Sunteți pe pagina 1din 17

OPT 2008

Amsterdam, The Netherlands, 27-28 February 2008

ECAs: Are they fit-for-purpose?


ANDREW COSHAM

Atkins Boreas, UK

ABSTRACT
Engineering critical assessments (ECAs) have increasingly become a routine part of pipeline
design to determine tolerable flaw sizes for weld defects. These assessments are now being
applied to pipeline systems in deeper water with increased loadings arising from responses
to thermal and pressure cycling. Often these are flowline systems in which fatigue damage
is exacerbated by the presence of aggressive internal conditions.
In these situations, ECAs can give 'alarming' results, indicating that only very small flaws
would be acceptable. In some cases, applying the same methodology to in-service pipelines
would suggest that the pipeline should have failed a long time ago, whereas in reality they
have not.
Therefore, a number of questions arise:
 are ECAs too conservative;
 are there situations where ECAs may be non-conservative; and
 do we fully understand what we are doing?
In this paper, these issues are illustrated by means of several examples and an attempt is
made to partly answer the above questions.

1 INTRODUCTION
Pipeline systems are being designed to operate in deep waters at high temperatures and
high pressures, in aggressive internal environments. Design issues for such pipeline
systems tend to arise in the flowlines rather than the risers or ‘platforms’. Fatigue can be a
significant design constraint.
The fatigue design of risers is typically governed by ‘hot spots’ at the top and bottom of the
riser due to loads arising from wind, wave and current loading. Wave loading is typically
7 8
predictable and of a low frequency, with of the order of 10 or 10 cycles over the design life
of the system. Current loading is more unpredictable. Factors of safety of five for flowlines
and ten for risers are typically applied in fatigue design.
Deep water flowlines operating at high temperatures and pressures need to be designed to
accommodate issues such as significant end expansion, walking and lateral buckling.
Insulation or direct electrical heating may be required for flow assurance. Shut-downs result
in significant pressure and temperature cycles. The fatigue loading is characterised by a
small number of large cycles, less than 10 3 cycles over the design life. The fatigue loading is
in a completely different regime to that in a riser , giving rise to a different set of design
challenges. The fatigue loading is also, in principle, under the direct control of the pipeline
operator because it is driven by variations in pressure and temperature, unlike environmental
loading. Factors of safety in fatigue design, and the associated ECAs, tend to be lower, in
part because there is perceived to be a higher level of confidence in the fatigue loading, but
also because higher factors of safety cannot be accommodated. The fatigue design is


Atkins Boreas, Churchill House, 12 Mosley Street, Newcastle upon Tyne, NE1 1DE, UK.
Tel: +44 (0)191 230 8098, Fax: +44 (0)191 261 0200, e-mail: andrew.cosham@atkinsglobal.com

1/17
complicated by issues such as: corrosion-fatigue, frequency and strain rate, in addition to the
effect of the internal environment on toughness.
Fatigue design is normally based on S-N curves (e.g. the class D and E design curves). The
ECA is a secondary consideration. S-N curves are only appropriate if the weld is free from
significant defects. An ECA is conducted to determine the tolerable flaw size, i.e. the size of
a significant defect. With the application of ECAs to pipeline systems in deeper water subject
to high thermal and pressure cycling and aggressive internal environments, it is important to
understand if ECAs can be overly conservative, or perhaps even non-conservative It is
therefore informative to examine a number of the issues surrounding ECAs, and their
relationship with S-N curves, and pipeline welding codes and standards.
Although spoken in a different context, the words of former secretary of defense are relevant:
“As we know, there are known knowns; there are things we know we know. We also know
there are known unknowns; that is to say we know there are some things we do not know.
But there are also unknown unknowns -- the ones we don't know we don't know.” [1].

2 ECAs IN CODES AND STANDARDS


An engineering critical assessment (ECA) is a method for assessing the acceptability of a
flaw in a structure, i.e. to demonstrate fitness-for-purpose. Fitness-for-purpose is defined in
BS 7910 : 2005 [2] as follows: “By this principle, a particular fabrication is considered to be
adequate for its purpose, provided the conditions to cause failure are not reached.”
BS 7910 describes in detail how to conduct an engineering critical assessment. Other widely
recognised codes and standards include API 579 and R6 [3,4], although these are less
commonly used in the pipeline industry. These generic codes and standards are
supplemented by additional guidance in pipeline design codes and standards.
The introduction to BS 7910 makes a number of statements regarding ECAs that are both
informative and sometimes not remembered. “… a proliferation of flaws, even if shown to be
acceptable by an ECA, is regarded as indicating that quality is in need of improvement. The
use of an ECA can in no circumstances be viewed as an alternative to good workmanship. …
[ECAs] are complementary to, and not a replacement for, good quality workmanship.”
Appendix A of DNV-OS-F101 2007 [5] gives additional guidance on conducting ECAs of girth
welds in offshore pipelines, intended to supplement that given in BS 7910. DNV-RP-F108 [6]
give specific guidance for installation methods that introduce cyclic plastic strain, i.e. reeling.
The pipeline welding codes BS 4515 and API 1104 also give relevant guidance [7,8]. In
addition, although developed for onshore pipelines, the EPRG guidelines for defects in
transmission pipeline girth welds are informative [9]. It is often instructive to compare the
results of ECAs with pipeline specific guidance.
ECAs are based on the application of the science of fracture mechanics. The conventional,
applied mechanics, approach to design is to consider the loads acting on the structure and
the tensile properties (yield strength, tensile strength, etc.) of the material. This approach is
not valid if the structure contains stress concentrations (e.g. notches) or defects (e.g. cracks).
Fracture mechanics considers the effect of a defect. In addition to the loads and the tensile
properties, fracture mechanics considers the toughness of the material and the defect
geometry.
Fracture mechanics is a relatively mature discipline, but there are new developments,
through joint industry projects, project-specific studies and academic research, and these are
gradually being incorporated into codes and standards. Issues such as constraint, the ‘local-
approach’, high strains and bi-axial loading are addressed to varying degrees.

2/17
3 FATIGUE
Fatigue is a damage process whereby a crack can form and grow under the action of
fluctuating (cyclic) loads. The fatigue life of a welded joint is lower than that of a plain plate
because of the presence of stress concentrations and crack-like discontinuities. In a welded
joint, fatigue crack initiation may occupy only a very small proportion of the fatigue life; it is
dominated by fatigue crack propagation (growth) [10].
The environment has a significant effect on fatigue. The fatigue life in a corrosive
environment is lower than that in a non-corrosive environment. Corrosion-fatigue can be
more severe than either corrosion or fatigue. The rate of crack propagation is higher and the
endurance limit (or threshold for the initiation of fatigue crack growth) is lower or non-
existent.
A corrosive environment can be created simply by the presence of pre-existing cracks or
crevices. Capillary condensation may cause there to be a corrosive environment in a crack,
even though the bulk environment is non-corrosive. This issue can be of particular concern
in sour environments, where corrosion-fatigue rates can be very high.
Frequency is another issue that needs to be considered. A number of fatigue tests, both
published and unpublished, have shown that in a corrosive environment, such as sea-water
or a sour fluid, the fatigue life is reduced as the loading frequency is reduced. These tests
tend also to show that there is a plateau in the frequency response. This is fortunate,
because loads associated with, say, lateral buckling have a frequency of the order of 10 -6 Hz,
or lower, which is too low for testing to be practical.
Fatigue is an important consideration in an ECA.
The two methods used to assess fatigue are:
 S-N curves; and
 fracture mechanics.
S-N curves are derived from endurance testing. In fracture mechanics based fatigue, fatigue
crack growth laws, also derived from testing, are used. It is important that the endurance
and fatigue crack growth rates tests are conducted in conditions (e.g. material, geometry,
frequency, temperature and environment) that are representative of the actual conditions.

4 PIPELINE WELDING CODES AND STANDARDS


Pipeline welding codes and standards, e.g. BS 4515-1, API 1104 : 2005 and DNV-OS-F101,
specify workmanship acceptance levels for welding defects in pipeline girth welds. These
acceptance levels represent what a ‘good’ welder should be able to achieve. They are not
fitness-for-purpose defect limits. An ECA is required determine if workmanship acceptance
levels are fit-for-purpose (see section 12). High static loads and/or high cyclic loads may
necessitate smaller acceptance criteria.
Welding defects can be categorised as planar or non-planar (these are sometimes referred
to as linear and volumetric, respectively). Non-planar flaws include: slag, inclusions and
porosity. Planar flaws include: incomplete (inadequate) penetration, lack of (incomplete)
fusion, undercut, and cracks.
Planar flaws are more significant than non-planar flaws, although non-planar flaws can be
indicative of poor workmanship and may mask the presence of more severe planar flaws.
There are differences between the workmanship acceptance levels in the various pipeline
welding standards, but broadly speaking, the workmanship acceptance levels for planar
flaws can be summarised as follows:

3/17
 the length of individual surface flaws should not exceed 25 mm (1 in.), and the total
length of such flaws in any 300 mm (12 in.) should not exceed 25 mm; and
 the length of individual embedded (also referred to as buried) flaws should not exceed
50 mm (2 in.), and the total length of such flaws in any 300 mm (12 in.) should not
exceed 50 mm.
Workmanship acceptance levels were developed at a time when welding was manual and
the completed welds were inspected using radiography. The type of flaw can be identified,
and the length of the flaw can be measured using radiography, but not the height (depth).
Consequently, workmanship acceptance levels were originally expressed in terms of the type
of flaw and the length of the flaw. Semi-automatic and automatic welding systems have
been developed, and automatic ultrasonic (AUT) inspection has been introduced. The profile
of manual and automatic (or semi-automatic) welds is different, and some types of weld
defect are more, or less, common depending upon the welding method. AUT inspection is
more effective at identifying planar defects than radiography, and is capable of measuring
the height, as well as length, of flaws.
Inspection methods have improved over time. Consequently, the size and types of welding
defect that can be found has been extended as inspection methods have improved.
The workmanship acceptance levels in BS 4515-1 can only be applied if a minimum Charpy
V-notch (CVN) impact energy requirement is satisfied. The average CVN impact energy at
the minimum design temperature should be at least 40 J, and the minimum at least 30 J.
Welding codes and standards also make reference to defect acceptance limits (or defect
limits) based on fitness-for-purpose. Appendix A of API 1104 gives alternative acceptance
criteria for girth welds, based on fitness-for-purpose methods. Appendix A is applicable if:
the maximum applied axial strain does not exceed 0.5%, and the CTOD at the minimum
design temperature is at least 0.127 mm (0.005 in.). Appendix A is not applicable to a
pipeline subject to fatigue loading in excess of a prescribed limit; the “spectrum severity” limit
in §A.2.2.1 is equivalent to a usage factor of 0.013 with respect to the class E design curve
(see section 9).
BS 4515-1 also allows acceptance criteria to be based on fitness-for-purpose, and refers to
the guidance in BS 7910 for conducting ECAs.

5 THE EPRG GUIDELINES


The EPRG guidelines for the assessment of defects in transmission pipeline girth welds give
simple defect limits, based on extensive small and full-scale testing, and fracture mechanics
analysis [9]. The EPRG reviewed workmanship acceptance levels and fitness-for-purpose
defect limits in various pipeline welding codes and standards, noted the differences and
inconsistencies, and subsequently developed guidelines incorporating workmanship
acceptance levels (Tier 1) and defect limits (Tiers 2 and 3).
The EPRG guidelines are not applicable to a pipeline subject to “onerous fatigue duty”, but
they can, nevertheless, be informative (see section 10).
Tier 2 of the EPRG guidelines is applicable if: the maximum applied axial strain does not
exceed 0.5%, the weld overmatches the pipe body, the yield to tensile ratio does not exceed
0.90, and the average Charpy V-notch impact energy at the minimum design temperature,
from a set of three tests, is at least 40 J, and the minimum is at least 30 J. The CVN impact
energy requirements in BS 4515-1 are taken from the EPRG guidelines. BS 4515-1 draws
the readers attention to the publications of the EPRG.
The Tier 2 defect limits for surface and embedded planar flaws are: flaw height less than or
equal to 3 mm (based on the typical height of a weld run), and flaw length less than or equal
to 7 times the wall thickness.

4/17
The Tier 3 defect limits are larger than the Tier 2 limits, and there are additional
requirements. Tier 3 of the EPRG guidelines is applicable if: the maximum applied axial
stress does not exceed the yield strength (similar to the Tier 2 requirement, in that API 5L
defines the yield strength as the stress at a total strain of 0.5% [11]), and, in addition to the
CVN requirement, the average CTOD (measured using a single edge notch bend specimen)
at the minimum design temperature, from a set of three tests, must be at least 0.15 mm, and
the minimum is at least 0.10 mm. The Tier 3 defect limits are larger than the Tier 2 limits, so
it is reasonable (and conservative) to consider that 40/30 J is equivalent to 0.15/0.10 mm.
The EPRG guidelines are partly based on wide plate, and full scale, testing, which means
that the implications of constraint are considered implicitly. Bi-axial is also addressed, in that
the guidelines are applied to transmission pipelines operating at hoop stresses up to 72%
SMYS (but the axial strain is limited to 0.5%).

6 S-N CURVES
S-N curves can be used to estimate the fatigue life of a welded joint. An S -N curve presents
the fatigue life, N, as a function of the applied stress range, S. S-N curves of welded joints
are based on endurance tests of workmanship quality welds. Welded joints are classified,
and for each class there is a different S-N curve, e.g. in PD 5500, the class E curve is for a
full penetration butt weld made from both sides, and the class F2 curve is for a full
penetration butt weld made from one side, without backing [12]. The class of an S-N curve
refers to a particular mode of fatigue failure, e.g. initiation at the weld toe, or the weld root, so
more than one class may apply to a particular welded joint. The effect of stress
concentrations due to the weld shape and type are included in the S-N curves of welded
joints.
Design S-N curves, such as those given in BS 7608, PD 5500 and DNV-RP-C203 [13], are
mean minus two standard deviation curves to the experimental data.
In addition to classifying the welded joint, to select the appropriate S-N curve, the effect of
plate thickness, the environment, and gross structural discontinuities and deviations from the
intended design shape, need to be taken into account. Some codes also indicate an effect of
material, if other than a ferritic steel at ambient temperature.
The endurance test specimens used in the fatigue tests upon which the S-N curves will have
included some degree of misalignment, and weld defects, but there is rarely sufficient
information in the published data for these effects to be quantified [14]. Also, it is common
practice in analysing endurance test data to plot the local stress range, corrected for
misalignment (from strain gauge measurements), not the nominal stress range. There is a
view that the class E design S-N curve includes an allowance for misalignment,
corresponding to an SCF equal to 1.3, but this is not universally accepted in design codes
[10]. BS 7608 states that the classifications for transverse butt welds allow for some degree
of misalignment, but only if the root sides of joints with single-sided preparations are back-
gouged [15]. Macdonald et al. (2000), in a review of S-N curves for girth welds, concluded
that the class E S-N curve could be used for full penetration girth welds made from one side,
in conjunction with a thickness correction and SCFs [14]. The guidance in DNV-RP-C203 is
based on the Macdonald et al. (2000) review.
PD 5500 does not require a factor of safety (or usage factor) to be applied to the fatigue life
estimated using the design S-N curves. DNV-RP-C203 and DNV-OS-F101 specify a factor
of safety, which depends on the safety class. For a ‘normal’ safety class, a factor of safety of
five must be applied to the fatigue life estimated using the design S-N curves. For a ‘high’
safety class, a factor of ten is applied. The fatigue limits in IGE/TD/1, a design code for
onshore transmission pipelines, incorporate a factor of safety of ten on fatigue life [16]. PD
8010-1 and -2 indicate that factors of safety in pipeline design typically range from 1 for non-
hazardous, non-critical pipelines to 3 to 10 for hazardous pipelines [17,18]. It further states

5/17
that if full penetration high quality welds are assured, then the relevant S-N curves from BS
7608 can be used without a factor of safety.
There are a number of reasons for the different approach es, including redundancy, ease of
inspection and weld quality. PD 5500 specifies acceptance levels that are more severe than
the typical workmanship acceptance levels in pipeline welding codes.

7 S-N CURVES AND ‘SIGNIFICANT DEFECTS’


S-N curves are only appropriate if the weld is free from significant defects. The question is:
what is a significant defect? It is commonly interpreted as workmanship acceptance levels,
but some care is required because different welding codes specify different workmanship
acceptance levels.
The fatigue strength of a weld will be reduced by the presence of defects that are too small to
detect with commonly available inspection methods. The improvement of inspection
methods over time, therefore has implications for the definition and sizing of ‘significant
defects’; the size of defect that can be detected and measured using modern non-destructive
inspection methods may be smaller than could have been detected in the ‘workmanship’
quality welds used in the development of S-N curves.
BS 7910 notes that the S-N curves for the various joint classes were derived from test data
for nominally sound welds made under laboratory conditions, and further states that it is
almost certain that none of the test welds had flaws of normally detectable size at the weld
toe [2].
Macdonald et al. (2000) states that design S-N curves for girth welds do not cover fatigue
failure from identifiable welding defects [14]. Reference is made to several tests of girth
welds with root defects in which relatively low fatigue lives were recorded. It is stated that
surface crack-like flaws are unlikely to be acceptable if a fatigue strength represented by any
of the design S-N curves for girth welds is required. Embedded volumetric flaws and shallow
weld toe undercut may be acceptable, and reference is made to the acceptance levels for the
‘quality categories’ given in BS 7910 (which are the same as those in PD 5500). It is further
noted that these are known to be conservative [14].
BS 7608 : 1993 states that welding workmanship should be in accordance with BS 5135 :
1984 (now superseded) [15,19]. Quality category A of BS 5135 does not permit cracks, lack
of fusion, lack of penetration, i.e. planar defects are not permitted.
Annex C of PD 5500, a design code for unfired fusion welded pressure vessels, gives
requirements for the fatigue design of pressure vessels. The guidance in this annex is
informative for understanding the applicability of S-N curves. The S-N curves in PD 5500
have a common background to those in BS 7608, Macdonald et al. (2000), DNV-RP-C203,
and UK HSE Offshore Guidance Notes [20].
PD 5500 states that the design S-N curves have been derived from fatigue test data obtained
from welded specimens, fabricated to normal standards of workmanship. The material must
have sufficient toughness to avoid brittle fracture.
The welds must be proved to be free from significant defects by non-destructive testing. PD
5500 defines three construction categories, which differ with respect to the level of inspection
(non-destructive testing), permitted material, maximum nominal thickness, and temperature
limits [Table 3.4-1]. The requirement to ensure that welds are free from significant defects
necessitates full (100 percent) inspection, implying construction category 1. Acceptance
criteria for weld defects revealed by visual examination and non-destructive testing are given
in §5.7 of PD 5500. Summarising the radiographic inspection levels for category 1 [Table
5.7-1]: planar defects (e.g. cracks and lamellar tears, lack of root, side or inter-run fusion,
and lack of root penetration) are not permitted, and the limits for non-planar defects are

6/17
similar to BS 4515-1. The acceptance levels for non-planar defects are summarised in Table
1, below. The important point to note is that PD 5500 does not permit planar defects. In
addition to these acceptance criteria, the criteria in §C.3.4.2 of PD 5500 apply to vessels
assessed to Annex C, i.e. vessels subject to fatigue loading. Acceptance levels for non-
planar embedded defects (i.e. inclusions and porosity) are given [Table C.4], with respect to
the weld classes, and are summarised in Table 2, below. The only significant difference
between the acceptance criteria in §5.7 and §C.3.4.2 are that the latter are more restrictive
on the length of slag inclusions; the criteria are compared in Table 1 and Table 2, below.
The weld defect acceptance criteria in §C.3.4.2 of PD 5500 correspond to the limits for non-
planar flaws for the ‘quality categories’ given in BS 7910. The ‘quality categories’ provide an
alternative method of assessing the fatigue life of a flaw in a weld, compared to a fracture
mechanics based fatigue calculation.
PD 5500 refers to BS 7910 for assessing the fatigue lives of defects, or determining the
tolerable defects for a given fatigue life.
The common thread in the above codes and standards is that ‘free from significant defects’
means no planar defects. This is significant when it is noted that BS 4515-1 does permit
planar defect.

code maximum length of slag maximum percentage area of


inclusion ,mm porosity on radiograph

PD 5500 [Table 5.7-1] [parent metal thickness] 2


BS 4515-1 50 2
API 1104 50 [the limit is expressed in a
different form that is not
readily translatable]
Table 1 Acceptance levels for non-planar defects, given in PD 5500, for pressure
vessels not subject to fatigue loading

class required maximum length of slag maximum percentage area


inclusion ,mm of porosity on radiograph

D (Q1) 2.5 3
E (Q2) 4 3
F (Q3) 10 5
F2 (Q4) 35 5
G (Q5) no limit 5
W (Q6) and lower no limit 5
NOTE 1 Tungsten inclusions in aluminium alloy welds do not affect fatigue behaviour and need not
be considered as defects from the fatigue viewpoint.
NOTE 2 For assessing porosity, the area of radiograph used should be the length of the weld
affected by porosity multiplied by the maximum width of weld.
NOTE 3 Individual pores are limited to a diameter of e/4 or 6 mm, whichever is the lesser.
NOTE 4 The above levels can be relaxed in the case of steel welds which have been thermally
stress relieved, as described in BS 7910.

Table 2 Acceptance levels for non-planar embedded defects, given in PD 5500


(and BS 7910), for pressure vessels subject to fatigue loading

7/17
8 WHAT IS AN ECA?
An ECA, in general terms, considers all of the modes of final failure of a flaw (e.g. fracture or
plastic collapse) and all of the possible material damage mechanisms that may lead to the
growth of a sub-critical flaw (e.g. fatigue) or deterioration in the material properties (e.g.
embrittlement in a hydrogen charged environment).
The maximum operating pressure of the typical pipeline is well below the creep regime.
Environmental crack mechanisms, such as stress corrosion cracking, tend to exhibit high
rates of crack propagation, so normal practice is to avoid the initiation of such mechanisms.
Therefore, an ECA in a typical pipeline will be concerned with two issues: failure by fracture
or plastic collapse (hereafter, referred to simply as fracture), and the growth of sub-critical
flaws by fatigue. The ECA is therefore concerned with static loads and cyclic loads. The
growth of sub-critical flaws by ductile tearing may be an issue if the static loads are high (e.g.
reeling). Corrosion-fatigue, embrittlement, and the initiation of environmental crack
mechanisms may be of concern in some environments. In simple terms, however, the
former is addressed by an increase in the rate of fatigue crack growth, whilst the latter two
are addressed by reducing the fracture toughness.
The steps in an ECA are summarised in Figure 1. The simplest case (and perhaps the case
that was originally envisaged when the concept of ECAs and fitness-for-purpose were first
developed) is the ECA of a known flaw, see Figure 1 a). A structure (such as a pipeline)
contains a known flaw (), detected by means of some inspection technique (e.g.
radiography, ultrasonics, or an intelligent pig). The limiting flaw size () in the structure is
then calculated, based upon the failure mode(s), applied loads and the material properties.
Whether the known flaw in the structure is ‘fit-for-purpose’ depends on the difference
between the known flaw size and the limiting flaw size (), taking into account material
damage mechanisms and the capabilities of the inspection technique. This may include
calculation of the remaining life of the known flaw, and even re-inspection at some regular
interval into the future.
It is a natural extension of this simple case to consider an ECA to determine a flaw
acceptance criteria, see Figure 1 b). The limiting flaw size () in the structure is calculated,
based upon the failure mode(s), applied loads and the material properties. In a design case,
this would be the flaw size at the end of the design life. The relevant material damage
mechanisms over the time period under consideration are identified, and their effect on the
limiting flaw size is taken into account (). In a design case, the time period would be the
design life of the structure. The result is a calculated flaw size (); if the structure contains a
flaw that is greater than or equal to this calculated flaw size, then the structure will fail before
the end of the time period under consideration. In a design case, the maximum tolerable
flaw size is equal to this calculated flaw size. Then the flaw acceptance criteria () is
determined, with reference to the tolerable flaw size, the capabilities of the inspection
technique(s), workmanship considerations, and other factors. One approach would simply
be to subtract the inspection tolerances from the tolerable flaw size. Another approach,
assuming that workmanship acceptance levels are less than the subtracting the inspection
tolerances from the tolerable flaw size, would be to apply workmanship acceptance levels,
with the results from the ECA used for concessions.
Figure 1 b) is a relatively simple illustration of the steps required to determine a flaw
acceptance criteria. In practice, it can be significantly more complicated, as illustrated in
Figure 1 c). Consider an offshore pipeline that is designed to accommodate lateral buckling.
Firstly, it may be necessary to consider both installation and operation. Secondly, it may not
be immediately obvious as to what is the limiting condition, necessitating a number of
different calculations. The limiting flaw size at the end-of-life (e-o-l) is calculated, based on
the end-of-life loads (). The corresponding flaw size at the start-of-life is then calculated by

8/17
2

limiting flaw size


3
1 ?

known flaw

flaw size (a & 2c)


a) ECA of a known flaw

4
1
acceptance criteria

limiting flaw size


2
? 3

calculated flaw

flaw size (a & 2c)


b) ‘simple’ ECA to determine a flaw acceptance criteria
flaw size after
limiting flaw size

installation

flaw size (a & 2c)


INSTALLATION OPERATION
limiting flaw size, e-o-l
s-o-l

8
3 1
acceptance criteria

5 2
? 7 4

calculated flaw

flaw size (a & 2c)


c) ‘complex’ ECA to determine a flaw acceptance criteria (e.g. offshore pipeline)
Figure 1 The steps in an ECA

9/17
taking into account the relevant material damage mechanisms over the design life (). The
structure may experience higher loads at the start-of-life than at the end-of-life, e.g. the
stresses and strains in a lateral buckle tend to be highest when the buckle first forms. The
limiting flaw size at the start-of-life (s-o-l) is calculated, based on the start-of-life loads ().
The tolerable flaw size, with respect to operation, at the start -of-life () is the lower of that
determined from  & , and . It is then necessary to consider the effect of the relevant
material damage mechanisms during installation (). In simple terms, this might only be
fatigue loading during installation. In general terms, there may be a requirement for a further
ECA to determine the tolerable flaw size for installation () and the flaw size after
installation, taking into account factors such as fatigue and ductile tearing (and even the
effects of installation on the material properties). The tolerable flaw size, with respect to
installation, at the start-of-life () is the lower of that determined from  & , and . Then
the flaw acceptance criteria () is determined, as above.

9 ECAs AND S-N CURVES


Fatigue design is normally based on S-N curves. The ECA is a secondary consideration.
A pipeline girth weld will be a class D or E weld (cap and root, respectively), with respect to
PD 5500 (or similar curves in DNV-RP-C203). A factor of safety of say five (or ten) would be
applied to this design curve. The allowable fatigue damage would then be split between
installation, as laid and operation; DNV-OS-F101 indicates that a split of 10%, 10% and 80%
is common.
An ECA is then conducted. It is not uncommon for this to indicate small tolerable flaw sizes.
The question that then arises is the relationship between ECAs and S -N curves.
S-N curves are only applicable if the weld is free from sig nificant defects; this effectively
means no planar flaws (see section 7). Fatigue crack initiation represents a small proportion
of the fatigue life. In an ECA, the initial flaw is assumed to be a fatigue crack. There is no
initiation. S-N curves are directly based on experimental data. An ECA is based on
mathematical models, with an indirect relationship with experimental data though the fatigue
crack growth law and its experimentally derived constants. Therefore, a direc t comparison of
S-N curves and ECA is not straightforward. However, it is instructive to calculate the initial
size of flaw that corresponds to a given design S-N curve, taking into account the safety
factors that would be applied in design to the S-N curve, as would be done in an ECA.
Figure 2 is a plot of the initial size of a surface flaw, in terms of flaw height and flaw length,
that corresponds to a given fraction of the class E design curve. The calculations have been
conducted in accordance with BS 7910, using the flat plate stress intensity factor solutions
(§M3.2), two-dimensional weld toe stress concentration factors, and the simplified (one
stage) in-air fatigue crack growth law. No stress concentration factors due to misalignment
have been applied to either the ECA or the S-N curve. No factor of safety has been applied
to the results of the ECA. To simply the calculations, the final flaw size is taken to be either
50 percent of the wall thickness, or 95 percent of the wall thickness. The difference between
the calculated initial flaws sizes is small compared to the final flaw sizes, and decreases as
the fatigue loading increases – in fatigue, the initial flaw size is more important than the final
flaw size.
Figure 2 a) is for a 15 mm thick pipe and Figure 2 b) is for a 25 mm thick pipe. The
calculated initial flaw size is larger for the thicker pipe – the cyclic stress intensity is related to
a/B. A set of curves are shown for fatigue lives ranging from 0.5 times the class E curve (i.e.
a factor of safety of two) to 0.05 times the class E curve (a factor of safety of twenty). The
initial flaw size increases as the factor of safety increases. The figures show that a factor of
safety of between 5 and 10 on the fatigue life predicted using the S-N curve, depending on
thickness, is required for the initial flaw size to exceed the typical workmanship acceptance
levels in pipeline welding codes.

10/17
6
B = 15 mm class E
5
flaw height, a (mm) 0.05 = f, 0.95xB = a f

4 0.05 = f, 0.5xB = af

3
0.1
0.1

0.2 0.2
1
0.3
0.4
0
0 10 20 30 40 50 60
flaw length, 2c (mm)
a)

6 0.05
0.05
B = 25 mm class E
5
flaw height, a (mm)

0.1 = f, 0.95xB = a f
4
0.1 = f, 0.5xB = af

2 0.2
0.2

1 0.3 0.3
0.4 0.4
0.5 0.5
0
0 10 20 30 40 50 60
flaw length, 2c (mm)
b)
Figure 2 Initial flaw sizes corresponding to the class E S-N curve

This simple comparison between the results on an ECA and an S-N curve illustrates the
following points:
 initial flaw size is more important than final flaw size;
 the influence of final flaw size decreases as the fatigue loading increases, indicating that
in a design subject to high fatigue loading factors that influence fatigue are more
important than those that affect the final flaw size; and

11/17
 if the fatigue design uses a high proportion of the allowable fatigue damage, then small
initial flaw sizes are inevitable.
It is also apparent that the factor of safety applied to S-N curves is, implicitly, also a ‘defect’
factor.

10 ARE ECAs TOO CONSERVATIVE?


An ECA should be based on conservative data and assumptions. The failure assessment
diagram (FAD) in BS 7910 is based on failure avoidance. Therefore, it is to be expected that
an ECA is conservative. All design is conservative. The question is whether ECAs are too
conservative. In part, this can be answered by comparing the results of an ECA conducted
in accordance with BS 7910, with other guidance in documents such as the EPRG guidelines
for the assessment of defects in transmission pipeline girth welds and Appendix A of API
1104.
Figure 3 compares the tolerable flaw sizes calculated in accordance with BS 7910 (using the
flat plate stress intensity factor solutions (§M3.2), two-dimensional weld toe stress
concentration factors, and the reference stress solution for a circumferential flaw in a curved
shell (§P.4.3.2)) with the Tier 2 and Tier 3 defect limits in the EPRG guidelines. Two wall
thicknesses: 15 and 25 mm, and three diameters: 8.625, 12.75 and 16 in., are considered.
The line pipe grade is X65. A fracture toughness of 0.15 mm is assumed. No stress
concentration factors due to misalignment are applied. The applied load (a primary
membrane stress) is equal to the specified minimum yield strength. Residual stresses due to
welding are assumed to be uniform, and nominally equal to the yield strength, but reduced in
accordance with the value of the reference stress.
The tolerable flaw height decreases as the length increases, and is less than the Tier 2 and
Tier 3 limits (and less than the equivalent API 1104 limits). Aside from the residual stresses,
the main reason for the conservative results is that constraint has not been considered. The
single edge notch bend test specimen for measuring the fracture toughness has a high level
of constraint and will give a conservative value of the fracture toughness, if the structure has
a lower level of constraint than the test specimen.
Figure 4 shows the required toughness to give a tolerable flaw size equal to the Tier 2 and
Tier 3 limits. It varies between approximately 1.5 and 2.25 times higher than 0.15 mm. This
is perhaps slightly higher than the difference in constraint between a single edge notch bend
test specimen and a girth weld in a pipeline. A more thorough (and complicated) calculation
than the simple calculation described here would further reduce the conservatism.
Tables A-2 to A-4 of Appendix A of DNV-OS-F101 give the required J-integral for a range of
tolerable flaw sizes for a “generic ECA”. This “generic ECA” is applicable if the total
longitudinal strain is less than 0.4 percent. Figure 5 compares the tolerable flaw sizes and
toughness requirements from this “generic ECA” with the Tier 2 and Tier 3 limits. The
required values of the J-integral have been converted to equivalent values of crack tip
opening displacement using the conservative expression given in Appendix A / E106 of DNV-
OS-F101. The required toughness is significantly higher than the 0.15 mm of Tier 3, and the
difference increases as the flaw length increases. The “generic ECA” is intended to be
conservative.

12/17
5 5
B = 15 mm, = 0.15 mm B = 25 mm, = 0.15 mm

4 4
flaw height, a (mm)

flaw height, a (mm)


3 3

2 2

D = 8.625, 12.75 & 16 in.


D = 8.625, 12.75 & 16 in.
1 1

API 1 104

API 11 04
Tier 3

Tier 3
EPRG, Tier 2 EPRG, Tier 2
0 0
0 50 100 150 200 0 50 100 150 200
flaw length, 2c (mm) flaw length, 2c (mm)
Figure 3 EPRG, API 1104 and BS 7910 defect limits
0.35 0.35
X65 X65 B = 15 mm
fracture tough ness, d (mm)
fracture to ughn ess, d (mm)

0.30 X = 2.0 0.30 X = 2.0


B = 25 mm

0.25 B = 15 mm 0.25

X = 1.5 B = 25 mm X = 1.5

0.20 0.20

EPRG, Tier 2 EPRG, Tier 3


0.15 0.15
0 200 400 600 800 1000 0 200 400 600 800 1000
diameter (mm) diameter (mm)

Figure 4 Fracture toughness corresponding to EPRG Tier 2 and 3 limits


1.0 1.0
B = 15 mm, a = 3 mm B = 25 mm, a = 3 mm 8.625 in.
12.75 in.
fracture toughness, d (mm )

fracture toughne ss, d (mm)

0.8 0.8 16 in.

8.625 in.
0.6 12.75 in. 0.6
16 in.

0.4 0.4

0.2 0.2
Tier 3

Tier 3

EPRG, Tier 2 EPRG, Tier 2


0.0 0.0
0 50 100 150 200 0 50 100 150 200
flaw length, 2c (mm) flaw length, 2c (mm)
Figure 5 Tolerable flaw sizes and toughness requirements in DNV-OS-F101 and
the EPRG Tier 2 limits

13/17
11 ARE ECAs UNCONSERVATIVE?
ECAs are intended to be conservative. This conservatism is intentional. Leaving aside
conservative data and assumptions, the conservatism can be attributed to: constraint, over-
matching, tearing resistance, residual stresses, the FAD, and the various stress intensity
factor and reference stress solutions.
To address over-conservatism, the sources of conservatism are addressed. It is possible
that hidden, or unrecognised sources of non-conservatism can be unintentionally unearthed
in the pursuit of non-conservatism. A potential example of this is the influence of bi-axial
loading.
Also of concern are assumptions made without sufficient justification, on the grounds that the
ECA is over-conservative. A potential example is the arbitrary selection of constraint factors.

12 ECAs AS A DESIGN TOOL


ECAs are used in design to determine the tolerable flaw size at the start-of-life, prior to
installation, commissioning and operation, and hence to inform the determination of the
acceptable flaw sizes. The steps in this process are summarised in Figure 1. Normally, the
ECA is concerned with the girth weld, although the same principles could be applied to the
seam weld (but noting that thick-walled, small diameter flowlines are seamless) or even the
pipe body.
An ECA requires a large amount of detailed information, e.g. geometry, material properties
(including fracture toughness), installation and operational loads, welding procedures,
environmental effects, etc. This requirement is not conducive to conducting ECAs early in
the design cycle. However, because the results of an ECA can have significant implications,
it is desirable to conduct the ECA early in the design cycle. Fatigue design using S-N curves
requires less information, but similar issues can arise.
Detailed welding procedures, and the results of welding procedure q ualification/production
tests are not available early in the design cycle. It is difficult to measure the fracture
toughness of the girth welds if representative welds are not available. Installation loads vary
significantly with installation method. Operational loads are refined as the design is
developed. Experimental testing to determine the effects of the environment on endurance
(i.e. S-N curves) and fatigue crack growth rates, toughness, and susceptibility to
environmental cracking is both complicated and time consuming. Test programmes can
easily take nine to twelve months to complete.
Modern pipeline girth welds should easily meet current workmanship standards. Therefore,
the important question to be answered early in design is: are workmanship acceptance levels
fit-for-purpose? This is essentially the approach recommended by SAFEBUCK. It is
possible to fabricate girth welds that are effectively ‘defect-free’, e.g. steel catenary risers,
but this can have significant cost implications, particularly if the requirement is not identified
early in the design cycle. Similarly, changing the S-N curve that the girth welds are required
to meet can have significant cost implications.
The commentary on the EPRG guidelines on the assessment of defects in transmission
pipeline girth welds offer some informative comments on how defect limits based on fitness -
for-purpose (i.e. ECAs) should be used during construction/installation:
 special applications where longer defects are anticipated (e.g. new processes);
 as a concession by the pipeline operator (in conjunction with penalty clauses) to avoid
unnecessary repairs; and

14/17
 as an insurance policy for cases where a defect is detected during post-construction
audit or during in-service inspection.
Limiting the objectives of the ECA to determining whether or not typical workmanship
acceptance levels (e.g. surface planar flaws limited to 3 mm deep by 25 mm long) are fit-for-
purpose, rather than precisely defining flaw acceptance criteria, means that the calculations
can be framed in terms of reasonable and conservative (but not overly conservative)
assumptions, and sensitivity studies. If workmanship acceptance levels are fit-for-purpose,
then the information that becomes available as the design progress is simply used to verify
the initial assumptions. Larger flaw acceptance criteria could be developed, but the question
is then: is this of benefit to the long term integrity of the pipeline?
If the ECA indicates that workmanship acceptance levels are not fit-for-purpose then there is
a problem. Identifying the problem early on in the design means that it is more likely that it
can be addressed without large cost implications, by re-design and/or raising the priority of
project-specific testing or other studies. In addition, identify those types of design that are
likely to be challenging in this regard (e.g. lateral buckling in sour environments, hydrogen
embrittlement in sour environments) means that a proactive approach to solving the
problems can be adopted.
Factors of safety are another problematic area. Factors of safety are lower on ECAs than on
the design S-N curves. BS 7910 does not require additional factors of safety provided that
the data and assumptions are conservative. It is not always clear how conservative the data
and assumptions are.

13 WHY THINGS GET COMPLICATED…


The design of pipeline systems in deeper water with increased loadings arising from
responses to thermal and pressure cycling, in the presence of aggressive internal conditions
is not straightforward. ECAs are further complicated.
A number of issues can cause problems. A sour environment implies higher rates of fatigue
crack growth (due to corrosion-fatigue) and a susceptibility to environmental cracking (if the
environment is not dry). The design premise is normally the avoidance of the initiation of
environmental cracking, specifically stress corrosion cracking. The temperature in the
pipeline system changes significantly during a thermal cycle associated with a shut-down.
The loads reduce as the temperature falls, but so does the toughness. It does not follow that
the worst case is at the maximum design temperature. The avoidance of the initiation of
stress corrosion cracking may not be limiting condition at lower temperatures. Hydrogen
embrittlement may be a concern (the toughness may be reduced by an order of magnitude or
more). If the toughness is very low then assuming lower bound tensile properties may be
non-conservative. Constraint is less relevant if the toughness is low. Operating at high
temperatures and high pressures is associated with significant end expansion, pipeline
walking and lateral buckling. Fatigue loading associated with thermal cycling is high, but low
frequency. Frequency effects in a corrosion-fatigue environment can significantly increase
the rate of fatigue crack growth. High cyclic strains may cause a combination of ductile
tearing and fatigue crack growth. High strains reduce the toughness. Bi-axial loading issues
increase as the strains increase. Strain localisations due to changes in coating type or
thickness, weak joints and counter boring (to reduce misalignment) may have a significant
effect. Over-matching may not be achieved over the stress-strain response of concern,
because of the limitations in the ways that over-matching is specified. The operating and
flow regime can give rise to unexpected sources of cyclic loading such as the effect of
changes in contents density due to slugging flow. The effect of what happens during
installation, both in terms of fatigue and high strains, on the subsequent response during
operation may become more important.

15/17
We stray into the murky waters of the known unknowns and the unknown unknowns…

14 CONCLUSIONS
The pipeline systems that are likely to be developed in the future will involve a variety of
challenges, from high pressures and high temperatures, to high fatigue damage in
aggressive internal conditions. It is important that ECAs are part of the solution, and not part
of the problem. Work is ongoing through various joint industry projects and project specific
studies. More work will be required.
Returning to the three questions posed at the start of this paper, the answers would appear
to be: sometimes, possibly and no.

NOMENCLATURE
a flaw height
2c flaw length
B wall thickness
CTOD crack tip opening displacement
CVN Charpy V-notch
ECA engineering critical assessment

REFERENCES
1. Department of defense news briefing, Donald H. Rumsfeld, February 12, 2002.
http://www.defenselink.mil/transcripts/transcript.aspx?transcriptid=2636
2. ANON.; Guide to Methods for Assessing the Acceptability of Flaws in Metallic Structures,
BS 7910 : 2005, British Standards Institution, London, UK, July 2005.
3. ANON.; Fitness-For-Service, API Recommended Practice 579, First Edition, American
Petroleum Institute, January 2000.
4. ANON.; Assessment of the Integrity of Structures containing Defects, R6-Revision 4,
British Energy, June 2005.
5. ANON.; Submarine Pipeline Systems, Offshore Standard DNV-OS-F101, Det Norske
Veritas, October 2007.
6. ANON.; Fracture Control for Pipeline Installation Methods Introducing Cyclic Plastic
Strain, Recommended Practice DNV-RP-F108, Det Norske Veritas, January 2006.
7. ANON.; Specification for Welding of Steel Pipelines on Land and Offshore. Carbon and
Carbon Manganese Steel Pipelines, BS 4515-1 : 2004, British Standards Institution,
London, UK, November 2004.
8. ANON.; Welding of Pipelines and Related Facilities, API Standard 1104, Twentieth
Edition, American Petroleum Institute, Washington, USA, November 2005.
9. KNAUF,G., HOPKINS,P.; The EPRG Guidelines on the Assessment of Defects in
Transmission Pipeline Girth Welds, 3R International, 35, Jahrgang, Heft, 10-11/1996, p.
620-624.
10. MADDOX,S.J.; Fatigue Strength of Welded Structures, Abington Publishing, Second
Edition, 1991.

16/17
11. ANON.; Specification for Line Pipe, Exploration and Production Department, API
Specification 5L, American Petroleum Institute, Forty Second Edition, 2000.
12. ANON.; Specification for unfired fusion welded pressure vessels, Published Document
PD 5500 : 2006, British Standards Institution, Third Edition, January 2006.
13. ANON.; Fatigue Design of Offshore Steel Structures, Recommended Practice DNV-RP-
C203, Det Norske Veritas, August 2005.
14. MACDONALD,K.A., MADDOX,S.J., and HAAGENSEN,P.J.; Guidance for Fatigue
Design and Assessment of Pipeline Girth Welds, Health and Safety Executive, Offshore
Technology Report OTO 2000 043, May 2000.
15. ANON.; Code of Practice for: Fatigue design and assessment of steel structures, BS
7608 : 1993, Incorporating Amendment No. 1, British Standards Institution, London, UK,
1993.
16. ANON.; Steel Pipelines for High Pressure Gas Transmission, IGE/TD/1 Edition 4 : 2000,
Recommendations on Transmission and Distribution Practice, Institute of Gas
Engineers, Communication 1670, 2001.
17. ANON.; Code of practice for pipelines - Part 1: Steel pipelines on land, PD 8010-1 :
2004, British Standards Institution, London, UK, 2004.
18. ANON.; Code of practice for pipelines - Part 2: Subsea pipelines, PD 8010-2 : 2004,
British Standards Institution, London, UK, 2004.
19. ANON.; Specification for arc welding of carbon and carbon manganese steels, BS 5135 :
1984, British Standards Institution, London, UK, 1984.
20. ANON.; Background to New Fatigue Guidance for Steel Joints and Connections in
Offshore Structures, Prepared by Failure Control Ltd. and MaTSU for the Health and
Safety Executive, Health and Safety Executive, Offshore Technology Report OTH 92
390, December 1999.

17/17

S-ar putea să vă placă și