RESEARCH REPORT 191Weld Repair Procedures and Techniques: These papers present weld repair...

50
HSE Health & Safety Executive Integrity of Repaired Welds (Phase 1) - Deliverable 5 Summary Report Prepared by Serco Assurance for the Health and Safety Executive 2004 RESEARCH REPORT 191

Transcript of RESEARCH REPORT 191Weld Repair Procedures and Techniques: These papers present weld repair...

  • HSEHealth & Safety

    Executive

    Integrity of Repaired Welds (Phase 1)

    - Deliverable 5 Summary Report

    Prepared by Serco Assurance for the Health and Safety Executive 2004

    RESEARCH REPORT 191

  • HSEHealth & Safety

    Executive

    Integrity of Repaired Welds (Phase 1)

    - Deliverable 5 Summary Report

    J K Sharples, L Gardner S K Bate, R Charles

    Serco Assurance Birchwood Park

    Warrington Cheshire WA3 6AT

    J R Yates The University of Sheffield

    Sheffield S1 3JD

    M R Goldthorpe M R Goldthorpe Associates

    The Grange 2 Park Vale Road

    Macclesfield Cheshire

    SK11 8AR

    This report summarises work that has been undertaken by Serco Assurance (formerly AEA Technology Consulting), The University of Sheffield and M R Goldthorpe Associates, on behalf of the Health and Safety Executive. It describes Phase 1 of a proposed multi-stage project aimed at

    (i) providing general guidance on when welded repairs may or may not be beneficial, and,

    (ii) proposing a suitable engineering procedural method for assessing the integrity of repaired welds ona case-bycase basis. Welds considered are appropriate to ferritic material.

    This report and the work it describes were funded by the Health and Safety Executive (HSE). Its contents, including any opinions and/or conclusions expressed, are those of the authors alone and do not necessarily reflect HSE policy.

    HSE BOOKS

  • © Crown copyright 2004

    First published 2004

    ISBN 0 7176 2800 0

    All rights reserved. No part of this publication may be

    reproduced, stored in a retrieval system, or transmitted in

    any form or by any means (electronic, mechanical,

    photocopying, recording or otherwise) without the prior

    written permission of the copyright owner.

    Applications for reproduction should be made in writing to:

    Licensing Division, Her Majesty's Stationery Office,

    St Clements House, 2-16 Colegate, Norwich NR3 1BQ

    or by e-mail to [email protected]

    ii

  • CONTENTS

    EXECUTIVE SUMMARY v

    INTRODUCTION 1

    TASK 1 – REVIEW OF CURRENT INDUSTRIAL PRACTICES AND PREVIOUS

    PROBLEMS AND ASSESSMENT OF INFORMATION CONTAINED IN THE

    LITERATURE 3

    TASK 2 – SCOPING CALCULATIONS TO ESTABLISH MATRIX OF CASES TO

    CONSIDER 5

    TASK 3 – WELD/SPECIMEN MANUFACTURE 7

    TASK 4 – MATERIAL CHARACTERISATION TESTS 9

    Tensile Tests 9

    Fracture Tests 9

    Fatigue Crack Growth Tests 10

    Metallography And Hardness Testing 10

    Microstructural Examination 10

    TASK 5 – RESIDUAL STRESS MEASUREMENTS 12

    TASK 6 – TESTS INVOLVING PHOTOELASTIC COATING AND THERMAL EMISSION

    METHODS 13

    TASK 7 – DEVELOPMENT OF FINITE ELEMENT MODELS 16

    Weld Modelling Technique 16

    Material Properties 17

    Results of Welding Simulations 18

    Analyses of Defects In The Simulated Welds 19

    TASK 8 – APPLICATION OF FINITE ELEMENT MODELS TO MATRIX CASES 21

    Edge Defects in the Welded Plate 21

    Equatorial Defects in the Welded Sphere 24

    Embedded Defects in the Welded Plate 25

    TASK 9 – ASSESSMENT BY ENGINEERING PROCEDURE METHODS 28

    General Methodology 28

    iii

  • 38

    Edge Cracks 29

    Embedded Cracks 36

    TASK 10 – PROVISIONAL GUIDANCE ON WELD REPAIRS

    Practical Issues 38

    Guidance Resulting From The Finite Element Calculations 39

    TASK 11 – PROVISIONAL GUIDANCE ON ENGINEERING PROCEDURE METHOD 41

    Route for assessing the significance of a flaw in a weld (as-welded, PWHT or

    repaired weld) 41

    Route for assessing whether repairing a weld Is likely lo be beneficial 41

    Critical Crack Size Evaluation 41

    Crack Growth Evaluation 42

    TASK 12 – RECOMMENDATIONS FOR FUTURE PHASES OF PROJECT 44

    REFERENCES

    FIGURES

    APPENDIX 1 – LITERATURE REVIEW

    APPENDIX 2 – MICROSTRUCTURAL EXAMINATION OF WELD SAMPLES

    UNDERTAKEN BY SHEFFIELD UNIVERSITY METALS ADVISORY CENTRE (SUMAC)

    iv

  • EXECUTIVE SUMMARY

    This report summarises work that has been undertaken by Serco Assurance (formerly AEA

    Technology Consulting), The University of Sheffield and M R Goldthorpe Associates, on behalf of

    the Health and Safety Executive. It describes Phase 1 of a proposed multi-stage project aimed at (i)

    providing general guidance on when welded repairs may or may not be beneficial, and, (ii) proposing

    a suitable engineering procedural method for assessing the integrity of repaired welds on a case-by-

    case basis. Welds considered are appropriate to ferritic material.

    The project has centred on detailed finite element modelling of a matrix of relevant un-repaired and

    repaired weld configurations. Development and validation of the finite element models have been

    undertaken by way of mechanical testing; involving photoelastic coating and thermal emission

    methods. A number of material characterisation tests have been performed and residual stress profiles

    have been evaluated experimentally. Metallurgical examination has also has also been carried out in

    order to examine the changes in microstructure resulting from the welding process.

    The project has involved the following tasks:

    Task 1 – Review of current industrial practices and previous problems and assessment of

    information contained in the literature.

    Task 2 - Scoping calculations to establish a matrix of cases to consider.

    Task 3 – Weld/specimen manufacture.

    Task 4 – Material characterisation tests.

    Task 5 – Residual stress measurements.

    Task 6 – Tests involving photoelastic coating and thermal emission methods.

    Task 7 – Development of finite element models.

    Task 8 – Application of finite element models to matrix cases.

    Task 9 – Assessment by engineering procedure methods.

    Task 10 – Provisional guidance on weld repairs.

    Task 11 – Provisional guidance on engineering procedure method.

    Task 12 – Recommendations for future phases of project.

    The report constitutes the final deliverable (Deliverable 5) of this phase 1 project. The main results of

    Deliverables 1 to 4 are summarised and the outline of the work and results are presented for Tasks 9,

    10, 11 and 12.

    v

  • 1. INTRODUCTION

    Repair welds are commonly carried out in industry on components where flaws or defects have been

    found in weldments during in-service inspection. However, in some cases the process may actually

    have a deleterious effect on the residual lifetime of a component. This can be due to metallurgical

    changes in the component material in the vicinity of the repair and because of very high residual

    stresses which can be introduced in the repaired region.

    A Serco Assurance (formerly AEA Technology Consulting) led consortium involving (in addition to

    Serco Assurance) The University of Sheffield and an independent consultant, M R Goldthorpe

    Associates, has undertaken Phase 1 of a proposed multi-stage project aimed at (i) providing general

    guidance on when welded repairs may or may not be beneficial, and, (ii) proposing a suitable

    engineering procedural method for assessing the integrity of repaired welds on a case-by-case basis.

    Welds considered are appropriate to ferritic material.

    The project has centred on detailed finite element modelling of a matrix of relevant un-repaired and

    repaired weld configurations. Development and validation of the finite element models have been

    undertaken by way of mechanical testing; involving photoelastic coating and thermal emission

    methods. A number of material characterisation tests have been perfiormed and residual stress

    profiles have been evaluated experimentally. Metallurgical examination has also has also been

    carried out in order to examine the changes in microstructure resulting from the welding process.

    The project has involved the following tasks:

    Task 1 – Review of current industrial practices and previous problems and assessment of

    information contained in the literature.

    Task 2 - Scoping calculations to establish a matrix of cases to consider.

    Task 3 – Weld/specimen manufacture.

    Task 4 – Material characterisation tests.

    Task 5 – Residual stress measurements.

    Task 6 – Tests involving photoelastic coating and thermal emission methods.

    Task 7 – Development of finite element models.

    Task 8 – Application of finite element models to matrix cases.

    Task 9 – Assessment by engineering procedure methods.

    Task 10 – Provisional guidance on weld repairs.

    Task 11 – Provisional guidance on engineering procedure method.

    Task 12 – Recommendations for future phases of project.

    1

  • The various components (i.e. Tasks) of the project, together with their dependencies, are contained in

    the flow diagram of Figure 1.

    Reports, constituting Deliverables 1 to 4, have previously been issued that outline the work and

    results of Tasks 1 to 8.

    This report constitutes the final deliverable (Deliverable 5) of this phase 1 project. The main results of

    Deliverables 1 to 4 are summarised and the outline of the work and results are presented for

    (i) asessment by engineering procedures (Task 9), (ii) provisional guidance on weld repairs (Task 10),

    (iii) provisional guidance on weld procedures (Task 11) and, (iv) recommendations for future phases

    of the project.

    2

  • 2. TASK 1 – REVIEW OF CURRENT INDUSTRIAL PRACTICES AND

    PREVIOUS PROBLEMS AND ASSESSMENT OF INFORMATION

    CONTAINED IN THE LITERATURE

    A draft report of the literature review carried out under Task 1 has been previously issued. An

    updated version of this report is included as Appendix 1.

    The papers reviewed can be categorised as folllows:

    Numerical analysis: These relate to the prediction of residual stresses in weldments.

    Case Studies: These papers discuss the metallurgical examination of repair welds and the evaluation

    of found defects.

    Weld Repair Procedures and Techniques: These papers present weld repair techniques.

    Performance of Repair Welds: An assessment of how various weld repairs have performed in service.

    The review has indicated that defects in welded structures can occur during the fabrication process

    due to ‘workmanship’ or in-service due to working conditions. During fabrication, PD5500 states that

    ‘unacceptable imperfections shall be either repaired or deemed not to comply with this standard’.

    Repair welds have to be carried out to an approved procedure and subjected to the same acceptance

    criteria as the original weld. Thus all welds have to satisfy the requirements of the design

    specification before acceptance by the purchaser or inspecting authority.

    For defects found in-service there are no standard guidelines available for utilities to use to make a

    decision on the need to carry out a weld repair. An industrial survey carried out by EPRI for utilities

    in the United States has shown that utilities will rely on the original manufacturer or outside vendors

    to assist on this decision. However, it is not clear that the assessment procedures used are consistent

    or are indeed reliable. In the UK, the repair of welds appears to rely on in-house experience in the

    absence of guidelines to follow. However, this review showed that re-cracking of repair welds still

    occurs due to lack of understanding on why original defects have occurred and how they should be

    repaired.

    Whilst the decision to repair a defect may be aided using an assessment procedure, the practical

    considerations identified in a paper by Jones could also usefully be considered. These show that

    repair welds should be considered on a case-by-case behaviour, therefore a definitive set of ‘rules’ can

    not be given. Instead, the guidelines need to be produced which provide good practice in assessing

    defects in welds and the requirements for carrying out a ‘safe’ repair.

    A number of References were found illustrating the capabilities of performing a repair weld without

    the need for PWHT. This was introduced by the half-bead technique defined in ASME XI primarily

    for the nuclear industry. This has been superseded by other temperbead techniques, which are all

    aimed at improving the properties within the weld HAZ, whilst saving time and costs by precluding

    the time for post-weld heat treatment (PWHT). There is evidence that this method is employed by

    other industries in the USA, but it is unclear on the use of this practice in the UK.

    In the references associated with case studies and the performance of weld repairs, only a few of them

    are related to residual stresses. These papers have indicated that the magnitude of residual stresses in

    3

  • repair welds can be of yield magnitude. The most recent advances in welding simulation were

    presented at an e IMechE conference in November 1999. The conference demonstrated the

    developments that had been made, mainly in the use of finite element analysis to predict residual

    stresses. Sufficient confidence in numerical analysis needs to be demonstrated by making comparison

    with measurement methods.

    When developing guidelines for the assessment of defects in repair welds, sufficient advice needs to

    be given to the user as to when residual stresses need to be considered in the assessment. Advice also

    needs to be provided on when the user should use simple approximations of the residual stress pattern,

    e.g. upper bound profiles given in BS7910, or to use finite element analysis techniques to predict the

    complex behaviour of the material during welding.

    4

  • 3. TASK 2 – SCOPING CALCULATIONS TO ESTABLISH MATRIX OF

    CASES TO CONSIDER

    A detailed section on this Task is included in the Deliverable D2 report previously issued (Reference

    1).

    That section outlines the matrix of analysis cases planned to be undertaken in Task 8. These cases

    were developed to illustrate the effect, on fatigue life or load margin, of either leaving a weld defect

    in place or repairing it in-situ. Only in-service repairs would be considered. Since such comparisons

    were only possible by considering the behaviour of defects, it was assumed that a defect inadvertently

    remains in the weld after “repair”. This may or may not grow by fatigue during subsequent service.

    The double V-preparation weld in plate (Figure 2) used in the experimental work of the project would

    be studied in Task 8. It is a relatively simple weld geometry, but this would enable a large range of

    analysis cases to be considered and so illustrate to non-experts, the effects that different parameters

    could have on the decision to repair a defective weld.

    The repair procedure carried out on the test plate in Task 3 is shown in Figure 3. This procedure was

    considered to be representative of an in-situ weld repair. The repair depth is 15 mm in order to

    simulate the grinding out of the weld 2 mm beyond an assumed defect with a depth of 13 mm. There

    was lower heat input than a shop repair, using no pre-heat and smaller electrodes. Strong-back plates

    were used to simulate the restraint on the surrounding structure and no PWHT was carried out.

    For Task 8, it was intended to carry out a variety of mainly two-dimensional, plane stress, finite

    element analyses. Comparisons would be made between simulations of un-repaired and repaired

    situations for a range of different parameters that affect fatigue life or margin on load. The intention

    was to illustrate the transition from cases where the defects are best left in place, to cases where repair

    is required. Since comparisons would be made between the un-repaired and repaired situations,

    simplified, two-dimensional, plane stress analysis would be capable of illustrating the role of different

    parameters in the repair decision.

    The base case would be a 40 mm thick plate, with an alternative thickness of 20 mm.

    For simplicity, defects would be considered to lie in a plane normal to the surface of the plate and

    through the middle of the weld. The repair evacuation would be symmetrical with respect to the

    middle of the weld.

    The base case for the un-repaired condition would be a surface breaking defect in the weld root as

    shown in Figure 4. The defect depth would be equal to one third of the plate thickness. Alternative

    cases would consider surface breaking weld root defects with different depths, covering the range

    from the minimum detectable by NDT (about 3 mm) to one half of the plate thickness.

    Embedded defects in the un-repaired condition (Figure 5) would also be considered. The base case

    would be an embedded defect having a total height equal to one third of the plate thickness and

    symmetrically positioned about the weld throat. Alternative cases would consider different defect

    heights and position relative to the weld.

    Figures 6 and 7 show the case of ‘wide’ and ‘narrow’ excavations that would be studied. These were

    considered to bound the repair procedure specified in Figure 3.

    5

  • Various defects remaining after the repair would be studied. Generally, these defects would be

    smaller in height than those in the un-repaired condition. The base case for repaired weld defects

    would be an embedded one caused by incomplete excavation, as indicated in Figures 6 or 7. Various

    defect heights would be considered, ranging from a minimum of 3 mm to a maximum smaller than the

    un-repaired size.

    Alternative analysis cases for the repaired condition would consider different surface defects

    remaining after improper repair of pre-existing surface defects (Figure 8) and embedded defects

    resulting from improper repair of embedded defects (Figure 9). Although in practice the former are

    likely to be weld toe cracks, the analyses would consider cracks situated in the middle of the weld.

    Figures 10 to 12 show the different defect configurations it was intended to analyse for the 40 mm

    thick plate, and Figures 13 to 15 show the defects for the 20 mm thick plate. Table 1 gives a summary

    of the un-repaired and repaired defect sizes with a code for each case. The finite element

    computations would actually consider a large range of defect sizes in order that calculations of fatigue

    crack growth could be undertaken.

    In addition to the geometrical parameters referred to above, the planned matrix of cases contained

    variations in tensile properties, fracture toughness, residual stresses and service stresses (service

    stresses would be simulated in the plate geometry by applying a tensile stress transverse to the weld).

    The variations in these parameters are included in Table 1.

    As will be seen in Section 9, the finite element analysis covered a good selection of the cases

    described above that were proposed under Task 2.

    6

  • 4. TASK 3 – WELD/SPECIMEN MANUFACTURE

    A detailed section on this Task is included in the Deliverable D2 report of Reference 1.

    Motherwell Bridge Group was contracted to prepare a suitable welded steel plate using materials and

    welding/repair processes typical of current industrial practice. They used available steel plate of

    thickness 40mm to BS1501 490 LT50. The weld procedure qualification record is shown in Figure 2.

    An asymmetric double “V” preparation was used with the weld root positioned 2/3 of the plate

    thickness from the surface of side 1, which was filled first. Typical pre-heat and interpass

    temperatures were used of 75°C and 250°C respectively. No PWHT was carried out. Visual

    inspection, Magnetic Particle Inspection (MPI) and ultrasonic testing confirmed that there were no

    detectable defects after welding.

    The test plate is shown in Figure 16 and comprised two 40 mm thick plates, with length 4000 mm and

    width 500 mm, welded together at the long edges. Half of the welded plate, (i.e. a 2000 mm length)

    was cut into five sections as shown to provide as-welded material for the experimental work under

    project Tasks 4 (material characterisation), 5 (residual stress measurements) and 6 (photoelastic and

    thermal emission experiments), along with two blanks for manufacture of further test specimens in a

    later phase of the project. Motherwell Bridge Group retained the remaining half of the test plate for

    repair weld processing described below.

    Strong back plates, made from the same material as the test plate, were used to restrain out of plane

    bending during welding. The strong-back plates formed 40 mm thick ribs, 400 mm high, running

    across the full 1000 mm width of the test plate on the opposite side to that being welded. Each

    strong-back plate was attached to the test plate by fillet welds, which extended for 300 mm from each

    end. A central 150 mm cut out was formed to accommodate pre-heaters in the case of the original

    weld only. Eight strong-back plates were used for the original welding of the 4000 mm long test

    plate, placed at 500mm intervals, commencing 250 mm from the end. The strong-back plates were

    fixed to test plate side 2 whilst welding side 1 and vice versa.

    The weld repair process carried out on the second 2000 mm length of test plate (Figure 16) was

    designed to simulate the site repair of a central root defect in the original weld. This involved typical

    grinding out from the narrower side of the weld (side 2) to a depth of 17 mm to ensure removal of a

    defect in the original weld root at a depth of 13.5 mm. The weld procedure qualification record for

    the repair weld is shown in Figure 3.

    To simulate a repair process being applied to a structure on site, rather than under ideal workshop

    conditions, some modifications were agreed to the weld procedure. Welding under more difficult

    access conditions was simulated by use of smaller electrodes and more rapid passes with less “weave”

    than was the case for the original weld. This process (known as “stringer bead” technique) resulted in

    a lower heat input than for the original weld. This was exacerbated by the omission of pre-heat for

    the repair, simulating a site situation where pre-heat could be difficult to apply effectively. Lower

    heat input results in more rapid cooling of the weld metal, which can lead to changes in the material

    properties. No PWHT was carried out following the repair welding. Visual inspection, MPI and

    ultrasonic testing confirmed that there were no detectable defects after repair.

    For the repair weld, four strong-back plates of the type used for the original weld were attached to

    simulate structural restraint. These were set at 500 mm spacing on the 2000 mm long test plate, fixed

    to side 1 only as the repair was single sided.

    7

  • 8

  • 5. TASK 4 – MATERIAL CHARACTERISATION TESTS

    A detailed section on this Task is included in the Deliverable D2 report of Reference 1. Initial

    material characterisation tests, covered in Reference 1, were those to determine tensile, fracture and

    fatigue crack growth properties. The results of metallography and hardness testing are also presented

    in Reference 1. Narrow bands of high hardness were measured in the heat affected zone (HAZ) of the

    samples (see below). To provide an understanding of the formation of these, it was decided to carry

    out a more detailed microstructural examination of the welded regions in samples for both the as-

    welded and weld repair specimens.

    Results of all the material characterisation tests are summarised as follows:

    5.1 TENSILE TESTS

    Tensile properties of the weld material in both as-welded and repair-welded conditions at room

    temperature were obtained from tests on 3.5 mm diameter round bar specimens.

    The tensile test results are listed in Table 2. True stress/true strain data are given in Reference 1.

    The results show that, in the as-welded state, the weld was overmatched by 46%, based on the 0.2%

    proof stress (PS) values of approximately 512 MPa and 350 MPa for weld and parent plate

    respectively. The parent material exhibited typical upper and lower yield point behaviour, which was

    not present in the weld metal results. The ultimate tensile stress (UTS) for the weld was 18% higher

    than that for the parent material, with average values of 622 MPa and 527 MPa respectively.

    For the repair weld material, higher values of 0.2%PS were obtained compared to the as-welded

    condition. The near surface average value for repair weld was 540 MPa compared to 512MPa for the

    as-welded condition (5% increase), whilst the near root average value for repair weld was 580MPa

    (13% increase). The UTS value obtained from near surface repair weld was similar to that for the as-

    welded condition (628 MPa against 622 MPa respectively), whilst the value for near root repair weld

    was 670 MPa (approximately 13% increase on as-welded). It should be noted that a spurious result

    was obtained from repair weld specimen WI12, due to failure outside the gauge length, and this has

    therefore been discounted.

    5.2 FRACTURE TESTS

    Fracture toughness J resistance curves at room temperature were obtained from single edge notch

    bend (SENB), side grooved, unloading compliance specimens to BS 7448 Part 4 for the original weld

    and the repair weld. Two specimens were tested in each condition. The specimen notch was aligned

    centrally in the through-thickness direction. The specimen orientation was selected, and the initial

    crack length, after fatigue pre-cracking, adjusted within the standard limits, to ensure that the crack tip

    lay in original weld or repair weld, as desired.

    The results are shown in the crack growth resistance curves of Figures 17 and 18, for as-welded and

    repair-welded material respectively. The results showed that the fracture toughness behaviour was

    similar in both the as-welded and repair-welded specimens, with initiation toughness J0.2 values of -2

    approximately 105 kJm-2

    and 102 kJm respectively (allowing for blunting, J0.2BL of approximately -2

    116kJm-2

    and 119kJm respectively).

    9

  • 5.3 FATIGUE CRACK GROWTH TESTS

    Fatigue crack growth properties at room temperature were obtained for the original weld and the

    repair weld using Compact Tension (CT) specimens in accordance with ASTM E647. The specimen

    notch was aligned centrally in the weld in the through-thickness direction. The specimen orientation

    was selected, and the initial crack length adjusted within the standard limits, to ensure that crack

    growth was obtained in original weld or repair weld, as desired.

    The results of the fatigue crack growth tests on weld metal are shown in the Paris Law plots of Figure

    19. The data indicate that similar fatigue crack growth behaviour was obtained with both the as-

    welded and repair-welded material. The slopes of the Paris Law plots are very similar, with some

    offset giving slightly higher growth rates with the as-welded material. The valid region of stress

    intensity factor range, DK, was from approximately 25 MPaÖm to 60 MPaÖm.

    5.4 METALLOGRAPHY AND HARDNESS TESTING

    Sections from the weld in the as-welded and repaired states were polished and etched to reveal the

    welds, macro photographs taken and hardness testing carried out. In addition to examination of

    transverse sections, the edges of the samples (i.e. the surface of the test plate) were also prepared by

    polishing down to the level of the plate surface. Surface hardness measurements were taken to

    compare with the sub-surface values obtained from the transverse sections. The Vickers Hardness

    surveys (Hv 10kg load) of the parent materials, welds and HAZs were carried out according to BS EN

    288-3.

    The original welds had typical, well-defined runs, with HAZs in the order of 2-3mm wide. The area

    of weld repair had a less well-defined weld run structure due to the larger number of smaller beads.

    The Vickers Hardness survey according to BS EN 288-3 showed no significant hard spots in any of

    the samples for the transverse sections. The hardness values in the unaffected parent material were in

    the region of approximately Hv140 to Hv180. The highest hardness values were recorded in the

    HAZ, as expected. The HAZ on the repair weld was slightly harder than the original weld, with

    maximum recorded values of Hv331, and Hv268 respectively. These levels are below the maximum

    permitted hardness value of Hv350 stated in BS EN 288-3 for this class of material.

    The results for the surface measurements show a similar but less pronounced variation in hardness to

    that recorded for the transverse sections. The maximum HAZ hardness values recorded were Hv258

    and Hv284 for the as-welded and repair-welded conditions respectively. This gives some confidence

    that increased hardness could be indicated by measurements on the accessible surface of a structure,

    but suggests that small, isolated areas of peak hardness may not be detected since they may occur sub-

    surface.

    5.5 MICROSTRUCTURAL EXAMINATION

    The more detailed microstructural examination was carried out by the Sheffield University Metals

    Advisory Centre (SUMAC). The details of this are given in Appendix 2. The SUMAC work

    consisted of examinations on both as-welded and repair-welded samples in terms of microstructural

    observations, standard hardness tests, microhardness surveys and microanalysis using dispersive x-

    rays.

    It was shown that the HAZ microstructure followed the typical pattern of a multi-pass weld with a

    zone of grain growth at the fusion line, backed by a band of recrystallization followed by a

    spheroidized/tempered zone before the unaffected matrix. Each weld pass imposed a further HAZ on

    the underlying weld (and it’s HAZ) leading to a refined microstructure at the overlap. The grain

    10

  • growth and recrystallization zones had a microstructure of grain boundary and Widmanstatten ferrite

    (the amount depending on the local austentising temperature and subsequent cooling rate) in a

    transformed matrix. In carbon and low alloy steels of this type, the matrix can be a mixture of the

    phases ferrite, pearlite bainite and martensite. The root run area was completely refined and tempered

    and contained no “hard spots”. The macro and micro-hardness testing indicated that the HAZ of the

    ‘toe’ welds in weld 2 (the smallest weld on the side containing the repair weld) of both the as-welded

    and repair-welded samples had higher hardness values than elsewhere. The microstructure, whilst not

    exhibiting defined ‘pools’ of hard phase, did show structural refinement and reductions in pro-

    eutectoid ferrite that could explain the increased hardness.

    The study concluded that both the as-welded and repair-welded samples passed the hardness

    requirement and some potentially high hardness values obtained by microhardness should not detract

    from this, particularly as they were in areas where this might be expected and were not found

    elsewhere in the weld.

    11

  • 6. TASK 5 – RESIDUAL STRESS MEASUREMENTS

    The destructive technique of block removal, splitting and layering was used to determine the through

    thickness residual stress distribution in the as-welded and repair-welded specimens. Further details of

    the procedure and the measured results are contained in Reference 1.

    The residual stress results for the as-welded condition are shown in Figures 20 and 21 for the

    Y direction (perpendicular to the weld) and X direction (parallel to the weld) respectively. The

    stresses in the Y direction are self-balancing through the thickness, with tensile values near the

    surfaces and compressive values in the central area. The stresses in the X direction are tensile

    throughout the thickness. The distributions are asymmetric as expected, considering the asymmetric

    weld preparation, with minimum values occurring at a depth of approximately 25 mm from weld side

    1, which corresponds to the location of the weld root. Stress maximum values occur at depths of

    approximately 5 mm and 35 mm. The peak tensile stress in the Y-direction (perpendicular to the

    weld) is ~220-350 MPa, and in the X-direction (parallel to the weld) ~500-580 MPa.

    The residual stress results for the repair-welded condition are shown in Figures 22 and 23 for the

    Y direction (perpendicular to the weld) and X direction (parallel to the weld) respectively. The form

    of the stress distributions is basically the same as for the as-welded condition (Figures 20 and 21).

    The stress minimum values are of similar magnitude to the as-welded but occur closer to the centre of

    the plate, corresponding to the location of the repair weld root. Also, the stress maximum values at

    depth of 5 mm show a noticeable increase over the as-welded for both the Y and X directions, whilst

    the maximum values at depth of 35mm remain at similar levels. The increase in peak tensile residual

    stress therefore occurs on the side remote from the weld repair, rather than on the repaired side. The

    peak tensile values at depth of 35 mm are 600 MPa and 750 MPa for Y and X directions respectively,

    the latter being in excess of the weld metal yield stress measured in the tensile tests. The reason for

    this high peak is not clear but the two sets of strain measurements taken in the X direction gave very

    similar results, which suggests that it is not due to an experimental error or test equipment fault.

    As a further check on the residual stress levels at the surfaces, measurements were made using the

    shallow hole drilling technique. This technique involves using a trepanning air-abrasive jet drilling

    technique, which has been shown to introduce practically no residual stresses into the component

    under test. The technique involves the drilling of a small blind hole (typically 1.8 mm diameter x 1.8

    mm deep) in the centre of a special three-element strain gauge rosette. Local strain relaxation is

    related to the initial stress state in the specimen and calibration using a known (usually uniform) stress

    field allows residual stresses to be calculated.

    The surface stresses evaluated from the shallow hole drilling technique are as follows:

    (1/3 weld side) (2/3 weld side)

    As-Welded: Perpendicular Stress (MPa) -94 365 368

    As-Welded: Parallel Stress (MPa) 225 138 181

    Repair-Welded: Perpendicular Stress (MPa) 34 280 386

    Repair-Welded: Parallel Stress (MPa) 181 -27 -162

    These values have been included in the residual stress distribution plots of Figures 20 to 23. It can be

    seen that the surface stresses obtained from the hole drilling method are generally consistent with the

    near-surface stress distributions evaluated from the block removal, splitting and layering technique.

    12

  • 7. TASK 6 – TESTS INVOLVING PHOTOELASTIC COATING AND

    THERMAL EMISSION METHODS

    Detailed information on this Task is contained in Reference 2. The Task focused on (i) quantifying

    the fatigue crack propagation rate in welded and repair welded steel plate, (ii) investigating the use of

    a full field photoelasticity technique to measure residual stresses in the plates, and (iii) investigating

    the use of a thermoelasticity technique to measure the true crack tip driving force (i.e. stress intensity

    factor) in the two types of weld.

    The specimens used for testing were obtained from the initial test plate as described in Section 4. The

    specimens tested were identical for both original and repair welds. The geometry used for the study of

    fatigue crack growth was a tension specimen, 41.5 mm wide (W) and 12 mm thick (t) with a 4 mm

    initial edge notch (a) spark machined in the side of the original or repair weld, as appropriate

    (Figure 24).

    Stress intensity factors were calculated using the following equation:

    K ID YDs= ap (1)

    where 2 3 4

    Y 231 0 12 1 ç æ

    -= .. a

    55 10 ç æ

    ÷ + ö

    . a ÷ ö

    72 21 ç æ

    - . a ÷ ö

    39 30 ç æ

    + . a ÷ ö

    (2) è W ø è W ø è W ø è W ø

    Such values are referred to as ‘DKI Theory’ so as to distinguish them from values determined by

    thermoelastic measurement.

    The tests carried out consisted of analysing the crack growth for a tensile edge cracked specimen

    using thermoelastic stress analysis. The machine used for this purpose was an ESH 100kN servo-

    hydraulic machine which allows the application of a cyclic load to the specimen at the frequency and

    load convenient for the thermoelastic test.

    Seven fatigue tests were carried out for different load conditions, as detailed below:

    Identifier Load range,

    kN

    R ratio Comments

    AEA1 32.4 0.13 Original weld

    AEA2 40 0.1 Original weld, 30kN range at R=0.1 applied for 800000

    cycles with no growth

    AEA3 33.0 0.13 Repair weld

    AEA4 39.6 0.1 Repair weld. Subsequently used for J test

    AEA_F2 36.9 0.28 Repair weld. Test run to fracture of specimen.

    AEA_F3 37.6 0.58 Original weld

    AEA_F1 37.6 0.58 Repair weld

    Images at different number of cycles during the tests were taken. At the same time, for every picture

    captured, the number of cycles and the crack length were noted. A vernier microscope was used to

    measure the rate at which the crack length had grown between different images.

    13

  • A non-standard J test was carried out by loading in four-point bending one of the edge cracked tensile

    specimen used for crack growth analysis. The results obtained were found to be similar to those

    previously obtained that are presented in Figure 17. In addition, one of the fatigue tests was run until

    failure. The loads at fracture were Pmin = 16.4 kN, Pmax = 51.6 kN, the final crack length being

    34.93 mm, including the initial 4 mm slit. Failure occurred after 536770 cycles.

    The stress intensity factor ranges were plotted against the crack length for different R values and for

    different specimens (original and repair welds). An example is shown in Figure 25 for the R = 0.13

    case of the repaired weld specimen, AEA3. “RAT” and “FGD” referred to in Figure 25 are the initials

    of the two different operators who processed the results. In all cases, experimental results were

    compared to the range of stress intensity factor calculated by Equations 1 and 2. It was observed that

    all experimental data lay below the theoretical values when the crack length is long enough. This is

    thought to be due primarily to the crack closure effect (see below), but other factors may also be

    influential. In particular, the large displacement of the crack at high stress intensity factors may well

    mean that the published stress intensity factor calibrations are erroneous at these levels.

    At the same time using information from the tests, the crack growth rate against the stress intensity

    factor was plotted for the different experiments. Figure 26 is an example of such a plot whereby the

    Paris law is presented using experimental values for the stress intensity factor (identified as

    ‘Deltatherm data’ in the Figure) and values predicted from Equations 1 and 2 and the crack growth

    rates obtained from experimental measurement. The “AEAT growth equation” curve included in

    Figure 26 has been derived from the data presented in Figure 19.

    Finally, an estimate of the closure level was made from the difference between the theoretical DKI and

    the value measured using Deltatherm. The values are shown in Figure 27 plotted against the crack

    length.

    In considering crack closure effects, it has previously been observed that non-linear crack opening

    behaviour results in a region of residual tensile deformation in the “wake” of a fatigue crack. The

    resulting permanent contact between the two crack faces results in a lowering of the crack opening

    displacement and consequently lower driving force for fatigue crack advancement.

    A large amount of research has been carried out on this topic during the last few years and the

    mechanisms involved have been described. These mechanisms suggest that several types of closure

    affect the rate of fatigue crack advance. The possible sources of crack closure are the following:

    - Plasticity induced crack closure, due to residual stress in the wake of the crack. - Oxide induced crack closure, due to the oxide layers formed inside the fatigue crack. - Roughness induced crack closure, due to the roughness of the fatigue fracture surface. - Viscous induced crack closure, due to the penetration of viscous fluids inside of the crack. - Transformation induced crack closure, due to phase deformations at the crack tip caused by stress

    or strain.

    In addition, the presence of non-uniform residual stresses in a structure will contribute to the crack tip

    driving force in addition to primary loads. These complex stresses may increase the stress intensity

    factor above that estimated from the external loading, or may decrease it, thereby having a similar

    effect on crack closure.

    Looking ahead to Figures 37 and 38, which present the finite element determined values of stress

    intensity factor for the residual stress fields (refer to section 9.1), it is evident that the KI values are

    positive for all crack sizes considered (crack depth, a, ranging from just over 2 mm to 20 mm). Crack

    14

  • opening, as opposed to crack closure, would therefore be expected to occur from the residual stress

    distribution.

    Some tests were undertaken using reflection photoelasticity with the intention of measuring residual

    stress in welds. Two different specimens were used from the original and the repair welds. The

    photoelastic results confirmed the previous measurements referred to in Section 6, whereby very little

    difference was observed between the residual stresses in the as-welded and repaired weld conditions.

    15

  • 8. TASK 7 – DEVELOPMENT OF FINITE ELEMENT MODELS

    Detailed finite element modelling of a matrix of relevant un-repaired and repaired weld configurations

    has formed a major part of the project. The work was mainly focussed on the modelling of a plate

    geometry but a spherical vessel geometry was also considered. This work (covering Tasks 7 and 8) is

    fully described in Reference 3 and summarised in the following sub-sections.

    8.1 WELD MODELLING TECHNIQUE

    In terms of the development of the finite element models, a weld bead lumping approach was used to

    model weldments in which a small number of lumped beads was modelled in both original and repair

    welds. A non-linear analysis of the welding process was carried out using a simplified ABAQUS

    finite element model of the parent plate and weld. In this analysis, the original weld was built up by

    the addition of each lumped weld bead in an incremental manner.

    A thermal transient analysis was first conducted in order to establish the temperature history of each

    point in the plate or sphere due to the addition of each weld bead. A subsequent elastic-plastic

    analysis used an almost identical finite element model to simulate the addition of the weld beads.

    This mechanical model was loaded by imposing, at each time increment, the temperature of each node

    from the above thermal transient analysis. Like the thermal analysis, the mechanical model was

    necessarily simplified, so the complex behaviour of the weld and parent metal near melting point was

    not considered. However, approximate temperature dependent mechanical properties were used.

    Low values of yield stress and perfectly plastic properties were used at temperatures near the melting

    point to reduce the loading on adjacent material. However, this did incur the penalty of producing

    unrealistically large plastic strains that cannot be annealed.

    After adding the final lumped bead of the original weld, the current state of the mechanical model

    (displacements, stresses, elastic and plastic strains etc.) was saved for subsequent restarts. Following

    this, the elements in the repaired areas were removed, and the lumped beads of the repair were added.

    The required state of the model was again saved for subsequent restarts.

    Figure 28 shows a part of the finite element mesh used to model a though-thickness section of the

    welded test plate in the region of the weld. For convenience, the mesh is shown rotated by 90o with

    respect to Figures 2 and 3. The plate thickness was 40 mm measured in the horizontal direction in

    Figure 28. The depth of the repair weld was 15 mm, this being slightly smaller than the 17 mm

    actually excavated in the real plate weld. The original weld comprised nine lumped beads, and the

    repair weld had four. The weld caps were not modelled. The plate width was measured in the vertical

    direction in Figure 28. Due to symmetry about the centre of the weld, only one half of the 1000 mm

    plate width was modelled.

    To make allowance for later generalisation, the finite element mesh was actually three-dimensional,

    but only a single element thickness was used in the plate height direction perpendicular to the plane of

    Figure 28. The strong back plates used during the actual welding were modelled as beam elements,

    with equivalent section modulus, running vertically along the appropriate side of the mesh in Figure

    28.

    A sphere was modelled with 40 mm thickness and 20 m diameter. The weld was considered to be a

    fully equatorial one, with dimensions and bead lumping exactly as modelled in the plate weld. The

    repair lay on the outside of the sphere. Figure 29 shows the axisymmetric finite element mesh used.

    Again due to symmetry about the centre of the weld, only one half of the sphere was modelled. No

    16

  • strong back restraints were used in the weld mechanical simulation. A further weld repair situation

    was also considered for the sphere. This involved a shallow excavation, with removal of only that

    material corresponding to repair weld beads 3 and 4 (see lower part of Figure 28) from the original

    weld and its replacement by simulated welding. The repair depth in this case was 6.6 mm instead of

    the standard 15 mm.

    After the modelling of the original weld in both geometries, a creep analysis was carried out to

    simulate a PWHT of four hours at 650oC. Here, the temperature of the model was raised from 20

    oC to

    650oC in five hours at a uniform rate. The temperature was held at 650

    oC for four hours and then

    reduced back to 20oC in six hours at a uniform rate. After the PWHT simulation, and also after the

    subsequent repair, the state of mechanical model was saved for subsequent restarts. These are, in fact,

    the welded plate and sphere states into which defects were inserted as discussed later.

    8.2 MATERIAL PROPERTIES

    The tensile properties of the parent plate and of the original and repair welds at ambient temperature

    are reported in Reference 1. Figure 30 shows the measured results for the parent steel, and in the

    centre of the original weld in terms of true stress versus true strain.

    The measured results were modified in order to be suitable for the weld simulation analysis. Firstly,

    the results were smoothed. Secondly, the Luders strain seen in the parent steel was removed since it

    was considered to be a consequence of the tensile test specimen geometry, rather than a real material

    phenomenon. The rate of strain hardening seen at the end of the tests, at about 3.5% strain, was

    extrapolated to larger strains. Thirdly, trial simulations of the original weld showed that an average

    equivalent plastic strain of about 12% remains in the weld. This strain could not be annealed within

    the analysis. To take account of this prior welding strain, the weld values in Figure 30 had 12% added

    to the plastic strain so that after the weld simulation, the stress versus additional strain followed the

    smoothed, measured curve. In Figure 31, the solid blue curve shows the tensile properties at 20oC

    used for parent plate in the weld simulation analyses. The solid red curve shows the properties of both

    the original and repair weld at this temperature.

    The resulting 0.2% and 1% proof stresses of the weld and parent materials are as follows:

    0.2% Proof Stress (MPa) 1% Proof Stress (MPa)

    Temperature (oC) Parent Weld Parent Weld

    20 371 362 418 362

    2000 5 5 5 5

    Values at intermediate temperatures were obtained by interpolation. Despite the apparent weld under-

    match (lower yield strength than parent) the plastic strain accumulated during the weld simulation

    analysis raised the yield stress; resulting in behaviour like that measured in Figure 30.

    The following Paris Law curve constants were used in the fatigue crack growth that were carried out:

    da = 10x 2 .1 -8 DK 77.2 (3)

    dN

    where DK is the stress intensity factor range during cyclic loading, measured in MPaÖm, and da/dN is

    in mm/cycle. It may be noted that Equation 3 represents the lower curve presented in Figure 19.

    As can be seen from Figures 17 and 18, fracture toughness at ambient temperature corresponding to

    0.2 mm of tearing, J0.2, was measured to be equal to 105 and 102 KJ/m2, respectively for original and

    repair weld material. These values translate to a toughness of about 150 MPaÖm in terms of KJ0.2. In

    17

  • this study, values of toughness were considered that ranged from 160 MPaÖm down to significantly

    lower levels of about 30 MPaÖm in the as-repaired condition. As discussed later, such low values of

    fracture toughness can result in repair welds due to a variety of circumstances.

    The properties used for the weld simulation thermal analysis and the creep properties used in the

    simulation of the intermediate post-weld heat treatment are described in Reference 3.

    8.3 RESULTS OF WELDING SIMULATIONS

    Figures 32(a-b) compare the through-thickness stress distributions at the middle of the weld with the

    measured results presented in Figures 20 to 23. It should be noted that, in these and subsequent

    similar Figures, the through-thickness distance is always measured from the non-repaired side 1. The

    experimental results are shown as solid lines and the predicted results are dashed lines. The stresses

    produced by the original weld are shown in blue, those caused by the repair weld are in red. Predicted

    results are in general agreement with the measurements, with tension near the plate surface and

    compression at mid-thickness. However, the numerical simulation was unable to predict the precise

    magnitudes and positions of stress peaks and troughs. This is not surprising, given the simplifications

    and approximations involved. It should also be noted that the predictions and measurements agree that

    a higher transverse stress occurs in the repaired weld, but on the un-repaired side 1. Both

    measurements and predictions show a similar magnitude of peak transverse stress on the repaired side

    2.

    In Figures 33(a-b), comparisons are made for through-thickness distributions of transverse and

    longitudinal stress across the middle of the weld between the four different numerical simulations.

    These cases are:

    (i) as originally welded (blue diamonds),

    (ii) as originally welded followed by post-weld heat treatment (green diamonds),

    (iii) as originally welded, followed by partial weld removal and repair welding (red circles),

    (iv) as originally welded, followed by post-weld heat treatment, partial weld removal and finally repair welding (orange circles),

    In case (ii), the effect of heat treating the original weld is apparent, with a large reduction of both

    components of stress compared with the as-welded case (i). In Figure 33(a) it is seen that the through-

    thickness transverse stresses in the weld for the two repair cases (iii) and (iv) are similar. The repair

    of the PWHT weld thus re-establishes a pattern of stress as if the original PHWT had not been carried

    out. Furthermore, close to the surface of the un-repaired side 1, the repair causes an increase in

    transverse stress to a higher peak level than the un-heat treated original weld (compare the orange

    with blue curves). Figure 33(b) shows that the longitudinal stress is affected by repair mainly on the

    repaired side 2 itself.

    Figures 34(a-b) compare the predicted residual stress results for the four simulation cases carried out

    on the sphere. The general pattern of results is similar to that of the plate in Figures 33(a-b).

    Figure 35(a) compares transverse stresses for cases (ii) and (iv) between the plate (open symbols) and

    sphere (filled symbols). For case (ii), the original PWHT weld shown in green, the peak transverse

    tensile stresses predicted in the sphere are about half those in the plate on the last welded side 2. This

    situation is reversed on the first welded side 1. The sphere therefore appears to have a component of

    through-wall bending stress. For case (iv), repaired stresses shown in orange, the sphere has higher

    18

  • values than the plate at the un-repaired side. Figure 35(b), showing longitudinal stresses, also

    illustrates slightly lower predictions in the sphere than the plate in respect of the PWHT original weld,

    case (ii).

    Comparisons of residual stresses for ‘deep’ and ‘shallow’ repairs in the sphere are shown in Figures

    36(a-b). These graphs show results for the original PWHT weld, the standard simulated repair of

    depth 15 mm and also for the shallower repair with a depth of 6.6 mm. On the repaired side of the

    weld, the shallow repair promotes peak values of transverse and longitudinal stress similar to the

    deeper repair. On the un-repaired side of the weld, the shallow repair gives peak stresses lying

    between the un-repaired PWHT cases and the deep repair case. Thus, shallow weld repairs can

    promote high local residual stresses if the component is not heat treated.

    8.4 ANALYSES OF DEFECTS IN THE SIMULATED WELDS

    Crack-like defects were inserted into the plate and sphere weld cases (ii) and (iv) of the previous

    section. Additional loads were applied to the models to give stresses on the defective section

    typically experienced by engineering structures, and crack driving forces (CDFs) were calculated.

    These parameters were then used to determine limiting or critical defect sizes for various values of

    weld fracture toughness in the two welded states. Comparisons were made between limiting defect

    sizes for defects in these heat treated and as-repaired situations.

    Using the CDFs, fatigue crack growth calculations were also carried out to determine the number of

    loading cycles required to reach the limiting condition for a range of initial defect sizes. Comparisons

    were made between fatigue lives of defects in the heat treated and as-repaired states for a range of

    initial defect sizes and fracture toughness.

    Some modelling simplifications were made in these analyses of defects in welds and these are

    explained in Reference 3.

    Defects were inserted into the plate model on the plane through the middle of the weld. The two

    configurations considered in the welded plate are actually those shown in Figures 4 and 5. In Figure 4,

    a surface defect of depth a is shown in the weld. In some cases, the tip reaches into the original weld

    (for the weld repair cases). In Figure 5, an embedded or internal defect is considered in the weld. In

    some of the weld repair cases, this also reached into the original weld. As for the surface defect, this

    defect was also considered as fully extended along the whole length of the weld. The defect is

    characterised by its depth, 2a, and the distance of its nearest tip from the repaired surface, p.

    In the welded sphere, surface defects were considered in the middle of the repair weld, like Figure 4.

    Since the repair was considered to lie on the outside of the sphere (Figure 29) and the finite element

    model was axisymmetric, this corresponds to a fully extended outer surface defect of depth a along an

    equatorial weld.

    Modelling of the defects was accomplished by removing the symmetry boundary conditions along the

    line of the defect. These restraints were replaced by equivalent forces that were reduced to zero in

    several subsequent elastic-plastic increments of the analysis. The created defect usually opened

    under the influence of the residual stress field. In some circumstances however, the defect closed over

    at least part of its depth due to a predominantly compressive residual stress. In such cases, the contact

    of the opposing faces of the defect was not modelled, so the defect was allowed to ‘over-close’.

    Simultaneous introduction of the entire crack surface is mechanistically different to the modelling of

    slow, sub-critical crack growth where the crack is introduced progressively. In the former, a zone of

    plastic deformation appears at the crack tip(s) only. In the latter, a wake of plastic deformation

    develops on the crack flanks as (each) crack tip moves forward.

    19

  • In terms of the primary loading, a remote, uniformly distributed, tensile load was applied to the top of

    the modelled plate, 500 mm away from the defect plane. This represents loading in the weld

    transverse direction normal to the plane of the defect; causing it to open further, or to open if closed in

    the residual stress field acting alone. Various magnitudes of remote membrane load were applied,

    with a maximum of 225 MPa. This load was considered to be the occasional ‘overload’ condition for

    which the possibility of ductile crack initiation or cleavage fracture was assessed. A remote load of

    180 MPa was considered to be the cyclic ‘operating’ load that causes fatigue crack growth. This value

    of nominal stress is about 50% of the 0.2% proof stress and 34% of the UTS of the parent plate, and

    so is typical of an engineering structure.

    An internal pressure was applied to the sphere. This results in an equi-biaxial stress in the spherical

    shell that acts to open the defect. Various magnitudes of pressure were applied, with a maximum of

    1.8 MPa corresponding to a meridional stress of 225 MPa according to thin shell theory. Again, this

    was considered as the overload condition. The operating condition was a repeatedly applied pressure

    of 1.44 MPa, causing a nominal stress of 180 MPa in the shell.

    Crack driving force was evaluated in terms of stress intensity factor. This parameter was evaluated

    both elastically (designated K) and from an elastic-plastic analysis (designated KJ). Because of the

    complexity of the finite element analyses, the conventional J-contour integral option with ABAQUS

    could not be accurately employed to evaluate K and KJ. The primary reason for this is that the

    contour integral calculation of J requires that significant unloading of the material does not take place.

    This was not the case in the present finite element analyses that simulated welding, heat treatment,

    material removal and repair welding. An alternative calibration approach, based on the crack opening

    displacements at the node immediately behind the crack tip, was therefore used as a proxy for J. Full

    details of this calibration procedure are contained in Reference 3. It may be noted that J was 0.5

    converted to K by the usual equation, K = [(EJ)/(1-n 2)] where E is Young’s modulus (taken as 200 GPa and n is Poisson’s ratio (taken as 0.3).

    20

  • 9. TASK 8 – APPLICATION OF FINITE ELEMENT MODELS TO

    MATRIX CASES

    This Task is associated with applying the finite element models and methodology referred to in

    Section 8 above to a matrix of cases. It should be noted that because of previously unforeseen

    complexities of the analyses (e.g. the requirement to develop the calibration method used to evaluate

    crack driving force), it was not possible to include all the cases that had originally been suggested

    under Task 2 (Section 3). A good selection of the cases was included in the analyses however.

    9.1 EDGE DEFECTS IN THE WELDED PLATE

    Figures 37(a-b) show results for elastic stress intensity factor K for various defect depths and levels of

    primary load in the welded plate in the un-repaired heat treated and the as-repaired states,

    respectively. The magnitude of primary load is indicated in the legends. 0 MPa corresponds to

    residual stress only. The stress intensity factors for the defect in the repaired weld are obviously larger

    than in the un-repaired PWHT case. The two curves for residual stress only show a tendency to rise

    with increasing defect depth and then gradually fall, reaching a maximum K for about 11 mm defect

    depth. This is a consequence of the residual stress fields presented in Figure 33(a) whereby the

    stresses are shown to start decreasing in magnitude after reaching tensile peak values at a distance of

    about 10 mm from the appropriate side of the plate. The other curves simply show that the additional

    stress intensity factor is proportional to the primary load applied.

    Figures 38(a-b) show results for KJ calculated from J obtained from elastic-plastic analyses. The

    curves for zero primary load are unchanged from Figures 37(a-b). With increasing crack depth and

    load, the value of KJ becomes larger than the corresponding value of K in Figure 37 due to plasticity

    effects. The KJ results in the as-repaired state are higher than in the PWHT state, particularly for

    intermediate defect depths and loads. For deeper defects and higher loads, the residual thermal strains

    arising from welding are reduced by the mechanical plastic strains, and so the difference in CDFs

    between the two welded states is reduced.

    Repeated loading and unloading between zero and 180 MPa was considered. Fatigue crack growth

    predictions are made using the Paris law, Equation 3, but with the more representative parameter

    DKJ=KJmax-KJmin used in preference to DK. Here KJmin is the crack driving force for the appropriate

    residual stress acting alone and KJmax is the total CDF for combined residual stress plus 180 MPa

    applied stress. Both these parameters are available in Figures 38(a-b). For each updated crack depth

    the value of KJ for an occasional 225 MPa applied stress was also available. This KJ was required to

    assess when the critical crack size had been reached during the fatigue crack growth calculations (i.e.

    fatigue crack growth was based on an applied stress range of 180 MPa and critical crack size was

    based on an overload stress of 225 MPa).

    Results of fatigue crack growth predictions are illustrated in Figures 39(a-b). These graphs show

    crack depth, a, as a function of the number of loading cycles, N, between zero and 180 MPa for the

    different initial defect depths indicated in the legends. Defects in the as-repaired weld, Figure 39(b),

    need fewer cycles to grow to a given depth compared with the PWHT state, Figure 39(a), since the

    value of DKJ is generally lower for the PWHT state (Figure 38).

    Ductile crack initiation, or cleavage failure in the ductile-to-brittle transition region of ferritic steels,

    is considered to occur when KJ is equal to a given fracture toughness KJc. No differentiation is drawn

    between these types of failure and the term ‘limiting condition’ is used hereafter. In Figures 40(a-b)

    21

  • results are presented for fracture toughness, KJc, versus the number of 0-180 MPa loading cycles, Nf,

    required to cause the limiting condition due to an occasional 225 MPa overload. Curves are shown for

    different initial defect depths. A comparison of the two graphs shows that for a given fracture

    toughness and initial defect, fewer cycles are required to grow to the limiting condition in the as-

    repaired weld.

    Figure 41 shows the relationship between critical defect depth, ac, at the limiting condition and

    fracture toughness in the two weld states. For a given toughness, the critical defect depth is smaller in

    the as-repaired weld. The difference in critical defect depth between the two welds depends on

    toughness. For example, for a weld toughness of 160 MPaÖm, the critical defect depth is about 17.5

    mm in the PWHT weld and 16.7 mm in the as-repaired case. This difference in depth is not

    significant. However, for a lower fracture toughness of 100 MPaÖm, the respective critical defect

    sizes are about 13.5 mm and 9.5 mm. This difference is more significant.

    Figure 42 shows curves of the ratio of the number of loading cycles to the limiting condition for a

    defect in the repair, Nf(repaired), to the number of cycles in the un-repaired PWHT state, Nf(un-

    repaired). These curves assume the same initial defect depth in both weld states. Each curve

    represents a different fracture toughness that is also assumed to be the same in both welds. So in this

    graph, a comparison is made of the fatigue life of the same size defect and same fracture toughness in

    the repaired and un-repaired welds. Values less than unity imply a worse life for the repair. Of course,

    in the majority of cases, this is the case due to the higher repair residual stresses. Some results are

    greater than unity for initial defects between 8 mm and 14 mm deep for high toughness. This occurs

    because of high values of KJ at zero load in the as-repaired state, Figure 38(b), giving lower values of

    DKJ in the as-repaired weld compared with un-repaired and so reduced fatigue crack growth rates.

    The series of graphs in Figures 43(a-e) also illustrate the ratio of operating cycles required to reach

    the limiting condition for repaired and un-repaired cases. These take account of different initial defect

    depths and fracture toughness in the two weld states. The scenarios are either: an edge defect is left in

    the (un-repaired) weld, or a repair is carried out that leaves the same size or shallower edge defect

    located in material with the same or reduced local fracture toughness. The trade-off is thus explored

    between introducing the same or shallower defect in the repair and higher levels of residual stress and

    lower fracture toughness in that weld.

    Firstly, Figure 43(a) shows comparisons between leaving un-repaired a 5 mm deep edge defect and

    inadvertently introducing either 5 mm, 4.2 mm or 3.3 mm deep defects in the as-repaired weld. Curves

    are shown of the ratio of operating cycles to reach the limiting condition in the repaired and un-

    repaired weld versus the percentage reduction in repaired fracture toughness from the original PHWT

    value. Each curve represents a combination of repair defect depth and original toughness. The highest

    values of PWHT fracture toughness are represented by blue curves, and the lowest by red. For

    example, the blue squares show the effect of leaving in the repair the same size 5 mm deep defect for

    an original PWHT fracture toughness of 160 MPaÖm; slightly greater than the initiation toughness of

    the plate test welds. The operating life of the repair is always lower than the un-repaired life (ratio of

    cycles to the limiting condition is less than unity). Repair life gets comparatively worse as the

    repaired toughness reduces. So, a 40% reduction of the repaired toughness, compared with the

    original PWHT value, leads to a halving of the repaired life compared with the life if left un-repaired.

    The open blue diamonds show the effect of introducing into the repair a 4.2 mm defect compared with

    leaving un-repaired the PWHT weld containing a 5 mm defect. The repaired life slightly exceeds the

    un-repaired life by only a small margin, though if the repaired toughness drops more than 20% below

    the original 160 MPaÖm, the life of the repair becomes less than the un-repaired life. The blue

    triangles show the comparison between having a 3.3 mm defect in the repair and leaving un-repaired

    22

  • the 5 mm defect. The repaired exceeds the un-repaired life until the repaired toughness drops below

    about 43% of the PWHT level.

    There are more interesting consequences for lower PWHT fracture toughness. Consider a toughness

    of 100 MPaÖm in the PWHT state; the three sets of orange curves and symbols in Figure 43(a). A

    defect in the repair having a depth of either 5 mm or 4.2 mm always has a shorter operating life than

    the 5 mm deep defect in the PWHT weld. A 3.3 mm deep repair defect, shown by orange triangles,

    gives a slightly longer life than the un-repaired 5 mm case for no reduction of toughness. However, a

    mere 10% or so reduction of toughness due to the repair results in a shorter operating life. For the

    lowest 80 MPaÖm PWHT toughness (red curves and symbols), all repaired defects from 3.3 mm to 5

    mm depth imply an inferior fatigue life even if the repaired toughness does not change. These results

    therefore demonstrate that repairing a shallow surface defect by re-welding is likely to result in a

    shorter operating life if it leaves a defect and reduces the fracture toughness. This is particularly

    apparent for materials with low original toughness. Although the repair surface defects considered

    here could be detected visually or by Magnetic Particle Inspection, it is considered that a defect about

    3 mm deep cannot be sized accurately by Ultrasonic Techniques.

    Figure 43(b) shows similar sets of predictions for a 6.7 mm deep original defect. Here a defect of

    depth 6.7 mm, 5 mm or 3.3 mm is considered left in the repair. The trend of the predictions is similar

    to the 5 mm case discussed above, but a larger reduction of toughness is needed to obtain a shorter life

    in the repaired situation. For example, the orange triangles show that for 100 MPaÖm toughness in the

    PWHT weld, a 45% reduction due to repair is required to give a shorter life for a 3.3 mm deep repair

    defect.

    Figures 43(c-e) however provide more support for repairing deeper surface defects. Figure 43(c)

    compares an un-repaired 9.2 mm defect with repaired defects of 6.7 mm, 5 mm or 3.3 mm. Note that

    not all symbols in the legend are seen on the graph because some initial defect/toughness

    combinations considered meet the limiting condition and so imply zero operating life (see Figure 41)

    or the repair life exceeds twice the un-repaired. The steeper angle of the curves suggests that, for these

    deeper initial defects, the effect of toughness reductions due to repair can be more severe. For

    example, the red diamonds compare the un-repaired 9.2 mm defect with 5 mm in the repair for a low

    PWHT toughness of 80 MPaÖm. If the toughness reduces by up to 10%, the life of the repair is still

    over twice the life if un-repaired. However, a toughness reduction of 30% due to repair causes the

    repaired life to drop drastically to about one quarter of that if the weld was left un-repaired.

    Figure 43(d) shows a comparison of the 10.8 mm deep un-repaired defect with 9.2 mm, 6.7 mm or 5

    mm defects in the repair. Since it is unlikely that a 9.2 mm defect is left in a repaired weld, the

    shallower depths are perhaps more feasible. Considering 160 MPaÖm PWHT toughness and a 5 mm

    defect in the repair (blue triangles) then a 60% toughness reduction due to repair (down to about 64

    MPaÖm) is required to obtain a shorter fatigue life in the repair. If the PWHT weld has a lower 80

    MPaÖm toughness (red triangles) then only a 30% reduction down to about 56 MPaÖm will give a

    worse or even no repair life.

    Finally, Figure 43(e) compares the un-repaired 13.3 mm deep defect with 9.2 mm 6.7 mm or 5 mm in

    the repair. Many of the ratios are zero or unreported because there is no un-repaired or repaired life:

    the initial defect is at or beyond the limiting condition. Obviously, this original 13.3 mm defect is

    more likely to warrant repair than the shallow ones discussed above. However, onerous welding

    conditions, giving the likelihood of poor toughness and a remaining defect, can result is a worse life.

    Leaving a mere 5 mm deep defect, and reducing an original toughness of 100 MPaÖm (orange dotted

    curve) by 45%, will lead to little or no life of the repair. It will reach the limiting condition on first

    overload. As seen in Figure 41, a 5 mm deep defect with a toughness of about 60 MPaÖm is near the

    limiting condition in the repair weld.

    23

  • 9.2 EQUATORIAL DEFECTS IN THE WELDED SPHERE

    The stress intensity factors for primary load alone are in good agreement with standard results for

    extended edge defects in spheres. Figures 44(a-b) give the results for the parameter KJ from the

    welded sphere simulations. It is apparent that the crack driving forces are lower in this geometry than

    previously seen for the plate. Due to algebraically more compressive residual stress at the defective

    side 2 of the PWHT sphere weld (Figure 35a) the crack driving forces are negative for the shallowest

    and also for the deepest defects in the absence of primary load, as seen in Figure 44(a).

    Repeated loading and unloading between nominal biaxial stresses of zero and 180 MPa was again

    considered, with fatigue crack growth predictions made using Equation 3 and DKJ=KJmax-KJmin. Only

    the range over which KJmin is positive contributes to fatigue since the crack is actually closed if KJmin is

    negative according to Figure 44(b). The value of KJ was calculated for an assumed overloading to a

    nominal stress of 225 MPa at each crack depth and associated number of cycles. Fatigue crack growth

    predictions are shown in Figures 45(a-b). As seen in the welded plate case earlier, defects in the as-

    repaired state need fewer cycles to grow to a given depth compared with the PWHT state.

    Figures 46(a-b) show results for fracture toughness versus the number of loading cycles required to

    cause the limiting condition at the 225 MPa load. As with the welded plate, fewer cycles are required

    in the as-repaired sphere weld to grow the defect to the limiting condition. Note that the range of

    toughness is shifted to lower values compared with the welded plate because of the lower crack

    driving forces in the sphere.

    Figure 47 plots critical defect depth at the limiting condition as a function of fracture toughness in the

    two weld states. Again for a given toughness, the critical defect depth is smaller in the as-repaired

    weld. However, the difference between the two cases is more significant due to the generally lower

    levels of toughness illustrated. For example, for a weld toughness of 100 MPaÖm, the critical defect

    depth is about 19 mm in the PWHT weld and 10.5 mm in the as-repaired case. The respective critical

    depths for the welded plate (Figure 41) are about 13.5 mm and 9.5 mm. Thus the difference between

    critical depths in the welded sphere is clearly more significant than for the plate. The green curve in

    Figure 47 suggests that for PWHT toughness close to 60 MPaÖm there is a large change in critical

    crack depth. This is due to the flat or falling CDF in Figure 44(a) arising from compressive PWHT

    residual stress at distances from side 2 greater than about 10 mm, see Figure 35(a).

    Figure 48 compares the fatigue life of the same initial size defect and fracture toughness in the

    repaired and un-repaired welds. Results are always less than unity; implying a worse life for defects in

    the repair.

    Figures 49(a-e) illustrate, for edge defects in the welded sphere, the trade-off between introducing the

    same or shallower defect in the repair and higher residual stress and lower toughness there. These

    graphs are similar to Figures 43(a-e) for the welded plate discussed earlier, with the exception that the

    maximum toughness examined here is lower due to smaller crack driving forces in the sphere.

    Figure 49(a) contrasts leaving un-repaired a 5 mm deep defect with introducing either 5 mm, 4.2 mm

    or 3.3 mm deep defects in the as-repaired weld. With few exceptions, the operating life of the repair is

    always lower than the un-repaired life for PWHT fracture toughness up to 110 MPaÖm. Repairing a

    long 5 mm deep surface defect in this weld geometry by re-welding without heat treatment is not

    beneficial if it is likely that a mere 3 mm or so deep surface defect can remain undetected after repair.

    The probability that the toughness will be reduced by a non-heat treated repair reinforces this

    conclusion.

    24

  • Figure 49(b) shows comparisons between leaving un-repaired a 6.7 mm deep defect and introducing

    6.7 mm, 5 mm or 3.3 mm deep defects in the repair. For the highest PWHT toughness of 110 MPaÖm

    and the smallest repair defect of 3.3 mm (blue triangles), the defective repair has a longer life unless

    the repair causes a toughness reduction of about 35% to 72 MPaÖm. For the lowest PWHT toughness

    examined of 70 MPaÖm, a mere 12% toughness reduction will give a lower life for a 3.3 mm deep

    defect in the repair (red triangles).

    Looking ahead to Figure 49(d), compares leaving un-repaired a 10.8 mm deep defect in the PWHT

    weld with having 9.2 mm, 6.7 mm or 5 mm defects in the as-repaired state. For 110 MPaÖm PWHT

    toughness and leaving the 5 mm defect after repair (blue triangles), a lower life is achieved by the

    repair should the toughness fall by more than 35% to about 72 MPaÖm. For the lowest considered

    PWHT toughness of 70 MPaÖm, only a 12% or so reduction in toughness will give a lower fatigue

    life for the 5 mm repair defect (red triangles).

    Finally, Figure 49(e) compares a 13.3 mm un-repaired defect with 9.2 mm, 6.7 mm and 5 mm defects

    in the repair. For 110 MPaÖm PWHT toughness and a 5 mm defect in the repair (blue triangles) then

    about 43% toughness reduction, to 63 MPaÖ m. is required to obtain a shorter life in the repair.

    9.3 EMBEDDED DEFECTS IN THE WELDED PLATE

    This section explores the behaviour of embedded defects in both the un-repaired and repaired weld in

    the plate. Two initial types of defect configuration were considered. In the first, labelled ‘p+2a=16.7

    mm’, the upper defect tip closest to side 1 (see Figure 5) lies at a depth of 16.7 mm from the repaired

    side 2 of the plate. Various initial defect heights 2a were examined. In the second configuration,

    ‘p+2a=10.8 mm’, the upper defect tip is 10.8 mm from side 2. Again, various initial defect heights

    were studied. In all cases examined here, only the lower tip of the defect closest to repaired side 2

    was considered. This necessary simplification meant that fatigue crack growth was not considered at

    the upper tip closest to side 1. This is not as approximate as it might at first appear, particularly for

    p+2a=16.7 mm, since the upper tip lies far from side 1 and generally experiences lower crack driving

    forces (and ranges) than the lower tip of the defect. Given the power law dependence of the Paris law,

    Equation 3, this leads to much lower rates of fatigue crack growth than experienced by the lower tip.

    Figures 50(a-b) gives some KJ crack driving force results for increasing height of an embedded defect

    in the un-repaired and repaired weld. These relate to the case p+2a=16.7 mm. A comparison of

    Figures 50(a-b) and Figures 38(a-b) shows that, for the same defect height/depth, the CDFs for the

    embedded cases are comparable to the edge cases, particularly for higher/deeper defects. At first

    sight, this appears to be inconsistent with what is generally understood: that edge cracks have higher

    CDFs than embedded cracks of the same depth. However, the embedded defect tip is developing

    towards the repaired surface and so experiencing an increasing tensile nominal stress field. By

    contrast, the edge defect results relate to the (only) tip of the defect in the ‘deep’ position which

    develops towards a more compressive stress field at plate mid-thickness. Should the 2a=14 mm high

    embedded defect break through the 2.7 mm remaining ligament to the repaired surface, it is re-

    characterised as a 16.7 mm edge defect. In both PWHT and as-repaired welds, the CDF will increase

    at the 220 MPa maximum applied load plotted: compare Figures 38(a-b) for a=16.7 mm with Figures

    50(a-b) for 2a=14 mm.

    Figures 51(a-b) plot fracture toughness versus number of loading cycles to the limiting condition for

    the case p+2a=16.7 mm. As expected, for the same toughness fewer cycles are achieved in the as-

    repaired weld. Figures 52 and 53 compare critical crack depths for un-repaired PWHT and as-repaired

    welds for the two embedded cases p+2a=16.7 mm and p+2a=10.8 mm, respectively. A lower range

    of toughness is displayed in these cases compared with the edge defects (Figure 41) consistent with

    the generally smaller crack driving forces obtained. The rapid change of critical defect height with

    25

  • toughness in Figure 53 compared with Figure 52 is due to the defect tip in question being closer to,

    and so more sensitive to, the tensile part of the residual stress field near the plate surface.

    Figure 54 plots the ratio of cycles to limiting condition in the as-repaired weld to cycles in the un-

    repaired condition as a function of toughness and initial defect height for the case p+2a=16.7 mm.

    This shows that, for the same initial defect size and toughness in both welds, a shorter life is generally

    obtained in the repair. A longer repair life is however seen for the very short initial defects examined.

    This behaviour is due to the defect tip of interest lying in the compressive part of the repair residual

    stress field, resulting in low rates of fatigue crack growth compared to the un-repaired case. Results

    for the case p+2a=10.8 mm are seen in Figure 55. There is a more restricted range of toughness to

    show here due to the low CDFs for this shorter defect.

    Figure 56(a) contrasts leaving un-repaired a 5 mm high defect with introducing either 5 mm, 4.2 mm

    or 3.3 mm high defects in the as-repaired weld for p+2a=16.7 mm. The squares show the effect of

    having the same size 5mm deep defect in the repair. Obviously the repaired life is always lower than

    the un-repaired life, and gets comparatively worse as the repaired toughness reduces. If the repaired

    defect is 4.2 mm high (diamonds), a reduction in toughness is needed to get a worse life out of the

    repair. The triangular symbols for the shallow 3 mm high defect in the repair are well over unity due

    to a large life of that repair. This is due to the defect tip in question lying well inside the compressive

    region of the repair residual stress field, giving low initial crack growth rates. However, the fatigue

    lives are generally very long for this un-repaired defect (see rightmost curve in Figure 51(a). Leaving

    un-repaired small height defects near the middle of the plate is therefore likely to be a reasonable

    course of action.

    Results comparing a 6.7mm high defect in the un-repaired weld with 6.7 mm, 5 mm or 4.2 mm high in

    the repair are illustrated in Figure 56(b). The diamonds show that introducing a smaller 5 mm defect

    in the repair always gives a shorter life. The triangles start to appear, showing the smallest repaired

    defect of 4.17 mm where the fatigue life ratios remain well above unity.

    Figure 56(c) compares the 9.2 mm high un-repaired defect with 9.2 mm, 6.7 mm or 5 mm in the

    repair. The diamonds have moved up slightly compared with the previous graph, but the squares have

    shifted downwards. This is an interaction between tip position and the associated residual stress field.

    The defect tip is growing towards the repaired surface, so initially higher (longer) defects experience

    more strongly the tensile region of the residual stress near the repair surface.

    In Figure 56(d), a 10.8 mm high un-repaired defect is compared with smaller 9.2 mm, 6.7 mm or 5

    mm defects in the repair. Leaving the same size in the repair (squares) always gives a shorter

    operating life, particularly for lower toughness. The 6.7 mm repair defect (diamonds) needs a large

    reduction in toughness to give shorter life than the un-repaired 10.8 mm defect.

    26

  • The final graphs, Figures 57(a-b) show similar comparisons for the embedded cases in which

    p+2a=10.8 mm. There is a more restricted range of defect heights and toughness to consider and so

    fewer points are plotted than in Figure 56. Also, the growing defect tip of interest lies at a

    comparatively shallow depth in the repair, so it tends to expe