[Elsevier] Review of Fatigue Assessment Procedures for Welded Aluminium Structures

download [Elsevier] Review of Fatigue Assessment Procedures for Welded Aluminium Structures

of 20

Transcript of [Elsevier] Review of Fatigue Assessment Procedures for Welded Aluminium Structures

  • 7/26/2019 [Elsevier] Review of Fatigue Assessment Procedures for Welded Aluminium Structures

    1/20

    International Journal of Fatigue 25 (2003) 13591378

    www.elsevier.com/locate/ijfatigue

    Review of fatigue assessment procedures for weldedaluminium structures

    S.J. Maddox

    TWI Ltd, Granta Park, Great Abington, Cambridge CB1 6AL, UK

    Received 5 February 2003; accepted 24 February 2003

    Abstract

    This paper presents a review of methods and corresponding Codes and Standards for the fatigue assessment of welded aluminiumalloy structures. Methods for the fatigue evaluation of welded aluminium structures are assessed from the viewpoints of originaldesign and estimation of the residual life of existing structures. Based partly on a literature search, but also reference to data usedin the formulation of recent fatigue design Standards, it goes on to review the information available for such assessments in designor guidance specifications in the light of relevant fatigue data. With regard to design specifications, particular attention is focussedon recent fatigue data obtained from structural components representative of actual structures. Recommendations are made forfuture research. 2003 Elsevier Ltd. All rights reserved.

    Keywords:Aluminium alloys; Cumulative damage; Design codes; Experimental data; Fatigue design; Fatigue crack growth; Fitness for purpose;

    Stress analysis; Structural fatigue tests; Variable amplitude fatigue; Welded joints

    1. Introduction

    There is growing interest in the structural use of alu-minium alloys, for such applications as automotive andrailway vehicles, bridges, offshore structure topsides andhigh-speed ships. In all cases, welding is the primary

    joining method and fatigue is a major design criterion.However, as is well known, welded joints can exhibitpoor fatigue properties. Thus, clear design guidelines areneeded to ensure that fatigue failures are avoided inwelded aluminium alloy structures. Apart from basicdesign of new structures, there is also increasing interestin methods for assessing the remaining fatigue lives ofexisting structures.

    Prompted by difficulties experienced in reaching aconsensus on fatigue design rules, extensive testing andanalysis of the fatigue performance of welded aluminiumalloys have been undertaken over the past 20 years. Ameasure of the research effort is the series of Inter-

    Corresponding author: Tel.: +44-1223-897762; fax: +44-1223-

    892588.

    E-mail address:[email protected] (S.J. Maddox).

    0142-1123/$ - see front matter 2003 Elsevier Ltd. All rights reserved.

    doi:10.1016/S0142-1123(03)00063-X

    national Aluminium Conferences (INALCO), which hasproduced seven volumes of papers since 1981. Thefatigue research work culminated in the production ofnewdesign specifications, notably BS 8118[1], Euroc-ode 9[2],the International Institute of Welding (IIW)[3]recommendations and specifications from the AluminumAssociation[4]in the USA and the Canadian StandardsAssociation [5]. In relation to ships, DNV issued sup-plementary guidance [6] to their Rules for the Classi-fication of High Speed and Light Craft based on theECCS recommendations [7], the forerunner to Euroc-ode 9.

    Thus, it is an opportune time to review the variousfatigue design procedures for welded aluminium alloystructures. This paper summarises methods for designand remaining-life assessments of fatigue-loaded alu-minium alloy welded structures and compares and con-trasts the information contained in the various recentfatigue design specifications. This includes assessmentof the corresponding design curves in the light of fatiguetest results, chiefly obtained in recent research pro-grammes where a major aim was to reproduce the fatigueperformance of full-scale welded structures, particularlywith respect to the effect of high tensile residual stresses.

  • 7/26/2019 [Elsevier] Review of Fatigue Assessment Procedures for Welded Aluminium Structures

    2/20

    1360 S.J. Maddox / International Journal of Fatigue 25 (2003) 13591378

    Nomenclature

    A constant in equation for SNcurve

    a crack sizeC constant in fatigue crack growth law

    K stress concentration factorKm stress magnification factor due to misalignmentm slope ofSNcurven exponent in fatigue crack growth law

    ni number of applied load cycles at stress rangeSi (i = 1, 2, 3,)

    N fatigue life in cyclesNi fatigue life at stress range Sip exponent in plate thickness correctionS stress rangeSeq equivalent constant amplitude stress rangeY correction factor in formula for stress intensity factor

    K stress intensity factor range

    Finally, areas still requiring further research are ident-

    ified.

    2. Fatigue assessment methods

    2.1. Background

    Broadly speaking, any fatigue assessment involves

    comparison of theactionswhich the component or struc-

    ture under consideration will be required to sustain dur-

    ing its design life (e.g. fatigue loading history, resultingstresses and the number of times they occur, anyenvironmental influences, etc.) with its resistance tofatigue. Clearly, the resistance must be sufficient to with-stand the actions without failure occurring. The form and

    source of the resistance data depend on the type of

    assessment being performed.

    A useful summary of the four main methods for

    assessing the fatigue lives of welded joints is contained

    in the new IIW fatigue design recommendations [3].

    They are:

    (a) SN curves for specific welded joints used in con-junction with nominal stresses.

    (b) SNcurves for welds used in conjunction with hot-

    spot stresses.

    (c) SN curves for materials used in conjunction withlocal notch stresses.

    (d) The fracture mechanics approach, whereby fatigue

    crack growth data are used in conjunction with the

    stress intensity factor to calculate the progress of a

    knownflaw.

    The first three are intended for application at the designstage and are described in detail in Section 2.2. The

    fourth is not generally used for design but for assessing

    known or assumed flaws. Thus, it would be applicablein an assessment of residual fatigue life, as described in

    Section 2.3.

    2.2. Design of a new structure

    2.2.1. Method

    Fatigue resistance data for design are usually

    expressed in terms of SN curves, relating nominal

    applied cyclic stress rangeSand the corresponding num-

    ber of cycles Nneeded to cause failure. In the simplestsituation, the designer would ensure that the number ofapplied load fluctuations, n, in the design life thatresulted in stress range Sdid not exceed N. In the more

    general case of a detail, which will experience a spec-

    trum of applied loads, the cumulative damage due to

    individual load cycles would need to be determined. The

    usual method is to apply Miners rule, which assumesthat the fatigue damage due to ni cycles of stress Si is

    directly proportional to ni/Ni. An important step in the

    assessment is estimation of the stress history that willbe experienced by the detail under consideration. In gen-

    eral, this involves identification of the loading history,conversion from loads to stresses (e.g. by finite elementanalysis (FEA) or strain gauge measurements) and,

    finally, extraction of recognisable stress cycles from thestress spectrum (the process of cycle counting) to pro-vide input to Miners rule. The full procedure is illus-trated in Fig. 1.

    2.2.2. SN curves used with nominal stresses

    The SNcurves used in fatigue design depend on theprocedure being used. Referring to those mentioned earl-

    ier, by far the most common approach is to use SN

    curves obtained from fatigue tests on specimens contain-

  • 7/26/2019 [Elsevier] Review of Fatigue Assessment Procedures for Welded Aluminium Structures

    3/20

    1361S.J. Maddox / International Journal of Fatigue 25 (2003) 13591378

    Fig. 1. (a) Miners linear cumulative damage rule for estimating fatigue lives under variable amplitude loading; (b) analysis of fatigue loadingfor cumulative damage calculations.

    ing the weld detail of interest (Fig. 2). Such SNcurvesappear in many codes and standards, including some that

    apply to welded aluminium alloys. The design curve is

    usually some statistical lower bound to published experi-

    mental data, typically mean2 standard deviations of

    logN. Since the SN curves refer to particular weld

    Fig. 2. Examples of designSNcurves for welded joints (from IIW

    recommendations for aluminium [3]).

    details, there is no need for the user to attempt to quan-tify the local stress concentration effect of the weld detail

    itself. Thus, the curves are used in conjunction with the

    nominal stress range near the detail. In codes and stan-

    dards, the curves are identified by arbitrary letters or,increasingly, by the fatigue strength at a particular life,

    usually 2 106 cycles. The current status of fatiguedesign rules for welded aluminium alloys is discussed

    in more detail later.

    2.2.3. Hot-spot stress approach

    The hot-spot stress method is an extension of the SN curve approach in that it makes use of SN curvesobtained from tests on actual welded joints. However,theSNcurve is based on the hot-spot stress range ratherthan the nominal. Nominal stress is easy to define insimple laboratory specimens. However, in real structures

    the presence of gross structural discontinuities, non-uni-

    form stress distributions and through-thickness stressgradients can be so complex that the nominal stress is

    no longer obvious. Experimental (e.g. strain gauges) and

    numerical (e.g. FEA) stress analysis methods are capable

  • 7/26/2019 [Elsevier] Review of Fatigue Assessment Procedures for Welded Aluminium Structures

    4/20

    1362 S.J. Maddox / International Journal of Fatigue 25 (2003) 13591378

    of providing detailed information about the stresses aris-

    ing near welded joints. In such circumstances, the struc-

    tural stress, which includes the effect of all sources of

    stress concentration except the weld itself, can be used.

    The hot-spot stress, which is discussed in more detail byNiemi [8], is the structural stress at the weld toe. It is

    usually necessary to determine it by extrapolation fromthe stress distribution approaching the weld (Fig. 3).

    However, parametric formulae exist for calculating the

    hot-spot stress in some tubular joints[9],and more suchformulae are likely to be developed as the hot-spot stress

    method becomes more widely used. A practical limi-

    tation is that the hot-spot stress method is only suitable

    for assessing weld details from the point of view of

    potential failure from the weld toe.

    Apart from tubular joints, there are no established SNcurves for use with the hot-spot stress. The SNcurvesfor use with the nominal stress are not generally suitable

    because they include some influence of the stress con-centration effect of the welded joint. Thus, for example,

    the SN curve for a fillet welded cover plate is belowthat for a simple fillet welded stiffener because of thegreater stress concentration effect of the cover plate. An

    obvious candidate for a hot-spot stress SNcurve is thatfor transverse butt welds, since there is essentially nostress concentration effect due to the joint (provided it

    is perfectly aligned), only the weld bead. Indeed this is

    the case for tubular joints.

    2.2.4. Notch stress approach

    While the notch stress approach applies only to assess-

    ments of potential failure from the weld toe or root, themethod attempts to include all sources of stress concen-tration, including the weld itself, in the stress used with

    the design SN curve. Thus, in principle, a single SN

    curve is sufficient for a given type of material. A practi-cal problem is that the local geometry of the toe or root

    Fig. 3. Stress distribution approaching a welded joint and the defi-

    nition of the hot-spot stress.

    of a weld is highly variable and, at the design stage, not

    known. In recognition of this problem, the weld

    geometry is normally idealised as having a particular

    shape and weld toe or root radius. The local stress is

    then calculated by numerical analysis. Alternatively,parametric formulae are available for a range of joint

    geometries[10].Until very recently[6,11],the notch stress method did

    not appear in any fatigue design specifications. Indeed,one of its protagonists [10]only recommends it for car-rying out comparative studies of the fatigue performance

    of different welded joint options. Furthermore, the

    method has not been developed to any extent for alu-

    minium alloys. Consequently, it is not considered further

    in this report.

    2.3. Remaining life of existing structures

    Broadly three approaches can be envisaged for thefatigue assessment of existing structures which have

    already experienced some service duty. The approach

    used will depend on the circumstances, particularly

    whether or not the structure was designed for fatigueloading, the time in service and what measures will be

    taken to assess its current condition with respect to

    potential fatigue damage already introduced during pre-

    vious service.

    Three assessment methods are described sub-

    sequently. Examples of their application or reference to

    their development may be found in Refs. [1214].

    2.3.1. Fatigue design assessmentThis method follows the procedure outlined in Section

    2.2.1 for original design. If the structure was designed

    for fatigue loading, the same actions can be assumed,after any modifications to allow for changes such asreduced severity of the stress history from reinforcement

    or a change in the operating conditions. Fatigue resist-

    ance is still represented by the design SN curves. Ifrepairs are introduced, the design curves may still be

    applicable, but a safety factor could be introduced if the

    repair was of uncertain quality. Post-weld improvement

    of repair welds, for example by toe grinding (to be dis-

    cussed in more detail later), would justify a higher SNcurve, which may be included in a design specificationor obtained from appropriate published information.

    Finally, assuming Miners rule (see Section 3.3.5 regard-ing validity of this assumption), it is used to calculate

    the fatigue damage introduced before and after the time

    of the assessment, on the basis that

    n/Nbefore

    n/Nafter

    1 .

    Then, the remaining life is given by:

    Remaining life (e.g. in years) (1)

  • 7/26/2019 [Elsevier] Review of Fatigue Assessment Procedures for Welded Aluminium Structures

    5/20

    1363S.J. Maddox / International Journal of Fatigue 25 (2003) 13591378

    life so far (years) 1n/N

    before

    n/Nafter

    per year .

    2.3.2. Fatigue design review

    The aim of a design review would be to improve theaccuracy of the original design process to provide a bet-

    ter estimate of the proportions of fatigue life used andremaining at the time of the assessment. A fatigue design

    process involves many assumptions and hence potential

    inaccuracies. When assessing an existing structure, there

    may be scope for improving the accuracy of some of

    those assumptions. For example, records of the service

    operation or even measurements made on the structure

    may enable a more precise definition of the stress his-tory. Knowledge of the actual structural arrangement and

    weld details used, including their quality (e.g.alignment), coupled with appropriate stress analysis may

    allow the more precise hot-spot stress approach to be

    used instead of the normal SN curves. A furtherrefinement might be the characterisation of actionsand/or resistance data in statistical terms to enable

    reliability methods to be used to assess the risk associa-ted with a particular estimate of remaining life. Some

    progress has been made in such an approach in the con-

    text of steel bridges [15].

    2.3.3. Fracture mechanics approach

    The third method specifically addresses circumstancesin which it has been found, or it must be assumed, thatflaws (e.g. fatigue cracks) have been introduced duringthe service life endured so far. Such flaws would bethose detected or measured by non-destructive testing(NDT), or assumed flaws corresponding to the limit ofdetection of the NDT methods used.

    A fracture mechanics assessment [16] utilises the

    same actions as those determined for design calculations.

    However, fatigue resistance is represented by fatigue

    crack growth rate data for the material under consider-

    ation, expressed in terms of the fracture mechanics stress

    intensity factor parameter K. K is a function of

    applied stress range (S), and crack size (a), such that:

    K YSa (2)

    where Yis a function of geometry and loading. The use

    ofKensures that the relationship between crack growth

    rate and Kcan be regarded as a law applicable to any

    geometry of the same material. The crack growth law

    approximates to a linear relationship (usually referred toas the Paris law):

    da

    dN C(K)n (3)

    with deviations as K approaches a threshold value

    below which crack growth is insignificant (Ko) and asKmax approaches the critical value for fracture, as illus-

    trated inFig. 4.In practice, the gradual transitions from

    the Paris law may be modelled more accurately bydefining several linear relationships. For a flaw size ai

    and a critical fatigue crack size of af, the remainingfatigue life N under stress range S is obtained by inte-

    grating Eq. (3):

    af

    ai

    da

    (YSa)n CN (4)

    For variable amplitude loading the integration will be

    performed for each individual cycle or block of equal

    stress cycles, to give:

    a1

    ai

    da

    (YS1a)n

    a2

    a1

    da

    (YS2a)n % etc. (5)

    CN

    3. Fatigue design data

    3.1. Design specifications

    As noted earlier, there is a wide choice of fatigue

    design specifications for welded aluminium alloys. Themain ones, in chronological order, are as follows:

    BS 8118:1991. Structural use of aluminiumPart 1Code of practice for design, BSI, London 1991.

    Fig. 4. Fracture mechanics fatigue crack growth relationship.

  • 7/26/2019 [Elsevier] Review of Fatigue Assessment Procedures for Welded Aluminium Structures

    6/20

    1364 S.J. Maddox / International Journal of Fatigue 25 (2003) 13591378

    ECCS. European recommendations for aluminium alloy

    structures, fatigue design, European Convention for

    Constructional Steelwork, Document No. 68, 1992.

    Canadian Standards Association CAN/CSA-S157-M92.

    Strength design in aluminium, 1993.The Aluminum Association. Specifications for alu-

    minium structures, Washington, DC, 1994.DNV. Class note: fatigue assessment of aluminium

    structures, Technical Report No. LIB-J-000010, 1995.

    International Institute of Welding. Fatigue design ofwelded joints and components, Abington Publishing,

    1996.

    Eurocode 9. Design of aluminium structures. Part 2.

    Structures susceptible to fatigue, ENV, 1999-2: 1998,

    CEN, Brussels, 1998.

    Apart from the DNV document, these all provide a

    selection of design SN curves expressed in terms of

    nominal stress ranges. In the DNV note, attention is con-

    fined to the use of the hot-spot stress range, a methodthat is also referred to in the IIW recommendations and

    Eurocode 9, but specific design data are not provided.There are significant differences between the SNcurvesin the rules and how they are used, and hence the differ-

    ent specifications will lead to different estimates offatigue life. In order to provide a basis for judging their

    applicability to welded aluminium structures, key fea-

    tures are compared and where possible assessed in the

    light of relevant published data.

    3.2. Historical developments

    In order to appreciate why so many fatigue designspecifications have been produced in recent years, it isuseful to review the developments over the past 20 years,

    mainly in Europe, which have influenced them.In 1979, it was decided that the British Standard

    design specification for aluminium, CP 118, which wasthe most comprehensive standard for aluminium at the

    time and used throughout the world, should be revised

    as a limit state Standard. This followed the publication

    of new steel bridge design rules, BS 5400, on the samebasis. With regard to fatigue, as a starting point the

    possibility was examined that the new rules for steels

    could be simply factored in accordance with the differ-

    ence in Youngs modulus between steel andaluminiumto provide the aluminium fatigue rules [17]. This

    approach stemmed largely from the good correlationbetween fatigue crack growth data for the two materials

    on the basis ofK/Eand the assumption that the fatigue

    lives of welded joints are dominated by fatigue crack

    growth[18].Thus, the fatigue design stresses for welded

    aluminium alloys would be obtained simply by dividingthose for steel by 3. A review of published data tended

    to support this approach and was adopted for the Draft

    for Public Comment of the Standard that would replace

    BS CP 118, BS 8118, in 1985. However, meanwhile

    some industries had drawn attention to the fact that the

    new rules were considerably more conservative than

    those in BS CP 118 in the high-cycle regime, while

    others felt that the steel/3 approach was too simplisticand effectively penalised aluminium alloys as compared

    with steel.The initial review of fatigue data for welded joints in

    aluminium alloys had drawn attention to the wide scatter

    in published data and the fact that most data referred tosmall-scale specimens of variable, unspecified quality.Another important characteristic of small-scale speci-

    mens, particularly those incorporating transverse welds,

    is that they will contain much lower tensile residual

    stresses than would be expected to be present in a real

    structure. It was felt that more realistic fatigue data rel-evant to actual welded structures were required. These

    arguments influence the newly formed ECCS Committeecharged with the task of drafting a European Standard.

    Consequently, they placed particular reliance on data

    obtained from realistic structural specimens, mainly

    beams. A large database was available from one source

    (Alusuisse) and this was made available to the Commit-

    tee. In addition, a number of new European projects pro-

    vided additional data that were taken into consideration.To some extent, the same results were used to review

    the BS 8118 Draft for Public Comment and the fatigue

    rules were revised slightly as a result.

    The resulting ECCS[7] and BS 8118[1] fatigue rules

    were finallyconsidered together as the basis of the newEurocode 9[2]in the early 1990s. Even more large-scale

    specimen data were available by then and so the finalform of Eurocode 9 is different from both BS 8118 and

    the ECCS specification.Other significant developments were the drafting of

    fatigue design rules in the USA[4]and Canada[5],bothof which are knownto have been influenced by the Euro-pean activities [19].

    3.3. Summary of design rules

    3.3.1. Design SN curves

    All the fatigue design specifications for welded alu-minium alloys present a series ofSNcurves for parti-cular weld details, with a classification scheme linkinga description of the welded joint with the appropriate

    design curve. Examples of the SNcurves provided areshown in Fig. 5. The classification usually depends onthe joint type, geometry, loading direction and mode of

    fatigue failure, as illustrated for one group of joint types

    in Eurocode 9 inFig. 2.Rather less comprehensive guid-

    ance is given by the Aluminum Association [4], whichonly refers to joint types and loading direction. The SNcurves are derived from linear regression analysis of

    logSversus logNfatigue data to establish mean curves

  • 7/26/2019 [Elsevier] Review of Fatigue Assessment Procedures for Welded Aluminium Structures

    7/20

    1365S.J. Maddox / International Journal of Fatigue 25 (2003) 13591378

    Fig. 5. Examples of fatigue designSN curves for aluminium alloys

    in recent specifications: (x) IIW [3]; (b) Aluminum Association [4];

    (c) Eurocode 9[7].

    and statistical lower bound, usually mean 2 standarddeviations of logN. The SNcurves have the form:

    SmN A (6)

    where A and m are constants. The curves are assumed

    to extend up to stress levels corresponding to the static

    design limit for the material, and down to a fatigue

    endurance limit. A constant amplitude fatigue endurance

    limit is introduced at an endurance of 5 106 cycles in

    all the aluminium specifications. Two basic approacheshave been used to define the design SNcurves:

    (a) An arbitrary grid of SN curves, usually equally

    spaced on loglog scales, is defined and the curveclosest to the selected lower bound to experimentaldata of a particular detail is allocated to that detail.

    The IIW recommendations [3] are based on thisapproach with the SN curves defined in terms ofthe stress range in N/mm2 at 2 106 cycles (see

    Fig. 5(a)).

    (b) The design SN curves are derived directly fromexperimental data. In some cases, the fatigue lives

    of different details are so similar that the experi-mental data can be combined to produce a single

    class for all of them. The resulting SN curves,which are not usually equally spaced, may be

    described in terms of the fatigue strength at 2 106 cycles, and possibly the slope m of the SNcurve as in draft Eurocode 9[2],or by arbitrary let-

    ters such as Class A, B, C, etc., in the Aluminum

    Associations specification [4] (see Fig. 5(b) and(c)).

    In most cases, the SN curves get progressivelysteeper as the fatigue strength of the detail decreases.

    The steepest curve usually has a slope which is consist-

    ent with crack growth data (i.e. m = n, typically 34),reflecting the fact that the lives of the low fatiguestrength details are dominated by crack growth [17,18].

    Apart from the implied need to utilise mechanised

    welding to achieve continuous welds without stop/starts,

    no distinction is drawn between different welding pro-

    cesses. The bulk of the test data upon which the design

    curves are based have been obtained from arc welds.

    However, as other processes became more viable forwelding aluminium, notably friction-stir welding [20]

    (see later) which seems to offer advantages from the

    fatigue viewpoint as well as production, it may become

    necessary to introduce new process-related categories.

    3.3.2. Effect of residual stress and mean stress

    Welding introduces tensile residual stresses, which

    modify the mean stress experienced by the welded joint

    under fatigue loading. Long-range, or reaction, residual

    stresses will also be introduced when welded sub-

    assemblies are connected together, due to imperfect fit-up. It is generally assumed that tensile residual stresses

    up to the proof strength of the material will be present

    in a welded structure. As a result, its fatigue life will be

    independent of mean stress and depend only on the

    applied stress range, even if this is compressive [21].Consequently, all the fatigue design specifications arebased on the use of full stress range regardless of

    whether it is tensile or compressive.

  • 7/26/2019 [Elsevier] Review of Fatigue Assessment Procedures for Welded Aluminium Structures

    8/20

    1366 S.J. Maddox / International Journal of Fatigue 25 (2003) 13591378

    3.3.3. Material

    A common feature of all the specifications is that nodistinction is drawn between different aluminium alloys

    when welded, unless they are exposed to a corrosive

    environment. This reflects the fact that fatigue crackgrowth rates are not significantly different in different

    alloys[22]and that fatigue crack growth dominates thefatigue lives of welded joints. Consequently, precise

    details of alloys used to produce welded test specimens

    discussed later are only given if they are significant.For unwelded material, some of the specifications

    [2,3] provide higher design stresses for high strength

    7000 series aluminium alloys as compared with all the

    other alloy types.

    3.3.4. Effect of plate thickness

    It is generally acknowledged that the fatigue strengths

    of welded joints failing from the weldtoe can decrease

    with increase in plate thickness [23]. This has led to

    thickness effect penalties, applied to the fatigue strength

    obtained from the SN curve, in many fatigue designrules for welded steel of the form (tref/t)

    p, where t is thethickness, tref the reference thickness (usually around 25

    mm) and the exponent p = 0.25. Recent work showed

    that the thickness effect also depended on the overallproportions of the welded joint[24,25].These influencesare incorporated in the fatigue rules in Eurocode 9. A

    further refinement in the IIW recommendations[3] is tomodify the thickness correction exponent p for different

    weld details. Values range from 0.3 to 0.1, reflecting thefact that the thickness correction also depends on the

    level of stress concentration introduced by the weldedjoint. In contrast, the Aluminum Association take theview that the database used to establish the design SNcurves covered the full range of thicknesses of alu-

    minium alloy likely to be used in practice [19]. Hence,

    there is no requirement to apply a thickness effect cor-

    rection. This assumption may be reasonable for some

    applications (e.g. automotive or railway vehicles where

    plate thickness is unlikely to exceed 25 mm) but not for

    large structures such as bridges or LNG tankers where

    plate thickness may be 100 mm or more.

    3.3.5. Cumulative damage

    Miners rule is universally adopted as the method forpredicting fatigue lives under variable amplitude loading

    using the constant amplitude design SNcurves. How-ever, the accuracy of the rule has been called into ques-tion in recent years as more and more fatigue tests

    obtained under random loading conditions have pro-

    duced failures in shorter lives than those predicted by

    Miners rule[26,27].It is thought that part of the reasonfor this is that the crack closure conditions for a givenstress fluctuation are different under constant and vari-able amplitude loading, with the result that a stress range

    may be more damaging in a variable amplitude sequence

    than it was under constant amplitude loading[27].How-

    ever, a second problem concerns the damaging effect of

    stress ranges below the constant amplitude fatigue limit.

    Some specifications take account of such stresses byassuming that the SNcurve extends below the constantamplitude fatigue limit at a shallower slope. For an S

    N curve of the form SmN = A, the extrapolated curvewould be of the form Sm + 2N= A, (seeFig. 5(c)). How-

    ever, on the basis of fatigue tests on large-scale welded

    beams (to be discussed later), the Aluminum Association[4]take the view that the SNcurve should be extrapo-lated indefinitely below the constant amplitude fatiguelimit without a slope change. The extent to which these

    modifications to the SN curve are successful will beconsidered later in the light of new experimental data.

    3.3.6. Hot-spot stress approach

    Only the DNV note[6] gives specific guidance on the

    use of the hot-spot stress fatigue design procedure. Thatguidance is related to four SN curves from the ECCSrecommendations [7], one for unwelded material, two

    for welded connections and the fourth for welds exposed

    to a corrosive environment (presumably seawater). The

    corresponding ECCS design curves are as follows:

    DNV class Material ECCS SNcurve

    I Unwelded Unwelded high

    strength 7020

    alloy

    II Welded Flush ground buttwelds

    III Welded As-weldedtransverse butt

    welds with good

    profileIV Welded, in Not included:

    corrosive 25% reduction in

    environment design stress from

    Class III curve

    Thus, for as-welded joints it is assumed that the butt

    weld design SN curve is applicable to both butt and

    fillet welds if the hot-spot stress is used. The reason forthe choice of the SN curve for the highest strength ofunwelded aluminium alloy (lower curves were provided

    for other aluminium alloys) is not known. The corre-

    spondingSNcurves in Eurocode 9 are lower than those

    in the ECCS rules.

    The curves are used in conjunction with specifiedstress concentration factors, K, by which stress ranges

    obtained from the specified design curve are divided, fora variety of typical welded connections used in ships.

  • 7/26/2019 [Elsevier] Review of Fatigue Assessment Procedures for Welded Aluminium Structures

    9/20

    1367S.J. Maddox / International Journal of Fatigue 25 (2003) 13591378

    The range is not totally comprehensive and, as will be

    seen later, no guidance is offered for some relevant

    details.

    3.3.7. Effect of environment

    As already mentioned, the DNV note provides guid-

    ance on the effect of immersion in a marine environmentby adopting an SNcurve which is 25% on stress below

    the lowest curve for welded joints. That same curve

    would be used for any detail, welded or unwelded. Thereduction in allowable stresses for as-welded joints of

    25% is in-line with the results of an extensive series of

    corrosion fatigue tests conducted in Norway in the early

    1980s[28].

    Eurocode 9 also provides guidance on the influenceof environment, industrial and marine. The basicapproach is to reduce the detail classification, by up tothree categories in the case of immersion in seawater,

    and to reduce the fatigue endurance limit, as illustrated

    in Fig. 6. The design penalty is most severe for 7000

    series alloys, which are also susceptible to stress cor-

    rosion cracking, while no reduction in design category

    is required for 3000 series and aluminiummagnesium5000 series alloys, although the fatigue endurance limit

    is still reduced. In general, the extent of the reductionin fatigue strength due to environment depends on the

    detail and endurance since the SNcurves are not paral-

    lel. It is not clear how the category-reduction approach

    should be applied to the lower category details, for

    which the required reduction would take them below the

    design categories provided.

    4. Comparison of design proposals and recent

    fatigue data

    4.1. Background

    All the recent design SNcurves for welded joints in

    aluminium alloys are claimed to have been derived from

    Fig. 6. Eurocode 9[7] allowance for reduction in fatigue strength due to marine corrosion.

    experimental data by linear regression analysis such that

    they represent approximately 97.7% probability of sur-

    vival [19,2931]. However, it is not always clear howthis has been achieved. It is evident that in most cases

    some judgement has been applied and even someassumptions, like the slope of the design SN curve,

    have been imposed. It is also claimed that special atten-tion was paid to the provision of data relevant to real

    structures, particularly with respect to the influence oftensile residual stresses. Thus, wherever possible themain basis of the design curves has been experimental

    results obtained from full-scale welded specimens, usu-

    ally beams, or from specimens tested under high tensile

    mean stress conditions to simulate the effect of high ten-

    sile residual stresses. In view of this, the design SN

    curves should be consistent with such data, includingdata generated since the curves were published. In order

    to provide a basis for judging the validity of the pro-

    posed design curves, relevant published data have been

    assembled and they are presented in comparison with

    some of the design curves. In this exercise, whenever

    possible, attention has been focused on test results

    obtained from specimens made from material 816 mmin thickness, since plate thickness is known to influencefatigue performance. Furthermore, only design curvesfrom the four most recent specifications, namely theAluminum Association specification, Eurocode 9, theIIW recommendations and the DNV design note, are

    considered. The majority of the experimental data were

    obtained under constant amplitude loading and presented

    in terms of nominal stress range. Hence, these will be

    used to assess the design SN curves intended for usewith nominal stress range. Many of the results for beamswere obtained from the compilation of data in Ref. [30],

    which does not always give full details of the source

    (which may have been an internal company report). Ref-

    erence is made to the relevant series in that reference.

    Limited data presented in terms of hot-spot stress

    range are also available and these will be used to assess

    the proposed hot-spot stress SNcurves. Finally, some

  • 7/26/2019 [Elsevier] Review of Fatigue Assessment Procedures for Welded Aluminium Structures

    10/20

    1368 S.J. Maddox / International Journal of Fatigue 25 (2003) 13591378

    data have been obtained under variable amplitude load-

    ing. These will be used to assess the validity of Minersrule and the methods proposed to modify the SNcurve

    below the constant amplitude fatigue limit to take

    account of the damaging effect of low stresses. The vari-able amplitude fatigue data will be considered in terms

    of the equivalent constant amplitude stress range, calcu-lated on the basis that Miners rule is correct using theconstant amplitude SNdata obtained in the same inves-

    tigation. This equivalent stress range is as follows:

    Seq mSmi n ini (7)where ni cycles were applied at stress range Si before

    failure, ni the total number of cycles to failure and m

    is the slope of the SN curve. In those cases where a

    significant number of applied stress cycles were belowthe constant amplitude fatigue limit, the SNcurve wasassumed to be extrapolated below this limit at a slope

    ofm + 2, as proposed in most of the specifications.

    4.2. Continuous longitudinal welds

    The stress concentrations introduced by continuous

    longitudinal butt and fillet welds, weld ripples or lumpsif stop/starts are present, are relatively minor. Their cor-

    responding fatigue performance is relatively good. The

    severity of a stop/start would be intensified if cratercracking occurred, which is certainly a possibility when

    welding aluminium alloys, but these would normally berepaired if found.

    In spite of the relatively good fatigue performance of

    continuous longitudinally loaded welds, they are

    important details, particularly in welded aluminium alloy

    structures. Such details may be the governing ones in

    well designed structures in which poorer transverse

    welds have been avoided or located in low stress regions.

    The potential for doing this is enhanced in the case of

    aluminium alloys because of the enormous scope for

    producing special extrusions, for example of sufficientrigidity in relevant parts to avoid the need for welded

    stiffeners.

    Continuous longitudinal welds are not explicitly

    included in the DNV note. Of the other specifications,both the IIW and Eurocode 9 distinguish between welds

    with and without stop/starts. The IIW recommendationsalso draw a distinction between butt and fillet welds. TheAluminum Association gives only one design category

    for both butt andfillet welds with and without stop/starts.Recent fatigue data obtained from structural compo-

    nents [30 (series C2, D1 and D2), 32,33] are confinedto I-section beams in which the test detail is either a

    continuous butt weld in the web or, in the case of fabri-

    cated beams, the web to flange weld. Data are available

    for 5000, 6000 and 7000 series alloys in thicknesses

    from 6 to 15 mm. They are shown together with relevant

    design SN curves in Fig. 7(a) for welds without

    stop/starts andFig. 7(b)for welds with stop/starts.

    Referring first to the results for welds withoutstop/starts(Fig. 7(a)), it will be seen that they are most

    consistent with the slope of the IIW design curves (i.e.m = 3), although only the Category 40 design curve is

    safe for all the results. The Eurocode 9 and Aluminum

    Association design curves appear to be too shallow. Thismay be a situation in which the slope of the design curve

    has been imposed. Both the Eurocode 9 and Aluminum

    Association specifications provide SN curves whichbecome gradually steeper as the fatigue strength of the

    detail decreases and the adoption of a slope of m = 3

    for such a high fatigue performance detail would intro-duce an anomaly into such a scheme. However, com-

    pared with actual data, the result is that both the Euroc-

    ode 9 and Aluminum Association design curves are

    particularly conservative in the high stress/low fatigue

    life regime.

    The case for a shallow SNcurve is better for welds

    containing stop/starts positions, as seen inFig. 7(b).This

    Fig. 7. Comparison of fatigue test results obtained from (a) continu-

    ous longitudinal butt and fillet welds without stop/starts and design

    curves and (b) continuous longitudinal fillet welds containing

    stop/starts and design curves.

  • 7/26/2019 [Elsevier] Review of Fatigue Assessment Procedures for Welded Aluminium Structures

    11/20

    1369S.J. Maddox / International Journal of Fatigue 25 (2003) 13591378

    situation arises mainly because of the wide scatter asso-

    ciated with thefillet weld results. There are some fatiguedata below all the design curves, the IIW curve being

    particularly non-conservative and the Eurocode 9 curve

    being the most suitable. Further investigation of the rea-son for the low [29] results would be worthwhile. By

    and large, it may be noted that the database does notprovide a strong indication that distinction should be

    drawn between butt and fillet welds in design specifi-cations, but the results do support the distinction betweenwelds with and without stop/starts.

    4.3. Transverse butt welds

    This section is concerned with transverse butt welds

    made from one or both sides, with the condition thatthey should be full penetration welds. A number of fac-

    tors will influence the fatigue performance of transversebutt welds and some of them influence the design curves.In particular, a distinction may be drawn between welds

    made from one or both sides and welds with different

    profiles (expressed in terms of the weld toe angle).Further conditions might be that the weld should be

    proved free from significant defects (i.e. those whichmight replace the weld toe as the site for crack initiationand lead to a lower fatigue life) by appropriate inspec-

    tion, and that the effect of misalignment as a source of

    secondary bending stress should be taken into consider-

    ation when calculating the stress experienced by the

    weld.

    A reasonable database from structural specimens con-

    taining transverse butt welds is available, mainly fromI-section beams [30 (series B7, B8, B10, B11), 33,34].These include specimens fabricated or extruded from

    5000, 6000 and 7000 series alloys in thicknesses ranging

    from 8 to 15 mm. In some cases, the weld toe angle is

    reported. The data are shown in comparison with rel-

    evant design curves in Fig. 8.

    There is some indication of an influence of weld toeangle in that the highest results were obtained from

    welds with a toe angle not exceeding 30, while the low-

    est were from welds with angles up to 60. However,some good profile welds also gave lives near the lowerbound and overall the results do not indicate a strong

    correlation between weld angle and fatigue strength.

    Similarly, the results do not provide support for dis-

    tinguishing between one- and two-sided full-pen-

    etration welds.The DNV note distinguishes between one- and two-

    sided welds and welds with different profiles. Defaultstress concentration factor values (by which stresses

    obtained from the DNV III curve are divided) of K =

    1.7 and 1.3, respectively, are given for weld toe anglesup to 50. The IIW recommendation distinguishes

    between one- and two-sided welds and different weld

    profiles, Eurocode 9 only distinguishes between one- and

    two-sided welds, but the Aluminum Association pro-

    vides only one design curve for any transverse butt weld.

    Referring to Fig. 8, it will be seen that the Aluminum

    Association Category C curve and the DNV curve for

    welds made from both sides are very similar and providereasonable lower bounds to the data. The Eurocode 9

    curves are unaccountably low while the IIW curvesappear to be too steep and, apart from FAT 28, too high.

    However, it is interesting to note that regression analysis

    of all the experimental data together results in a meanSN curve with slope of m = 2.95, very similar to the

    assumed slope ofm = 3 in the IIW recommendations.

    4.4. Transverse butt welds made on permanent

    backing

    One technique for ensuring full penetration for butt

    welds from one side only is to use a permanent backing

    bar or, in the case of aluminium alloys, backing lip

    included in the extrusion. For joints in steel, the fatiguestrength of the resulting joint is lower than that obtained

    from butt welds made from both sides, due to the severestress concentration introducedat the weld root between

    the main plate and backing bar [21].Since this is a geo-

    metric effect, it would be expected that the same would

    be found from aluminium alloys. To some extent this is

    the case, but the database is surprisingly limited in view

    of the potential for extruding aluminium sections

    incorporating backing lips. In fact, only one reference to

    tests on structural components [35] could be found. In

    view of this, data obtained from specimens are also con-sidered. The data found in the literature search are given

    in Fig. 9 together with the appropriate design SN

    curves. These refer to plate specimens in 6005 and 7020

    alloys [30 (series B4)], extruded bridge deck panels in

    6005 alloy and specimens extracted from such panels

    [35].In fact, these specimens were reported to be sever-ely misaligned (angular distortion) with the result that

    secondary bending occurred at the weld. The corre-

    sponding stress magnification factor Km was estimatedby the authors and the results are presented in terms of

    Km nominal stress range in Fig. 9.

    It will be seen that most of the data lie above the DNV

    and Aluminum Association design curves, which are

    shallower than the IIW and Eurocode 9 curves. The datatend to follow the slope of the shallower curves, but with

    such a limited database confined to a very limited rangeof relatively low endurances this may be a misleadingimpression. Certainly, in the light of experience of joints

    in steel, the slopes of the IIW and Eurocode design

    curves seem to be more appropriate, but further experi-

    mental data are needed to confirm this.

    4.5. Transverse cruciform joints

    Fatigue data are available for I-section beams incorpo-

    rating cruciform joints [30 (series F1), 33] in 15 mm

  • 7/26/2019 [Elsevier] Review of Fatigue Assessment Procedures for Welded Aluminium Structures

    12/20

    1370 S.J. Maddox / International Journal of Fatigue 25 (2003) 13591378

    Fig. 8. Comparison of fatigue test results obtained from transverse butt welded 5000, 6000 and 7000 series aluminium alloy beams and design

    curves.

    Fig. 9. Comparison of fatigue test results obtained from transverse

    butt welds made on permanent backing and design curves.

    thick5083 alloy and 8 mm thick 6005 aluminium alloy

    [34].In all cases, fillet and full penetration joints, failurewas by fatigue cracking from the weld toe. Also, some

    data have been obtained from a model of a structural

    connection used in an aluminium alloy ship [36]. Thiswas essentially the joint between the hull, T-section

    longitudinal stiffeners and a transverse bulkhead (see

    Fig. 10). The specimens were made in 6.4 mm thick

    5086 H116 alloy. In all cases, the fillet weld sizes weresufficient to avoid failure in the weld throat in preferenceto failure from the weld toe.

    All the results are plotted in Fig. 11 in comparison

    with the appropriate design SNcurves. The detail is not

    explicitly covered in the Aluminum Association specifi-cation but it has been assumed that the design curve for

    transverse fillet welded stiffeners is appropriate. TheEurocode 9 design curve depends on plate thickness and

    joint proportions and the curve shown is applicable to

    the sizes of specimens used to generate the data

    presented. As will be seen, the data are widely scattered.

    However, both the IIW and Eurocode 9 curves appear

    to be representative of the slope of the data and close to

    the lower bound. The DNV and Aluminum Association

    curves are similar, but both seem to be too shallow.

    4.6. Transverse fillet welded attachments and

    stiffeners

    Transverse non-load carrying fillet welded attach-ments and stiffeners are very common in actual struc-

    tures. Like transverse butt welds, small-scale specimensare unlikely to contain high tensile residual stresses andhence be representative of real structures from this view-

    point. Therefore, fatigue test results obtained from struc-

    tural specimens are particularly valuable.

    A reasonable database now exists, as shown in Fig.

    12. Most of the results were obtained from beams with

    full or partial depth web stiffeners in 1115 mm thick5000, 6000 and 7000 series alloys [30 (series E1), 37].

    In all cases, fatigue failure was from the weld toe in the

    flange. In addition, a few results were obtained from I-section beams in12 mm thick 6061-T6[34] or 15 mm

    thick 7020 alloys[30 (series E8)]with simple transverse

    attachments on the tension flange.Fig. 12also includes the relevant design curves from

    the four specifications being considered. As will be seen,the data strongly support the slopes of the Eurocode 9and Aluminum Association SN curves, between m =

    3.2 and 3.6, and indeed those design curves are close to

    the lower bound to the data. The DNV curve appears to

    be too shallow, with a result that it is unduly conserva-

    tive in the short life regime. The IIW curve, on the otherhand, seems to be too steep and over-conservative in the

    long life regime. However, introducing a set of results

    [38] obtained from small-scale specimens makes the

  • 7/26/2019 [Elsevier] Review of Fatigue Assessment Procedures for Welded Aluminium Structures

    13/20

    1371S.J. Maddox / International Journal of Fatigue 25 (2003) 13591378

    Fig. 10. Structural detail representing intersection of hull, longitudinal stiffener and transverse bulkhead fatigue tested by Beach et al.[36].

    Fig. 11. Comparison of fatigue test results obtained from structural

    specimens incorporating cruciform joints and design curves.

    slope of the IIW curve seem more reasonable. Thesespecimens were tested with the maximum stress heldconstant at a value close to proof strength, to simulate

    the presence of high tensile residual stresses. As seen,

    they are rather similar to the results for beams with sim-

    ple attachments. Clearly, more data for the long life

    regime are needed to clarify the slope issue.

    4.7. Longitudinal non-load carrying fillet welded

    attachments

    Specimens incorporating longitudinal non-load carry-

    ing fillet welded attachments offer the advantage that

    Fig. 12. Comparison of fatigue test results obtained from beams with

    transverse fillet-welded attachments or web stiffeners, or plates with

    transverse fillet-welded attachments, and design curves.

    high tensile residual stresses exist even in small-scale

    specimens[18].The detail itself is not particularly com-

    mon in real structures, except perhaps as gussets tostiffen corners. Fatigue failure occurs by crack growth

    from the weld toe at the end of the attachment, the

    fatigue life being similar whether or not the weld is con-

    tinued around the ends of the attachment. However, the

    fatigue life varies with attachment size, the stress con-centration effect at the end of the attachment increasing

    with attachment length.

    Fatigue data are available for beams, extruded and

  • 7/26/2019 [Elsevier] Review of Fatigue Assessment Procedures for Welded Aluminium Structures

    14/20

    1372 S.J. Maddox / International Journal of Fatigue 25 (2003) 13591378

    fabricated, with attachments fillet welded to the tensionflange [30 (series E6, E7), 34,39]. These have beenobtained using specimens made from 5000, 6000 and

    7000 [34] series alloys in thicknesses between 11 and

    15 mm. Some of these were tested under variable ampli-tude loading[39].The loading spectrum used was based

    on strain measurements on a railway freight wagon.Finally, partly to extend the range of endurances, but

    also because they include further fatigue test results

    obtained under variable amplitude loading, recent resultsobtained from 10 mm thick 7019 alloy specimens can

    be considered[38].

    All these data are plotted in Fig. 13, together with

    the appropriate design SN curves, including the curves

    extrapolated below the constant amplitude fatigue limit

    for use when performing cumulative damage calcu-lations. As will be seen, the SNcurves are rather similar

    above the constant amplitude fatigue limit, with slopes

    that are consistent with the experimental data. The

    design curves also lie close to the lower bound to the

    data. Thus, any of the design curves could be supported

    by the database.

    Comparing the data in Fig. 13 with those in Fig. 7,

    which are representative of beams without stiffeners or

    other attachments, provides a striking illustration of thedetrimental effect of attachments on fatigue perform-

    ance. In practice, they should be avoided in highly

    fatigue-loaded areas.

    The variable amplitude fatigue test results for both

    beams and specimens are consistent with the respective

    constant amplitude data, supporting the validity of

    Miners rule for the spectra used. Furthermore, the beamdata extend well below the constant amplitude fatiguelimit and hence provide a useful check on the validity

    of the extrapolated SNcurve. Again, the data could be

    used to support any of the proposals, but the fact that

    they appear to be consistent with the same SN curve

    as the constant amplitude data provides support for the

    Aluminum Associations approach for extrapolating the

    Fig. 13. Comparison of fatigue test results obtained from beams and

    plates with longitudinal fillet-welded attachments and design curves.

    SN curve without any slope change. More results in

    the high-cycle regime are needed to check this further,

    particularly for other loading spectra.

    It will be noted that the results obtained from small-

    scale specimens are entirely consistent with thoseobtained from beams, confirming that the two types of

    specimen incorporated similar high tensile residualstresses. This adds confidence to the use of small-scalespecimens for investigating the fatigue performance of

    this particular detail.

    4.8. Beams with cover plates

    Large cover plates welded to beam flanges representextremely high stress concentrations and consequently

    result in the lowest fatigue performance for welds failingfrom the weld toe. Consequently, beams with cover

    plates have been widely investigated for providing

    design data. This has resulted in a large database for

    beams in aluminium alloys, chiefly from fabricatedbeams in 1015 mm thick 5000, 6000 and 7000 seriesalloys [30 (series F3), 37]. A large number of results

    have also been obtained from smaller beams [40,41], in

    4 mm thick 6261-T6 aluminiumalloy. All these results

    are plotted together in Fig. 14, along with the relevantdesign SN curves from the four specifications con-sidered.

    It will be seen that the IIW and Eurocode 9 design

    curves are very similar and consistent with the database

    in terms of slope and position. The Aluminum Associ-

    ation curve is slightly lower but of similar slope, while

    the DNV curve appears to be too shallow with a resultthat it is excessively conservative in the short liferegime.

    This detail is one in which the thickness effect would

    be expected to apply, and indeed is evident from a com-

    parison of the results obtained from 4 mm thick speci-

    mens with the remainder. This thickness effect is incor-

    porated in both the IIW recommendations and Eurocode

    Fig. 14. Comparison of fatigue test results obtained from beams with

    cover plates and design curves.

  • 7/26/2019 [Elsevier] Review of Fatigue Assessment Procedures for Welded Aluminium Structures

    15/20

    1373S.J. Maddox / International Journal of Fatigue 25 (2003) 13591378

    9 and higher design curves would be used for the 4 mm

    thick specimens.

    4.9. Fatigue data expressed in terms of the hot-spot

    stress range

    Although guidance exists for the determination ofthehot-spot stress, notably the IIW recommendations [8],

    there is still the need for corresponding SNcurves. Pre-

    liminary proposals for weldable aluminium alloys havebeen made by Partanen and Niemi [42]on the basis of

    fatigue data generated at their university over a period

    of years. The data were obtained from a variety of butt-

    and fillet-welded specimens in 5000 and 6000 seriesalloys, including a model of the structural connection

    between the deck, longitudinal stiffeners and a transversebulkhead in a naval ship. The range of plate thicknesses

    was 36 mm. In all cases, hot-spot stresses were determ-ined from FEA or strain gauges using the procedures in

    the IIW recommendations. The results(Fig. 15)were in

    reasonable agreement and the authors proposed that the

    IIW FAT40 design curve for transverse butt welds,

    which is included in this figure, was suitable as a hot-spot stress SN curve for both butt and fillet weldedjoints in plate thicknesses up to 6 mm.

    This reference to thickness is important because the

    thickness effect discussed earlier will still exist even if

    fatigue strength is expressed in terms of the hot-spot

    stress. This is evident from another set of results for a

    wider range of thicknesses, 324 mm, also expressed interms of the hot-spot stress range [25]. The test speci-

    mens were all 6061-T6 aluminium alloy plates withtransverse fillet welded attachments. The results areshown in Fig. 16. Since the lowest SN curve, for 24

    mm thick specimens, happens to correspond exactly with

    the FAT40 design curve, it is tempting to conclude that

    these data support Partanen and Niemis proposal. How-ever, the results clearly show a thickness effect that justi-

    fies different hot-spot stress SN curves for different

    Fig. 15. Fatigue data presented by Niemi and Partanen[44]as a basis

    for the hot-spot stress SNcurve for thickness up to 6 mm.

    Fig. 16. Fatigue test results obtained from 6061-T6 beams withfillet-

    welded attachments expressed in terms of the hot-spot stress range

    which illustrate a thickness effect [25].

    thicknesses. A much larger database is needed to estab-

    lish such SNcurves.

    Other variables that might need to be considered

    further are the influence of the joint type and thethrough-thickness stress gradient. Referring to the data

    from Partanen and Niemi, there is a tendency for the

    higher results to be for butt welds and the lower ones

    for fillet welds, suggesting that there may be a need fordifferent hot-spot stress SNcurves for the two types of

    joint. A notable exception is the single lap joint that gave

    the highest results. This probably reflects the influenceof bending stress gradient, which would have been parti-

    cularly high in these joints due to their inherent mis-

    alignment. The effect, which is particularly significant

    in thin sections, is to increase fatigue resistance. Thus,this is another thickness effect that needs to be con-sidered when selecting data from which hot-spot stress

    design curves could be deduced. In general, data for high

    stress gradients should be excluded unless the hot-spot

    stress curve will only be used for similar conditions.

    Finally, inorder to assess the DNV hot-spot stressS

    Ncurves[6] mentioned earlier, they are included inFig.

    15.As will be seen, both curves look reasonable. How-

    ever, it should be mentioned that many of the results in

    Fig. 16, virtually all those for 24 mm thick specimens,lie below curve III, suggesting that the thickness effect

    correction should be introduced at a lower value than

    that currently specified (i.e. 25 mm) in the DNV note.

    4.10. Effect of marine environment

    Marine corrosion fatigue (full immersion and a saline

    atmosphere) of weldedaluminium alloys was studied in

    some depth in Norway [28]in the early 1980s. Fatigue

    tests were performed on transverse butt- and fillet-welded joints in 812 mm thick 5052, 5083, 6351 and7004 aluminium alloys. The tests on specimens in saline

    atmosphere or immersed in seawater were carried out at

    the low frequency of 1 Hz to allow time for the corrosion

  • 7/26/2019 [Elsevier] Review of Fatigue Assessment Procedures for Welded Aluminium Structures

    16/20

    1374 S.J. Maddox / International Journal of Fatigue 25 (2003) 13591378

    reaction to take place. The tests were conducted in bend-

    ing which meant that relatively high fatigue lives were

    obtained, as compared with those expected for axial

    loading. Based on comparison of the fatigue perform-

    ance in air and seawater, the results were consistent withearlier studies by Sanders and McDowell [43]on 5000

    series alloys. However, the effect of the environmentvaried with alloy type, 5052 and 7004 alloys being more

    susceptible to environment than the others. In general,

    immersion in seawater produced fatigue lives approxi-mately one-third of those obtained in air, corresponding

    to a 25% reduction in fatigue strength. A saline atmos-

    phere was generally less harmful, but not always. It pro-

    duced a similar reduction in fatigue life to full immersion

    in seawater in the 5052 alloy, while it produced an order

    of magnitude reduction in fatigue life in the case of buttwelds in 7004 alloy. The effect was less severe in filletwelds. The influences of environment and alloy typeseen in this study are reflected in Eurocode 9 (seeFig. 6).

    Fatigue crack growth studies in Russian AlMg5 and

    AlZnMg alloys immersed in 3% sodium chloride sol-

    ution[44]showed rather similar effects of environment

    for both alloy types. Crack growth rate was increased,

    but only significantly, by up to seven times, at relativelyhigh crack growth rates, with little effect of environmentnear the threshold. The threshold itself was effectively

    independent of environment. These results suggest that

    the effect of environment on SNdata referred to earlier

    may have been largely associated with crack initiation,

    which might also explain why butt welded 7004 alloy

    was more susceptible to environment than fillet welds.

    5. Friction-stir welding

    All the data presented so far were obtained from arc

    welded specimens. A new welding process that offers

    considerably better fatigue performance is friction-stir.

    Friction-stir welding (FSW) was invented at TWI, and

    the first patent application was filed in December 1991.The process is an entirely new method of making con-

    tinuous welds in several configurations using a solid-state process. The concept of FSW is illustrated in

    Fig. 17(a). This shows a rotating tool that consists of a

    shoulder and a pin. The former is pressed against the

    surface of the materials being welded, while the pin is

    forced between the two components by a downward

    force. The rotation of the tool under this force generatesfrictional heat which softens the work-piece, and the

    movement of the rotating tool along the joint line causes

    softened material to flow from the region ahead of thetool to the region behind, consolidating to form a solid

    phase weld. The process uses no filler, and for mostmaterials a shielding gas is not required. As the process

    does not melt the materials being joined, materials such

    as series 2000 and 7000 aluminium alloy, which are

    Fig. 17. Friction-stir welding: (a) FSW process; (b) FSW joint in alu-

    minium sheet.

    often difficult to weld by fusion processes due to solidi-fication problems, are readily weldable. Experience hasshown that as the process is fully mechanised, high lev-

    els of consistency can be obtained in weld quality. Inaluminium, it is possible to make full penetration singlepass butt welds in thicknesses of less than 1 mm to over

    50 mm.

    The process is used commercially by an ever-growing

    list of companies in the aerospace, shipbuilding, railway

    and automotive sectors. Almost all of the current com-

    mercial usage involves aluminium alloys, although some

    copper and magnesium alloys are also being welded. The

    joining of other materials, including titanium alloys,

    steel and nickel alloys, is under development.FSW of aluminium alloys produces joints of high

    quality with static mechanical properties that equal, or

    generally exceed, those of competing processes, butwith

    lower scatter. An example of a weld is shown in Fig.

    17(b).In view of the favourable profile, it is not surpris-ing to find that, under similar conditions, the fatigueproperties of friction-stir welds in aluminium alloys

    compare very favourably with those for welds made by

    MIG, the normal alternative. There are several examples

    in the literature, all relating to 6000 alloys (since 2000

    and 7000 alloys cannot be welded easily by the MIGprocess), although they are mainly confined to tests onrelatively thin specimens[4547].A typical example, for5 mm thick 6082 alloy tested under the relatively severe

  • 7/26/2019 [Elsevier] Review of Fatigue Assessment Procedures for Welded Aluminium Structures

    17/20

    1375S.J. Maddox / International Journal of Fatigue 25 (2003) 13591378

    loading condition of R = 0.5[45] is shown in Fig. 18.

    In the same investigation FSW joints in 6005 alloy gave

    fatigue test results close to those obtained from the

    unwelded material, which is deliberately low to allow

    for possible weld root flaws.In principle, it is possible to make any weld design

    which does not require the addition of a filler materialby FSW, although most experience to-date relates to

    butt- and lap-joints. A particular benefit is that full pen-etration butt welds are readily achieved in joints madefrom one side only, whereas such joints are highly vul-

    nerableto root flaws in welds made by other processes.Fig. 18 includes the current Eurocode 9 design curves

    for transverse butt welds made from one or two sides,

    for comparison with the test data. It will be clear that

    friction-stir welds made from one side can achieve con-siderably better fatigue lives than those indicated by the

    design curve. Root flaws can still arise in FSWs, butthey can be relatively large before they affect the fatigue

    performance of the joint [48]. Unfortunately, FSW can-

    not be used to make fillet welds, as no filler is used.Therefore comparative fatigue data for fillet welds donot exist.

    6. Fatigue life improvement methods

    Fatigue life improvement techniques play an

    important part in achieving higher design stresses when

    the fatigue lives of structures are restricted by the pres-

    ence of low fatigue strength details like fillet-welded

    attachments. They may also be needed to ensure that aweld repair of fatigue damage will survive longer thanthe original detail. Thus, they are relevant to both the

    original design and life extension of existing structure.

    There are two main principles behind the various

    improvement techniques [21]:

    (a) Reduction of the stress concentration due to weld

    Fig. 18. Comparison offatigue data for 5 mm thick 6082 alloy butt

    welded by MIG or FSW[45].

    geometry, e.g. dressing by machining, grinding or

    TIG remelting.

    (b) Introduction of compressive residual stresses, e.g.

    peening (hammer, needle, shot and brush), ultrasonic

    impact treatment.

    In general, both types of improvement technique areonly readily applicable to surface stress concentrations,

    notably weld toes. Improvement techniques have been

    widely studied in the context of welded steel, but less

    so in aluminium alloys. However, the general principles

    should be applicable to aluminium, as confirmed in arecent review by Hobbacher[49]. However, the reviewshowed that most published data referred to butt welds,

    whereas in practicefillet welds present the greater poten-tial fatigue problem. It was concluded that a fatigue

    strength improvement at 2 106 cycles of around 1.4 or

    more was justified for all the joint types reviewed, lead-

    ing to the recommendations summarised in the follow-ing table:

    Structural detail Treatment Fatigue strength

    method improvement

    factor

    Transverse butt Laser dressing 1.4welds TIG dressing

    Brush peening

    Shot blasting

    Cruciform joint Hammerfillet welds peening

    Longitudinal Grindingfillet-welded Hammerstiffener peening

    Transverse fillet- Brush peeningwelded stiffener

    A practical problem with aluminium that can arise in

    the case of dressing techniques concerns porosity. Flush-

    grinding of butt welds or TIG dressing of butt or filletwelds can result in previously embedded pores being

    exposed.TIG dressing can actually cause surface-break-

    ing pores[50].They then act as crack initiation sites andcan actually reduce the fatigue life of the weld detail.

    Although some design codes refer to the use of

    improvement techniques, none provides recommen-

    dations on the improvement in fatigue life to be achi-

    eved. This partly reflects uncertainty about the correctapplication of the techniques and their effect on the

    fatigue performance of real structures. In relation to the

    last point, it is known that the fatigue performance of

  • 7/26/2019 [Elsevier] Review of Fatigue Assessment Procedures for Welded Aluminium Structures

    18/20

    1376 S.J. Maddox / International Journal of Fatigue 25 (2003) 13591378

    welds treated by the residual stress techniques are sig-

    nificantly affected by mean stress, any benefit disap-pearing at stresses approaching yield[51].Thus the tech-

    niques may not be suitable for large structures containing

    high tensile residual stresses or there may be doubtsabout their benefits in situations where the mean stress

    is not known. The IIW is currently addressing both theprovision of specifications for the application ofimprovement techniques and corresponding benefits interms of revised SNcurves[52], including preliminaryrecommendations for welded aluminium alloys.

    7. Fracture mechanics assessment of fatigue

    Fracture mechanics offers the ability to assess thefatigue performance of aluminium structures containing

    known or assumed flaws. In the context of the originalstructure, these could be manufacturing flaws, when theassessment might be required if the flaw exceeds themanufacturing quality standard being worked to. In the

    context of the assessment of existing structures, they are

    likely to be cracks formed during previous service by

    fatigue or other mechanisms.

    Eurocode 9 contains an Appendix with guidance onthe use of fracture mechanics for assessing welded alu-

    minium alloys. This includes recommended fatigue

    crack growth laws based on a large database produced

    by Alusuisse [22]. These are expressed as a series of

    Paris laws, to model the data more accurately than a

    single Paris law, as illustrated inFig. 19.The same data

    are referred to in BS 7910 [16] and the correspondingIIW recommendations [53]. These also contain verydetailed guidance on the use of fracture mechanics for

    assessing welded structures, including a comprehensive

    Fig. 19. Example of multi-stage fatigue crack growth relationships

    for aluminium alloys proposed by Jaccard[22].

    set of stress intensity factor solutions for the types of

    crack and welded joint geometry which are likely to be

    encountered. One advantage of the procedure in the IIW

    recommendations is that it is linked with the IIW fatigue

    design curves for welded aluminium alloys to facilitatedirect comparison of the fatigue performance of flaws

    and that of basic design details in the same structure.However, BS 7910, which relates the fracture mechanics

    assessment to British Standard design SN curves, is

    more up-to-date.The multi-stage Paris law crack growth data fromEur-

    ocode 9 have also been presented as polynomials [54],

    in order to deduce better estimates of the probability of

    failure associated with upper-bound curves. However,

    regardless of the method of presentation, little has been

    done to confirm that the same complex, multi-stagecrack growth relationship is a general law, applicable to

    any cracks in real structures.

    8. Future research

    A number of aspects of both the design specificationsand residual life assessment methods considered in this

    review would be improved by further research. The fol-

    lowing are suggested as being the most important:

    (a) Provision of fatigue data for non-arc welding pro-

    cesses, particularly friction-stir but also laser weld-

    ing, and their incorporation in design specifications.

    (b) Study of cumulative damage under realistic stressspectra, with particular emphasis on the high-cycleregime and the damaging effect of stresses below

    the constant amplitude fatigue limit.

    (c) Further fatigue tests and FEA of structural details to

    establish hot-spot stressSNcurves and guidance on

    the practical application of the approach.

    (d) Identification of potential fatigue design improve-ments that could be achieved by better use of special

    extrusions, and generation of appropriate fatigue

    data. Data are also required for transverse butt weldsmade on the backing provided by an extruded lip.

    (e) Establishment of specifications for applyingimprovement techniques (of particular relevance for

    life extension) to welded aluminium and experi-

    mental confirmation of their value under realisticloading conditions (e.g. mean stress, loadingspectrum).

    (f) As far as the use of fracture mechanics for estimating

    residual fatigue life is concerned, the information

    incorporated in BS 7910 probably represents the cur-

    rent state of the art. However, it does place particularemphasis on steel and experimental work to decide

    on the choice of fatigue crack growth relationships

    appropriate for aluminium alloys and experimental

  • 7/26/2019 [Elsevier] Review of Fatigue Assessment Procedures for Welded Aluminium Structures

    19/20

    1377S.J. Maddox / International Journal of Fatigue 25 (2003) 13591378

    validation of the fracture mechanics approach in gen-

    eral would be useful.

    9. Conclusions

    Based on a review of published information on fatigueassessment procedures for welded aluminium structures

    and supporting experimental data, the following con-

    clusions can be drawn:

    (a) Of the three fatigue design assessment procedures

    described, that using the nominal stress SNcurves

    is the most developed and standardised, but the hot-

    spot stress approach will probably prove to be the

    most valuable for structural design in future.(b) Several national and international fatigue design

    specifications have been published in recent years.Eurocode 9 and the IIW recommendations are the

    most comprehensive.

    (c) Even so, most design data refer to arc welded joints

    and there is a need for corresponding data for other

    welding processes (e.g. friction-stir).

    (d) There are significant differences between all the pro-posed design SN curves and the fatigue test data-base available for large-scale structural specimens,

    mainly due to the choice ofSN curve slope. Thus,

    some specifications are unduly conservative in thelow endurance regime and others in the high-cycle

    regime. Eurocode 9 curves were generally the most

    consistent with experimental data.

    (e) There is a scarcity of data for weld details incorpor-ating special extrusion shapes, (e.g. built in backinglip) and little effort seems to have been made to

    investigate improved fatigue performance of welded

    aluminium structures by better use of extrusions.

    (f) The fatigue performance of commonly welded alu-

    minium alloys is not greatly influenced by immersionin a marine environment. The recommendations in

    Eurocode 9 are consistent with relevant published

    data.

    (g) There is little information in the literature on remain-ing life assessment procedures. The most appropriate

    are the use of design SN curves or, if allowance

    must be made for damage sustained during previous

    service, fracture mechanics. Comprehensive guid-

    ance and recommended input data for the application

    of fracture mechanics are contained in BS7910:1999.

    Acknowledgements

    This work described in this paper was funded partly

    by the Australian Maritime Engineering CRC Ltd. and

    partly by Industrial Members of TWI Ltd. The author is

    also grateful to his colleague Dr P.L.Threadgill for his

    assistance with the section on friction-stir welding.

    References

    [1] BS 8118:1991. Structural use of aluminiumpart 1 code of prac-tice for design. London: BSI, 1991.

    [2] Eurocode 9. Design of aluminium structures: part 2: structures

    susceptible to fatigue. Brussels: CEN, 1998 ENV, 1999-2.

    [3] International Institute of Welding. Fatigue design of welded joints

    and components. Abington, Cambridge: Abington Publishing;

    1996.

    [4] The Aluminum Association. Specifications for aluminium struc-

    tures. Washington, DC: The Aluminum Association; 1994.

    [5] Canadian Standards Association CAN/CSA-S157-M92. Strength

    design in aluminium. Canada: CSA; 1993.

    [6] DNV. Class note: fatigue assessment of aluminium structures.

    Technical Report No. LIB-J-000010; 1995.

    [7] ECCS. European recommendations for aluminium alloy struc-

    tures. In: European convention for constructional steelwork.

    Document No. 68. Brussels: ECCS; 1992.[8] Niemi E. Stress determination for fatigue analysis of welded

    components. Abington, Cambridge: International Institute of

    Welding, Abington Publishing; 1995.

    [9] Lloyds Register of Shipping. Stress concentration factors for sim-

    ple tubular joints, assessment of existing and development of new

    parametric formulae. HSE Report No. 0TH91 354, 1991.

    [10] Radaj D. Design and analysis of fatigue resistant welded struc-

    tures. Abington, Cambridge: Abington Publishing; 1990.

    [11] Fatigue strength of welded ship structures. Bureau Veritas Docu-

    ment NI 393 DSM ROIE; July 1998.

    [12] IABSE Workshop, Lausanne 1990. Remaining fatigue life of

    steel structures. IABSE Report, vol. 59. Zurich:IABSE;1990.

    [13] Hadley I, Manteghi S. Remnant life of semi-submersible rigs. In:

    Proceedings of the Seminar on Remnant Life Prediction, Nov-

    ember. London: IMechE; 1997.[14] Maddox SJ, Kenzie BW. Fatigue assessment of ageing ship struc-

    tures. In: Proceedings of the 11th International Maritime and

    Shipping Conference on Ships, the Ageing Process. London:

    Institute of Marine Engineers; 1997.

    [15] Moses F. Bridge load models for fatigue. In: Proceedings of the

    IABSE Workshop on Remaining Fatigue Life of Steel Structures.

    IABSE Report, vol. 59. Zurich: IABSE; 1990.

    [16] BS 7910:1999. Guide on methods for assessing the acceptability

    offlaws in metallic structures. London: BSI; 1999.

    [17] Maddox SJ. Fatigue design of welded aluminium alloy structures.

    In: Proceedings of the Second International Conference on Alu-

    minium Weldments. Dusseldorf: Aluminium-Verlag; 1982.

    [18] Maddox SJ, Webber D. Fatigue crack propagation in AlZnMg

    alloyfillet welded joints. In: Fatigue testing of weldments. Phila-

    delphia, PA: ASTM; 1977 [ASTM STP648].[19] Menzemer C, Fisher J. Revisions to the Aluminum Association

    fatigue design specifications. In: Proceedings of the Sixth Inter-

    national Conference on Aluminium Weldments, Cleveland, OH,

    35 April 1995. Miami, FL: American Welding Society; 1995

    [ISBN 0-87171-458-2].

    [20] Dawes CJ, Thomas WM. Friction stir process welds aluminium

    alloys. Welding Journal 1986;75(3):415.

    [21] Gurney TR. Fatigue of welded structures, 2nd ed. Cambridge:

    Cambridge University Press; 1978.

    [22] Jaccard R. Fatigue crack propagation in aluminium. IIW

    Doc.XIII-1377-90; 1990.

    [23] Gurney TR. The influence of thickness on the fatigue strength of

    welded joints. In: Proceedings of the Second International Con-

    ference on Behaviour of Offshore Structures, London. 1979.

  • 7/26/2019 [Elsevier] Review of Fatigue Assessment Procedures for Welded Aluminium Structures

    20/20

    1378 S.J. Maddox / International Journal of Fatigue 25 (2003) 13591378

    [24] Maddox SJ. The effect of plate thickness on the fatigue strength

    of fillet welded joints. Abington, Cambridge: Abington Pub-

    lishing; 1987.

    [25] Maddox SJ. Scale effect in fatigue of fillet welded aluminium

    alloys. In: Proceedings of the Sixth International Conference on

    Aluminium Weldments, Cleveland, OH, 35 April 1995. Miami,

    FL: American Welding Society; 1995 [ISBN 0-87171-458-2].

    [26] Gurney TR, Maddox SJ. An alternative to Miners rule for cumu-lative damage calculations? In: Proceedings IABSE Workshop

    Remaining Fatigue Life of Steel Structures, vol. 59. Zurich:

    IABSE; 1990.

    [27] Niemi E. Random loading behaviour of welded components. In:

    Proceedings of the IIW International Conference on Performance

    of Dynamically Loaded Welded Structures. New York: Welding

    Research Council; 1997.

    [28] Engh B, Solli O, Arbo S, Paauw AJ. The influence of a corrosive

    environment on the fatigue life of aluminium weldments. In: Pro-

    ceedings of the Third International Conference on Aluminium

    Weldments. Dussedorf: Aluminium-Verlag; 1985.

    [29] Ogle MH, Maddox SJ. Design rules for fatigue of aluminium

    structures according to BS8118 Part 1. In: Proceedings of the

    Fifth International Conference on Aluminium Weldments. Alu-

    minium-Zentrale/Technical University of Munich; 1992.[30] Jaccard R, Kosteas D, Ondra R. Background document to fatigue

    design curves for welded aluminium components. IIW Doc.

    No.XIII-1588-95; 1995.

    [31] Hobbacher A. The development of the new IIW fatigue rec-

    ommendations. In: Proceedings of the IIW International Confer-

    ence on Performance of Dynamically Loaded Welded Structures.

    New York: Welding Research Council; 1997.

    [32] Mazzolani F, Grillo M. Fatigue strength of longitudinally welded

    aluminium alloy structures. In: Proceedings of the Sixth Inter-

    national Conference on Aluminium Weldments, Cleveland, OH,

    35 April 1995. Miami, FL: American Welding Society; 1995

    [ISBN 0-87171-458-2].

    [33] Orjaster O. Fatigue of large weldments in AlMgSi1 and

    AlMg4.5Mn aluminium. In: Kosteas D, Ondra R, Ostermann F,

    editors. Proceedings, 5th INALCO Conference on AluminiumWeldments, Munich, 2729 April 1992. Munchen, Germany:

    Technische Universitat Munchen; 1992.

    [34] Soetens, F.,van Straalen, I.J., Dijkstra, O., European research on

    fatigue of aluminium structures. In: Proceedings of the Sixth

    International Conference on Aluminium Weldments, Cleveland,

    OH, 35 April 1995. Miami, FL: American Welding Society;

    1995 [ISBN 0-87171-458-2].

    [35] Haagensen PJ, Raines M, Kluken AO, Kvale I. Fatigue perform-

    ance of welded aluminium deck structures. In: Proceedings of the

    15th International Offshore Mechanics and Arctic Engineering

    (OMAE) Conference, Materials Engineering, vol. 3. New York:

    ASME; 1996.

    [36] Beach JE, Johnson RE, Koehler FS. Fatigue of 5086 aluminium

    weldments. In: Proceedings of the Second International Confer-

    ence on Aluminium Weldments. Dusseldorf: Aluminium-Ver-lag; 1982.

    [37] Menzemer CC. Fatigue behaviour of welded aluminium struc-

    tures. PhD thesis, Lehigh University, Bethlehem, PA, 1992.

    [38] Webber, D., Gurney, T.R., The effect of some programme vari-

    ables on fillet weld cumulative fatigue damaged test results. In:

    Kosteas D, Ondra R, Ostermann F, editors. Proceedings, 5th

    INALCO Conference on Aluminium Weldments, Munich, 2729

    April 1992. Munchen, Germany: Technische Universitat

    Munchen; 1992.

    [39] Hirt, M.A., Kimberley, G.J., Smith, I.F.C., Fatigue behaviour of

    aluminium beams with welded attachments. In: Kosteas D, Ondra

    R, Ostermann F, editors. Proceedings, 5th INALCO Conference

    on Aluminium Weldments, Munich, 2729 April 1992. Munchen,

    Germany: Technische Universitat Munchen; 1992.

    [40] James MN, Lambrecht HO, Patterson AE. Fatigue strength of

    welded cover plates on 6261 aluminium alloy I-beams. Inter-

    national Journal of Fatigue 1993;15(6):51924.[41] James MN, Patterson AE, Sutcliffe N. Constant and variable

    amplitude loading of 6261 aluminium alloy Ibeams with

    welded