Вы находитесь на странице: 1из 41

Pipeline Defect Assessment

Phil Hopkins
Penspen Ltd., UK
(p.hopkins@penspen.com)

Contents

1.

INTRODUCTION.....................................................................................................2

2.

HOW SAFE IS A TRANSMISSION PIPELINE?......................................................4


2.1 GENERAL........................................................................................................... 4

3.

DEFECTS IN A TRANSMISSION PIPELINE SYSTEM : THE NEED FOR FITNESSFOR-PURPOSE......................................................................................................5


3.1 DEFECTS IN PIPELINES.......................................................................................5

4.

CAN I APPLY, AND DO I NEED TO USE, FITNESS-FOR-PURPOSE METHODS?1


4.1 PHASE 1 - APPRAISAL........................................................................................1
4.2 PHASE 2 - ASSESSMENT ....................................................................................1
4.3 PHASE 3 - SAFETY FACTORS AND PROBABILISTIC ASPECTS.................................2
4.4 PHASE 4 - CONSEQUENCE.................................................................................2
4.5 PHASE 5 - REPORTING.......................................................................................2

5.

ASSESSING DEFECTS AND DAMAGE IN A TRANSMISSION PIPELINE SYSTEM


................................................................................................................................ 2
5.1 INTRODUCTION...................................................................................................2
5.2 DECIDING ON THE CORRECT LEVEL OF ASSESSMENT .........................................3
5.3 FITNESS-FOR-PURPOSE ASSESSMENT METHODS FOR PIPE UNDER INTERNAL
PRESSURE LOADING .................................................................................................4
5.4 CONSTRUCTING DEFECT ASSESSMENT PLOTS ....................................................3

6.

NEW STRUCTURAL RELIABILITY METHODS - LIMIT STATE ANALYSIS.........4


6.1 RELIABILITY-BASED LIMIT STATE DESIGN.............................................................4

7.

CONCLUSIONS......................................................................................................6

8.

REFERENCES.....................................................................................................10

Penspen Group

Rev 2004/1

1.

INTRODUCTION

These course notes explain how to conduct fitness-for-purpose assessments (sometimes called
Engineering Critical Assessments (ECA)) of defects in a transmission pipeline, and present a summary
of the methods available. The methods cover a wide range of defects, and can be applied to both onshore
and offshore oil and gas pipelines. The following types of defect are discussed:
Gouges, Dents, Dents and Gouges
Corrosion
Girth Weld Defects
Stress Corrosion Cracking
Material Defects
Construction Defects
1.1 How do Pipelines Fail?
Oil and gas pipelines carry hazardous products, but it is important to remember that it is the product that
is hazardous, not the pipeline. The pipeline will have high reliability if it is correctly designed,
maintained and operated.
Pipelines can fail due to:

natural disasters or acts of God,

gross human error,

sabotage/wars

existing defects in a new pipeline or introduced during operation

Engineers can do little to prevent i iii., but have many tools available to reduce iv.
1.2 Dealing with Risks

There are millions of kilometres of transmission pipelines around the world. The oil and gas transmission
system in Western Europe alone is over 150,000 km in length. A pipeline, and all its associated plant,
must be operated safely and efficiently. There are four key issues ('risks') in the operation of these
systems:
1. Safety - the system must pose an acceptably low risk to the surrounding population,
2. Security of Supply - the system must deliver its product in a continuous manner, to satisfy the
owners of the product (the 'shippers') and the shippers' customers (the 'end users'), and have low
risk of supply failure,
3. Cost Effectiveness - the system must deliver the product at an attractive market price, and minimise
risk of losing business,
4. Regulations - the operation of the system must satisfy all legislation and regulations, and this
usually requires the operator to assess, and mitigate, pipeline-associated risks.

Penspen Group

Rev 2004/1

An operator must ensure that all risks are acceptably low. For most of the lifetime of a pipeline, this will
not be a problem, but occasionally, an operator will detect, or become aware, of defects in their pipeline.
In the past, this has usually led to expensive shut-downs and repairs. However, recent years have seen
the increasing use of fitness-for-purpose methods to assess these defects. These course notes present
some of the fitness-for-purpose methods applicable to defects in transmission pipelines.
1.2 Fitness for Purpose (FFP)
During the fabrication of a pipeline, recognised and proven quality control (or workmanship) limits will
ensure that only innocuous defects remain in the pipeline at the start of its life. These control limits are
somewhat arbitrary, but they have been proven over time. However, a pipeline will invariably contain
larger defects at some stage during its life, and they will require an engineering assessment to determine
whether or not to repair the pipeline. This assessment can be based on fitness for purpose, i.e. a failure
condition will not be reached during the operation life of the pipeline.
The fitness for purpose of a pipeline containing a defect may be determined by a variety of methods
ranging from previous relevant experience, to model testing, to engineering critical assessments, where
a defect is appraised analytically, taking into account its environment and loadings (for example the UK
code, British Standard BS7910, or the USA code from the American Petroleum Institute, API 579 see
later). It should be noted that fitness for purpose is not intended as a single substitute for good
engineering judgement; it is an aid.
1.2.1 Generic FFP
There are various technical procedures available for assessing the significance of defects in a range of
structures. These methods use fracture mechanics; for example, BS 7910 contains detailed engineering
critical assessment methods, and can be applied to defects in pipelines. Also, API 579, which has similar
methods to BS 7910, can be used: but this standard has a bias towards process plant.
1.2.2 Pipeline-Specific FFP
The above standards are generic; they can be conservative when applied to specific structures such as
pipelines. Therefore, the pipeline industry has developed its own fitness for purpose methods over the
past 35 years. However, it should be noted that they are usually based on experiments, with limited
theoretical validation (i.e. semi-empirical). This means that the methods may become invalid or
unreliable if they are applied outside these empirical limits.
The pipeline industry has used their fitness for purpose methods to produce specific guidelines for the
assessment of defects in pipelines. These methods and guidelines are based on pioneering work at
Battelle Memorial Institute in the USA on behalf of the Pipeline Research Council International (see
below), with the more recent additions of ad hoc guidelines for the assessment of girth weld defects,
mechanical damage and ductile fracture propagation produced by the European Pipeline Research Group
(see below)).
Best practices in structural assessments of defects in pipelines are now emerging (see Section 7, later),
and a Joint Industry Project sponsored by 14 major oil and gas companies produced a state-of-the-art
Pipeline Defect Assessment Manual in 2003 (see Section 7, later).

Penspen Group

Rev 2004/1

2.

HOW SAFE IS A TRANSMISSION PIPELINE?

2.1

General

Pipelines are the safest form of transporting energy[1]. A review[2] on failure rates in pipelines in Western
Europe and North America concluded:

T ABLE 1 - Failure rates in transmission pipelines [2]


FAILURES
Failure Rate per 1000 km year
Any Incident Requiring a Repair to the Pipeline
4
Incident causing Loss of Product
0.6
Incident causing Loss of Product PLUS a
0.16
Casualty or High Damage Cost
The risk resulting from pipeline failures are very low when compared with the risks from other forms of
failures/disasters[3]. This can be illustrated by considering the 450,000 mile long North American gas
transmission system where there have been 30 fatalities, and 106 casualties between 1987 and 1993.
These casualty figures are very low compared with other risks to the general public in the USA (Table
2).

T ABLE 2 - Relative risks of fatality in USA In 1992 [1]


RISK
Gas Transmission Pipelines
Liquid Lines
Distribution Lines
Accidents in the Home
Falls
Accidental Poisoning
Firearm Accidents

FATALITIES IN 1992
4
0
13
19 500
6 200
4 100
700

Pipeline failures can also cause environmental damage. Figure 1 shows the total amount of oil spills into
the marine environment between 1990 and 1999. Nearly 1,000 million gallons (3 million tonnes) were
spilled in this period, but over 50% was due to tanker spills: pipelines accounted for only 6.4% of total
spills. Sub-sea pipelines can and do fail (see Figure 2, later), but they tend to spill less product that a sea
tanker. The tankers at sea today are massive: the Exxon Valdez that spilled oil in 1989 released 11
million gallons (257,000 barrels or 38,800 metric tonnes) of its 53 million gallon cargo. 11 million
gallons would fill 125 Olympic swimming pools.

Penspen Group

Rev 2004/1

60
50
40
30
20
10
0

Figure 1. Amount of Oil Spilled in Marine Waters Worldwide (1990-99)

3.

(68)

DEFECTS IN A TRANSMISSION PIPELINE SYSTEM : THE


NEED FOR FITNESS-FOR-PURPOSE

Fitness-for-purpose methods for assessing defects in pipelines use fracture mechanics as their technical
basis[4]. Therefore, they are applied mechanics methods.
However, because they require an engineer to work outside existing codes, and because the defect may
be large and a potential safety hazard, they must be used in a structured and systematic manner, by
suitably qualified engineers.
When a defect is assessed using fitness-for-purpose methods, it should not decrease the overall integrity
of the pipeline. As the name suggests, a fitness-for-purpose assessment will always ensure that the
pipeline is left in a fit and safe state.
3.1

Defects in Pipelines

The major cause of damage and failures in transmission pipelines in Western Europe and North America
is external interference ('mechanical damage), e.g. a farmer gouging a pipeline accidentally, while
ploughing, or a supply boat denting an offshore pipeline by dragging an anchor across it [5]. Table 3
summarises some failure data, and the major causes of failures in certain pipeline systems.

Penspen Group

Rev 2004/1

TABLE 3. Major Causes of Failures in Pipeline Systems [5]


COUNTRY
USA (Onshore gas)
USA (Offshore gas)
USA (Onshore & Offshore oil)
Western Europe (Onshore gas)
Western Europe (Onshore oil)
Hungary (Onshore gas)
Poland (Onshore gas)
CIS (Onshore gas)
Czechoslovakia (Onshore gas)

MAJOR CAUSE OF
FAILURE
External Interference
Corrosion
External Interference
External Interference
Corrosion
Girth Weld Defects
Corrosion
Construction/Material Defect
Construction/Material Defect

FAILURE RATE
(per 1000 km year)
0.16
0.70
0.56
0.60
0.80
0.10
0.08
0.33
0.13

The above table shows various causes of failure and failure rates. There are more detailed breakdowns
of failure data: Figure 2 shows a detailed breakdown of failures in offshore pipelines in the UK sector of
the North Sea. This is taken from the Pipeline and Riser Loss of Containment 2001 Report, Energy
Institute, UK.
The differing failure rates are probably due to differing definitions of failure (see Reference 5 and Table
1). The variety of failure causes is not surprising as pipelines operate in a variety of hostile
environments. Eastern Europe has problems with construction and material defects, and also with girth
weld defects. However, problems with girth weld defects are not confined to Eastern Europe; one
Western European operator has reported the following defects in girth welds[4]:

T ABLE 4 - Older Girth Welds in a Transmission System [4]


GIRTH WELDS
Fabricated Before 1968
Fabricated Between 1968 and 1972
Fabricated After 1972

% UNACCEPTABLE TO CURRENT CODES


70
10
0

Clearly, older girth welds may contain unacceptable defects. These welds are unacceptable to
workmanship standards, but they are not necessarily unfit-for-purpose. Clearly, there is a need for
detailed fracture mechanics analyses to determine the scale of these type of problems, and avoid
unnecessary repairs (see Section 5.3.9, later).

Penspen Group

Rev 2004/1

400

Number of
incidents 350

250
INCIDENTS

OPERATION

200

300
250

Fig
ure 1. Failures in offshore

150

200
100

150
100

50

50
0
Supply
boat
(anchors) and
trawl
boards
(impact)
are
main causes in
this category

CONSTRUCTION

OPERATION

50

PIPELINES

160
LEAKS

40

FITTINGS

PIPELINES

140
120
100

30

80
20

60
40

10
0

20
ANCHO R/IMPACT

MATERIAL/
CORROSION

'O THER'

Figure 2. Failures in Sub-sea Pipelines In UK Sector of North Sea

Penspen Group

Rev 2004/1

LEAKS

NO LEAKS

Data is for steel


and flexible
pipelines:
24,837km in
length,
328,858km-years.
An incident is
an
occurrence
which
directly
results in or
threatens to result
in loss of
containment.

4.

CAN I APPLY, AND DO I NEED TO USE, FITNESS-FORPURPOSE METHODS?

Any engineer with a potential defect problem should structure their assessment as follows:
4.1

Phase 1 - Appraisal

4.1.1 Is it Really There, and Could It Go Away?

Is it really a defect, or is it some feature of the inspection method I am using (e.g. a pig report)?

Are the operating conditions able to create such a defect?

If I've detected a defect, can I leave it alone? For example, is it within design and fabrication
acceptance levels?

What experience of the defect exists? For example, have other companies faced this problem, and
produced a solution that concludes that the defect is acceptable?

4.1.2 Is It A Defect, Or A Mess?

Do I know how the defect was formed, and how it may develop? For example, if it is a design fault,
the last thing I want to do is apply fitness-for-purpose methods - I need to re-design.

Is the defect indicative of malpractice? For example, are there so many defects, or are the defects so
large, that they warrant a re-appraisal of the design or operational practices?

4.1.3 Why Me, And Can I Do It?

Where do I stand legally (professional liability), and with my Regulatory Body (statutory liability),
and who is responsible for the structure (legal liability), and any defect assessment relating to it?

Am I capable and experienced enough to apply fitness-for-purpose methods?

4.1.4 Is It Worth The Effort?

4.2

Is it cheaper to repair than assess?


Phase 2 - Assessment

4.2.1 Can Fitness-For-Purpose Methods Reliably Solve It?

Can fitness-for-purpose methods solve my problem? For example, are the methods robust for my
particular defects?

What data do I have, and how reliable are they? If I do not have any data, I will need to use much
judgement, or conduct special tests.

The key to a safe and reliable fitness-for-purpose assessment is good input data. Given good
consistent data, most fitness-for-purpose methods will be safe and reliable.

Penspen Group

Rev 2004/1

4.3

Phase 3 - Safety Factors And Probabilistic Aspects

4.3.1 What Safety Margins Are Needed?

If I'm going to apply fitness-for-purpose, what safety factors am I going to use?

How am I going to set these safety factors, and would it be better to conduct a probabilistic
analysis?

4.4

Phase 4 - Consequence

4.4.1 What Are The Consequences Of Getting It Wrong?

4.5

What are the consequences of my getting it wrong? Would a risk analysis that takes into account
consequences be more appropriate? This can be simplistic, e.g. a leak before break criterion, or a
full quantitative risk analysis.
Phase 5 - Reporting

4.5.1 Who Needs To Know, And What Detail Is Needed?

Any fitness-for-purpose analysis of a defective structure should be fully reported, with details of all
assumptions, safety factors, etc., and evidence of both quality assurance and accuracy checks. The
report should be sent to the owners and operators of the structure, and included in the structure's
design file. This is particularly important if the analysis has recommended monitoring, or if the
defect will restrict any possible uprating of the structure. Additionally, the report may need to be sent
to Regulatory Authorities, etc..

5.

ASSESSING DEFECTS AND DAMAGE IN A TRANSMISSION


PIPELINE SYSTEM

5.1

Introduction

The first consideration when applying fitness-for-purpose methods to any defect detected or expected in
a transportation system is the predominant micromechanism of fracture.
Transmission pipelines usually have relatively thin wall thicknesses (most are less than 25 mm), and the
steels have been produced for many years to recognised international standards (e.g. API 5L). This
ensures that our line pipe is ductile, and will not exhibit low toughness or brittle behaviour. This means
that the line pipe we use in pipelines is ductile.
Transmission pipelines are usually buried or subsea, and therefore they are not generally subjected to
low temperatures. Accordingly, the failure of part wall defects (such as corrosion pits) in (parent)
transmission pipeline material, is by tensile necking of the ligament remaining below the defect. Sharper
defects in parent material, such as scratches or gouges, fail by progressive ductile crack growth
(microvoid coalescence).
Defects in older pipeline steels (e.g. those produced in the 1950s), and the thicker walled pipelines, will
generally show ductile fracture initiation (although they may eventually fail by dynamic brittle crack
propagation). Consequently, defects in transmission pipelines are generally assessed using solutions
based on limit loads, or plastic collapse[6-8].

Penspen Group

Rev 2004/1

Defects in longitudinal (i.e. factory produced) pipeline welds usually fail by a ductile fracture
mechanism, although defects in circumferential (field) welds can fail by brittle fracture and may require
brittle fracture calculations[9]. However, new guidelines, specifically for transmission pipelines are
available (see Section 5.3.9, later).
Other pipeline defect such as loss of wall thickness defects within dents, can be highly constrained, and
can exhibit low ductility failures. These defects are often assessed using fracture mechanics based
models that incorporate the effect of the material toughness[9] (see Section 5.3.4, later).
5.2

Deciding on the Correct Level of Assessment

Having decided that a defect assessment can be conducted, it is now necessary to determine the level of
detail and complexity that is required.
Different levels of defect assessment, ranging from simple screening methods (e.g. the ASME B31.G,
Section 5.3.5, later) to very sophisticated finite element stress analyses or probabilistic/risk methods are
available. The methods used depend on the defects detected, or the type of pipeline, or the customers
requirements. Figure 3 summarises the differing levels of defect assessments, and the required data.
Generally, defect assessments are conducted up to Stage 3. If defects still remained unacceptable at this
Stage, a higher level assessment, or repair would be necessary.
The higher levels may require risk analyses. Risk is a function of the probability of failure and the
consequences of failure. Such analyses are becoming increasingly popular [10], but are also very
complicated. Similarly, limit state analyses, which also work out the probability of failure, are also
becoming popular[11,12] (see Section 6).
If an operator requires a risk analysis, then both probability and consequences of failure need to be
considered in a probabilistic and detailed manner. Usually a fitness-for-purpose analysis of defects does
not entail a risk analysis, although due account of the consequences of failure will be taken in a
qualitative manner, and the recommended safety factor will reflect this. The fitness-for-purpose
assessment will usually involve a deterministic assessment of the defects, on a go/no go basis. If there is
a problem with the defects, or with defects with a significant consequence on failing, then a risk analysis
may be recommended.

Penspen Group

Rev 2004/1

Figure 3. Various Defect Assessment Levels


FIGURE 1. PIPELINE DEFECT ASSESSMENT:

THE FIVE STAGES

DATA

STAGE 1
QUALITATIVE -

ACCEPT

e.g. COMPANY OR CODE


WORKMANSHIP LEVELS

REJECT

AS ABOVE, PLUS,
PIPE DATA,
PIPE PRESSURE

STAGE 2
QUANTITATIVE -

ACCEPT

e.g. ASME B31.G


CODE

REJECT

AS ABOVE,
BUT ADDITIONAL
DEFECT, PIPE &
MATERIAL DATA

STAGE 3
QUANTITATIVE e.g. FRACTURE
MECHANICS CALC.
REJECT

AS ABOVE, PLUS,
PIPE SAMPLE
OR MATERIAL
PROPERTIES

ACCEPT

REJECT

STAGE 4b
QUANTITATIVE STAGE 4a
EXPERIMENTAL-

ACCEPT

NUMERICAL
ANALYSIS
REJECT

REJECT

MODEL/FULL
SCALE TESTING

REJECT

ACCEPT
REJECT

AS ABOVE, PLUS,
DISTRIBUTIONS
OF PIPE, MATERIAL
& DEFECT DATA

STAGE 5
PROBABILISTIC (USING
LIMIT STATE ANALYSIS), OR

RISK ANALYSIS
REJECT

REJECT
5.3

ACCEPT

<......SIMPLE STAGES.......> <........................................................EXPERT LEVELS........................................................>

DEFECT
SIZE & TYPE

STAGES

ACCEPT

Fitness-For-Purpose Assessment Methods For Pipe Under Internal Pressure


Loading

This Section gives a summary of the fitness-for-purpose methods available to assess the variety of
defects that may occur in a pipeline. Defect assessment guidelines and limits for pipelines have been
available for many years, either in national codes, or in company specifications.

Penspen Group

Rev 2004/1

When summarising fitness-for-purpose methods, it is best to start off with an assessment of the failure
stress of a defect-free pipe. This gives a benchmark failure stress for any pipeline.
It should be noted that safety factors are not given or recommended in the following Sections - they will
be dependent on the type of defect, the reliability of the data used in the assessment and the assessment
method, and the consequences of the failure of the defect. It is the responsibility of the engineer
conducting the assessment to derive an appropriate safety factor.
Before we talk about defect failure criteria we need to know how a defect in a pipeline fails.
A transmission pipeline can fail in a variety of ways: by internal pressure bursting the pipe, by axial
overload caused by earthquake, etc.. However, the most common failure (Figure 4) is by internal
pressure loading on a part wall defect or pipe damage.
A defect in a pipeline with internal pressure will be subjected to two main stresses: hoop and axial. The
hoop stress is the largest for internal pressure.
hoop
axial

A part wall defect in ductile line pipe under internal pressure fails as follows:

PART WALL DEFECT:


o

The defect bulges as the pressure in the pipeline is increased.

The ligament below the defect plastically deforms.

Stable crack growth may start, as the pressure continues to increase.

Unstable crack growth, through the wall, lead to the creation of a through wall defect.

THROUGH-WALL DEFECT:

This through-wall defect can fail either as a:

LEAK (its length does not increase) or

RUPTURE1 (its length does increase), depending on its initial length


and the pipeline pressure (see Figure 5, later).

The rupturing defect can either:


o

ARREST (the rupturing defect quickly stops increasing in


length).

PROPAGATE (the rupturing defect continues to increase in


length to create a propagating fracture

Sometimes called a break.

Penspen Group

Rev 2004/1

a. Pipeline contains a

b. If the stress in the Pipeline is above a critical value, then the remaining
ligament below the Part Wall Defect fails

Part Wall Defect

and produces a Through-Wall

l
d

Defect

c. A Through Wall Defect in a Pipeline.

d. The through Wall Defect causes a Leak


if the defect is Short, or if the pressure is Low.

e. The Through Wall Defect causes a Rupture


if the defect is Long, or if the pressure is High.

f. The Through Wall Defect ruptures, but

g. The Through Wall Defect ruptures, and Propagates


if the pressure is high, and/or if the pipe has a Low Toughness.

Arrests
if pressure is low, and/or pipe is High Toughness, or
if the product is a liquid.

Figure 4. How a Pipeline Defect Fails


5.3.1 Defect-Free Pipe under Internal Pressure
t

Pipe Dimensions
The simplest and, in general, the most conservative formula for the range of transmission pipeline D/t
ratios is given by using u in the simple Barlow equation:

Pf

2t u
D

(1)

where:
Pf = failure pressure (MPa)
t = pipewall thickness (mm)
D = outside diameter of pipe (mm)
u = ultimate tensile strength (MPa)
Penspen Group

Rev 2004/1

For a more accurate assessment, the method given in ASME VIII[13] is:
K 1
for K 1.5
0.6 K 0.4 u

(2)

K 2 1

K 2 1 u

(3)

Pf

Pf

for K > 1.5

where:

Ro
Ri

(4)

Ro = outside radius of pipe D/2 (mm)


Ri = inside radius of pipe, Ro-t (mm)
More complex and accurate methods are available[14].
5.3.2 Gouges or Similar Metal Loss Defects2
External interference, or damage during construction, can cause gouges or scratches on the pipes
surface. These metal loss defects may be accompanied by local plastic deformation.
If this deformation has caused a dent, then the gouge must be assessed using sophisticated fracture
mechanics methods (see Section 5.3.4).

Even
if

Denting
Hardened
layer

Pipe

the

Cracking
gouge
is
not
associated with a dent, there may be a work hardened layer at the base of the gouge which may reduce
the local ductility and it may contain cracking.
It is good practice to remove any surface hardening by grinding, although special procedures are
required for dressing damage on a live pipeline. Following dressing, the gouge can be assessed using its
dressed dimensions.
2

See Section 5.3.10.2 for methods for assessing the fatigue life of these type of defects.

Penspen Group

Rev 2004/1

5.3.2.1 Axially-Orientated Gouges


In ductile line pipe, the failure stress of an axially-orientated gouge subject to internal pressure loading is
described by[6]:
2c

d
t

Defect Dimensions

Ao
A 1
M
Ao

M 1 0.26

2c

Rt

or

t
d 1
M
t

(5)

(6)

where:
Pf f

t
R

f = failure stress (MPa)


M = bulging factor (Eq. 6 is an example there are many other formulations)

= flow stress (MPa)


d = maximum or average depth of part wall defect (mm)
2c = defect axial length (mm)
A = area of metal loss in the axial plane through the wall thickness (mm2)
Ao = original area (d2c)
R = outside radius of pipe (mm)
Equation 6 can also be used to determine the failure of a through wall defect: this failure of a through
wall defect will be a rupture as the through wall defect is already leaking (see Figure 3). The stress to
fail (rupture) a through wall defect is

Penspen Group

Rev 2004/1

M 1

(6a)

Assessments of part wall defects are more accurate if the area ratio (A/Ao) is used. Using the maximum
defect depth will be conservative (using the average defect depth is equivalent to using the area ratio).
Figure 5 gives a simple representation of Equations 5 and 6a.
Gouges can be assessed using the above equations, providing your pipeline has a toughness >20J (19).
Note that a gouge needs to be checked for possible fatigue crack growth in some pipelines (e.g. some
liquid lines).
Allowance (e.g. adding 0.5mm to defect depth) for the hard layer or sub-surface cracking is advisable, if
they are to be left in the pipeline, but you should ensure there is no risk of environmental cracking, and
no residual denting, and no problems from cyclic loading.
Note that the assessment of gouges in pipeline is often not allowed by codes and regulations [64-66] due to
concern over the difficulty of detecting associated denting and cracking. Gouges in these
codes/regulations require repair.
5.3.2.2 Circumferentially-Orientated Gouges
For a circumferential gouge (orientated at an angle of 90o to the pipeline axis) in ductile linepipe, the
following plastic collapse failure criterion due to Kastner [15] may be used to calculate the axial failure
stress:
2c

R
t

Defect Dimensions

z
1

21 sin

c
R

d
t

where:
z = axial failure stress (MPa)

= flow strength (MPa)

Penspen Group

Rev 2004/1

d = maximum or average depth of part wall defect (mm)


c = half circumferential length of defect(mm)
t = pipe wall thickness (mm)
R = outside radius of pipe (mm)

Penspen Group

Rev 2004/1

2c (l)

M
t

1 0.40

2c

Rt

Flow strength = 1.15y

0.8

1 - (d/t) = 0.6

0.6

0.5
0.4

0.4

0.3
0.2

0.2

0.1
0.05

0
0

1.2

2c/(Rt)^0.5
F
a
i
l
u
r
e
S
t
r
e
s
s
/
Y
i
e
l
d
S
t
r
e
n
g
t
h

F
a
i
l
u
r
e
S
t
r
e
s
s
/
Y
i
e
l
d
S
t
r
e
n
g
t
h

1.2

f
M 1

This boundary is not sensitive to pressurising medium

2c or l

0.8

RUPTURE

0.6

LEAK

0.4

0.2

2c
0
0

2c/(Rt)^0.5

Figure 5. Equations 5 (left) and 6a (right) as Universal Graphs

Penspen Group

Rev 2004/1

5.3.3 Dents (with no associated defects)


5.3.3.1 Burst Strength of Plain Dents
A dent in a pipeline is a permanent plastic deformation of the circular cross section of the pipe. A plain
dent (sometimes referred to as a smooth dent) is defined as damage which causes a smooth change in
curvature of the pipe wall without a reduction in pipe wall thickness, i.e. it contains no defects or
imperfections, such as a girth or seam weld. A kinked dent contains rapid changes in contour.
Dents in pipelines are assessed using data derived from full scale tests. Large dents can be tolerated [16],
although their behaviour under cyclic loads, or when they coincide with seam welds, remains a
problem[16].
The effect of a plain dent (i.e. one with no associated loss of wall thickness defect, and of smooth
shape) is to introduce high localised stresses and cause yielding in the pipe material. The high stresses
and strains caused by the dent are accommodated by the ductility of the pipe. Full scale test results have
confirmed this by showing that plain dents do not generally affect the burst strength of the pipeline[16-22].
On pressurisation the dent attempts to move outward, allowing the pipe to regain its original circular
shape. Provided that nothing restricts the movement or acts as a stress concentration (e.g. a gouge or a
kink), then the dent will not reduce the burst strength of the pipe.
Empirical limits for plain dents under static internal pressure loading have been derived from extensive
full scale testing. It should be noted that all of the dent depths in the full scale tests were measured at
zero pressure. Based on these full scale tests, BG (formerly British Gas) quote that a plain dent of less
than 8% of the pipe diameter has little effect on the burst strength of pipe[16]. However, full scale tests on
plain dents on welds and dents containing defects have demonstrated very low burst pressures. There
have been no burst tests on dented girth welds. Therefore, the burst strength (and the fatigue strength) of
dents containing welds cannot be reliably predicted, and caution is recommended with this type of
damage.
It should be noted that a dent of depth 8% of the pipe diameter may restrict product flow, and the
passage of pigs in the pipeline. Consequently, the maximum dent depth usually accepted by pipeline
companies is 6% pipe diameter, and this is the depth that is used in codes (64, 65) and new USA pipeline
safety regulations (66).
5.3.3.2 Fatigue Life of Plain Dents
Under cyclic pressure loading there will be large cyclic stress and strains localised in the dent. Full scale
fatigue tests[16-22] on plain dents indicate that they reduce the fatigue life compared to plain circular pipe.
The greater the dent depth the shorter the fatigue life. No fatigue failures occurred in those tests where
the pipe was hydrotested prior to fatigue cycling, because the dent was pushed out, mitigating its stress
concentrating effect.
A semi-empirical model for predicting the fatigue life of a plain dent subject to cyclic pressure loading
has been developed[19-21]. One of the relationships developed is:

N 2.0 x10
p
6

1
p 11400

3.74

(7)

where
N

= fatigue life

Penspen Group

Rev 2004/1

= stress intensification factor


= cyclic pressure, psi

The reader is directed towards the original references[19-21] if they wish to conduct an assessment. The
fatigue model is based on an S-N curve, modified for the stress concentration due to the dent. A stress
intensification factor has been derived from FE analysis to account for the stress concentration due to
the dent. It is a function of the diameter to wall thickness ratio (D/t), the ratio of the dent depth to
nominal diameter, and the average pressure.
However, it should be emphasised, that for the fatigue design of pipelines based on S-N curves, it is
normal practice to take an S-N curve based on the mean minus two standard deviations (i.e.
approximately 97.7% one tail confidence limit) and to then apply an allowable damage ratio of 0.1
(equivalent to a factor of safety of ten on fatigue life[23]).
5.3.4 Plain Dent Containing a Defect
5.3.4.1 Burst Strength
Dents containing defects can record low failure pressures. This is because a defect in the dent is affected
by the stress concentration and the large strains within the dent; this causes ductile tearing of the defect
through the remaining ligament. The structure comprising the dent and the defect is complex and
unstable.
Full scale tests and ring tests investigating the burst strength of combined dents and defects have been
undertaken[16-22]. Based on full scale and model vessel tests, BG developed a semi-empirical fracture
model for predicting the failure stress of a combined dent and gouge[24]. The model includes the effect of
the material toughness (measured in terms of the Charpy impact energy).
The fracture model is based on tests in which the damage was introduced into unpressurised pipe,
therefore the dent depth measured at zero pressure must be used[17].
The fracture model is defined as follows (in imperial units):
D

R
t

Defect Dimensions

1.5E
1.8 D0
2
R D0

1
cos exp 2
Y2 10.2

Y1 1

2R
t 2R

Ad

where:

Penspen Group

Rev 2004/1

ln(C v ) K 1
exp

K2

(8)

115
. SMYS 1

(lbf/in2)

d
d
Y1 112
. 0.23 10.6
t
t
d
d
7.32
t
t

Y2 112
. 1.39

d
131
.
t

21.7
2

d
14.0
t

30.4

K1 1.9

K 2 0.57
(K1 and K2 are non-linear regression parameters)
f = hoop stress at failure (lbf/in2)

= plastic collapse stress of infinitely long gouge (lbf/in2)


A = fracture Area of Charpy (0.083 in2 for a 2/3 Charpy specimen)
E = Youngs Modulus (30,000,000 lbf/in2)
Cv = 2/3 Charpy toughness (ftlbf)
d= maximum or average depth of part wall defect (in)
D0 = dent depth measured at zero pressure (in) (See Reference 17)
t = pipe wall thickness (in)
R = outside radius of pipe (in)
This failure criterion for a dent containing a metal loss defect does not give a lower bound failure stress.
It is a mean predictive model. Additionally, the model is semi-empirical and therefore limited by the
bounds of the original test data[16-22].
The model is well known to give large scatter in its predictions[63], and it is considered more of a research
tool than a practical model, hence, its use in the field is no recommended without expert help.
5.3.4.2 Fatigue Life
The fatigue life of a dent containing a gouge is difficult to predict. Full scale tests indicate that the
fatigue life of a combined dent and gouge can be of the order of between ten and one hundred times less
than the fatigue life of an equivalent plain dent. In some cases even shorter fatigue lives have been
observed during testing.
5.3.5 Dents on Welds
5.3.5.1 Burst Strength
Full scale tests have shown that dented seam welds can exhibit very low burst pressures [16,17]; the
minimum burst pressure in one test was 24 percent of the SMYS. The low burst pressures occur due to
cracking in the weld as a result of the large stress and strains associated with the denting process. The

Penspen Group

Rev 2004/1

burst strength of a dented weld is critically dependent on whether or not the weld cracks during the
denting process, and whether or not the weld contains any fabrication defects. There are no methods for
reliably predicting the failure pressure of a dented weld. Therefore, dented welds are usually repaired if
found in an operational pipeline.
5.3.5.2 Fatigue Life
There have been a number of fatigue tests on pipe rings containing dented seam welds [16,17], and fatigue
tests on vessels containing dented seam welds and dented girth welds[20,21]. These tests have shown that
the fatigue life of a dent containing a weld can be considerably lower than the fatigue life of an
equivalent plain dent. There are no methods for reliably predicting the fatigue life of a dented weld.
Reference 63 gives full advice on the fatigue behaviour and Reference 65 gives more practical guidance,
but the following is a list of key facts about seam welds in dents:
-

Tests on low frequency ERW and DSAW welds in dents, and deep dents on girth welds have
shown low fatigue lives.

Old welds such as acetylene welds or severely flawed arc welds may have poor fatigue lives.

If the longitudinal seam or girth weld is good quality, ductile and free from major defects (e.g.
the girth weld is to API 1104), the fatigue life can be good

Some tests on (the better quality) high frequency ERW welds in dents have shown good fatigue
lives.

Work in the 90s for OPS in the USA also supported the view that shallow rock dents on girth
welds in gas lines were not a major problem

Generally, dents on girth welds or ERW, or DSAW welds have fatigue lives of about 20% that of
a similar plain dent with no weld.

5.3.6 Note of the Assessment of Gouges and Dents


If a gouge or dent is detected in the field, an engineer needs to check:
- DENTS OR GOUGES CHECK FOR SURFACE CRACKING - There may be some cracklike indications (spalling) caused by the damaging object. If the cracking is deep, it may be
indicative of a gouge that has cracked due to denting (the denting may not be visible as it may
have been pushed out (24,25)). This is severe, and requires repair.
- GOUGES CHECK FOR EVIDENCE OF DENTING - The impact may have also dented the
pipe. Residual denting around a gouge is severe see above.
Finally, an engineer should always think carefully of the consequences of getting things wrong. If your
damage is in a pipeline in a high consequence area, you should inspect the damage closely before
assessment, and include appropriate safety factors in your assessment.

Penspen Group

Rev 2004/1

5.3.7 Corrosion3
There are several approaches that have been used to characterise the behaviour of both through and part
wall corrosion defects.
5.3.7.1 ANSI/ASME B31G
The most well-known document for the assessment of the remaining strength of pipelines with smooth
corrosion is ANSI/ASME B31.G[26]. This supplement to B31 was developed over 20 years ago [25],
although it has recently been re-issued[27,28]. It is based on an empirical fit to an extensive series of full
scale tests on vessels with narrow machined slots. The basis of the equation used in B31G is relatively
simple and involves:

Assuming that the maximum pipe hoop stress is equal to the pipe materials yield strength, and,

Characterising the corrosion geometry by a projected parabolic shape for relatively short corrosion,
and a rectangular shape for long corrosion.

The equation used is the same as in Section 5.3.2:

Ao

(5)

A 1
M
1

Ao

In the B31G code, a simplified equation is provided which represents the defect as a parabolic shape
(hence the two thirds factor in the equation):

2d
3t

2d 1
1
3 t M
1

(9)

The flow strength ( ) is defined by 1.1xSMYS (specified minimum yield strength). In more recent
years a flow strength of 1.15xSMYS has been used to accommodate the added ductility in modern steels.
The parabolic shape of the projected area is used as an approximation to the actual defect, and a
modified bulging factor is used:

M 1 0.8

2c

Dt

(10)

It is stated in the code that the above equations should only be applied to corrosion defects which have a
maximum depth greater than 10% of the nominal wall thickness, and less than 80% of the nominal wall
thickness. Furthermore, the relative longitudinal extent should satisfy the following equation:

M 1 0.8

2c

Dt

(11)

4.12

Usually, the most difficult data to obtain when assessing corrosion, is the expected corrosion growth rate. This
is important, because most assessments of corrosion are based on intelligent pig data, where the defect must be
assessed over its whole life, and its size at the end of the pipelines design life needs to be used in the
calculations.

Penspen Group

Rev 2004/1

In Reference 26 the above limit on the defect length is expressed in terms of B:

B 4. 0
where
B

M 2 1

The above equation limits the use of the parabolic shape formulation because when M is greater than
4.12 (i.e. long corrosion), the approximation of a parabolic shape is no longer adequate. Instead a
rectangular shape is used. Accordingly the failure equation is replaced by the following equation:
d

f 1
t

(12)

5.3.7.2 Modified B31.G/RSTRENG


The B31G criterion has been used successfully in the pipeline industry for many years. The method has
been proven, in general, to be conservative and as a result an improved method was developed which
modified the existing B31G guidance, and removed some of the excess conservatism[27]. The modified
B31G/RSTRENG[27,28] method has recently been adopted as the preferred method for the fitness for
purpose assessment of corrosion defects in the ANSI/ASME B31 Code[28].
The main modifications to the original B31G method were to change the definition of the flow stress, the
definition of the Folias factor, and to replace the parabolic area assumption (the two thirds factor) with
an arbitrary shape correction factor (taken to be equal to 0.85).
The modified B31G equation is given by:

d
t
d 1
1 0.85
t M

1 0.85

(9a)

where

= SMYS + 68.95 MPa (10 ksi)


M=

1 0.6275

Dt

0.003375

2c

Dt

Dt

2c

Dt

for

2c

2c

Dt

3.3

M = 0.032

2c

for

50

(13)

50

(14)

Both the original B31G criterion and the modified B31G criterion define simple approximations to the
exact corroded area, based on the maximum length and the maximum depth of the defect. In addition, the
modified B31G criterion describes a method based on the effective area and effective length of the defect,
incorporated into software known as RSTRENG. The RSTRENG method, based on an iterative
algorithm, was developed to allow the actual profile of the corrosion defect to be considered, thereby
giving more accurate predictions of the failure pressure of the corrosion defect. The modified B31G
method, including RSTRENG, has been validated against 86 burst tests on pipe containing real
corrosion defects.

Penspen Group

Rev 2004/1

The effective area method is incorporated into software referred to as RSTRENG. For simple hand
calculations the geometric shape approximation is recommended (this is the arbitrary assumption that
the defect area is 0.85dL).
The RSTRENG method is based upon the effective area and effective length of the corrosion defect. The
shape of the corrosion defects is not important, all that is required is a profile of the defect (i.e. a series
of length and depth measurements along the length of the defect). Determining the corrosion defect
profile requires a large number of depth measurements to be taken at regular intervals along the length of
the defect. The defect can be a single defect or a composite defect formed through defect interaction. The
procedure is based upon considering various subsections of the total defect profile, and predicting the
corresponding failure pressure. This process is repeated for all possible combinations of the various
subsections. In most cases, although not all, RSTRENG predicts a minimum failure pressure that is less
than the value predicted using the exact area, total length method[27].
The modified B31G equation has been incorporated into software referred to as RSTRENG, which
allows the assessment of irregular-shaped corrosion defects. The RSTRENG method is permitted as an
approach for use within a new (proposed) ANSI/ASME B31.8 fitness-for-purpose addendum, as is the
use of the finite element stress analysis method[28].
5.3.7.3 Line Pipe Corrosion Group Sponsored Project
New methods for the assessment of corrosion defects under internal pressure loading have been
developed through a Group Sponsored Project, undertaken by BG Technology (formerly British Gas).
Over 70 full scale tests on single, interacting and complex shaped defects (specifically, pits in areas of
general corrosion) and a large number of three-dimensional, non-linear, elastic plastic finite element
analyses, were carried out in the course of the development and validation of the assessment methods[29].
The project has produced guidance for the assessment of single defects and interacting defects, and a
method for assessing the actual shape of a corrosion defect. The underlying failure model has the same
form as the original Battelle part wall failure criterion (as incorporated into the B31G and modified
B31G methods), but the geometry correction factor (i.e. the Folias factor) has been modified). In
addition, the flow stress has been defined in terms of the ultimate tensile strength of the linepipe steel,
rather then in terms of the yield strength, based on an analytical model of the ultimate pressure of a fully
restrained pipe containing long corrosion patches[14].
The single defect equation is given by:

d
1 t
f

1 d 1

t M

(5)

where
M

1 0 .31

Dt

(15)

= 0.9UTS
The method for taking into account the actual profile of the corrosion defect is, like RSTERNG, an
iterative procedure, but based on the principle of considering the actual profile as a collection of pits
within patches. The corrosion defect is divided into a number of increments, based on depth, and

Penspen Group

Rev 2004/1

modelled by an idealised patch containing a number of idealised pits. The assessment method
determines whether the defect behaves as a single irregular patch, or whether local pits within the
patch dominate. Potential interaction between the pits is also assessed. The failure pressure is taken as
the minimum failure pressure from the analysis of all of the depth increments.
The methods developed from this project, together with those from a DNV project (see below), have
been incorporated into a DNV Recommended Practice, RP-F101[30].
5.3.7.4 DNV Joint Industry Project
A Joint Industry Project undertaken by Det Norske Veritas (DNV) has also produced guidelines for the
assessment of corrosion defects, but considering axial and bending loads in addition to internal
pressure[31]. In the course of the project, 12 full scale tests on axial and circumferential single defects
subjected to internal pressure, and axial and bending loads were carried out, together with a large
number of three-dimensional, non-linear, elastic plastic finite element analyses.
The results of the DNV project were guidance for assessing single corrosion defects under both internal
pressure and combined loading. In addition, a probabilistic calibration exercise was undertaken to
produce partial safety factors to be used with the assessment method. The intention of providing partial
safety factors, rather than a single safety factor, was to give a more consistent level of safety over a wide
range of defect sizes and pipeline geometries.
The methods developed from this project, together with those from a BG project (see above), have been
incorporated into a DNV Recommended Practice, RP-F101[30]. The basic equation is:

Pf 0.9 U

d
t
d 1

1
t Q

2t
D t

Q 1 0.31

(16)

(17)

Dt

5.3.7.5 Summary of Methods


METHOD
Original

FLOW
STRESS

Battelle
(5)

SMYS
69MPa

Battelle

B31.G

(9)

Modified
B31.G
RSTRENG
DNV

BASIC
EQ.
(No)

Battelle
(9a)
Battelle
(9a)

RP- Battelle

Penspen Group

1.1SMYS

DEFECT
SHAPE
+

FOLIAS FACTOR

Defect Area

1 0.8( 2c /

Dt ) 2

Parabolic
[2/3(d/t)]

1 0.8( 2c /

Dt ) 2

SMYS
69MPa

+ Arbitrary
[0.85(d/t)]

1 0.6275

2c

Dt

0.003375

2c

Dt

SMYS
69MPa

+ Defect
Profile

1 0.6275

2c

Dt

0.003375

2c

Dt

0.9SMTS

Rectangular

1 0.31( 2c /

Rev 2004/1

Dt ) 2

F101

(16)

(and Defect
Profile)

The main methods (ASMEB31G, modified B31G and DNV are given graphically in Figure 6.

Penspen Group

Rev 2004/1

1.2

d
1
f
t

d 1

1
t M

Original
Battelle
(NG18)
Equation

F a ilu re S tre s s /Yie ld Stre n g th

0.8

0.6

0.4

0.2

1.0

0.9

0.8
0.7
1.0

0.6

0.9

REJECT

0.4
0.3
0.2

ACCEPT

0.1
0.0
0

Defect Length, 2c/(Rt)^0.5

ASME B31G

0.8

1.0

0.7

0.9

0.6

0.8

0.5

D e fe c t D e p th /W a ll T h ic k n e s s

0.5

D e fe c t D e p th /W a ll T h ic k n e s s

D e fe c t D e p th /W a ll T h ic k n e s s

2c/(Rt)^0.5

0.7

0.4

0.6

0.3
8

0.5

0.2

0.4

0.1

0.3

0.0
0

2c/(Rt)^0.5

0.2
0.1

Mod ASME B31G

0.0
3

2c/(Rt)^0.5

DNV
Figure 6. Schematics of the Main Corrosion Assessment Methods

Penspen Group

Rev 2004/1

5.3.8 Girth Welds


Pipeline girth welds have a good operating record and are not a major cause of pipeline failures [4,32].
However, defects are sometimes detected during service, and they require assessment. Additionally,
operators may wish to set different defect acceptance levels (based on fitness-for-purpose, rather than
workmanship limits) at the construction stage of their pipeline.
Girth welds are fabricated to stringent standards (e.g. API 1104, BSI4515, CSA Z184 [33-36]). These
standards contain acceptance levels for defects based on workmanship considerations and fitness for
purpose criteria. A review[33] conducted on behalf of the European Pipeline Research Group (EPRG)
highlighted the wide differences in girth weld acceptance levels in various company and national pipeline
welding codes. This applies to both the workmanship limits and the fitness for purpose criteria.
Workmanship standards are by their nature subjective so the difference is unsurprising. Fitness for
purpose criteria are based on structural analysis and should result in more consistent limits. However
different approaches have been adopted in various standards giving rise to different limits.
The standards require the repair of non-planar defects such as slag and porosity which are generally
accepted as innocuous[35], although they may mask the presence of more serious planar flaws. Many
planar defects are also considered to be unacceptable despite full scale tests which show high failure
stresses even under the most severe loadings at low temperatures[37].
In the absence of universally accepted defect acceptance criteria and the stringent requirements in
standards regarding otherwise innocuous defects, the EPRG produced a set of independent guidelines for
the assessment of girth weld defects[33-36]. The guidelines consist of three tiers. Tier 1 is workmanship
limits, derived from existing welding standards. The Tier 2 limits are semi-empirical, the workmanship
limits are extended based on the results of wide plate tests. For both Tier 1 and Tier 2 a minimum
Charpy toughness is specified. For Tier 3, fitness for purpose failure criteria have been used to derive
defect limits which have been validated by full scale tests. Consequently the defect limits are semiempirical. A minimum Charpy toughness and a minimum CTOD toughness is required for Tier 3.
The EPRG guidelines can be used on new pipelines (to set weld defect acceptance levels), but can also
be applied to the assessment of defects in existing pipelines[4,34]. The reader is directed towards these
guidelines and their background literature[33-36].
For girth welds and defects outside the scope of the EPRG guidelines more advanced assessment
methods are available, and guidance is given in BS 7910[9].
5.3.9 Stress Corrosion Cracking
Stress corrosion cracking has been well known for many years in pipelines. The older type of stress
corrosion cracking (high pH) has been studied in various research programmes, and some guidance on
its assessment is given in Reference 28, based on work reported in Reference 38.
In recent years, there have been an increasing number of pipeline failures caused by a different type of
stress corrosion cracking[39]. This is a 'new' type of SCC that attacks the external surface of the pipeline,
but it can occur at relatively low temperatures, and in environments with relatively low pH (near neutral)
values[39]. No reliable fracture mechanics model has yet been developed, but the basic fracture mechanics
model reported in References 28 and 38 is likely to provide the basis.

Penspen Group

Rev 2004/1

5.3.10 Material Defects


Defects in the parent plate of pipe steels (e.g. laps. slivers, laminations, etc.) can be assessed using the
methods outlined in Section 5.3.2, if they result in a loss in the cross-sectional area of the pipe. If they do
not cause a reduction in the cross-sectional area (e.g. a planar, radial-orientated lamination), they will
not cause a reduction in the pressure containment of the pipe. However, such defects can cause problems
when intelligent pigs are run, by producing confusing results.
5.3.11 Construction Defects
Construction defects will generally be gouges, or perhaps dents. The assessment methods detailed in
Sections 5.3.2 - 5.3.4 above can be used to assess these construction defects. However, it should be
emphasised that construction defects detected prior to operation can be easily repaired, and may be more
of a contractual/legal issue than a fitness-for-purpose one.
5.3.12 Other Fitness-For- Purpose Methods For Transmission Pipelines.
5.3.12.1 Preventing Dynamic Fracture
Running fractures - both ductile & brittle - are a risk in gas and multiphase pipelines. Consequently,
design codes specify toughness requirements, based on impact tests (DWTT and C v), for the pipeline
material. These requirements are based on full scale testing in both North America and Europe. The
simple Cv test remains the definitive test for fracture propagation. Guidance on this type of assessment is
given elsewhere in the course notes and in References 40 and 41.
5.3.12.2 Fatigue Calculations
Pipelines experience changes in pressure due to varying customer demand, or differing batch flows. This
means the pipeline can experiencing fatigue cracking. Loss of wall thickness defects can be assessed for
fatigue using classic fracture mechanics calculations[4 ,9, 42].
There are two basic approaches that can be adopted to determine the fatigue life of an engineering
structure, either using S-N curves, or fracture mechanics.
Fatigue life predictions using S-N curves are based on experimental tests on specimens of similar
materials, geometries, types of weld, etc. In the case of welded structures, a large number of fatigue
tests have been carried out on a wide variety of different types of weld to produce curves relating the
fatigue life to the applied cyclic load. Because S-N curves have been derived from experimental results
which demonstrate scatter, they are typically presented with varying degrees of statistical confidence.
Design curves are typically presented as the mean minus two standard deviations.
In order to assess the fatigue life of a structure using fracture mechanics based methods it is necessary to
know the initial defect size, the final defect size (i.e. the limiting defect size), and a fatigue crack growth
law. The final defect size may be calculated using an appropriate failure criterion and the maximum
operating load. The initial defect size may have been obtained from the results of an inspection, or be an
upper bound estimate based on a known prior proof load (for example).
The fatigue crack growth law defines how the defect grows from the initial defect size to the final defect
size. It relates the rate of crack growth to the applied cyclic stress intensity factor. The most commonly
used fatigue crack growth law is the Paris Law[43], which takes the following form:
da
m
C K
dN

Penspen Group

(18)

Rev 2004/1

K Y

(19)

where:
N = fatigue life
a = defect size
= applied cyclic stress
K = applied cyclic stress intensity factor
Y = compliance factor, a function of the defect size and the geometry of the structure
C,m = fatigue crack growth constants
The constants C and m are defined experimentally. They depend upon the material and the environment.
A plot of da/dN against K will normally be a sigmoidal curve (if log-log scales are used). The Paris
Law only applies to the central linear portion of the curve. For low values of K the rate of crack
growth may decrease rapidly if there is a threshold stress intensity factor below which fatigue crack
growth will not occur. As the defect approaches failure, the rate of crack growth increases rapidly. The
use of the simple Paris Law is acceptable under most circumstances. It is conservative to neglect
threshold effects.
More complicated fatigue crack growth laws are available which can model the whole of the fatigue
crack growth regime, and which incorporate mean stress effects[44].
5.4

Constructing Defect Assessment Plots.

Once you have decided on your assessment method (which will be governed by the defects in question),
the basic equations are used to construct acceptance curves.
Figure 6 shows such curves. This figure shows that a number of gouges (square points) have been
detected on a pipeline. The gouges measurements are all maximums. Using Equation 5, two curves are
constructed. The first one calculates the failure stress of gouges in the pipeline at the maximum
operating pressure (MAOP), and the other curve shows the size of gouges that would fail at the preservice hydrotest level for this pipeline.
As can be seen, only one defect is predicted to fail at the MAOP. This defect would need to be repaired.
In Figure 7, the engineer conducting the assessment, has decided that a reasonable safety factor for this
pipeline could be based on the hydrotest, i.e. they have decided that any gouge that has a failure stress
above that at hydrotest is acceptable. Any defect with a lower failure stress, but one above the failure
stress at MAOP, would be considered for reassessment using more sophisticated methods, or more
accurate defect measurements.
These type of plots can be constructed for gouges, corrosion (taking into account corrosion growth [4, 45]),
etc.. This subject is covered in more detail elsewhere in the course notes.

Penspen Group

Rev 2004/1

7. Defect
Assessment
Chart
FIGURE 3. Figure
EXAMPLE
OF DEFECT
ASSESSMENT
PLOT
12in x 9.5mm, X52 Pipeline (MAOP = 104 bar = 46% SMYS)
DEFECT DEPTH / WALL THICKNESS (nom)
1

0.8

0.6

REPAIR
FITNESS-FOR-PURPOSE LINE
BASED ON MAOP

0.4

REASSESS
FITNESS-FOR-PURPOSE LINE
BASED ON HYDROTEST

0.2

ACCEPT
0
0

100

200

300

400

500

DEFECT LENGTH, mm

6.

NEW STRUCTURAL RELIABILITY METHODS - LIMIT STATE


ANALYSIS

A fitness-for-purpose assessment based on a deterministic approach does not quantify the safety of the
pipeline. Instead, it needs to impose arbitrary safety factors, and use conservative methods, to ensure
that a 'safe' pipeline will emerge. A major problem with fitness-for-purpose methods is setting the safety
factor, and agreeing this factor with the operator and regulatory authorities. In many ways, the safety
factor is better termed an ignorance factor; the larger the safety factor used, the greater the
ignorance[11].
It is now possible to quantify the safety of a pipeline using reliability-based limit state design (also
known as load and resistance factor design, LRFD [11, 12, 46-51]).
The term Limit State Design used in the pipeline industry is misleading as it describes the application
of two distinct concepts to a pipeline design, i.e. limit state (structural engineering) analysis and
reliability theory. This Section introduces the reader to this leading-edge technology.
6.1

Reliability-Based Limit State Design

The strength of the pipeline is calculated, using limit load (or 'state') analysis. Limit 'state' has been
defined in various ways; in the draft European Standard prEN 1594 [48] for gas pipelines it is defined as 'a
condition in which the pipeline can be considered unserviceable or unsafe'. It is the theoretical
relationship between pipeline parameters that exists at the onset of failure. A separate limit state is

Penspen Group

Rev 2004/1

required for each credible failure mode. Limit state analysis requires sophisticated, state-of-the-art
plastic collapse models and elastic-plastic fracture mechanics models. Consequently, limit state
calculation methods can go beyond simple elastic stress analysis, and into strain-based design. However,
the fitness-for-purpose methods detailed in Section 5 above, all constitute limit state models for
individual failure modes.
Limit state design relates a load parameter, L, to a structural resistance factor, R. If both L and R were
deterministic, failure would occur if L>R, and the structure would not fail if L<R. This can be visualised
as an operating hoop stress in a pipelines of 72% SMYS (=L) and a material yield strength of SMYS
(=R).
However, all parameters are variables - pipeline pressure, material yield strength, wall thickness (t),
diameter (D), etc., all vary. Therefore, if we assume that a pipeline will fail due to internal pressure (P R)
when the yield strength (Y) of the material is reached:
PR = 2Yt/D

(14)

we need to know how all the parameters P R, Y, t, and D vary, i.e. we need to know their distributions. If
we know their distribution, we can calculate the probability of failure.
Figure 8[46] shows how the load on the structure (L) is a variable, and the resistance (R) to failure of the
structure is a variable. Failure occurs when the two distributions overlap.
Limit state design is covered elsewhere in this course, and hence is not dealt with in this section of the
course notes. The course attendee is directed towards the course notes on limit state design or the
literature[46-52].

Penspen Group

Rev 2004/1

Figure 8. Standard Load-Resistance Curve (39)


FIGURE 4. STANDARD LOAD-RESISTANCE DIAGRAM (39)

Mean Safety Margin

L = R

LOAD (L)
DISTRIBUTION

Nominal Safety
Margin,

Mean Resistance

Nominal Load
e.g. 80% SMYS

Nominal Resistance
e.g. SMYS

Partial Safety Partial Safety


Factor,
Factor,

Mean Load

PROBABILITY

RESISTANCE (R)
DISTRIBUTION

Overlap is
failure
probability

The nominal load and resistance values are selected using probabilistic criteria,
e.g. the nominal resistance is defined as the strength below which not more than
5% of all possible test values would fall (referred to as the 5% fractile). The nominal load is
often defined as the mean load (for fixed loading) or a specified annual probability of
exceedance (for variable loads), e.g. 100 years.
Once the nominal loads and resistances are defined (they will be dependent on
the variability of loads and resistances and the model uncertainties), the partial
safety factors are used to achieve a sufficient separation between the load and resistance
probability distribution, to keep the probability of failure below an acceptable target value.

7.

BEST PRACTICES

The pipeline industry has used their fitness for purpose methods to produce generic guidelines for the
assessment of defects in pipelines. These methods and guidelines are based on pioneering work at
Battelle Memorial Institute in the USA on behalf of the Pipeline Research Council International (6,53,54)
with the more recent additions of ad hoc guidelines for the assessment of girth weld defects, mechanical
damage and ductile fracture propagation produced by the European Pipeline Research Group (36,55,56).
Best practices in structural assessments of defects in pipelines are now emerging (e.g. 57-59), although
the recognised generic publications (9, 60) are still applicable.
A Joint Industry Project sponsored by 14 major oil and gas companies produced a state-of-the-art
Pipeline Defect Assessment Manual in 2003 (59, 61-63).

Penspen Group

Rev 2004/1

8.

CONCLUSIONS

1.

Fitness-for-purpose methods are available for a wide variety of defects in transmission pipelines.
Some of these methods have been available, and tried-and-tested, for many years, and their use
should be encouraged.

2.

There are five levels (or stages) of defect assessment, ranging from simple methods detailed in
current codes, to highly sophisticated risk analyses, and limit state methods. A very high level of
expertise is required on the higher levels.

3.

The defect assessment will only be as reliable and as accurate as the method used, and the data that
is available.

4.

Corrosion in pipelines can be reliably assessed, but account should be taken of the accuracy and
reliability of the inspection data, and any future corrosion growth.

5.

External interference can result in severe defects such as combined dents and gouges. These type of
defects must be assessed with caution, as their behaviour with time (i.e. fatigue) is not wellunderstood, and dents associated with welds are very difficult to assess.

6.

Defect assessment in pipelines is moving into reliability-based limit state design and risk analysis.
These methods will allow even more sophisticated assessment, providing accurate data is available.

9.

RECOMMENDATIONS

This paper will end by giving engineers general advice on assessing defects in pipelines (67):

1. ALWAYS THINK SAFETY Pipeline codes are safety standards, and an engineers
prime role in any industry is to ensure safety. Structural assessments are safety
assessments.
2. PIPELINE DESIGN CODES ARE DAY 1 CODES Pipeline design codes give us
minimum structural integrity requirements they are the starting point for integrity, not
the end point. A pipeline designed and built to code will be very safe on its first day in
service. However, after day 1 the pipelines management dictates its safety, not the
design code. Therefore, good managers and good management systems4 are the key to
pipeline structural integrity. This means a continual appraisal of technical issues such as
smart pigging, risk management programmes, correct routeing etc., to achieve high
integrity. But remember - only good management will guarantee integrity
3. PIPELINES ARE SAFE IF WE ADOPT A HOLISTIC APPROACH Pipelines fail for
a reason, and most of these reasons can either be avoided or mitigated by good pipeline
management. Structural assessments must address all elements of the problem including
consequences of failure and the management systems that may affect the assessment or
need to be in place to implement any recommendation. Therefore, pipeline integrity
management must consider all aspects of our pipeline system, as it is an integrated
process, where all elements affect safety.
4. DO NOT DO THE MINIMUM - Codes. Regulations, etc., are minimum
requirements. Aim to do more than your peers. It makes sense if you do only one
thing more than your peers, his/her pipeline has a higher probability of failing first. You
4

Management Systems are covered References 64 and 65.

Penspen Group

Rev 2004/1

will learn from their experience, and act accordingly, so it does not happen to you. They
will then copy your practices, but you will then move one step ahead again and then
their pipeline will again fail before yours!
5. DO NOT AUTOMATICALLY SELECT THE LOWEST BID IN ENGINEERING - A
trained monkey can select the smallest of three objects. An engineer can select the safest
and best.
6. CHANGE IS DIFFICULT It is difficult to change pipeline engineering practices in
companies, because it is difficult to change people. However, as pipelines age we must
continually change our approach to pipeline structural integrity.
7. REGULATIONS AND COMPANY PRIORITY - Your company/work will be limited
by both company codes/practices and your national/state legislation/regulation. These
may not allow fitness for purpose assessments, or the defects you consider acceptable.
You must always check local and national limits and legislation, before applying fitness
for purpose:
a. can I apply it, and
b. can I use the results?
8. CONSEQUENCES - Always appreciate the consequences of your assessment. If I have
made an error, if the defect measurement is wrong, etc., and the pipeline fails, what are
the consequences? Consequences will dictate your safety margin on your calculations.
9. NEVER ASSESS A MESS There will be some situations where an engineer is asked
to assess the structural significance of defects, where the defects are extensive, or the
structure itself is in poor condition. Sometimes even the best surgeon has to pronounce
the patient dead.
10. CALCULATIONS ARE NOT ENGINEERING - Calculations convey the thought
process and design intent, and are an essential part of any engineering appraisal. The
quality, etc. of calculations indicates the level of care and diligence; however,
calculations substantiate, but do not substitute, for judgement.
11. ENGINEERING RESPONSIBILITY - It is your responsibility to ensure that any
fitness for purpose assessment is correct. Try and use the best possible practices
available. Check calculations, inputs and assumptions. Use all available data (historical,
current and circumstantial) when you assess a defect inspection data, operations
records, maps, etc., may give you more information about the defect, its cause, etc..
12. PROFESSIONAL AND PERSONAL INTEGRITY Engineers performing structural
assessments, where decisions can effect company profits, career progression,
environment, etc., must put safety first, and remember that engineers have professional
(institutional requirements), legal (standard of care) and ethical (moral) criteria to also
satisfy in all aspects of their work.
ACKNOWLEDGEMENTS
The author would like to thank colleagues at Penspen and Andrew Palmer & Associates.

Penspen Group

Rev 2004/1

10.

REFERENCES

1.

Anon., 'Interstate natural gas pipelines - Delivering energy safely', Interstate Natural Gas
Association of America Report, USA, 1994.

2.

Hopkins, P., 'Transmission pipelines: How to improve their integrity and prevent failures', 2nd Int.
Conference. on Pipeline Technology, Ostende, Belgium, 1995.

3.

Fearnehough, G.D., 'The control of risk in gas transmission pipelines', Institution of Chemical
Engineers, Symposium Series No. 93, Manchester, UK, 1995.

4.

Hopkins, P., Haswell, J., The Practical Assessment Methods for Application to UK Gas
Transmission Pipelines, The Institute of Materials 2 nd Griffith Conference, Sheffield, UK,
September 1995.

5.

Hopkins, P., Ensuring the Safe Operation of Older Pipelines, Int. Pipeline and Offshore
Contractors Association, 28th Convention, Acapulco, Mexico, Sept. 1994.

6.

Kiefner, J. F. et al., 'Failure Stress Levels of Flaws in Pressurised Cylinders', ASTM STP 536, pp
461-481, 1973.

7.

Shannon, R. W. E., 1974, 'The Failure Behaviour of Linepipe Defects', Int. J Press Vessel & Piping,
(2), pp 243-255.

8.

Miller, A. G., 'Review of Limit Loads of Structures Containing Defects', Int. J of Pressure Vessels
& Piping, (32), nos1-4, p195, 1988.

9.

Anon., 'Guidance on Methods for the Derivation of Defect Acceptance Levels in Fusion Welds', BSI
7910, British Standards Institution, London, 1999. Formerly: Anon., 'Guidance on Methods for the
Derivation of Defect Acceptance Levels in Fusion Welds', BSI PD6493, British Standards
Institution, London, 1991

10. Hopkins, P., Hopkins, H., Corder, I., The Design and Location of Gas Transmission Pipelines
Using Risk Analysis Techniques, Risk and Reliability Conference Aberdeen, UK, May 1996.
11. Hopkins, P., New Design Methods for Quantifying and Reducing the Number of Leaks in Offshore
and Onshore Transmission Pipelines, European Pipeline Leakage Prevention Conference, IChemE,
London, May 1997.
12. Hopkins, P., Haswell, J., The Practical Application of Structural Reliability Theory and Limit State
Concepts to New and In-service Transmission Pipelines, Int. Seminar on Industrial Applications of
Structural Reliability Theory, ESReDA, Paris, France, October 1997.
13. Anon., ASME Boiler and Pressure Vessel Code Section VIII, Unfired Pressure Vessel Division 2,
1980.
14. Stewart, G., Klever, F., Ritchie, D., An Analytical Model to predict the Burst capacity of
Pipelines, Offshore Mechanics and Arctic Engineering, Houston, Texas, 1994.
15. Kastner,W., Rohrich,E., Schmitt,W. and Steinbuch,R.; Critical Crack Sizes In Ductile Piping, Int.
Journal Press. Vess. and Piping, (9), 1981, pp197-219.
16. Hopkins, P., Jones, D. G., Clyne, A. C., 'The Significance of Dents and Defects in Transmission
Pipelines', Conference. on 'Pipework Engineering and Operation', IMechE, London, Paper
C376/049, 1989.

Penspen Group

Rev 2004/1

17. Hopkins, P., Corder, I., Corbin, P.; The Resistance of Gas Transmission Pipelines to Mechanical
Damage, Paper VIII-3, International Conference on Pipeline Reliability, Calgary, Canada, June
1992.
18. Jones, D.G.; The Significance of Mechanical Damage in Pipelines, 3R International, 21, Jahrgang,
Heft, 7, July 1982.
19. Fowler, J.R.; Criteria for Dent Acceptability in Offshore Pipelines, OTC 7311, 25 th Offshore
Technology Conference, Houston, Texas, 3rd-6th, pp 481-493, May 1993.
20. Fowler, J.R. Alexander, C.R., Kovach, P.J., Connelly, L.M.; Fatigue Life of Pipelines with Dents
and Gouges Subjected to Cyclic Internal Pressure, PD-Vol. 69, Pipeline Engineering, ASME 1995.
21. Fowler, J.R. Alexander, C.R., Kovach, P.J., Connelly, L.M.; Cyclic Pressure Fatigue Life of
Pipelines with Plain Dents, Dents with Gouges, and Dents with Welds, AGA Pipeline Research
Committee, Report PR-201-927 and PR-201-9324, June 1994.
22. Wang, K.C., Smith, E.D., The Effect of Mechanical Damage on Fracture Initiation in Linepipe
Part I - Dents, CANMET, Canada, Report ERP/PMRL 82-11 (TR), January 1982.
23. Anon., Rules for Submarine Pipeline Systems, Det Norske Veritas (DNV) , December 1996.
24. Hopkins, P., 'The Application of Fitness for Purpose Methods to Defects Detected in Offshore
Transmission Pipelines', Conference. on 'Welding and Weld Performance in the Process Industry',
London., 1992.
25. Kiefner, J.F., Duffy, A.R.; Criteria for Determining the Remaining Strength of Corroded Areas of
Gas Transmission Pipelines, AGA Operating Section on Transmission, Conference, AGA 1973.
26. Anon., 'Manual for determining the Remaining Strength of Corroded Pipelines', ANSI/ASME B31
G-1984, 1984.
27. Kiefner,J.F., Vieth,P.H.; A Modified Criterion for Evaluating the Strength of Corroded Pipe, Final
Report for Project PR 3-805 to the Pipeline Supervisory Committee of the American Gas
Association, Battelle, Ohio, 1989.
28. Rosenfield,M.J., Kiefner,J.F.; Proposed Fitness-for-Purpose Appendix to the ASME B31.8 Code
for Pressure Piping - Section B31.8, Gas Transmission and Distribution Systems, Kiefner and
Associates, Worthington, Ohio, Jan. 1995.
29. Kirkwood, M., Senior, S., Bin, Fu., 'Improved Guidance for Assessing the Integrity of Corroded
Pipelines', ASME/JSME Pressure Vessel and Piping Conference, Honolulu, Hawaii, 1995.
30. Anon, Corroded Pipelines, DNV Recommended Practice RP-FP101, Det Norske Veritas, 1999.
31. Bjrny,O.H., Cramer,E.H., and Sigurdsson,G.; Probabilistic Calibrated Design Equation for Burst
Strength Assessment of Corroded Pipes, International Conference on Offshore and Polar
Engineering, ISOPE97, Honolulu, USA, 1997.
32. Vieth,P.H.; Analysis of DOT Reportable Incidents for Gas Transmission and Gathering Pipelines,
Tenth European Pipeline Research Group/American Gas Association Seminar, Cambridge, UK,
April 1995.
33. Roodbergen, A. H., Denys, R., 'Limitations of Fitness for Purpose Assessments of Pipeline Girth
Welds', Int. Conference. on Pipeline Technology, AIM, Italy, 1987.
34. Hopkins, P., 'The European Pipeline Research Group's Guidelines on Acceptable Girth Weld Defects
in Transmission Pipelines', 8th Symposium on Linepipe Research, American Gas Association,
Houston Texas, 1993.

Penspen Group

Rev 2004/1

35. McDonald, K., Hopkins, P., The significance of embedded non-planar defects in transmission
pipeline girth welds: a literature review, Pipes and Pipelines International, March-April (Part 1),
May-June (Part 2), 1995.
36. 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, p620-624.
37. Hopkins,P., Pistone,V., and Clyne,A.J.; A Study Of The Behaviour Of Defects In Pipeline Girth
Welds : The Work Of The European Pipeline Research Group, Conf. On Pipeline Reliability,
Calagary, Canada, June 1992.
38. Leis, B. N., Mohan, R., 'Failure Criteria for Axial Flaws', 8th Symposium on Linepipe Research,
American Gas Association Catalogue No L51680, USA, 1993.
39. Delanty, B. S., OBeirne, J., Oil and Gas Journal, June 15, 39, 1992.
40. Vogt, G., et al., EPRG Recommendation for crack arrest toughness for high strength linepipe
steels, 3R International, 34, Jahrgang, Heft 10-11/1995, p607-611.
41. Vogt, G., et al., EPRG Report on toughness for crack arrest in gas transmission pipelines, 3R
International, 22, 1983, p3-10.
42. Hopkins, P., Jones, D.G., The Design of Gas Storage Systems Using Fracture Mechanics, The 6 th
Int. Conference on Fracture, New Delhi, India, 1984.
43. Paris,P., Erdogan,F., A Critical Analysis of Crack Propagation Laws, Journal of Basic
Engineering, Trans, ASME Volume 85, 1963.
44. Irving,P.E., McCartney,L.N.; Prediction of Fatigue Crack Growth Rates: Theory, Mechanisms and
Experimental Results, Met. Sci., 11, 8/9, 1977.
45. Hopkins, P., The Assessment of Pipeline Defects Detected During Pigging Operations, Pipeline
and Pigging and Integrity Monitoring Conference, Aberdeen, UK, 1990.
46. Anon., Draft Appendix C, Oil and Gas Pipeline Systems, Canadian Standards Association, CSA
Z662-94, 1994 (reported in Zimmermann, T., et al Target Reliability Levels for Pipeline Limit
State Design, International Pipeline Conference - Vol. 1, ASME, p.111).
47. Anon., Rules for Submarine Pipeline Systems, Det Norske Veritas, DNV1996, Norway (Reported
in Jiao, G., et al, The Superb Project: Wall Thickness Design Guideline for Pressure Containment
of Offshore Pipelines, Offshore Mechanics and Arctic Engineering Conference, OMAE 1996,
Florence, Italy.
48. Anon., Pipelines for Gas Transmission, European Committee on Standardisation, peEN 1594,
1994 (Draft).
49. Anon., Pipeline Transportation Systems for the Petroleum and Natural Gas Industries,
International Organisation for Standardisation, ISO CD 13623, March 1995, (Draft).
50. Anon., Eisen Voor Stalen Transportleidingsystemen, Netherlands Code NEN 3650, NNI,
September 1992.
51. McKinnon, C., et al, A Case Study of the Britannia Gas Export Line Reliability Based Limit State
Design, Conference on Risk and Reliability and Limit States in Pipeline Design and Operations
Conference, Aberdeen, May, 1996.
52. Palmer, A, et al, Introduction to Limit State Reliability Based Pipeline Design, American Gas
Association, Pipeline Research Committee, Report PR-234-9425, May 1995.

Penspen Group

Rev 2004/1

53. W A Maxey, J F Kiefner, R J Eiber, A R Duffy, Ductile Fracture Initiation, Propagation and Arrest
in Cylindrical Vessels, ASTM STP 514, American Society for Testing and Materials, Philadelphia,
1972, pp. 70-81.
54. Pipeline Research Committee of the American Gas Association. Proceedings of the Symposia on
Line Pipe Research, USA, 1965 onwards. www.prci.com.
55. R Bood, M Galli, U Marewski, P Roovers, M Steiner, M Zarea, EPRG Methods for Assessing the
Tolerance and Resistance of Pipelines to External Damage (Part 2), 3R International, 12/1999, pp.
806-811.
56. G Re, V Pistone, G Vogt, G Demofonti, D G Jones, EPRG Recommendation for Crack Arrest
Toughness for High Strength Line Pipe Steels, Paper 2, Proceedings of the 8th Symposium on Line
Pipe Research, American Gas Association, Houston, Texas, 26-29 September 1993.
57. P. Hopkins, A Cosham, How To Assess Defects In Your Pipelines Using Fitness-For-Purpose
Methods Conference on Advances in Pipeline Technology 97, Dubai, IBC, September 1997.
58. A Cosham, M Kirkwood, Best Practice In Pipeline Defect Assessment Proceedings of IPC 2000,
International Pipeline Conference, October 2000; Calgary, Alberta, Canada, Paper IPC00-0205.
59. A Cosham, P Hopkins, A New Industry Document Detailing Best Practices In Pipeline Defect
Assessment, Fifth International Onshore Pipeline Conference Amsterdam, The Netherlands,
December 2001.
60. Anon., Recommended Practice for Fitness for Service, American Petroleum Institute, USA, API
579, 2000.
61. A. Cosham, P. Hopkins, The Pipeline Defect Assessment Manual, International Pipeline
Conference (IPC 2002), Calgary, September 29 to October 4, 2002.
62. A. Cosham, P. Hopkins, The Assessment Of Corrosion In Pipelines Guidance In The Pipeline
Defect Assessment Manual (PDAM), International Colloquium Reliability of High Pressure Steel
Pipelines, 27-28th March 2003 Prague, Czech Republic.
63. A. Cosham, P. Hopkins, The Effect Of Dents In Pipelines Guidance In The Pipeline Defect
Assessment Manual, Proceedings ICPVT-10, July 7-10, 2003, Vienna, Austria.
64. Anon., Managing System Integrity for Hazardous Liquid Lines, 1st Ed., ANSI/ASME Standard
1160-2001, November 2001.
65. Anon., Managing System Integrity of Gas Pipelines, ASME B31.8S 2001. Supplement to ASME
B31.8, ASME International, New York, USA.
66. Anon., US Department of Transportation. Federal Register Parts 49 CFR 195.452 (May 2001) and
Part 49 CFR 192.763. (January 2003).
67. P. Hopkins, The Structural Integrity of Oil and Gas Pipelines, In Comprehensive Structural
Integrity, Volume 1, Elsevier Science Ltd., UK, 2003.
68. D S Etkin et al. Estimates Of Oil Entering The Marine Environment In The Past Decade:
GESAMP Working Group 32 Project. 1999 International Oil Spill Conference. Paper 166.
69.

Penspen Group

Rev 2004/1

Вам также может понравиться