................................................................................................................................143 A special issue of Journal of Pipeline Engineering on fracture-toughness testing, evaluation, and application for pipeline steels Dr Xian-Kui Zhu...........................................................................................................................................145 Fracture-toughness (K, J) testing, evaluation, and standardization Dr William R Tyson, Dr Guowu Shen, Dr Dong-Yeob Park, and James Gianetto...........................................157 Low-constraint toughness testing Dr Su Xu, Dr William R Tyson, and Dr C H M Simha..................................................................................165 Testing for resistance to fast ductile fracture: measurement of CTOA Dr Robert Eiber............................................................................................................................................175 Drop-weight tear test application to natural gas pipeline fracture control Dr Brian N Leis ............................................................................................................................................183 The Charpy impact test and its applications Dr Robert L Amaro, Dr Jeffrey W Sowards, Elizabeth S Drexler, J David McColskey, and Christopher McCowan............................................................................................................................199 CTOA testing of pipeline steels using MDCB specimens Prof. Claudio Ruggieri and Leonardo L S Mathias......................................................................................... 217 Fracture-resistance testing of pipeline girth welds using bend and tensile fracture specimens Dr He Li, Qiang Chi, Jiming Zhang, Yang Li, Lingkang Ji, and Chunyong Huo..............................................229 Fracture-toughness evaluations by different test methods for the Chinese Second West-East gas transmission X-80 pipeline steels Dr Philippa Moore and Dr Henryk Pisarski...................................................................................................237 CTOD and pipelines: the past, present, and future Dr Rudi M Denys, Dr Stijn Hertelé, and Dr Antoon A Lefevre......................................................................245 Use of curved-wide-plate (CWP) data for the prediction of girth-weld integrity Dr Xian-Kui Zhu and Dr Brian N Leis...........................................................................................................259 Ductile-fracture arrest methods for gas-transmission pipelines using Charpy impact energy or DWTT energy OUR COVER PHOTO shows a single-edge-notched specimen in tension (SENT, or equivalently SE(T)) under load in CANMET’s laboratory. For further details, see the paper by William Tyson et al. on pages 157-163. The photo is reproduced with kind permission of CANMET; it was originally published in one of the research organization’s recent technical reports. The Journal of Pipeline Engineering has been accepted by the Scopus Content Selection & Advisory Board (CSAB) to be part of the SciVerse Scopus database and index. 142 The Journal of Pipeline Engineering T HE Journal of Pipeline Engineering (incorporating the Journal of Pipeline Integrity) is an independent, international, quarterly journal, devoted to the subject of promoting the science of pipeline engineering – and maintaining and improving pipeline integrity – for oil, gas, and products pipelines. The editorial content is original papers on all aspects of the subject. Papers sent to the Journal should not be submitted elsewhere while under editorial consideration. rib ut io n Authors wishing to submit papers should do so online at www.j-pipeng.com. The Journal of Pipeline Engineering now uses the Aires Editorial Manager manuscript management system for accepting and processing manuscripts, peer-reviewing, and informing authors of comments and manuscript acceptance. Please follow the link shown on the Journal’s site to submit your paper into this system: the necessary instructions can be found on the User Tutorials page where there is an Author's Quick Start Guide. Manuscript files can be uploaded in text or PDF format, with graphics either embedded or separate. Please contact the editor (see below) if you require any assistance. The Journal of Pipeline Engineering aims to publish papers of quality within six months of manuscript acceptance. 4. Back issues: Single issues from current and past volumes are available for US$87.50 per copy. or d 1. Disclaimer: While every effort is made to check the accuracy of the contributions published in The Journal of Pipeline Engineering, Great Southern Press Ltd and Clarion Technical Publishers do not accept responsibility for the views expressed which, although made in good faith, are those of the authors alone. is t Notes ot f 5. Publisher: The Journal of Pipeline Engineering is published by Great Southern Press Ltd (UK and Australia) and Clarion Technical Publishers (USA): Great Southern Press, PO Box 21, Beaconsfield HP9 1NS, UK • tel: +44 (0)1494 675139 • fax: +44 (0)1494 670155 • email:[email protected] • web:www.j-pipe-eng.com • www.pipelinesinternational.com py -n 2. Copyright and photocopying: © 2013 Great Southern Press Ltd and Clarion Technical Publishers. All rights reserved. No part of this publication may be reproduced, stored or transmitted in any form or by any means without the prior permission in writing from the copyright holder. Authorization to photocopy items for internal and personal use is granted by the copyright holder for libraries and other users registered with their local reproduction rights organization. This consent does not extend to other kinds of copying such as copying for general distribution, for advertising and promotional purposes, for creating new collective works, or for resale. Special requests should be addressed to Great Southern Press Ltd, PO Box 21, Beaconsfield HP9 1NS, UK, or to the editor. co Editor: John Tiratsoo • email: [email protected] pl e Clarion Technical Publishers, 3401 Louisiana, Suite 110, Houston TX 77002, USA • tel: +1 713 521 5929 • fax: +1 713 521 9255 • web: www.clarion.org Associate publisher: BJ Lowe • email:[email protected] Sa m 3. Information for subscribers: The Journal of Pipeline Engineering (incorporating the Journal of Pipeline Integrity) is published four times each year. The subscription price for 2013 is US$350 per year (inc. airmail postage). Members of the Professional Institute of Pipeline Engineers can subscribe for the special rate of US$175/year (inc. airmail postage). Subscribers receive free on-line access to all issues of the Journal during the period of their subscription. v 6. ISSN 1753 2116 v v www.j-pipe-eng.com is available for subscribers 3rd Quarter, 2013 143 Guest Editorial A special issue of Journal of Pipeline Engineering on fracture-toughness testing, evaluation, and application for pipeline steels T Guest Editor delivers a brief review of historical efforts as well as recent advances in the development of the critical K and J testing, resistance-curve testing, experimental estimation, and standardization by the American Society for Testing and Materials (ASTM). The ASTM standard specimens include the three-point bend and compact-tension specimens. More than ten experts in this area were invited to write papers for this special issue, including selected organization representatives, influential scientists, engineers, and academics who have made significant contributions to the development of fracture-toughness test methods and are recognized internationally in the area of fracture-toughness testing and evaluation, and its application to pipeline steels. These authors come from different countries and regions in the world, including North America (USA and Canada), Europe (UK and Belgium), South American (Brazil), and Asia (China). Eleven papers were selected to cover fracture-toughness test methods, test procedures, experimental techniques, experimental evaluations, and standard developments, as well as their applications to transmission pipelines for the six fracture-toughness parameters of CVN, DWTT, K, J, CTOD, and CTOA. In order to better characterize dynamic ductile fracture toughness for pipeline steels, the CTOA was proposed as an alternative fracture parameter, and different CTOA test methods have been developed with use of different laboratory specimens. In the third paper, Dr Su Xu et al. describe the CTOA test procedures and ‘round-robin’ results using the DWTT specimens and a simplified single-specimen method developed at Canmet Materials, with a discussion on the application to determine a critical CTOA with typical results and step-by-step procedures. The sixth paper, by Dr Robert Amaro et al., summarizes the CTOA testing of pipeline steels at quasi-static and dynamic rates using modified double-cantilever-beam (MDCB) specimens that have been done at NIST at its Boulder, Colorado, facility between 2006 and 2012. rib ut io n HIS SPECIAL ISSUE of the Journal of Pipeline Engineering is dedicated to the topic of fracturetoughness testing, evaluation, and application for pipeline steels. It is well known that fracture toughness is the most important material property required by fracturemechanics’ methods. For pipeline steels, the commonlyused fracture-toughness parameters are Charpy-V notch (CVN) energy, drop-weight tear test (DWTT) energy, stress-intensity factor K, J-integral, crack-tip-opening displacement (CTOD), and crack-tip-opening angle (CTOA). These parameters have been extensively used in the oil and gas pipeline industry for engineering design and structural-integrity assessment, including material selection; material-performance evaluation; defect assessment; fatigue-life estimation; crack, leak, or rupture determination; engineering-critical analysis; fitness-for-service analysis; fracture-initiation control; and fracture-propagation control. Thus, fracturetoughness testing and evaluation are critical to pipeline steel manufacture and to structural-integrity assessment and management. Sa m pl e co py -n ot f or d is t Fracture toughness is known to depend on the crack-tip constraint due to geometry and loading configurations. To provide a more meaningful measure of toughness, ‘low-constraint’ tests are developing using single-edge notched tensile (SENT) specimens. In the second paper, Dr William Tyson et al. describe the development of SENT tests in terms of J and CTOD for determining crack-growth-resistance curves that can be used to assess the tolerance of weld flaws to tensile loads. The seventh paper by Professor Claudio Ruggieri presents the J-integral resistance-curve testing for pipeline girth welds using the conventional bending and SENT specimens. Among the six fracture parameters, the K and J are true fracture parameters and have special significance. The K factor, proposed in 1957 to describe the elastic cracktip field, symbolized the birth of linear-elastic fracture mechanics, and the J concept, proposed in 1968 to characterize the elastic-plastic crack-tip field, symbolized the birth of elastic-plastic fracture mechanics. Over a subsequent half century, numerous efforts have been made to develop valid fracture-toughness test methods and standards. In the first paper, the undersigned The DWTT energy is an apparent toughness parameter that has been used in the pipeline industry since the 1960s. In the fourth paper, Robert Eiber reviews the need, development, and application of DWTT energy for controlling fracture propagation in natural gas transmission pipelines. He summarizes the incidents that started the research leading to the development of the DWTT from 1960 to present. The initial goal of the DWTT was to define the ductile-to-brittle transition temperature of pipeline steels to facilitate the specification of transition temperature below the 144 The Journal of Pipeline Engineering io n rib ut It is expected that this special issue will serve as a technical document for tracking the historical efforts and developments of fracture-toughness testing, evaluation, and application to pipeline steels, for understanding/ using appropriate fracture-toughness parameters as well as the corresponding test methods, and also for further improving the fracture-toughness test standards in the future. As such, it will provide a useful technical source for researchers and engineers in the oil and gas pipeline industry. ot f The eighth paper, by Dr He Li et al., presents fracturetoughness evaluations using the CVN and DWTT energies for the X80 pipeline steels used for the Chinese second west-east gas transmission pipeline. Based on a variety of test data, these authors compare the 2-mm striker and 8-mm striker CVN energies over a range of temperatures, and the DWTT energy with the 8-mm striker CVN energy at the room temperature, and thus obtain useful relationships between these toughness parameters. The last paper, by the Guest Editor and Dr Brian Leis, discusses the applications of CVN impact energy and DWTT energy on ductile-fracture-propagation control, and reviews the existing ductile-fracture arrest methods in terms of CVN and DWTT toughness parameters for predicting the arrest-fracture toughness of gas transmission pipelines, including modern high-strength pipeline steels. is t In the 1950s, the CVN impact energy was the first apparent fracture-toughness parameter to be used to characterize toughness of linepipe steels. In the fifth paper, Dr Brian Leis reviews the CVN test and standard development, and assesses its use to characterize fracture resistance in applications from vintage to modern toughness pipeline steels. He concludes that where tough materials are involved, alternative testing practices are needed that are better adapted to the specific loading and failure response of the structure of interest. In addition to the fracture-toughness tests, Professor Rudi Denys developed a curved-wide-plate (CWP) specimen for conducting a quasi-structure test to provide a rational basis for predicting girth-weld integrity for both stress- and strain-based designs, for establishing material requirements, and for validating numerical models or fracture-mechanics’-based defect assessments. These are described in the tenth paper. or d operating temperature range for linepipe. Today, the need for a measure of the steel toughness has emerged to control ductile-fracture-propagation arrest, leading to examination of the DWTT energy as a substitute for the CVN energy. Sa m pl e co py -n The CTOD is an engineering-fracture parameter that was proposed in 1963, and this parameter has been widely used for structural integrity analysis in the oil and gas industry ever since. In the ninth paper, Drs Philippa Moor and Henryk Pisarski present a review of CTOD testing and its application to pipelines in the past, present, and future, and describe the development of standardized fitness-for-service assessment procedures from the use of the CTOD design curve to the failure-analysis diagram approach in the CTOD British Standard. Thanks are given to the editor who made this special issue possible. He provided helpful advice in paper solicitation, organization, review, and final submission. His exceptional support made this special issue a recordbreaker in the history of this Journal. Dr Xian-Kui Zhu Fellow of ASME Principal Engineer, Structural Integrity and Modelling Edison Welding Institute Columbus, OH 43221, USA Email: [email protected] 3rd Quarter, 2013 145 Fracture-toughness (K, J) testing, evaluation, and standardization by Dr Xian-Kui Zhu io n Battelle Memorial Institute, Columbus, OH, USA T ot f or d is t rib ut HE STRESS-INTENSITY factor K and the J-integral are the two most important parameters in fracture mechanics, serving as the material properties to quantify the toughness or resistance of materials against fracture in linear elastic and elastic-plastic conditions, respectively.The fracture-toughness characterized by K and J has been widely used as the essential material properties in fracture-mechanics’ design and structural-integrity assessment for pipelines and other structures. Experimental testing and evaluation has played a central role in providing reliable fracture-toughness for fracture-mechanics’ analysis of structures containing cracks. Since the K-factor and the J-integral concepts were proposed, numerous investigations have been made to develop valid experimental test methods, test techniques, evaluation procedures, and test-method standardization, as evident in ASTM E399 and ASTM E1820 – two commonly used fracture-toughness test standards. In recent years, important improvements for the KIc testing and significant progresses for the J-integral testing have been achieved. To better understand these two fracture-toughness parameters and to properly use the associated test standards, the present paper delivers a brief review of historical efforts as well as recent advances in the development of the K-factor and the J-integral experimental estimation and standard testing. F Fracture-toughness measurement and evaluation is required by the fracture-mechanics’ methods for quantifying material toughness and for predicting crack failure in structural design, fitness-for-service analysis, material-performance evaluation, and structural-integrity assessment for a variety of engineering components or structures, such as nuclear pressure vessels and piping, chemical plant vessels, petroleum storage tanks, and oil and gas pipelines. Fracture-toughness testing, experimental estimation, test procedure, and data evaluation all are essential to experimental fracture mechanics, and have been investigated extensively for more than a half of century by numerous scientists and researchers around the world. Sa m pl e co py -n RACTURE TOUGHNESS IS the capacity of a material containing a crack to absorb applied energy or to resist fracture, and is one of the most important properties of any material for many design applications. Fracture resistance is a measure of fracture toughness to describe the increasing resistance of the material against crack growth. Several fracture parameters have been proposed to characterize the fracture toughness or resistance of metallic materials. The stress-intensity factor K and the J-integral are the two most often used fracture-toughness parameters [1], where the K-factor was proposed in 1957 by Irwin [2] for describing the intensity of an elastic crack-tip field, and the J-integral was proposed by Rice [3] in 1968 for describing the intensity of an elastic-plastic crack-tip field. Now, the K-factor is the symbol of the linear-elastic fracture mechanics, whereas the J-integral is the symbol of the elastic-plastic fracture mechanics. All fracture-mechanics’ methods require fracture toughness for crack assessments, and thus fracture-toughness testing and evaluation are very important. Typically, KIc and JIc represent fracture toughness at the onset of a mode-I crack in plane-strain conditions, while K-R and J-R curves describe the material resistance against crack growth during loading, respectively, for elastic-dominated cracks and for ductile tearing cracks. Author’s contact details: Note: Since writing this paper, the author has moved to EWI in Columbus, OH, USA, where his contact information is: tel: +1 614 688 5135 email: [email protected] Over the years, a large number of experimental investigations have been performed to measure reliable fracture toughness, including development of experimental methods, test techniques, test devices and specimens, test procedures, data instrumentation, estimation methods, experimental evaluation, and testmethod standardization. In recent years, improvements in K testing and significant progress for the J-integral testing have been achieved, as evident in the fracturetoughness test standards ASTM E399 [4] and ASTM E1820 [5] that are updated periodically. ASTM E399 is a standard test method for linear-elastic plane-strain fracture-toughness KIc testing, and ASTM E1820 is a standard test method for the measurement of combined fracture-toughness parameters including J and K. These 146 The Journal of Pipeline Engineering Fracture toughness KIc (GIc) testing -n io n ot f In the fracture-mechanics’ community, the work of Griffith [6] in 1920 has been regarded as the start of the fracture-mechanics’ method, whereas actually the work of Irwin [2] and Williams [7] in 1957 led to the development of linear-elastic fracture mechanics’ theory based on the crack-tip field analysis. Griffith introduced the concept of the elastic energy-release rate G, while Irwin proposed the concept of the stress-intensity factor K. For an elastic crack in model-I loading, Irwin [2] established a simple relationship between the stressintensity factor KI and the elastic-energy release rate G as: In 1958, ASTM established a technical committee E24 on fracture testing of metals (now E08 on fatigue and fracture mechanics) to study the fracture-toughness problem and to develop a standard fracture-test method for metallic materials. Srawley and Brown [11], the two pioneers who made eminent contributions to fracture-toughness testing, drafted the first standard method for the KIc testing. The draft test method included the details of specimen design, specimen size requirement, fatigue pre-cracking, test device, fixture design, crack-opening-displacement gauge, data instrumentation, load transducer, test procedure, and the K calibration and calculation method. This draft method became the first widely recognized fracturetoughness standard ASTM E399-70T for the KIc testing, and ASTM E399 became the model for all subsequent standard test methods of fracture-toughness measurement in ASTM and other standard organizations in the world. E399 has undergone many revisions and updates over the years, and the current version is ASTM E399-09e2 [4]. In this standard, a conditional fracture toughness KQ is defined as: rib ut K (G) testing The lower-bound value KIc (or GIc) is called ‘plane-strain fracture toughness’ for the mode-I cracks, denoting the minimum fracture resistance of the material. is t To better understand the two most important fracturetoughness parameters and their test standards, the present paper delivers an overview of historical efforts and recent advances in the development of the K-factor and the J-integral experimental test methods and standards in ASTM. This includes the review of K (G) testing, early J experimental estimation and testing, advances of J experimental estimation and evaluation, and development of the K-factor and the J-integral based fracture-test standardization. Thus, the thickness dependence of the critical Kc is similar to that for the critical Gc. or d standards were developed by ASTM (American Society for Testing and Materials), and have been used worldwide for measuring the critical value of K and J at the onset of fracture, and resistance curves (R-curves) during crack growth. is for the plane stress conditions, and is for the plane-strain conditions. co where E’ = E E’ = E (1 – ν2) py K I2 / E ' = G(1) Sa m pl e In order to predict fracture failure, Irwin and Kies [8] first introduced the concept of fracture resistance and used it as a material property based on the energy approach. Krafft et al. [9] obtained the earliest R-curves using a set of centre-cracked thin plates for 7075-T6 aluminium alloy in terms of the elastic energy release rate G and the absolute crack extension Δa. Fracture instability was determined at a critical value of Gc, where a drive force intercepts or tangents the resistance curve. Irwin et al. [10] showed that the Gc depends on the initial crack size and specimen thickness, and observed that the Gc decreases with increasing thickness, and reaches a lower bound value denoted by GIc for a large thickness where the plane-strain conditions prevails. In the 1960s, R-curves were first developed in terms of the energy-release rate G; later, the stress-intensity factor K was often used in lieu of G. Basically, a K-R curve and a G-R curve are equivalent to each other for an elastic material due to their relationship in Eqn 1. KQ = PQ B W f (a / W ) (2) where PQ is a measured critical load, and f(a/W) is a geometry function of a/W with a the crack size and W the specimen width. Different fracture specimens have been developed, but this paper is central to the commonly-used compact tension (CT) and single-edgenotched bend (SENB) specimens. In order to validate the conditional toughness KQ as a fracture toughness KIc result, the initial crack size of 0.45 ≤ a0/W ≤ 0.55 and the following two validity requirements must be met: B, b ≥ 2.5 ( K Q / σ ys ) 2 (3a) Pmax ≤ 1.1PQ (3b) where b = W-a is the specimen ligament, Pmax is the maximum load, and σys is the yield stress of tested material. In general, the size requirements in Eqn 3a make it difficult to measure KIc, because the material must be either relatively brittle or very large with a specimen width exceeding 1 m for the 95% secant offset procedure in ASTM E399 to accurately estimate the load PQ at the crack initiation. The second requirement, in Eqn 3b, is an attempt to assure that the nonlinearity 3rd Quarter, 2013 147 io n Applications showed that the fracture toughness KIc is good to use only for cracks in elastic conditions, and thus a new facture-toughness parameter was sought from the 1960s to the 1970s to characterize fracture toughness of a crack in elastic-plastic conditions. Such a fracture parameter was proposed by Rice [3] in 1968 based on the deformation theory of plasticity, and is known as the J-integral, a path-independent mechanics’ parameter. Originally, this parameter was used to measure the intensity of the HRR singular crack-tip field developed by Hutchinson [16], and Rice and Rosengren [17], for elastic-plastic hardening materials. Finite-element analysis showed that the J-integral can well describe the stresses, strains, and other mechanics’ behaviours at the crack tip for ductile metals. The analytical and numerical results encouraged the early experimental investigations on the J-integral testing to develop a viable test method for evaluating its critical value. Among the early investigators, Begley and Landes [18, 19] successfully measured the J-integral and its critical value at the onset of ductile-fracture tearing using multiple laboratory specimens in mode-I loading. Since then, the J-integral has become a measurable material parameter and obtained extensive applications in characterizing the fracture toughness of ductile materials. -n ot f Similarly, Joyce [60] discussed more KIc issues. In addition, Wallin [13] suggested determining KQ at a fixed amount of crack growth of 0.5 mm or 2% ligament size of 1T specimen to eliminate the size effect on the KIc value. Further discussions of the validation requirements and Wallin’s suggestions to improve the KIc testing are continuing in the ASTM E08 committee. J-integral estimation for stationary cracks rib ut a. the ligament size b is the controlling dimension in a standard specimen, and KIc is a function of b because of the 95% secant procedure; b. the coefficient 2.5 in Eqn 3a may be reduced to 1.5 or smaller; and c. the Pmax/PQ requirement in Eqn 3b only ensures that the resistance curve between PQ and Pmax has a ‘power’ of less than 0.1, and it has no significance to define KIc. Early J experimental estimations is t The issues related to the validation requirements in Eqns 3 have been discussed by the ASTM Technical Committee E08 on fatigue and fracture mechanics for many years with an aim to improve the KIc testing and to determine a size-independent KIc value. Recently, Wallin [12] analysed a large amount of the KIc test data for different metals including aluminium, steel, alloy, and titanium, and concluded that: sheet metallic materials if the thickness requirement for maintaining crack-tip plane-strain conditions is disregarded. It was concluded that a J-R curve is more general and useful than a K-R curve for metallic materials. or d observed in a load-displacement record relates to crack initiation and not just to the plastic zone. However, experiments have showed that the Pmax/PQ criterion in Eqn 3b mostly invalidates good test data [1, 60]. K-R curve testing Sa m pl e co py In the 1970s, after a substantial progress was made on the KIc testing of thick-section specimens, the attention was refocused on the critical Kc testing for thin-section specimens, where plane stress prevails. Utilizing the K-factor, R-curves were able to obtain for the thin-sections using small-scale laboratory specimens, such as the CT specimens and central cracked panels. A fracture test method of K-R curves using small specimens was standardized and published in 1974 by ASTM with designation E561-74, and the current version is ASTM E561-10e1 [14]. This standard was developed particularly for K-R curve measurements of mode-I cracks for thin-sheet materials under plane-stress crack-tip conditions. Materials that can be tested for the K-R curve development are not limited by strength, thickness, or toughness, so long as specimens are of sufficient size to remain predominately elastic to the effective crack extension value of interest. Recently, Zhu and Leis [15] showed that ASTM E561 can be equivalent to ASTM E1820 (to be discussed next in detail) to determine a size-independent R-curve for the thin-section specimen of low-toughness aluminium alloy, and ASTM E1820 would be applicable to thin- In the earliest experimental evaluation, the J-integral was interpreted as a strain-energy release rate, or work done to the specimen per unit fracture surface area in a material given by: J= − dU (4) Bda where U is the strain energy, a is the crack length, and B is the specimen thickness. Begley and Landes [18] tested a series of fracture specimens of the same geometry with different crack sizes and instrumented load-displacement data. From their test data, the energy absorbed by each specimen was determined, and the J was calculated from Eqn 4. This approach was rudimentary and has shortcomings, such as multiple specimen tests and complicated experimental analysis for determining a critical Jc. Therefore, a simple experimental technique was sought for estimating the J-integral only from a single-specimen test. 148 The Journal of Pipeline Engineering io n ot f where A is the total area under a load versus load-line displacement (LLD) that represents the work done to the specimen or the energy absorbed by the specimen as a result of the presence of a crack. η is a geometry factor that is a function of a/W. Clarke and Landes [23] and Sumpter [24] obtained expressions of the η factor using the limit-analysis method for CT and SENB specimens. The J-integral estimation Equns 5 or 8 are valid only for stationary cracks in an experimental evaluation of the J-integral to obtain its critical value at ductilefracture initiation. However, the earliest J-R curves were constructed simply using the J-integral values that were calculated by Eqn 5 in terms of the original crack size and crack extension that was measured using an unloading compliance technology proposed by Clark et al. [26]. The resulting resistance curve tends to overestimate J for a growing crack because the crackgrowth correction was not taken into account. To allow crack growth, Eqn 5 or 8 has been extended in different ways to obtain a crack-growth-corrected J as needed in an accurate J-R curve evaluation. Two typical improved equations for considering the crack-growth correction are of incremental functions, where test data are spaced at small intervals of crack extension and the J is evaluated from the previous step. The first J-integral incremental equation was proposed by Garwood et al. [27] and improved by Etemad and Turner [28], for a single-edge-notched bend specimen with a deep crack. At the nth step of crack growth, the total J-integral was estimated by: rib ut η A(5) Bb J-integral estimation for growing cracks is t J= where BN is a net thickness for the specimen with side grooves, η denotes a plastic geometry factor, and Apl is the plastic area under the load-LLD curve obtained in a fracture test. Equations 6-8 were adopted in the first ASTM fracture-toughness test standard E813-81 [25] that was published in 1982, and now are used in the basic procedure of the current version E1820-09e2 [4] to evaluate the plain-strain initiation toughness JIc, when a crack-growth resistance is not desired. or d Among many others, Rice et al. [20] showed that the J-integral can be simply determined from a loaddisplacement curve obtained in a single-specimen test using an approximate evaluation formula. They proposed several simple J evaluation equations for different specimens they considered. Here, only the single-edge-notched bend (SENB) specimen and compact-tension (CT) specimen in mode-I loading are discussed. Through further investigations by Landes et al. [21] and Merkle and Corten [22], a more general equation for estimating the J-integral in a singlespecimen fracture test by use of the SENB and CT specimens was developed as: py -n For convenience, a total load-line displacement is often separated into an elastic component and a plastic component. On this basis, the total J-integral can be split into elastic and plastic components that are determined separately: = J J el + J pl (6) co ηn ∆U n ,( n −1) g (η ) n = J n J n −1 1 + (an − an −1 ) + (9) (W − an ) B(W − an ) The objective of the above separation is to improve the m pl e accuracy of the J-integral estimation, and to obtain the consistent value of J when deformation is near linear elastic conditions. In Eqn 6, the elastic component Jel can be calculated directly from the K-factor for a plane strain crack: K 2 (1 −ν 2 ) (7) E Sa J el = where E is Young’s modulus and ν is Poisson’s ratio. For a stationary crack, the plastic Jpl is determined from Eqn 5 as: J pl = η Apl BN b The second J-integral incremental equation was proposed by Ernst et al. [29] based on the principle of variable separation. Since the J-integral was developed in reference to the deformation theory of plasticity, it was shown that the J is independent of the loading path leading to the current values of loadline displacement and crack size in the J-controlled crack-growth conditions. As a result, the deformation J-integral is a unique function of two independent variables: load-line displacement and crack length. On these bases, Ernst et al. [29] obtained an incremental equation to evaluate the total J-integral at the ith step of crack growth in the form of: η γ J i J i −1 + i −1 Ai −1,i 1 − i −1 (ai − ai −1 ) (10) (8) = Bb b i −1 i −1 3rd Quarter, 2013 149 (11) io n rib ut where the CMOD-based plastic geometry factor was obtained in Ref.31 from their FEA results. Advances in J experimental estimations More accurate J-integral incremental equations In the experimental evaluation of J-R curves, the LLD-based J-integral incremental Eqn 11 has been used widely as an ‘accurate’ expression for more than 30 years because it considers crack-growth correction and was adopted by ASTM E1820. In contrast, the other incremental Eqn 9 did not receive much attention until 2008 when two similar equations were proposed by Neimitz [32] and Kroon et al. [33]. However, Zhu and Joyce [34] showed that the two ‘new’ equations are similar and equivalent to the Garwood-type Eqn 9. In addition, Tyson and Park [35] proposed a modified ASTM E1820 incremental J-integral equation in order to allow larger crack-growth increments between any two unloading-reloading cycles in a fracture test using the elastic-compliance method. In comparison to Eqn 11, it is seen that their expression is too complicated to be used in practice. -n i −1,i where the incremental plastic area Apl is calculated by: i −1,i A= pl pl K 2 (1 −ν 2 ) ηCMOD ACMOD + (13) E Bb ot f γ η = J pl (i ) J pl (i −1) + i −1 Aipl−1, i 1 − i −1 (ai − ai −1 ) BN bi −1 bi −1 J= is t Based on the incremental equation (10) for a J-integral evaluation, the first J-R curve test standard ASTM E1152-87 [30] was standardized and published in 1987. This standard separated the J-integral into elastic and plastic components and determined them separately and incrementally at each loading step. The elastic component of J is calculated directly from the K-factor using Eqn 7, and the plastic component of J is determined incrementally from Eqn 10 in the form of: specimen. Following the basic idea of Sumpter, Kirk, and Dodds [31] studied several possible J-integral estimation approaches for shallow cracked SENB specimens using detailed elastic-plastic finite-element analyses (FEA). They found that the LLD-based J-estimation equation could give inaccurate results for hardening materials because the LLD-based plastic η factor is very sensitive to the strain-hardening exponent for SENB specimens with shallow cracks of a/W < 0.3. In contrast, for the same geometry, the CMOD-based plastic η factor is nearly insensitive to the strain-hardening exponent, when a similar η-factor equation was used with the plastic area being obtained under a load-CMOD curve. Thus, Kirk and Dodds [31] concluded that the CMOD-based J-estimation is the most reliable, and suggested use of the following equation in a J-integral evaluation for SENB specimens: or d where γ is a geometry factor related to the plastic η factor, Ai-1,i is the incremental area under an actual load-displacement record from step i-1 to i. Both incremental equations in Eqns 9 and 10 consider the crack-growth correction on the J-integral from the last step. Moreover, Eqn 10 makes the correction on the incremental work done to the specimen, but Eqn 9 does not. Consequently, a larger J is likely estimated from Eqn 9 than from Eqn 10, as shown experimentally in Ref. 29. In general, these two incremental formations of the J-integral equation are applicable to any specimen, provided the two geometry factors are known for each specimen. 1 ( Pi + Pi −1 ) ( ∆ pl (i ) − ∆ pl (i −1) ) (12) 2 Sa m pl e co py where Δpl is the plastic component of load-line displacement. Accurate estimation of the plastic component Jpl(i) at each loading step using Eqn 11 requires small and uniform crack-growth increments. Accordingly, a loading increment between the two loading-unloading cycles must be small, and usually 30 to 60 cycles are sufficient if the elastic-compliance method prescribed in E1820 is used. Equivalently, a crack-growth increment is required be less than 1% of the crack ligament size in order to use Eqn 11 for an accurate J-integral evaluation [34]. With the calculated values of Ji and the measured values of crack extension (Δa = ai - a0), where a0 is an original crack length, a J-R curve is obtained by plotting J against successive increments of crack growth from a single-specimen test. CMOD-based J estimation for stationary cracks Experiments showed that an accurate measurement of load-line displacement (LLD) is more difficult than that of crack-mouth-opening displacement (CMOD) for the SENB specimens in three-point bending, particularly for a shallow crack. Sumpter [24] first used the load-CMOD data directly in a J-integral evaluation using a bending To obtain a more-accurate J-integral incremental equation for a growing crack, Zhu and Joyce [34] developed different mathematical models and physical models, and obtained the corresponding incremental J-integral equations. Three physical models were assumed to approximate the integration path of a differential of the J-integral along the actual load-displacement curve obtained in a fracture test for a growing crack. For convenience, these physical models were referred to as the upper-step-line approximation (USLA), the lowerstep-line approximation (LSLA), and the mean-step-line 150 The Journal of Pipeline Engineering For the LSLA model, the J-integral incremental equation is: γ η = J pl (i ) J pl (i −1) 1 − i (ai − ai −1 ) + i Aipl−1,i bi BN bi (15) For the MSLA model, the J-integral incremental equation is: 1γ γ = J pl (i ) J pl (i −1) 1 − i −1 + i (ai − ai −1 ) 2 b b i i −1 1 ηi −1 ηi i −1,i 1 γ i −1 γ i + + + Apl 1 − 2 BN bi −1 bi 4 bi −1 bi (16) (ai − ai −1 ) Typically, to obtain an adequate normalization function, a blunted crack size is used, and measured loads are normalized: PNi = Pi WB [1 − abi / W ] η (17) where i refers to the i-th loading point, PNi is a normalized load and abi is the blunting corrected crack length. In the same time, the measured plastic displacement is normalized: -n ot f Comparisons of Eqn 14 with Eqn 11, and Eqn 15 with Eqn 9, show that the J-integral incremental equation for the USLA model is the same as the Ernst-type equation, and the incremental equation for the LSLA model is identical to the Garwood-type equation. Equation 16 for the MSLA model is a new incremental equation that is equivalent to the average of Eqns 14 and 15. Furthermore, Zhu and Joyce [34] and Zhu [36] showed that: io n (14) rib ut γ η = J pl (i ) J pl (i −1) + i −1 Aipl−1,i 1 − i −1 (ai − ai −1 ) B b b N i −1 i −1 is t For the USLA model, the J-integral incremental equation is: The concept of the normalization method was proposed by Herrera and Landes [37] for determining a J-R curve directly from a load-displacement record in a single-specimen test. This method requires an adequate calibration function to fit the relation between the normalized load and the normalized plastic displacement. Different calibration functions were investigated, such as a power-law function, and a combined function of power law and straight line. A three-parameter LMN function proposed by Landes et al. [38] was found to be appropriate. Joyce [39] improved the LMN function as a four-parameter normalization function, and the corresponding normalization method was finally accepted by ASTM E1820-01 and all later versions in Annex 15 Normalization data reduction technique. or d approximation (MSLA). For each physical model, the authors developed an incremental equation for estimating the J-integral with considering the crack growth correction. pl e co py a. the Garwood-type incremental equation (15) could overestimate a true J-R curve; b. the Ernst-type incremental equation (14) always underestimates the true J-R curve; and c. the new Eqn 16 determines a J-R curve that well matches the true curve with much higher accuracy than the two existing incremental J-integral equations. m Normalization method Sa The two conventional test techniques, i.e. the elasticunloading-compliance method and the electric-potentialdrop method, are often used for growing-crack-size measurement. They can be difficult or impractical to implement under severe test conditions, such as high loading rate, high temperature, or aggressive conditions. A normalization method was then developed as an alternative approach for directly estimating instantaneous crack lengths from a load vs load-line-displacement curve in conjunction with the use of initial and final measurements of physical crack sizes. This method does not require any test devices for online monitoring crack growth, and thus the test costs are reduced. ∆= pli ∆ pli ∆ i − PC i i = (18) W W where Ci is the specimen load-line compliance using the blunting corrected crack length abi. Using Eqns 17 and 18, the measured load and displacement data, up to but not including the maximum load, are normalized. The final load-displacement pair is normalized using the same equations except for the final crack length without blunting correction. From the final normalized point, a tangent line is drawn to the normalized load-displacement curve to define a tangent point. Using the normalized load-displacement pair PNi, ∆ pli, a normalization function can be fitted using the least-squares regression in the form of: PN = c1 + c2 ∆ pl + c3 ∆ 2pl c4 + ∆ pl (19) where c1, c2, c3, and c4 are the fitting coefficients. With this normalization function, an iterative procedure is further used to force all (PNi, ), and ai data at each loading point to lie on the fitted function as in Eqn 19 by adjusting ai. In this way, crack lengths at all data points can be determined, the J is calculated from Eqns 7, 8, and 11, and a J-R curve is obtained. 3rd Quarter, 2013 151 Modified-basic method is t The elastic component and total value of the J-integral are still calculated by Eqns 6 and 7, respectively. Note that an equation similar to Eqn 21 was recently proposed by Cravero and Ruggieri [45] in a different analysis for a single-edge-notched tension (SENT) specimen. For a special case with an equal LLD and CMOD, such as for CT specimens where LLD could be estimated directly from CMOD gauges mounted at the loadline, the two incremental Eqns 11 and 21 become identical to each other. In general, Eqn 21 can be used for any specimen, provided that the corresponding geometry factors ηCMOD and γCMOD are known a priori for that specimen. ot f J pl (i ) (a0 ) α − m ∆a (20) 1+ ⋅ α + m b0 1 ( Pi + Pi −1 ) (Vpli − Vpli−1 ) (22) 2 -n = J i ( ∆a ) J el (i ) (a0 ) + i −1,i A= V pl or d To unify the different fracture-test standards developed in Europe and in the USA, Wallin and Laukkanen [43] proposed a new evaluation procedure to correct ductile-crack growth in a J-R curve evaluation. This procedure is regarded as an improved basic method of ASTM E1820, and so it is referred to as a ‘modifiedbasic’ method. In this method, four steps are needed to determine a final crack-growth corrected J-R curve: rib ut io n For SENB specimens in three-point bending, successful of a crack-growth corrected J-R curve. To this end, Zhu et applications of the normalization method were al. [44] developed a CMOD-based J-integral incremental demonstrated by the present author and his coauthors: equation for determining the plastic component of the Zhu and Joyce [40] for HY80 steel, Zhu and Leis J-integral that is similar to ASTM E1820 LLD-based [41] for X-80 pipeline steel, and Zhu et al. [42] for J-integral incremental equation: A285 carbon steel. They compared experimental J-R i −1 i −1 curves obtained using the normalization method with ηCMOD γ CMOD i −1, i A = J J + 1 (ai − ai −1 ) (21) − those obtained using the elastic-compliance method or pl (i −1) V pl pl ( i ) bi −1 BN bi −1 the electrical-potential method. Combined with other comparisons, they concluded that the normalization method is equivalent to the unloading-compliance In this equation, ηCMOD and γCMOD are two CMOD-based method and the potential-drop method in a J-R curve plastic geometry factors, denotes the incremental area under the P-Vpl curve (where Vpl is plastic CMOD), evaluation from a single-specimen test. and is calculated by: py where α =1 for SENB specimens and α = 0.9 for CT specimens; m is a curve-fitting parameter from experimental data. m pl e co The new correction procedures have been developed for standard CT and SENB specimens, and are generally valid for both LLD-based and CMOD-based J-integral calculations. The procedures are applicable to both single-specimen tests and multiple-specimen tests, and have the same or better accuracy as the crack-growth correction used in the present ASTM E1820. Therefore, this modified-basic method was adopted by ASTM E1820-05 and its later versions in Annex A16 Evaluation of crack growth corrected J-integral values. Sa CMOD-based J-integral incremental equations Since CMOD measurements are generally more accurate than LLD measurements, a fracture test using SENB specimen favours CMOD gauges for measuring displacement and specimen compliance. Using loadCMOD data, a crack-growth corrected J-R curve can be determined using the modified-basic method outlined above. However, the suggested correction procedure is indirect and involves multiple steps in determining a crack-growth corrected J-R curve. A direct CMOD method has been desired for a long time for the determination Due to the more-accurate CMOD data are used in Eqns 21 and 22, this new J-integral incremental equation is able to determine a more-accurate J-R curve in a single-specimen test. Moreover, because LLD data are not needed in the CMOD-based J-integral estimation, LLD gauges are not required. Thus, the fracture test becomes more cost-effective. Plastic geometry factor determination for SENB specimens In the LLD- and CMOD-based J-integral incremental Eqns 11 and 21, two plastic geometry factors η and γ are involved. An accurate J-R curve evaluation needs accurate functions of these geometry factors, and thus their determination become very important. A brief review of determining these geometry factors was given by Zhu and his co-workers [40, 44]. The slip-line solution and the elastic-plastic finite-element calculation have been used to determine these geometry factors for the conventional fracture specimens. However, inaccurate functions of η and γ were found in the existing solutions for the SENB specimens in both LLD- and CMOD-based formulations. More-accurate functions of these geometry factors were thus determined by Zhu and Joyce [40] for 152 The Journal of Pipeline Engineering io n With the development of experimental estimation methods and experimental techniques, numerous efforts were contributed to standardize the K-factor and J-integral-based fracture-test methods. Zhu and Joyce [1] presented more detailed reviews on the K and J test method standardization. Landes [54] presented an interesting review of historical development of J-integral fracture mechanics’ and experimental testing at ASTM that involved important events, places, and people. A related review was also made in Ref.55. pl e co py -n ot f In the 1980s, the constraint-effect or size-dependence of J-R curves was not fully understood and it was thought that the deformation J-integral was incapable of describing fracture resistance of ductile materials at large crack extensions. Accordingly, Ernst [51-52] proposed a so-called modified J-integral, JM, for use in characterization of fracture resistance at large crack extensions beyond the limits of the deformation J-R curves, and this author showed that JM-R curves were consistent at the early stages with the JM vs tearingmodulus curves correlated closely for C(T) specimens. However, in the late 1980s and in the early 1990s, many experimental results showed that the modified JM-R curves were still size-dependent and may even behave worse than the deformation J-R curves. As a result, the deformation J, after correction for crack growth, continues to be used in E1820 for the J-R curve testing today. Development of fracture-toughness test standards rib ut Extensive experiments [46-48] showed that the crack-tip constraint has significant effects on J-R curves, with lower curves for standard deep cracks in bending and elevated curves for non-standard shallow cracks in bending or any cracks in tension. The present author [49, 50] proposed a constraint-correction method to determine a constraint-corrected J-R curve in terms of fracture-toughness testing, the two-parameter fractureconstraint theories, and the finite-element calculations. This constraint-correction method is a viable technology to solve the transferability of the J-R curves using small laboratory specimens to those for cracks of real structures. More detailed review on this topic can be found in Ref.1. From these observations, it should be recognized that the constraint effect on resistance curves is a natural consequence of the crack-tip constraint or triaxiality effect on the crack-tip field. Thus, the deformation J-R curve is adequate to describe the constraint effect, and it should continue to be used. is t Constraint effect and modified JM resistance curves 3. A JM-R curve is always higher than the deformation J-R curve, and may become upward hooking beyond 30% crack extension for small or deeply cracked specimens. Such behaviour may lead to nonconservative results for the integrity assessment. or d SENB specimens with a wide range of crack lengths in pure bending conditions. Zhu et al. [44] obtained more-accurate functions of both LLD- and CMOD-based η and γ factors for SENB specimens with deep and shallow cracks in three-point bending conditions. The newer functions of the plastic geometry factors were adopted in the current version of ASTM E1820-11 [5]. Sa m Recently, the present author [53] obtained an incremental equation for calculating the modified J-integral in development of a resistance curve and re-evaluated the deformation vs modified J-integral resistance curves using historical test data and newly developed experimental data for different structural steels. The results showed that: 1. A JM-R curve is essentially the corresponding J-R curve without crack-growth correction, and thus is not a result corresponding to the deformation plasticity theory. 2. The JM-R curves are also dependent on specimen size, geometry type or loading mode, and specimen length. Thus, they are not ‘size-independent’. It is found that the process of ASTM standardization for the first KIc test method was excessively long and it took about 10 years to draft and to publish the test standard. Srawley and Brown [11] took many years to draft/revise the first KIc test method and published it in 1965 [11]. ASTM assigned it a temporary standard designation of ASTM E399-70T in 1970, and officially published it in 1972 [1]. This standard became the model for all subsequent fracture-toughness test standards in ASTM. In contrast, the first ASTM R-curve test standard E561-74 took a shorter time and was published in 1974. These two standards are in the present versions of ASTM E399-09e2 and ASTM E561-10e1. In fact, the latest version of E399 is E399-12e1 as shown only on the ASTM website. In this newest version, the Pmax/PQ criterion might be removed and some other improvements may be updated for the KIc testing. Similar to ASTM E399, the standardization of the first JIc test method took about 10 years to draft and to publish [1]. The first JIc test standard was published in 1982 with a designation of ASTM E813-81 [25], in which only the test result of the critical J-integral was accepted as the fracture toughness of materials. Similarly, the first J-R curve test standard ASTM E1152-87 [30] also took about 10 years to be developed and got published in 1987. Then about 10 years later, ASTM E1737-96 [57] was formed in 1996 by merging E813 for the JIc testing and E1152 for the J-R curve testing. At the same time, a common combined fracture-test standard ASTM E1820-96 [58] was published for 3rd Quarter, 2013 153 Conclusions Sa m pl e co py -n io n rib ut ot f This paper reviewed the historical efforts and recent advances in development of the K-factor and J-integralbased fracture-toughness testing, experimental estimation, and standardization at ASTM in the USA. While the KIc and JIc are used to characterize fracture-initiation toughness, a K-R curve is used to characterize the fracture resistance for thin-sheet materials with the elastic deformation dominated at the crack tip and a J-R curve is used to characterize the fracture resistance for thicker specimens of ductile materials. The traditional J-R curve evaluation was LLD-based, and has been used for more than 30 years. A moreaccurate J-R curve estimation method was recently developed by use of CMOD only. In addition, this review discussed the KIc issues including the Pmax/PQ criterion used in ASTM E399, and the developments of the normalization method, modified-basic method, more-accurate J-integral incremental equations, CMODbased J-R curve evaluation, more-accurate functions of the plastic-geometry factors, and other progress recently made for ASTM E1820. It is anticipated that this review will help users to better understand and use ASTM E399 for K testing, and ASTM E1820 for the J testing. 1. X.-K.Zhu and J.A.Joyce, 2012. Review of fracture toughness (G, K, J, CTOD, CTOA) testing and standardization. Eng Fract. Mech., 85, pp 1-46. 2. G.R.Irwin, 1957. Analysis of stresses and strains near the end of a crack traversing a plate. J.Applied Mechanics, 24, pp 361-364. 3. J.R.Rice, 1968. A path independent integral and the approximate analysis of strain concentration by notches and cracks. Idem, 35, pp 379-386. 4. ASTM, 2012. ASTM E399-09e2: Standard test method for linear-elastic plane-strain fracture toughness KIc of metallic materials. American Society for Testing and Materials, West Conshohocken, PA, USA. 5. Ibid., 2012. ASTM E1820-11: Standard test method for measurement of fracture toughness. American Society for Testing and Materials, West Conshohocken, PA, USA. 6. A.A.Griffith, 1920. The phenomena of rupture and flow in solids. Philosophical Transactions of the Royal Society of London, Series A, 221, pp 163-197. 7. M.L.Williams, 1957. On the stress distribution at the base of a stationary crack. J.Applied Mechanics, 24, pp 109-114. 8. G.R.Irwin and J.A.Kies, 1954. Critical energy rate analysis for fracture strength, Welding Journal - Research Supplement, 19, pp 193-198. 9. J.M.Krafft, A.M.Sullivan, and R.W.Boyle, 1961. Effect of dimensions on fast fracture instability of notched sheets. Cranfield Symposium, I, pp 8-28. 10.G.R.Irwin, J.A.Kies, and H.L.Smith, 1958. Fracture strengths relative to onset and arrest of crack propagation. Proc. American Society for Testing and Materials, 58, pp 640-660. 11. J.E.Srawley and W.F.Brown, 1965. Fracture toughness testing methods. In: Fracture toughness testing and its applications. ASTM STP 381, American Society for Testing and Materials, pp 133-145. 12. K.Wallin, 2012. E399 size requirements and relevance of the Pmax/PQ criterion, ASTM E08 Workshop on KIc testing, May, Phoenix, Arizona, USA. 13. Ibid., 2005. Critical assessment of the standard ASTM E399. Journal of ASTM International, 2, JAI12051. 14.ASTM, 2012. ASTM E561-10e1: Standard test method for K-R curve determination. American Society for Testing and Materials. 15.X.-K.Zhu and B.N.Leis, 2009. Revisit of ASTM round robin test data for determining R curves of thin sheet materials. Journal of ASTM International, Paper ID JAI102510. 16.J.W.Hutchinson, 1968. Singular behavior at the end of a tensile crack in a hardening material. J. Mechanics of Physics and Solids, 16, pp 13-31. 17.J.R.Rice and G.F.Rosengren, 1968. Plane strain deformation near a crack tip in a power law hardening material. Idem, 16, pp 1-12. is t The different versions of ASTM E399 and E1820 reviewed here reflect the improvements and updates of these fracture-toughness test standards made by ASTM over the past 50 years or so. Experimental technique and development of fracture-toughness testing are not reviewed in this paper, but can be found in an ASTM manual by Joyce [59]. References or d measuring the critical values of J, K, and δ (the cracktip-opening displacement) as well as J-R curve and δ-R curve. Due to the E1820 publication, E1737 was withdrawn in 1998. The latest version ASTM E182011 [5] incorporates many recent updates, such as the normalization method in annex A15, the modifiedbasic method in A16, the CMOD-based η-equation for basic procedure in A1, the CMOD-based incremental equation for the resistance curve procedure in A1, and the more-accurate expressions of the plastic η and γ factors in A1. Most recently, the E1820 also added two new annexes: A17, a fracture-toughness test method at impact loading rates using pre-cracked Charpy-type specimens, and X2, the guidelines for measuring the fracture toughness of materials with shallow cracks. Acknowledgements The author is grateful to Professor James Joyce in the US Naval Academy for his useful discussions on the historic efforts of fracture-toughness testing and standardization in ASTM, and to Jesse Zhu for his helpful manuscript editing. 154 The Journal of Pipeline Engineering or d is t rib ut io n 34. X.-K.Zhu and J.A.Joyce, 2010. More accurate approximation of J-integral equation for evaluating fracture resistance curves. Journal of ASTM International, 7, 1, Paper ID JAI102505. 35.W.R.Tyson and D.-Y.Park, 2009. Modified E1820 J-integral equation (A1.8) to allow larger increments between unloading. ASTM E08.07.05 Task Group Meeting, Atlanta, GA, USA, November. 36.X.-K.Zhu, 2012. Improved incremental J-integral equation for determining crack growth resistance curves. J. Pressure Vessel Technology, 134, Paper ID: 051404. 37.R.Herrera and J.D.Landes, 1988. Direct J-R curve analysis of fracture toughness test. J. Testing and Evaluation, 16, pp 427-449. 38.J.D.Landes, Z.Zhou, K.Lee, and R.Herrera, 1991. Normalization method for developing J-R curves with the LMN function. Idem, 19, pp 305-311. 39.J.A.Joyce, 2001. Analysis of a high rate round robin based on proposed annexes to ASTM E 1820. Idem, 29, pp 329-351. 40.X.-K.Zhu and J.A.Joyce, 2007. J-resistance curve testing of HY80 steel using SE(B) specimens and normalization method. Eng Fract. Mech., 74, pp 2263-2281. 41.X.-K.Zhu and B.N.Leis, 2008. Fracture resistance curve testing of X80 pipeline steel using SENB specimen and normalization method. Journal of Pipeline Engineering, 7, pp126-136. 42. X.-K.Zhu, P.S.Lam, and Y.J.Chao, 2009. Applications of normalization method to experimental measurement of fracture toughness for A285 carbon steel. Int. J. Pressure Vessels and Piping, 86, pp 599-603. 43.K.Wallin and A.Laukkanen, 2004. Improved crack growth corrections for J-R curve testing. Eng Fract. Mech., 71, pp 1601-1614. 44. X.-K.Zhu, B.N.Leis, and J.A.Joyce, 2008. Experimental evaluation of J-R curves from load-CMOD record for SE(B) specimens. Journal of ASTM International, 5, Paper ID: JAI101532. 45. S.Craveroand C.Ruggieri, 2007. Further developments in J evaluation procedure for growing cracks based on LLD and CMOD data. Int. J. of Fracture, 148, pp 387-400. 46.J.A.Joyce E.M.Hackett and C.Roe, 1993. Effect of crack depth and mode of loading on the J-R curve behavior of a high-strength steel. In: Constraint effects in fracture, ASTM STP 1171, American Society for Testing and Materials, pp 239-263. 47.J.A.Joyce and R.E.Link, 1995. Effects of constraint on upper shelf fracture toughness. In: Fatigue and fracture mechanics: 26th Volume, ASTM STP 1256, American Society for Testing and Materials, pp 142-177. 48.Ibid., 1997. Application of two parameter elasticplastic fracture mechanics to analysis of structures. Eng Fract. Mech., 57, pp 431-446. Sa m pl e co py -n ot f 18.J.A.Begley and J.D.Landes, 1972. The J-integral as a fracture criterion. Fracture mechanics, ASTM STP 515, pp 1-23. 19.J.D.Landes and J.A.Begley, 1972. The effect of specimen geometry on JIC. Idem, pp 24-39. 20. J.R.Rice, P.C.Paris, and J.G.Merkle, 1973. Some further results of J-integral analysis and estimates. Progress in flaw growth and fracture toughness testing, ASTM STP 536, pp 231-245. 21. J.D.Landes, H.Walker, and G.A.Clarke, 1979. Evaluation of estimation procedures used in J-integral testing. Elastic-plastic fracture, ASTM STP 668, pp 266-287. 22. J.G.Merkle and H.T.Corten, 1974. A J integral analysis for the compact specimen, considering axial force as well as bending effects. J. Pressure Vessel Technology, 96, pp 286-292. 23.G.A.Clarke and J.D.Landes, 1979. Evaluation of the J-integral for the compact specimen. J. Testing and Evaluation, 7, pp 264-269. 24.J.D.G.Sumpter, 1987. JC determination for shallow notch welded bend specimens. Fatigue and Fracture of Engineering Materials and Structures, 10, pp 479-493. 25.ASTM, 1982. ASTM E813-81: Standard test method for JIc, a measure of fracture toughness. American Society for Testing and Materials, West Conshohocken, PA, USA. 26. G.A.Clarke, W.R.Andrews, P.C.Paris, and D.W.Schmidt, 1976. Single specimen tests for JIC determination. Mechanics of crack growth, ASTM STP 590, pp 27-42. 27.S.J.Garwood, J.N.Robinson, and C.E.Turner, 1975. The measurement of crack growth resistance curves (R-curves) using the J integral. Int. J. Fracture, 11, pp 528-530. 28.M.R.Etemad, S.J.John, and C.E.Turner, 1988. Elasticplastic R-curves for large amounts of crack growth. Fracture Mechanics: 18th Symposium, ASTM STP 945, pp 986-1004. 29. H.A.Ernst, P.C.Paris, and J.D.Landes, 1981. Estimations on J-integral and tearing modulus T from a single specimen test record. Fracture Mechanics: 13th Conference, ASTM STP 743, pp 476-502. 30.ASTM, 1987. ASTM E1152-87: Standard test method for determining J-R curves. American Society for Testing and Materials, West Conshohocken, PA, USA. 31.M.T.Kirk and R.H.Dodds, 1993. J and CTOD estimation equations for shallow cracks in single edge notch bend specimens. J. Testing and Evaluation, 21, pp 228-238. 32.A.Neimitz, 2008. The jump-like crack growth model, the estimation of fracture energy and JR curve. Eng Fract. Mech., 75, pp 236-252. 33. M.Kroon, J.Faleskog, and H.Oberg, 2008. A probabilistic model for cleavage fracture with a length scale – parameter estimation and predictions of growing crack experiments. Idem, 75, pp 2398-2417. 3rd Quarter, 2013 155 or d is t rib ut io n 55.X.-K.Zhu, 2009. J-integral resistance curve testing and evaluation. Journal of Zhejiang University Science A, 10, pp 1541-1560. 56. ASTM, 2013. ASTM E399-12e1: Standard test method for linear-elastic plane-strain fracture toughness KIc of metallic materials. American Society for Testing and Materials, West Conshohocken, PA, USA. 57.Ibid., 1996. ASTM E1737-96: Standard test method for J-integral characterization of fracture toughness. Idem. 58.Ibid., 1996. ASTM E1820-96: Standard test method for measurement of fracture toughness. Idem. 59.J.A.Joyce, 1996. Manual on elastic-plastic fracture: laboratory test procedures. ASTM Manual Series: MNL27. 60.Ibid., 2012. Background of KIc evaluation issues. ASTM E08 Workshop on KIc testing, May, Phoenix, Arizona, USA. Sa m pl e co py -n ot f 49.Y.J.Chao and X.-K.Zhu, 2000. Constraint-modified J-R curves and its applications to ductile crack growth. Int. J. Fracture, 106, pp 135-160. 50. X.-K.Zhu and B.N.Leis, 2006. Application of constraint-corrected J-R curve to fracture analysis of pipelines. J. Pressure Vessel Technology, 128, pp 581-589. 51.H.A.Ernst, 1983. Material resistance and instability beyond J controlled crack growth. In: Elastic plastic fracture: Second Symposium, Vol. I – Nonlinear crack analysis, ASTM STP 803, American Society for Materials and Testing, pp I191-I213. 52.H.A.Ernst and J.D.Landes, 1986. Elastic-plastic fracture mechanics methodology using the modified J, JM, resistance curve approach. J. Pressure Vessel Technology, 108, pp 50-56. 53.X.-K.Zhu and P.S.Lam, 2012. Deformation versus modified J-integral resistance curves for ductile materials. Proc. ASME Pressure Vessels and Piping Conference (PVP2012), July, Toronto, Canada. 54.J.D.Landes, 2000. Elastic-plastic fracture mechanics: Where has it been? Where is it going? In: Fatigue and fracture mechanics: 30th Volume, ASTM STP 1360, pp 3-19. e pl m Sa py co io n rib ut is t or d ot f -n 3rd Quarter, 2013 157 Low-constraint toughness testing 1 CanmetMATERIALS, Natural Resources Canada, Ottawa, ON, Canada 2 CanmetMATERIALS, Natural Resources Canada, Hamilton, ON, Canada 3 CanmetMATERIALS, Natural Resources Canada, Calgary, AB, Canada rib ut P io n by Dr William R Tyson1*, Dr Guowu Shen2, Dr Dong-Yeob Park3, and James Gianetto2 E This factor, along with the increasing importance of strain-based design allowing applied strains above yield, has led to the development of low-constraint-toughness tests that much better represent the actual loading situation for girth-weld flaws. Sa m pl e co py -n ot f NSURING THE INTEGRITY of pipeline girth welds is a vital aspect of fracture control. Defects are difficult to avoid, especially when welding in harsh environments with short construction seasons, and it is important to develop repair criteria that are safe but not overly conservative. Workmanship standards are still important to ensure that the best procedures are followed, but development of criteria based on fracture mechanics is gaining increasing acceptance. This involves comparison of measured imperfection size with a critical size (or ‘acceptable’ size including suitable safety factors) to determine whether the imperfection could impair the integrity of a pipeline under worstcase loading situations and therefore must be rejected or repaired. The determination is done by comparing the worst-case crack-driving force (the crack-tip-opening displacement CTOD, or J integral) with the fracture toughness (CTOD, or J integral resistance) of the region of the weld where the imperfection is located. In the pipeline industry, toughness is commonly measured using three-point-bend specimens which provide high constraint (high triaxial stresses at the crack tip) and therefore conservative toughness estimates. Constraint has a substantial effect on toughness. Girth-weld flaws are subjected to primarily tensile loading in service which is a low-constraint situation, and the use of highconstraint toughnesses could unnecessarily penalize the materials by dictating a need for unnecessary repairs. or d is t IPELINES IN UNSTABLE terrain are sometimes subjected to large-scale bending and tensile deformation, which places the girth welds in tension.These welds may contain small weld flaws, and it is important to evaluate the resistance to growth (toughness) of these flaws under the relevant stresses. Current standards require that toughness be evaluated in ‘high-constraint’ bending tests, and this can significantly underestimate the crack-growth resistance in tension.To provide a more meaningful measure of toughness,‘low-constraint’ tests are being developed at a number of laboratories around the world using single-edge-notched tensile specimens (SENT, or SE(T)). The intent of this paper is to describe these developments and to indicate the state-of-the-art in measuring crack-growth-resistance curves (R-curves) that can be used to assess the tolerance of weld flaws to tensile loads. * Corresponding author’s contact details: tel: +1 613 992 9573 e-mail: [email protected] Background It has been known for some time that shallow cracks and tensile loading decrease constraint from that of deeply cracked bend specimens and thereby increase the resistance (toughness) of steel (see, for example, [1]). A practice for testing single-edge-notched specimens in tension (SENT, or equivalently SE(T); the latter term will be used here) was introduced by Det Norske Veritas (DNV) in 2006 [2]. However, this practice required the testing of multiple specimens to generate data of toughness as a function of crack growth (Resistance, or R-curve) which is costly in terms of testing time and material. Tests requiring only a single specimen have been developed in several laboratories around the world, and those of CANMET [3] and ExxonMobil [4] are nearing standardization. This paper is meant to provide an introduction to the state-of-the-art in low-constraint toughness testing suitable for linepipe steel rather than an in-depth review of the evolution of low-constraint toughness tests. It is heavily weighted by the authors’ experience, and apologies are extended to the many scientists around the world who are involved in developing this technology but whose work is not referenced for lack of space in this brief article. 158 The Journal of Pipeline Engineering is t rib ut io n selected to simulate the constraint experienced by a circumferential surface flaw in a girth weld1 as closely as possible. It has been shown by FEA that the crack-tip stress state in such a specimen is similar to that in a surface-cracked pipe for cracks of the same height a/W, where a is the crack size and W is the specimen width [5]. In order to ensure transferability of R-curves between test specimen and pipe, it is important that W be the same for both geometries, i.e. that W (= B for B x B specimens) be as close as possible to the pipe wall thickness. The specimen is instrumented with a single clip gauge to measure CMOD in the CANMET procedure, and with two clip gauges mounted at different heights above the specimen surface in the ExxonMobil procedure. A specimen instrumented with a single gauge is shown being tested in Fig.1; the formation of a plastic hinge in the ligament in the plane of the crack is easily visible. Fig.1. SE(T) specimen under load. -n ot f or d In both procedures, unloading compliance (UC) is used to measure crack size, although other methods are allowed if they can be shown to give good estimates of crack size. Because of the significant rotation of the specimen evident in Fig.1, it is important to correct the compliance for rotation; that is, the unloading compliance of a deformed specimen differs from the compliance of a straight specimen with the same crack size because of the different geometries. The correction can be made as a function of the ratio of the applied load to the ligament limit load [6]. py SE(T) test methods Sa m pl e co In addition to the DNV multi-specimen procedure, two single-specimen procedures are in circulation. They are very similar in most respects, and the major difference between them is in the clip-gauge instrumentation and the choice of toughness parameter. The ExxonMobil procedure [4] uses a double-clip-gauge arrangement directly to measure the CTOD at the original cracktip location, while the CANMET (Canada Centre for Materials and Energy Technology) procedure uses a single clip gauge to measure the crack-mouth-opening displacement (CMOD) and thereby calculate the J integral from load-CMOD data. In the CANMET procedure, the CTOD is calculated from the J integral using the relationship between the two parameters derived from finite-element analysis (FEA). The specimen design is identical in the two procedures. This is a square-cross-section bar (B x B, where the specimen width W = B, and B is the specimen thickness) clamped between fixed grips with a ‘daylight’ H between the grips of ten times the specimen width (H = 10B). This specimen design and fixed-grip configuration was In both procedures, side grooves are recommended to generate a stress state close to that of plane strain, which simulates the stress state along most of the crack front of a circumferential surface flaw in a pipe. Side-groove depths of 5% and 7.5% on each side are recommended by ExxonMobil and CANMET respectively. In the ExxonMobil procedure, the CTOD at the position of the original crack tip is measured by triangulation from the openings of the two clip gauges, and a plot of the CTOD as a function of the crack growth Δa then yields the CTOD R-curve. In the CANMET procedure, the area under the load vs CMOD curve is used in conjunction with equations developed using FEA to calculate the J integral; a plot of the J integral vs crack growth then gives the J-R curve. The CTOD can be calculated from J, following general procedures that have been adopted by ASTM for single-edge bend SE(B) tests [7], using relations between J and CTOD again developed for SE(T) specimens using FEA. Details 1. Extensive FEA calculations have been done to confirm similarity of constraint. The alternative gripping arrangement – by pin loading – is not only impractical because of the geometry (i.e. surface rather than throughthickness notches) but, owing to unrestrained rotation of the specimen, pin loading is accompanied by a large amount of bending which increases the constraint beyond that experienced in the field. 159 rib ut io n 3rd Quarter, 2013 co py -n ot f ) or d is t Fig.2. J-R curves for an X-100 pipe steel, with rotation correction. PS: plain-sided; SG: side-grooved (from [6]). pl e Fig.3. Crack size for clamped SE(T) specimens: ao, af : initial and final crack size (SG: side-grooved, PS: plain-sided) (from [11]). Sa m of the procedures (equations for calculation of crack size from compliance and for calculation of CTOD and J) can be found in the published recommended practices [3, 4]. A typical set of J-R curves is shown in Fig.2 [6]. The sided-grooved specimens show, on average, slightly lower R-curves than the plain-sided specimens, but the effect is not large. The main advantage of the side grooves is to promote straightness of the growing crack. Determination of crack size by optical measurements on the fracture surface to ensure correspondence between the size estimated by unloading compliance and the actual size follows well-established methods such as found in ASTM standard E1820 [7]. Good agreement can be found with careful experimentation as shown in Fig.3 [11]. Finally, it is important to appreciate the magnitude of the difference in R-curves resulting from differences in constraint, and Fig.4 [11] illustrates this for a highstrength pipe steel. All specimens were pre-cracked to a/W = 0.5, and the curves with symbols in the figure show data for tensile SE(T) tests, and the continuous curves show results for bend SE(B) tests. Clearly, the SE(T) tests exhibit higher toughnesses than the SE(B) tests. Some of the specimens were plain-sided (PS), others were side-grooved (SG). For the bend specimens (solid-line curves), there is a clear difference between The Journal of Pipeline Engineering ot f or d is t rib ut io n 160 -n the two conditions, the side-grooved specimens having lower R-curves. There is also a reduction in toughness with side grooves for the tensile specimens, although the difference is much less pronounced. Sa m pl e co py The general conclusion is that tensile loading on SE(T) specimens is reflected in higher R-curves than threepoint bending on SE(B) specimens (i.e. imposes less constraint than bending), at least for deep cracks. For shallow cracks, the situation is not so clear: experience shows that SE(T) R-curves are significantly less sensitive to a/W than is the case for SE(B) R-curves, so that the difference between tensile and bending loading diminishes as a/W decreases. It may be the case that shallow-cracked-bend specimens yield R-curves that are only slightly more conservative than tensile specimens at the same crack size, but this remains to be established. In any case, it is certain that the R-curves generated by testing shallow-cracked specimens in tension (simulating service loading of girth-weld flaws) are substantially higher than R-curves from deep-cracked specimens tested in bending (as now used to characterize weld toughness). Some data from reference [5] shown in Table 1 emphasize this point, showing that the toughness for a deep crack in bending (SE(B)) is only approximately half that of a shallow crack in tension (SE(T)). The data in the figures above have been obtained using homogeneous material (base metal). The reason for this is that the tests were done to illustrate the effect of constraint on R-curves and to demonstrate that it Specimen type Fig.4. J-R curves measured using SE(T) and SE(B) specimens, a/W = 0.5 (from [11]). a/W PS SG SE(B) 0.5 650 450 SE(T) 0.25 1100 900 Table 1. J values (kJ/m2) at Δa = 0.5 mm. is feasible to perform low-constraint SE(T) tests. Of course, the practical application of the test methods will be to characterize welds. In fact, SE(T) testing has been used to derive R-curves for weld metals (WMs) and heat-affected zones (HAZs) as well as base metals (BMs); for example, see [10] from which Fig.5 has been taken. Special problems associated with toughness testing of WMs and HAZs are location of the notch and path taken by the crack during tearing propagation. Pros and cons Of the three current test procedures (DNV, CANMET, ExxonMobil), the latter two offer economies in test time and material consumption compared to the former because they require testing of only one B x B specimen rather than the minimum of six 2B x B specimens required for the DNV procedure. Regarding crack-size measurement, a major difficulty with using the UC technique is the prevalence of apparent negative crack growth on initial loading. This 161 is t rib ut io n 3rd Quarter, 2013 ot f or d Fig.5. SE(T) J-R curves for X-100 BM,WMs, and HAZs for single-torch welds (R1) and dual-torch welds (R2); H/W = 10, a/W = 0.34 (from [10]). The advantage of this procedure is that corrections can be made to the J integral for crack growth, using deformation plasticity [8], enabling the actual driving force to be found. The disadvantage is that the value of m must be estimated, and because m is a function of mechanical properties (primarily the work-hardening coefficient) and crack size, approximations must be made in estimating CTOD from J. Additionally, the CTOD in this procedure is somewhat artificial since it is a calculated property rather than a physical measurement on the actual specimen. For both of these reasons it would be better to use J as the driving force rather than CTOD. Also, for the tough steels used in current pipeline construction there is extensive plasticity in the ligament, making the definition of J and its calculation using FEA questionable. Nevertheless, small-geometrychange FEA yields consistent values of far-field J that can be used for both small-scale measurement procedures and large-scale (for example, cracks in pipe) geometries. With the definition of CTOD used in the ExxonMobil procedure, the advantage is that an actual physical quantity is measured that can be readily interpreted, both in small-scale and large-scale situations. co py -n phenomenon is familiar in SE(B) testing as well, and is difficult to deal with since it appears seemingly randomly and is not consistently connected with any procedural detail. Nevertheless, ASTM E1820 [7] contains a methodology to correct for it. Both CANMET and ExxonMobil procedures use a similar technique to derive a corrected initial crack size a0q when apparent negative growth is encountered. Results show that this technique works well, with a0q being very close to the optically-measured initial crack size. Sa m pl e Both of the single-specimen procedures deliver similar results for CTOD at small amounts of crack growth (defined in this paper to include blunting). In this region, both procedures apply a practically equivalent definition of the CTOD. However, with larger amounts of growth – roughly, larger than 0.5 mm – it is expected (although this has not been demonstrated) that there will be a difference in the R-curves. This is because the ExxonMobil procedure measures the opening at the original crack tip while the CANMET procedure derives the CTOD that would be appropriate for a crack of the current size at the same load. That is, the CANMET procedure uses the J integral as the crack driving force variable with the CTOD being estimated from the relation J = mσYδ where σY is the flow stress (average of yield and tensile strength), δ is the CTOD, and m is a parameter obtained from FEA calculations. The CTOD in this case is defined as the opening between the intercepts of the crack flanks with lines drawn at ±45º from the crack tip. Demonstration and application All candidate procedures have been used in a variety of contexts, up to and including practical material assessment in strain-controlled situations. To obtain general acceptance of a test procedure, however, it is first necessary to demonstrate the practicality and 162 The Journal of Pipeline Engineering as possible the actual field conditions. This objective has been addressed vigorously with the development of toughness tests that have been described briefly in this paper. A multi-specimen test has already been published as a Recommended Practice, and more-efficient singlespecimen tests are moving steadily toward standardization. The intent of the present paper is to draw attention to these developments and to highlight the relevant literature where details may be found. To use the R-curves measured in SE(T) tests, it is necessary to compare the material resistance (toughness as a function of crack growth) with the crack-driving force in the structure (CTOD or J as function of crack size and load or strain). The advantage of the CANMET procedure is that the driving force can be calculated using FEA with the same definitions of J and CTOD as in the SE(T) tests. However, using the ExxonMobil definition of CTOD (the opening at the original crack tip), there is no straightforward way to derive the CTOD numerically for comparison with the R-curve without simulating the crack growth. That is, to estimate the driving force in terms of CTOD as the crack grows it is necessary to model the crack growth in the FEA calculations, or to directly measure the CTOD on experimental test pipes. It is to be expected, although this has not been demonstrated, that the CTOD at the original crack tip would increase more rapidly with applied strain in the presence of crack growth than if the crack were stationary. In other words, the definition of CTOD significantly affects the procedure that is used to predict crack growth and instability in service. Acknowledgements io n repeatability of the method. The DNV procedure is already published as a Recommended Practice and has been used by the pipeline industry. A round-robin has been completed for the CANMET procedure applied to pipe steel base metal [9], and another round-robin that will amalgamate both the CANMET and ExxonMobil procedures and demonstrate practical application to welds has been launched with the support of members of PRCI. -n ot f or d is t rib ut This work has been made possible by the support of the Federal Program on Energy Research and Development (PERD). The research upon which the present review is based has been supported by PERD as well as a large number of organizations, notably the Pipeline and Hazardous Materials Safety Administration of the US Department of Transportation (DOT) and the Pipeline Research Council International, Inc. (PRCI). The authors have benefited by discussions with leading researchers from laboratories around the world including ExxonMobil Upstream Research Co, the Center for Reliable Energy Systems (CRES), the National Institute of Standards and Technology (NIST), the Lincoln Electric Co (LECO), SINTEF, TWI, and others. The views and conclusions in this paper are those of the authors and should not be interpreted as representing the official policies of any of these organizations. pl e co py In other words, methods to use the R-curves to predict material performance (strain capacity) are still under development. In particular, there is not yet consensus on key measures to be taken from the R-curves. Leading candidates are toughness values at fixed amounts of crack growth (such as 0.5 or 1 mm) and parameters of power-law curve fits to R-curves, but toughness at the maximum load in the SE(T) test could also be considered. m Closing remarks Sa To ensure integrity of high-performance pipelines, engineers are increasingly using strain-based design (SBD) which enables the pipe to withstand strains substantially beyond yield strain. For efficient utilization of the high strength available in current materials, this requires accurate and realistic assessment of both the crack-driving force and the material resistance. In particular, reliable measurement of the toughness (R-curve) of pipe steel and welds is required for accurate prediction of the effect of conceivable flaws on the limiting strain capacity of the pipe and circumferential girth welds. To make the most efficient use of available materials, it is important that the small-scale test conditions represent as closely We would especially like to thank the staff of CanmetMATERIALS (CMAT) for their part in developing SE(T) test procedures, and the participants of the CANMET round-robin for helping to demonstrate the practicality of SE(T) testing. As noted in the introduction, apologies are extended to the many scientists around the world who are involved in developing SE(T) technology and whose work is not referenced for lack of space in this brief article. References 1. J.A.Joyce, E.M.Hackett, and C.Roe, 1993. Effects of crack depth and mode of loading on the J-R curve behaviour of a high strength steel. In: J.H.Underwood, K.-H.Schwalbe, R.H.Dodds (Eds), Constraint effects in fracture. ASTM STP 1171, American Society for Testing and Materials, Philadelphia, pp 239–263. 2. DNV Recommended Practice DNV-RP-F108, 2006. Fracture control for pipeline installation methods introducing cyclic plastic strain, Det Norske Veritas, Norway. 3. (i) G.Shen, J.A.Gianetto, and W.R.Tyson, 2008. Development of procedure for low-constraint toughness testing using a single-specimen technique. MTL Report No. 2008-18(TR); (ii) W.R.Tyson, G.Shen, J.A.Gianetto and D.-Y.Park, 2011. 3rd Quarter, 2013 163 or d is t rib ut io n 6. G.Shen and W.R.Tyson, 2009. Crack size evaluation using unloading compliance in single-specimen singleedge-notched tension fracture toughness testing, J. Testing and Evaluation, 37, 4, paper ID JTE102368, ASTM International. 7. ASTM, 2011. ASTM E1820-11: Standard test method for measurement of fracture toughness. ASTM International, West Conshohocken, PA, USA. T.L.Anderson, 2005. Fracture mechanics: 8. fundamentals and applications. CRC Press. 9. W.R.Tyson and J.A.Gianetto, 2013. Low-constraint toughness testing: results of a round robin on a draft SE(T) test procedure. Proc. ASME Pressure Vessels & Piping Division Conference (PVP2013), July, Paris, France, paper PVP2013-97299. 10.D.-Y.Park, W.R.Tyson, J.A.Gianetto, G.Shen, and R.S.Eagleson, 2012. Fracture toughness of X100 pipe girth welds using SE(T) and SE(B) tests. Proc. International Pipeline Conference (IPC2012), September, Calgary, Canada, paper IPC2012-90289. 11. G.Shen, J.A.Gianetto, and W.R.Tyson, 2009. Measurement of J-R curves using single-specimen technique on clamped SE(T) specimens. Proc. International Offshore and Polar Engineering Conference, Osaka, Japan, June. Sa m pl e co py -n ot f Development of a low-constraint SE(T) toughness Test. 10th International Conference on Fracture and Damage Mechanics, September, Dubrovnik, Croatia; (iii) Y.-Y.Wang, H.Zhou, M.Liu, B.Tyson, J.Gianetto, T.Weeks, M.Richards, J.D.McColskey, M.Quintana, and V.B.Rajan, 2011. Weld design, testing, and assessment procedures for high strength pipelines. PRCI contract [277-PR-348-074512], Summary Report 277-S-01, Section 4, Low-constraint SE(T) test protocol. 4. (i) H.Tang, M.Macia, K.Minaar, P.Gioielli, S.Kibey, and D.Fairchild, 2010. Development of the SENT test for strain-based design of welded pipelines. Proc. International Pipeline Conference (IPC2010), September, Calgary, Canada, paper IPC2010-31590; (ii) D.P.Fairchild, S.A.Kibey, H.Tang, V.R.Krishnan, X.Wang, M.L.Macia, and W.Cheng, 2012. Continued advancements regarding capacity prediction of strain-based pipelines. Proc. International Pipeline Conference (IPC2012), September, Calgary, Canada, paper IPC2012-90471. 5. G.Shen, R.Bouchard, J.A.Gianetto, and W.R.Tyson, 2008. Fracture toughness evaluation of high-strength steel pipe. Proc. ASME PVP2008 Conference, Chicago, July, paper PVP2008-61100. Held under the Patronage of His Excellency Shaikh Ahmed bin Mohamed Al Khalifa, Minister of Finance, Minister in Charge of Oil and Gas Affairs, Chairman of National Oil & Gas Authority, Kingdom of Bahrain rib ut 20–23 October 2013, Bahrain io n Keynote speaker: Mr Abdulrahman Al-Wuhaib, Senior Vice President for Downstream, Saudi Aramco Opening address by Mr Abdulhakim Al-Gouhi, General Manager - Pipelines Department, Saudi Aramco GULF CONVENTION CENTRE, BAHRAIN ORGANIZERS PLATINUM ELITE SPONSOR is t or d W ® ® Global Webb PLATINUM SPONSOR -n SUPPORTING ORGANIZATIONS ot f SILVER SPONSORS py Join leaders in the international pipeline industry as they converge for the Best Practice in Pipeline Operations and Integrity Management Conference and Exhibition in Bahrain. co CONFERENCE Sa m pl e Technical streams presented by industry leaders covering a wide range of subjects will run over the two and a half day event. Some of the subjects to be discussed; • Planning, design, construction and materials • Operations and maintenance • Asset integrity management • Inspection and cathodic protection • Repair and rehabilitation • Automation and control • Leak detection Paper abstracts are now being accepted. EXHIBITION A comprehensive exhibition will be part of the event, allowing companies from around the world to showcase their products and services. Visit our website to book your space. NETWORKING Throughout the event there will be ample opportunities to network with participants to further your business relationships. Meet with industry leaders from around the world. W O N R E T REGIS www.pipelineconf.com 3rd Quarter, 2013 165 by Dr Su Xu*1, Dr William R Tyson2, and Dr C H M Simha1 rib ut 1 CanmetMATERIALS, Natural Resources Canada, Hamilton, ON, Canada 2 CanmetMATERIALS, Natural Resources Canada, Ottawa, ON, Canada io n Testing for resistance to fast ductile fracture: measurement of CTOA E ot f or d is t NSURING ARREST OF a fast ductile fracture (i.e. a running shear fracture) is an essential design requirement for high-pressure natural gas and other (for example, CO2 and hydrogen) pipelines. Fracture-arrest toughness has traditionally been measured and specified using standard Charpy absorbed energy (CVN), such as used in the Battelle two-curve method (BTCM). But shortcomings of the Charpy test have become evident when used to characterize modern high-strength and/or high-toughness steels. The crack-tip-opening angle (CTOA) has been proposed as a better fracture-propagation toughness parameter. In order to measure CTOA using a laboratory-scale specimen, a simplified single-specimen method (S-SSM) has been developed. The S-SSM uses the familiar drop-weight tear test (DWTT) specimen, and CTOA is calculated from instrumented outputs of force vs force-line displacement. This method is being evaluated in an international round-robin project and has been proposed to ASTM for consideration of adoption. This paper describes the test development and application with typical results, and step-by-step procedure. On-going finite-element (FE) modelling work to support development of the CTOA procedure is discussed. n: exponent in hardening model P: force Pm: maximum force applied by tup during test rp: plastic rotation factor S: specimen span between two supports (S = 254 mm for DWTT tests) SE(B): single-edge-bend specimen S-SSM: simplified single-specimen method t: thickness of specimen T: temperature TM: melting temperature TR: reference temperature Sa m pl e co py a: crack length b: remaining specimen ligament B: specimen thickness C1, C2: constants in flow stress model CTOA:crack-tip-opening angle CTOAc: crack-tip-opening angle in the steady-state stage CVN: Charpy absorbed energy D: scalar damage variable DWTT: drop-weight tear test g: function modifying damage evolution law m: exponent in damage evolution law -n Alphabetical listing of abbreviations and principal symbols used P REVENTION OF A FAST ductile fracture (i.e. a long-running shear fracture) is the key element of fracture control and design of high-pressure natural gas and other (for example, CO2 and hydrogen) transmission pipelines. The required toughness (i.e. the arrest toughness) has traditionally been measured and specified using standard Charpy V-notch impact * Corresponding author’s contact details: tel: +1 905 645 0815 e-mail: [email protected] W: specimen width α: specimen rotation angle δ: constant in function g εp: plastic strain έ: plastic strain rate έo: reference rate σf: average of the yield and ultimate tensile strengths Δ: force-line displacement (LLD) Δm: force-line displacement (LLD) at maximum force ζ: negative slope of the Ln (P/ Pm) vs (Δ – Δm)/S curve ξ: thermal softening exponent absorbed energy in semi-empirical fracture-arrest models, such as the Battelle two-curve method (BTCM) [1,2]. But shortcomings of the Charpy test have become evident when used to characterize modern high-strength and/or high-toughness steels because much of CVN is related to plastic (bending) deformation and crack initiation. Also, the Charpy specimen is a small sub-pipe-thickness specimen. The use of CVN in crack-arrest prediction for high-toughness steels requires an empirical correction factor [for example, Ref.3]. The Journal of Pipeline Engineering rib ut io n 166 co py -n ot f or d is t Fig.1. Force vs displacement curve of a DWTT. Sa m pl e The DWTT specimen is full-thickness and the test is performed under impact loading [4,5]. Following experience of a number of running brittle fractures in pipelines, the DWTT has been used to ensure ductile rather than brittle fracture propagation. Historically, the DWTT was developed to determine fracture appearance and ductile-to-brittle transition temperature. Thus, pipeline standards refer only to fracture appearance rather than energy, and 85% of shear area indicating ductile fracture at the operating temperature is the usual requirement. In this case, fast fracture would be ductile tearing. To obtain a more-reliable estimation of ductile-fracture-arrest conditions, work (summarized in review [6] and [7-10]) is under way to use data from full-thickness dynamic DWTTs rather than CVN to characterize ductile propagation including using (i) DWTT total energy, (ii) DWTT propagation energy (i.e. from post-maximum-force energy, or from dynamically tested static-pre-cracked specimens), and (iii) CTOA. Fig.2. Ln(P/Pm vs (Δ-Δm)/S for all data points beyond the maximum force point (Pm, Δm) and determination of slope for CTOAc in the central part of the specimen. The CTOA has been proposed as the most promising of the fracture parameters used in ductile-fracturepropagation control. In a CTOA-based fracture-arrest methodology, the applied CTOA, usually derived from a numerical method as a function of pipe geometry and operating conditions, is compared with the material CTOA (CTOAc) that must be measured experimentally. The fracture-arrest criterion is met if the applied CTOA is equal or less than CTOAc (i.e. CTOAapplied ≤ CTOAc) [8,9]. A major challenge in applying the CTOA approach to fast-ductile-fracture control is to simplify and standardize the measurement of CTOAc in a procedure suitable for a mill test. Although CTOAc may be measured directly on lab. specimens, the optical method requires high-speed camera monitoring during DWTT, is time-consuming to analyse the images, and does not reflect the effects of constraint (i.e., variation of the CTOA through the thickness); therefore the optical CTOA method is 3rd Quarter, 2013 167 Velocity CTOA from Optical Measurement (°) (m/s) Surface Mid-thickness CTOA from S-SSM (°) X100 5×10-5 19.0 (20.0,18.0) * 11.0 (10.0,12.0) 10.4 5.1 13.0 (12.0, 14.0) 9.5 (9.0, 10.0) 9.7 5×10-5 12.0 5.5 (6.0, 5.0) 4.6 5.1 10.5 (10.0, 11.0) 6.5 (6.5, 6.5) 7.0 5.1 11.8 (11.0, 12.5) 7.3 (7.0, 7.5) rib ut X52 io n Steel 7.5 or d It has been found that the CTOA and crack velocity remain reasonably constant (steady-state stage) while the crack propagates through the central part of the ligament, and the slope of the corresponding linear part of a Ln(P/Pm vs (Δ-Δm)/S plot yields the critical CTOA value: ξ= 4r * tanγ (4) ot f not suitable for a pipe-mill test [10-14]. To overcome these shortcomings, a simplified single-specimen method (S-SSM) has been developed at CANMET [10-14]. In the S-SSM, CTOA is calculated using outputs from instrumented DWTTs of force vs force-line displacement. This indirect method is suitable for a practical mill test, and has been proposed to ASTM for consideration of adoption. The procedure is being evaluated in an international round-robin. This paper describes the test development and application with typical results, the basic step-by-step test procedure, and the on-going modelling effort. is t Table 1. CTOA of interrupted DWT/CTOA tests.Values in brackets are measurements from opposite side surfaces or sections. -n or Simplified single-specimen method co py The CTOA is related to specimen geometry via a hinge model in which rotation occurs about a centre in the ligament. It may be shown [10] that: e dα CTOA tan = rp b 2 da (1) pl The crack length is deduced from the ligament size b (b = W-a) which is in turn obtained from the force and specimen geometry as follows. Assuming that the flow strength (σf) is constant in the steady-state stage, the limit force can be written as: CTOA= c 8rp 180 ⋅ ξ π (°) (5) Figures 1 and 2 show typical force vs displacement and Ln(P/Pm vs (Δ-Δm)/S plots of an X-100 pipe steel tested using a DWTT specimen of thickness 8 mm tested at the usual impact rate (tup velocity 5.1 m/s). To apply the S-SSM, a reference point (Pm and Δm) is chosen, taken for convenience to be the maximum force. The S-SSM has been applied to typical pipe steels [13, 14]. The main conclusions are: 4rp P (∆ − ∆ m ) ln = − (3) tan(CTOA/2) S Pm Table 1 shows a comparison of CTOA values from the optical method and from the S-SSM method of typical high-strength and low-strength pipe steels [14]. Sa m where A* is a constraint factor. It follows from these equations that P, Δ, and CTOA after the peak force (Pm) are related as: crack-front tunnelling exists in DWTT specimens. The extent of tunnelling varies with steel and loading rate (Fig.3 [14]); CTOA values measured optically at the surface are higher than those measured at mid-thickness. The difference between surface and interior CTOAs increases with the extent of tunnelling; and CTOAs derived from force vs force-line displacement (i.e. using the S-SSM) are in good agreement with values measured at mid-thickness (i.e. giving conservative values for fracture control). P= 4A*σ f Bb 2 (2) S i. ii. iii. 168 The Journal of Pipeline Engineering S-SSM procedure io n The development of the S-SSM using DWTT specimens has been driven by the need to design for fast-ductilefracture arrest of axial-running cracks in steel highpressure gas pipelines [11]. The purpose has been to develop a better test to characterize fracture-propagation resistance than the traditional Charpy test in a form suitable for use as a pipe-mill test. The recommended practice measures fracture-propagation resistance in terms of CTOA, and can be used to characterize structural steels. rib ut (a) X-100, quasi-static load The S-SSM test procedure has been summarized in a step-by-step format [15], and is being evaluated in an international round-robin with seven participants. This step-by-step procedure is described below. is t Scope and summary co py -n (b) X-100, impact load ot f or d The S-SSM procedure below describes a method to determine fracture-propagation toughness in terms of the critical crack-tip-opening angle (CTOAc) using the drop-weight tear test (DWTT). The method is intended to be used for structural steels, and draws for apparatus, specimen geometry, and test procedures from ASTM E436: Standard test method for drop-weight tear tests of ferritic steels [5] and API 5L3: Recommended practice for conducting drop-weight tear tests on line pipe [4]. Sa m pl e (c) X-52, quasi-static load (d) X-52, impact load Fig. 3. Crack tunnelling of DWTT specimens [14]. In addition to data (P vs Δ) from the tests, calculation of CTOA requires only an empirically determined parameter, the plastic-rotation factor (rp). For typical structural steels, plastic-rotation factor values have been estimated experimentally, and will be discussed below. The CTOA values derived according to this recommended practice are representative of the average through-thickness CTOA values, dominated by the high-constraint middle-thickness region, and are usually lower than the surface CTOA values measured optically [13,14]. This reflects the effects of through-thickness constraint and the resulting crack-tip tunnelling. Crack velocities in the steady-state stage of crack propagation through DWTT specimens for a tup velocity of 5 m/s usually range from 12-20 m/s. DWTT specimens of modern steels tested at room temperature usually exhibit shear fracture under impact loading. For pipe steels, this mimics the fracture mode observed in full-scale pipe burst tests. The apparatus, specimens, and test procedures are consistent with those described in API RP 5L3 or ASTM E436-03, with the intent to employ standardized DWTT procedures and machines (such as hammer, anvil, and support span) to the fullest extent possible. 169 is t rib ut io n 3rd Quarter, 2013 or d Fig.5. An example of anti-buckling guide for pipe steel specimens of wall thickness less than 10 mm. Fig.4. DWTT/CTOA machine-notched specimen and dimensions. The specimens are flattened unless a straightening technique [17] is desired to leave the centre part of the specimen unflattened, which may be desirable for small-diameter pipes. The DWTT/CTOA specimen is schematically shown in Fig.4. The recommended notches are machined to a depth of 10 mm or pressed to a depth of 5 mm as required by API RP 5L3 or ASTM E436. It has been shown for typical high-strength pipe steels that the pressed-notch and machined-notch DWTT specimens produce similar CTOAc values [18]. Because the data required to derive fracture-propagation toughness (CTOA) are obtained after peak force, other notch types, sometimes used to promote easy fracture initiation (such as Chevron notch or static pre-crack), are acceptable but have little or no effect on CTOAc. ot f Apparatus co py -n The test is normally conducted using either a vertical drop-tower type or a pendulum-type-impact test machine as described in API RP 5L3 or ASTM E436-03; the key dimensions are support span S = 254 mm (10 in), impact tup radius = 25.4 mm (1 in), and fixed support anvils radius = 19.05 mm (0.75 in). The machine must provide sufficient energy to completely fracture a specimen in one impact. The absorbed energy of typical pipe steels with thickness less 20 mm is usually less than 10,000 J. pl Specimen e The initial velocity of the hammer at impact must be at least 4.88 m/s1. Sa m For pipe fracture-arrest applications, specimens are removed from the pipe such that the length of the DWTT/CTOA specimen is in the circumferential direction. For straight seam-welded pipes, the specimens are taken from the 90° position with respect to the pipe seam weld [16]. No maximum impact speed is specified in API RP 5L3 or ASTM E436. The CTOA is not sensitive to velocity in the range of 4 to 20 m/s, which is the range easily accessible in lab testing. With a drop tower, from Newton’s second law we have v = √(2ax), where v is velocity, a is acceleration due to gravity (9.8 m/s2), and x is drop height; v = 4.88 m/s corresponds to a drop height of 1.22 m (= 4 ft). 1 The specimens from steel pipe are of full thickness to ensure that, because CTOA is dependent on thickness, the CTOA measured in the test reflects the value that would be obtained for fast ductile fracture in the fullthickness pipe. Tests are considered to be invalid if the specimen buckles during impact. For specimens of thickness less than 10 mm, anti-buckling guides may be needed to prevent buckling and sliding. Simple guides acting at the ends of the DWTT specimens mounted on the anvils have proven to be effective. Figure 5 shows an example of an anti-bucking fixture used at CanmetMATERIALS. For very thin specimens, the buckling forces may be less than the crack-initiation forces and it may be necessary to reduce the width to avoid buckling. For 170 The Journal of Pipeline Engineering rib ut io n The plastic-rotation factor (rp) stems from the ‘plastic hinge’ model [20,21] that assumes the two arms of a single-edge-notched bend (SE(B)) specimen rotate symmetrically about an axis of rotation centred in the uncracked ligament. The ratio of the distance between the crack tip and the hinge point to the length of the remaining ligament is defined as the plastic rotation factor; its value is usually determined experimentally or numerically. Values of rp for structural steels in the recommended practice are: 0.57 (CVN > 100 J) rp = 0.54 (CVN ≤ 100 J) (6) These values are empirical, and reflect the change in geometry of the ligament from a rectangular to a neartrapezoidal shape. The change in geometry (thickening at the loading point and thinning at the notch) depends on the deflection of specimens before crack initiation and therefore on initiation toughness. For pipe steels, it has been shown that fracture initiation scales with notch toughness and the step change at 100 J (Eqn 6) has been chosen as suitable based on evidence in the literature [3]. ot f The CTOA test requires measurement of initial hammer velocity and of force vs time during the impact test. The force vs deflection relation can be deduced from the force vs time relation and the initial velocity of the hammer. During an impact test, the energy absorbed is supplied by the kinetic energy of the hammer, and the instantaneous velocity and displacement can be calculated based on the energy-conservation principle. High-speed data-acquisition systems and instrumented hammer tups are required. Systems commercially available for instrumented Charpy tests are acceptable for the DWTTs. Force-time data acquisition at a rate of 5 x 106/s has proven sufficient. High-speed video equipment (10,000 frames/s) can be used to monitor surface-crack propagation during impact tests but is not necessary for the determination of CTOAc according to this method. The data-acquisition system and the high-speed camera can be triggered via a photo-diode that can also be used to record initial impact velocity. is t Procedure (i.e. Ln(P/Pm)) corresponding to this range of a/W was calculated from the limit force equation (i.e. P proportional to b2) [10,12]. The best-fitting slope corresponds to the average CTOA in the steady-state region and this technique has been shown to smooth CTOA scatter in the steady-state region very effectively. or d a 3-mm thickness, the DWTT specimen width had to be reduced from 76 mm to 31 mm to prevent buckling [13]; critical CTOA values were determined in the same a/W range as the standard 76-mm-wide DWTT specimens as described below. Calculation py -n Data from an instrumented DWTT can be imported to and processed in commonly available office software (such as MS Excel). The only information required for CTOAc calculation are force vs deflection; an example of such data obtained from a test on a high-strength pipe steel is shown in Fig.1. co The steps in the recommended practice for calculating CTOA are: Sa m pl e 1. Find the maximum force value (Pm) and the corresponding force-line displacement value (Δm) (Fig.1). 2. Calculate Ln(P/Pm) and plot vs (Δ – Δm )/S for all data points beyond the maximum force point (Pm, Δm) (i.e. the data relevant to fracture propagation, see Fig.2). 3. Find the slope of the Ln(P/Pm) vs (Δ – Δm )/S curve corresponding to Ln(P/Pm) values between -0.51 and -1.21 (i.e. between the dashed lines in Fig.2). 4. Calculate CTOAc according to Eqn 5. The specified force range (i.e. corresponding to Ln(P/Pm) values between -0.51 and -1.21) spans the steady-state region ([12]; see Fig.2). Over this range, the crack traverses the central part of the specimen, i.e. 0.32 ≤ a/W ≤ 0.53 [19]; the force ratio Validation The CANMET recommended practice is intended for ductile-fracture propagation and is not valid for cleavage fracture. Cleavage fracture manifests itself as a sudden drop of force in the force vs displacement curve; CTOA values in cleavage fracture are very small. Report The values of CTOAc are reported to one decimal place. The report must contain a summary including, as a minimum: material and specimen ID, wall thickness, specimen orientation, test temperature, initial tup velocity, and fracture appearance (i.e. percent ductile fracture). A graph of the force vs displacement curves (see Fig.1) must be included in the report. Finite-element (FE) modelling FE modelling provides insight into the specimen response (stress distribution, through-thickness effects, temperature rise, etc.) and enables extrapolation into regions not accessible to the DWTT test (notably, 171 rib ut io n 3rd Quarter, 2013 higher strain rates and larger specimens). It is integral to the development of the procedure. or d is t Fig.6. Force-displacement curve and contour plots of damage for X-70 DWTT specimen. The flow stress of the matrix material, accounting for strain rate and temperature effects, is assumed to be given by: Sa m pl e co py -n ot f Recent progress using the modified Xue-Wierzbicki damage model [22] to simulate DWTTs is described ξ below. The material is assumed to be an elastic-plastic ε p T − TR n σ σ o (1 + C1ε p ) 1 + C2 1 − (9) material and the strength is governed by work hardening, = εo TM − TR strain rate, and temperature. Damage in the material is modelled using a damage-mechanics’ approach. Damage is assumed to be a scalar isotropic damage variable that where the first terms on the right model the is used to degrade the strength and elastic modulus quasi-static hardening curve, which is described by of the material. Accordingly, damage, is assumed to the yield strength σo, hardening constant C1, and evolve by a phenomenological evolution law given in the hardening exponent n. The second set of terms incremental form as: in parentheses model a logarithmic dependence of strength on strain rate, which is described by a constant C2 and the ratio of plastic strain rate to a reference (m −1) 1 εp 1 rate ε p / εo . During plastic straining, 90% of the plastic dD = m dε p (7) g(εp ) ε f εf work is dissipated as heat, which leads to softening; the terms in the final parentheses in Eqn 9 relate where is our modification for strain-rate effect, εp is the to model softening owing to the corresponding equivalent plastic strain, εf is the failure strain which temperature increase, which is assumed to depend on depends on the Lode angle and mean stress, and m temperature, T, reference temperature TR, and melting is an exponent. Function g is taken to be temperature TM. Linear thermal softening is assumed in our calculations, and the softening exponent is set to unity. εp g(εp )= 1 + δ log( ) εo (8) The model was implemented via a user-subroutine in Abaqus/Explicit finite-element software and computations to simulate the DWTT were performed. where ε p / εo is the ratio of the plastic strain rate to Element removal, when D = 1, in the element was a reference rate, and δ is a constant. The reference used to model crack propagation. Strength properties rate is usually chosen as the rate of the quasi-static of a typical X-70 pipe steel were used for the tensile test. Notice that for positive δ, the introduction matrix. By trial and error, the model parameters were of function g is equivalent to assuming a decrease in calibrated and some preliminary results are displayed in Fig.6. damage rate with increase in strain rate. 172 The Journal of Pipeline Engineering Summary References is t rib ut io n 1. W.A. Maxey, 1974. Fracture initiation, propagation, and arrest. Proc. 5th Symposium on Line Pipe Research, American Gas Association, Paper J, Arlington, AGA Catalog Number L301175. 2. R.J.Eiber, T.A.Bubenik, and W.A.Maxey, 1993. Fracture control technology for natural gas pipelines. NG-18 Report #208, PRC of AGA, Catalogue No. L51691. 3. B.N.Leis, R.J.Eiber, L.Carlson, and A.Gilroy-Scott, 1998. Relationship between apparent (total) Charpy vee-notch toughness and the corresponding dynamic crack-propagation resistance. Proc. 3th International Pipeline Conference (IPC 1998), ASME, pp 723731. 4. API Recommended Practice 5L3, 1996. Recommended practice for conducting drop-weight tear tests on line pipe, 3rd Edn. American Petroleum Institute, Washington, DC, USA 5. ASTM, 2008. ASTM E436-03: Standard test method for drop-weight tear tests of ferritic steels. ASTM International, West Conshohocken, PA, USA. 6. A.B.Rothwell, 2000. Fracture propagation control for gas pipelines – past, present and future. Proc. 3rd International Pipeline Conference, pp 387-405. 7. H.Makino, T.Kubo, T.Shiwaku, S.Endo, T.Inoue, Y.Kawaguchi, Y.Matsumoto, and S.Machida, 2001. Prediction for crack propagation and arrest of shear fracture in ultra-high pressure natural gas pipelines. ISIJ International, 41, pp 381-388. 8. G.Demofonti, G.Buzzichelli, S.Venzi, and M.Kanninen, 1995. Step by step procedure for the two specimen CTOA test. Proc. 2nd Pipeline Technology Conference: Pipeline Technology, Vol. II, Ed. R.Denys, pp 503-512. 9. M.Di Biagio, A.Fonzo, G.Mannucci, A.Meleddu, M.Murri, M.Tavassi,and L.N.Pussegoda, 2005. Ductile fracture propagation resistance for advanced pipeline designs. GRI Report No. GRI-04/0129, Gas Research Institute, March. 10. S.Xu, R.Bouchard, and W.R.Tyson, 2007. Simplified single-specimen method for evaluating CTOA. Eng. Fract. Mech., 74, pp 2459-2464. 11.S.Xu and W.R.Tyson, 2008. CTOA measurement of pipe steels using DWTT specimen. Proc.7th International Pipeline Conference (IPC 2008), ASME, IPC2008-64060. S.Xu, W.R.Tyson, and R.Bouchard, 2009. 12. Experimental validation of simplified single-specimen CTOA method for DWTT specimens. Proc. 12th International Conference on Fracture (ICF12), Paper T35.018. co py -n ot f A test procedure for measurement of CTOA for resistance to fast ductile fracture has been developed and demonstrated. It is being evaluated in an international round-robin project organized by CanmetMATERIALS and results to date (i.e. the results from three of the seven participants) showed good repeatability (within ±10% of the mean). The recommended CTOA test procedure is being studied by the relevant ASTM committee. On-going FE modelling has shown good agreement with experiment and excellent promise to provide further insight into the test and to extend investigations into regions of specimen size and loading conditions not accessible by laboratory tests. editor, Dr Xian-Kui Zhu, for helpful suggestions to improve the manuscript. We would also like to thank the CTOA round-robin participants and ASTM E08.07 for their contributions. or d In Fig.6, open symbols indicate the force and displacement corresponding to contour maps of damage. Post-peak, at a force of approximately 150 kN, the propagating fracture changes from initially flat (tunnelling) to slant, and subsequently, the slant fracture becomes fully developed. The CTOAc value and fracture front shape are in agreement with experimental observations of the DWTT specimen. Following the article by Xue and Wierzbicki [22], the adoption of Lode-angle dependence in the failure strain is the key to modelling slant fracture. To estimate CTOA in the computation, the mesh is sectioned at the mid-plane: the plane is such that the long dimension of the sample lies in it. Two vectors are drawn along the flanks of the crack and the angle between them estimated and taken as the CTOA. Preliminary estimates of the CTOA yield 15o and this compares favourably with the value of 12.5o obtained by carrying out the analysis described in the previous sections on the force-displacement curve, for this X-70 steel, by Xu and Tyson [11]. Details of our continuing modelling effort will be reported in a forthcoming publication. Sa m pl e The long-term objective of the work is to validate a CTOA-based method to predict the toughness required for crack arrest which has become problematic for highstrength steels, and to incorporate this approach in pipeline standards. This will be done by: standardization of crack-tip-opening angle (CTOA) measurement using the drop-weight tear test (DWTT) specimen; and review and development of crack-arrest toughness specification, by analysis of transferability of CTOA from small-scale specimen to full-scale pipe. Acknowledgement This work forms part of a CanmetMATERIALS project on fracture arrest toughness measurement and specification supported by the Federal Program on Energy Research and Development (PERD). The authors would like to thank their colleagues R.Bouchard, R.Eagleson, D.Y.Park, J.Liang, J.Sollen, and R.Guilbeault for their contribution to experimental work, and to the guest 3rd Quarter, 2013 173 is t rib ut io n 18.S.Xu, J.Sollen, J.Liang, R.Zavadil, and W.R.Tyson, 2012. Effects of notch type and loading rate on CTOA of an X65 pipe steel for CO2 pipeline. CanmetMATERIALS Report, CMAT-2012-01(TR). 19.S.Xu, R.Bouchard, and W.R.Tyson, 2004. Flow behaviour and ductile fracture toughness of a high toughness steel. Proc. 5th International Pipeline Conference (IPC 2004), IPC04-0192. 20. R.W.Nichols, F.M.Burdekin, A.Cowan, D.Elliot, and T.Ingham, 1969. Practical fracture mechanics for structural steel. Ed.: M.O.Dobson, United Kingdom Atomic Energy Authority. 21.British Standard Draft for Development D19, 1972. Methods for crack opening displacement (COD) testing. 22. L.Xue and T.Wierzbicki, 2009. Numerical simulation of fracture mode transition in ductile plates. Int. J. Solids and Structures, 46, pp 1423–1435. Sa m pl e co py -n ot f or d 13. S.Xu, W.R.Tyson, R.Eagleson, C.N.McCowan, E.D.Drexler, J.D.McColskey, and Ph.P.Darcis, 2010. Measurement of CTOA of pipe steels using MDCB and DWTT specimens. Proc.8th International Pipeline Conference (IPC 2010), ASME, IPC201031076. 14.S.Xu, R.Eagleson, W.R.Tyson, and D.-Y.Park, 2011. Crack tunnelling and crack tip opening angle in drop-weight tear test specimens. Int. J. Fracture, 172, pp 105-112. 15.S.Xu and W.R.Tyson, 2011. Recommended practice for determination of crack-tip opening angle of structural steels using DWTT specimens. CANMETMTL Report, 2011-03(TRR). 16. CSA Z245.1-07, 2007. Steel pipe. Canadian Standards Association. 17.BS 7448, 1997. Fracture mechanics toughness tests, part 2, method for determination of KIC, critical CTOD and critical J values of welds in metallic materials. British Standards Institution, London. e pl m Sa py co io n rib ut is t or d ot f -n 3rd Quarter, 2013 175 Drop-weight tear test application to natural gas pipeline fracture control io n by Dr Robert Eiber rib ut Robert J Eiber Consultant Inc., Columbus, OH, USA T ot f or d is t HE DROP-WEIGHT tear test (DWTT) has had a significant positive impact on the fracture properties of linepipe steels. This review summarizes the incidents that started the research leading to the development of the DWTT from 1960 to present. The initial driver for the development of the test was an incident that involved 8.3 miles (13.3 km) of brittle fracture during pre-service testing of a natural gas pipeline with gas. The initial goal of the DWTT was to accurately define the ductile-to-brittle transition temperature of pipeline steels to facilitate the specification of transition temperatures below the operating temperature range for linepipe. As the pipeline industry used the low-transition-temperature steels, the need for a measure of the steel toughness emerged to control ductile-fracture propagation arrest leading to examination of the DWTT energy as a substitute for the Charpy V-notch energy which had been identified as the way to define the steel fracture toughness. T Control of brittle fracture propagation Sa m pl e co py -n HE GOAL OF this paper is to review the need for, development of, and application of the drop-weight tear test (DWTT) for controlling fracture propagation in natural gas transmission pipelines. In the 1950 to 1960 time period, gas transmission pipelines were being installed that used 16 to 30-in (406 to 917-mm) diameter pipes. As the pipe producers were able to produce larger-diameter pipes and higher-strength steels, a problem occurred during pre-service testing of the pipes using natural gas as the pressurizing medium. The problem was the unstable propagation of brittle fractures that could propagate for hundreds of feet before arresting. In 1960, a new 36-in (914-mm) diameter line involving 56,000-psi (386-MPa) yield-strength steel was being constructed and, in moving gas from one section to another for gas testing, a fracture initiated at a shipping fatigue crack alongside the doublesubmerged-arc seam weld at a pressure of 63% SMYS (specified minimum yield strength). One to six fractures occurred running a total of 8.3 miles (13.3 km) before arresting. This test failure stimulated a need to solve the brittle-fracture problem and research was initiated to define the problem and develop a solution to achieve fracture control. Author’s contact details: tel: +1 614 538 0347 email: [email protected] Research was performed by the AGA (American Gas Association) Pipeline Research Committee at Battelle which indicated that – depending on the brittleness of the fracture surface – low-shear-area fractures could run at speeds of 1100 to 2300 fps (335 to 701 mps), and since the natural gas can only escape from the pipe at its acoustic velocity of about 1300 fps (396 mps), the pressure driving the fracture could not decay and was the initial line pressure. The fracture appearance of the high-speed fractures was 10-20% of the shear area. The initial experimental results indicated that if the minimum pipe temperature was at or above a shear area of 85% in the Charpy V-notch (CVN) impact test, the fracture would be assured of arresting in a pipeline incident. This indicated that the way to ensure brittle-fracture-propagation control in the event of an incident was to require the linepipe steels to have a transition temperature, as defined by the CVN impact test shear area of 85%, below the minimum operating temperature of the pipeline. The use of increasing wall thicknesses in pipelines lead to discovery that the 2/3 thickness Charpy V-notch impacttest specimen did not accurately predict the transition temperature of pipe thicknesses greater than 0.375 in (9.5 mm). This led to the development of a test specimen that used the full thickness of the pipe wall in order to get a 1:1 correlation of the laboratory test shear-area- The Journal of Pipeline Engineering io n 176 is t The DWTT test was standardized in ASTM E436 [1] and API 5L RP5L3 [2]. The full-scale test research indicated that this should solve the problem of fracture propagation control in natural gas transmission pipelines by making sure the pipe steel was ductile. Sa m pl e co py -n ot f • The fatigue pre-cracked Charpy specimen was examined but did not show an improved correlation over the standard mill-notched specimen, mainly because the specimen was a constant thickness for all pipe wall thicknesses. • The US Navy tear test was modified by pressing a sharp notch in place of the drilled hole. The test was slow and expensive and the correlation was poor because of the static loading. • The Robertson thermal-gradient test was found to exhibit crack-arrest temperatures approximately 30oF (17oC) below the full-scale crack-propagation transition temperature because there was not enough energy in the test specimen to drive the fracture in contrast to a fracture in a gas pipeline. • The US Naval Research Laboratory (NRL) dropweight test (ASTM E 2081) aimed at defining a nil-ductility transition (NDT) temperature was examined, but the NDT temperature was found to be approximately 60oF (34oC) below a pipe’s ductile-to-brittle transition temperature. The main concern was that the test did not measure the ductile-to-brittle transition but rather the temperature where ductility started to appear in a steel sample under a low strain rate. • The dynamic tear test (ASTM E 604) was examined as it was being developed by the NRL. One of its problems was that the specimen was a constant thickness; in discussions, NRL staff suggested the concept of the drop-weight tear test which was explored and developed. drop-weight tear test (DWTT) because falling weighted tups were used to break the specimen over a range of temperatures. Figure 1 shows a comparison of the CVN specimen with a DWTT specimen size and orientation in the pipe: the test specimens are selected to measure the properties in the longitudinal direction of the pipe as this is the direction that a fracture propagates. or d transition temperature with the pipe-transition temperature. There were a number of specimens examined in the development of the new test: rib ut Fig.1. Comparison of DWTT with a 2/3 thickness CVN specimen (photo courtesy of PRC International). A 3 x 12-in (76 x 305-mm) full-wall-thickness specimen was found to accurately predict the pipe-transition temperature regardless of the wall thickness. The test is called the Figure 2 presents the dimensions of the DWTT specimen and support for the impact test. The notch is pressed into the test sample using tool steel machined with a 45o angle and sharpened to a fine point. The resulting notch radius has some variability but generally is 0.001 in (0.025 mm), and the velocity of the hammer at impact is to be greater than 16 ft/sec (4.88 mps). The energy available in the hammer must be adequate to break the specimen into two pieces. As the toughness of steels has increased, the available energy of pendulum testing machines has had to keep increasing. The energy available to break a specimen should be at least 25% greater than the energy absorbed by the specimen, but this has not been examined in great detail. Figure 3 shows a series of DWTT specimens tested over a range of temperatures (shown in Fahrenheit) in the figure. The significant feature of the fracture surfaces is that even at the highest temperatures the fracture was initiated as a brittle fracture shown by the light grey fracture surface. The pressed notch cold works the steel at the notch, resulting in an embrittled region where the fracture initiates as a brittle fracture. Figure 4 compares the results of a 2/3-thickness CVN specimen and a full-thickness DWTT specimen from a 30in (762-mm) diameter 0.375-in (9.5-mm) wall thickness 177 io n 3rd Quarter, 2013 ot f or d is t rib ut Fig.2. DWTT specimen and support dimensions (figure courtesy of PRC International [3]). py -n Fig.3. DWTT specimens tested over a range of temperatures (Note: pressed notch is at the bottom of the fracture) (courtesy of PRC International [3]). co X-52 pipe. The DWTT results exhibit an abruptness in the transition from brittle to ductile behaviour over a narrow temperature range as contrasted to the CVN results. Sa m pl e The question was which test represented the fracturesurface appearance in pipes pressurized with natural gas. To resolve this, four full-scale fracture tests were conducted over a range of temperatures on a single length of 30-in diameter by 0.375-in wall thickness X-52 pipe as indicated by the four solid data points in Fig.5. The fractures were initiated in the pipes at temperatures from 27o to 107oF (-2.7o to 42oC), and the calculated stress levels varied from 73 to 78% SMYS. It can be observed that the DWTT reproduces the abrupt ductile-to-brittle transition region displayed by the fractures in the full-scale pipe fracture tests. Figure 6 presents the appearance of fractures in pipe tests (right side photos) that were approximately 15-ft (4.5-m) long pipe specimens pressurized with 6 to 10% nitrogen. The sections through the fractures (left side photos) show the difference in thickness reduction with increasing ductility of the fracture. To evaluate the effect of pipe thickness on the transition temperature, a 0.750-inh (19-mm) steel plate had DWTT specimens prepared from it in a 0.750-in (19-mm) thickness as well as 0.5-in (12.5-mm) and 0.375-in (9.5-mm) thicknesses. The reduced-thickness specimens were all machined from the mid-thickness of the plate to keep the microstructure as constant as possible. In addition, 2/3 thickness CVN specimens were also prepared from the mid-thickness of the plate for comparison. The results, presented in Fig.7, show that as the DWTT thickness decreased the transition temperature identified by the DWTT specimen also decreased. The CVN curve was the same for all three DWTT thicknesses since the specimen was a constant thickness. The transition temperature shift from 0.375 in (9.5 mm) to 0.50 in (12.5mm) was 12oF (7oC), and from 0.50 in (12.5 mm) to 0.75 in (19 mm) was 16oF (9oC), indicating that the transition temperature shift is not linear with thickness increases. The temperature in the DWTT at 85% shear area was selected as the indicator of a pipe’s transition from ductile to brittle behaviour. The question that occurred The Journal of Pipeline Engineering is t rib ut io n 178 e co py -n ot f or d Fig.4. Comparison of DWTT and CVN test results on 0.375-in wall thickness X-52 pipe (figure courtesy of PRC International [3]). m pl was “Is this all that is necessary for the specification of pipe that would control fracture propagation and avoid long fractures in service?” The concern was whether it was also necessary to specify a toughness level. Sa To evaluate this concern, several full-scale fracture tests were conducted which involved 673 ft (205 m) of pipe with a 150-ft (46-m) test section in the centre in which a fracture was initiated with an explosive linear cutter. The tests were conducted at stress levels from 70 to 72% SMYS on X-52 and X-60 pipe, and in the tests the fractures were initiated above the DWTT transition temperature. The fractures all arrested quickly in each direction from the origin, and this appeared to confirm that all that was necessary was to order pipe with a DWTT transition temperature below the lowest operating temperature. Fig.5. Comparison of full-scale fracture appearance versus DWTT and 2/3 CVN (figure courtesy of PRC International [3]). At this time, the pipe manufacturers started to produce pipe with low transition temperatures regardless of the CVN energy level. Control of ductile fracture propagation In December, 1968, there was a service incident involving an 850-ft long fracture in a new 36-in (914mm) diameter by 0.375-in (9.5-mm) wall thickness X-65 pipe pressurized with natural gas. The fractures in all pipe lengths were 100% shear. The obvious question was why the fractures had not arrested as they had in prior full-scale burst tests. The one piece of evidence that pointed to the problem was that the pipes through which the fractures propagated had relatively low 2/3 CVN plateau energy levels of 11 to 179 is t rib ut io n 3rd Quarter, 2013 py -n ot f or d Fig.6. Photos of fractures in the pipe-fracture tests in Fig.5 (photos courtesy of PRC International [3]). e co Fig.7. Effect of thickness on transition temperature based on the DWTT (figure courtesy of PRC International [3]). Sa m pl 13 ft-lbs1 (15 to 18J), and the two end or arrest pipes had 14 and 22 ft-lb (19 to 30J) 2/3 CVN plateau energy levels. The 2/3 CVN plateau energy of the origin pipe length was 18 ft-lbs (24J) which may have been high enough to have arrested the fracture, but the pressure driving the fracture in the origin length was higher than after steady-state fracture propagation was established which is the reason for the propagation of the fracture through the origin pipe length. Research was again initiated by the AGA Pipeline Research Committee at Battelle to develop a means of controlling long-ductile-fracture propagation. The CVN energies at the failure temperature were on the plateau and thus the same as the values quoted. 1 Full-scale burst tests were conducted using 36-in (914-mm) diameter pipe similar to that involved in the service failure, which lead to the development of a model that has been called the two-curve model (TCM). Figure 8 presents the results of the TCM for 36-in (914-mm) diameter by 0.375-in (9.5-mm) wall thickness X-70 pipe at 67% SMYS. The predicted CVN arrest toughness is 43 ft-lbs (58J) determined when the two curves are tangent, which is the borderline condition for arrest. Higher-toughness pipes (i.e. 53 and 63 ft-lbs [72 and 85J]) do not intersect the gas-decompression curve, indicating arrest, while lower-toughness curves (i.e. 33 ft-lbs [45J]) intersect the gas-decompression curve, indicating the fracture would propagate at the intersection speed (i.e. 575 fps [175 mps]) for a considerable distance. The Journal of Pipeline Engineering Fig.8.Two-curve model prediction of arrest toughness. Sa m pl e co py -n ot f or d is t rib ut io n 180 With the knowledge that a given CVN energy level was necessary for the arrest of a ductile fracture, it became apparent that not only was it necessary for the pipe to be operating above its DWTT transition temperature, but also that it needed to have sufficient CVN impact energy to arrest a ductile fracture. Thus, the use of both the DWTT transition temperature and the CVN energy/toughness in a pipeline design can control brittle and ductile fracture propagation, which provided fracturepropagation control for the vintage pipeline steels. Fig.9. Comparison of the three correction factors. Recent observations with newer high-toughness steels The pipe manufacturers armed with the knowledge that CVN plateau energy was important began to produce high-toughness steels with low transition temperatures. These high-toughness steels were achieved by removing impurities of carbon and sulphur, with the result that controlled-rolled or thermo-mechanically-processed (TMP), steels can routinely achieve at least 200 ft-lb 3rd Quarter, 2013 181 io n rib ut CVNTCM-Leis = CVNTCM + 0.002 CVNTCM2.04 – 21.18(1) where CVNTCM J is the Charpy energy predicted for arrest using the TCM. For X-80 steels, the Leis equation has been modified as shown in Equn 2 [4]. The terms are the same as in Equn 1, and the equation only applies to CVN energies greater than 70 ft-lbs (95J). ot f • Pre-compressed regions in the area of the notch to embrittle the steel. Unfortunately, this did not solve the problem. • TIG (tungsten inert gas) spot welds which embrittled the area under the notch, but which did not solve the problem. • Brittle weld beads (similar to the brittle weld bead in a drop-weight test, ASTM E 208) across the notch location with a machined notch to initiate the fracture. This was effective in initiating a brittle fracture most of the time but the behaviour was erratic and occasionally the complete weld would pop out of the specimen during impact testing, and the idea was therefore discarded. • Fatigue-cracked pressed notch involved a pressednotch specimen being subjected to a cyclic bending load to fatigue crack the notch. The fatiguing, because of the residual stress field from specimen flattening and the notch pressing, did not result in a uniform depth fatigue crack and therefore was not acceptable. • Chevron notch was machined into the specimen with a depth of 0.2 in (5 mm) at mid-thickness with a notch included angle of 90o through the thickness. The problem that was evident from examination of the fracture surface was that reducing the thickness of the notch tended to introduce more ductility into the steel, which was the wrong direction to go. • Pre-cracked pressed notch involves statically loading a standard PN (pressed-notch) specimen in the normal testing fixture used to hold the specimen for impact testing until the static bending load develops a 0.2-in (5-mm) crack depth below the tip of the pressed notch which usually develops after a load decrease of 1.5% after the maximum load has been reached. Another problem that has been encountered with the new high-toughness high-strength linepipe steels is that the TCM-predicted CVN energy level has been found to under-predict the toughness of some of the newer TMP steels (ranging from X-70 to X-100). As a result, various factors to correct the TCM prediction of arrest CVN energy have been proposed [4]. The correction factor developed by Leis for X-704 is represented in Equn 1. (Note: The Leis correction equation is only applicable if the measured CVN toughness is greater than 70 ft-lbs (95J).) is t One of the problems was that it became more difficult to initiate a brittle fracture from the notch that would propagate all the way across the specimen. This led to examination of other notch types to see if an improved notch could be found that would reduce the energy required to initiate a fracture and introduce a brittle fracture. The notch types examined were: a brittle crack, the PN-DWTT specimens could be pre-cracked and then tested without having to make new test specimens. or d (271J) CVN energy levels on the upper plateau. Also, the steels are generally thicker because of the trend toward higher operating pressures in gas transmission pipelines. However, the combination of the higher toughness and increased thickness caused problems in using the DWTT for their evaluation. Sa m pl e co py -n CVNTCM-Leis = CVNTCM + 0.003 CVNTCM2.04 – 21.18(2) The most promising notch revision was the pre-cracked pressed notch. One advantage was that if the initial testing indicated a problem with the initiation of The third correction factor is 1.7 times the CVN TCM prediction as proposed by CSM [4] and shown in Equn 3. Factors have also been developed for X-80 and above, as indicated in Ref. 4. CVNTCM CSM = 1.7 CVNTCM(3) Figure 9 shows the trends of the three correction factors. When compared to the burst-test results, the Leis X-70 and X-80 correction factors generally split the non-conservative propagate points and the conservative arrest data points from full-scale burst tests. The CSM X-70 correction is somewhat more conservative than the Leis X-70. The trend is that as the pipe grade (yield strength) increases, the magnitude of the factor also needs to increase. The reader should examine the individual grade correlations developed in Ref.4, which appear to be grade-dependent as well as possibly mill-dependent. It would be nice if a single correlation could be developed for a test such as the DWTT that could be used to predict the transition temperature and the ductile-fracture arrest energy. Additional development 182 The Journal of Pipeline Engineering The DWTT has been a major step forward in solving pipeline fracture-control problems considering where the industry was in the 1950s and 1960s. The DWTT is used by the pipeline industry internationally for specification of transition temperatures. io n 1. American Society for Testing and Materials. Book of Standards, Part 10. 2. American Petroleum Institute. Specification for Line Pipe, Recommended Practice 5L3. 3. R.J.Eiber, T.A.Bubenik, and W.A.Maxey, 1993. Fracture control technology for natural gas pipelines. AGA NG-18 Report 208, Catalog L51691, December. 4. R.J.Eiber, 2008. Fracture propagation 1, Oil & Gas Journal, Fracture Propagation, 20 Oct; and Fracture Propagation – Conclusion. Idem, 27 Oct. Sa m pl e co py -n ot f There are still problems occurring with the use of the test on high-toughness heavy-wall-thickness steels. The potential usage of the DWTT test energy as a substitute for the CVN impact energy would be ideal as this would eliminate the need for two tests to be conducted in the pipe mill. At the moment, the industry needs to conduct full-scale burst tests to validate fracture-control plans for the newer high-strength and high-toughness grades of steel. References rib ut Conclusions Which of the parameters CVN or DWTT should be used, and how to define an arrest toughness for describing the ductile-fracture propagation behaviour for extremely high-toughness steels in Grade X-100 or above, is still a challenge. The main conclusion is that both a transition temperature and a toughness level (impactenergy level) are required to achieve fracture control. is t As far as is known, no company has tried to order pipe to a DWTT energy criterion as of this date. One potential problem is that the measurement of the DWTT energy has not been standardized. In order to implement a DWTT energy requirement, this will be a necessity. One of the developments in the recent years is that with increasing fracture toughness of the pipe steel, there has been a tendency to specify only the CVN energy necessary for ductile fracture arrest. This is a mistake, and can lead to undesirable consequences, as the transition temperatures of some of the newer steels have tended to increase if a DWTT-based transition temperature is not specified. or d is needed before we can reach this point. One issue is that the DWTT energy measurement has not been correlated with ductile-fracture arrest, nor has a machine design been proposed that will produce reproducible energy values. 3rd Quarter, 2013 183 The Charpy impact test and its applications by Dr Brian N Leis io n B N Leis, Consultant, Inc, Worthington OH, USA T ot f or d is t rib ut HIS PAPER REVIEWS the Charpy V-notch (CVN) impact test and assesses its utility to characterize fracture resistance in applications to modern tough materials in contrast to those encountered prior to the availability of such materials.The origin of the CVN test and its development into a standard for use with metallic materials is discussed, with brief reference also made to application-based standards for use with other engineering materials. Thereafter, the evolution of mechanical and other properties motivated by industry demands is illustrated in regard to strength and toughness.The interpretation of the CVN test in regard to (1) the force-displacement and compliance response that develops during the test, and (2) factors affecting the energy measured and controlling failure of the CVN specimen, are discussed, including the tup design and the use of sub-size specimens. The utility of CVN testing is illustrated and discussed in the context of pipeline and other applications involving tough steels. Finally, the implications of evolution in material properties is assessed for impact-test practices including ASTM E23 and ISO 148-1, which are specific to the CVN practice, and the drop-weight tear test. It is concluded that where tough materials are involved, alternative testing practices are needed that are better adapted to the specific loading and failure response of the structure of interest. A more-resistant ones would only bend, such that a notch was introduced to facilitate initiation circa the 1890s [3]. The use of a pendulum-based impact test – whose use permitted determination of the energy absorbed in failure – is first evident in Russell’s work circa the mid 1890s [4]. Like the earlier drop-weight machines, his practice made use of un-notched samples. Soon after the turn of the century, Charpy [5] improved Russell’s method by introducing a redesigned pendulum, a notched sample, and more-precise specifications, thereby moving the impact test toward a more-standardized practice. Charpy’s name appears to be associated with impact testing more because of his efforts to standardize the practice over the ensuing years than because of his role in the development of the hardware. Sa m pl e co py -n WEBCRAWL ON the topic of impact testing and related standards develops more than 850,000 hits which, after allowing for duplicated hits and/or otherwise similar publications, makes clear that this topic has a broad following. Consideration of the dates of the citations found in such a search range from current times, with the history uncovered in this process indicated [1] to trace back to the min-1800s. Accordingly, impact testing also reflects an enduring concern that addresses the integrity of structures. Because the origins of impact testing have recently marked a century of development, reviews have been written that document this history (Ref.1, for example), which note the early work was motivated by a need to qualify materials resistant to impulse loadings. It is apparent that the early work centred on military and railroad applications. The need to screen materials for their resistance to impulse loadings led to the development of the first dropweight machine, which emerged circa 1860 [2], although it was used then in testing smooth rectangular bars (i.e. without a notch or starter crack). While that practice was found adequate for the less-resistant materials, the Author’s contact details: email: [email protected] Recognizing that impact testing was first developed to screen ferrous structural materials in the context of what has since been termed brittle-fracture resistance, a host of variations on the initial impact-test practice evolved to quantify resistance, as did other practices designed to understand the failure mechanism. The documents evaluated indicate that variations developed in the test practices related to the nature of the imposed loading and test-specimen geometry, as well as in the notch shapes and the use of localized treatments to affect desired changes in the nature of the failure process. As the significance of strain rate became evident, practices were introduced to vary the speed of the impact, with The Journal of Pipeline Engineering is t rib ut io n 184 (a) (b) (for example, Refs 10 and 11). Some factors known to affect variability include the dynamic stiffness of the system, differences in the tup geometry (radius and overall width and angle to the striking edge), and the use of sub-size (reduced-thickness) specimens, with the nature of the material being characterized also being a major consideration in regard to the energy required to fail the specimen, and steel cleanliness. Suffice it here to note that the significance of some of these factors is considered later in more detail. pl e co py -n ot f slow-bend practices introduced in complement to the impact machines. In a related vein, it became evident that sufficient energy was required to continue to drive the failure, such that ‘excess energy’ became a consideration. In this context, it has been noted [6] that “more than 50 types or variations of tests can be clearly identified for evaluating the susceptibility of materials to brittle fracture” that include well-known through now-obscure practices. Examples of some well-known tests aside from the Charpy V-notch (CVN) practice include the Izod test, the explosion bulge test, the Pellini crack-starter test, the NRL-NDT drop-weight test, and the Battelle drop-weight tear test (DWTT), while some lesser-known practices include the Kahn tear test, the Bagsar cleavage-tear test, and the Charpy keyhole and U-notch tests. These and other practices can be found with more details in the citations listed in Reference 6, beginning at page 316. While many such impact tests exist, the focus hereafter is on the CVN test. or d Fig.1.Trends in CVN energy over time, adapted from Refs 18 and 19: (a) energy trends with grade, circa 1969; (b) historical view of plateau energy and grade. Sa m CVN test development and applications-based standards The CVN test is a simple, low-cost, and reliable test method that today is commonly required by regulatory and other codes for fracture-critical structures such as pipelines, bridges, and pressure vessels. Today, three standards are prominent in the general use of the CVN test [7-9] that are relevant for applications to ‘metallic materials’, and provide a viable basis for the consistent use of this practice. In contrast, the history of the CVN practice from about 1900 through to 1960 is marked in regard to data scatter and variability, more generally in regard to the factors that adversely affect its outcome Realizing that it took until the 1960s before the CVN test transitioned from a qualitative to a practically useful quantitative method, issues due to factors such as those just cited persisted long after the CVN practice became a tentative standard of the American Society for Testing and Materials (ASTM) in 1933. Then designated E 23-33T1, this tentative practice somewhat loosely covered two notched-bar impact practices that differed significantly in regard to the methods to support the test specimen (i.e. the CVN test involving three-point beam-bending and the Izod involving cantilevered beam-bending). The tentative standard also did not adequately control the striking edge or ‘tup’ for either test practice, and it allowed flexibility in regard to the notch geometry. Adequately managing such differences was a key to move the CVN practice from one viewed as qualitative, in part due to its scattered results, to one that is quantitative and so can be practically useful aside from screening materials for fracture mode. For those less familiar with the ASTM process, the tentative designation is no longer part of its standards’-development process. 1 3rd Quarter, 2013 185 Evolution in properties: implications for ASTM E23 or ISO 148-1 io n Evolution of materials and properties is t rib ut While much has been done to ‘standardize’ and ‘specify’ the essential details of the CVN test practice over the years, the metallic materials being developed for use today differ significantly from those in the era that the CVN test transitioned from qualitative into what is considered today a quantitative test. Figure 1, which adapts trends from Refs 18 and 19, serves to illustrate and emphasize this point. Figure 1a summarizes trends in CVN energy for data developed up through the late 1960s. The y-axis in the figure is CVN energy, which is shown as a function of test temperature on the x-axis. To facilitate comparing trends for which differences exist in the transition temperature, this figure centres them at close to the 50%SA temperature, such that this axis is labelled ‘relative temperature’. As is evident from the figure, these trends are presented only in the units reported, because the format of the plot precludes presenting dual units. Results are shown for several steels whose yield stresses, denoted sy, range from 40 to 200 ksi (276 to 1378 MPa). These trends are shown as the dash-double-dot lines. Results also are presented for two aluminium alloys with yield stresses of 38 and 75 ksi (261 to 517 MPa), which are shown as dashed trends toward the lower edge of this figure. py -n ot f Thus, although today the ASTM E23 standard still covers these quite different bend-bar impact tests, and also still allows different notching practices, it now makes clear the differences between these tests and the notching practices, and addressed those differences through the requirements of the standard. Unfortunately, differences remain today between the geometry of the ‘tup’ in what might be considered a North American [7] and European standards [8], which can open to differing outcomes particularly when evaluating more resistant materials. because the force versus load-point displacement record from such set-ups helps understand the extent to which the substituted material behaves differently from the material it is replacing. or d Over the years, aspects unique to the specimen geometries and loadings have been identified and understood such that today the standard embeds clear requirements for the test specimens, as well as for the testing procedures and the reporting. Significantly the reporting requirements prescribe clear identification of the type of test and specimen, as well as other details that could promote what otherwise lead to what could be misconstrued as data scatter. Annexes to the standard have been developed that address the test machines, including verification of the Charpy impact machine. These annexes also outline optional specimen configurations, and discuss CVN pre-cracking, specimen orientation, determination of the percent of shear fracture area (%SA), and methods to measure the centre of strike. Finally, the significance of notched-bar impact testing is considered in regard to the correlation between the test practice and its outcomes with the service situation. There it notes that the Charpy or Izod tests may not directly predict the ductile or brittle behaviour of steel – an aspect that is elaborated later in the section that illustrates the application of the CVN test to specifying linepipe steels. Sa m pl e co In addition to the above-noted standards that focus on metallic materials, such as steel, other standards have evolved based on the basic tenets of ASTM E23 that focus on miniaturized testing and address materials other than the types of materials that were targeted by the original standard (for example, 12-16). Such standards have evolved to address the unique concerns of polymers and cast iron, for example, and are developed to meet the needs specific to the application involved. Needs for an application-specific alternative to E23 can become apparent as new materials emerge that will find use under impulse-loading conditions, or under circumstances that otherwise could promote a change in fracture resistance due, for example, to cold-temperature service, or irradiation. Materials substitution also can create applications wherein an existing material is replaced by an emerging material, which needs to be screened for changes in its fracture resistance over the range of operational circumstances that are unique to its adapted use. Note that, in this context, the use of an instrumented testing system, consistent with standards [17], can facilitate the use of impact testing in cases involving material substitution or new applications concerned with impulse loads. This is It is evident from Fig.1a in regard to the data shown for the steels that these CVN trends up through the plateau (upper-shelf) energy, denoted CVP, indicate that the resistance to fracture (or toughness as characterized by CVP) decreases as the yield stress increases, and that this decrease is significant. This same trend is also evident for the aluminium alloys. While the trends are similar for the steels with sy = 200 ksi (1378 MPa), it is apparent that the energy dissipated in a CVN test is not uniquely quantified relative to the grade of steel. Rather, experience indicates that the energy dissipated reflects the chemistry and processing history of the steel on a case-by-case basis. Finally, these data indicate that for steels produced up through the late 1960s a five-fold increase in yield stress (grade) could affect more than a ten-fold reduction in toughness. This tendency reflects the well known tradeoff between strength and ductility that was characteristic of the carbon-manganese (C-Mn) steels produced prior to the advent of the high-strength low-alloy (HSLA) grades that emerged commercially in the 1960s. The Journal of Pipeline Engineering (b) X-100 specimen @ 118 ft-lb (160J) py -n ot f or d is t (a) X-70 specimen @ 250 ft-lb (339 J) rib ut io n 186 co (c) fracture features @ 250 ft-lb (339 J) (d) fracture features @ 18 ft-lb (~24 J) Fig.2. Differences in CVN response with toughness, adapted from Refs 22 and 23. Sa m pl e HSLA steels have evolved over time, coupling chemistry and processing changes to achieve strength while limiting the frequency and size of particles that nucleate voids and trigger the onset of plastic collapse, through the use of cleaner steels and sulphide shape control [20]. Rolling practices also evolved through use of controlrolled and thermal-mechanically controlled process (TMCP) techniques that coupled controlled rolling with accelerated cooling, which subsequently gave way to high-temperature-processed (HTP) steels. Thus, since the mid-1960s, it has become possible to couple strength gains with improved toughness and ductility to enhance fracture resistance. Such changes have developed in the context of automotive steels and linepipe steels, for which weldability also was a major consideration. Thus, in contrast to Fig.1a, the trends in part (b) of this figure show a reversal of the decrease in ductility with increasing grade, with this shift apparent in Fig.1b circa more or less the mid 1960s. Note that in Fig.1b that the data points flagged as Q&T (quenched and tempered) that lie toward the bottom of the interval for the mid-1960s lie at quite low toughness values, which is as expected for higher-strength steels based on the results evident in Fig.1a. Figure 1b, which presents data for linepipe steels produced up through circa 2000, indicates that CVP energies on the order of 250 ft-lb (339 J) were being achieved toward the turn of the millennium. But, in comparison with such gains over about a 30-year period, it is not unusual currently to see energy values approaching double that level in grades up to X-80 (551 MPa) – all within the last 10 or so years. It is also apparent that for grades above X-80, as for example X-120, which are produced by approaches that differ from those used to make modern X-80, it is much more difficult to achieve 3rd Quarter, 2013 187 io n rib ut Figures 2c and 2d contrast the fracture features for results such as those shown in Figs 2a and 2b, respectively. While the view in Fig.2d shows evidence of crack initiation and propagation, as has typically been observed and expected in the context of the interpretation of the CVN test as evident for example in ASTM E23, the image in Fig.2c for the much-tougher steel does not show the same traits. Rather, it shows significant local stretching at the notch, very large lateral contraction below the notch, with thickening occurring toward the back face, and evidence of some stable tearing and stretching occurring in lieu of what might be called cracking. ot f It follows from the significant increase in toughness evident in Fig.1b, for steels produced recently in contrast to those of late 1960s, that for materials where high CVN energies have been achieved the failure response of a CVN specimen could differ radically from that typically seen in the late 1960s. The results in Fig.2, which are adapted in part from Ref.21 and from unpublished work [22], clearly support this expectation. The images in Figs 2a and 2b are post-test perspective views of full-size CVN specimens that both showed 100% SA, following testing in a 512 ft-lb capacity pendulum machine equipped with unmodified anvils and a Dynatup striker. The machine used had been certified by the National Institute of Standards and Technology (NIST) practice [23], and maintained in calibration. While clearly apparent from the back-face deformation for the specimen shown as Fig.2a, but less evident for that shown in Fig.2b, during the test both specimens had been wedged onto and around the tup as they were pushed through the machine. Thus, these specimens were intact after testing – although to a much different extent for what was the higher toughness case as compared to that requiring much less energy to push it through the anvils – as becomes clear, as follows. is t Implications for standardized impact testing face and flanks of the tup, with the remainder of the specimen otherwise largely un-deformed, although the ligament remains intact. Nevertheless, close examination of the image reveals the imprint of the tup shape on the back face. As such, this specimen rode the tup through the anvils, in like manner to that for the higher-toughness steel. And while a crack did develop, which was quite deep, manually closing the angle between the back faces after the test did not break the specimen into two pieces. A later section further considers that and other aspects that appear to limit the utility of the CVN test in applications to highertoughness materials. or d the high toughness levels necessary to fully capitalize on the strength gained for the applications for which such grades were designed. Sa m pl e co py -n The image in Fig.2a reflects a measured energy of about 250 ft-lb (339 J) for a CVN specimen made of X-70 steel produced circa the mid-1990s. This image focuses on the area of a specimen in the vicinity of the notch and fracture, to illustrate the complex nature of tearing and the significant lateral and local back-face deformation. As noted in the above discussion, this image clearly shows that the back-face of the specimen has deformed around and about the face and flanks of a tup whose size and shape conformed to ASTM E23, whereby the striker radius is 8 mm. Had the narrower, smaller-radius (2-mm) tup of ISO 148-1 been used, less wedge force would have developed, such that less energy would have been required to push the specimen through the anvils. In this context, the total measured energy involves the energy to deform the specimen, to initiate cracking or tearing/stretching, and to propagate that cracking or tearing/stretching through the depth of the specimen, along with the energy required to push the specimen through the anvils. Further on this aspect follows later in the section that deals with factors affecting the measured energy and controlling failure of a CVN specimen. Suffice it here to conclude that such observations indicate that the test in applications to higher-toughness materials no longer develops features typical of those observed throughout its historic applications, with a subsequent section further considering factors affecting the measured energy and controlling failure of CVN specimen, which closes considering the implications in regard to the utility of the CVN test as standardized by ASTM E23 [7] or ISO 148 [8]. Figure 2b reflects a measured CVN energy of about 118 ft-lb (160 J) for a specimen made of X100 steel, which was produced just into the new millennium. This image is in strong contrast to that in Fig.2a, as only limited local distortion has occurred around the The CVN test: its interpretation and utility The CVN test can be implemented using either a drop-weight system, which requires an instrumented tup, or a pendulum set-up than can make use of an instrumented tup, but does not require it to determine the dissipated energy. While an instrumented set-up is not required using the a pendulum, the results of measured force as a function of the load-point displacement do provide much insight into the dissipative processes that contribute to the energy required to fail the specimen. Output from an instrumented test is shown in Fig.3 as the basis for further discussion on the test and its interpretation, based on results reported in Ref.24. The results in this figure reflect transverse CVN samples cut transverse to the longitudinal axis of a joint of linepipe, with the notch oriented through the thickness of the pipe. This orientation corresponds to testing in the y-x direction as identified in ASTM E23. The Journal of Pipeline Engineering rib ut io n 188 (b) is t (a) Force-displacement and compliance response bending and tearing coupled with stretching in tougher steels, or cracking in less-tough steels – the sample moves on through un-modified anvils. While less-tough steels fail as two pieces that can be forcefully ejected from the machine, as indicated in Fig.2a, tougher steels can be found still wedged around the tup. In extreme cases, specimens can bend around the tup to an extent that unloads the face of the striker – which is also discussed in a later section. Because the format of this figure precludes the use of dual units, the results are presented using in the units used to quantify this response. Sa m pl e co py -n ot f Force as presented on the y-axis of Fig.3 is as measured by a Dynatup striker (as made circa the early 1990s) as a function of the load-point displacement (determined from pendulum arm rotation) that is shown on the x-axis. The secondary y-axis is a measure of the specimen’s compliance, as it presents the x-axis displacement value divided by the corresponding load. Force – as shown on the y-axis of Fig.3 – has been divided by the initial net-section area of the CVN specimen, with that three-point bending net-section stress then divided by the average value of the ultimate tensile stress (UTS) for the steel involved. The value of the UTS used is based on companion round-bar tension tests made from flattened blanks, which also were cut in the transverse (hoop) direction, and tested in accordance with ASTM E8 [25]. On this basis, the value of the y-axis should normalize the results for this steel and others up through either the limit load or fracture, with the outcome shown on the y-axis being a multiple of the ratio of the instantaneous load relative to the limit-load that depends on the bending stiffness of the CVN sample. or d Fig.3.Typical response trends developed in instrumented CVN testing: (a) moderate to higher toughness; (b) a range from low to high toughness. Figures 3a and 3b present results that show the specimen’s initial elastic response where, after yielding occurs, plastic bending ensues as the load increases through the limit load, beyond which the load falls off as it would in a tensile test as necking develops. The analogue to necking for tough steels leads to localized thinning across the notch face of the CVN specimen, with the response under quasi-static loading similar to that under impact loading evident – as presented in a later section. Lateral thinning along the notch is accompanied by back-surface expansion, during which – after sufficient The plot shown in Fig.3a presents results for a controlrolled X-80 steel (labelled #3) cut from mill-expanded line pipe with 0.61-in (15.5-mm) thick wall that was produced in 1992. Figure 3b supplements this result with data for an accelerated-cooled X-70 steel from mill-expanded linepipe with 0.75in (19.1-mm) thick wall that also was produced in 1992 (labelled #2), a control-rolled X-60 steel from mill-expanded linepipe with 0.546-in (13.9-mm) thick wall that was produced in 1970 (labelled #7), and a conventionally-rolled X-52 steel from mill-expanded linepipe with 0.375-in (9.5-mm) thick wall that was produced in 1960 (labelled #8). All involved testing of full-size samples except for the X-52, which made use of a 2/3 thickness specimens. All results reflect tests that developed 100 %SA, such that these trends reflect fully ductile response. In reference to Fig.3a, a series of dotted vertical lines have been added to the test record, which have been located to tie key changes in compliance to events evident in the force-displacement response. Note in this context that if the stress-strain response is mapped onto this coordinate system, the locus traced (assuming the 3rd Quarter, 2013 189 is t rib ut io n In contrast, the force-displacement response developed by steel #2 is comparable to that shown by steel #3 in regard to Fig.3a, except that this steel does not show evidence of a significant reduction in the force required to move the specimen through the anvils over the course of the testing. As for the other steels, analysis indicates that the peak load for this case correspond closely with what is inferred as the limit load for simple bending relative to this steel’s UTS. And as for the other steels, it is evident that the force-displacement response for this steel is normalized as expected by the use of the ordinate noted above, with the difference in response up to the limit load reflecting its unique flow response and strainhardening rate. However, in contrast to the results for steel #3, which shows the force-displacement response in the CVN test drops below the locus of the stress-strain response of the steel when mapped on to this figure, that trend for steel #2 remains at or very close to the locus of the stress-strain curve. As such, what in 1997 was partitioned as largely deformation energy with a smaller initiation component (for example, see Ref.28) would more correctly be represented as all deformation energy. ot f The end of the regime associated with ‘initiation’ is marked by the second vertical line from the left side of the plot, which lies at the onset of what is identified in Fig.3a by the label ‘propagation’ – that for this steel occurred ductile tearing. Beyond the area labelled propagation, the response for this specimen becomes complex in regard to the force-displacement response and the change in compliance. Fractography and observations made using a quasi-static set-up infer that the tearing quickly transitions into rotation and stretching. While this transition is ill-defined by the dynamic force-displacement behaviour and related change in compliance, the quasi-static results indicate the transition first involves the back-face of the specimen wrapping around the tup, after which the associated wedging load causes failure while the specimen passes through the anvils, and eventually breaks in two. Accordingly, the response shown in Fig.3a beyond that labelled propagation is divided into regions labelled deformation (as the specimen wraps around the tup) and propagation (as it fails between the anvils). ordinate discussed earlier, with the differences in trends up to the limit load reflecting the inherent differences in their flow response and hardening rates. or d specimen geometry remained as it began) would begin to drop at the same displacement that marks the peak load, which is marked by the first of the vertical lines. The locus tracked by the stress-strain response depends on the actual strain-hardening behaviour of the steel, and for this test tracks at a higher loads beyond the peak load. It follows that, for this test, the peak load evident in Fig.3a corresponds to the bending limit load, and that the response up through the limit load reflects energy dissipated in deforming the specimen. As such, this initial regime has been labelled ‘deformation’. Because of the difference in load drop for this test as compared to the locus tracked by the stress-strain response, it also follows that this loss in stiffness (or increase in compliance) evident beyond the limit load is associated with local stretching and thinning and the initiation of tearing, such that this regime has been labelled ‘initiation’. This has been done in spite of the observation that a major component of that energy is due to continued plastic flow. pl e co py -n Not surprisingly, steels #7 and #8 both broke cleanly in two pieces. In contrast, steel #2 simply deformed and showed local stretching and lateral local thinning at the notch, and expansion toward the back face sufficiently so to pass between the anvils – and showed no evidence of the potential to break as a result of the testing. In all three cases, the compliance trends are as anticipated. For the pair of steels with the lowest toughness the compliance increasing sharply once the peak load is reached for each of these steels. In contrast, the compliance for steel #2 increases only slightly from its initial trend, which signifies that little change in bending stiffness occurred over the course of the test. Sa m As the trends for the other steels shown in Fig.3b indicate, none shows the complexity evident for the steel just discussed. The energies measured relative to these trends for steels #2, #3, #7, and #8 are respectively in ft-lb 250, 192, 17.5, and 36 (in J, 339, 259, 23.7, and 49) – the last of which is a linearly scaled full-size equivalent (FSE) energy. The lowest energy pair of results (steels #7 and #8) developed force-displacement response that is characteristic of steels and other metallic materials that fail at relatively low energies, which was commonly the case through the 1960s. Analysis indicates that these peak loads correspond closely with what is inferred as the limit load for simple bending relative to their respective values of the UTS. It is apparent that their force-displacement response is normalized as expected by the use of the It follows from the pair of results that developed for the very tough steels that the CVN test is tending toward the limits of its utility, whereas as history has shown it is a viable basis to quantify the response at much lower toughness levels. Utility of CVN testing in pipeline applications A webcrawl also was done on the topic of Charpy impact testing, but for the purposes of this section it was filtered with the terms of pipeline and applications. This search led to well in excess of 600,000 hits, many of which are duplicated or repeated in some manner. However, the number of results, the dates of the citations, and content and scope of the documents make clear that this testing practice is broadly still used and that the CVN test remains relevant in the pipeline industry. The Journal of Pipeline Engineering Fig.4. Energy components from instrumented CVN testing partitioned as in Fig.3. ot f or d is t rib ut io n 190 e co py -n It is apparent in this context that the CVN test remains the basis for fracture-control plans for the industry. The American Petroleum Institute (API) and others adopt the CVN test as the basis to quantify fracture resistance (for example, Ref.26), whereas the Battelle drop-weight tear test [27] (B-DWTT) – that subsequently simply is identified by the acronym DWTT – is used to characterize fracture mode. At the same time, literature based on full-scale testing has been developing that is consistent with the above noted results (such as Ref.28), indicating that the CVN test has reached the limits of its utility in applications to the tough and weldable high-strength steels being adopted by the pipeline industry. Sa m pl Figure 4 (adapted from Ref.28) was derived from the components of dissipated energy as partitioned in the context of Fig.3 to shed light on the future utility of the CVN test in applications to high-toughness steels. The y-axis in this figure is the relative fraction of the total energy dissipated in each of several tests involving a total of nine steels whose toughness as quantified by CVP ranged up to 260 ft-lb (352 J), values for which are shown on the x-axis. For each test, these fractions must sum to unity. As often occurs for CVN tests, the results in Fig.4 show scatter, which is made worse for the present analysis by the gradual changes in compliance that make it difficult to identify the breakpoints between the energy components. Although there is scatter, the key point in regard to Fig.4 is not clouded by such concerns: it is clear from the figure that the trend for propagation energy decreases to zero a value of CVP at about 250 ft-lb (339 J) and above, as indicated by the data points circled in the figure. This outcome is anticipated in light of Fig.3b, and the related discussion, because the force-displacement response for steel #2 (which had a CVP of 350 ft-lb (339 J)) involved only bending and stretching, with lateral contraction across the notch and expansion on the back face – but no propagation. In contrast, the models developed in the 1970s to characterize the fracture-arrest process, which remain in use today to quantify the toughness required to arrest the phenomenon termed running fracture, were calibrated in a framework that involved only lower-toughness steels that Fig.4 shows involved a significant component of crack-propagation energy. Clearly one can question the use of the above-noted fracture-arrest models that assume fracture controls failure in applications where the CVN test does not produce cracking. One could also question the initial use of the CVN test – which involves a notched beam in simple bending – for use in applications for which little or no bending occurs. Justification for that choice was as easy then as it is today – as the models were empirically calibrated such that the value of CVP could have been a surrogate for some other related metric. In addition, the basic concepts and tools of fracture mechanics were in the early stages of development – but more critically the phenomenon that was then termed ‘propagating fracture’ is complex and since has defied a 3rd Quarter, 2013 191 rib ut io n (a) is t (b) the use of full-thickness samples, such as the DWTT. It is, however, clear that as time has passed the utility of the DWTT even to quantify the fracture mode, which is its purpose, is open to question. Issues with so-called inverse fracture have become widely evident since about 2000, and as the steels have become increasingly tougher the DWTT practice is also developing gross stretching and evidence of overall distortion – as similarly has occurred for the CVN practice. -n ot f first-principles’ formulation. Today it is understood that this phenomenon first involves axial straining ahead of the crack in a plastic zone that propagates along the pipeline, with symmetric thinning leading to ‘propagating shear’ as shown in Fig.5, with in-plane stretching occurring until shear failure (plastic collapse) occurs through-wall. or d Fig.5. Cross sections through propagating shear contrasted to tearing shear (from Ref.27): (a) symmetric shear (as polished, observed during steady-state propagation): t = 0.560 in (14 mm); (b) asymmetric tearing shear (rough grind, observed during ring-off arrest): t = 0.560 in (14 mm). m pl e co py The significant in-plane stretching evident in Fig.5a results in symmetric through-wall thinning and eventual collapse via shear – much as could occur in a tensile test. This symmetric process develops during steady-state propagation, which corresponds to the scenario the CVN test has used to characterize the arrest toughness. Figure 5b shows the transition that occurs in this failure behaviour as steady-state transitions to arrest, which often results in the ring-off of the failure path. While, as Fig.5b indicates, this transition leads to asymmetric (so-called tearing) shear, the failure process during arrest remains shear dominated – with much less thinning occurring, as now the driving force for the process is greatly diminished. Sa It follows that at least for pipelines made of highertoughness steels the usual CVN test practice appears to have limited utility. Whereas the steady-state pipelinefailure process primarily involves stretching and thinning that is largely free of bending, the CVN test relies on bending with thickening developing as the failure process in the specimen continues to the back face. In this context, any bending-impact test falls short of emulating the structural response. In spite of this fact, it has been asserted that the issues with the CVN test in pipeline applications evident in the stretching, tearing, and lateral flow parallel to the notch could be offset by Figure 6 (courtesy of Bernard Hoh) provides graphic evidence of the gross distortion that can develop through the use of the DWTT in applications involving tough steels. While these images reflect heavy-wall linepipe made with a thickness of 1.5 in (38 mm) in X-70 steel, the results reflect testing done in a machine with adequate excess energy. Thus, while considered by some a potential alternative to the CVN practice in anticipation that its increased depth and larger overall size would offset such concerns, the DWTT practice does not provide a path forward for such applications, even as a stopgap solution. On this basis, work is needed to better understand the factors controlling the failure process such that an appropriate mill test can be identified in lieu of the bending impact tests in use today. Utility of CVN testing in other applications involving tough steels Central to the discussion of the CVN test in its application to pipelines was the thesis that the issues discussed can be traced to the toughness of the steel. This section further considers this hypothesis in regard to the classes of steel in use in the ground-vehicle industry. The Journal of Pipeline Engineering or d is t rib ut io n 192 py -n ot f (a) co (b) Sa m pl e Like the pipeline industry, the automotive / groundvehicle industry has pushed the development of tough, strong, weldable steels. As such, it is not surprising that about the same time that issues with the CVN test were emerging in applications to the pipeline industry, similar concerns with the CVN test became evident in that industry. For example, a 1987 paper [29] authored from within the Structures and Dynamics Division of Caterpillar Inc. notes that “recently, two new families of steels …… have become available” that were of interest to that industry. After considering those steels in light of several basic materials’ tests, the paper concludes by noting that some tests are “reliable indicators of material performance in components while Charpy V-notch energy is of little value for quantitative engineering analysis”. The paper goes on to state that an “alternative impact test which more closely represents material behaviour in components” is needed. Fig.6. Unusual fracture and gross distortion in a DWTT of high-toughness steel: (a) ‘fracture’ features (notch is up); (b) perspective view of the tested specimen. Another paper associated with the evaluation of highpower beam welds (such as electron-beam or laser welds) noted that the failure of such welds might result in a plastic constraint loss around both the notch and crack tip, making it difficult to evaluate fracture performance of girth-welded pipe joints. This paper determined that intrinsic fracture toughness was lower than the results of standard Charpy specimens and fatigue pre-cracked three-point bend specimens. It was concluded that differences in plastic constraint between the structure and a three-point bend specimen underlie this situation. Thus, in addition to applications where plasticity acts to limit the utility of the CVN test, as occurred for the pipeline and the ground-vehicle cases, issues can develop that limit the direct transferability of the CVN test results due to constraint – in the absence 193 rib ut io n 3rd Quarter, 2013 py -n ot f or d is t (a) co Fig.7. Energy measured in a quasi-static evaluation of forcedisplacement response [21]: (a) photographs of the set-up; (b) results of the comparison. pl e (b) Sa m of the significant effects of plasticity. It follows that care should be exercised in adapting what otherwise begins as screening tests to more general applications Factors affecting the energy measured in a CVN specimen Earlier discussion alluded to the use of subsize specimens and differences in tup design as drivers for variability in the measured energy and differences in the failure response in a CVN specimen. As both of these aspects have been the subject of related research, this section briefly elaborates on such effects in regard to measurements directed at assessing their significance. Tup design issues Consider first the role of tup geometry, which for the present is evaluated specifically in regard to the energy measured via ASTM E23 (8-mm radius striker) rather than contrasting such results relative to the outcome for ISO 148-1 – with the results shown in Fig.7 developed in the late 1990s being instructive in this context. Figure 7a shows a view of the specimen after loading well beyond the limit load, whereas part (b) of this figure shows the force-displacement response as measured by a Dynatup striker mounted in series with a load cell and a linear variable differential transformer. It is evident The Journal of Pipeline Engineering ot f or d is t rib ut io n 194 Sa m pl e co py -n from images synced to specific loads that over the course of the loading the specimen begins to wrap around the tup, which occurs as the rotation increases just the limit load. As this happens the load, as measured by the tup, falls off in comparison with the load measured by the load cell. In view of the inset in Fig.7a, this decrease could reflect the observation that the face of the tup is unloaded as the specimen rotation about the tup increases. It is further evident from the inset that, as the rotation increases, the load is increasingly being transferred to the back face of the CVN specimen along the outer extremities of the radiused tup face. Such behaviour indicates that as the rotation increases the CVN specimen is subjected to a wedge-opening load that is not quantified in simple terms by the measured load. Accordingly, while the response evident in Fig.7b indicates that the measured load underestimates the vertical component of the loading by about 7%, the wedge-opening load is not quantified. It follows that aspects such as tup wear could lead to variability, particularly if this wear occurs local to the shoulders of the ASTM E23 tup design. Likewise, it follows that the use of a smaller-radius tup that makes contact over a narrower zone, as occurs for the ISO 148-1 tup design in comparison to that for E23, would lead to a reduction in the wedge-opening load. Finally, it follows that the observation of a threshold toughness below which tup geometry has no apparent Fig.8. Comparison of subsize and full-size CVN energy for one X-70 steel. effect (for example, Ref.31) reflects the observation that less-tough steels do not survive in the test to the point beyond which rotation and contact with the flanks of the tup occur to a significant extent. Thus, care must be taken when interpreting the energy measured in a CVN test for higher-toughness steels, as it can underestimate the actual force involved, while at the same time fail to reflect the effects of wedge-opening loading. Subsize specimen issues Results have been reported since the 1950s (such as Ref.32) concerning the effects of subsize specimens, with references cited in that work that date to the late 1940s. As such, the effect is not new. While the topic has been investigated for many decades, the Appendices of ASTM E23 have not provided a conclusive view on the role of thickness in the form of a correlation, suggesting that some uncertainty remains as to the circumstances that control the effect, and whether there is a consistent pattern. Issues in the context of reproducibility have been reported for the CVN test longer than the effects of subsize specimens. For example, work also reported in the 1950s (such as Ref.33) that cite “poor steel, poor heat treatment, or both” and note “variations….caused by poor testing techniques, and the poor condition of the Charpy machine, or both” as underlying causes for those issue, 195 or d is t rib ut io n 3rd Quarter, 2013 (a) (b) ot f Fig.9. Force-displacement response for CVN specimens made of X-70 steel: (a) full-size specimen; (b) 2/3-thick specimen. separated into two pieces – which, for the samples tested in the 516-ft-lb (700-J) Charpy machine (most of them), was well below 80% of that machine’s capacity. co py -n all of which could contribute to the lack of clarity in regard to thickness effects. Also possibly contributing to this circumstance is the host of factors that have more recently been considered in the context of ASTM STP 1248 [34], which considered the specimen, the anvil and striker, the test procedures, and still other topics. Accordingly, this section adds only to the knowledge base – particularly given it relies on the ASTM E23 tup and considers subsize effects in reference to tough steel. Sa m pl e Figure 8 presents results for 1990s’ vintage X-70 steel cut from large-diameter pipe that had a wall thickness of 0.560 in (14.2 mm), comparing the energy to fail for sub-thickness and full-size thickness CVN specimens. In this figure, the x-axis represents the energy from testing full-size specimens whereas the y-axis presents the FSE energy for the 2/3-thick specimens, where FSE energy is linearly scaled energy density per unit area. While some of the specimens were tested in a 264-ft-lb (358-J) Charpy impact machine, the majority were evaluated in a 516-ft-lb (700-J) Charpy machine. For the higher-toughness cases, some of the specimens did not separate into two pieces. In particular, for cases where the FSE energy was about 200 ft-lb (270 J), the 2/3-thick specimens did not separate into two pieces, whereas those that were made full thickness did. However, at FSE energy levels above 250 ft-lb (339 J), neither the full-thickness nor the 2/3-thick specimens This same comparison after excising results where the measured energy exceeded 80% of the machine’s capacity (tests in the 264-ft-lb (358-J) machine) suggest that for energies less than 180 ft-lb (244 J) linearly scaling the 2/3-size Charpy data to an FSE value slightly underestimates the result measured on full-thickness specimens. In contrast, for energies greater than 180 ft-lb (244 J), linearly scaling the 2/3-thick specimen data overestimates the result measured on full-thickness specimens, which could be due to constraint developing in the specimens at higher-toughness levels. As the degree of constraint decreases, the extent of the through-thickness thinning along the notch increases, which adds to the fraction of the deformation component of energy. Fig.9 illustrates this point, and makes clear the earlier assertion that instrumented testing is essential for understanding the outcomes of such testing. Figure 9a presents the force-displacement response of a full-size CVN specimen with toughness greater than 170 ft-lb (230 J), while Fig.9b presents the comparable result for a 2/3-thick specimen made of the same steel. When these results are plotted with the y-axis used in Fig.3 (here the UTS is the same for both), 196 The Journal of Pipeline Engineering io n is t rib ut This paper also considered the utility of standardized impact testing such as ASTM E23 and ISO 148-1 in the context of applications involving tough steels, with the potential utility of the DWTT practice also considered. It was apparent in that discussion that the response of bending-impact specimens is, as expected, a consequence of the loading and the structural geometry involved, and thus a poor analogue for the circumstances that significantly differ. This aspect was discussed in regard to pipelines and the ground-vehicle industry, with the conclusion that bending geometries are a poor analogue for the fracture-propagation process that develops in pipelines, and also for some other applications. Given that structural geometry and the loadings, in conjunction with the material’s properties, dictate the failure response, where critical differences exist in that context, alternative testing practices should be developed. Sa m pl e co py -n ot f Inlaid into the force-displacement for these figures is a cross-hatched estimate of the energy dissipated in propagation – which for the subsize specimen excludes the component of energy due to deformation induced by the decreased constraint. Each part of Fig.9 also indicates the ratio of propagation to total energy decreases based on that estimate. This ratio indicates that as the specimen thickness is reduced, the deformation component for the 2/3-thick specimen has increased as compared to the full-size specimen. While the scope of the effort for the client that supported this work precluded more comprehensive analysis, it was expected that if the energies are separated into initiation, deformation, and propagation components as just noted, a direct correlation would emerge between the propagation energies for 2/3-thick and full-size specimens. In that context, comparing the outcomes for the results shown in Fig.9, the propagation energy density for the standard thickness specimen is 717 ft-lb/in2 (150 J/cm2), while that for the 2/3-thick specimen is 704 ft-lb/in2 (148 J/cm2). While not an exact match, these results are well within material scatter limits, and suggested further consideration of this plausible correlation – which unfortunately was beyond the scope of that effort. materials. The origin of the CVN test and its development into a standard was discussed, and the evolution of mechanical and other properties motivated by industry demands was evaluated. Interpretation of the CVN test was discussed in regard to the force-displacement and compliance response that develops during the test, and factors affecting the energy measured and controlling failure of the CVN specimen were assessed, including the tup design and the use of subsize specimens. or d the trends for both tests are comparable up through the peak load and beyond the compliance change that indicates ‘propagation’ has begun. However, during the course of the propagation phase, the trends deviate. It is apparent that for the full-size specimen the propagation trend in Fig.9a continues in the same manner evident for the data shown in Fig.3. As in those trends, the energy dissipated in propagation tracks along the slope shown until the change in compliance indicates a transition to deformation has occurred, which (as in Fig.3) is associated with final rotation and eventual separation of the halves. In contrast, the trend for the subsize specimen shown in Fig.9b indicates that after some propagation occurs as for the full-size specimen, the energy in what was the propagation phase begins to increase. As suggested above, this reflects the decreased constraint, which admits greater through-thickness thinning along the notch, and so increases the relative fraction of deformation energy during what was otherwise steady propagation in the thicker specimen. This deformation component continues to increase through the transition to overall deformation energy, flagged by the vertical trend, after which rotation and separation occur as for the fullthickness specimen. Summary and discussion This paper has reviewed the Charpy V-notch impact test and assessed its utility to characterize fracture resistance in applications to modern tough materials in contrast to those encountered prior to the availability of such Conclusions While a number of conclusions have been drawn as the paper developed, two primary conclusions bear repeating here: • because the behaviour in an impact test can be complex, data interpretation and assessment of their practical implications is best based on data developed using a well-instrumented machine; and • because structural geometry and the loadings act in conjunction with the material’s properties to control the failure response, where critical differences exist in that context, alternative testing practices are needed that are adapted to the specific loading and failure response of the structure of interest – which is now the case for fracture propagation in pipelines where tough materials are involved. Acknowledgments The data reported herein in Figs 7, 8, and 9 were generated under contract to Alliance Pipeline, as part of developing its fracture-control programme in the late 1990s. Permission to release that data is gratefully acknowledged, as is the author’s related collaboration with David Rudland, then with Battelle. 3rd Quarter, 2013 197 Sa m pl e co py -n io n rib ut is t ot f 1. T.A.Siewert, M.P.Manahan, C.N.McCowan, J.M.Holt, F.J.Marsh, and E.A.Ruth, 1999. The history and importance of impact testing. In: Pendulum impact testing: a century of progress. ASTM STP 1380, American Society for Testing and Materials, pp 3-16. 2. A.E.White and C.L.Clark, 1925. Bibliography of impact testing. Dept of Engineering Research, University of Michigan. 3. A.LeChatalier, 1892. On the fragility after immersion in a cold fluid. French Testing Commission, 3. 4. S.B. Russell, 1898. Experiments with a new machine for testing materials by impact. Trans ASCE, 39, p 237, June. 5. M.G.Charpy, 1901. Note sur léssai des métaux à la flexion par choc de barreaux entaillés. Soc. Ing. Français, pp 848–877. Reprinted in ASTM STP 1380, American Society for Testing and Materials, 2000. 6. W.J.Hall, H.Kihara, W.Soete, and A.A.Wells, 1967. Brittle fracture of welded plate. Prentice Hall. 7. ASTM E23 - 12c. Standard test methods for notched bar impact testing of metallic materials. American Society for Testing and Materials, 03.01. 8. BS EN ISO 148-1: 2009. Metallic materials - Charpy pendulum impact test - Part 1: Test method; Part 2: Verification of testing machines; and Part 3: Preparation and characterization of Charpy V-notch test pieces for indirect verification of pendulum impact machines, International Standards Organization. 9. EN 10045-1: 1990. Charpy impact test on metallic materials. Test method (V- and U-notches) – now withdrawn and superseded by BS EN ISO 148-1. 10.ASTM, 1938. Symposium on impact testing. Proc. 41st Annual Meeting of the American Society for Testing and Materials, Atlantic City, 28 June. 11.ASTM, 1955. Symposium on impact testing. Proc. 58th Annual Meeting, Atlantic City, 27 June. ASTM STP 176, American Society for Testing and Materials, 1956. 12.ASTM E2248 – 13. Standard test method for impact testing of miniaturized Charpy V-notch specimens. American Society for Testing and Materials, 03.01. 13.ASTM D6110 – 10. Standard test method for determining the Charpy impact resistance of notched specimens of plastics. American Society for Testing and Materials, 08.03. 14. ASTM E1253 – 13. Standard guide for reconstitution of irradiated Charpy-sized specimens. American Society for Testing and Materials, 12.02. 15.ASTM F2231 - 02(2008). Standard test method for Charpy impact test on thin specimens of polyethylene used in pressurized pipes. American Society for Testing and Materials, 08.04. 16.ASTM A327 / A327M – 11. Standard test methods for impact testing of cast irons. American Society for Testing and Materials, 01.02. 17.ASTM E2298 – 13. Standard test method for instrumented impact testing of metallic materials. American Society for Testing and Materials, 03.01. 18.W.T.Matthews, 1969. The role of impact testing in characterizing the toughness of materials. ASTM ST 466, American Society for Testing and Materials, pp 3-20. 19. B.N.Leis and T.F.Forte, 2005. Managing the integrity of early pipelines – crack growth analysis and revalidation intervals. Battelle Final Report to the Research and Special Projects Agency, Contract DTRS56-03-T-0003, February. 20. W.J.McGregor Tegart, 1966. Elements of mechanical metallurgy. Macmillan. 21.D.L.Rudland and B.N.Leis, 2007. Pipeline dynamic fracture test program. Battelle’s Final Report to R.J.Eiber, Consultant, and Alliance Pipeline, 7 February: filed in the Archives of the Canadian National Energy Board (NEB) Hearing GH 3-97. 22.R.D.Galliher and B.N.Leis, 2003. Evaluation of fracture resistance via CVN and DWTT practices for recently produced X80 and X100 line pipe steels. Battelle’s Final (Proprietary) Report to a commercial client, 23 October. 23.NIST impact verification program: http://www. nist.gov/mml/acmd/structural_materials/charpyverification-program.cfm 24.B.N.Leis, 1993. Instrumented CVN and tensile testing of nine line-pipe steels in grades from X42 to X80. Battelle IR&D Report, April. 25.ASTM E8 / E8M – 11. Standard test methods for tension testing of metallic materials. American Society for Testing and Materials, 03.01. 26. B.N.Leis, 1977. Relationship between apparent (total) Charpy Vee-notch toughness and the corresponding dynamic crack-propagation resistance. Battelle’s Final Report to R.J.Eiber, Consultant, and Alliance Pipeline, 7 February: filed in the Archives of the Canadian NEB Hearing GH 3-97. 27. B.N.Leis, R.J.Eiber, L.E.Carlson, and A. Gilroy-Scott, 1998. Relationship between apparent Charpy Veenotch toughness and the corresponding dynamic crack-propagation resistance. International Pipeline Conference, ASME, Calgary, pp 723-732. 28.B.N.Leis, 1999. Characterize fracture features for samples from the first and second burst tests. Battelle’s Final Report on Task 6, to R.J.Eiber, Consultant, and Alliance Pipeline, 7 May: filed in the Archives of the Canadian NEB Hearing GH 3-97. 29.D.J.Lingenfelser, 1987. Application of basic material tests for evaluating new engineering materials. SAE Technical Paper 870801. 30.M.Ohata, M.Toyoda, N.Ishikawa, and T.Shinmiya, 2007. Significance of fracture toughness test results of beam welds in evaluation of brittle fracture performance of girth welded pipe joints. J. Pressure V. Tech., 129, 4, pp 609-618. or d References 198 The Journal of Pipeline Engineering 33.D.E.Driscoll, 1956. Reproducibility of Charpy impact test. ASTM Special Technical Publication 176, American Society for Testing and Materials, pp 70-74. 34. T.A.Siewert and A.K.Schmieder, 1995. Eds: Pendulum impact machines: procedures and specimens for verification. ASTM STP 1248, American Society for Testing and Materials. Sa m pl e co py -n ot f or d is t rib ut io n 31.E.A.Ruth, 1995. Striker geometry and its effect on absorbed energy. ASTM STP 1248, American Society for Testing and Materials, pp 101-110. 32.R.S.Zeno, 1956. Effects of specimen width on the notched bar impact properties of quenched and tempered and normalized steels. ASTM STP 176, American Society for Testing and Materials, pp 59-69. 3rd Quarter, 2013 199 CTOA testing of pipeline steels using MDCB specimens1 io n by Dr Robert L Amaro, Dr Jeffrey W Sowards, Elizabeth S Drexler, J David McColskey, and Christopher McCowan* rib ut NIST, Applied Chemical and Materials Division, Boulder, CO, USA T ot f or d is t HE CRACK-TIP-OPENING angle (CTOA) is used to rank the relative resistance to crack extension of various pipeline steels. In general, the smaller the CTOA value, the lower the resistance to crack extension. It is unclear, however, whether CTOA is a material property that is valid for all thicknesses and rates of crack growth. Historically, drop-weight tear tests (DWTT) and modified double-cantilever beam (MDCB) specimens have been used for measuring CTOA. Tests using either specimen may be conducted at quasi-static and dynamic rates. The fastest displacement rates achieved in our laboratory were near 14 m/s, resulting in crack extension rates near 30 m/s for high-toughness linepipe steels. In-service crack extensions for ductile-crack fracture can be more than 100 m/s. The failure mode at this rate is plastic collapse, and it is uncertain if correlations can be drawn between in-service failures and laboratory tests conducted on thinner material tested at slower rates.We describe the evolution of our test method using MDCB specimens from 2006 to 2012 and the direction we anticipate for future CTOA research. T worked on developing new measures of fracture control. Among these, crack-tip-opening angle (CTOA) is one alternative for characterizing fully plastic fracture [6-7], especially for running ductile cracks in pipes [2,7-13]. In cases where the fracture process is characterized by a large degree of stable tearing, CTOA has been recognized as a measure of the resistance of a material to fracture [6,9]. The main advantages of CTOA are that it can be directly measured from the crack-opening profile and can be related to the geometry of the fracturing pipe. However, there are difficulties in determining CTOA with a simple measurement technique that would be widely available to many material-test laboratories. In addition, the CTOA criterion can be implemented in finite-element models of the propagating-fracture process [6,9,13,14]. co py -n HE INCREASING DEMAND for natural gas as an alternative energy source implies continued growth of gas pipeline installations and the qualification of materials in the actual pipeline network. A difficult problem to be solved for the economic and safe operation of high-pressure gas pipelines is the control of ductile-fracture propagation [1]. As a result, the accurate prediction of the resistance to fracture and ductile-fracture arrest in pressurized gas pipelines are currently important issues. Sa m pl e Initially, the measure of a material’s fracture resistance was determined on the basis of Charpy V-notch (CVN) shelf energy, such as that used in the Battelle two-curve model [1]. Later correlations were developed between Charpy and dynamic drop-weight tear test (DWTT) data. The Battelle two-curve model worked well for many years, but when applied to modern higher-strength pipeline steels, significant errors are apparent [2–5]. Correction factors have been developed [1,4,5] for high-strength steels; however, use of these correction factors adds further uncertainty to the estimates. Thus, in parallel with the CVN- and DWTT-based fracture strategies, pipeline designers have 1. Contribution of NIST, an agency of the US government: not subject to copyright in the United States. *Corresponding author’s contact details: tel: +1 303 497 3699 email: [email protected] The literature contains a number of different specimen geometries for studying ductile-fracture propagation with the CTOA criterion, such as middle-tension specimens, M(T) [6,15,16], compact-tension specimens, C(T) [6,15,16], DWTT specimens (with methodologies based on one specimen [1,8,9,10] or two specimens [17]), three-point bend specimens, 3-PB [7,12], and modified doublecantilever beam specimens, MDCB [3,13]. Our efforts have focused on test methods using the MDCB specimen [3,13,18-20] that is promising for CTOA measurement in pipeline steel, because this specimen design allows an extended region for steady-state crack growth and for larger plastic deformation at the crack tip. This may 200 The Journal of Pipeline Engineering API Designation SMYS MPa (ksi) O.D. mm(inch) Thickness mm 1 N/A (~X70) 517 (75) 0.51 (20) 9.7 2 X52 359 (52) 0.51 (20) 8.0 3 Grade B 244 (35) 0.56 (22) 7.4 4 N/A (~X52) 335 (48) 0.51 (20) 7.9 5 N/A 281 (40) 0.56 (22) 7.8 6 X65 448 (65) 0.61 (24) 7 X65 448 (65) 0.51 (20) 8 X65 448 (65) 0.76 (30) 17.0 9 X100 689 (100) 1.32 (52) 20.6 10 X100 689 (100) 1.22 (48) 20.0 11 X100 689 (100) 1.22 (48) 20.0 12 X70 spiral 483 (70) 0.91 (36) 13.7 is t rib ut 25.0 • Steel #2 is an API X-52 characterized by a ferrite-pearlite structure, with a significantly larger ferrite grain size than steel #1. This steel has the most pronounced banding (of pearlite) of the steels evaluated here. • Steel #3 is an API Grade B ferrite-pearlite steel without banding. • Steel #4 is a ferrite-pearlite steel with low banding. • Steel #5 is a ferrite-pearlite steel without banding. • Steel #6 is an API X-65 grade of ferrite-pearlite steel, which might be better described as ferritecarbide, because there is very little pearlite in the microstructure. The grain size of this steel was not measured, but the ferrite grain size is similar in size to that of steel #1. • Steel #7 is an API X-65 grade with no pearlite and a fine non-equiaxial microstructure. • Steel #8 is an API X-65 grade with a ferritepearlite microstructure and heavy bands of pearlite. • Steel #9 is an API X-100 grade: this is an experimental alloy that was used for full-scale testing. • Steels #10 and #11 are two modern API X-100 bainitic steels. • Steel #12 is an API X-70 spiral pipe steel, the microstructure of which is not known. -n ot f simulate the conditions surrounding running cracks on pipelines as they exhibit plastic regions on the order of 2.5 pipe diameters ahead of the crack tip and 0.3 diameters on each side of the crack line [10]. Moreover, the MDCB specimen can be cut directly from pipe with no subsequent flattening required, which avoids potential load-history effects due to pre-straining the material. 31.5 or d Table 1. Information on pipeline steels tested. io n ID Number co py NIST has a history of conducting CTOA tests with the MDCB specimen. The gripping mechanism has evolved to increase constraint, methods to mark grids on the specimens have improved, and loading and recording systems were developed to conduct tests at high rates [21-31]. Sa m pl e In this summary of our CTOA testing, data from 12 pipeline steels are presented. They are described here and referenced by ID number in the sections that follow. Table 1 summarizes the dimensions of the pipes from which all samples were extracted. The specified minimum yield strength (SMYS) and the API designations are also provided in the table. In Table 2, the chemical compositions of the steels are given, while in Table 3, the grain size and pearlite volume fraction are given for the ferrite/pearlite steels (#1 – #5). Microstructures The microstructures of the 12 pipeline steels tested are briefly described as follows: • Steel #1 is a ferrite-pearlite steel with low carbon (low pearlite) content and a fine ferrite grain size. This steel represents a modern, fine-grained ferrite pipeline steel. Tensile properties The tensile properties of steels #1 to #5 were measured with flat tensile specimens (due to plate thickness), while round tensile specimens (6-mm diameter) were tested for steels #6 to #11. The flat specimens were 3rd Quarter, 2013 201 #1 #2 #3 #4 #5 #6 Al 0.031 B <0.0002 0.06 0.24 0.27 0.18 0.25 0.07 Co 0.006 0.025 0.007 0.014 0.025 0.003 Cr 0.02 0.024 0.029 0.021 0.019 0.12 Cu 0.11 0.038 0.015 0.054 0.046 Mn 1.46 1.03 0.36 0.52 0.97 Mo 0.025 0.016 0.007 0.009 0.017 Nb 0.054 0.007 0.005 0.005 Ni 0.10 0.064 0.021 0.021 P 0.01 0.016 0.005 0.026 0.013 0.008 S <0.01 0.013 0.015 0.010 0.012 0.004 Si 0.28 0.057 0.009 0.043 0.061 0.094 or d #7 #8 Al 0.030 0.039 B <0.0002 0.0002 C 0.07 0.08 #9 Co 0.002 Cr 0.13 Cu py 0.066 0.17 0.03 0.04 #10 #11 #12 0.025 0.012 0.039 <0.0001 0.0003 0.064 0.04 0.084 0.001 0.003 0.03 0.021 0.023 0.07 0.09 0.30 0.286 0.28 0.31 1.59 1.56 1.90 2.092 1.87 1.56 0.006 0.15 0.127 0.23 0.20 0.005 0.003 0.008 0.041 0.017 0.069 0.003 pl Mo 0.04 0.10 e Mn 0.07 0.003 0.002 ot f 0.002 -n 0.045 co V 1.48 rib ut 0.003 Ti 0.12 0.007 is t N io n C 0.03 0.04 Ni 0.14 0.21 0.50 0.501 0.47 0.11 P 0.009 0.011 0.008 0.10 0.009 0.010 S 0.004 0.003 0.0005 0.002 <0.001 0.009 Si 0.092 0.325 0.10 0.108 0.099 0.24 Ti 0.02 <0.01 0.007 0.17 0.013 V 0.04 0.04 0.006 0.002 0.003 m Nb Sa N Table 2. Chemical composition of the pipeline steels tested, by mass. Column numbers give identification number for the steel, as defined in Table 1. 202 The Journal of Pipeline Engineering Steel # Ferritic grain size (μm) Pearlite volume fraction (%) 1 2 3 4 5 6.5 11.8 10.8 N/A 22.2 5 37.1 25.3 37.9 17.1 Orientation E (GPa) σ 0.2 (MPa) σ UTS (MPa) σ0.2 / σUTS eu (%) ef (%) 1 L 211* 517 611 0.846 6.7% 35.0% 0.19 T N/A 543 606 0.896 8.0% 27.4% 0.29 L 211* 360 556 0.647 12.3% 32.7% 0.38 T N/A 448 576 0.777 11.1% 25.6% 0.43 L 212* 244 451 0.541 19.6% 37.8% 0.52 T N/A 255 459 0.555 18.8% 38.0% 0.49 L 210* 335 535 0.626 12.9% 34.9% 0.37 T N/A 428 560 0.764 10.5% 22.0% 0.48 L 214 265 454 0.583 16.0% 38.0% 0.42 T NA 248 453 0.547 19.5% 35.0% 0.56 L 201 460 534 0.870 8.2% 24.7% 0.33 T 218 497 560 0.890 7.7% 15.9% 0.48 L NA 502 570 0.880 6.8% 25.7% 0.26 T N/A 511 577 0.885 7.2% 20.9% 0.34 L 217 522 618 0.844 10.1% 27.3% 0.37 T N/A 576 644 0.894 6.9% 24.8% 0.28 L N/A 694 801 0.910 4.6% 20.3% 0.23 N/A 797 828 0.966 4.1% 19.3% 0.21 192 722 855 0.844 4.6% 17.8% 0.26 T 213 912 916 0.995 2.6% 18.0% 0.14 L 198 729 838 0.869 5.8% 20.5% 0.28 T 207 833 868 0.989 4.7% 17.5% 0.27 T NA 576 650 0.940 NA NA NA 5 6 7 8 9 T L co 10 m pl 12 e 11 rib ut is t or d 4 ot f 3 py 2 eu/ef io n Steel # -n Table 3. Measurements of the grain size and ferrite fraction for the ferrite/pearlite steels. *Average determined from dynamic elastic modulus test Sa Table 4.Tensile properties of the materials. (Note: * = average determined from dynamic-elastic-modulus test.) 6 mm wide. Full-thickness specimens (Table 1) were tested for the longitudinal orientation, and typically 3-mm thick specimens were tested for the transverse orientation. All specimens had a gauge length of 25.4 mm. Experiments were performed either in a screwdriven tensile testing machine of 100-kN capacity, or a closed-loop servo-hydraulic machine of 100-kN capacity. Tests were conducted in displacement control at rates of 0.25 mm/min for the flat specimens and 0.1 mm/ min for the round specimens. The measured mechanical properties of the steels are given in Table 4, where E is the Young’s modulus, σ0.2 the yield stress, σUTS the ultimate strength, eu the uniform elongation, and ef the fracture elongation. In addition to the standard properties, the ratios of σ0.2/σUTS (stress ratio) and eu/ef (strain ratio) are also given in Table 4. These two parameters indicate the strain-hardening potential of the steel. As shown in Fig.1, the stress ratio increases as the strain ratio decreases. 3rd Quarter, 2013 203 CTOA test matrix rib ut io n CTOA tests were conducted on X-52, X-65, X-70, and X-100 pipeline steels and other pipeline grades not identified with an API designation (Table 1). The tests were conducted by tensile loading MDCB specimens at actuator rates ranging from 0.002 to 14,000 mm/s. The 8,000 and 14,000-mm/s displacement rates were attained with a disc spring set-up [22]. Early tests were quasi-static, and later a series of tests were conducted on X-65 (#6) and X-100 (#9) with changing actuator rates. High-rate tests were also conducted on two additional X-100 steels (#10 and #11), and also on an X-70 steel (#12). CTOA test specimen py is t -n ot f • It can be cut directly from a pipe, without flattening; • The width and thickness are limited by pipe curvature and wall thickness; • The long ligament in the gauge section allows for the CTOA to be measured multiple times and averaged; • High constraint in the test section is promoted by two thicker loading arms; • The test section does not restrain the transition to slant mode shear fracture; • The test section is flat near the crack tip for ease of CTOA measurement. Fig.1. Strain ratio versus stress ratio. or d A modified double-cantilever beam (MDCB) specimen (Fig.2) was used to conduct the CTOA test. The specimen exhibits the following characteristics: e co The test specimens were cut with the notch direction along the axis of the pipe. The thickness of the curved plate was reduced by machining to obtain a flat plate, which eliminated the probable residual plastic strains that would be caused by flatting the plate by use of a straightening procedure. Sa m pl The specimens were fatigue pre-cracked following the ASTM standard procedure for conducting crack-tipopening displacement (CTOD) tests [32]. The precracking loads were selected to ensure that the ratio of stress intensity factor range to the Young’s modulus (∆K/E) remained below 0.005 mm-2. All specimens were fatigue pre-cracked at a ratio of R = 0.1 [13], to a crack-to-width ratio of a0/W = 0.3 to 0.5 [with a specimen width, W, equal to 182 mm, and a0 equal to the machined notch length (60 mm) plus the initial fatigue pre-crack length (approx. 10 mm)]. Methods and procedures Two apparatuses were used for CTOA testing, a ‘quasistatic’ set-up and a dynamic set-up. Fig.2. CTOA specimen, with dimensions in millimetres. Quasi-static apparatus For quasi-static testing (0.002 to 3 mm/s), a 250-kN uniaxial servo-hydraulic test machine was used. Tests were conducted in displacement control. As shown in Fig.3a, the load line ran through the centreline of the first pair of holes in the specimen. A digital camera and frame-capture software/hardware were used to capture images. The camera was mounted on an XYZ stage, which provided a stable platform to follow the crack tip. The image acquisition was controlled by a personal computer with image-analysis software: the captured images had a size of 2048 pixels × 1536 pixels, which resulted in a resolution of about 32 pixel/mm. Images were acquired and stored, along with time, load, and displacement data as the crack propagated across the specimen. Tests were stopped at 80 mm of crack extension beyond the machined notch tip. Details of the set-up have been reported previously [21]. Dynamic apparatus Tests with actuator rates of 3, 30, and 300 mm/s were performed on a 500-kN uniaxial servo-hydraulic test machine shown in Fig.3b [22]. As with the quasi-static 204 The Journal of Pipeline Engineering Data processing rib ut io n tests, the load line was located at the centreline of the first pair of holes in the specimen. The machine was adapted with large-capacity servo-valves to accomplish rapid loading. For actuator rates greater than 300 mm/s, the 500-kN uniaxial test frame was configured with a set of disc springs; the potential energies stored in these springs for the X-65 and X-100 tests were 5.6 kJ and 7.5 kJ, respectively [22]. Higher crack velocities were obtained by further increasing the stored energy with the use of sacrificial links1, which were made of aluminium alloy 7075-T6. In this configuration, grip-displacement rates up to 14 m/s were attained. or d is t Once images were captured, the CTOAs were measured using data within the distance from the crack-tip ranges prescribed by the ISO draft standard [35] and the ASTM standard [36]. Within these ranges on the samples, we used the following four approaches to measure the CTOA [23]: -n ot f • Method 1 used an algorithm that located the crack tip in the data, and then selected pairs of points along the crack profile at prescribed distances from the tip to calculate CTOA. The crack tip was always included in this calculation of the CTOA. • Method 2 used data-point pairs that were within the range 0.1 mm to 0.2 mm behind the crack tip to fit lines within this region (Fig.5a). This method never included the crack tip. • Method 3 used data points marking the upper and lower grid lines to fit lines for CTOA calculation. Each line was fitted with 2 to 10 points, located within the increment 0.5 to 1.5 mm from the crack tip (Fig.4b). • Method 4 used all of the profile data in the interval 0.5 to 1.5 mm to define the two bestfit lines associated with the upper and lower crack-tip profiles to calculate the CTOA. In this case, typically 100 to 200 data points were used for each line fit (Fig.4b). Highspeed camera 2 Specimen Sa m pl e co py Discspring setup Highspeed camera 1 Fig.3. Set-ups for (a - top) ‘static’ set-up with camera on a motion-control XYZ stage, and (b - bottom) dynamic set-up with high-speed cameras and springs. The software required the operator to trace the profile of the crack tip, and mark data points along the closest set of upper and lower grid lines, as shown in Fig.4. The CTOA values for each method were then calculated from the collected images from each specimen, and an average CTOA for each method was determined. 1 Sacrificial links were inserted into the load line and loaded to compress the springs. The links were calculated to fail at the load required to full compress the springs, which resulted in abruptly releasing the stored energy onto the CTOA specimen. 205 io n 3rd Quarter, 2013 or d is t rib ut Fig.4. Showing (a - left) crack edge traced by the operator and points marked on the gridlines adjacent to the crack, and (b - right) two sets of lines fitted with grid points and crack trace respectively (grid = 1 mm x 0.5 mm). ot f (a) (b) Results and discussion py Steady-state region -n Fig.5. (a) Method 1 and (b) Method 2 for determining the CTOA. For both Methods, n was set equal to 3, L1 = r1 + r0 = 0.5 mm, L2 = r2 + r0 = 1 mm, and L3 = r3 + r0 = 1.5 mm (r0 was set to 0.15 mm). co CTOA values reached a constant at crack lengths ranging from about 3 to 5.2 mm, which is 1 to 1.8 times the specimen thickness. This result is consistent with those observed by Mannucci et al. [9], Shterenlikht et al. [18], and Hashemi et al. [3,19,20]. pl e Comparison of CTOA algorithms Sa m For methods that measure CTOA directly on the fracture surface (Methods 1, 2, and 4) the scatter in the CTOA decreased with an increased measurement basis (Fig.5). For example, increase in the measurement basis r reduced the standard deviation of the CTOA data from 2.11o to 0.97o for steel #1, and from 6.57o to 2.30o for steel #3. For method 1, increase in L decreased the standard deviation of CTOA measurement from 1.52o to 0.90o for steel #1, and from 1.35o to 0.76o for steel #3. This result is attributed to factors such as: the difficulty of identifying the exact location of crack edges and the crack tip; the local deformations in regions adjacent to the apparent crack tip; and the effect associated with the longer line segments. Typically, the crack edges appear to be irregular, which is a natural result of the ductile- fracture mechanism, and when all the profile data (0.5 to 1.5 mm interval) are used to fit lines and calculate CTOA for Method 4, the result compares well with the average CTOA calculated for Method 2, as expected. Method 1 had the highest scatter in CTOA. Of the five ferrite-pearlite steels tested, only steel #1 had a standard deviation less than 1°; for the other steels, the standard deviations were between 2.2° and 4.1°. Method 1 also resulted in the highest average CTOA value for the steels. For steel #2, for example, the average CTOA value by Method 1 was 5.8° higher than that by Method 2. Method 1 depends on accurately locating the apparent crack tip, and is the most sensitive method for local deformations as well as operator judgment. Method 2 had standard deviations in CTOA measurements between 0.7° and 0.81°, and the CTOA values consistently agreed with Method 3 more than with Method 1. Method 3 had the smallest standard deviation in CTOA values (0.46° and 0.64°). Method 4 tends to track well with Method 2 results. More discussions of selecting a proper measurement basis L or r were given by Heerens et al. [36]. It can be expected that an increase of L or r may give rise to size and geometric effects. In our analysis, L or r values 206 The Journal of Pipeline Engineering Stable CTOA (o)* Standard deviation (o) Cross head mm/s Crack velocity (mm/s) 1 11.7 2.04 0.05 0.22 2 9.1 1.71 0.05 0.26 3 9.8 1.39 0.05 0.20 4 10.0 2.00 0.05 0.28 5 9.51 0.05 6 11.4 0.02 7 9.9 NA 8 NA 0.002 9 8.6 1.42 NA 10 7.8 1.9 0.02 11 8.2 2.3 12 11.9 1.3 6-X65 11.7 6-X65 0.22 is t rib ut NA NA 0.02 NA 1.2 0.002 0.004 11.4 1.2 0.02 0.044 6-X65 10.5 1.0 0.2 0.5 6-X65 11.6 2.2 5 9.2±0.6 6-X65 11.0 2.4 30 45.5±1.5 6-X65 11.2 1.1 300 594±8 6-X65 11.3 1.7 8000 6500±600 9-X100 8.6 1.1 0.002 0.008 8.3 1.8 0.02 0.088 9.3 1.1 0.2 0.66 9.4 1.0 3 6.7±0.7 ot f -n py 9-X100 co 9-X100 or d 0.02 9-X100 8.1 1.0 300 762±35 9-X100 8.8 1.6 30 118±3 9-X100 8.6 1.1 8000 7250±605 9-X100 9.8 7500 5500 10-X100 7.3 2.3 10000 13000 10-X100 10.6 5.3 20000 29000 11-X100 8.1 11-X100 8.9 2.5 8000 7000 11-X100 9.3 3.2 20000 20000 12-X70 10.2 1.8 NA 5467 m pl e 9-X100 Sa io n Steel # 2900 Table 5.The CTOA values (Method 4) calculated for the steels at various testing rates. 207 is t rib ut io n 3rd Quarter, 2013 co py -n ot f or d Fig.6. Ranking of steels by CTOA results. e Fig.7. CTOA versus crack velocity for steel #6 (X-65) and steel #9 (X-100). m pl within the range from 0.5 to 1.5 mm were chosen to calculate CTOA data. This is a reasonable compromise between the demands for minimizing scatter and possible size effects in calculating CTOA for pipeline steels. Sa CTOA ranking of the pipeline steels tested The results for the 12 steels are summarized in Table 5 and plotted in Fig.6, where all CTOA values were determined with Method 4 and averaged multiple specimen results. In Fig.6 the baintic steels (#9, #10, #11) have the lowest CTOA values and the highest strengths. For the ferrite-pearlite steels, we see that increasing strength by increasing the volume fraction of pearlite did not result in lowering the CTOA (#3 and #5 compared with #2 and #4). Not unexpectedly, reducing the grain size for strengthening (#1) resulted in significant increase in CTOA, as compared with steels with larger grain size but higher ferrite content at similar strength levels (#1 and #6 compared with #2 and #4). These data raise the question whether fine-grained ferritic steels with higher pearlite contents (or another strengthening addition) can provide an improvement in the strength to CTOA ratio. Influence of loading rate on CTOA Actuator displacement rates covering nearly seven orders of magnitude – from 0.002 mm/s to approximate 8,000 mm/s – are shown in Fig.7 for the X-65 (#6) and X-100 (#9) steels [22 and 24]. The average crack-growth velocities for the test matrix are given in Table 5. The Journal of Pipeline Engineering Fig.8. Crack velocity vs displacement rate. m pl e co py -n ot f or d is t rib ut io n 208 Sa As shown in Fig.7, the CTOA results indicate that the X-65 steel (#6) consistently had a higher resistance to cracking than the X-100 steel (#9). The CTOA for the X-65 is typically more than 2° higher than the CTOA for the X-100 throughout the range of rates evaluated. Neither steel showed a trend for CTOA with actuator or crack velocity. Additional data for X-100 steels #10 and #11 provide data at rates up to 29 m/s (Table 5), and the CTOA values range from 7.3° to 8.2° with no clear trend with velocity. The X-70 steel at higher velocity (#12) had a CTOA of 10.2, and an increased crack velocity did not result in decreased CTOA. Fig.9. Idealized slant and flat fracture-mode morphologies are illustrated on the left, and traces of actual fractures are shown on the right: (A) full slant; (B) double slant; and (C) flat. Given that the measured CTOA values are independent of the crack-tip velocities, material-specific correlations were sought between test variables (load, displacement, displacement rate) and the measured crack-tip velocity. Not surprisingly, as shown in Fig8, the displacement rate correlates extremely well with the measured cracktip velocity. However, for the same actuator rates, the X-65 specimens typically had a slightly lower crack velocity than the X-100 specimens, which indicates that the X-65 exhibits higher resistance to crack growth than does the X-100. It is surprising, however, that the correlation was found to be so valid for all 209 rib ut io n 3rd Quarter, 2013 fracture, the characteristic features of the CTOA specimens have much in common with the morphology-associated uniaxial tensile failures of ductile steels; however, in the case of the CTOA specimens, the morphology is cup-cup rather than cup-cone. The shear-oriented regions associated with the cup-cup fractures are formed by plastic flow that results in fractures with a knife-edge morphology. An important point associated with this observation is that the CTOA angle measured on the outside surface of a specimen in a flat-fracture mode is the angle formed between the two knife-edges as final fracture occurs. This is not an angle formed by the interior fracture planes; it is more like a plastic hinge. This is not the case for the slant-fracture mode, for which the fracture planes intersect the outside surface of the specimen. ot f materials tested. One may infer from Fig.8 that the crack-tip velocity is primarily a function of the far-field loading rate, for the materials and loading rates tested here. or d is t Fig.10. Cross sections of CTOD specimens of X-100 (top) and X-65 (bottom). From left to right the crosshead displacement rates applied to these specimens range between 0.002 mm/s and 8,000 mm/s. The thickness at the bottom edge of the cross sections is approximately 8 mm. -n Macroscopic fracture modes e co py The macroscopic failure mode for pipes and CTOA specimens is often described as either a flat or a slant fracture mode (Fig.9). However, mixed-mode (flat and slant) fracture morphologies are observed for both field fractures and laboratory fractures. The range of fracture modes observed in our studies for 8-mm thick MDCB specimens is shown Fig.10: the basic slant fracture occurs on a single macroscopic shear plane through the thickness of the sample, but double-slant fractures and mixed-mode fractures are not uncommon. Sa m pl Flat and mixed modes were the typical fracture modes for CTOA specimens tested at crosshead displacement rates of 300 mm/s and less (Fig.10), while at rates near 8,000 mm/s, slant fractures were typically observed. This is in agreement with fracture modes observed for full-scale, high-rate tests of the X-100. However, the fracture surface features can differ significantly, so it is useful to look a bit closer to determine whether the fracture mechanisms for laboratory fractures are representative of field fractures. Both slant and flat fracture modes have significant areas of their fracture surfaces on angles near 45° to the applied loading. The fracture-surface features on these two types of shear-oriented surface, however, indicate that they are formed by different mechanisms. For flat Details of the fracture modes Considering details for the fracture modes (Figs 11 and 12), flat fracture initiates in the centre of the specimen thickness on a plane perpendicular to the applied tensile force and grows to form an internal void. As this void grows, it effectively divides the specimen thickness into two thinner thicknesses, with lower constraint. These two thinner plates deform on shear planes until they thin down and fail in a knifeedge morphology. The ductile dimples on these thinned shear planes are characteristically elongated along the primary loading direction. But, unlike the case for cupcone shear rupture, ductile dimples on the knife-edge cup-cup fracture surfaces do not have ‘mating’ dimples on the opposing fracture surface characteristic of shear failure and void sheet coalescence. The Journal of Pipeline Engineering (a) or d is t rib ut (b) io n 210 (d) co py -n ot f (c) (e) (f) Sa m pl e Fig.11. Fracture-surface features associated with flat fracture.The overview (a) shows the ‘flat’ central portion of an X-65 fracture, bounded at both surfaces by shear regions.The central region (b and c) is a mixture of large ductile dimples, elongated in the direction of crack growth (and plate rolling), surrounded by smaller equiaxial dimples.The knife-edge final-fracture region has a shear-fracture region for a distance of about 100 μm into the specimen (d) on which shear dimples are apparent (e). There is a gradient in texture on the surface of the final fracture, with a smoother shear dimple surface near the outside edge of the specimen (e), and a more textured equiaxial dimple surface toward the centre of the specimen (f). Due to the extensive plastic flow associated with formation of the knife-edge regions for the flat-fracture morphology, the surface roughness of these regions is smooth. As shown in Fig.11d, the surface texture near the outside surface is smooth, and this region becomes smoother and extends into the specimen further with increased testing rate. This trend is evident for both the X-65 and the X-100 steels tested, and is noticeable with the unaided eye (due to the increased light reflection for the smoother surfaces). So, flat-fracture morphologies like these like these might be interpreted to some extent to determine whether fracture occurred dynamically or not, and may give some guidance on the relative rate of fracture. These results also point out that a model of ductile fracture in pipeline steels should take into consideration the true failure mode of the steel. A full-slant fracture mode results in the fracture surface on a single plane, tilted at an angle of 45° to the primary stress on the CTOA specimen. Details of the fracture features on these slant planes differ from those 211 (b) Centre or d is t rib ut (a) Edge io n 3rd Quarter, 2013 (d) Centre co py -n ot f (c) Centre (e) Centre (f) Edge m pl e Fig.12. Details of a slant-fracture mode from an X-100 CTOA specimen showing ductile dimple morphologies: (a) region very near the outside surface of specimen, (b, c, d, e) regions through the thickness, not near the final fracture regions, (f) higher magnification of a region very close to the final fracture of the knife-edge showing shear dimples. Sa for both cup-cup planes, formed with the flat-fracture morphology, and shear planes formed by cup-cone failure modes (Fig.12). The ductile dimple morphology over most of the slant surface is indicative of the ductile rupture typically observed on the interior ‘flat’ portion of the fracture that is normal to the applied load, rather than the rupture on the shear lips of the tensile specimen. Elongated shear dimples are found only very near the outside edge of the shear planes on CTOA specimens. Across most of the slant failure, dimples are typically equiaxial and have full rims. If they are elongated, the elongation is in the direction of crack growth, as is the case for the central region of flat failure modes. This indicates that mode-I loading is the primary influence. In general, evidence of shear dimple failure (mode III) is limited to regions very near the outside surface of the specimen. Crack-front shape The ‘crack tip’ measured at the outside surface of the specimen in the CTOA test is not the tip of the crack in the interior of the specimen. For example, in Fig.13 the tip of the crack front is about 1.5 mm ahead of the intersection of the crack front with the surface of the specimen. 212 The Journal of Pipeline Engineering The extent of crack-tip tunnelling varied on test specimens from about 1.5 mm to 8 mm, and the shape of the crack fronts varied from a gentle curve to an arrowhead-like shape. Occasionally crack fronts with irregular features were observed, and crack-front markings were not clear on all fracture surfaces. is t Intermittent crack growth Intermittent crack growth was sometimes observed for the X-100 specimens tested at quasi-static rates. Dark and light bands on the fracture surface mark this behaviour, as shown in Fig.14. In general, the leading edge of the crack front is coincident with the centreline of the plate (which is not always in the middle of the CTOA specimen), and intermittent crack growth does not always occur on both sides of the centreline. py -n ot f or d Fig.13. Crack-front trace as observed by markings on the fracture surface of a CTOA specimen.The regions of the specimen are (1) the outside surface, (2) the ‘flat’ portion of the fracture, (3) the ‘knife-edged’ portion of the surface, and (4) the final fracture region. rib ut io n Castings of CTOA specimens were made under loading to evaluate the 3D shapes of voids formed by tunnelling cracks in CTOA test specimens1. In the case of ‘flat’ fractures, the casting shapes indicate that the ‘flat’ fracture surfaces in the centre of the specimen thickness formed an interior CTOA of 9.2o (X-100), which compares reasonably well with the CTOA measured on the surface of the X-100 specimens. Castings also showed that the final fracture planes (knife-edge planes) on the X-100 samples were typically 45° to 50°. Sa m pl e co Fig.14. Intermittent bands of fast and slow ductile fracture. Details on the fracture surface in the banded regions show regions that resemble stair steps of quasi-ductile fast fracture followed by ductile re-blunting regions [27]. The appearance of the quasi-ductile regions is similar to the details on the surfaces of secondary cracking (splits) in the burst test fractures for this X-100 steel. Since the ‘riser’ sections of the stair step would have the same orientation as secondary cracking, this is not too surprising. In the ‘tread’ orientation, ductile dimpling is apparent and indicates more ductility for this orientation. These banded regions are observed in the higher-strength materials during CTOA testing for X-80 [37] and X-100 [27, 37]. The lower-strength grades (X-65 and X-70) did not exhibit banding. Both reports suggested that the higher-strength grades may exhibit the contrasting bands due to alternating regions of quasi-cleavage and ductile fracture. Quasi-cleavage regions are likely associated with brittle microstructural constituents associated with rolling and segregation during production of higher-strength pipes. Comparison to full-scale, high-rate test The example shown (Fig.15) for the X-100 steel that failed in a full-scale burst test has two slant-fracture regions separated by a region of flat fracture. This is Fig.15. Cross section of an X-100 steel (#9) fractured in a full-scale test. Profiles of the fracture vary with position along the length of the fractured pipe. Some regions have flat regions joined to the outside by a slant fracture. Other regions show a full slant fracture mode. 1 Castings were material used for region when the further extended, made with a polysiloxane precision-impression material, a dental casting. The material was injected into the crack-tip CTOA specimen was under load and when the crack was the casting was released. 213 or d is t rib ut io n 3rd Quarter, 2013 ot f Fig.16. Changes in the elongation of the X-100, X-65, and X-70 steels with test rate. of around 25% while the X-100 has an elongation around 20%. As testing rates increase, the percent elongation decreases slightly for all three alloys (Fig.16). The results show that the deformation of the grid is in the loading direction and the grid rectangles rotate or deform in the cracking direction. The deformation of the grid lines shows non-uniform ‘elongation’ at the ‘necked’ regions. -n something of a mixed mode, but is generally characterized as a slant fracture, because failure is essentially on a single shear plane, with no cup-cup or cup-cone shear region associated with the fracture. In addition, some regions along the full-scale tested pipe fractured in a pure slant mode, with little or no flat-fracture regions. co py For the CTOA specimens, the slant-mode failures tend to have very little or no ‘flat’ fracture. The constraint differences and higher rate of crack growth in the fullscale pipe may have influenced the observed difference in fracture mode. Sa m pl e Differences in the appearance of the fracture surfaces from the full-scale test and laboratory CTOA tests are also apparent. No laboratory tests have produced fractures that reproduce all the features observed on the full-scale, high-rate test fractures. Interestingly, some of the quasi-static test fractures have more in common with appearance of the burst-test fracture than do the highest-speed CTOA fractures compared in this study. This is because a number of the X-100 tests conducted at 0.02 mm/s had intermittent crack growth, which has regions of dynamic crack growth that might be similar to the actual fracture conditions of the full-scale test. Plastic deformation Deformation through the gauge length of the reduced section indicates that the X-65 and X-70 steels are more ductile than the X-100 steel, with quasi-static values Generally the X-100 and X-65 specimens show similar profiles for thinning due to plastic flow. The shoulder where the specimen thickness changes from 8 mm to 15 mm constrains the plastic flow, and thinning is limited in the first 6 mm or 7 mm from the shoulder, and then increases in a similar manner for all of the alloys and test rates evaluated. Both fracture modes follow the same basic trend, although the slant-shear fracture mode has less thinning during final fracture than the flat-fracture mode, which ‘necks’ during final fracture. The grids show little rotation, indicating that most of the plastic deformation is parallel to the load line. Numerical modelling The stable tearing behaviour of the CTOA test was modelled. Finite-element analysis (FEA) was applied to predict the applied load vs crack extension behaviour of steels #1 – #5, and showed correlation coefficients between the experimental and FEA results of between 0.92 and 0.993 [21]. This FEA model under-predicted the initial crack extension, when the crack extension was 214 The Journal of Pipeline Engineering io n rib ut ot f • The energy dissipation rate R reaches a minimum value in the case of slant fracture for a final tilt angle equal to 45°. This result is consistently obtained for different material hardening or damage parameters. The energy dissipation rate correlates well with CTOA values. • Stress and strain states in the stable tearing region hardly depend on the assumed tilt angle. • The CTOA on the surface of the specimen is close to the CTOA at the centre of the specimen (steady-state propagation). 1. A.B.Rothwell, 2000. Fracture propagation control for gas pipelines - past, present, and future. Pipeline Technology, 1. Elsevier, Netherlands, pp 387-405. 2. G.M.Wilkowski, Y.-Y.Wang, and D.L.Rudland, 2000. Recent efforts on characterizing propagating ductile fracture resistance of linepipe steels. Idem, pp 359-386. 3. S.H.Hashemi, I.C.Howard, J.R.Yates, R.M.Andrews, and A.M.Edwards, 2004. A single specimen CTOA test method for evaluating the crack tip opening angle in gas pipeline steels. Proc. International Pipeline Conference, pp 0610.1-7. 4. G.Demofonti, G.Mannucci, C.M.Spinelli, L.Barsanti, and H.-G.Hillenbrand, 2000. Large diameter X100 gas linepipes: fracture propagation evaluation by full-scale burst test. Pipeline Technology, 1. Elsevier, Netherlands, pp 509-520. 5. B.N.Leis, R.J.Eiber, L.Carlson, and A.Gilroy-Scott, 1998. Relationship between apparent (total) Charpy V-notch toughness and the corresponding dynamic crack propagation resistance. Proc. International Pipeline Conference, 2, pp 723-731. 6. J.C.Newman, Jr, and M.A.James, 2001. A review of the CTOA/CTOD fracture criterion – why it works! Proc. 42nd AIAA/ASME/ASCE/AH/ASC Structures, Structural Dynamics, and Materials Conference and Exhibit, paper AIAA-200-1324, Seattle, Washington, USA, pp 1042-1051. N.Pussegoda, S.Verbit, A.Dinovitzer, W.Tyson, 7. A.Glover, L.Collins, L.Carlson, and J.Beattie, 2000. Review of CTOA as a measure of ductile fracture toughness. Proc. International Pipeline Conference, 1. pp 247-254. 8. D.J.Horsley, 2003. Background to the use of CTOA for prediction of dynamic ductile fracture arrest in pipelines. Eng Fract. Mech. 70, 3-4, pp 547-552. 9. G.Mannucci, G.Buzzichelli, P.Salvini, R.J.Eiber, and L.Carlson, 2000. Ductile fracture arrest assessment in a gas transmission pipeline using CTOA. Proc. International Pipeline Conference, 1, pp 315-320. 10.R.Jones and A.B.Rothwell, 1997. Alternatives to Charpy testing for specifying pipe toughness. Fracture control in gas pipelines. WTIA/APIA/ CRC-MWJ International Seminar, pp 5-1-21. O.E.O’Donoghue, M.F.Kanninen, C.P.Leung, 11. G.Demofonti, and S.Venzi, 1997. The development and validation of a dynamic fracture propagation model for gas transmission pipelines. Int. J.Pressure Vessels Piping, 70, 1, pp 11-25. 12.N.Pussegoda, L.Malik, A.Dinovitzer, B.A.Graville, and A.B.Rothwell, 2000. An interim approach to determine dynamic ductile fracture resistance of modern high toughness pipeline steels. Proc. International Pipeline Conference, 1, pp 239-45. is t A recent model also focused on X-100 steel [26] via the computational cell technique in simulation of slant-crack advance. The dependence of crack-growth parameters on the tilt angle was systematically investigated, and a simple GTN model was used to simulate ductile damage growth within the computational cells. The main results are summarized as follows: References or d less than twice the specimen thickness, and accurately described the crack extension behaviour beyond the peak stress. In another model, the focus was on X-100 steel, and the failure parameters of the JohnsonCook and Hooputra et al. models were evaluated by parametrical computation [28]. From this work, the equivalent plastic-strain parameters of damage, for both the Johnson-Cook and Hooputra et al. models, were defined for X-100 steel. -n Future work Sa m pl e co py The difficulty in producing accurate and reproducible direct CTOA measurements on the surface of test specimens is clear from our results, and suggests the need for robust, automated measurement procedures and evaluation of indirect estimations of CTOA from load-displacement data. Furthermore, cross-head displacement has been used as a proxy for crack-mouth-opening displacement (CMOD), as the clip gauges typically used in these tests are prone to slip and lag behind when loading rates are increased. This indicates the need for improved dynamic displacement measurement which will require a novel test apparatus and specimen-preparation modifications, so that more-accurate correlations of crack location, velocity, crack-path, and crack morphology can lead to the understanding of the fracture mechanisms and their associated changes. Correlating crack-tip velocity to the applied load and far-field deformation for various pipeline steels, and methods of indirectly calculating CTOA, will likely be of interest for future FE-modelling efforts. Acknowledgments The authors thank the many guest researchers who were involved in developing and conducting CTOA tests over the years: P.P.Darcis, G.Kohn, A.Bussiba, A.Shtechman, R.Reuven, J.M.Treinen, and H.Windhoff. 3rd Quarter, 2013 215 or d is t rib ut io n 25. S.Xu, W.R.Tyson, R.Eagleson, C.N.McCowan, E.S.Drexler, J.D.McColskey, and PhP Dacis, 2010. Measurement of CTOA using MDCB and DWTT specimens. Idem, IPC2010-31076. 26.J.Besson, C.N.McCowan, and E.S.Drexler, 2013. Modeling flat to slant fracture transition using the computational cell methodology. Eng Fract. Mech., 104, pp 80-95. 27. J.W.Sowards, C.N.McCowan, and E.S.Drexler, 2012. Interpretation and significance of reverse chevronshaped markings on fracture surfaces of API X100 pipeline steels. Mat. Sci. Eng A – Struct., 551, pp 140-148. M.Szanto, C.N.McCowan, E.S.Drexler, and 28. J.D.McColskey, 2011. Fracture of X100 pipeline steel: combined experimental-numerical process. NIST BERB publication (B2011-0116). 29. R.J.Fields, J.D.McColskey, C.N.McCowan, P.P.Darcis, E.S.Drexler, S.P.Mates, and T.A.Siewert, 2012. Mechanical properties and crack behavior in line pipe steel. Final Report to the Department of Transportation. http://primis.phmsa.dot.gov/matrix/ FilGet.rdm?fil=7883. 30. E.S.Drexler, PhP.Darcis, C.N.McCowan, J.W.Sowards, J.D.McColskey, and T.A.Siewert ,2011. Ductilefracture resistance in X100 pipeline welds measured with CTOA. Weld J., 90, 12, pp 241-s - 248-s. 31.PhP.Darcis, C.N.McCowan, J.D.McColskey, and R.Fields, 2008. Crack tip opening angle measurement through a girth weld in an X100 steel pipelines. Fatigue Fract. Mater. Struct.. 31, pp 1065-1078. 32.ASTM, 1999. Standard E1290-99. Standard test method for crack tip opening displacement (CTOD) fracture toughness measurement. ASTM Book of Standards, West Conshohocken, PA, USA. 33.K.-H.Schwalbe, J.C.Newman, Jr, and J.L.Shannon, Jr, 2005. Fracture mechanics testing on specimens with low constraint–standardisation activities within ISO and ASTM. Eng Fract. Mech., 72, pp 557-576. 34.K.-H.Schwalbe, J.Heerens, U.Zerbst, H.Pisarski, and M.Kocak, 2002. EFAM-GTP 02, The GKSS test procedure for determining the fracture behaviour of materials, 2nd issue. GKSS Report 2002/24, GKSS-Forschungszentrum Geesthacht. 35.ISO/FDIS 22889, 2013. Metallic materials: method of test for the determination of resistance to stable crack extension using specimens of low constraint. J.Heerens and M.Schodel, 2003. On the 36. determination of crack tip opening angle, CTOA, using light microscopy and delta-5 measurement technique. Eng Fract. Mech., 70, 3-4, pp 417-426. 37.S.H.Hashemi, 2012. Comparative study of fracture appearance in crack tip opening angle testing of gas pipeline steels. Mat. Sci. Eng A – Struct., 558, pp 702-715. Sa m pl e co py -n ot f 13. R.M.Andrews, I.C.Howard, A.Shterenlikht,and J.R.Yates, 2002. The effective resistance of pipeline steels to running ductile fractures; modelling of laboratory test data. In: ECF14, Fracture mechanics beyond 2000. EMAS Publications, Sheffield, UK, pp 65-72. 14.D.S.Dawicke, 1996. Residual strength predictions using a CTOA criterion. Proc. FAA-NASA Symposium on Continued Airworthiness of Aircraft Structures, Atlanta, GA, USA. 15.K.-H.Schwalbe, J.C.Newman, Jr, and J.L.Shannon, Jr, 2005. Fracture mechanics testing on specimens with low constraint–standardisation activities within ISO and ASTM. Eng Fract.Mech., 72, pp 557-576. 16.ASTM ,2006. Standard E2472-06. Standard test method for determination of resistance to stable crack extension under low-constraint conditions. ASTM Book of Standards, West Conshohocken, PA, USA. 17. G.Demofonti, G. Buzzichelli, S.Venzi, and M.Kanninen, 1995. Step by step procedure for the two specimen CTOA test. Pipeline Technology, 2. Elsevier, Netherlands, pp 503-512. 18. A.Shterenlikht, S.H.Hashemi, I.C. Howard, J.R.Yates, and R.M.Andrews, 2004. A specimen for studying the resistance to ductile crack propagation in pipes. Eng Fract. Mech., 71, pp 1997-2013. 19.S.H.Hashemi, R.Gay, I.C.Howard, R.M.Andrews, and J.R.Yates, 2004. Development of a laboratory test technique for direct estimation of crack tip opening angle. Proc. 15th European Conference of Fracture, Stockholm, Sweden. 20.S.H.Hashemi, I.CHoward, J.R.Yates, R.M.Andrews, and A.M.Edwards, 2004. Experimental study of thickness and fatigue precracking influence on the CTOA toughness values of high grade gas pipeline steel. Proc. International Pipeline Conference, pp 0681.1-8. P.P.Darcis, G.Kohn, A.Bussiba, J.D.McColskey, 21. C.N.McCowan, R.Fields, R.Smith, and J.Merritt, 2006. Crack tip opening angle: measurement and modeling of fracture resistance in low and high strength pipeline steels. Idem, IPC2006-10172. 22. A. Shtechman, C.N.McCowan, R.Reuven, E.Drexler, Ph.Darcis, J.M.Treinen, R.Smith, J.Merritt, T.A.Siewert, and J.D.McColskey, 2008. Dynamic apparatus for the CTOA measurement in pipeline steels. Idem, IPC2008-64362. P.P.Darcis, C.N.McCowan, H.Windhof f, 23. J.D.McColskey, and T.A.Siewert, 2008. Crack tip opening angle optical measurement methods in five pipeline steels. Eng Fract. Mech., 75, pp 245-246. 24.R.Reuven, E.Drexler, C.McCowan, A.Shtechman, P.Darcis, M.Treinen, R.Smith, J.Merritt, T.A.Siewert, and J.D.McColskey, 2008. CTOA results for X65 and X100 pipeline steels: influence of displacement rate. Proc. International Pipeline Conference, IPC2008-64363. io n rib ut -n ot f or d is t THE PROFESSIONAL INSTITUTE OF PIPEINE ENGINEERS Sa m pl e co py The global organization for oil and gas pipeline engineers • Recognizing your skills and status • Providing a professional network • Promoting pipeline engineering Sign up today! www.pipeinst.org 3rd Quarter, 2013 217 by Prof. Claudio Ruggieri* and Leonardo L S Mathias io n Fracture-resistance testing of pipeline girth welds using bend and tensile fracture specimens rib ut Department of Naval Architecture and Ocean Engineering, University of São Paulo, São Paulo, Brazil S ot f or d is t TRUCTURAL-INTEGRITY ASSESSMENTS of pipe girth welds play a key role in the design and safe operation of pipeline systems. Current practices for structural-integrity assessments advocate the use of geometry-dependent resistance curves so that crack-tip constraint in both the test specimen and the structural component is similar. Thus, testing standards now under development to measure fracture resistance of pipeline steels often employ single-edge-notched (SE(T)) specimens under tension.This paper presents an investigation of the ductile-tearing properties for a girth weld of an API 5L X-80 pipeline steel using experimentally measured crack-growth-resistance curves (also termed J-R curves).Testing of the girthweld pipeline steels employed clamped SE(T) specimens and three-point bend (SE(B)) specimens with weld centreline cracks to determine the J-resistance curves. The experimental toughness data enables further evaluation of crack growth resistance properties of pipeline girth welds. T high-toughness steels, often undergo significant stable crack growth (a) prior to material failure. Under sustained ductile tearing of a macroscopic crack, large increases in the load-carrying capacity for the flawed structural component are possible beyond the limits given by the onset of crack-growth initiation. Simplified engineering approaches for defect assessments, such as BS7910 [4] and API579 [5] methodologies, among others, incorporate the effects of ductile tearing on crack-driving forces to evaluate the severity of crack-like flaws in structural components in terms of crack-growth resistance (J-a) curves (also often denoted J-R curves) in which the J-integral fracture parameter characterizes the significant increase in toughness over the first few millimetres of stable crack extension (Da) [1,3]. Sa m pl e co py -n HE ENGINEERING APPLICATION of fracture mechanics remains invaluable in assessments of macroscopic fracture behaviour and defect-analyses procedures of safety-critical structural components, including pressure vessels and pipeline girth welds. Fitness-for-service (FFS) fracture-assessment approaches, also referred to as engineering-critical assessment (ECA) procedures, rely upon a single parameter to define the crack-driving force and to characterize fracture resistance of the material [1-3]. These approaches provide a means for introducing acceptance criteria in cracked structural components by relating the operating conditions to a critical applied load or critical crack size. While a oneparameter description of applied loading in terms of the J-integral or the crack-tip-opening displacement (CTOD) and their corresponding macroscopic measures of fracture toughness (Jc or δc) [3] usually suffices to characterize the essentially stress-controlled failure by cleavage mode, quantitative analyses of fracture preceded and accompanied by extensive plastic deformation becomes more complex. Low-constraint and structural components containing defects and flaws, including circumferentially cracked pipelines and their weldments made of high-grade, *Corresponding author’s contact details: tel: +55 11 3091 5184 email: [email protected] Standardized techniques for crack-growth resistance testing of structural steels, including ASTM 1820 [6], routinely employ three-point bend (SE(B)) and compact-tension (C(T)) specimens containing deep, through cracks (a/W ≥ 0.45-0.50). However, a variety of crack-like defects are most often surface cracks formed during in-service operation and exposure to aggressive environment or during welding fabrication. Structural components falling into this category include girth welds made in field conditions for high-pressure piping systems and steel catenary risers (SCRs). These crack configurations generally develop low levels of crack-tip stress triaxiality which contrast sharply to The Journal of Pipeline Engineering rib ut io n 218 -n ot f or d is t (a) (b) pl e co py conditions present in deeply cracked specimens [7]. Recent defect-assessment procedures advocate the use of geometry-dependent fracture-toughness values so that crack-tip constraint in both test specimen and structural component is similar. In particular, fracture-toughness values measured using single-edge-notch tension (SE(T)) specimens appear more applicable for characterizing the fracture resistance of pressurized pipelines and cylindrical vessels than standard, deep-notch, fracture specimens under predominantly bend loading [8-10]. Sa m Recent applications of SE(T) fracture specimens to characterize crack-growth-resistance properties in pipeline steels [11] have been effective in providing larger flaw tolerances while, at the same time, reducing the otherwise excessive conservatism which arises when measuring the material’s fracture toughness based on high constraint, deeply-cracked SE(B) or C(T) specimens. However, some difficulties associated with SE(T) testing procedures raise concerns about the significance and qualification of measured crack-growth-resistance curves. While slightly more conservative, testing of shallow-crack-bend specimen configurations may become more attractive due to its simpler testing protocol and laboratory procedures, and the much smaller loads required to propagate the Fig.1. (a) Partial unloading during the evolution of load with displacement; (b) definition of the plastic area under the load-displacement curve. crack. Although deeply-cracked SE(B) specimens are the preferred crack geometry often adopted in conventional defect-assessment methods, recent revisions of ASTM 1820 [6] and ISO 15653 [12] have also included J-estimation equations applicable to shallow-crack-bend specimens. Consequently, use of smaller specimens which yet guarantee adequate levels of crack-tip constraint to measure the material’s fracture toughness becomes an attractive alternative. This work presents an experimental investigation of the ductile-tearing properties for a girth weld made of an API 5L X-80 pipeline steel in terms of crackgrowth-resistance curves. Use of these materials is motivated by the increasing demand in the number of applications for manufacturing high -strength pipes for the oil and gas industry. Testing of the pipeline girth welds employed side-grooved, clamped SE(T) specimens and three-point-bend SE(B) specimens with a weld-centreline notch and varying crack sizes to determine the crack-growth-resistance curves using the unloading compliance (UC) method and a single specimen technique. Recently developed compliance functions and η-factors applicable for SE(T) and SE(B) fracture specimens are introduced 3rd Quarter, 2013 219 to determine crack -growth -resistance curves from laboratory measurements of load-displacement records. Overall, these experimental results provide toughness data which enable further evaluation of crack-growthresistance properties of pipeline girth welds. standards such as ASTM E1820 [6]) to evaluate J with crack extension follows from an incremental procedure which updates Je and Jp at each partial unloading point, denoted k, during the measurement of the load vs displacement curve depicted in Fig.1(b) as: Experimental evaluation of J-resistance curves k J= J ek + J pk (1) io n where the current elastic term is simply given by: rib ut K2 J ek = I (2) E ′ k is t and the current plastic term follows an incremental formulation which is applicable to CMOD data in the form [14-16] in Equn 3 below, in which the geometry factor γLLD is evaluated from: or d bk −1 dη Jk −−1LLD k −1 γ LLD = −1 + η Jk −−1LLD − (4) k −1 Wη J − LLD d ( a W ) In the above expressions, KI is the elastic stress intensity factor for the cracked configuration, Ap represents the plastic area under the load-displacement curve, BN is the net specimen thickness at the side groove roots (BN = B if the specimen has no side grooves, where B is the specimen gross thickness), and b denotes the uncracked ligament (b = W - a, where W is the specimen width and a is the crack length). In the above Eqn 2, plane-strain conditions are adopted such that E’ = E/(1-ν)2, where E and ν are the (longitudinal) elastic modulus and Poisson’s ratio, respectively. The factor ηJ in Eqns 3 and 4 represents a non-dimensional parameter which relates the plastic-deformation contribution to the strain energy for the cracked body and J. Figure 1a illustrates the essential features of the estimation procedure for the plastic component Jp, where Ap (and consequently ηJ) is defined in terms of load-LLD (or Δ) data or load-CMOD (or V) data. These geometry factors are denoted ηJ-LLD and ηJ-CMOD, respectively, when LLD or CMOD are used. ot f Conventional testing programmes to measure crack-growth resistance (J-Δa) curves in metallic materials routinely employ the UC method based on a single specimen test. The estimation procedure used in ASTM E1820 standard [6] predominantly employs load-line displacement (LLD) records to measure fracture-toughness-resistance data incorporating a crack-growth correction for the J-integral. An alternative method which potentially simplifies the test procedure involves the use of crack-mouth-opening displacement (CMOD) to determine both the J-integral and the amount of crack growth. This section provides a brief overview of the analytical framework needed to evaluate data for common fracture specimens from laboratory measurements of load-displacement records. Attention is directed to an incremental procedure to obtain estimates of J and crack length for the SE(T) and SE(B) configurations based on CMOD data. Incremental estimation procedure of the J-integral m pl e co py -n The procedure to estimate crack-growth-resistance data considers the elastic and plastic contributions to the strain energy for a cracked body under Mode I deformation [3] so that the J-integral conveniently derives from its elastic component, Je, and plastic term, Jp, as J = Je + Jp. Here, an estimation scheme for the plastic component employs a plastic η-factor introduced by Sumpter and Turner [13] to relate the crack-driving force to the plastic area under the load versus LLD (or CMOD) for cracked configurations (see also Refs [14-16]) as illustrated in Fig.1(a). The procedure to estimate Jp based on the η-methodology has proven highly effective in testing protocols to measure fracture toughness in stationary cracks while, at the same time, retaining strong contact with the deformation plasticity definition of J. Sa However, the area under the actual load-displacement curve for a growing crack differs significantly from the corresponding area for a stationary crack (upon which the deformation definition of J is based) [3]. Consequently, the measured load-displacement records must be corrected for crack extension to obtain an accurate estimate of J-values with increased crack growth. A widely used approach (which forms the basis of current The incremental expression for Jp defined by Eqn 3 coupled with Eqn 4 contains two contributions: one is from the plastic work in terms of CMOD and, hence ηJ-CMOD, and the other due to crack growth correction in terms of LLD by means of ηJ-LLD; evaluation of Eqns 3 and 4 is relatively straightforward provided that these two geometric factors are known. For the clamped SE(T) specimens with H/W = 10 and the conventional SE(B) specimen with S/W = 4 utilized in this study, γ k −1 η k −1 = J pk J pk −1 + J −CMOD ( Apk − Apk −1 ) 1 − LLD ( ak − ak −1 ) (3) bk −1 BN bk −1 220 The Journal of Pipeline Engineering −1 Be= B io n By measuring the instantaneous compliance during unloading of the specimen (see Fig.1b), the current crack length follows directly from solving the functional dependence of crack length and specimen compliance in terms of μCMOD. For the clamped SE(T) and SE(B) specimens analysed here, the corresponding compliance expressions are given by Cravero and Ruggieri [14] and Appendix X.2 of ASTM E1820 [6] respectively as: a 1.921 − 13.219µ + 58.708µ 2 − 155.282 µ 3 W = SET (12) + 207.399 µ 4 − 107.917 µ 5 0.1 ≤ a / W ≤ 0.7 a 1.019 − 4.537 µ + 9.01µ 2 − 27.333µ 3 W = (13) SEB ot f Current testing protocols to measure the crack-growthresistance response using a single-specimen test are primarily based on the unloading compliance (UC) technique to obtain accurate estimates of the current crack length from the specimen compliance measured at periodic unloadings with increased deformation. Figure 1b illustrates the essential features of the method. The slope of the load-displacement curve during the k-th unloading defines the current specimen compliance, denoted Ck, which depends on specimen geometry and crack length. For the clamped SE(T) specimen with H/W = 10 and the SE(B) specimen with S/W = 4 analysed here, the specimen compliance is often defined in terms of normalized quantities expressed as [6, 14]: 2 rib ut Crack-length estimation ( B − BN ) (11) B− is t Equations 7 and 8 applicable to SE(B) specimens agree very well with the η-factors in the revised J-integral expressions developed by Zhu et al. [19] which form the basis of current ASTM E1820 [6] and ISO 15653 [12] standards using CMOD records. CCMOD denotes the specimen compliance in terms of CMOD (CCMOD = V/P) and the effective thickness, Be, is defined by: or d a convenient polynomial fitting of the results given by Cravero and Ruggieri [14], Donato and Ruggieri [17], and Ruggieri [18] provides the corresponding η-factor equations for homogeneous materials in the form shown in Equns 5-8 below. + 74.4µ 4 − 71.489µ 5 0.05 ≤ a/W ≤ 0.45 Effect of weld strength overmatch on plastic η-factors and Current test standards employ J-estimation expressions which are mainly applicable to fracture specimens made of homogeneous materials. For a given specimen geometry, mismatch between the weld metal and base-plate strength affects the macroscopic mechanical behaviour of the specimen in terms of its load-displacement response, with a potentially strong impact on the coupling relationship between J and the near-tip stress fields. −1 (10) py EWBeCCMOD SEB µCMOD = 1 + S 4 -n SET µCMOD = 1 + EBeCCMOD (9) e co where μSETCMOD and μSEBCMOD define the normalized compliances for the SE(T) and SE(B) specimens. In the above expressions, E is the longitudinal elastic modulus, 2 3 4 a a a a a + 7.81 − 18.27 + 15.30 − 3.08 W W W W W pl m η JSET = 1.07 − 1.77 −CMOD 5 (5) Sa 0.2 ≤ a / W ≤ 0.7 η JSET − LLD = −0.62 + 9.34 2 3 4 a a a a a − 4.58 − 47.96 + 87.70 − 44.88 W W W W W 5 (6) 0.2 ≤ a / W ≤ 0.7 2 a a η = 3.65 − 2.11 + 0.34 (7) W W 0.1 ≤ a / W ≤ 0.7 SEB J −CMOD 2 3 4 5 a a a a a η = 0.02 + 18.09 − 73.26 + 152.22 − 159.777 + 66.88 (8) W W W W W SEB J − LLD 0.1 ≤ a / W ≤ 0.7 221 rib ut io n 3rd Quarter, 2013 where σMBys and σWMys denote the yield stresses for the base-plate metal and the weld metal. ot f Accurate estimation formulas for J more applicable to welded fracture specimens may become important in robust defect-assessment procedures capable of including effects of weld-strength mismatch on fracture toughness. or d is t Fig.2. Geometry of tested fracture specimens with weld-centreline notch and B x B cross section: (a) clamped SE(T) specimen with a/W = 0.4 and H/W = 10; (b) three-point SE (B) specimen with a/W = 0.25 and S/W = 4. Material description and welding procedures co py -n Previous work by Donato et al. [20], and Paredes and Ruggieri [21], introduced a functional dependence of geometry factor ηJ-CMOD on crack size and weld-strength mismatch for weld centreline fracture specimens. The expressions of factor ηJ-CMOD for the weld cracked SE(B) with S/W = 4 and SE(T) with H/W = 10 are summarized as follows: Experimental details = 3.88 + 0.22 η JSEB −CMOD 3 a a − 5.01 W W 2 (14) e a + 4.02 − 0.41M y − 0.05M y2 W m pl 0.2 ≤ a/W ≤ 0.7, 1.0 ≤ My ≤ 1.5 Sa η JSET −CMOD = −0.36 + 11.69 a a − 23.59 W W 3 2 (15) a + 13.90 − 0.28M y − 0.03M y2 W 0.2 ≤ a/W ≤ 0.7, 1.0 ≤ My ≤ 1.5 In the above expressions, the mismatch ratio, My, is defined as: My = σ WM ys σ ysMB (16) The material used in this study was a high-strength, low-alloy (HSLA), API grade X-80 pipeline steel produced as a base plate using a control-rolled processing route without accelerated cooling. The mechanical properties and strength/toughness combination for this material are mainly obtained by both grain-size refinement and second-phase strengthening due to the small-size precipitates in the matrix. A 20-in diameter pipe with longitudinal seam weld from which the girth weld SE(T) and SE(B) specimens were extracted was fabricated using the UOE process. The tested weld joint was made from the API X-80 UOE pipe having wall thickness tw = 19 mm. Girth welding of the pipe was performed using the FCAW process in the 1G (flat) position with a single V-groove configuration in which the root pass was made by GMAW welding. The main weld parameters used for preparation of the test weld using the FCAW process are: (i) n umber of passes = 12 (including the root pass made by the GMAW process); (ii) welding current = 165 A; (iii)welding voltage = 23 V; (iv) average heat input = 1.5 kJ/mm. The Journal of Pipeline Engineering rib ut io n 222 Sa m pl e co py -n ot f or d is t Fig.3. Measured load-CMOD curve for the tested X-80 pipeline girth weld using clamped SE(T) specimens with a/W = 0.4 and three-point SE(B) specimen with a/W = 0.25. Mathias et al. [22] provide the tensile properties for the tested pipeline girth weld and base material which include: σWMys = 715MPa σWMuts = 750 MPa σBMys = 609 MPa σBMuts = 679 MPa. Here, σys and σuts represent the material’s yield stress and tensile strength, and WM and BM denote the weld metal and the base plate. Fig.4. J-resistance curves for the clamped SE(T) specimens with a/W = 0.4 and H/W = 10. Specimen geometries Mathias [23] conducted unloading compliance tests at room temperature on weld-centreline-notched SE(T) specimens with fixed-grip loading to measure tearing resistance curves in terms of J – Δa data. The clamped SE(T) specimens have a fixed overall geometry and crack length to width ratio defined by a/W = 0.4, H/W = 10 with thickness B = 14.8 mm, width W = 14.8 mm (W = B) and clamp distance H = 148 mm (refer to Fig.2a). Here, a is the crack depth and W is the 223 is t rib ut io n 3rd Quarter, 2013 mode on the J – Δa data. We first draw attention to the load-carrying capacity for the bend and tension configurations. Figure 3 shows a typical load-CMOD curve measured from tests of the SE(T) specimen with H/W = 10 and a/W = 4, and the SE(B) specimen with S/W = 4 and a/W = 0.25. The strong effect of loading mode (tension vs bending) associated with specimen geometry is evident in this plot. At similar levels of CMOD, the applied load for the SE(T) specimen increases approximately by a factor of four compared to the load response for the SE(B) specimen. co py -n ot f specimen width which is slightly smaller than the pipe wall thickness, tw. The UC tests at room temperature were also conducted on weld-centreline-notched SE(B) specimens with a/W = 0.25, specimen thickness B = 14.8 mm, width W = 14.8 mm (B = W) and span S = 4W (refer to Fig.2b). Conducted as part of a collaborative research programme at the University of São Paulo on structural-integrity assessment of marine steel-catenary risers (SCRs), testing of these specimens focused on the development of accurate procedures to evaluate crackgrowth resistance data for pipeline girth welds. Mathias et al. [22] provided more details on the materials used and fracture testing of the X-80 pipeline girth weld. or d Fig.5. J-resistance curves for the three-point SE(B) specimens with a/W = 0.25. Sa m pl e All specimens, including the SE(T) configuration, were pre-cracked in three-point bend conditions. After fatigue pre-cracking, the specimens were side-grooved to a net thickness of approx. 85% the overall thickness (7.5% side-groove on each side) to promote uniform crack growth along the crack front, and tested following the general guidelines described in ASTM E1820 [6]. Records of load vs CMOD were obtained for the specimens using a clip gauge mounted on knife edges attached to the specimen surface. Crack-growth-resistance curves Effect of specimen geometry on J-resistance curves The framework for determining J-resistance curves based on CMOD from conventional fracture specimens described previously provides the basis for evaluating the ductile-fracture response of the tested material and assessing effects of specimen geometry and loading Evaluation of the crack-growth-resistance curve follows from determining J and Δa at each unloading point of the measured load-displacement data based upon the previous formulations for the η-factors and compliance functions. Figures 4 and 5 compare the measured crack-growth-resistance curves for the SE(T) and SE(B) specimens: it can be seen that the resistance curves for the shallow-crack SE(B) specimen are comparable to the J-R curves corresponding to the deeply-cracked SE(T) specimen. Here, the average J-values at fixed amounts of crack growth, Δa, for both crack configurations are reasonably similar. Another significant features associated with these plots include: • The average slope of the J-R curve for the SE(B) specimen (which is also related to the material’s tearing modulus) is slightly higher than the corresponding slope for the SE(T) specimen. • The estimated value of the J-integral at onset of ductile tearing, JIc, is fairly independent of specimen geometry and loading mode. The Journal of Pipeline Engineering rib ut io n 224 Configuration Measured post test Compliance estimation ot f Specimen or d is t Fig.6. J-resistance curves for the SE(T) specimen and SE(B) specimen based upon η-factors for homogeneous materials and overmatched welds. Deviation (%) a0 (mm) af (mm) Δa (mm) af (mm) Δa (mm) 8.79 3.13 8.77 3.11 0.78 8.66 2.55 8.56 2.45 3.87 6.29 9.32 3.03 9.20 2.92 3.75 6.70 10.59 3.89 10.59 3.89 0.03 SE(T) H/W=10 5.66 SET2 H10 SE(T) H/W=10 6.11 SET3 H10 SE(T) H/W=10 SET4 H10 SE(T) H/W=10 SEB1 SE(B) a/W=0.25 4.38 6.28 1.90 5.89 1.39 26.84 SE(B) a/W=0.25 4.99 7.00 2.01 6.38 1.48 26.37 SE(B) a/W=0.25 4.48 6.65 2.17 6.12 1.78 17.97 SEB4 SE(B) a/W=0.25 3.75 6.50 2.75 5.84 1.94 29.45 SEB5 SE(B) a/W=0.25 3.93 6.25 2.32 5.30 1.65 28.88 py m pl e SEB3 co SEB2 -n SET1 H10 Sa Table 1. Predicted and measured crack extension for tested fracture specimens. Unfortunately, the measured resistance curves are perhaps somewhat more scattered than we would expect for these specimens, particularly for the bend configuration. While we did not thoroughly investigate such behaviour, the crack-front measurements conducted by Mathias [23] revealed a somewhat highly uneven crack advance, thus providing some explanation for the scatter of the measured resistance curves. However, it is evident that the J-resistance data for the SE(B) configuration compare relatively well with the SE(T) specimen results. Effect of weld strength overmatch on J-resistance curves The effect of weld-strength mismatch on the fractureresistance response characterized by the J - Δa data is examined here for the tested SE(T) and SE(B) specimens with weld-centreline notch. The primary interest is to assess the potential deviation that arises from evaluating the J-resistance curves using η equations developed for homogeneous materials. 225 Figure 6 compares the J-resistance curves for the shallow-crack SE(B) specimen and deep-crack SE(T) specimen based on η-factors for homogeneous materials and overmatched welds, as represented by open and solid symbols. The η-values for the overmatch condition are determined from using the estimation Eqns 14 and 15 provided previously, with My = 1.18 (refer to Eqn 16). To facilitate comparison, only the lowest and highest resistance curves for these crack configurations are included in the plot. It can be seen that the fracture-resistance curves derived from η-factors for overmatched welds are practically indistinguishable from the curves evaluated with η-factors for homogeneous materials. Here, the use of η-factors for homogeneous materials (i.e. not taking into account the degree of weld strength overmatch) leads to slightly non-conservative (higher) estimates of the resistance curve (we should emphasize that the larger the levels of weld-strength mismatch, the larger the degree of non-conservativeness). or d is t rib ut (a) Crack-length estimates (b) py -n ot f After testing, all specimens were subjected to heattinting treatment (300°C for 30 min), and then air cooled before being broken apart. Following standard methods based on the nine-point average technique, such as the procedure given by ASTM E1820 [6], the initial and final crack length measured after the test by means of an optical method were compared with crack length estimates derived from the UC method. Table 1 provides the predicted and measured crack extension for all tested fracture specimens, where the deviation is defined as: io n 3rd Quarter, 2013 co Δameasured – Δapredicted/ Δameasured (c) e The significant features that emerge from these results include: Sa m pl • predictions of crack extension based on the UC procedure for the SE(T) specimens are in close agreement with experimental measurements with a level of accuracy of approx 5%; • crack-extension predictions for the shallow-crack SE(B) configuration derived from the UC procedure are not in good agreement with the measured amount of ductile tearing; here, the UC method underestimates the nine-point average crack extension by approx. 25-30%, which produces apparent higher J-resistance curves. The crack-growth behaviour for the shallow-crack SE(B) configuration can be explained in terms of the uneven crack advance and a rather irregular crack-front profile observed in these specimens. Figure 7 shows typical crack surfaces for the SE(B) and SE(T) configurations: (d) Fig.7.Typical fracture surfaces of tested crack configurations: (a, b) three-point SE(B) specimen with a/W = 0.25; (c, d) clamped SE(T) specimen with a/W = 0.4. 226 The Journal of Pipeline Engineering co py -n This study described an experimental investigation of the ductile-tearing properties for a girth weld made of an API 5L X-80 pipeline steel and experimentally measured crack growth resistance curves (J – Δa curves). Testing of the pipeline girth welds used side-grooved, clamped SE(T) specimens and shallow-crack-bend SE(B) specimens with a weld-centreline notch to determine the crack-growth-resistance curves based upon the unloading compliance (UC) method using a single specimen technique. The work described here supports the following conclusions: m pl e • Shallow-crack SE(B) specimens (a/W = 0.25) provide crack-growth-resistance curves which are comparable to the J-resistance curves for deepcrack SE(T) specimens (a/W = 0.4). Despite the relatively larger scatter of the J – Δa data, the fracture resistance for the shallowcrack SE(B) configuration at a fixed amount of crack growth, Δa, is relatively similar to the corresponding fracture resistance for the SE(T) specimen. • Levels of weld-strength overmatch within the range of 10-20% (approx.) overmatch do not significantly affect J-resistance curves derived from using η-values applicable to homogeneous materials. While the fracture resistance curves based on η-values for homogeneous materials are slightly higher than those based on η-factors for overmatched weldments, differences are nevertheless small and within acceptable limits. Sa io n rib ut ot f Concluding remarks While the analyses described here clearly provide support for using shallow-crack-bend configurations as an alternative fracture specimen to measure crack-growth properties for pipeline girth welds and similar structural components, they are also suggestive of the need for more experimental studies to validate the UC-based procedure for estimating J-resistance curves of SE(B) configurations. In particular, more-accurate techniques for crack-length estimations in small-size-bend specimens appear central to develop a robust and efficient J-resistance evaluation procedure. Additional work is in progress along these lines of investigation. is t The UC procedure described previously to estimate the current crack length involves the assumption of a straight crack front. Consequently, the compliance equations described by Eqns 12 and 13 should be viewed as idealized solutions providing estimates for the average crack extension. Moreover, the inaccurate estimate of crack extension resulting from these analyses is also suggestive of a strong effect of the bend-loading mode on crack-length predictions. Indeed, previous studies [24, 25] have already indicated that use of the UC method with three-point-bend specimens underestimates crack extension when compared with optically measured values of crack length; this effect appears to be more pronounced for SE(B) specimen with a shallow crack. • Crack-extension predictions based on the UC procedure agree well with experimental measurements for the SE(T) specimens. In contrast, the unloading-compliance method underestimates the nine-point average crack extension for the shallow-crack SE(B) specimen by 25-30% (approx.). This rather strong under-prediction of crack extension for this crack configuration produces apparent higher J-resistance curves and, at the same time, underlies some limitations of current UC estimation equations to predict crack length in small-sized-bend specimens. or d it can be seen that the bend specimens exhibited a highly non-uniform fatigue pre-crack compared to the SE(T) specimens. Such a feature could be caused by unexpected misalignment between the specimen and the rollers, thereby affecting the crack-tip stresses and strains driving the ductile-fracture process. However, crack tunnelling is less pronounced for the SE(B) configuration than for the SE(T) specimen after some amount of ductile tearing. Acknowledgments This investigation is primarily supported by Fundação de Amparo à Pesquisa do Estado de São Paulo (FAPESP) through Grant 2009/54229-3 and by Agência Nacional de Petróleo, Gás Natural e Biocombustíveis (ANP). The work of CR is also supported by the Brazilian Council for Scientific and Technological Development (CNPq) through Grants 304132∕2009-8 and 476581∕20095. The authors acknowledge Tenaris-Confab Brasil and Lincoln Electric Brasil for providing support for the experiments described in this work. References 1. J.W.Hutchinson, 1983. Fundamentals of the phenomenological theory of nonlinear fracture mechanics.’ J. Applied Mechanics, 50, pp 1042-1051. 2. U.Zerbst, R.A.Ainsworth, and K.-H.Schwalbe, 2000. Basic principles of analytical flaw assessment methods. Int. J. Pressure Vessels and Piping, 77, pp 855-867. 3. T.L.Anderson, 2005. Fracture mechanics: fundaments and applications. 3rd Edn, CRC Press, New York. 4. British Standard Institution, 2005. Guide to methods for assessing the acceptability of flaws in metallic structures, BS7910. 5. American Petroleum Institute, 2007. API Standard 579-1/ASME FFS-1, Fitness-for-service, 2nd Edn. 6. ASTM, 2011. ASTM E1820: Standard test method for measurement of fracture toughness 3rd Quarter, 2013 227 or d is t rib ut io n 17.G.H.B.Donato and C.Ruggieri, 2006. Estimation procedure for J and CTOD fracture parameters using three point bend specimens. Proc. International Pipeline Conference (IPC 2006), American Society of Mechanical Engineers, Calgary, Canada. 18.C.Ruggieri, 2012. Further results in J and CTOD estimation procedures for SE(T) fracture specimens - Part I: homogeneous materials. Eng Fract. Mech., 79, pp 245-265. 19. X.-K.Zhu, B.N.Leis, and J.A.Joyce, 2008. Experimental estimation of J-R curves from load-CMOD record for SE(B) Specimens. J. ASTM International, 5, JAI 101532. G.H.B.Donato, R.Magnabosco, and C.Ruggieri, 20. 2009. Effects of weld strength mismatch on J and CTOD estimation procedure for SE(B) specimens. Int. J. Fracture, 159, pp 1-20. 21.M.Paredes and C.Ruggieri, 2012. Further results in J and CTOD estimation procedures for SE(T) fracture specimens - Part II: center cracked welds. Eng Fract. Mech., 89, pp 24-39. 22. L.L.S.Mathias, D.F.B.Sarzosa, and C.Ruggieri. Effects of specimen geometry and loading mode on crack growth resistance curves of a high-strength pipeline girth weld. Int. J. Pressure Vessels and Piping (submitted for publication). 23.L.L.S.Mathias, 2013. Experimental evaluation of J-R curves in X80 pipeline girth welds using SE(T) and SE(B) fracture specimens. Master’s Thesis, Polytechnic School, University of São Paulo (in Portuguese). 24.P.A.J.M.Steekamp, 1988. J-R curve testing of threepoint bend specimen by the unloading compliance method. Fracture Mechanics: 18th Symposium, ASTM STP 945, American Society for Testing and Materials, Philadelphia, pp 583-610. 25.J.Dzugan, 2003. Crack length calculation by the unloading compliance technique for Charpy size specimens. Wissenschaftlich-Technische Berichte FZR-385, Forschungszentrum Rossendorf (FZR). Sa m pl e co py -n ot f 7. G.H.B.Donato and C.Ruggieri, 2008. Constraint effects and crack driving forces in surface cracked pipes subjected to reeling. Proc. ASME International Conference on Offshore Mechanics and Arctic Engineering (OMAE 2008), American Society of Mechanical Engineers, Lisbon, Portugal. 8. B.Nyhus, M.Polanco, and O.Orjasæter, 2003. SENT specimens as an alternative to SENB specimens for fracture mechanics testing of pipelines. Proc. ASME International Conference on Offshore Mechanics and Arctic Engineering (OMAE 2003), American Society of Mechanical Engineers, Cancun, Mexico. 9. S.Cravero and C.Ruggieri, 2005. Correlation of fracture behavior in high pressure pipelines with axial flaws using constraint designed test specimens - Part I: Plane-strain analyses. Eng Fract. Mech., 72, pp 1344-1360. 10. L.A.L.Silva, S.Cravero, and C.Ruggieri, 2006. Correlation of fracture behavior in high pressure pipelines with axial flaws using constraint designed test specimens - Part II: 3-D effects on constraint. Idem, 73, pp 2123-2138. 11.D.Y.Park, W.R.Tyson, J.A.Gianetto, G.Shen, and R.S.Eagleso, 2010. Evaluation of fracture toughness of X100 pipe steel using SE(B) and clamped SE(T) single specimens. Proc. International Pipeline Conference, Calgary, Canada. 12. International Organization for Standardization, 2010. Metallic materials − method of test for the determination of quasistatic fracture toughness of welds, ISO 15653. 13.J.D.G.Sumpter and C.E.Turner, 1976. Method for laboratory determination of Jc. Cracks and Fracture, ASTM STP601, American Society for Testing and Materials, pp 3-18. 14. S.Cravero and C.Ruggieri, 2007. Estimation procedure of J-resistance curves for SE(T) fracture specimens using unloading compliance. Eng Fract. Mech., 74, pp 2735-2757. 15.Ibid., 2007. Further developments in J evaluation procedure for growing cracks based on LLD and CMOD data. Int J. Fracture, 148, pp 387-400. 16.X.-K.Zhu and J.A.Joyce, 2012. Review of fracture toughness (G, K, J, CTOD, CTOA) testing and standardization. Eng Fract. Mech., 85, pp 1-46. io n rib ut is t or d Houston, February 10–13, 2014 -n ot f . . r e e f n n c o e C Exhibition s e s r u Co its 26 year, the PPIM Conference is recogniz ed as Now entering tional forum for sharing and learning ab a n r e t n i t s o m e r out best the fo aintenance and condition-monitoring te m e m i t e if l n i s e chnology practic for natural gas, crude oil and product pipelines. www.clarion.org Sa m pl e co py th The international gathering of the global pigging industry! Organized by 3rd Quarter, 2013 229 is t Tubular Goods Research Institute, CNPC, Xi’an, China rib ut by Dr He Li*, Qiang Chi, Jiming Zhang,Yang Li, Lingkang Ji, and Chunyong Huo io n Fracture-toughness evaluations by different test methods for the Chinese Second West-East gas transmission X-80 pipeline steels F -n ot f or d RACTURE RESISTANCE DETERMINATION was one of the most important aspects during the research and development of the Chinese 2nd West-East X-80 pipeline steels. More than 30 kinds of longitudinal submerged-arc-welded (LSAW) X-80 pipes were tested using the 2-mm striker Charpy V-notch (CVN) test, the 8-mm striker CVN test, and the instrumented drop-weight tear test (DWTT). It was found that the threshold energy was about 200 J for the 2-mm striker and 8-mm striker CVN tests. Below this threshold, the difference between the 2-mm striker CVN energy and the 8-mm striker CVN energy was small. Above the threshold, however, the difference between the 2-mm striker CVN energy and the 8-mm striker CVN energy increases as the CVN energy increases. It was also found that there is a linear relation between the DWTT energy and the 8-mm striker CVN energy when the latter is lower than 400 J/cm2. Otherwise, the DWTT energy trends to reach a plateau. I e co py N ORDER TO MEET the sharply increased gas consumption in China, high-grade X-80 linepipes have been used for the second West-East gas pipeline (WEGP). During the research and development of Chinese X-80 linepipe steels for the WEGP, fractureresistance determination was been taken as one of the most important issues. Sa m pl World-wide, Charpy V-notch (CVN) impact tests are undertaken as the standard method to determine the fracture resistance, and ASTM E23 [1] and ISO 148 [2] have been issued by American Society for Testing and Materials (ASTM) and International Organization for Standardization (ISO), respectively. In China, the national standard GB/T 229 [3] has been widely used with procedures and specifications similar to ISO 148. Unfortunately, there are two different specimen striker/hammer designs: ASTM E23 adopts the Charpy striker with a tup radius of 8 mm, while ISO 148 and *Corresponding author’s contact details: tel: +86 29 8872 6155 email: [email protected] GB/T 229 adopt the Charpy striker with a tup radius of 2 mm. Consequently, the equivalence of the test results is questionable. Some previous studies showed that there is an increasing CVN energy difference between the 2-mm striker and the 8-mm striker after a threshold energy. The threshold energy appears to be material-dependent and is related to the fracture characteristics of the material [4, 5]. On the other hand, the CVN specimens were considered by some researchers to be too small to develop significant propagation and therefore might be inappropriate to quantify fracture resistance for today’s linepipes [6]. This leads to the adoption of the drop-weight tear test (DWTT) energy as an alternative fracture resistance [7]. However, the choice of DWTT or CVN energy for fracture control of pipelines is still controversial. In the present study, a series of tests were performed for the 2nd WEGP X-80 pipeline steels. The energy difference determined by the 2-mm and 8-mm striker CVN tests, as well as the energy difference determined by the 8-mm striker CVN test and instrumented DWTT, are discussed and clarified. 230 The Journal of Pipeline Engineering Nb+V Ceq + Ti C Si Mn P S Mo Cu Ti Nb Requirements of 2nd WEGP, maximum 0.09 0.42 1.85 0.022 0.005 0.35 0.3 0.025 0.11 – – 0.23 Experimental materials, average 0.05 0.24 1.76 0.012 0.001 0.28 0.23 0.01 0.073 0.106 0.46 0.19 Pcm io n Element Table 1.The chemical composition of 2nd WEGP X-80 pipeline steels (wt %). ASTM E23-06 rib ut GB/T 229-2007 ISO148 Allowable deviation (mm) Nominal dimension (mm) Allowable deviation (mm) Length (L) 55 ±0.6 55 +0, -2.5 Height (H) 10 ±0.075 10 ±0.075 Width (W) 10 ±0.11 Angle of notch (o) 45 ±2 Height below notch 8 ±0.075 45 ±1 ±0.075 – – ±0.025 0.25 ±0.025 ±0.42 – – ±2° ot f 90° ±2° 90° ±2° 90° ±10’ Distance of plane of symmetry of notch from ends of test piece 27.5 Angle between plane of symmetry of notch and longitudinal axis of test piece 90° -n 0.25 or d 10 Radius of curvature at base of notch Angle between adjacent longitudinal faces of test piece is t Nominal dimension (mm) pl e co py Table 2. Dimensional requirements of Charpy V-notch specimens according to GB/T 229-2007 and ASTM E23-06. (b) m (a) Sa Fig.1. Photos of (a) CVN specimen and (b) pressed-notch DWTT specimen. Materials and experimental procedures The 2nd WEGP longitudinal submerged-arc welded (LSAW) X-80 pipes with an outside diameter of 1219 mm (48 in) and wall thickness of 18.4 mm were used as the experimental materials. The average chemical composition of the linepipe material is shown in Table 1. Initially, the materials were tested by a 750-J CVN machine using 2-mm and 8-mm impact strikers according to GB/T 299 and ASTM E23, respectively. Among these Charpy impact tests, three sets were carried out at temperatures of 20oC, 0oC, -10oC, -20oC, -40oC, and -60 oC, while other tests were carried out at room temperature (approx. 20oC).Secondly, the materials were tested by a 50,000-J instrumented DWTT machine at room temperature (approx. 20oC). The load-displacement curves were recorded during the tests. The CVN specimens had a size of 55 mm x 10 mm x 10 mm with a 2-mm machined notch, as shown in Fig.1a. 3rd Quarter, 2013 231 CVN energy (J) Specimen 1 Temperature (˚C) Specimen 2 Specimen 3 8-mm striker 2mm striker 8mm striker 2mm striker 8mm striker 20 305 400 263 389 302 431 0 300 420 270 396 295 -10 295 385 250 349 288 -20 290 350 208 278 -40 295 356 207 189 -60 285 330 98 128 io n 2-mm striker 396 rib ut 415 340 165 217 171 158 is t 291 or d Table 3. Comparison between CVN energy determined by the 2-mm and 8-mm strikers for specimens #1, #2, and #3. ot f The DWTT specimens had a size of 305 mm x 76.2 mm x 18.4 mm with a 5-mm pressed notch, as shown in Fig.1b. The DWTT specimens can be seen to be much larger than the CVN specimens. py -n Table 2 shows the dimension requirements by GB/T 229-2007 (or ISO148) and ASTM E23-06, and Fig.2 shows the schematic of the 2-mm and 8-mm strikers. As shown in Table 2, the Charpy test standards GB/T 229-2007 (or ISO148) and ASTM E23 have similar size requirements; the most significant difference is the pendulum striker radius, as shown in Fig.2. co The relationship between the 2-mm and 8-mm striker CVN energy Sa m pl e Table 3 lists the CVN energy tested by the 2-mm and 8-mm strikers at six different temperatures for specimens #1, #2, and #3. Hereafter, KV2 is used to indicate the CVN energy tested by the 2-mm striker, while KV8 is used to indicate the CVN energy tested by the 8-mm striker. Figure 3 shows that for specimen #1, the KV2 values at different temperatures are almost equal (approx. 290 J), while the KV8 (approx. 400 J) values at higher temperatures (0oC and 20oC) are obviously higher than those (approx. 340 J) at lower temperatures (-60oC, -40oC, and -2oC). And, KV8 is higher than KV2 at all test temperatures from -60oC to +20oC. For specimens #2 and #3, the trend of CVN change with temperature is similar. At lower temperatures (-60oC and -40oC), the KV2 and KV8 are almost equal, while at higher temperatures (-10oC, 0oC, and 20oC), KV8 is much higher than KV2. It is noted that the KV8 and KV2 values have the similar ductile-brittle transition behaviour, and both KV8 and KV2 reach the upper shelf energy at temperatures above -10oC. (a) (b) Fig.2. Schematic of Charpy impact strikers: (a) 2-mm striker used in ISO148; and (b) 8-mm striker used in ASTM E23. The Journal of Pipeline Engineering rib ut io n 232 py -n ot f or d is t (a) Sa m pl e co (b) (c) Fig.3. Average CVN energy versus temperature for (a) specimen #1, (b) specimen #2, and (c) specimen #3, tested by the 2-mm and 8-mm strikers. 233 rib ut io n 3rd Quarter, 2013 co py -n ot f or d is t Fig.4. CVN energy determined by the 8-mm striker vs CVN energy determined by the 2-mm striker. pl e Fig.5. DWTT total energy versus CVN energy (KV8). Sa m Figure 4 shows the relationship between KV2 and KV8 for the 30 specimens tested at room temperature in the present study. A general equivalence between the KV2 and KV8 is observed for the CVN energy up to about 200 J, after which KV8 is greater than KV2 with the difference increasing with increasing energy. The present results are consistent with Naniwa’s observations [8] on carbon and low-alloy steels. The energy of 200 J as the threshold was also reported in Ref.9. Nevertheless, other thresholds such as 60 J [10] and 100 J [11] have been reported for different steels. So, it is reasonable to assume that the threshold energy is material-dependent. Li’s study [12] showed that the KV8 is almost the same as the KV2 if the specimens were broken into two parts; otherwise, the KV8 is significantly larger than KV2 if the specimens were unbroken. Similar observations have been found in the present study: (a) when both KV2 and KV8 were below 200 J, all of the specimens could be completely broken; (b) when both KV2 and KV8 were above 200 J, the specimen occasionally could not be completely broken; and (c) when the KV2 is above 300 J, the specimen frequently could not be completely broken. It is believed that the energy fraction of specimen bending and plastic deformation will increase with the total energy increasing, and consequently the large deformation causes the specimen not to be broken. Meanwhile, the 8-mm striker will cause a much higher bending and plastic energy dissipation due to the larger The Journal of Pipeline Engineering is t rib ut io n 234 edge radius. So, it is not surprising that the KV8 has a higher value and the difference between KV8 and KV2 increases with increasing energy/temperature. Once the KV8 energy is greater than 400 J/cm2, the CVN specimens cannot frequently be completely broken in the test, and Fig.7 shows the comparison between the unbroken specimens and the fully broken specimens. In contrast, all of the DWTT specimens could be fully broken. So, when the KV8 energy is greater than 400 J/cm2, the major energy dissipation in the CVN specimens might not be the crack-propagation energy, but the crack-initiation energy, plastic-deformation energy, and friction dissipation between the specimen and the anvils. It is believed that the real crack-propagation resistance may not increase with increasing KV8. On the other hand, since the fracture-propagation energy is still the major part of the DWTT total energy, the plateau in the DWTT total energy occurs at very high energies for the Chinese X-80 steels. It is noted that this observation is subject to further verification. ot f Relationship between the 8-mm striker CVN energy and the DWTT energy or d Fig.6. DWTT propagation energy versus CVN energy (KV8). Sa m pl e co py -n 25 specimens were tested by both the 8-mm striker CVN test and the instrumented DWTT with the standard pressed notch. As shown in Fig.5, the DWTT total energy has a good linear relation with the KV8 energy for the KV8 below 400 J/cm2. The linear relationship can be expressed as DWTTtotal = 2.3 CVN + 63.04 (J/cm2), which is almost the same as Wilkowski’s result of DWTTtotal = 2.53 CVN + 63.04 (J/cm2) [7]. In Wilkowski’s study, the maximum value of KV8 is below 320 J/cm2, and thus only a linear relation between the DWTT total energy and the KV8 energy was obtained. In contrast, the present study has a maximum KV8 of 571 J/cm2, and so the initial linear relationship, an overall non-linear relationship, and a final plateau when the KV8 energy is above 400 J/cm2, are observed. Furthermore, the DWTT-propagation energy is calculated as the area under the load-displacement curve by integration between the maximum load and a load of 0.3 times the maximum load. As shown in Fig.6, a good linear relation is again observed and can be expressed as DWTTpropagation = 0.9 CVN + 30 for KV8 below 400 J/cm2. The slope in Fig.6 is lower than that in Fig.5; it seems that the propagation energy fraction in the DWTT total energy decreases with KV8 increasing, and thus the DWTT deformation energy increases with the increase of the total Charpy energy. Conclusion More than 30 kinds of X80 LSAW pipe steels made for the Chinese 2nd WEGP were experimentally investigated by different test methods using the 2-mm striker CVN test, the 8-mm striker CVN test, and the instrumented DWTT. From the tests and discussions presented in this work, the following conclusions are obtained: • The differences between KV2 and KV8 can be ignored for Charpy impact energy below 200 J. However, when the impact energy is greater than 200 J, the differences between KV2 and KV8 increase with increasing energy, and this is usually accompanied by the increase of unbroken specimen occurrence. • Both KV2 and KV8 have the similar change trends with the change of test temperature. 235 is t rib ut io n 3rd Quarter, 2013 co py -n ot f or d (a) Fig.7. Photos of fracture appearance for (a) unbroken CVN specimens, and (b) fully broken CVN specimens. m pl e (b) Sa • The DWTT total energy has a linear relationship with the KV8 for the KV8 energy values less below 400 J/cm2, and is described by DWTTtotal = 2.3 CVN + 63.04 (J/cm2). However, for KV8 energies above 400 J/cm2, the DWTT total energy trends to reach a plateau. • The DWTT propagation energy has a linear relationship with the KV8 for the KV8 energy values below 400 J/cm2 and is described by DWTTpropagation = 0.9 CVN + 30 (J/cm2). For KV8 energies above 400 J/cm2, the DWTT propagation energy also trends to reach a plateau. Acknowledgments This project was supported by Natural Science Basic Research Plan in Shaanxi Province of China (Program No. 2011JQ6017) and by China National Petroleum Cooperation (CNPC). References 1. ASTM E23-06. Standard test methods for notched bar impact testing of metallic materials. 2. ISO 148-1:2006. Metallic materials – Charpy pendulum impact test method. 236 The Journal of Pipeline Engineering rib ut io n 8. T.Naniwa, M.Shibaike, et al., 1990. Effects of the striking edge radius on the Charpy impact test. ASTM STP 1072, pp 67-80. 9. C.N.McCowan, J.Pauwels, G.Revise, and H.Nakano, 2000. International comparison of impact verification programs. ASTM STP 1380, pp 73-89. 10.O.L.Towers, 1993. Effects of striker geometry on Charpy results. Metal Construction, 15, 11, pp 682686. 11.T.A.Siewert and D.P.Vigliotti, 1995. The effect of Charpy V-notch striker radius on the absorbed energy. ASTM STP 1248, pp 140-152. 12.H.P.Li, X.Zhou, and W.C.Xu, 2011. Correlation between Charpy absorbed energy using 2mm and 8mm striker [J]. J.ASTM International, 8, 9, 1-4. Sa m pl e co py -n ot f or d is t 3. GB/T 229-2007. Metallic materials – Charpy pendulum impact test method. 4. R.K.Nansta and M.A.Sokolov, 1995. Charpy impact test results on five materials and NIST verification specimens using instrumented 2-mm and 8-mm strikers. ASTM STP 1248, pp 111-139. 5. M.Tanaka, Y.Ohno, H.Horigome, et al., 19995. Effects of the striking edge radius and asymmetrical strikes on Charpy impact test results. Idem, pp 153-167. 6. G.M.Wilkowski, W.A.Maxey, and R.J.Eiber, 1978. What does the Charpy test really tell us? American Society for Metals, pp 201-225. 7. G.M.Wilkowski, D.L.Rudland, H.Xu, and N.Sanderson, 2006. Effect of grade on ductile fracture arrest criteria for gas pipelines. Proc. International Pipeline Conference, Calgary, Paper No: IPC2006-10350. 3rd Quarter, 2013 237 CTOD and pipelines: the past, present, and future by Dr Philippa Moore* and Dr Henryk Pisarski io n TWI Ltd, Cambridge, UK C -n ot f or d is t rib ut RACK-TIP-OPENING displacement (or CTOD) has been the most widely used fracture-toughness parameter within the oil and gas industry for nearly 50 years. Originally developed from research at TWI in the UK during the 1960s, CTOD was an ideal parameter for characterizing the fracture toughness of medium-strength carbon-manganese steels used in pressure vessels, offshore platforms, and pipelines where the application of linear-elastic fracture mechanics was insufficient to account for their ductility. Once fracture-toughness testing (CTOD testing) became standardized within BS 7448, ASTM E1290, ISO 12135, and ISO 15653, the CTOD concept enjoyed an established international reputation.The development of standardized fitness-for-service assessment procedures, initially through the use of the CTOD design curve, and then to use of the failure-analysis diagram approach described in BS 7910, also allowed CTOD to be used directly to determine tolerable flaw sizes to assess the structural integrity of welds. In more recent times, single-edge-notched tension specimen (SENT) testing has been enthusiastically adopted by the pipeline industry in place of the traditional single-edge-notched bend (SENB) specimen used for standard CTOD tests. However, currently there is no national standard describing SENT testing, although this is being developed. SENT testing is particularly advantageous when pipeline girth welds are subjected to plastic straining, and a number of assessment procedures based on CTOD have been and are being developed to define strain capacity and flaw-acceptance criteria. C Sa m pl e co py RACK-TIP-OPENING DISPLACEMENT (or CTOD) has been the most widely used fracture toughness parameter within the oil and gas industry for nearly 50 years. Originally developed from research at TWI in the UK during the 1960s, CTOD was an ideal parameter for characterizing the fracture toughness of medium-strength carbon-manganese steels used in the manufacture of pressure vessels where the application of linear-elastic fracture mechanics was insufficient to account for their ductility. The development of North Sea oil and gas from the 1970s onwards hastened the application of CTOD testing and analysis concepts for application to the construction of steel-jacket production platforms, and pipelines. The fracture-toughness testing of single-edge-notched bend specimens (or the ‘CTOD test’ as it is sometimes called) is the standard method to measure it. However, as further progress is made in the development of fracture mechanics, both testing and assessment, the CTOD concept must change and adapt to keep up. To illustrate the story of CTOD over the years we have highlighted some of the key people at TWI who have been involved in this work; however it *Corresponding author’s contact details: tel: +44 1223 899000 email: [email protected] must be recognized that the research and development of this field has been the collaboration of a much larger group of engineers from many institutions worldwide. The origins of the CTOD concept Fracture mechanics as an engineering discipline was conceived just after the Second World War as a result of the Liberty Ship fractures. Of 2700 ships fabricated using the new welding technology during the war, around 400 had fractures, 90 were considered serious, and about ten ships fractured completely in half [1]. This failure rate had driven the US Naval Research Labs to research the effect of cracks in steels, and by the 1950s it had developed the linear-elastic fracture mechanics’ (LEFM) description of cracks in brittle materials in work led by Dr George L Irwin [2]. However, the stress-based LEFM did not sufficiently describe the behaviour of more ductile materials, such as medium-strength structural steels. The UK had chosen to begin its own investigations into brittle-fracture issues after the War, driven by the UK Admiralty Advisory Committee on Structural Steels, who held conferences at the University of Cambridge in 1945 and 1959 attended by many of those who would become eminent in the fields of structural engineering, metallurgy, and fracture mechanics, including George 238 The Journal of Pipeline Engineering io n In the case of the CTOD test, the specimen size is usually representative of the full material thickness. The CTOD test piece originally had a saw-cut notch but later used fatigue pre-cracking to produce a sharp notch [5]. The crack mouth is instrumented with a clip gauge to measure the crack-mouth opening, and then loaded under quasi-static three-point bending to enable a load versus crack-mouth opening displacement trace to be plotted (Figs 2 and 3). Fig.1. Alan Wells at TWI, UK, in the 1990s. Sa m pl e co py -n ot f Irwin and Dr Alan Wells from the British Welding Research Association [3]. Alan Wells (Fig.1) had taken a sabbatical at the US Naval Research Lab in 1954 and worked with George Irwin in that time. After his return to the UK and the British Welding Research Association (BWRA), which later became TWI, Alan Wells proposed an alternative model of fracture to LEFM in 1961 [4]. Wells developed the crack-opening displacement (COD), later the crack-tip-opening displacement (CTOD), model of fracture mechanics from an observation of the movement of the crack faces apart during plastic deformation of notched test pieces. He showed that fracture would take place at a critical value of COD, and for calculations below general yield, this was proportional to the square of the critical stress intensity factor divided by the yield strength. Furthermore, he showed that the critical value of COD determined in bend specimens and wide-plate specimens (representing structural components) of the same thickness were equivalent. Thus he was able to demonstrate transferability of fracture toughness determined from test specimens to other structural geometries. This was to have far-reaching implications on the development of fitness-for-service concepts for welded structures for the avoidance of fracture. As a result of this, the CTOD parameter was used extensively in the UK for elastic-plastic fracture mechanics (EPFM) analysis of welded structures from the 1960s especially once the development of North Sea oil reserves in the 1970s was driving much of the fracture research at that time [1]. or d is t rib ut As confidence grew in the ability of the small-scale CTOD test to predict the fracture conditions of a crack in a full-scale structure, the test method became standardized. BSI published a Draft for Development (DD19) on applied force for CTOD testing in 1972. This became a standard in 1979 as BS 5762 [6], which described ‘Methods for crack opening displacement (COD) testing’. This was then superseded by BS 7448 Part 1 in 1991 [7] as a ‘Method for determination of KIC, critical CTOD and critical J values of metallic materials’. Part 2 of BS 7448 provided an equivalent method for welds in metallic materials when it was published in 1997 and was largely based on TWI’s experience in testing welds. This was developed further by an ISO committee, with TWI representation by Henryk Pisarski (Fig.4), and in 2010 BS 7448 Part 2 was superseded by BS EN ISO 15653 [8], although Part 1 still remains current for CTOD testing of parent metals. Standard methods for CTOD testing In essence, the fracture-toughness test specimen comprises a rectangular bar of material that is notched into the appropriate region (with respect to a welded joint). Once fracture-toughness testing became standardized within BS 7448 [7], ASTM E1290 [9], E1820 [10], ISO 12135 [11], and ISO 15653 [8], the CTOD concept enjoyed an established international reputation. CTOD had been established as the fracture-toughness parameter for the oil and gas industry so thoroughly, that often the phrase ‘CTOD test’ has been used interchangeably with the more precise ‘fracture-mechanics’ test’ by those in that industry. Definition of CTOD In the days of development of the CTOD testing standard BS 5762 within TWI, both Mike Dawes (Fig.5) and Alan Wells put forward formulae to determine CTOD from the test result, based respectively on either the load and crack-mouth opening, or the crack-mouth opening alone. The Dawes approach [12] which combined separate elastic and plastic components of the crack-tip-opening displacement was that which was ultimately adopted by the British Standard, and in the early editions of ASTM E1290. The equation to determine CTOD from bend specimens in the current fracture-toughness-testing standards ISO 12135 and BS 7448 Part 1 comprises an elastic component and a plastic component which are added together. The elastic part is based on the applied force 239 rib ut io n 3rd Quarter, 2013 co py Fig.3. Example of a load vs crack-mouth opening displacement (CMOD) trace measured during a fracture-mechanics’ test (‘CTOD test’) of a single-edge-notched bend specimen. -n ot f or d is t Fig.2. Fracture-mechanics’ test (‘CTOD test’) of a single-edge-notched bend specimen instrumented with a double-clip gauge. Sa m pl e (F) and a function of initial crack length to specimen width ratio (a0/W), as well as the specimen dimensions, while the plastic component is determined using the plastic component of the clip-gauge displacement (Vp) and the height of the clip gauge above the crack mouth (z), in addition to specimen dimensions. The current CTOD equation in ISO 12135 and BS 7448 Part 1 for bend specimens is given in the following formula: 2 S F a (1 −ν ) CTOD or δ = × f 0 0.5 W 2σ YS E W ( BBNW ) 0.4 (W − a0 )V p + 0.4W + 0.6a0 + z 2 (1) where S is the loading span of the specimen, W is the specimen width, B is the thickness, BN is the net section thickness (accounting for side grooving), ν is Poisson’s ratio, and E is Young’s modulus. The CTOD can also be related to the stress-intensity factor, KI, using the following formula, given in BS 7910 [13]: CTOD or δ = K I2 (2) mσ YS E ' where E’ is the Young’s modulus under plain strain conditions, equal to E/(1-ν2), and m is a geometric and material factor which is usually between 1 and 2. In the new editions of BS 7910 which is due to be published in 2013, a more-precise value of m is given which is a function of the ratio of material yield to tensile strength for deeply notched specimens. The crack-tip-opening displacement can be conceptually understood as the amount that a crack tip needs to be opened up (or the distance the crack faces need to be moved apart) before unstable propagation of the crack occurs (Fig.6a). However, several alternatives have The Journal of Pipeline Engineering is t rib ut io n 240 Fig.5. Mike Dawes at TWI in the 1980s. been put forward as to how to exactly define CTOD, illustrated in Fig.6. For materials with lower ductility, the original definition (Fig.6a) is fine, but for moreductile materials, unstable fracture may not occur and the CTOD is determined from the point of maximum load in the test: the CTOD can be considered the opening displacement of the deformed crack at the tip position of the original crack (Fig.6b). A third definition is that most often used when performing numerical models of cracks, which defines the CTOD as the displacement at the points where a 90° angle at the crack tip intersects with the crack sides (Fig.6c). to be used to more accurately determine tolerable flaw sizes to assess the structural integrity of welds. Published standard fitness-for-service (FFS) assessment procedures also cemented the power of fracture-mechanics’ testing. Successful experience using the CTOD concept to determine tolerable flaw sizes over almost a decade led to the development of an FFS Published Document (PD 6493) by the British Standards Institution. Initially published as PD 6493:1980 [17] ‘Guidance on methods for assessing the acceptability of flaws in fusion welded structures’, the procedure was revised in 1991 and subsequently became a standard, BS 7910 [13], in 1999. -n ot f or d Fig.4. Henryk Pisarski at TWI in 2013. py Fitness-for-service assessment Sa m pl e co One of the earliest codified applications of the CTOD concept was to provide alternative flaw-acceptance criteria to those in Appendix A of API 1104 [15] in the 1970s. The Welding Institute’s CTOD design curve was developed by Mike Dawes (Fig.5) and colleagues, at a time when very little guidance was available on the application of elastic-plastic fracture-mechanics’ (EPFM) analyses to common materials, particularly to welded structures with high residual stresses and stress concentrations. The CTOD design curve [16] was intended to provide a logical, simple, and rapid means of determining the allowable crack sizes in welded structures subjected to normal design loads. The importance of CTOD was that it could be used directly to calculate the maximum tolerable flaw size for a given weld. The CTOD design curve approach was intended to be used as a first, coarse, filter in fitness-for-purpose assessments. The subsequent development of standardized fitness-for-service assessment procedures, initially through the use of the CTOD design curve and then to the use of the failure-analysis diagram approach, allowed CTOD FFS assessment allowed larger flaws to be shown as tolerable, for example in offshore platforms and pipelines, compared to the small flaw sizes permitted by applying ‘workmanship’ flaw limitations imposed by welding standards. At TWI, John Harrison (Fig.7) had performed numerous FFS assessments using both the CTOD design curve and PD6493, and demonstrated the effectiveness of these methods to industry, particularly for oil and gas [18]. He became heavily involved in the standardization of PD 6493 into BS 7910. These FFS methods were used for a number of offshore installations in the 1970s and 1980s where it had been necessary to demonstrate avoidance of brittle fracture in as-welded joints in thicker section (40-120 mm). Similar to pressurevessel practice, highly stressed welds for operation at sub-zero temperatures would normally be subjected to post-weld heat treatment (PWHT) when section thicknesses exceeded 40 mm, and avoiding PWHT needed this kind of rigorous justification. These issues were addressed in greater detail in the UK Department of Energy’s Guidance Notes on the design, construction, and certification of offshore structures, which was first published in 1990 [19]. This document had a large influence on the design of steel jackets operating in the North Sea. The sections 241 (a) (b) (c) io n 3rd Quarter, 2013 rib ut Fig.6. Some definitions of CTOD [14]: (a) early idealization; (b) CTOD at original crack-tip position; (c) CTOD at positions subtending 90° at crack tip. ot f or d is t dealing with toughness requirements for steels, avoidance of brittle fracture, and post-weld heat treatment were largely based on work conducted by John Harrison and Henryk Pisarski [20], and utilized concepts described in PD 6493. Where problems had arisen with flaws having been detected and repair strategies needing to be decided, an FFS assessment could justify whether repair without subsequent PWHT was acceptable or not. With the standard methods, performing an FFS assessment became a regular part of the preparation for any new pipeline installation in order to set the fabrication flawacceptance criteria. SENT testing co py -n The desire to extend the application of pipelines to ever-more-challenging loading conditions during both during installation and operation has been driving the development of fitness-for-service methods to reduce their over-conservatism while remaining confident that the structure will be safe. Pipelines intended for deeper water, higher pressure, or installation methods or upset conditions (such as ground movement) involving strains beyond yield, impose greater challenges to integrity. Sa m pl e The development of fracture-toughness testing had traditionally been through the use of the deeplynotched bend (SENB) tests, which are intended to impose a high degree of crack-tip constraint, and hence provide a lower-bound estimate of fracture toughness. Recognition that flaws in pipeline girth welds are subjected to lower crack-tip constraint has led to the development of the single-edge-notch tension (SENT) test which has lower constraint than the SENB specimen. Collaboration between Henryk Pisarski (Fig.6) at TWI in the UK and SINTEF and DNV in Norway, along with a group of industry partners, produced guidance for fracture control for pipeline-installation methods introducing cyclic plastic strain, which became DNV’s Recommended Practice DNV-RP-F108 in 2006 [21]. The method intended for pipe installation by methods such as reeling used SENT specimens to measure fracture toughness from a notched specimen whose constraint more closely matched that of a flaw in a girth weld. Fig.7. John Harrison at TWI in 2007. The method used multiple-specimens to produce a J-R curve which was then used in an assessment procedure based on BS7910 to generate flaw-acceptance criteria. The higher value of fracture toughness that can be obtained from a SENT specimen compared to a SENB specimen has led to a rapid growth of interest in using these specimens for other fracture-mechanics’ assessment and testing. SENT testing has been enthusiastically adopted by the offshore pipeline industry (for example, in DNV-OS-F101) and gradually been accepted by the pipeline industry in general. However, the initial testing procedures for SENT specimens gave the fracture toughness exclusively in terms of J-R curves rather than CTOD. Although developed for pipeline installation, there is growing evidence that the biaxial-loading experience by pipelines during operation may also exhibit similar ductiletearing resistance as the R-curve measured from SENT testing [22], making SENT specimens appropriate for analysis of pipeline operation as well. In its Appendix A, the DNV-OS-F101 standard for submarine pipeline The Journal of Pipeline Engineering rib ut io n 242 or d CTODmat ≤ 0.1 mm; X = 1.8 0.1 mm < CTODmat < 0.4 mm; X = 1.9 - CTODmat CTODmat ≥ 0.4 mm; X = 1.5 CTODmat = -n CTOD from SENT tests to be iteratively improved using the following clauses to determine the parameter X. ot f systems [23] describes a fracture-mechanics’ method to determine the acceptability of flaws. The 2012 edition requires the fracture-toughness testing to be performed on SENT specimens, and expresses the fracture-toughness requirements in terms of J. It allows the fracture toughness to be expressed as CTOD only if the procedure for calculating CTOD is demonstrated to be conservative. is t Fig.8. Comparison of methods to determine CTOD from J relative to double-clip measurements of CTOD for SENT specimens. m pl e co py Despite its growing popularity, there is currently no standardized procedure for carrying out SENT tests, or for determining CTOD in a SENT specimen. This gap in CTOD knowledge led to a study by TWI to validate methods for determining CTOD in the SENT specimen [24]. The work compared direct measurements of CTOD made by silicone-crack infiltration to a finite-element model of the SENT specimen to predict CTOD, and the double-clip gauge method to determine CTOD using the equivalent triangles rule. These methods were also compared to CTOD from J equations in recent literature to improve upon the over-conservative equation given in the 2010 edition of DNV-OS-F101 [25]. Sa Using silicone-rubber crack infiltration allowed direct measurement of CTOD to be made from replicas of the SENT specimen notch, although the method is not practical for routine testing. FEA models also give a reliable way to determine CTOD, but require too much analytical processing to be practical for determining CTOD for every fracture-mechanics’ test. These methods were used to compare the effectiveness of other simpler methods for calculating CTOD. DNV-OS-F101:2012 [23] now also gives a revised method to calculate CTODmat from Jmat (Equn 3), which uses the material yield strength (YS) and an estimate of CTOD J mat X .σ YS (1 −ν 2 ) (3) Both the FEA model and the crack-infiltration methods agreed fairly well with the double-clip method, giving confidence that the double-clip method can give reliable values of CTOD for SENT specimens with a/W ratios of between 0.3 and 0.5. When the equations for calculating CTOD from J for SENT specimens were compared against the double-clip (Fig.8), the one given by Shen and Tyson [26] offers the best alternative method to calculate CTOD from J compared to the over-conservative approach given in DNV OS F101 from 2010. The Moreira and Donato method [27] may show more benefit when applied to weld specimens, not just parent metals. When the newer method from DNV-OS-F101 from 2012 is also included, the improvement in accuracy of the new formula can be seen, but seems to be still consistently slightly over-conservative. The intention of this comparison was to provide confidence in the value of CTOD that is determined when using SENT-test specimens, so that the validation of the SENT-specimen approach keeps pace with the pipeline industry’s need to continue to define fracture toughness in terms of CTOD, while using the most modern test methods. 3rd Quarter, 2013 243 io n rib ut ot f Nevertheless, the CTOD fracture-toughness parameter is being withdrawn from the fitness-for-service assessment standard BS7910 in the 2013 edition, which will directly use only K to determine the fracture ratio. However, the intuitive understanding of CTOD as fracture toughness by a wide range of industries will mean that it will continue to be used to describe fracture toughness for some time to come. Indeed, there is current discussion on whether J or CTOD are the appropriate fracturecharacterizing parameters when considering assessments in the plastic regime. For example, the strain-based assessment procedure used by ExxonMobil uses CTOD [26]. Pipeline-assessment research and development continues towards strain-based methods and the CTOD parameter is well suited for this application. is t Further research into SENT testing with the intention of developing a complete and thorough testing standard is being carried out at TWI as part of a Group Sponsored Project, while research on SENT testing is also active in the US, Brazil, and Canada. The authors of this paper are involved in the British Standards’ committee which will eventually publish the SENT testing standard as BS 8571 in the near future. 5. P.Houldcroft, 1996. Fifty years of service to industry – a brief and occasionally lighthearted history of BWRA and The Welding Institute. TWI. 6. BS 5762:1979. Methods for crack opening displacement (COD) testing. British Standards Institution. 7. BS 7448-1:1991. Fracture mechanics toughness tests: Part 1: Method for determination of Kic, critical CTOD and critical J values of metallic materials. Idem. 8. BS EN ISO 15653:2010. Metallic materials - method of test for the determination of quasistatic fracture toughness of welds. Idem. 9. ASTM, 2008. ASTM E1290-08e1: Standard test method for crack-tip opening displacement (CTOD) fracture toughness measurement. (Withdrawn 2013), American Society for Testing and Materials. 10.Ibid., 2011. ASTM E1820-11e1: Standard test method for measurement of fracture toughness. Idem. 11.ISO 12135:2002, 2008. Metallic materials - unified method of test for the determination of quasistatic fracture toughness. International Standards Organization. 12.M.G.Dawes, 1979. Elastic-plastic fracture toughness based on COD and J-contour integral concepts. In: Elastic-plastic fracture. ASTM STP 668, American Society for Testing and Materials, pp 307-333. 13.BS 7910:2005. Guide to methods for assessing the acceptability of flaws in metallic structures. British Standards Institution. 14.J.D.Harrison, 1980. The “state of the art” in crack tip opening displacement (CTOD) testing and analysis. TWI Members Report 108/1980, April. 15.API 1104:1973. Standard for welding pipelines and related facilities. 13th Edn (superseded), American Petroleum Institute. 16.F.M.Burdekin and M.G.Dawes, 1971. Practical use of linear elastic and yielding fracture mechanics with particular reference to pressure vessels. In: Proc. Institution of Mechanical Engineers Conference on Practical Application of Fracture Mechanics to Pressure Vessel Technology, London, 3-5 April, pp 28-37. 17.PD 6493:1980. Guidance on some methods for the derivation of acceptance levels for defects in fusion welded joints (superseded). British Standards Institution. J.D.Harrison, M.G.Dawes, G.L.Archer, and 18. M.S.Kamath, 1979. The COD approach and its application to welded structures. In: Elastic-plastic fracture, ASTM STP 668, American Society for Testing and Materials, pp 606-631. UK Department of Energy, 1984 and 1990. 19. Standard: Offshore installations: guidance on design, construction and certification. 3rd and 4th Edns, HMSO, London. 20.J.D.Harrison and H.G.Pisarski, 1986. Background to new guidance on structural steel and steel construction standards in offshore structures. HMSO, London. or d The future py -n TWI is proud of its history and involvement in the story of CTOD, and of individual colleagues whose work in those early years allowed CTOD to become such a versatile fracture parameter, well established within the oil and gas industry and beyond. The present is a time of further changes in both fracture-mechanics’ testing and assessment and standardization, but offers a bright future for further research and development in this field, not just at TWI, but worldwide. co Acknowledgements pl e The authors are grateful to Dr Mike Dawes for providing a photograph and some biography text for this paper, despite his retirement. m References Sa 1. T.L.Anderson, 1995. Fracture mechanics – fundamentals and applications. 2nd Edn, CRC Press. 2. G.Irwin, 1957. Analysis of stresses and strains near the end of a crack traversing a plate. J. Applied Mechanics, 24, pp 361-364. 3. J.Knott, 1997. From CODs to CODES (the realisation of fracture mechanics in the UK): Fracture research in retrospect – an anniversary volume in honour of R.George Irwin’s 90th Birthday, Ed. H.P.Rossmanith, A.A.Balkema, Netherlands, ISBN 9054106794. 4. A.A.Wells, 1961. Unstable crack propagation in metals: cleavage and fast fracture. Proc. Crack Propagation Symposium, Cranfield, UK, 2, p 210. 244 The Journal of Pipeline Engineering or d is t rib ut io n Dr Henryk Pisarski is a TWI Technology Fellow and works in the Structural Integrity Technology Group of TWI. He has been with TWI since 1973. His interests include the application of fracture-mechanics’ based assessment and testing methods to assure the integrity of welded structures with respect to fracture avoidance and to demonstrate fitness-for-service. He is also involved in the development of strain-based procedures for the assessment of pipeline girth welds. He has managed projects applying fracture-mechanics’ testing and assessment methods to a wide range of engineering structures including ships, offshore structures, subsea components, pipelines, and pressure vessels. In addition, he has contributed to standards’ bodies on fracture-toughness testing and flaw assessment (ISO 15653, BS 7448, and BS 7910). He has published a number of papers on these subjects and also carried out expert-witness work. Currently, he is the UK delegate to IIW Commission X (Structural performance of welded joints – fracture avoidance) and is contributing to the revision of BS 7910 (flaw assessment). Dr Mike G Dawes was involved in the development of the CTOD approach to fracture avoidance at TWI from 1968 until he retired in 2000. By that time his CTOD design curve and fracture-toughness test method relationships were used in all the corresponding BSI, ASTM, and ISO standards. Much of his work, including his PhD thesis, was concerned with applications to welded-metal structures. This included development of the local compression treatment, which for the first time enabled acceptable shapes of fatigue pre-cracks to be obtained in fracture-toughness-test specimens extracted from as-welded joints. In 1993, in recognition of his work on international standards, he received the ASTM’s highest award, the Award of Merit, and the title of Honorary Fellow. -n ot f 21.DNV-RP-F108:2006. Recommended Practice F108: Fracture control for pipeline installation methods introducing cyclic plastic strain. January. 22. K.A.MacDonald, 2011. Fracture and fatigue of welded joints and structures. Chapter 2 Constraint fracture mechanics: test methods. By K.A.MacDonald, E.Ostby, and B.Nyhus, Woodhead Publishing Ltd. 23. DNV–OS-F101:2012. Offshore Standard F101: Submarine pipeline systems. August. 24.P.Moore and H.Pisarski, 201. Validation of methods to determine CTOD from SENT specimens. In: Proc. ISOPE-2012, The 22nd International Offshore (Ocean) and Polar Engineering Conference, Rhodes, Greece, 17-22 June. 25.DNV-OS-F101:2010. Offshore Standard: Submarine pipeline systems (superseded). 26.G.Shen and W.R.Tyson, 2009. Evaluation of CTOD from J-integral for SE(T) specimens. In: Proc. Pipeline Technology Conference, Ostend, 12-14 October. 27. F.Moreira and G.Donato, 2010. Estimation procedures for J and CTOD fracture parameters experimental evaluation using homogeneous and mismatched clamped SE(T) specimens. In: Proc.ASME 2010 Pressure Vessels and Piping Conference. PVP2010, Bellevue, Washington, USA, 18-22 July. 28.H.Tang, M.Macia, K.Minnaar, P.Gioielli, S.Kibey, and D.Fairchild, 2010. Development of the SENT test for strain-based design of welded pipelines. In: Proc. 8th International Pipeline Conference, Calgary, Canada, 27 Sept.-1 Oct. Appendix – Biographies of TWI engineers cited Sa m pl e co py Dr Alan Wells died at the end of 2005. By the time he formally retired from TWI in the summer of 1988, Alan Wells had notched up 25 years at TWI’s site in Great Abington, the last 11 as Director General. His relationship with TWI, in the guise of the British Welding Research Association, dated back to 1950, a few months after completing his PhD in soil mechanics. As a Fellow of the Royal Society, Dean of Faculty for four of his years at Queen’s University in Belfast, a recipient of the Order of the British Empire, and one of the very few non-Americans to have contributed to the US Navy’s brittle-fracture work, he believed his career peaked relatively early. In 1954 he was able to accept an invitation to work with Dr George Irwin at the Naval Research Laboratory in Washington, DC. That placement was key in his role in developing the CTOD concept, but in addition Alan Wells also worked to establish the Wells wide-plate test. Wells later spent 13 years as the Head of Civil Engineering at Queen’s University Belfast before returning to TWI as Director General. Upon retirement he was involved in several major failure investigations including a fracture under hydrostatic test of one of the Sizewell power station boilers and the Westgate bridge collapse in Australia. Dr John Harrison joined TWI in 1962, when he worked on fatigue of welded structures, concentrating in particular on the significance of weld defects, the topic upon which he obtained his PhD from Cambridge University. He then became responsible for TWI’s fracture research, eventually being promoted Head of Engineering Research, covering design engineering, fatigue, fracture, and non-destructive testing. In 1985, the Engineering and Materials Group was formed, comprising the Design Engineering, Fatigue, Fracture, Non-Destructive Testing, and Materials Departments, and John was appointed Group Manager of a total of 90 staff. John Harrison has published over 50 papers on fatigue and fracture and has been extensively involved in national and international committees. In 1988 he resigned Chairmanship of Commission XIII of the International Institute of Welding which deals with the fatigue behaviour of welded structures and components, a position which he held for 15 years. He retired from TWI in 1998. 3rd Quarter, 2013 245 Use of curved-wide-plate (CWP) data for the prediction of girth-weld integrity io n by Dr Rudi M Denys*, Dr Stijn Hertelé, and Dr Antoon A Lefevre rib ut Laboratorium Soete, Universiteit Gent (UGent), Gent, Belgium A E ECA produces conservative estimates provided the input data and the variability of these data are known. The conservatism increases for defects in overmatched welds [21,24,34,36]. ot f VEN WHEN THE most stringent welding procedures are used, the occurrence of girth-weld defects is unavoidable during the construction of a pipeline. The acceptance or rejection of a detected defect can be assessed using workmanship or fitness-for-service based acceptance criteria [1-20]. However, the curved-wide-plate (CWP) test is an alternative tool to determine the effect of a defect on weld performance [21-36]. or d is t MULTITUDE OF interdependent material and geometric factors determine the response of a girth weld containing a defect under installation or service loads. A carefully designed curved-wide-plate (CWP) test enables a direct assessment of these factors. Consequently, the CWP test provides, as a predictor of the failure conditions, a tempting tool to assess girth-weld integrity, establish material requirements, and validate numerical models or fracture-mechanics’-based defect assessments. However, multi-disciplinary skills are required to explore this potential. This paper outlines the CWP testing requirements and the material data required to obtain representative information.The evolution, and the current and future roles, of CWP testing are also discussed. -n The CWP test, developed in 1979 by the Soete Laboratory at the UGent, Belgium, incorporates all factors affecting weld integrity and allows for the determination of the critical tensile load and strain capacity at fracture or at load instability. Accordingly, the CWP test provides a convenient tool to quantify the safety margin or the ratio between the predicted and the critical defect size involved in an ECA. pl e co py For conventional stress-based designs (axial remote strain < 0.5%)1, acceptance of a girth-weld defect is usually assessed against conservative weld-quality or workmanship (WMS) based go / no-go acceptance criteria. These criteria have no scientific basis and do not account for pipe grade, weld-strength mismatch, and wall-thickness effects. However, the application of these criteria prevents slipshod welding practices. Sa m If the defect is outside the WMS limits, contemporary welding standards allow that the rejected defect may be acceptable when assessed on the basis of a performancebased or fitness-for-service acceptance criterion. A fitnessfor-service assessment, or an engineering-critical assessment (ECA), uses fracture-mechanics’-based analytical equations and / or a plastic collapse criterion to determine the allowable dimensions of a planar defect. By design, an The 0.5% strain limit coincides with the strain at which the value of the specified minimum yield strength (SMYS) is defined. 1 *Author’s contact details: email: [email protected] After a brief discussion on the conservatism of ECA methods, the next sections are focused on the historical and future role of CWP testing in material selection and defect acceptance. Further, a brief review of the key factors to be considered in CWP testing is given. In this context, many of useful papers dealing with the details of the subject matter have been published. For space reasons, however, only a selected number of these papers is cited. Conservatism of ECA methods The quest for perfecting the accuracy of the predicted allowable defect limits has led to the development of a multiplicity of ECA methods. In particular, the pipeline industry can make use of both generic [9-12] and pipeline-specific ECA methods [13-20] which give different allowable defect sizes as the provisions, the assumptions, and the treatment of the input data differ 246 The Journal of Pipeline Engineering rib ut io n a. the predicted dimensions when the test is intended to determine the safety margin; b. the worst-case dimensions that could be encountered or escape detection; or c. for strain-based designs, the test performance criterion. is t For stress-based designs, it is recommended to select defect dimensions ensuring the onset of pipe yielding. Note also that the effects of weld reinforcement, weld-joint design (shape of the weld bevel), and weld misalignment on test performance are directly accounted for in the test. The test is conducted at the minimum design temperature until failure occurs in the notched section or by necking in the weakest pipe. When the toughness of the notch region shows transitional behaviour at the minimum design temperature, a lower test temperature is used to ensure lower-bound results. Note that specimen cooling needs due attention [48-49]. ot f If traditional ECA methodologies are used as a design tool, the user has to accept that the inherent conservatism of an ECA may produce over-conservative requirements which may lead to difficulties in weldingprocedure qualification testing. In addition, it may also be a challenge to obtain and ensure the required toughness for higher-grade pipe in production welds if the assessment is uniquely focused on the three-factor relationship between toughness, applied stress, and defect size. This problem can be solved if the many more interdependent factors affecting the complex relationship between crack-driving force and fracture resistance or fracture toughness are accounted for in the assessment. For example, the crack-driving force can be reduced by using overmatched weld metals [47-48]. In this context, it should also be noted that the latest ECA updates are mainly based on numerical simulations involving material models which do not necessarily represent the stress-strain characteristics of contemporary high-strength pipe materials. Therefore, the veracity of the assumptions for modern high-strength pipeline materials also requires more experimental validation. Depending on the properties of the weld region, the notch is placed at the weld-metal centreline or along the fusion line (HAZ) in either the weld root or cap. Multiple notched specimens can also be tested. The selection of the notch dimensions can be based on: or d [37-46]. Consequently, there exists no generally accepted ECA. This situation confuses potential users, and in particular the regulatory bodies. co CWP test py -n As no single ECA provides information on the safety margin, ECA defect assessments do not necessarily produce the most economic solution. To overcome this limitation, the CWP test, as a predictor of the failure conditions, is considered to be fit for setting allowable defects for a pre-determined level of safety. The CWP test can also be considered as a necessary step to rationalize existing ECA methods. Sa m pl e The CWP test involves an axially tensile-loaded curved (un-flattened) girth-welded pipe segment containing either a real defect or an artificial circumference surface-breaking crack-like notch1 at mid-length along the weld. As discussed in a later section, the test incorporates all influential factors intervening in the fracture/failure process. The overall dimensions of the test specimen are typically 1200 mm (axial length) by 420 mm (circumferential direction). Comparison of CWP and full-scale pipe bend-test data has demonstrated that testing of an axially tensile-loaded pipe segment with an arc length of 300 mm (or about 10% of the pipe circumference in the gauge section) is a conservative means to determine the failure characteristics of defective girth welds in a large-diameter ( > 36-in) pipeline [23,27]. In CWP testing, the term ‘notch’ represents a real defect, crack, flaw, or any other discontinuity. 2 To obtain structurally representative results, the CWP specimen must also be carefully instrumented. The analysis of the test results is a further challenge, since the effect of the multiple interdependent factors on the test results have to be accounted for [51-61]. Hence, users not well versed in the factors affecting girth-weld performance might unintentionally either produce structurally irrelevant CWP test data or not fully explore the potential of a CWP test. CWP testing Since its inception in 1979, UGent uses the CWP test to study girth-weld integrity for both stress-based and strain-based (axial remote strain > 0.5%) designs. At present, more than 1100 CWP tests on old-vintage low-strength and modern high-strength pipe grades up to API 5L X-120 have been conducted [32]. Historical developments in CWP testing The results of the early-1980s CWP tests demonstrated that: a. low fracture toughness (CTOD < 0.05 mm) does not automatically disqualify girth welds overmatched in strength; b. the first-generation, CTOD-design-curve based, ECA defect-acceptance methods over-estimate the effect of toughness on defect tolerance; and c. toughness, applied stress, and defect size are not the only factors determining weld integrity [2-7, 21,24]. 3rd Quarter, 2013 247 CWP tests on matched welds also illustrated that the highly constrained standard three-point bend CTOD toughness test underestimates the fracture resistance of tensile loaded girth welds [24]. In 1992, CWP tests on girth welds of moderate toughness confirmed that the workmanship limits for stress-based designs are very safe because CWP failure occurred after pipe yielding [22]. Since 2001, CWP instrumentation and material-testing requirements have gradually been fine-tuned to cover these issues [52-61]. Current use of CWP testing for stress-based designs Nowadays, the CWP test is applied to: a. determine the safety margin implied by the commonly used ECA defect assessments for stress-based designs; b. study factors affecting girth-weld integrity that cannot simply be modelled (such as tearing behaviour, weld-strength mismatch, and heterogeneity; c. determine the axial strain capacity of a girth weld with postulated defect dimensions [21-36]; and d. optimize material requirements for both stressand strain-based designs. io n In the mid 1990s, using the results of 186 CWP tests, UGent developed, on behalf of EPRG: co py is t -n ot f The analysis of 93 multiple-notched CWP specimens in 2005 led to the development of girth-weld-specific defect-interaction criteria which are now incorporated in the recently revised 2013 ERPG guidelines [29]. These guidelines also include the defect length-acceptance limits for 4- and 5-mm deep defects which can be applied for pipe grades up to X-80 [14]. It is of interest to note that the application of the EPRG defect-acceptance criteria is less complicated than the ECA methods using CTOD or J as fracture-resistance parameter [9-12, 15-20]. In 2011, 480 CWP test results were used to develop, through a lower-bound curve-fitting, the UGent defect-acceptance model for strain-based designs [62-63]. Furthermore, UGent also performs low-temperature (-50/-60°C) CWP tests to quantify the defect sizing capabilities of AUT inspection [28,31]. In contrast to the classical ‘salami’ technique, the low-temperature CWP test provides a direct visual access to the whole defect and prevents the actual defect size from being incorrectly estimated. In addition, the low-temperature CWP test allows the assessment of the severity of natural defects at lower-shelf toughness conditions [28,31,36]. or d CWP tests on 150-mm wide specimens extracted from thin-walled (5-mm) small-diameter pipe were used to adapt the ECA-based defect-acceptance procedures for small-diameter pipelines in the Australian pipeline code [26]. rib ut a. the world-wide accepted EPRG 30/40-J Charpy-V girth-weld toughness requirement where the axial (tensile and bending) strain is less than or equal to 0.5%; and b. the EPRG-Tier 2 length limit for 3-mm deep defects in pipe grades up to API 5L X-70 [13-14]. m pl e The results of the EPRG studies, the introduction of high-strength pipeline steels, and the increasing demand to study girth-weld integrity for strain-based designs, paved the way for the world-wide use of the CWP test [25, 30-36]. These research efforts disclosed that: Sa a. the level of weld-metal strength mismatch has a strong effect on the crack driving force; b. the pipes at either side of a girth weld have different tensile properties; c. strain-hardening capacity as measured by the Y/T ratio has a strong effect on ductile tearing behaviour and weld performance; and d. the correct interpretation of CWP test performance of specimens extracted from production welds or high-strength pipes requires significantly more pipe material and weld metal tensile testing than specified in existing material testing standards [49]. In contrast with an ECA assessment, wide-plate-test results allow for the determination of the critical conditions and the selection of the desired safety factor. Therefore, CWP testing is increasingly applied to formulate tailor-made weld-defect-acceptance criteria when production girth welds fail to meet the specified toughness or strength mismatch requirements, and an ECA, using the material properties measured during weld qualification or production testing, produces unduly restrictive defect limits. In addition, the decision for conducting CWP testing during construction is also driven by either the costs involved by production delays or the costs associated with the re-qualification of the materials until satisfactory results are obtained. CWP testing and strain-based design Numerical studies and pressurized (bi-axially loaded) full-scale tension (FST) tests have shown that internal pressure causes a higher axial stress for an equal axial strain [64-70]. This effect elevates the crack-driving force and reduces the strain capacity in the post-yield loading range. This means that the FST is the most suitable test to assess defect acceptance for strain-based designs. However, FST testing is expensive and, moreover, the material properties controlling the crack-driving force in the FST test have to be derived from a dummy weld. The Journal of Pipeline Engineering Fig.1.Typical sampling plan (not to scale). or d The UGent guidelines provide information on the specimen design, instrumentation, testing requirements, and the material properties that need to be measured for the interpretation of the CWP failure characteristics. The guidelines illustrate that the CWP test cannot be standardized in the same way as, for example, a CTOD test. Unlike the CTOD test, many more factors affect CWP test performance, among which are: ot f Even if the dummy weld is made with pipe pup pieces sampled from the same (parent) pipe and the same welding procedure, the inherent weld-to-weld variability makes it difficult to obtain the required information with sufficient precision. The problem is that the actual material properties determine test performance in the post-yield loading range [52,67,70]. is t rib ut io n 248 Sa m pl e co py -n Since, leaving aside the difference in crack-driving force, the CWP and FST test performance are controlled by the same factors, it is recommended to use the (uni-axially loaded) CWP test for strain-based designs. However, this option requires the internal pressure effect to be accounted for in the translation of the test results. For matching welds, the measured strains have to be reduced by a factor of two [62]; the correction decreases for overmatched welds [63]. Besides the practical and economic advantages, CWP testing allows the pipe material and weld-metal properties affecting CWP strain capacity to be obtained from test specimens taken out adjacent to the CWP specimen. Because of the better accessibility of the weld-root region, it is also easier to place the flaw/notch tip in the target weld metal (WM) and HAZ microstructure. Considering the variability of the material properties around the circumference, and since several CWP specimens can be removed from a single weld, it is also possible to conduct a sensitivity analysis. Guidance for CWP testing Using UGent’s vast experience in CWP testing for both stress-based and strain-based designs, the authors developed the UGent guidelines for CWP testing in 2009 [49]. Currently, DOT-PHMSA and PRCI are also preparing a CWP testing procedure, although to date there is no generally accepted standard for CWP testing. • notch location and dimensions (length and height); • ratio of notch depth to wall thickness; • geometric aspects such a pipe-wall thickness variations, high-low weld misalignment, weldbevel geometry, and weld reinforcement; • toughness properties of the notched region; • post-yield stress-strain responses (strain-hardening characteristics, Y/T ratio, etc.) of the materials in the notched region and the (remote) pipe sections; • uniform strain (uEL) capacity of the pipe metal; • ductile-tearing resistance of the material in the notched region; • level of yield, flow, and tensile strength mismatch. In the following, the UGent CWP testing guidelines and the recent modifications that require due considerations are briefly discussed. Material characterization and specimen sampling Literature on CWP testing reveals that different CWP test specimen dimensions are used [58], and the differences in instrumentation complicate the comparison of published CWP results. Added to this, the material properties controlling CWP performance are either 249 is t rib ut io n 3rd Quarter, 2013 Fig.2. CWP dimensions and instrumentation layout for deformation measurements. or d specimen. The minimum length of the prismatic test section is 3W, with the girth weld at mid-length. This length avoids the yielding bands emanating from the notch tips and the specimens ends affecting the strain measurements in the pipe body. ot f not reported or poorly documented. The issue is that the tensile and toughness properties vary in the circumferential direction [6, 10-11]. This explains, for example, why the interpretation of CWP results based on material properties inferred at a distance away from the CWP specimen may lead to misleading conclusions. Consequently, the pipe-material and weld-metal properties in the immediate vicinity of the CWP specimen need to be known in the evaluation of CWP test results. That is, material characterization in terms of tensile, hardness, and toughness cannot be derived from the testing of randomly sampled specimens. This problem can be simply solved when the small-scale material characterization testing samples are taken out adjacent to the CWP specimen (Fig.1). Medium CWP specimens py -n As discussed, W is normally equal to 300 mm, and this arc length ensures that the CWP test conservatively models the failure behaviour of girth welds in a largediameter pipe under bending loads. For some users, the load capacity of the test machine is a limiting factor for the testing of 300-mm wide specimens. When a medium, narrow, sub-sized, or ‘mini’ CWP specimen is tested, the test provides a distorted picture of the strain capacity and tearing behaviour occurring in wider specimens. Decreasing specimen width causes an increase of the net-to-gross section stress ratio with the result that the critical condition in a narrow specimen occurs at a lower level of applied load, applied strain, and CMOD, while the onset of ductile tearing occurs at a higher level of applied (gross) stress. However, medium CWP specimens provide structurally relevant information for small-diameter pipe when the arc length (width) of the prismatic section meets the 10%-circumference rule. m pl e co Pipes of different tensile properties are welded in a random order in the pipeline string, and so it is incorrect to assume in a CWP test assessment that the pipes on either side of the girth weld have the same or very comparable tensile properties. As shown in Fig.1, the pipe-material tensile properties at either side of the girth weld must also be determined. Sa CWP test specimen For large-diameter transmission pipelines, the following practical guidelines can be used to generate structurally representative CWP test results. Specimen dimensions The nominal dimensions of the ‘standard UGent’ dog-bone-shaped CWP specimen are shown in Fig.2. The minimum CWP dimensions are 1.4W (loading pull-tabs) by 4W (overall specimen length), where W is the width (arc length) of the prismatic part of the Notch location The notch is oriented parallel to the welding direction. In the region of the notch, the weld reinforcement is ground flush to facilitate the accurate placement of the notch. For HAZ testing, the notch can be placed in the HAZ of either the strongest or weakest pipe. As discussed in a later section, hardness testing can serve to assist notch-tip placement. Similarly, when significant wall-thickness variations exist, the HAZ of the thinnest The Journal of Pipeline Engineering tests showed that for notch-area ratios smaller than 7% failure occurs by plastic collapse. For such toughness levels, crack tip blunting causes a fatigue crack to act in a similar manner to a machined notch. Consequently, when the CVN toughness of the notch region meets this threshold, a machined notch is used, made with a thin (0.15/0.20-mm wide) jewellers’ cutting wheel or by electro-erosion. ot f or thickest pipe can be tested. When the HAZ of the strongest (and/or thickest) pipe is tested, the effect of the minimum level of weld-strength overmatch on strain capacity is evaluated. One can also test the HAZ/weldmetal interface of the weakest (thinnest) pipe, as the highest deformations will be concentrated in this area. Fig.3. Digital-image correlation picture of longitudinal strain distribution in a field weld. or d is t rib ut io n 250 -n Notch dimensions Sa m pl e co py The selection of the notch dimensions (height and length) are based on either the allowable defect size as predicted by a fracture-mechanics’-based assessment, the worst-case weld anomaly that might occur, or the maximum flaw dimensions that could escape detection by the inspection method used. When the latter two options are used, the non-destructive inspection tolerance is added to the selected notch dimensions. Further, it is of interest to mention that a remote strain of at least 0.5% will be obtained for notch area ratios smaller than 7% of the load-bearing cross-sectional area provided the threshold notch toughness of 30 J (min.) / 40 J (average) is met [13,51-52]. The 7% limit assumes matched or overmatched welds and applies for pipe metal Y/T ratios not exceeding 0.90. For Y/T ratios greater than 0.90, the notch-area ratio must be reduced, while for overmatched welds, larger notch-area ratios can be tested. Moderate to ample cap/root weld reinforcement in thin-wall pipe has a similar effect as weld-strength overmatch, and thus equally allows for a relaxation of the 7% limit. Notch-tip acuity The early CWP tests were carried out with a pre-fatigued sharpened notch. Provided that the Charpy toughness (CVN) of the notch area exceeds the 30/40-J threshold, Deformation measurements To capture the effects of the strength differences of the neighbouring pipes, and the possible wall-thickness differences (which can vary up to 1.5mm in the pipeline), the overall remote axial strain (gauge length: l0) as well as the strains of the pipe sections at either side of the girth weld (gauge lengths lA and lB, Fig.2) are measured. For room-temperature CWP tests, a more comprehensive picture of the complex strain distribution occurring in a CWP specimen / field weld can be obtained by means of digital-image correlation (DIC), Fig.3 [60], which illustrates that a relatively small difference in yield strength (22 MPa in the case shown) can cause very large differences in remote strains; the weaker pipe strains three times more than the stronger pipe. The crack-mouth-opening displacement (CMOD) is monitored by a clip-on measuring device which follows the relative displacement of two steel pins straddling the notch at mid-length. However, CMOD is not directly comparable with CTOD, as measured in a CTOD bend test. For an experienced eye, a CMOD-elongation/strain plot provides a simple means to explain the effects on strain capacity of toughness, weld-strength mismatch, 251 or d Curve A (failure mode: net section collapse) – undermatched girth weld is t Fig.4. Dependence of the CMOD on remote strain and weld-strength mismatch: rib ut io n 3rd Quarter, 2013 Curves B through F (failure after remote yielding) M = level yield of strength mismatch overmatched welds: MF > ME > MD > MC > MB ot f Curve F: CWP specimen failed by necking in the pipe. Note:The pipe material in girth weld E exhibited discontinuous yielding (Lüders plateau). Using established relationships between hardness and tensile strength, hardness-test results allow a quick screening to determine the possible strength differences between adjacent pipe metals, to identify the weakest pipe, and to obtain an estimate of the level of tensile strength mismatch, and thus to verify whether the girth weld is under- or overmatched. py -n and other factors. For example, Fig.4 shows the CMOD remote-strain responses of CWP specimens containing an identical notch where the girth welds were made with consumables of different strengths. These records provide experimental evidence that: Sa m pl e co a. the relationship between toughness in terms of CMOD and CWP strain capacity strongly depends on the level of strength mismatch; b. strength mismatch has a clear effect on strain capacity; c. reliance on toughness alone as the quantifying parameter in a defect assessment can disqualify acceptable welds; and d. toughness requirements can be related to weldstrength mismatch (overmatched welds require less toughness than matched welds). In this respect, note that the beneficial strength mismatch effects on strain capacity gradually break down with increasing defect size. Macrosectioning and hardness testing Transverse-weld cross section(s) are extracted to illustrate the configuration and sequence of the weld runs within the girth weld, to facilitate notch-tip placement in the CWP specimen and to perform hardness tests. Hardness testing is performed according to current welding-qualification standards. Using a 5-kg pyramiddiamond indenter (HV5), the indentations are made along traverses both at the cap and root side, 2 mm below the pipe surface in the unaffected pipe material, the weld metal, and the HAZ at either side of the girth weld. Alternatively, a hardness map provides a far better picture of the hardness variation across the girth weld, Fig.5: a typical hardness map involves between 700 and 1200 HV5 measurements. The girth-weld hardness maps shown in Fig.5 provide information with respect to the strength variations in the pipe materials, the heat-affected zones, and the weld metal. For example, both parts of the figure illustrate that the pipes have different tensile properties, and that the weld metal is matched (left) and undermatched (right) in tensile strength. Also, Fig.5b shows that, because of the large hardness differences in the through-wall direction, the root area in the SMAW weld is either The Journal of Pipeline Engineering ot f or d is t rib ut io n 252 -n Fig.5. Macrosections and corresponding hardness maps of a GMAW and SMAW girth weld. py matched (right) or undermatched in tensile strength (left), whereas the weld cap is overmatched. Furthermore, for the GMAW weld, the HAZ regions are overmatched to both the pipe materials and the girth weld (Fig.5a). co Material tensile properties Sa m pl e Unless the CVN toughness of the weld region is lower than the 30/40-J threshold, the CWP specimen usually fails by plastic collapse in either the notch section or by necking in the soft pipe. In this case, test performance is controlled by the axial tensile properties of the girth weld and the adjacent pipe pup pieces. The quantification of these failure modes requires the stress-strain response and the strain-hardening rate or Y/T ratio of the materials. The information is also essential to facilitate the quantification of the level of yield, flow, and tensile-strength mismatch [47-48]. The percentage elongation at the uniform strain of the pipe materials, uEL, is another essential factor because it relates to the achievable strain capacity. Pipe-metal tensile tests Full-thickness longitudinal specimens, extracted from both pup pieces in the axial direction, are tested. Round-bar (RB) specimens may not be used because the material is taken from the mid-wall thickness location so that the ‘stronger’ inner and outer pipe surfaces are not tested. The use of full-thickness specimens can avoid an overestimated level of strength mismatch. In addition, since pipes in the coating-aged condition can exhibit higher yield strength and Y/T ratio values and lower uniform elongations than bare – uncoated – pipes, the test coupons should have undergone a thermal cycle as used in the plant and/or field coating [69]. All-weld-metal tensile tests The tensile properties of the girth weld are determined by either all-weld-metal RB or rectangular specimens. It must also be ensured that the parallel length of the specimen consists entirely of weld metal corresponding to a particular extraction position. Hardness measurements can be used to assist the placement of RB specimens in the through-thickness direction. Practical considerations The variation of the tensile properties of the pipe in the axial and circumferential directions is often neglected in the assessment of CWP test results. The associated difference in post-yield stress-strain response is another 253 rib ut io n 3rd Quarter, 2013 Fig.6.The effect of sampling position on weld yield strength and Y/T ratio of an undermatched weld (specimens were taken from the 9 and 10 o’clock positions). or d is t threshold by a factor of two, it is considered that the effect of sampling location on toughness does not need further consideration. When the CWP test is aimed at the validation of numerical or analytical defect assessments, complementary fracture-mechanics’ tests are performed to determine the fracture toughness in terms of K, J, and/or CTOD. When ductile tearing intervenes in the failure process, other fracture-mechanics’ parameters such as the R-curve can be used to characterize material toughness. For girthweld defects, the SENT test is a better tool than the SENB test to estimate the onset of stable tearing and tearing resistance under tensile load. The single-specimen R-curve testing requires measurements of crack growth by means of techniques such as unloading compliance or (direct-current) potential drop. Similar crack-growth measurements can be performed on CWP tests [71-72]. co Toughness testing py -n ot f significant factor requiring consideration [48,55,57]. The scatter of the all-weld (AW) metal tensile properties is an even more complicated issue. Aside from the variations around the circumference, the microstructural variations within the weld deposit or through-thickness direction make it more difficult to determine representative AW tensile properties [48]. The nature of this variation also depends on the welding process. For SMAW welds, the tensile properties of the weld root are lower than those of the fill passes. The weld passes in the root region of narrow-gap GMAW welds are stronger than the fill and cap passes. Consequently, specimen geometry, sampling location, and specimen dimensions affect the measured AW tensile properties [36,48]. Therefore, a post-test verification of the microstructure(s) sampled by AW specimen is recommended when significant strength variations are measured, Fig.6. Sa m pl e Toughness testing can be limited to Charpy V-notch (CVN) toughness testing because correlations between CWP performance and CVN impact energy have shown that the CVN test is suitable for predicting the expected CWP failure behaviour [51-52]. As discussed, failure by plastic is ensured if the toughness of the weld region exceeds the 30/40-J threshold. However, SENT tests are recommended to assess the tearing resistance of high Y/T ratio welds. In the material-screening phase, CVN toughness testing is concentrated on the girth-weld centreline and the fusion line of the low-strength pipe, as the highest plastic strains in CWP testing occur in this area. Unless the weld cap requires a specific assessment, it is a standard practice to focus testing on the weld root. The CVN impact test is conducted at the minimum design temperature using full-size standard (10 x 10 mm2) test pieces. If the CVN properties exceed the 30/40-J CWP test performance criteria Upon completion of the test, the overall and remote strains at specimen failure are either compared to a predetermined performance criterion or used as input in a fracture-mechanics’-based assessment (ECA validation). After the test, the notched region and the fracture surfaces of the broken specimens are subjected to a fractographic examination. The broken as well as unbroken HAZ specimens are then sectioned for post-test metallography to identify the microstructure sampled by the notch tip. Strain criterion Unless the CWP data are used to validate fracturemechanics’ methods, it is safe for stress-based designs to require that the remote longitudinal strain meets or exceeds the 0.5% level [13,14]. In this context, the 0.5% criterion has been applied to establish the Charpy requirements of pressure-vessel steels [73]. The rationale for adopting the 0.5% performance criterion The Journal of Pipeline Engineering rib ut io n 254 -n Post-test fractography and metallography or d circumferential surface-breaking defect(s) for both stressbased and strain-based designs. Since the CWP test captures the influential factors affecting weld performance, the test data play a useful role in identifying the limitations inherent in the commonly used ECAbased defect assessments and associated material-test requirements. At this time, CWP testing of a carefully designed and instrumented specimen is often used as a reliable tool for developing tailor-made weld-defectacceptance criteria, which account for the specified girth-weld performance requirements and the available material properties. However, for strain-based designs, the effect of internal pressure on the crack-driving force must be accounted for by applying a correction factor. ot f is both to provide an adequate margin of safety for conventional stress-based pipeline designs, and also to exclude toughness-dependent fracture. As discussed, for strain-based designs, the CWP test does not capture the internal pressure effects on the crack-driving force. For matched welds, this deficiency can be accounted for by reducing the measured strains by a factor of two; this factor can be reduced for overmatched welds [49-50]. is t Fig.7. Cross sections of valid and invalid notch locations in CWP specimens which failed by pipe necking. Note that significant blunting occurred prior to failure. pl e co py For broken specimens, the initial notch geometry, the extent of blunting and ductile tearing, the crack dimensions at failure, and the fracture faces are evaluated under a stereomicroscope. In addition, macrographic and micrographic examinations of HAZ-notched specimens are performed to verify whether the notch tip effectively intercepted the target microstructure (validity check), and to evaluate the fracture path in the through-thickness direction relative to the fusion boundary. The photographs in Fig.7 illustrate the locations of the notch tips in two unbroken CWP specimens. Sa m UGent’s experience has shown that the CWP test results of CGHAZ / fusion-line notched specimens are ‘metallurgically valid’ when the distance between the target microstructure (the fusion line) and the original (machined or pre-fatigued) notch tip is smaller than 0.50mm (Fig.7a). When this requirement is not achieved (Fig.7b), engineering judgment is needed to quantify the practical significance of the CWP test result. Concluding remarks This paper gives a brief overview of the developments and potential of CWP testing to assess the integrity of girth welds containing either a single or a multiple The correct interpretation of CWP test results requires that the material properties controlling test performance are known. In particular, the material characterization in the vicinity of the CWP specimen is critical to the assessment of CWP test results. The issue is that the quantification of the pipe material and weld-metal properties requires significantly more testing than specified in existing material-testing standards. Concluding matters, the CWP test is not a routine test. Its execution is complex and the analysis and understanding of the test results requires a multi-disciplinary expertise and a wide range of skills because multiple interdependent factors influence the CWP test performance. References 1. R.Denys, 1999. ECA-based girth weld defect acceptance levels. Pipeline & Gas Journal, Dec., 226, 9, p 60, Part I. 2. E.L.von Rosenberg, 1980. Alternative girth weld defect assessment criteria for pipelines. Proc. Conf. on Pipeline and Energy Plant and Piping, pp 319328. Permagon Press. 3rd Quarter, 2013 255 or d is t rib ut io n 23.P.Hopkins, G.Demofonti, G.Knauf, and R.Denys, 1991. An experimental appraisal of the significance of defects in pipeline girth welds. Proc. EPRG/ NG18 8th Biennal JTM on Line Pipe Research, Paris, May 14-17, paper 20. 24.R.Denys, 1993. Fracture behavior of large diameter pipeline girth welds: effect of weld metal yield strength – Volume II. AGA Catalog No. 202-922. F.Minami, M.Ohata, M.Toyoda, T.Tanaka, 25. K.Arimochi, A.G.Glover, and T.H.North, 1995. The effect of weld metal yield strength on the fracture behaviour of girth welds in grade 550 pipe. Proc. Pipeline Technology Conf., pp441-461, Ostend, Belgium. 26.G.F.Bowie, F.J.Barbaro, C.J.Moss, L.Fletcher, and R.Denys, 1997. Defect acceptance criteria in thin walled pipes girth welds. Proc. 2nd Symposium on Mis-Matching of Interfaces and Welds, pp573586, GKSS Geesthacht. 27. W.Schipaanboord, R.Denys, A.Lefevre, and P.Roovers, 1999. Validation tests to sSupport the EPRG – Tier 2 guidelines. Proc. 12th Biannal JTM on Pipe Line Research, Groningen, 17-21 May. 28.R.Denys, A.Lefevre, C.De Jaeger, and S.Claessens, 2000. Failure characteristics and defect tolerance levels of girth welds in large diameter X65/ X70 steel pipelines: experimental verification through wide plate testing and comparison with ECA prediction models. Proc. 3rd International Pipeline Technology Conference, Elsevier Science, Amsterdam, I, pp151-167, Brugge, Belgium. 29.R.Denys, W.De Waele, and A.Lefevre, 2005. Input to code revision for interaction of girth weld defects under plastic collapse conditions. Contract GRI 8766, January. 30.H.G.Pisarski, A.Muhammed, and A. Staccy, 2000. Using wide plate test results to improve predictions from probabilistic fracture mechanics. ECF 13, San Sebastian, Spain. 31.R.M.Andrews and L.L.Morgan, 2003. Integration of automated ultrasonic testing and engineering critical assessment for pipeline girth weld defect acceptance. 14th JTM on Line Pipe Research, Berlin, paper 36. 32. M.L.Macia, S.D.Papka, D.P.Fairchild, R.Denys, and A.Lefevre, 2004. Verification of fracture toughness and defect tolerance of X120 girth welds. Proc. 4th International Pipeline Technology Conference, I, Ostend, Belgium, 9-13 May. 33.M.V.Hukle, A.M.Horn, D.S.Hoyt, and J.B.LeBleu, Jr., 2005. Girth weld qualification for high strain pipeline applications. ASME, Int. Conf. on OMAE, Paper No 2005-67573, pp1-6, Halkidiki, Greece. R.Denys, S.Hertelé, and M.Verstraete, 2010. 34. Longitudinal strain capacity of GMAW welded high niobium (HTP) grade X80 steel pipes. Proc. of HSLP, Xi’an, China. Sa m pl e co py -n ot f 3. API, 1983. API-1104 Standard for field welding pipe lines and related facilities – Appendix A. 4. BSI, 1984. BS-4515 Specifications for field welding of carbon steel pipelines – Appendix H. British Standards Institution, London. 5. CSA, 1986. CSA-Z184 Appendix K, 1986, Standard for welding pipelines and related facilities – Appendix K. Canadian Standards Institution, Mississauga. 6. P.D.Hilton and R.A.Mayville, 1985. An evaluation of girth weld defect acceptance criteria. Arthur Little Inc., Cambridge, MA. 7. BSI, 1980 and 1991. BSI-PD 6493, 1980 and 1991, Guidance on some methods for the derivation of acceptance levels for defects in fusion welded joints. British Standards Institution, London. 8. British Energy, 2005. R6-Rev 4 Assessment of the integrity of structures containing defects. 9. BSI, 2005. BS 7910:2005 Guide on methods for assessing the susceptibility of flaws in structures, incorporating amendment No.1. British Standards Institution, London. 10. API, 2000. API-579 Fitness-for-service: API recommended practice. 11.M.Kodak, S.Webster , J.J.Janosch, R.A.Ainsworth, and R.Koers, 2008. FITNET fitness-for-service (FFS) procedure (Vol.1). GKSS Research Centre, Geesthacht. 12.M.Kodak, L.Hadley, S.Szavai, Y.Tkach, and N.Taylor, 2008. Ibidem, Annex (Vol.2). 13.G.Knauf and P.Hopkins, 1996. EPRG guidelines on the assessment of defects in transmission pipeline girth welds. 3R International, 35, 10-11, pp 620-624. 14.R.M.Denys, R.M.Andrews, M.Zarea, and G.Knauf, 201. EPRG Tier 2 guidelines for the assessment of defects in transmission pipeline girth welds. Proc. IPC, Calgary, Canada. 15. Standards Australia, 2002. AS 2885.2-2002 Pipelines – gas and fluid petroleum, Part 2: welding. Homebush, New South Wales. 16.EN 12732, 2002. Gas supply systems – welding steel pipework – functional requirements. London. 17.BSI, 2004. BS 4515 Specification for welding of steel pipelines on land and offshore. Carbon and carbon manganese steel pipelines. British Standards Institution, London. 18. API, 2005. API-1104 Welding of pipelines and related facilities. 20th Edn, Washington, USA, November. 19.Det Norske Veritas, 2007. DNV-OS-F101 Submarine pipeline systems, offshore. 20.CSA, 2007. Z662 Oil and gas pipeline systems. Mississauga, Ontario. 21.B.Lian, R.Denys, and L.Van de Walle, 1986. An experimental assessment on the effect of weld metal yield strength overmatching in pipeline girth welds. Proc. 3rd Conf. on Welding and Performance of Pipelines, TWI, London, November. 22.R.Denys, 1992. Fracture behaviour of girth welds containing natural defects and comparison with existing workmanship standards. AGA No. 202-009, pp102, 9 Feb. 256 The Journal of Pipeline Engineering or d is t rib ut io n 49.R.Denys and A.Lefevre, 2009. Gent guidelines for curved wide plate testing. Proc. 5th International Pipeline Technology Conference, Ostend, Belgium, 12-14 October. 50.Y.-Y.Wang, M.Liu, Y.Chen, and D.Horsley, 2006. Effects of geometry, temperature, and test procedure on reported failure strains from simulated wide plate tests. Proc. IPC, paper IPC2006-10497, Calgary, Canada. 51.R.Denys, 1995. Toughness requirements for pipeline integrity. Proc. 13th OMAE Conference, III, Copenhagen. R.Denys and A.Lefevre, 2006. Material test 52. requirements for strain-based pipeline design. Proc. Int. Symposium on Microalloyed Steels for the Oil and Gas Industry, Araxa, MG, Brazil, TMS (The Minerals, Metals & Materials Society) Publication, Warrendale, USA, ISBN 978-0-87339-656-1, pp513-533. 53. D.P.Fairchild, W.Cheng, S.J.Ford, K.Minnaar, N.E.Biery, A.Kumar, and N.E.Nissley, 2008. Recent advances in curved wide plate testing and implications for strain-based design. Int. J. of Offshore and Polar Engineering, 18, 3, pp161-170, September. 54.S.Hertelé, W.De Waele, R.Denys, and M.Verstraete, 2010. Investigation of pipe strain measurements in a curved wide plate specimen. Proc. IPC, paper IPC2010-31292, Calgary, Canada. 55.R.Denys, S.Hertelé, and M.Verstraete, 2010. Strain capacity of weak and strong girth welds in axially loaded pipelines. Proc. IPTC 2010, International Pipeline Conference, Beijng, China. T.Weeks, D.McColskey, M.Richards, Y.-Y.Wang, 56. H.Zhou, W.Tyson., and M.Quitana, 2011. Weld design, testing, and assessment procedures for highstrength pipelines curved wide plate tests. US DOT Contract No. DTPH56-07-T-000005, final report to US DOT and PRCI, September. 57.S.Hertelé, W.De Waele, R.Denys, and M.Verstraete, 2011. Sensitivity of plastic response of defective pipeline girth welds to the stress-strain behaviour of base and weld metal. Proc. 30th OMAE Conference, paper 2011-49239, Rotterdam, June. 58. S.Hertelé, M.Verstraete,K.Van Minnebruggen, R.Denys, and W.De Waele, 2012. Curved wide plate testing with advanced instrumentation and interpretation. Proc. IPC, paper IPC2012-90591, Calgary, Canada. 59. S.Hertelé, W.De Waele, R.Denysm and M. Verstraete, 2012. Full-range stress-strain behaviour of contemporary pipeline steels: Part I: model description. Int. J. Press Vessels and. Pipe, 92, 34-40. 60.S.Hertelé, W.De Waele, R.Denys, and M.Verstraete, 2012. Investigation of strain measurements in a (curved) wide plate specimen using digital image correlation and finite element analysis. J. Strain Anal. Eng. Des., 47, 5, 276-88. 61. S.Hertelé, N.P.O’Dowd, R.Denys, K.Van Minnebruggen, and W.De Waele, 2013. Effects of pipe steel heterogeneity on the tensile strain capacity of a flawed pipeline girth weld. Submitted to Eng. Fract. Mech. Sa m pl e co py -n ot f 35. Y.-Y.Wang, M.Liu, X.Long, M.Stephens, R.Petersen, and R.Gordon, 2011. Validation and documentation of tensile strain limit design models for pipelines. US DOT Contract No. DTPH56-06-T000014, 2 August. 36.T.Swankie, 2012. Validation of assessment methods for production scale girth welding of high strength steel pipelines with multiple pipe sources. US DoT, DTPH56-07-T-000006 WP #275, March. 37.G.M.Wilkowski and R.J.Eiber, 1978. Review of fracture mechanics approaches to defining critical defect size girth weld discontinuities. Weld. Res. Bull., 329, July. 38.A.H.Roodbergen and R.Denys, 1987. Limitations of fitness-for-purpose assessment of pipeline girth welds. Int. Conf. on Pipeline Technology, AIM, Rome. 39.J.R.Gordon, 1993. A review of fracture assessment procedures and their applicability to welded structures: fracture control in welded structures: codes and standards. Special Issue: Canadian Metallurgical Quarterly, 32, 3, July-September, pp195-204. 40.P.M.Scott, 1993. Interpretative study of published and recent research on the applicability and limitations of current fracture prediction models for girth welds. Ibidem, pp223-238. 41.R.Denys, A.Lefevre, J.T.Martin, and A.G.Glover, 1995. Yield strength mis-match effects on fitness for purpose assessments of girth welds. PRC/EPRG 10th Biennial Joint Technical Meeting on Line Pipe Research, AGA, Paper 3, pp1-15, Cambridge, UK. P.M.Scott, T.L.Anderson, D.A.Osage, and 42. G.M.Wilkowski, 1998. Review of existing fitnessfor-service criteria for crack-like flaws. WRC Bulletin, 430, 155p April. Y.-Y.Wang, M.Liu, and D.Rudland, 2007. A 43. comprehensive update in the evaluation of pipeline weld defects. US DOT Agreement # DTRS5603-T-0008, PRCI Contact # PR-276-04503. 44.A.Cosham, 2008. ECAs: are they fit-for-purpose? OPT 2008, Amsterdam, Netherlands, February. 45.G.Wilkowsky, D.Rudland, D.Shim, and F.W.Burst, 2008. An advanced integration of multi-scale mechanics and welding process simulation in weld integrity assessment. Final Technical Report DOE DE-FC36-04Go14040, June. 46.R.Denys and W.De Waele, 2009. Comparison of API 1104 - Appendix A Option 1, and EPRG-Tier 2 defect acceptance limits. Proc. Pipeline Technology Conference, Ostend, Belgium. 47.R.Denys, W.De Waele, and A.Lefevre, 2004. Effect of pipe and weld metal post-yield characteristics on plastic straining capacity of axially loaded pipelines. IPC, paper IPC04-0768, Calgary, Canada. 48.R.Denys, 2008. Weld metal strength mismatch: past, present and future. Proc. Int. Symp. to Celebrate Prof. Masao Toyoda’s Retirement from Osaka University, Osaka, Japan pp115-148. 3rd Quarter, 2013 257 or d is t rib ut io n 68.X.Li, L.Ji, W.Zhao, and H.L.Li , 2007. Key issues should be considered for application of strain-based designed pipeline in China. Proc. ISOPE2007, Lisbon, Portugal. 69. L.Fletcher, 2003. Strain limits for hydrostatic testing: Revision to AS 2885.1. Issue Paper 1.12, Revision: B, 20 June. 70.Y.-Y.Wang, M.Liu, J.Gianetto, and W.Tyson, 2010. Considerations of linepipe and girth weld tensile properties for strain-based design of pipelines. Proc. IPC, Calgary, Canada. 71.W.Cheng, P.C.Gioielli, H.Tang, K.Minnaar, and M.L.Macia, 2009. Test methods for characterization of strain capacity – comparison of R curves from SENT/CWP/FS tests. Proc. Pipeline Technology Conference, Ostend, Belgium. 72. H.Tang, K.Minnaar, S.Kibey, M.L.Macia, P.C.Gioielli, and D.P.Fairchild, 2010. Development of SENT test procedure for strain-based design of welded pipelines. Proc.IPC, Calgary, Canada. 73. R.Denys, 1990. Wide plate testing of weldments, Parts I, II, and III. Fatigue and Fracture Testing of Weldments, ASTM STP 1058, pp157-228, ASTM, Philadelpia. Sa m pl e co py -n ot f 62.R.Denys, S.Hertelé, M.Verstraete, and W.De Waele, 2011. Strain capacity prediction for strain-based pipeline designs. Proc. Workshop Welding of High Strength Pipeline Steels, Araxa, Brazil. 63.M.Verstraete, W.De Waele, R.Denys, and S.Hertelé, 2012. Pressure correction factor for strain capacity predictions based on curved wide plate testing. Proc. IPC, paper IPC2012-90592, Calgary, Canada. 64.J.R.Gordon, N.Zettlemoyer, and W.C.Mohr, 2007. Crack driving force in pipelines subjected to large strain and biaxial stress conditions. Proc. IPC, 4, pp3129-3140, Lisbon, Portugal. 65.K.Minnaar, P.C.Gioielli, M.L.Macia, F.Bardi, and W.C.Kan, 2007. Predictive FEA modelling of pressurized full-scale tests. ISOPE, Lisbon, Portugal. 66.S.Kibey, S.P.Lele, H.Tang, M.L.Macia, D.P.Fairchild, W.Cheng, W.C.Kan, M.F.Cook, R.Noecker, and B.Newbury, 2011. Full scale test protocol for measurement of tensile strain capacity of welded pipelines. Proc. ISOPE, Maui, Hawaii, USA. 67.D.B.Lillig, D.S.Hoyt, M.W.Hukle, J.Dwyer, and A.M.Horn, 2006. Materials and welding engineering for ExxonMobil high strain pipelines. Proc. 16th Int’l Offshore and Polar Eng. Conf., pp8-17, San Francisco, USA. rib ut io n DON’T MISS AN ISSUE! Sa m pl e co py -n ot f or d is t SUBSCRIBE TODAY Purchase your Pipelines International subscription online at pipelinesinternational.com/subscribe 3rd Quarter, 2013 259 is t 1 Edison Welding Institute, Columbus, OH, USA 2 B N Leis Consultant, Inc, Worthington, OH, USA rib ut by Dr Xian-Kui Zhu*1 and Dr Brian N Leis2 io n Ductile-fracture arrest methods for gas-transmission pipelines using Charpy impact energy or DWTT energy T or d HE STANDARDIZED CHARPYV-notched (CVN) impact energy has been used by the pipeline industry since the 1960s to characterize fracture toughness of pipeline steels, and is central to the fracturecontrol technology developed for gas transmission pipelines. The drop-weight tear test (DWTT) has been standardized to assess fracture mode in such applications, with DWTT energy suggested as a means to quantify toughness, although not standardized for such a use. py -n ot f Battelle developed its two-curve model (BTCM) in the early 1970s to determine the required toughness to arrest ductile fracture in gas transmission pipelines in terms of CVN impact energy.The BTCM has been found viable for pipeline grades up to X-65, but issues have emerged in applications to higher grades. Thus different correction methods were proposed over the years to improve the BTCM predictions.This paper reviews the use of CVN and DWTT energy in conjunction with the BTCM to predict arrest toughness to control running fractures in gas transmission pipelines, and evaluates correction methods adopted to extend its use to X-80 and above. The correction methods include the Leis correction factor, the CSM factor, the Wilkowski DWTT method, and others. These methods are evaluated through analysis and comparison of predictions with full-scale experimental data. Suggestions to further improve the BTCM also are discussed. D Sa m pl e co UCTILE FRACTURE PROPAGATION control is a major concern for the safe design and operation of modern gas transmission pipelines at high internal pressure. Technology to ensure such control is critical for the structural integrity and safety of the gas pipeline because the possibility of a running fracture opens-up the catastrophic long-running failure of a gas pipeline which poses a societal threat involving the public and property, as well as to the environment local to the pipeline. Fracture resistance to a running fracture is an important material property, with the minimum resistance to affect arrest defining the arrest toughness. The absorbed energy obtained in a Charpy V-notched (CVN) impact test has been used to characterize the ductile fracture resistance since the 1960s in the pipeline industry, *Corresponding author’s contact details: tel: +1 614 688 5135 email: [email protected] while a drop-weight tear test (DWTT) has been used to assess the fracture mode, with DWTT energy also suggested to quantify toughness. Both CVN and DWTT have played a key role in that context as part of the technology developed for fracture-propagation control (FPC) for gas transmission pipelines [1]. Based on the fracture resistance quantified by CVN energy, Battelle developed its two-curve model (denoted here as BTCM) in the mid-1970s to predict arrest toughness for ductile pipeline steels. Maxey [2] detailed the technology that underlies the BTCM and its use for dynamic fracture control. The fracture and the gas-decompression behaviours in a gas pipelines were uncoupled, being described by two independent curves. The gas-decompression curve was quantified by the code GASDECOM [3] developed at Battelle, which was valid for a wide range of gas compositions; the fracture-resistance curve was described by a semiempirical fracture model in terms of CVN energy. It has been found that the CVN-energy-based BTCM is The Journal of Pipeline Engineering rib ut io n 260 co py -n ot f Much has been done since the 1980s to improve the predictability of BTCM in applications at the highertoughness levels commonly available for modern highstrength steels. Such work includes corrections and correlations: viz. the Leis correction [4], CSM factor [5], Wilkowski pressed-notch DWTT method [6], and Japanese pre-cracked DWTT method [7]. In a soon-tobe-released update of the PRCI report on Fracture control technology for transmission pipelines, Leis and Eiber [8] outline the history, technology, methods, and progress for ductile FPC for gas-transmission pipelines. Further details of these topics can be found in that report. design against this threat. In the 1960s, as the need for such technology became clear, the potential of fracturemechanics’ technology was evident, but the capabilities were rudimentary as compared to the complexity posed by ductile running fractures. More critically, the CVN test had been adopted by the pipeline industry to quantify fracture resistance: accordingly, the technology developed to quantify arrest toughness was formulated in reference to CVN impact energy, as follows. or d accurate only for pipeline grades up to X-65 with low to intermediate fracture toughness, with issues developing for higher toughness levels, which commonly develop for higher-strength pipeline grades. is t Fig.1. Fracture and gas-decompression velocity curves. Battelle two-curve model Battelle’s two-curve model (BTCM) was presented by Maxey [2] in 1974. This model considers the gas decompression (fracture-driving curve) and the dynamic crack-propagation resistance (fracture-resistance curve) as uncoupled processes, with both related to decompressed pressure local to the crack tip. The gas-decompression curve is determined by the code GASDECOM developed at Battelle that is viable for a wide range of gas compositions [3]. The fracture-propagation curve (or fracture-resistance curve) is determined by the following semi-empirical equation: Sa m pl e This paper reviews the use of CVN and DWTT energy in conjunction with the BTCM to predict arrest toughness to control a running fracture in gastransmission pipelines, and evaluates correction methods adopted to extend its use to X-80 and above. The 1/ 6 objective is to assess the utility of the methods available σf P = V C − 1 (1) f to assess arrest toughness for modern pipelines made R Pa of high-toughness and high-strength steels. Corrections and correlations are evaluated through analysis and comparison of predictions with full-scale experimental where: data. Suggestions to further improve the BTCM also are discussed. Vf is the fracture propagation velocity in m/s (or ft/s); CVN impact energy and ductile C is a backfill parameter with a constant fracture arrest methods value of 2.75 (or 0.648) and 2.34 (or 0.47), respectively for no backfill in air conditions Because high internal pressure can cause rapid axial and soil backfill; ductile fracture propagation along a gas-transmission σf = σy + 69 MPa (or 10 ksi) is a flow stress pipeline, technology has evolved as noted above to in MPa (ksi); 3rd Quarter, 2013 261 et al. [10] curve-fit a simple formula to the results of parametric BTCM analyses permitting direct calculation of the arrest toughness for such applications. Based on the results developed via the BTCM for a wide range of mechanical properties and CVN values, the arrest fracture toughness quantified relative to a 2/3-size CVN specimen was obtained as a function of the initial hoop stress and pipe geometry as: io n R = CV/Ac is fracture resistance with CV the full-size Charpy (CVN) impact energy at the upper shelf in Joule (or ft-lb) and Ac the ligament cross-section area of Charpy specimen with a value of 80 mm2 (or 0.124 in2); P is the instant decompressed pressure near the crack tip in MPa (or psi); and Pa=2tσa/D is the arrest pressure at the crack tip in MPa (or psi). The arrest hoop stress is determined by: CV ( 2 /= 7.2 × 10−3 × σ h2 ( Rt ) 3) 2σ f π RE σa = ](2) arccos exp[− 2 24σ f Dt / 2 3.33π where CVN energy is in ft-lb, the hoop stress σh is in ksi, and the pipe radius R and wall thickness t are in inches. Because of its simplicity, this simplified form of the BTCM has been adopted in a variety of codes and standards for gas-transmission pipeline design [for example, see Ref. 8]. Concurrently many simplified equations (SEs) have emerged in a form similar to Equn 3 through work done in the North America, Europe, and Japan. While their predictive quality varies, these SEs provide a simple basis for estimating arrest toughness. Nine such equations can be found in Ref. 8. (US units) (3) is t or d Since the mid-1970s, pipeline-steelmaking technology has been improved remarkably and thus fracture toughness of pipeline steels has been significantly increased. It was found that the BTCM and all simplified models predicted non-conservative arrest toughness in comparison to measured Charpy energy for high-strength pipeline steels with CVN energies larger than 70 ft-lb (or 95 J) [4]. Therefore, a variety of corrections, correlations, and modified methods have been proposed to improve the BTCM predictions, as reviewed next. ot f where E is the elastic modulus in GPa (or ksi), D is the pipe diameter in mm (or inches), and t is the pipe wall thickness in mm (or inches). Equn 2 was originally developed in Ref. 9 for fracture-initiation control for pipelines by use of the rudimentary fracture-mechanics’ method. In this equation, a linear relationship between the CVN energy and the fracture toughness GC was assumed. The linear relationship was obtained by calibration of full-scale hydrostatic burst data for pipeline steels dominated by grades ≤ X65. rib ut 1/ 3 Sa m pl e co py -n When the fracture curve determined by Equn 2 and the gas-decompression curve determined by GASDECOM are tangent, as shown in Fig.1, the minimum fracture-arrest toughness in terms of CVN energy is determined for the pipeline under consideration. If the CVN toughness increases, the fracture curve moves up to be above the gas curve, and a running fracture will arrest because the fracture velocity is slower than the gas-decompression velocity at all pressure levels. On the other hand, if the CVN toughness decreases, the fracture curve moves down, and the running fracture will continue to propagate because the fracture velocity is faster than the gasdecompression velocity. Note that the gas-decompression code GASDECOM is quite general in its applicability because it embeds the BWRS equations of state and accounts for both single-phase and two-phase decompression behaviours of gases. In contrast, Equns 1 and 2 used for determining the fracture-resistance curve were calibrated with experimental data available in the late 1960s to the early 1970s, such that lower-toughness linepipe steels up to grade X-65 were involved. Therefore, the BTCM may be inaccurate when used beyond its calibration database. Simplified equation to represent the BTCM Use of the BTCM to predict the arrest toughness depends on its software. While DYNFRAC developed at Battelle [3] is available today, its initial use required iterative computer-based analysis. To facilitate such analysis when dealing with single-phase decompression, Maxey Leis correction It has been found the proportion of propagation to initiation energy dissipated in the CVN specimen varies greatly with increasing toughness for ductile pipeline steels [for example, Ref.4]. Experimental data showed this proportion was fairly constant for the lower-toughness steels, with the propagation component tending to zero as the toughness increases toward 250 ft-lb (340 J). From these observations, and based on the energy-dissipation principle, Leis [4] developed a correction to the BTCM for the Alliance pipeline project. In this correction, the arrest toughness in terms of CVN energy is found to be the same as that determined by the BTCM if the measured CVN energy is less than 95 J (or 70 ft-lb); otherwise, a non-linear correction between actual arrest toughness and the BTCM-predicted arrest toughness is needed. The Leis correction can be expressed mathematically as: ( CV )arrest = ( CV )BTCM for (CV) < 95 J (4a) The Journal of Pipeline Engineering rib ut io n 262 or d The factor was determined as the to affect a correct prediction via the CSM factor is expressed as a between the arrest toughness predicted toughness: multiplier required the BTCM. Thus, linear relationship and the BTCM ot f See also Equn 4b (below) where (CV)arrest is the CVN full-size equivalent (FSE) energy in Joules required for arrest, and (CV)BTCM is the CVN FSE arrest energy in Joules, calculated using Equn 3. Because the ratio of (CV)arrest/(CV)BTCM determines a correction factor to the BTCM prediction, the Leis correction was often referred to as the Leis correction factor. is t Fig.2. Comparison of arrest Charpy (Cv) toughness predictions by four improved methods. py -n Because the characteristics of the flow and fracture responses of the steels involved in developing Equn 4 had been limited to grades X-70 and below, this equation was limited in its use to X-70 and below [8]. Care must be taken if this correction is used for higher grades of X-80 and above. co Modified Leis correction Sa m pl e Based on recent experimental burst data for X-70 and X-80 pipeline steels, Eiber [11, 12] showed that the Leis correction is accurate for X-70 steels, but was less so for some X-80 steels. If the coefficient of 0.002 in Equn 4b was empirically set at 0.003, Eiber determined that the following modified equation better predicted the arrest toughness for a set of full-scale experimental data covering a range of X-80 pipeline steels, shown in Equn 5 below. (CV)arrest = k (CV)BTCM(6) Such multiplicative factors were determined first for X-80 steels, and then again as X-100 and X-120 pipes were subjected to test. The constant factor k was found to be 1.43 for X-80 and 1.7 or higher for X-100, as given by Demofonti [13]. In contrast to the non-linear form of the Leis correction factor, the CSM factor is linear and specific to the grade and full-scale tests that underlie its determination. Statistical factor As the CSM factor depends on full-scale burst tests and is grade-specific, a more-general factor was sought via a statistical analysis. Wolodko and Stephens [14] at C-FER obtained a statistical correction for testing involving single-phase decompression in grades from X-70 to X-100 in the form: CSM factor = (1.5 + 0.29nsd )( CV ) BTCM (7) ( CV )arrest CSM [13] proposed a simple factor determined by comparing the actual arrest toughness based on results of full-scale burst data to the corresponding BTCM prediction of the minimum CVN arrest energy (CV)BTCM. where nsd is the multiplier on the standard deviation of the model error that can be selected to achieve the desired probability of non-arrest of a ( CV )arrest = ( CV )BTCM + 0.002 ( CV )BTCM − 21.18 2.04 for (CV) ≥ 95 J ( CV )arrest = ( CV )BTCM + 0.003 ( CV )BTCM − 21.18 for (CV) ≥ 95 J 2.04 (4b) (5) 263 rib ut io n 3rd Quarter, 2013 1/ 6 Vf = C σf P − 1 K R Pa (8) ot f Figure 2 compares the arrest-toughness predictions by the four correction methods, i.e. the Leis correction in Equn 4, the modified Leis correction in Equn 5, the CSM factor in Equn 6, and the statistical factor in Equn 7. It is assumed here that X-80 pipeline steels are considered in this figure, and so the factor k = 1.43 in Equn 6, and the factor in Equn 7 is taken as 1.935. For X-80 steels, the arrest toughness generally ranges from 130 to 270 J, and the CSM simple factor gives a reasonable prediction, as shown in Ref. 13. On this basis, Fig.2 shows that the modified Leis correction in Equn 5 predicts a good result of arrest toughness that is close to the CSM factor result from Equn 6. The original Leis correction in Equn 4 slightly underestimates – and the statistical factor in Equn 7 overestimates – the arrest toughness for X-80 gas pipeline steels. As such, the statistical-factor equation not recommended be used, at least for X-80 pipeline steels. Based on their test data, these authors modified Equn 1 of the fracture curve as: or d running fracture. For example, when nsd = 1.0, 1.5, and 2.0, corresponding to the probability of non-arrest of 16%, 6.7%, and 2.3%, the factor between the required arrest toughness and the BTCM prediction used in Equn 7 will be 1.79, 1.935, and 2.08, respectively. is t Fig.3. Effect of speed-dependent fracture toughness on fracture-speed curves (from [17]). m pl e co py -n where: Sa Backfill correction In the fracture curve of the BTCM (Equn 1), the effect of backfill on the fracture velocity is lumped into a ‘backfill coefficient’ that is empirically based, and does not distinguish between different soil types or strengths. In Equn 1, a constant power-law exponent of 1/6 is fixed for all kinds of backfill. In order to improve this, and to consider the backfill effect due to different backfill depths and different soil types, Rudland and Wilkowski [15, 16] conducted a series of burst tests for gas pipelines. K = 0.275 Hactual / Hnominal + 0.725; Hactual is the actual backfill depth used in the burst test of gas linepipes; and Hnominal = 30 in, as was employed in the early gas burst tests conducted by Battelle in the 1970s to calibrate the backfill coefficient C. This modification in Equn 8 appears reasonable, but its practical use is limited because the burst tests in calibration of Equn 8 were conducted only for pipes with toughness CVN < 100 J. Speed-dependent toughness method In the original BTCM, the fracture toughness CVN was assumed as a constant material resistance. Since experiments showed that fracture toughness or resistance is dependent on the fracture speed, Duan and Zhou [17, 18] at TransCanada modified the fracture resistance as a speed-dependent value: R = R0V f− a(9) where: ; Rref is a reference fracture resistance at the reference speed Vref; and α is a fracture-speed-dependent index. The Journal of Pipeline Engineering is t rib ut io n 264 steels. Later, the DWTT was also evaluated as the basis for measuring the ductile-fracture resistance. The larger DWTT specimen with its full-size wall thickness has been considered superior to the smaller CVN specimen in quantifying fracture resistance for ductile pipeline steels with high toughness and large plastic deformation, particularly for modern high-strength pipeline steels. Correlations between the DWTT and CVN energies were developed as the basis for adapting the BTCM for use with DWTT energy, as outlined next. ot f Figure 3 show the effect of speed-dependent fracture toughness on fracture-speed curves for an X-80 pipeline steel, where α = 0 represents the case of constant fracture toughness as used in the BTCM. When α = 0.2, the predictions are in good agreement with full-scale burst tests of X-80 pipes. or d Fig.4. Correlation between Charpy and DWTT specific energies (E/A) (from [20], (1ft-lb/in2 = 0.0021J/mm2). -n Reformulated BTCM Sa m pl e co py As the BTCM remains the only viable tool to quantify arrest toughness, consideration has been given to its reformulation to extent its use for tough higher-grade linepipe steels. Very recently, in a work funded by the China National Petroleum Cooperation [19], the present authors adopted modern elastic-plastic fracture mechanics and in part reformulated the original BTCM. This reformulated BTCM better predicted arrest toughness as compared to the modified Leis correction for high pipeline grade X-80, and also was effective for X-100 when compared to full-scale burst-test data. A simplified equation also was developed, whose predictions are comparable to those by the Reformulated BTCM. However, as this work 1s proprietary, the present paper simply notes it, in passing. Research exploring DWTT-based ductile-fracture-arrest technology Early DWTT-energy methods A drop-weight tear test (DWTT) specimen was developed at Battelle, which was the first candidate considered to replace the Charpy impact specimen to quantify the ductile-to-brittle transition temperature and determine the fracture mode (brittle versus ductile) for pipeline Early DWTT method at Battelle Wilkowski et al. [20, 21] at Battelle developed a linear correlation between the standard pressed-notch (PN) DWTT energy density and the CVN energy density for conventionally rolled steels and quenched and tempered steels in the late 1970s: E E = 3 + 300 A DWTT A CVN (ft-lb/in2) (10) where E is the total fracture energy in ft-lb, A is fracture area of the specimen ligament in in2, and E/A denotes the energy density (or the energy per unit area) in ft-lb/in2. Figure 4 shows the linear relationship between the CVN and DWTT energy densities that is good for these selected steels. When the minimum CVN energy for a fracture arrest is predicted using the BTCM, the minimum DWTT energy for the fracture arrest can be determined from Equn 10. 265 is t rib ut io n 3rd Quarter, 2013 Early DWTT method at British Gas slope of the linear-correlation function continues to decreases from 2.94 for grade X-60 to 1.91 for grade X-100. Therefore, a general correlation between CVN and DWTT is a non-linear function for high-toughness pipeline steels. This observation is consistent with that by Leis [23]. m pl e co py -n ot f Fearnehough et al. [22] at British Gas were among the early investigators who developed test methods and showed that the propagation energy in highertoughness pipeline steels was not linearly related to the Charpy energy. Their work, which preceded that of Wilkowski, involved a series of DWTT specimens that were pre-cracked to different crack lengths – from short to deep – under quasi-static loading, and then impacted under drop-weight dynamic loading. This test was called an ‘interrupted DWTT’ practice. The energy density (i.e. E/A) was determined for this series of interrupted DWTT specimens with differing crack lengths, and compared with the Charpy energy density for the same steel, with the result as shown in Fig.5. This figure indicates that the two energy densities are linearly related for the lower-toughness steels (i.e. data in groups A and B). However, for the higher-toughness steels (with CVN energy of 70 J or more) the E/ADWTT energy becomes strongly nonlinear with E/ACVN (i.e. data in group C). or d Fig.5.Variation of DWTT (E/A) and Charpy (E/A) as developed by Fearnehough (from [22]). Sa Leis [23] discussed the linear correlation between DWTT and CVN energies in Equn 10 for pipeline grades up to X-70, and found that most burst-test data for Alliance pipeline steel X-70 did not follow the linear trend, but behaved in a non-linear relation similar to that in Fig.5. This raises questions about generality of any linear correlation. For modern high-toughness pipeline steels with CVN energy larger than about 100 J, experiments showed that the relation between DWTT and CVN energies deviates from the linearity. Recently, Wilkowski et al. [6] showed that the pipeline grade has significant effect on their correlation. The Brittle-notch DWTT specimens In the 1970s, investigators believed that the non-linear relationship might be caused by the large initiation energy obtained by the standard PN DWTT specimens for tougher steels. Accordingly, Battelle modified the standard PN DWTT specimen in an attempt to reduce the initiation energy from the total DWTT energy – so that the propagation energy is dominant – through use of a brittle-notch (BN) DWTT specimen. Figure 6 shows the results obtained by Wilkowski et al. [20, 24] in 1977 using the BN DWTT specimens for pipeline grades up to X-70. Based on these test data, Wilkowski et al. proposed a curve-fitted nonlinear function between the BN and PN DWTT energy densities: 0.385 E E = 175 − 1500 (11) A BN − DWTT A PN − DWTT where the energy densities and the constant are in ft-lb/in2. An alternative to the brittle-notch DWTT specimen was a termed the static-precracked (SPC) DWTT specimen, which is the interrupted test of Fearnehough, but uses The Journal of Pipeline Engineering is t rib ut io n 266 175 E E = 1.3 3 A DWTT A CVN (W 2000 ) Wilkowski DWTT methods pl e co py -n In order to reflect the non-linear relationship between the DWTT and CVN energies, Wilkowski et al. [6] proposed two new non-linear correlations. They assumed that the standard PN DWTT specimen used for fitting Equn 10 and the BN DWTT specimen were equivalent (because both specimens have less initiation energy in the total absorbed energy). In this case, they replaced the PN DWTT energy in Equn 10 with the BN DWTT energy in Equn 11 and obtained the following so-called Wilkowski 1977 prediction of CVN energy from the standard DWTT tests: 0.385 − 600.0(12) Sa m 175 E E = 3 A DWTT A CVN (W 1977 ) this statistical factor with Equn 12 gives the Wilkowski 2000 prediction of the CVN energy from the standard DWTT tests: ot f a fixed notch length. These DWTT specimens have crack-like notches. It has been shown that the SPC DWTT specimen leads to results similar to the BN DWTT [6]. or d Fig.6.Variation of brittle-notch and pressed-notch DWTT energies for pipeline grades up to X-70 (from [24]). where the specific energy and the constant are in ft-lb/in2, (E/A)DWTT is the total PN DWTT energy density, and (E/A)CVN(W 1977) is the total CVN energy density. When compared with the full-scale burst-test results for pipeline steels of X-52, X-60, X-65, and X-70, Wilkowski et al. [25] found in 2000 that the estimation from Equn 12 resulted in an overestimation of arrest DWTT energy in comparison to the full-scale DWTT data. A statistical factor of the overestimation was determined as 1.291. Combining 0.385 − 600.0 (13) This equation, as well as Equn 12, can be used to estimate the required arrest toughness in terms of DWTT energy when the minimum CVN energy is predicted by the BTCM. From Equns 12 or 13, Wilkowski et al. [6] also obtained the arrest CVN energy by use of the BTCM-predicted CVN toughness, although the authors did not describe the procedure they used for such a prediction. As a result, this predicted CVN result may rely on the quality of the correlation of CVN energy with DWTT energy, and embed the related uncertainty, which later discussion in regard to Fig.10 indicates can be very large. Kawaguchi DWTT method Kawaguchi et al. [26] considered the Wilkowski DWTT equations for use with X-80 linepipe steels. They found that Equn 11 did not match their test data for X-80 because this curve-fit equation was based on test data only up to X-70 (see Fig.6); they therefore proposed the following relationship to correlate the SPC DWTT and the PN DWTT energy densities: 0.9563 E E = 0.9431 A SPC − DWTT A PN − DWTT (ft-lb/in2) (14) 267 is t rib ut io n 3rd Quarter, 2013 Wilkowski CVN correction ot f Following Wilkowski’s assumption and concepts that generated the non-linear correlation as Equn 12, using Equns 10 and 14 Kawaguchi obtained the following correlation between the Charpy and DWTT energy densities for X-80 steels: or d Fig.7. Comparison of three correlations between DWTT and Charpy specific energies. -n 0.9563 2 − 100 (ft-lb/in ) (15) py E E = 0.3144 A A CVN DWTT The literature also includes a so-called Wilkowski CVN correction equation. Motivated by the Leis correction Equn 4b, Papka et al. [27] at ExxonMobil proposed an alternative correction in 2003 in reference to experimental data and the DWTT correlation obtained by Wilkowski et al. [20]. From Equns 10 and 12, and assuming (Cv)(W 1977) = (Cv)BTCM, after the DWTT term was eliminated, the following ‘correction’ for the BTCM-predicted CVN arrest energy was obtained (see Equn 16 below) where the CVN energy and constant are in ft-lb. co Note that both Equns 14 and 15 are nearly linear because the value of the exponent is close to unity. Sa m pl e Figure 7 compares the four correlations between DWTT and Charpy energy densities determined by the Battelle linear correlation in Equn 10, the Wilkowski 1977 correlation in Equn 12, the Wilkowski 2000 correlation in Equn 13, and the Kawaguchi correlation in Equn 15. It is seen from this figure that Equns 12 and 13 have a non-linear dependence between the DWTT and CVN energy densities that diverges from the experimental trend observed in Fig.4. Equation 15 is nearly linear with another slope value, and this opens questions concerning the viability of the three correlations proposed by Wilkowski [6] and that of Kawaguchi [26], with further discussion of this aspect following later. ( CV )arrest = 0.04133 0.138 ( CV )BTCM + 10.29 ( CV )arrest = 0.056 0.1018 ( CV )BTCM + 10.29 2.597 2.597 Wolodko and Stephens [14] at C-FER in 2006 converted this Wilkowski CVN ‘correction’ in Equn 16 from Imperial units to SI units as shown in Equn 17 below, where the CVN energy and the constant are in Joules. Both Equns 16 and 17 have been presented for use in determining the arrest toughness when the BTCM-predicted CVN toughness is used applications involving high-strength pipeline steels. Eiber [11, 12] considered Equn 17 in the evaluation of arrest CVN toughness for X-70 and X-80 pipeline steels, and showed its overestimation of arrest toughness for such steels. − 12.4 (16) − 16.8 (17) The Journal of Pipeline Engineering rib ut io n 268 where: Vc is the fracture velocity in m/s; σf = (σy + σuts)/2 is the flow stress; R = Dp/Ap is the material resistance in J/mm2; Dp is the estimated total energy of PC DWTT specimen in J; Ap is the fracture area of PC DWTT specimen in mm2; P is the decompressed pressure at the crack tip in MPa; Pa is the arrest pressure in MPa; D is the pipe diameter in mm; and t is the pipe thickness in mm. co py -n ot f Figure 8 compares the predictions of minimum CVN arrest toughness obtained by the four CVN-correction methods in applications to X-80 pipeline steel: i.e. the Leis correction factor in Equns 4, the Modified Leis correction in Equn 5, the CSM factor in Equn 6, and the Wilkowski CVN ‘correction’ in Equn 17. It is apparent that the CSM factor prediction is most accurate, which follows because it was developed based on these data. Figure 8 indicates that the Wilkowski ‘correction’ in Equn 17 gives a conservative result that is larger than that obtained by the Modified Leis correction in Equn 5 or the CSM factor in Equn 6. Because the assumption of (Cv)(W 1977) = (Cv)BTCM is questionable, the resulting correction in Equns 16 or 17 is equally questionable, as are its predictions. or d is t Fig.8. Comparison of arrest CVN toughness predictions by four CVN-correction methods Japanese HLP model Sa m pl e In the late 1970s, in parallel to the work done at Battelle on use of the DWTT specimens, Japanese researchers [7, 28] began a large research programme referred to as ‘HLP’. This involved extensive experimental and analytical work that sought to extend the fracture model developed in the BTCM. This work relied on fracture resistance expressed in terms of pre-cracked (PC) DWTT energy. For the soil backfill conditions, fracture velocity in the PC DWTT-based HLP model was determined as: σf P V f = 0.670 − 1 R Pa tσ f Pa = 0.382 D 0.393 (18) 3.81×107 R (19) ] arccos exp[− σ 2f Dt The adaptation of DWTT energy and recalibration of the constant and exponent in the fracture velocity equation for direct use with DWTT energy is one of the features of the HLP model. In order to estimate associated arrest toughness in terms of CVN energy, the Japanese HLP Committee developed a correlation between the CVN energy and PC DWTT energy in the form: D p (estimate) = 3.29t1.5CV 0.544 (20) This correlation was developed based on test results from a variety of linepipes in grades from X-60 to X-100, and with wall thicknesses from 10 mm to 32 mm [29]. Because Equn 18 was calibrated in reference to the full-scale burst-test data for X-70 pipeline steels, its prediction of arrest toughness in terms of DWTT energy is accurate for X-70 pipeline steels, but inaccurate for X-80 pipeline steels [29]. Thus, the applications of the Japanese HLP model to high-strength pipeline steels of grade X-80 and above are not recommended. 269 rib ut io n 3rd Quarter, 2013 Modified HLP model or d is t Fig.9. Actual vs predicted CVN energies by BTCM for X-80 and X-100 high-strength pipeline steels (taken from [13]). γ= 3 t/D 3.22 + 0.20 (25) t0 / D0 ot f To improve the HLP model, Makino et al. [29, 30] investigated the effect of pipe geometry on the prediction of arrest toughness. It was found that the accuracy of arrest toughness prediction by the HLP model depends on pipe diameter to wall thickness ratio D/t, with the modified the HLP model referred to as the Sumitomo model. A new fracture-velocity curve equation was proposed in a general form: 3.42 -n In these three expressions, the reference pipe diameter was set as D0 = 1219.2 mm (48 in) and the reference wall thickness was set as t0 = 18.3 mm. β co 4.57 ×107 R ](22) arccos exp[− σ 2f Dt e tσ f Pa = γ 0.382 D py σf P = Vf α − 1 (21) R Pa Sa m pl where the three parameters α, β, and γ were assumed as a function of the pipe diameter D and the wall thickness t. Note that the coefficient 4.57 in Equn 22 is different from 3.81 in Equn 19. Based on available full-scale burst-test data for high-strength pipeline steels up to X-100, the curve fitting determined these three parameters as: 1/ 4 Dt α = 0.670 D0t0 (23) D β = 0.393 D0 5/ 2 t t0 −1/ 2 (24) The results in Refs 29 and 30 showed that use of the newly developed Equns 21 to 25 led to improved predictions of arrest toughness for high-strength pipeline grades X-100 and X-120 in comparison to the predictions by the original HLP model. However, comparisons of Equns 2, 19, and 22 show that both HLP and improved HLP models (a) use the same BTCM fracture model for determining arrest pressure, and (b) assume that the fracture resistance R = Dp/Ap from PC DWTT energy is equal to R = Cv/Ac in terms of CVN energy. This implies that the PC DWTT energy density is equal to the CVN energy density – which is contrary to the non-linear relationship between these two energy parameters expressed in Equn 20. Thus, the DWTT-based HLP model and the improved HLP model appear to be viable for the grades used in their calibration, but are likely to suffer the same issues that occur for the BTCM in applications to circumstances beyond their calibration database. Further discussion of arresttoughness prediction methods CVN energy-based methods Figure 9 shows the measured CVN energy versus the BTCM-predicted CVN energy for high-strength pipeline steel grades X-80 and X-100 presented by Demofonti et The Journal of Pipeline Engineering rib ut io n 270 The backfill-correction method proposed by Rudland and Wilkowski [15, 16] considered this aspect over a limited range of backfill depths and was focused on pipeline steels with CVN energy less than 100 J. If the backfill height used in an actual burst test for high-strength steels is equal to the original backfill height as used in the BTCM, Equn 8 reduces to Equn 1. As such, there is little improvement made in such applications, which implies this correction has limited general utility. -n ot f al. [13], where experimental burst data were extracted from the CSM database. As evident in this figure, the BTCM underestimates the actual arrest-CVN toughness and leads to non-conservative predictions. Similar results were observed for the simplified BTCM model. Thus, as discussed in regard to Equn 6, the arrest toughness predicted by the BTCM must be multiplied by 1.43 for X-80, and 1.7 for X-100. or d is t Fig.10. Correlation between PN DWTT and CVN energies (from [13]). Combining the results in Figs 2 and 9 shows that: pl e co py a. the modified Leis correction in Equn 5 leads to viable predictions of arrest CVN toughness in comparison to the CSM linear-factor method; b. the original Leis correction factor in Equn 4 slightly underestimates the actual arrest CVN energy (this is not surprising because Equn 4 was calibrated for grades up to X-70); and c. the statistical factor method overly estimates the arrest CVN energy. Sa m While not evident in this paper due to proprietary restrictions, it was found that the reformulated BTCM provides viable predictions for the available full-scale database. It follows that the modified Leis correction factor and – once published – the reformulated BTCM, are the best methods to predict the arrest for X-80 steels. However, these methods are not broadly validated for predictions of the arrest toughness for X-100 steels, and therefore more work is needed in the context of such grades. This is particularly the case in applications where the flow response of the steel shows limited strain hardening, the extensive presence of splits, and limited strain to failure – or has limited through-thickness strength. The speed-dependent toughness method proposed by TransCanada considered the effect of fracture speed on the fracture resistance in terms of CVN or DWTT energy, which is consistent with the common understanding that dynamic fracture toughness depends on the strain rate. While such considerations offer the potential to improve predictions based on the BTCM through use of a speed-dependent resistance, it is not clear how to quantify in general the reference resistance, the reference speed, and the fracture speed and a constant index. DWTT energy-based methods Figure 10 shows broad experimental relations between the standard PN DWTT energy density and the CVN energy density for various high-strength pipeline grades. This figure includes data for X-80 and X-100 high-strength pipeline steels as developed by Demofonti et al. [13] (for which the full-scale data were extracted from the CSM database), along with the linear correlation of Equn 10. It is apparent that the experimental trend in Fig.10 begins to deviate strongly from linear response at a CVN energy density of approx. 120 J/cm2 (or 570 ftlb/in2), which corresponds to a FSE CVN energy of 3rd Quarter, 2013 271 io n rib ut References 1. X.-K.Zhu, 2013. Existing methods in ductile fracture propagation control for high strength gas transmission pipelines. Proc. ASME Pressure Vessel and Pipeline Conference (PVP 2013), July, Paris, France. 2. W.A.Maxey, 1974. Fracture initiation, propagation, and arrest. Proc. 5th Symposium on Line Pipe Research, November, Houston, USA. 3. R.J.Eiber, T.A.Bubenik, and W.A.Maxey, 1993. Fracture control technology for natural gas pipelines. NG-18 Report 208, Pipeline Research Council International, Project PR-3-9113, Battelle. 4. B.N.Leis et al., 1998. Relationship between apparent Charpy Vee-notch toughness and the corresponding dynamic crack-propagation resistance. International Pipeline Conference, Calgary Canada, pp. 723-732. 5. G.Mannucci, G.Demofonti, D.Harris, L.Barsanti, and H.-G.Hillenbrand, 2001. Fracture properties of API X100 gas pipeline steels. Proc. 13th Biennial Joint Technical Meeting on Pipeline Research, 30 April-4 May, New Orleans, USA. 6. G.Wilkowski, D.L.Rudland, H.Xu, and N.Sanderson, 2006. Effect of grade on ductile fracture arrest criteria for gas pipelines. Proc. International Pipeline Conference, Canada. Paper IPC006-10350. 7. E.Sugie et al., 1982. A study of shear crack propagation in gas-pressurized pipelines. J. Pressure Vessel Technology, 104, p 338. 8. B.N.Leis and R.J.Eiber, 2013. Fracture control technology for transmission pipelines. Battelle Report on PRCI Projects PR-003-00108 and PR003-084506 (update of Reference 3). py -n ot f Beyond the extent of linearity between the CVN and DWTT energy densities, the trend of the experimental results shown in Fig.9 becomes non-linear, and this bent-over trend follows the same tendency to the experimental trend presented for the interrupted (precracked) DWTT results shown in Fig.5. In contrast, the three correlations between DWTT and CVN energy densities proposed by Wilkowski and by Kawaguchi have a non-linear dependence that deviated from linearity in the opposite direction – that is it is bent-up. This opens to question the utility of Equns 12 to 15: similar uncertainty exists in regard to Equns 16 and 17. It was noted that the modified Leis correction coupled with the BTCM and a newly developed reformulated BTCM provide viable methods to determine arrest toughness, at least for high-strength pipeline steels (X80). However, given that such methods are not broadly validated for predictions of the arrest toughness for X-100 steels, more work is needed in the context of such grades. This is particularly the case in applications where the flow response of the steel shows limited strain hardening, the extensive presence of splits, and limited strain to failure – or has limited throughthickness strength. is t In reference to Equn 10, the just-noted limit on linearity at a CVN E/A of approx. 120 J/cm2 (or 570 ft-lb/in2) corresponds to a PN DWTT E/A of 420 J/cm2 (2000 ft-lb/in2). It is evident in Fig.6 that the BN DWTT E/A response also begins to deviate from linearity as a function of CVN E/A at roughly this same level (i.e. 2000 ft-lb/in2). It follows that the trends for the PN DWTT and the BN DWTT do not differ significantly as a function of the CVN E/A. The onset of this nonlinearity occurs at a FSE CVN energy of approx. 95 J (or a CVN E/A of approx. 120 J/cm2 (570 ft-lb/in2)), or a PN DWTT energy density of 420 J/cm2 (approx. 2000 ft-lb/in2). This implies that little to no benefit accrues to the use of the BN (or a PC) DWTT in lieu of a PN DWTT or a CVN specimen. In turn, this indicates that the use of such a geometry/test practice will not offset the inherent calibration issues with the BTCM as the toughness increases. The results showed that some of the existing non-linear models for correlating DWTT and CVN energy densities are open to question. For example, the trends for the PN DWTT and the BN DWTT were shown do not differ significantly as a function of the CVN E/A, which opens to question the utility of the BN DWTT practice, and correlations that embed it. In turn, it was indicated that the use of such a geometry/test practice will not offset the inherent calibration issues problems with the BTCM as the toughness increases. or d approx. 95 J (or 70 ft-lb). Leis and Eiber [8] present an equally extensive dataset in the context of grade X-65 and below, which also shows this same trend. pl e co In addition to the discussions and analyses above, other relevant discussions can be found in a recent review by Mannucci and Demofonti [31]. In particular, these authors discussed the applications of the CVN energy and the DWTT energy in the ductile-fracture propagation control for X-80 gas transmission line pipes. m Conclusions Sa This paper discussed work regarding the use of CVN and DWTT energy in applications to the arrest of ductile-fracture propagation, and evaluated related methods to quantify arrest toughness for hightoughness, higher-strength, gas-transmission pipelines. The BTCM was reviewed along with corrections and correlations to offset its shortcomings as toughness increases. These included the Leis correction, the CSM factor, and Wilkowski’s DWTT correlations, which were evaluated through comparison with the full-scale fracture-propagation test data for higher-grade pipeline steels including X-80 and X-100. 272 The Journal of Pipeline Engineering or d is t rib ut io n 21. G.Wilkowski, 1979. Fracture propagation toughness measurements. Proc. 6th Symposium on Line Pipe Research, Paper K, Houston, USA. 22.G.D.Fearnehough, D.T.Dickson, and D.G.Jones, 1976. Dynamic toughness determination in ductile materials. The Dynamic Fracture Toughness Conference, London, UK. 23.B.N.Leis, 2002. Evolution of line-pipe steel and its implications for transmission pipeline design. Proc. International Pipeline Conference, Calgary, Canada. 24.G.Wilkowski, W.A.Maxey, and R.J.Eiber, 1978. Problems in using the Charpy, dynamic tear test and drop weight tear test for high toughness quenched and tempered steels. In: What does the Charpy test really tell us? ASM. 25.G.Wilkowski, Y.Y.Wang, and D.L.Rudland, 2000. Recent development on determining steady-state dynamic ductile fracture toughness from impact tests. Proc. 3rd International Pipeline Technical Conference, Bruges, Belgium, 1, pp 359-386. 26.S.Kawaguchi et al., 2004. Application of X80 in Japan: fracture control. Proc. 4th International Conference on Pipeline Technology, Ostend, Belgium. 27.S.D.Papka et al., 2003. Full-size testing and analysis of X120 linepipe. Proc. 13th International Offshore and Polar Engineering Conference, Hawaii, USA. 28.H.Makino et al., 2001. Prediction for crack propagation and arrest of shear fracture in ultrahigh pressure natural gas pipelines. ISIJ Int., 41, pp 381-388. H.Makino, I.Takeuchi, and R.Higuchi, 2009. 29. Fracture arrestability of high pressure gas transmission pipelines by high strength pipelines. Proc. Pipeline Technology Conference, Ostend, Belgium. 30.R.Higuchi, H.Makino, and I.Takeuchi, 2009. New concept and test method on running ductile fracture arrest for high pressure gas pipelines. Proc. 24th World Gas Conference, Buenos Aires, Argentina. 31.G.Mannucci and G.Demofonti, 2011. Control of ductile fracture propagation in X80 gas linepipe. Journal of Pipeline Engineering, 10, pp 133-145. Sa m pl e co py -n ot f 9. W.A.Maxey, J.F.Kiefner, R.J.Eiber, and A.R.Duffy, 1972. Ductile fracture initiation propagation and arrest in cylindrical vessels. Fracture toughness, ASTM STP 514, Part II, pp 347-362. 10. W.A.Maxey, J.F.Kiefner, and R.J.Eiber, 1975. Ductile fracture arrest in gas pipelines. Final Report for PRC Project NG-18, Report no 100, December. 11.R.J.Eiber, 2008. Fracture propagation – 1: Fracturearrest prediction requires correction factors. Oil & Gas Journal, 106, 39, 20 October. 12.Idem, 2008. Ibidem, 40, 27 October. 13.G.Demofonti, G.Mannucci, and P.Roovers, 2007. Existing methods for the evaluation of material fracture resistance for high grade steel pipelines. Proc. PRCI-EPRG-APRA 16th Biennial Joint Technical Meeting on Pipeline Research, March, Canberra, Australia. 14.J.Wolodko and M.Stephens, 2006. Applicability of existing models for predicting ductile fracture arrest in high pressure pipelines. Proc. International Pipeline Conference, Calgary, Canada. Paper IPC2006-10110. 15.D.L.Rudland and G.Wilkowski, 2007. Effects of backfill soil properties and pipe grade on ductile fracture arrest. Proc. PRCI-EPRG-APRA 16th Biennial Joint Technical Meeting on Pipeline Research, March, Canberra, Australia. 16.Idem, 2006. The effects of soil properties on the fracture speeds of propagating axial crack in line pipe steels. Proc. International Pipeline Conference, Calgary, Canada. Paper IPC2006-10086. 17.D.M.Duan and J.Zhou, 2009. Speed dependent fracture toughness and the effect on fast ductile fracture propagation in gas pipelines. Proc. 12th International Conference on Fracture, July, Ottawa, Canada. 18.D.M.Duan, J.Zhou, D.J.Shim, and G.Wilkowski, 2010. Effect of fracture speed on ductile fracture resistance – Part 2: Results and application. Proc. 8th International Pipeline Conference, Calgary, Canada. X.-K.Zhu and B.N.Leis, 2012. Fracture arrest 19. toughness analysis for high-grade gas pipelines. Battelle Final Report to Tubular Goods Research Center of China National Petroleum Cooperation, August. 20. G.Wilkowski, W.A.Maxey, and R.J.Eiber, 1977. Use of a brittle notch DWTT specimen to predict fracture characteristics of line pipe steels. ASME 1977 Energy Technology Conference, Paper 77Pet-21, Houston, USA, September Are you up to speed? Training courses – 2013 io n 0 0 0 2 0 1 2 6th Pipeline Technology Conference (Ostend, Belgium) 20-23 October 2nd Best Practices in Pipeline Operations and Integrity Management Conference (Bahrain) 11-12 November DOT Pipeline Safety Regulations - Overview and Guidelines for Compliance (Houston) 11-12 November Pipeline Repair Methods, Hot Tapping, and In-Service Welding (Houston) 11-13 November Pipeline Integrity Management (Houston) 11-15 November Onshore Pipeline Engineering (Houston) 11-15 November November Practical pigging training course (Houston) 11-15 November Subsea Production Systems Engineering (Aberdeen) 11-15 November Advanced Pipeline Risk Management (Houston) 11-15 November Stress Corrosion Cracking in Pipelines (Houston) 18-20 November ot f or d is t Subsea production systems (Houston) 6-9 October -n NOV 2013 rib ut TRAINING OCT 2013 1-4 October 2013 18-22 November Subsea Pipeline Engineering (Houston) Defect Assessment Calculations Workshop (Houston) Pigging & In-line Inspection (Calgary) 2-4 December Defect Assessment in Pipelines (Calgary) 4-6 December Pipeline Integrity Management (Calgary) 4-6 December Defect Assessment Calculations Workshop (Calgary) 5-6 December Geohazard Management for Pipeline Engineers (Calgary) Sa m pl e DEC 2013 2-3 December co py 20-22 November Defect Assessment in Pipelines (Houston) Working with a faculty of 38 leading industry experts, Clarion and Tiratsoo Technical are privileged to provide some of the best available industry based technical training courses for those working in the oil and gas pipeline industry, both onshore and offshore. Complete syllabus and registration details for each course are available at: www.clarion.org e pl m Sa py co io n rib ut is t or d ot f -n
© Copyright 2024