SAFEBUCK III Design Guideline Rev A E1 PDF
SAFEBUCK III Design Guideline Rev A E1 PDF
SAFEBUCK III Design Guideline Rev A E1 PDF
Design Guideline
SAFEBUCK III
5087471/ 01/ A
Design Guideline
Prepared for
SAFEBUCK III
COMMERCIAL IN CONFIDENCE
TABLE of CONTENTS
1. GENERAL INFORMATION .......................................................................... 1
1.1. BACKGROUND ............................................................................................ 1
1.2. OBJECTIVE OF GUIDELINE ........................................................................ 1
1.3. SCOPE OF DOCUMENT .............................................................................. 3
1.4. STRUCTURE OF GUIDELINE ...................................................................... 3
1.5. DESIGN CODES........................................................................................... 3
1.6. APPLICABILITY ............................................................................................ 4
1.7. DESIGN EXAMPLE ...................................................................................... 7
1.8. NOTATION, ABREVIATIONS AND DEFINITIONS........................................ 7
1.9. UNITS ......................................................................................................... 13
1.10. DISCLAIMER .............................................................................................. 13
2. BACKGROUND TO LATERAL BUCKLING ............................................... 14
2.1. LATERAL BUCKLING ................................................................................. 14
2.2. BUCKLE DEVELOPMENT .......................................................................... 16
2.3. PIPELINE EXPANSION RESPONSE.......................................................... 18
2.4. VIRTUAL ANCHOR SPACING.................................................................... 22
2.5. PIPE WALKING .......................................................................................... 23
3. DESIGN PROCESS .................................................................................... 24
3.1. METHODOLOGY OVERVIEW .................................................................... 24
3.2. SUSCEPTIBILITY TO BUCKLING .............................................................. 25
3.3. ASSESSMENT OF UNCONTROLLED BUCKLING .................................... 26
3.4. ASSESSMENT OF CONTROLLED BUCKLING ......................................... 27
3.5. BUCKLE INTERACTION AND PIPELINE WALKING .................................. 29
4. CALCULATION OF CHARACTERISTIC VAS ............................................ 32
4.1. OVERVIEW OF METHODOLOGY .............................................................. 32
4.2. CHARACTERISTIC VAS: DETERMINISTIC DEFINITION .......................... 32
4.3. CHARACTERISTIC VAS: PROBABILISTIC DEFINITION ........................... 34
4.4. STRUCTURAL RELIABILITY APPROACH ................................................. 35
4.5. BUCKLE FORMATION MODEL .................................................................. 36
4.6. INPUT PROBABILITY DISTRIBUTIONS ..................................................... 42
4.7. PROBABILISTIC SIMULATION .................................................................. 47
4.8. PROBABILISTIC RESULTS........................................................................ 47
4.9. FE ASSESSMENT ...................................................................................... 49
5. DESIGN ANALYSIS ................................................................................... 51
5.1. SYSTEM MODELLING: VAS ANALYSIS .................................................... 51
5.2. SYSTEM MODELLING: INTERACTION AND WALKING ............................ 53
5.3. DESIGN LOADS ......................................................................................... 55
5.4. CLAD AND LINED PIPES ........................................................................... 57
5.5. MECHANICAL PROPERTIES ..................................................................... 58
5.6. PLASTICITY MODELLING .......................................................................... 61
5.7. PIPE SOIL INTERACTION.......................................................................... 66
6. LIMIT STATES ........................................................................................... 74
6.1. APPROACH TO PIPELINE SAFETY .......................................................... 74
6.2. LOCAL BUCKLING ..................................................................................... 75
6.3. FATIGUE AND FRACTURE ........................................................................ 78
6.4. FATIGUE: S-N APPROACH ....................................................................... 80
6.5. ECA ............................................................................................................ 85
1. GENERAL INFORMATION
1.1. BACKGROUND
This document considers the design of on-bottom submarine pipeline systems that are
susceptible to Euler buckling in the horizontal plane (lateral buckling), as well as the
associated pipeline walking phenomena.
Lateral buckling can be a safe and effective way to accommodate the thermal expansion of a
hot pipeline. However there is considerable uncertainty in the initial buckle formation
process and the loads developed in the buckles. High stresses and strains can develop in
the buckles and a conventional stress based design approach is generally not suited to the
design of a laterally buckling pipeline. Consequently, within this guideline the conventional
stress (or moment) limit is relaxed and replaced by a strain limit. However, in doing so the
design must address the following issues:-
1. The amount of feed-in accommodated by each lateral buckle must not be excessive, so
that the bending strains and deformations at the lateral buckle are tolerable. The
following failure modes or limit states must be avoided:
a) Local buckling or wrinkling;
b) Fracture at girth welds, or;
c) Fatigue failure due to low cycle loading associated with operating cycles involving
bending and partial straightening of the line.
2. If a buckle fails to form at an intended location, this not only results in much higher axial
force, but it can also lead to:
a) Excessive feed-in displacement into another buckle, and any of the failure modes
associated with this (under item 1 above), or;
b) Excessive expansion displacements at the ends of the line, which could
overstress the end spools.
3. Unplanned (Rogue) buckles could:
a) Lead to damage of the pipe (item 1 above) if the conditions for lateral buckling
are less favourable, for example due to weight coating or other restraint to free
lateral movements of the pipe, or;
b) Induce a buckle formation failure (item 2 above) by reducing the axial load in the
line.
In theory a single analysis of the entire pipeline (e.g. a finite-element simulation) could
demonstrate that all the above failure modes will be avoided, but in practice there is
considerable uncertainty (and spatial variation) regarding matters such as pipe-soil
interaction, and the as-laid out-of-straightness of the pipe. Furthermore choosing an input
parameter to be conservative for one failure mode can make it non-conservative for another.
Consequently, a more robust design strategy is required. This is the subject of this design
guideline.
BOEMRE
1.6. APPLICABILITY
In developing the guideline a range of pipeline parameters and design scenarios have been
considered. Strictly the guideline is therefore developed from, and is valid within, the set of
parameters outlined here.
The basic approach to the design problem will be applicable to many pipelines that
do not strictly comply with these limitations. However, in this case careful
interpretation of the guideline is required to ensure the inherent assumptions
employed in its development are not compromised and detailed verification of its
applicability is essential.
1.6.4. Fluid
The guideline is intended for pipelines transmitting hydrocarbon or water based products.
Project-specific testing may be required if the guidelines are to be applied to hydrocarbon
pipelines under sour service conditions.
This limitation is driven by the fatigue and fracture limit states, for which the presence
of H2S has a detrimental effect.
The SAFEBUCK JIP has performed extensive testing to investigate the effect of sour
service on fatigue and fracture. This information is included in the guideline as
information. However, the range of tests is insufficient to provide general guidance
and projects should investigate the effect based on their own environment and
loading conditions.
1.6.6. Inspection
The guidance given assumes that the pipeline girth welds are subject to 100% non-
destructive testing. The inspection method (e.g. radiography, manual ultrasonics or
automatic ultrasonics) must be appropriate to the size and type of flaws that need to be
detected, as determined in the ECA.
When defect acceptance criteria based on ECA are employed ultrasonic inspection is
normally required. However, radiography may be sufficient if the acceptable defect
size complies with the workmanship levels defined in the relevant welding
specification.
If there is certainty about where the pipeline will buckle, then it is possible to designate such
areas as „fatigue sensitive zones‟ and apply enhanced inspection methods in those regions
only.
1.6.7. Over-matching
The guidance given assumes that the girth welds are over-matched.
In this context the girth weld strength should be greater than the pipe strength over
the whole strain range of interest and over the whole thermal range of interest. This
is a more demanding requirement than simply over matching the SMYS.
1.8.1. Notation
The following notation is employed in this design guideline.
Roman Symbols
A Constant in SN relationship -
2
Ae External cross sectional area of pipe m
2
Ai Internal cross sectional area of pipe m
2
As Steel cross sectional area of pipe m
-1
Aα Constant in definition of coefficient of thermal expansion °C
C Constant in stress-strain curve -
D Pipeline steel mean diameter m
Deff Effective PIP diameter m
Dfat Fatigue damage ratio -
Di Pipeline steel inside diameter m
Do Pipeline steel outside diameter m
DOC Pipeline overall diameter including external coatings m
e Weld centre-line misalignment m
e Exponential constant -
2
E Young‟s modulus N/ m
EA Pipeline axial stiffness N
2
EI Pipeline bending stiffness Nm
FL Cumulative probability distribution for total length of pipe
F100 Cumulative probability distribution for 100m length of pipe
f Axial pipe soil resistance N/m
fB Bauschinger factor -
fUB Maximum axial pipe soil resistance N/m
fLB Minimum axial pipe soil resistance N/m
fp Strain enhancement factor due to internal overpressure -
f2 API 1111, operational bending safety factor -
g API 1111, collapse reduction factor -
H Height of vertical trigger m
H Lateral pipe soil resistance N/ m
HBE Best estimate lateral breakout resistance N/ m
HLB Minimum lateral breakout resistance N/ m
Greek Symbols
Coefficient of thermal expansion -
αfat Allowable fatigue damage ratio -
δ Pipe ovality, defined in accordance with API 1111. -
Δθ Operating temperature difference (from installed temperature) °C
εb Bending strain -
εc Critical buckling strain -
ε2 Allowable bending strain -
εeq Equivalent strain -
p
εeq Equivalent plastic strain -
εeqN Neuber equivalent strain -
εL Engineering Lüder strain -
p
εH Hoop direction plastic strain -
p
εL Longitudinal direction plastic strain -
p
εR Radial direction plastic strain -
εy Engineering yield strain
-1
κ Maximum bending curvature at buckle crown m
λ Logarithmic strain -
λL Logarithmic Lüder strain -
λp Plastic logarithmic Lüder strain -
λY Logarithmic yield strain -
μA Axial friction factor (or equivalent coefficient) -
μL Lateral friction factor (or equivalent coefficient) -
μmax Maximum axial friction factor -
μXnb Mean value of XNB -
ν Poisson‟s ratio -
2
ζeq Equivalent stress N/ m
2
ζeqN Neuber equivalent stress N/ m
2
ζf Flow stress N/ m
2
ζh Hoop stress N/ m
2
ζLf Longitudinal Flow stress N/ m
2
ζR Nominal axial stress range N/ m
2
ζRL Local axial stress range N/ m
2
ζy Yield stress N/ m
2
ζyc Compressive Yield strength in the hoop direction N/ m
θ Pipeline Temperature °C
θL Pipeline Temperature during installation (when constraint occurs) °C
1.8.2. Abbreviation
The following abbreviations are employed in this design guideline.
1.8.3. Definitions
The following terms are used in this design guideline.
1.9. UNITS
All information is presented in metric units.
Unless stated otherwise, all parameters (stress, strain, force) are tension positive.
The main area where this convention is violated is in the definition of effective axial
force, which is sometimes tension positive and sometimes compression positive. To
clarify the position the symbol S is used when the force is tension positive and N is
used when the force is compression positive.
1.10. DISCLAIMER
All reasonable efforts were made to ensure that the work contained within this guideline
conforms to accepted scientific and engineering practices, but The SAFEBUCK JIP makes
no other representation and gives no warranty with respect to the reliability, accuracy,
validity or fitness of the information presented. Any use or interpretation of the information
contained in the guideline is undertaken at the user‟s own risk.
L
L
Mode 2
Mode 1
Mode 4
Mode 3
L L
i
However Mode 1 buckling is the normal mode of upheaval buckling. A mode 1 shape could only
arise when the lateral deformation is developed from an initial vertical (upheaval buckling)
displacement, which facilitates more concentrated reactions at the touchdown points (perhaps
through local pipe embedment). This has been observed in an operating pipeline, but is unusual.
Some lateral buckles can appear to have a mode 1 shape, particularly when the friction over the
central lobe is very low because it is resting on a sleeper off the seabed, while the outer lobes
become deeply embedded.
ii
In practice, local variations mean that the mode shapes are slightly different to the theoretical
shapes. Nevertheless it is generally possible to categorise actual buckles using these idealised
modes.
3.5 Lateral
Displacement (m)
3
2.5
Full Load
2
Shutdown
1.5
1
0.5
0
-100 -80 -60 -40 -20 -0.5 0 20 40 60 80 100
-1.5
iii
The cyclic stress range can be as high as twice the uniaxial yield without involving cyclic plastic
strain. In reality the maximum elastic stress range is further limited by the Bauschinger effect to a
maximum of 1.4 to 1.6 times the uniaxial yield stress.
Snap
Force
Nmin Remote from
Buckle
In Buckle
A B C B
Amplitude Amplitude
(a) Effect of Imperfection Size (b) Force Remote from and in Buckle
Figure 2.4 Response of Imperfect Pipe
The bold curve is the analytic equilibrium solution for a perfect pipe outlined by Hobbs iv. If a
small imperfection is present in the pipeline, the response is illustrated by Curve A. The
force in the pipeline increases as the system heats up. Very little movement occurs and the
force builds up to a value in excess of the minimum equilibrium force. At some point the
force approaches (but doesn‟t quite reach) the equilibrium curve. At this side of the
minimum force curve, the pipeline is in unstable equilibrium. Any increase in the force in the
pipeline causes a bifurcation and the pipeline snaps through to a stable equilibrium on the
other side of the U shaped curve, as illustrated by the arrow. Further increase in imposed
force causes further (stable) increase in the amplitude of the lateral buckle.
A different response occurs if large imperfections are present in the system; this is illustrated
by curve B or C. Initially the force increases as for curve A, however, initial movement now
occurs at a force below the minimum force. In addition, the displacement is not
accompanied by a snap type bifurcation. Instead the amplitude increases gradually towards
the stable theoretical equilibrium solution.
The force presented in Figure 2.4a is the force outside the buckle at the anchor point. The
variation in force within the buckle is illustrated in Figure 2.4b. The figure shows that the
force within the buckle continues to drop as the amplitude increases. A force difference
develops between the anchor point and the buckle as soon as movement commences, i.e.
the reduction in driving force occurs as soon as the buckle starts to move, irrespective of
snap behaviour. This force difference implies feed-in to the buckle along a slip zone. The
difference between the force within the buckle and at the anchor point continues to increase
– this means that the slip length into the buckle becomes ever longer as the driving force
increases.
iv
This curve describes the post buckle equilibrium configuration for the pipeline. The curve is U-
shaped, and the minimum value on the curve is sometimes called the “safe” force. For forces below
the safe force it is not possible to develop a stable post buckled configuration (that complies with the
inherent assumptions of the analysis).
The formation of the buckle is the key uncertainty in the lateral buckling design. The
buckling process is very sensitive to initial pipe imperfections and interaction with the
seabed; neither of these parameters is known with any certainty. Consequently, it is
impossible to define exactly where, or how many, buckles will form. Within the SAFEBUCK
methodology, a structured approach to buckle formation is adopted to ensure that the design
is sufficiently reliable.
μA
Expansion
Compressive
S Sw pe A e pi A i .... 2.1
For a straight pipeline with fixed ends the axial strain is zero everywhere and the force
developed in the system is known as the fully constrained effective force, which is usually
defined as:-
S 0 SL pi piL A i 1 2 E A s .... 2.2
v
Route changes in the pipeline may modify this force development.
The fully constrained effective force does not depend upon the external pressurevi. Since
pressure and temperaturevii vary along the pipeline length, the fully constrained force also
varies with length (as the pipe cools and the temperature falls the effective axial force
becomes smaller, as shown in Figure 2.5).
For the pipeline with free ends the effective axial force is zero at the ends viii and gradually
increases due to the frictional restraint of the seabed. The slope of the force profile in the
slip zones is defined by the axial frictionix, AW s. At some point the frictional restraint is
sufficient to suppress any expansion and the axial strain in the pipe is zero. Clearly, the
force at this point is the fully constrained effective force defined above.
A similar response occurs at the cold end of the pipeline and the figure shows that the
system develops a fully constrained section in the middle of the pipeline. This is bounded by
points known as virtual anchor points. In the absence of buckling, there is no displacement
over the central section (the pipe is fully constrained) and the pipe expands between the
virtual anchor and the pipe end, reaching a maximum displacement at the pipe end (Figure
2.5b).
X1 X2 X3
Effective Axial Force _
Expansion
Virtual Distance
Lateral Anchors
Buckle
vi
However, the external pressure still acts to develop hoop stress, and the wall force is still calculated
from the effective force through equation 2.1.
vii
Strictly, this is the pressure and temperature difference (the difference between the values during
operation and at installation).
viii
If the pipeline is connected into expansion spoolpieces (or PLETS), there will be a reaction to
expansion that means that the effective force at the end has a small compressive value. If the
pipeline is directly connected to a steel catenary riser, the effective force at the end will be tensile.
ix
This assumes that the friction is fully mobilised. For very small axial slip (or large mobilisation
distances) the slope can be lower than the limiting value, i.e. aW s is not fully attained.
In the example shown the pipeline has formed three buckles (the number of buckles is
entirely case dependant) and the force profile is somewhat more complex. At each buckle
site the axial force drops and the pipe feeds-in from each side to provide the additional pipe
length required to form the buckle. The force in each buckle varies from site to site;
depending upon the initial out-of-straightness, seabed frictional response and buckle
spacing.
The expansion behaviour along the pipeline is illustrated in Figure 2.6b. In all slip zones the
slope of the force profile is the samex – governed by the axial friction. Clearly, between
adjacent buckles there must be a point at which the direction of pipe expansion changes –
this is termed a virtual anchor point. The system is effectively divided into three short
pipelines anchored at each end. These virtual anchors are positions of zero displacementxi
but not zero strain. The VAS associated with each buckle varies, and is labelled X 1 to X3 in
Figure 2.6b.
X1
Effective Axial Force _
Expansion
Compressive
Buckle
x
This is true if the axial resistance does not vary with location. In practice the axial resistance will
vary from point to point, with local soil properties and embedment.
xi
These points are not actually positions of zero displacement. Movement can occur during buckle
development.
A free-ended short pipeline never reaches a position of full constraint; instead the pipeline
forms a virtual anchor at the centrexii and expands from this point. The maximum axial force
in the pipeline can be significantly below the fully constrained force, but can still be sufficient
to cause buckling. In the event of buckling, the pipe will expand into the buckle and out
towards the ends. There must be a point at which the direction of pipe expansion changes –
i.e. a virtual anchor point. The system effectively contains a short pipeline anchored at each
end – as before the virtual anchors are points of zero displacement but not zero strain. The
VAS associated with the buckle is labelled X1 in Figure 2.7.
In the absence of very high end reactions (from the flowline spoolpieces), the maximum
possible buckle spacing for the short pipe is half of the overall length of the flowline (in
practice the force within the buckle normally reduces this maximum spacing to less than half
of the overall length). The worst place for a buckle to form is normally at the centre of the
flowline.
X1 X2
Figure 2.8 Effective Force Profile for Buckles in a Lowly Loaded Pipeline
In this case, although the force in the buckle drops it is a relatively high proportion of the fully
constrained force. The VAS is defined by the intersection of the feed-in zones and the fully
constrained force. Two buckles are illustrated in Figure 2.8, although any number is
possible, and these are isolated by a section of fully constrained, straight pipe. This is the
problem of an isolated buckle in an infinitely long pipeline, originally considered by Hobbs [4].
In this case the VAS cannot exceed the value defined by the fully constrained force profile,
although it still varies with position as illustrated.
xii
The virtual anchor does not necessarily form exactly at the centre; variations in axial friction or spool
piece restraint act to move it away from the pipeline centre.
The Characteristic VAS will be reduced in two ways. Firstly, any significant vertical OOS
may promote buckling. Secondly, significant spans will absorb expansion in a similar way to
engineered triggers. If the reliability of these effects can be demonstrated, then this can be
incorporated into the calculation of Characteristic VAS.
Seabeds with severe bathymetric variation have not been considered in detail in the
SAFEBUCK JIP. However, there is no reason why the design methodology cannot
be applied to these cases.
3. DESIGN PROCESS
3.1. METHODOLOGY OVERVIEW
The design methodology is summarised in Figure 3.1.
Select basic design
parameters
Modify design parameters Modify design parameters
No
No Is pipeline
Is walking OK? susceptible to
buckling?
Yes
Yes
Is
Yes uncontrolled
buckling
acceptable?
No
Yes
Is Can
controlled No initiation No
buckling strategy be
acceptable? improved?
Yes
buckle
No
interaction and
walking OK?
Yes
Design Complete
Once an acceptable lateral bucking strategy has been developed, then buckle interaction
and pipeline walking are evaluated (Section 3.5).
The basic design methodology is applicable to either conceptual or detailed design; the
difference between the phases of design is simply the analysis tools employed and the
relevant acceptance criteria.
The maximum compressive effective axial force in the system is defined as:-
Where:-
L
Nf max Nend fUB .... 3.4
2
The fully constrained effective force (N0) governs for long pipelines and the maximum
available frictional force (Nfmax) governs for short pipelines. If the pipeline is anchored
at both ends, the maximum force in the system is the fully constrained effective force
irrespective of length.
The forces outlined in Equations 3.1 to 3.6 are compression positive. Consequently,
in equation 3.3 the lay tension, NL, would normally be negative.
In equation 3.3 the internal pressure term is the difference in internal pressure
between operation and installation. Under normal circumstances the internal
pressure at installation is surface ambient pressure and is usually neglected.
However, if the pipe is laid flooded then the internal pressure at installation is the
local seabed ambient pressure; this is beneficial in reducing the fully constrained
force, buckle susceptibility and associated end expansion.
The critical buckling force is defined as:-
Ncr min 0.65 N , NcrB
Where:-
EI HLB
N 3.86 .... 3.6
D
Where HLB is the minimum lateral break-out resistance and N∞ is the minimum force under
which a buckle in a straight pipeline can develop; this is derived from the equations of Hobbs
for an infinite mode buckle.
Equation 3.5 defines the critical buckling force associated with any large radius bend
in the pipeline route.
In evaluating Equation 3.5 and 3.6 the minimum pipeline submerged weight should be
employed. This should include an allowance for hydrodynamic lift during storm events.
Storm events may lead to breakout and lateral buckling, particularly if the pipe is not
vertically stable under maximum wave forces. This is a complex dynamic interaction
that requires specific study.
Tentatively it is suggested that the maximum lift force associated with the appropriate
return period significant wave is employed in the design check.
Yes
Max Max
UC <1 Calculate UC >1
Can the VAS UC for each Decrease VAS
be increased? limit state
Max
No UC =1
Within Figure 3.2 there is a check on whether the VAS can be increased. There is an upper
limit on VAS; in long lines this is normally the unconstrained VAS (Section 2.3.4); in short
lines it is related to the pipeline length (normally half the pipe length). If all unity checks are
acceptable at this upper limit then buckling is acceptable and no initiation strategy is
required. This situation may apply over the full length in a lowly loaded pipeline, or over part
of the length for a pipeline whose operating conditions fall with distance.
If the design process is followed this situation should not arise for the rogue buckles,
as this implies that uncontrolled buckling is acceptable (assessed in Section 3.3 ).
However, this situation may apply to the response at the triggers.
The lateral buckling analysis should be undertaken in accordance with the modelling
requirements outlined in Section 5.
The assessment will usually involve an analysis of buckling for a range of VAS. This is
desirable, and allows the sensitivity of the pipeline integrity to buckle spacing to be clearly
demonstrated.
The calculations should be performed at sufficient locations along the pipeline to develop the
variation of Tolerable VAS with KP.
For short pipelines or highly insulated pipelines, the tolerable VAS may not vary
significantly. In this case a single calculation, at inlet conditions would be adequate
The Tolerable VAS must be calculated for a planned buckle and a rogue buckle.
In general these will differ. A rogue buckle is an on-bottom buckle whose capacity is
influenced by the lateral pipe-soil resistance. Most initiation strategies modify the
influence of lateral restraint. As a result the feed-in capacity of the engineered and
the rogue buckle is different, and the Tolerable VAS will be different.
In general the trigger spacing should be no more than the minimum of the engineered buckle
Tolerable VAS or twice the rogue buckle Tolerable VAS.
This is based on the assumption that all the triggers fire. In this case the maximum
feasible VAS for a rogue buckle is half the distance between triggers. This strategy
will be adequate so long as the probability of a trigger failing is low. This is
addressed in the calculation of Characteristic VAS (Section 4).
It is possible that the tolerable VAS at the trigger may be much greater than the
tolerable VAS for rogue buckling. This can occur for initiation techniques that reduce
the loading within the buckle, for example distributed buoyancy. In this situation the
definition of trigger location will be driven entirely by the capacity of rogue buckles.
The location of the triggers may also be influenced by sensitive areas of the pipeline.
For example, the triggers may be employed to prevent buckling at a pipeline crossing
or a mid-line tie-in location.
The selection of trigger locations must ensure that the Tolerable VAS is greater than the
Characteristic VAS at all points along the pipeline.
The aim of the analysis is to confirm the overall expansion behaviour and evaluate
pipeline walking. It is not to assess the acceptability of the design checks for other
failure modes; that is the role of the VAS analysis, Section 3.3 and 3.4. However, if
the analysis highlights undesirable behaviour (for example pull out of a route curve)
then it may be necessary to reassess the results of the VAS analysis.
If the pipeline is restrained to prevent unacceptable walking, for example by
anchoring the flowline, then it is essential to carry out cyclic loading analysis. In
particular the presence of a pipeline anchor can cause route curve pull-out, or buckle
pullout, after the anchor becomes fully loaded over a number of operating cycles;
while end expansion will usually take some further cycles to stabilise.
A suitable design methodology for evaluating cyclic behaviour is outlined in Section 5.2.
In this stage the pipeline walking response should be evaluated. Pipeline walking is not a
limit state; however, the very significant axial displacements involved can lead to
overstressing and subsequent failure of pipeline risers, jumpers or spoolpieces. The
incremental axial displacement associated with each start-up cycle must be evaluated. The
total axial displacement over the life of the pipeline must be calculated based upon this
incremental displacement and the best estimate of the number of start-up/shut-down cycles
anticipated over the design life.
In the assessment of walking it is important that the number of major shut-down
cycles is not overestimated. The number of cycles should be a realistic best
estimate, not an upper bound. Historical information from similar projects can be
used to assist in quantifying the appropriate number of cycles.
The number of cycles need not be the same as that assumed for low cycle fatigue
design.
The behaviour of the pipeline is acceptable if:-
The total axial displacement over the life of the pipeline is within the design capacity of
any pipeline connections;
Pipeline connections include off-line tie-ins via jumpers or spools, or in-line
connections such as to a SCR.
The assessment must also confirm the integrity of any mechanical connectors under
the extreme displacement and cyclic loading.
If the number of cycles is large, then very low walking rates per cycle could lead to
very significant movement over the life of field. However, pipeline walking is not well
understood and there is limited operational verification of the design predictions.
When the predicted walking rate is very low, the model uncertainty means there is
little confidence in the prediction. Tentatively, if the walking rate is below 5 mm per
cycle no further assessment is required. However, this may be revised as operational
experience grows.
The loads within any lateral buckle remain sustainable;
Route curve pull-out does not occur.
Some movement of the route curves may be acceptable, so long as the walking
response is tolerable and the route curve has stabilised within a reasonable number
of cycles.
If the pipeline connections or lateral buckles are not able to tolerate the total axial
displacement developed over the life of the pipeline, remedial measures must be developed
to control the walking response, for example anchoring.
There is significant uncertainty over how accurate the pipeline walking predictions are.
Consequently, an acceptable approach is to recognise the potential for walking, but to delay
significant remedial measures until the actual behaviour of the pipeline has been
established. This approach must be integrated into the pipeline integrity monitoring system
(Section 7.3.3).
Given the uncertainty at low rates of walking, the current recommendation is to
implement monitoring procedures rather than pre invest in expensive mitigation
techniques. However, in this case the monitoring procedures must be capable of
quantifying the pipeline behaviour in a way that allows mitigation measures to be
implemented as necessary.
In addition, if this approach is adopted, the design should identify how any
operational mitigation could be implemented (for example, it may be sensible to
incorporated suitable anchor attachment points into the design).
Where
XL is the Characteristic VAS limited by pipeline length
XDF is the Characteristic VAS limited by the driving force envelope
XS is the Characteristic VAS limited by sharing
N NBH
X DF 2.588 L BH 2 0 .... 4.2
fLB
Where the Hobbs buckle length, LBH, is calculated by solving the following equation:-
2 5
N0 34.06
EI
1.294 fLB L BH 1 1.668x10 4 EA HUB Re s L BH 1
.... 4.3
2
L BH fLB EI 2
EI
NBH 34.06 .... 4.4
L2BH
This solution is the Hobbs isolated mode III solution [4] for an elastic pipeline. LBH is
the length of the central buckle lobe and 2.588LBH is the overall length of the lateral
buckle.
N̂ N
X S CR BH 2.588 L
BH .... 4.5
fLB
The upper bound critical buckling force for a straight laid pipeline can be calculated from:-
EI HUB EI HBE
N̂cr Max 4.4 ,7.1 .... 4.6
D D
This equation is based upon the critical buckling force model outlined in Section
4.5.2.1 and 4.6.4.1. The upper bound lateral break-out condition is combined with
the best estimate out of straightness parameter and the best estimate lateral break-
out condition is combined with the upper bound out of straightness parameter.
N̂ NBT L
X S CRT BT .... 4.7
fLB
2
If the trigger spacing is less than the sharing limit associated with the trigger, then the
Characteristic VAS must be evaluated using the Probabilistic Definition (or the trigger
spacing increased).
Note 1 If there are no engineered triggers, then the full length of pipe is the total pipeline length. If there are engineered
triggers, then the full pipe length is the total length of pipe between initiators.
6000
5000
Characteristic VAS (m)
4000
3000
0
0 5 10 15 KP 20
The red curve shows the Characteristic VAS for a pipeline in which the temperature
does fall significantly. The Characteristic VAS reduces significantly with length and in
this example buckling is not a concern beyond KP 14 (the probability of buckling in
any km is less than 1%).
When a buckle initiation strategy is adopted, the calculation of Characteristic VAS must
incorporate the influence of the triggers on buckle formation.
If the triggers are well designed (fire reliably), then the Characteristic VAS associated
with any rogue buckle will reduce, this is illustrated in Figure 4.2.
4 Sleepers (rogues)
6000
No Triggers
5000 Triggers
Characteristic VAS (m)
4000
3000
2000
1000
0
0 5 10 15 KP 20
Define a quantitative model of the underlying process - in this case the buckle formation
process (Section 4.5);
Establish relevant probability distributions for the key input parameters (Section 4.6);
Undertake the probabilistic simulation (Section4.7);
Extract probabilistic results (Section 4.8).
Since frequent buckle formation is desirable, it is important that the simulation does not
overestimate the likelihood of buckling. Specific issues which can compromise buckle
formation are:-
Higher than anticipated residual lay tension
A high residual lay tension will reduce the driving force for buckling. The lay tension
can be higher than anticipated due to residual thermal loads; installation of in-line
structures or flooded pipeline installation
Longer than anticipated time delay between installation and start-up
If there is a significant delay between installation and start-up, the on-bottom
condition of the pipe may change. Buckle formation could be compromised if the
pipeline embedment increases significantly or significant self-burial occurs. A similar
concern arises if the pipeline initially operates in a low temperature regime, which
means that the lateral restraint may be higher than anticipated once full operating
conditions are imposed.
Each of these areas of concern can be evaluated using the buckle formation model.
In the implementation the limit state can be recast as a load resistance ratio:-
N
R .... 4.10
Ncr
Where
EI Ws
Nchar 3.86 .... 4.13
D
XNH is a normalised critical force parameter. Guidance on the behaviour of XNH is given in
Section 4.6.4.1.
If XNH is equal to unity, then the critical force is the Hobbs minimum force for an
infinite mode buckle. For XNH below 1 buckling would not be feasible based on the
Hobbs analysis. In practice buckling could occur if significant OOS is present. As
XNH increases above 1, lower and lower level of OOS could initiate a buckle.
Consequently, the parameter XNH is a measure of pipeline OOS.
For a PIP system, an effective diameter can be used in equation 4.13:-
8 EI
D eff
EA
Where EI is the total bending stiffness of the PIP and EA is the total axial stiffness of
the PIP.
For a given target bend radius, the actual as-laid shape will contain features that are more
severe than the target. The parameter XNB is essentially a measure of this variation from
nominal radius. Guidance on the behaviour of XNB is given in Section 4.6.4.2.
EI Ws
NV 4 .... 4.16
H
The parameter XNV describes the lateral OOS in the vicinity of the vertical imperfection.
Guidance on the behaviour of XNV is given in Section 4.6.4.3.
In practice the local lateral OOS will be very important and the buckling force will be
related to the lateral resistance provided at the vertical upset. As the model does
not consider the impact of lateral OOS explicitly, the lateral resistance is not a
parameter. Care is required in the use of this relationship.
Where
EI WsB
NcharB 3.86 .... 4. 18
D
1
Pre-buckle
0.9
1st buckle stage 1
0.8 Buckle #1 Pre-buckle 2
0.7
N/No max
As the start-up continues and the operating conditions become more onerous, the buckle
begins to develop. The force within the buckle drops as the pipe feeds-in, and the slip zones
become more developed. The figure shows two stages in the heat-up post buckling. As the
buckle develops the locations of the anchor points change; essentially the term anchor point
is somewhat of a misnomer. Prior to buckling every point between the hot anchor point and
the end moves towards the end of the pipeline as it expands (i.e. to the left in Figure 4.3).
Consequently, when a buckle appears, the first virtual anchor forms at a location that has
already translated away from the buckle site; this effects the feed-in to the buckle, but is
conservatively neglected within the VAS concept.
The green profile in the figure is taken to be the point just before a second buckle initiates.
Here the force developed is higher than that attained in the original pre-buckled profile; again
this may not be the case – the details of whether a second buckle forms when the force is
higher or lower than the initial critical force is entirely determined by the distribution of out-of-
straightness (and seabed frictional response). Further progress of the start-up process is
illustrated in Figure 4.4.
1 Pre-buckle 2
Buckle #1 Buckle #2
0.9 Buckle#2 stage 1
0.8 Pre-buckle 3
0.7
N/No max
0.6
Fully Constrained
0.5 Force
0.4 As system heats-up
0.3
0.2
0.1
0
0 0.2 0.4 x/L 0.6 0.8 1
1
0.9
Pre-buckle 3
0.8
Fully developed
0.7
Steady State No
N/No max
1
0.9
Fully developed
0.8
Steady State No
0.7
N/No max
0.6
D1 D2 D3 D4
0.5
0.4
0.3
0.2
0.1
0
0 0.2 0.4 x/L 0.6 0.8 1
The information provided here is based on data gathered for a small number of
projects. In order to gain confidence in the methodology and develop a more robust
definition of OOS it would be necessary to repeat the process with a large number of
different data sets. Although the analysis undertaken provides useful support to the
proposed methodology, care should be taken in its application. In general project
specific work should be performed to support the chosen distributions.
3
Recommended
2.5 1 km pipe Length OOS1
OOS7
2 OOS8
OOS9
Probability
OOS25A
1.5 OOS25B
0.5
0
0 0.5 1 XNH 1.5 2 2.5
The results are consistent with the anticipated behaviour. The distribution implies
that for a 1 km length of pipe the most likely buckling force is the Hobbs safe force
(XNH~1). It would be expected that 1 km of pipe would contain at least one area with
sufficient OOS to trigger buckling at this level. However, it is possible for the force to
be below the Hobbs safe force – in areas of high OOS – but not too far below it.
Conversely, very straight sections of pipe will not buckle until the force is significantly
above the Hobbs safe force.
A total of 158 km of data has been analysed to produce the distributions. The data
covers a large range of pipe bending stiffness; from 44 MNm2 to 822 MNm2. The
variation in the mean value of XNH for the individual pipelines is illustrated in Figure
4.8.
1.5
1.5
1.4
1.4
1.3
1.3
1.2
1.2
Mean XNH
1.1
Mean XNH
1.1
1
1
0.9
0.9
0.8
0.8
0.7 0.7
0.6 0.6
Nchar (kN)
EI (MNm2)
0.5 0.5
0 500 1000 0 5000 10000
In this equation s, the bend arc length, must be defined in km. The distribution is suitable
for conceptual design, but should be verified for use during detailed design.
The statistics are based on a review of high quality as laid data from three pipelines.
The data is summarised in Figure 4.9.
1.4
1.2
0.8 5% Prob ex
X NB
0.6
0.4
0
0 0.2 0.4 0.6 0.8 1 1.2
Arc Length (km)
1.2
3.5
1 3
0.8 2.5
Probability
2
NV
0.6
X
1.5
0.4
1
0.2
0.5
0 0
0.4 0.6 0.8 1 1.2 0 0.5 1 1.5
Height (m) X NV
4.6.4.5. Bathymetry
Lateral buckling can initiate from vertical OOS. This is an important driver for buckle
formation, especially if significant spanning can occur. It is not possible to provide generic
data as this is entirely site specific.
If good quality bathymetric data is available to the project then this can be analysed using FE
analysis to provide an estimate of the severity of vertical imperfections. These analyses can
be used to:-
Update the inherent OOS distribution (XNH) to incorporate the vertical OOS;
Identify specific significant features which can be incorporated into the buckle formation
analysis in the same way as engineered triggers.
One method to do this is to divide the bathymetric data into discrete sections
(typically a 1 km long) and deform an initially straight pipe to the seabed data. The
operating load in the pipe is then increased until bucking occurs. This will identify the
most severe OOS features. If these are to be incorporated into the buckle formation
assessment, then a number of sensitivity cases should be undertaken to assess the
variation in buckling force with key parameters (for example vertical seabed stiffness,
survey data error, lateral resistance). The results of these analyses can then be
used to define a suitable distribution of critical buckling force for any bathymetric
feature.
Identify the influence of free spanning on the force profile in the pipeline.
If the bathymetry leads to pipeline free spans, then the expansion behaviour of the
pipeline will be modified. The pipeline expansion will feed in to the spanning
sections. This will absorb some of the feed-in that would otherwise drive lateral
buckling, and is generally beneficial. Very significant spans will modify the force
profile in a similar way to lateral buckling, i.e. the effective axial force will drop at the
span as the pipe feeds-in to it and virtual anchor points can be set up between spans
(or between spans and lateral buckles). The effect of this is to reduce the available
driving force for lateral buckling and reduce the Characteristic VAS should lateral
buckling occur.
Probability of exceeding
0.15 1.E-01
pdf
0.1
1.E-02
0.05
1.E-03
0
0 5 10 15 20
1.E-04
Number of Buckles in Pipeline
VAS
0 1000 2000 3000 4000 5000
1.E+00
1.E-01
P (X>VAS)
1.E-02
1.E-03
1.E-04
The VAS cumulative probability distribution is shown as the blue line within Figure
4.12. In this example the trigger is quite reliable - there is a 7% chance that the
trigger fails to initiate a buckle - so the distribution starts with a probability of
exceedance very close to 1. For a trigger, the Characteristic VAS has a 10% chance
of being exceeded; this level of confidence is shown within Figure 4.12 as the solid
black line. The Characteristic VAS can be extracted from the distribution, which in
this example is 2800 m.
For a rogue buckle the distribution of VAS can be obtained by recording the maximum VAS
in each 1 km of pipe in each simulation. The Characteristic VAS is extracted from the
resulting VAS distribution.
In each simulation within the Monte-Carlo simulation, if a buckle forms anywhere
along the 1 km section of pipe, the VAS associated with the buckle should be
recorded. If no buckle forms within that km of pipe then the VAS should be recorded
as zero (this allows the probability of no buckling to be identified). The resulting VAS
distribution can then be plotted, Figure 4.13.
VAS (m)
0 2000 4000 6000 8000
1.E+00
1.E-01
P (X>VAS)
1.E-02
1.E-03
1.E-04
This analysis approach may be performed using the anticipated bathymetric profile.
This can provide useful insight into the key vertical OOS features (see Section
4.6.4.5) but it is important that the uncertainty in the bathymetry and other data is
fully explored.
To demonstrate robust formation under all conditions, it is necessary to perform extensive
analysis to address a significant number of additional issues, for example:-
The effect of global variation in friction; a conservative choice of friction from a pipeline
integrity point of view is not necessarily conservative from the buckle formation point of
view;
The effect of local variations in friction, including increased lateral restraint in
undesirable locations;
The effect of undesirable variations in the engineered OOS;
The effect of large OOS in undesirable locations;
The effect of survey errors in vertical OOS;
The effect of vertical seabed stiffness.
The aim of the analysis is to demonstrate that under adverse combinations of the various
parameters, buckles will form as frequently as required.
5. DESIGN ANALYSIS
5.1. SYSTEM MODELLING: VAS ANALYSIS
The pipeline should be modelled with an isolated buckle between virtual (or real) anchor
points. The model length should be equal to the appropriate VAS.
For a PIP system both the jacket pipe and flowline must be incorporated within the model,
i.e. an equivalent section model is not acceptable.
For a PIP system both the jacket pipe and flowline must be incorporated within the model.
The interaction between the two pipes must be carefully modelled to incorporate all relevant
effects (e.g. insulation stiffness and strength, gap modelling, centralisers, structural
connectivity between pipes and internal friction).
The design equations outlined in Section 6 assume that the loads are calculated in
accordance with certain modelling assumptions. If they are not, the design process may be
non-conservative. For example the use of a strain limit for local buckling requires that the
potential for strain localisation be fully quantified; this is achieved using the assumptions
outlined in this section.
The following modelling features must be incorporated into the FE model:-
The stresses associated with the as-laid condition;
This requirement applies to major as-laid features, for example route or snake bends
or vertical stress over sleepers and bathymetry if there is significant undulation. It is
not necessary to model the stress associated with minor bathymetric variation in the
VAS model. If it is significant, the influence of bathymetry can be evaluated in the
buckle interaction model (Section 5.2 ).
The element length at the crown of the buckle must be small enough to identify the
curvatures developed within the buckle. The user should perform a sensitivity study to
demonstrate that the chosen mesh is adequate. In the absence of such a study, the
maximum element length in the vicinity of the buckle must be limited to one pipe
diameter.
A fully non-linear pipe-soil interaction model must be employed, in accordance with
Section 5.7.
The potential for strain localisation at the crown of the buckle due to pipeline field joint
stiffness discontinuity must be fully quantified.
For single pipe systems significant stiffness discontinuity could occur in a concrete
coated pipe[14-16] or a thick insulation coated pipe. The effect of the field joint on strain
localisation should be evaluated through detailed FEA or a test programme.
For a PIP system, the field joint stiffness discontinuities should be modelled.
Any other source of significant strain concentration should also be incorporated into
the design, for example buckle arrestors or change in cross-section.
It is not necessary to model the influence of normal joint to joint strength mismatch on
strain localisation.
The influence of these discontinuities can be incorporated into the assessment either by use
of appropriate SNCF or the inclusion of a length of elements with reduced bending capacity.
If weak elements are incorporated into the FE model, it is prudent to undertake some
analysis without the weak joint. This allows the importance of the weak joint to be
isolated and may assist in defining suitable tolerances within the pipeline
specification.
If a SNCF is employed, it should be evaluated using detailed local FE models of the
buckle crown. The SNCF should be appropriate to the level of pressure, axial force
and bending anticipated at the buckle crown; this may change as a function of
imposed load. With the SNCF approach care should be taken to account for the
change in global curvature caused by the presence of the local discontinuity.
For fatigue loading the analysis should incorporate the potential for stress concentration due
to the field joint discontinuity.
If the tension on unload in the vicinity of the route bend exceeds SRC, then consideration
should be given to increasing the route bend radius.
Simple models to address walking have been developed within the SAFEBUCK JIP[8,17].
These models are recommended for conceptual design and are outlined in Appendix D.
The selection of appropriate pipe-soil parameters for walking interaction analysis is
discussed in Section 5.7.4
In all cases the user must ensure that the simplifications do not compromise the applicability
of the analysis.
For example, a pipe-soil model may be employed that does not incorporate the effect
of berms. However, in this case, the model may overestimate the ratcheting growth
of the buckles, which in turn will influence the walking behaviour of the pipeline.
The selection of pipe-soil parameters for walking analysis is discussed in Section 5.7.5
Sensitivity analyses should be undertaken to evaluate the influence of relevant parameters.
Full length models are complex, and for all but the shortest pipelines will involve very
long analysis run time. Generally it is not practical to undertake a large number of
sensitivity analyses using a full length FE model. The use of the VAS concept within
the SAFEBUCK design methodology is intended to minimise the number of analyses
required using full length FE models. Consequently, if the design guideline is
followed, it is anticipated that only one or two sensitivity configurations will require
investigation. A suitable number of sensitivity analyses can be defined by the project
team.
The relevant parameters will be different for each project. Typical parameters
include; axial and lateral pipe-soil resistance; vertical seabed stiffness; bathymetry;
route bend radius; trigger design and location; operating cycles.
1
C
B F
0.5
D A E
0
-1.5 -1 -0.5 0 0.5 1 1.5
-0.5
Normalised Axial Stress
-1
-1.5
s ts c tc
eq 5.1
ts tc
Where ζ is the property of interest (Young‟s Modulus, Poisson‟s ratio and coefficient of
thermal expansion), t is the nominal wall thickness. The subscript s refers to the base pipe
and c refers to the cladding.
The submerged weight of the system must incorporate the weight of the cladding.
A m .... 5.2
The CRA data in Table 5.2 are tentative and taken from supplier data sheets. These
should only be used in conceptual design, and in the absence of any more accurate
information.
550 Engineering
Stress (MPa)
500
450
Plateau
Roundhouse
400
350
300
0 1 2 3 4 5
Engineering Strain (%)
Stress (MPa)
450
250
50
-150
Twice Yield
Strain (%)
-350
compressive cycle
-550 tensile cycle
E Y
Q QY Y L ..... 5.3
C np L
In this model the true stress, Q is given in terms of the logarithmic strain λ. The real
material will exhibit a plateau in the engineering stress-strain response in both
tension and compression. This leads to conflicting requirements in true stress-strain
space (unless a non-monotonic stress-strain curve is adopted[23,24]). Equation 5.3
defines a plateau in the true-stress true-strain response; this is a compromise
between the observed conflicting responses and should be adequate for most design
purposes.
This model provides a good representation of the stress-strain response from first load to
ultimate tensile strength.
The model employs four parameters; the engineering yield strength, Young‟s Modulus, yield
to tensile strength ratio and the engineering Lüder strain.
From the engineering yield stress and the engineering Lüder strain the true stress and strain
parameters are:-
y
Q y y 1 y y 1
E
..... 5.4
y
Y ln1 ..... 5.5
E
The constant C and the hardening co-efficient n can be calculated from the other parameters
using the following equations:-
QY
C n
..... 5.7
Q
L Y
E
n n
QY e Q
L Y ..... 5.8
UTS n E
Suitable parameters for X60 and X65 linepipe steel at room temperature are given in Table
5.4
Table 5.4 Suitable Parameters for X60 and X65 Steel at Room Temparature
This model of stress strain response is based upon a review of extensive project data
undertaken for the lateral buckling SRA[9]; the review covered API X60 and X65
linepipe steel. The recommended curves are based on the SMYS, combined with
the mean plus 1SD YT ratio and Lüder strain observed in the data set.
The YT ratio was found to be a strong function of the material yield stress. The data
is well represented by a normal distribution with mean given by equation 5.9 and a
CoV of 2.5%.
y
YT 0.65 0.16
..... 5.9
ref
Where σref is 400 MPa.
The Lüder strain was well represented by a normal distribution with mean 1.6% and
CoV of 15%.
Wherever possible the stress-strain curve parameters should be confirmed by project
specific testing.
The Ramberg Osgood model cannot give an accurate model of the stress-strain
response over the whole strain range of the material. The degree of strain hardening
must be carefully chosen and be appropriate to the strain range of interest. Basing
the hardening on SMYS and SMTS can produce unrealistic curves. The strain
hardening should be fully supported by testing to ensure that an excessively
optimistic model is not employed.
In the absence of project specific data, the parameters defined in Table 5.5 can be
used to represent X65 linepipe.
Yield strength α n
Low hardening 450 MPa 1.4 27
Median hardening 450 MPa 1.35 21
High hardening 450 MPa 1.3 16
These parameters are developed from a brief review of mechanical test data
undertaken as part of the lateral buckling SRA[9].
Much of the data available to SAFEBUCK JIP is related to very soft high plasticity
deepwater clays. Limited test data is available for high strength soils, such as stiff-
clays and sandy soils. This is an area of ongoing research.
2
Axial pipe-soil resistance
1.5
0.5
-0.5
-1
Undrained Brittle
-1.5 2nd Cycle
-2 Undrained Ductile
Drained
-2.5
Axial Displacement
High breakout resistance will inhibit pipe walking under thermal transient loading - if
the breakout occurs on all cycles. However, in most cases, thermal start-up
transients follow a long period of pipe cooling, during which the pipe is continuously
contracting; so a breakout response is unlikely to occur on all walking cycles.
A high breakout resistance can produce a higher driving force that will tend to increase the
propensity for buckling. This is particularly true at hydrotest, where a long set-up time is
likely. If it can be demonstrated that long sections of the pipeline remain fully constrained
(with no axial displacement) prior to buckle formation then incorporating a breakout
resistance into the force profiles may be warranted, provided there is evidence to support a
high breakout resistance.
In most cases accounting for axial breakout is potentially non-conservative for buckle
formation and walking predictions. For this reason modelling the axial breakout
resistance should only be undertaken with great care.
Sensitivity checks on route curve pull-out should include axial breakout resistance in
cases where a long set-up, following an extended shutdown, and a subsequent
reduction in effective force (due to cooling for example) could increase the peak axial
tension along the route curve.
Residual resistance
Horizontal
,
displacement
Accumulating passive
Horizontal
resistance
displacement
Lateral buckles with uneven support conditions require careful consideration of the
effect of increasing contact pressure at support points and reducing contact pressure
at low points along the lateral buckle. This will occur with „heavy‟ pipelines but also
where a pipeline passes over a feature such as a sleeper or a vertical out-of-
straightness in the seabed.
In the absence of project specific soils tests, reference should be made to Appendix B for a
representative approach to establishing the values of key parameters such as cyclic frictional
behaviour and berm resistance.
6. LIMIT STATES
The relevant failure modes within a lateral buckle are:-
Local buckling (Section 6.2);
Fatigue (Section 6.4);
Fracture (Section 6.5).
In addition to these limit states, the pipeline material specification must be consistent with
the high levels of imposed load; this is addressed in Section 6.6.
The limits outlined in Section 6.1 to 6.6 are intended for general application. It is possible
that project specific assessments of the limit states will lead to enhanced system capacity.
Guidance on the potential for this approach is presented in Section 6.7.
All other limit states, for example pressure containment or hydrostatic collapse,
should be evaluated in accordance with the governing design code.
The design equations presented in this section are based on the requirements of API 1111[2].
For pipelines designed in accordance with DNV OS-F101, relevant modifications and
interpretations required for application of DNV-OS-F101 are given in Appendix C.
The approach outlined here is only intended for use in the assessment of the conditions in a
laterally buckled pipeline.
To date at least four pipelines have ruptured as a result of lateral buckling; another
has been abandoned early in life with little chance of meeting its original design life.
Three failures involved pipelines in which lateral buckling was not seriously
considered as a design criteria, and two in which it was - albeit with a seriously
flawed design process.
Basing actual failure statistics on these failures alone indicates a historical failure rate
in excess of the target values defined in Table 6.1. Clearly, this is not a like for like
comparison and there are generally a number of causes to any failure. Nevertheless,
this does indicate that previous design practice was not consistent with the desired
levels of reliability. The guidance in this recommended practice is intended to ensure
that the pipeline design does meet these targets.
t
c 0.5 .... 6.1
D0
The superior strain hardening behaviour of CRA materials means that Equation 6.1
may be unduly pessimistic. A project specific investigation may produce a significant
enhancement in the buckling strain limit.
The imposed strain should comply with the following equation:-
c
2 .... 6.2
f2
D
f2 1.4 1 0.02 0 .... 6.3
t
D
f2 1.2 1 0.02 0 .... 6.4
t
The strain to be employed, ε2, is the absolute value of the maximum compressive axial
mechanical strain, i.e. any thermal strain should be removed from the total compressive axial
strain prior to use in Equation 6.2.
Alternatively the designer can employ the bending strain in the design check:-
D0
b .... 6.5
2
This approach is consistent with the test data base used to develop the strain limits.
1
g .... 6.7
1 20
And the collapse factor for combined loading is defined by:-
1.2 fo
fc .... 6.8
g
The collapse factor, fo, is equal to 0.6 for cold expanded pipe and 0.7 for seamless pipe.
The collapse pressure is defined by:-
p YC pE
pC .... 6.9
p 2YC pE2
Where:-
t
p YC 2 yc .... 6.10
Do
3
2E t
pE
1 2
Do
.... 6.11
Dmax Dmin
.... 6.12
Dmax Dmin
This definition differs from the DNV-OS-F101. For the same cross-section the
API 1111 ovality is half the DNV-OS-F101 ovality, i.e. the pipe ovality must be less
than 3% based on the DNV-OS-F101 definition.
For reeled pipelines, the post installation ovality and any potential for a reduction in
yield stress due to reeling should be employed in the design check.
In applying Equation 6.6 the minimum level of internal pressure that can occur during
operation should be employed. The minimum internal pressure may result from a blowdown
condition, although such a condition may be associated with low operating temperature.
c
2 fp .... 6.13
f2
D
1.6 1 0.02 0
f2 t
..... 6.16
25 fp D0 / t
1 0.45 tanh
2 .5 f
p
The suitability of the local buckling design equations for use at a lateral buckle has
been evaluated in the SAFEBUCK JIP[9]. This found that the basic equations are not
acceptable when failure is at strains below the Lüder strain. For internal over
pressure this is only likely to apply to for pipes with D/t>30 and σh/σy<0.5.
The SAFEBUCK SRA performed a comprehensive assessment for the internal
overpressure condition. For external overpressure a very limited evaluation was
performed[10]. This showed that the buckling capacity is lower and is significantly
compromised by the Lüder strain. Insufficient work was undertaken to develop
general design guidance. If this limit state is critical the project should undertake
specific work to ensure that the design is conservative.
A corrosion allowance is routinely added to carbon steel pipelines, but the corrosion
mechanism is rarely considered in structural design. For example, pitting is a
common form of corrosion for which the corrosion allowance might be intended. It is
probable that small isolated pits will have little or no effect on the buckling resistance
of the pipeline. Corrosion involving more significant areas of metal loss, for example
six o‟clock grooving, may have a greater effect on the collapse resistance. However,
even this might have a limited effect on the bending buckling resistance of a pipeline
when the bending is in the horizontal plane (as it is for lateral buckling). Only if the
corrosion anticipated involved 360º metal loss would removal of the allowance in the
resistance calculation be wholly required.
In addition, corrosion is a time dependant phenomena. If the operating conditions
reduce with time; combining the early life loads with late life corrosion is
unnecessary.
In all cases, the imposed load should be calculated based upon the nominal pipeline
thickness.
Undertaking the load calculations with a corroded pipe cross section will normally
reduce the loads. Using these reduced loads in the design is generally non-
conservative and not representative of the pipe response.
LATERAL
BUCKLING
DESIGN
S-N ECA
S-N pipeline geometry
REVISE weld geometry
DESIGN*
IS FATIGUE tensile stress
DAMAGE N tensile properties
ACCEPTABLE?
Y stress range
number of cycles
ECA stress concentration factor
REVISE
fracture toughness
DESIGN*
IS FLAW SIZE S-N curve fatigue crack
ACCEPTABLE growth law
N AND
PRACTICAL? correction to S-N curve / fatigue crack
Y growth law / toughness for environment,
temperature and frequency
FATIGUE & FRACTURE factor of safety*
LIMIT STATES (*this may be different for the S-N and the ECA)
ACCEPTABLE
Figure 6.1 The Fatigue and Fracture Limit States and the S-N and ECA Calculations
The two methods used to assess fatigue are:
S-N curves; and
Fracture mechanics.
The fatigue limit state should first be addressed through the S-N approach to fatigue. If the
results of the S-N calculations are acceptable, then an ECA should be conducted. The ECA
addresses both the fatigue limit state, through the fracture mechanics approach to fatigue,
and the fracture limit state. Although the ECA calculations require more information than the
S-N calculations, most of the required information is common to both calculations, see
Figure 6.1.
The two sets of calculations are complimentary. S-N curves are only appropriate if the weld
is free from significant defects. The presence of flaws may reduce the fatigue life below that
predicted using S-N curves. One of the objectives of the ECA is to determine this significant
flaw size. The guidance in Section 6.4 is mainly concerned with the S-N calculations and
that in Section 6.5 with the ECA calculations.
Fracture mechanics calculations should not be used to justify a fatigue life in excess
of that predicted using S-N calculations.
In the fatigue assessment, it is necessary to consider all sources of fatigue loading, from
installation through to the end of the design life of the pipeline in order to identify the limiting
case. Typically, in a pipeline designed to buckle laterally, the limiting condition will be
defined by the fatigue loading arising from the movement of the buckle due to start-up and
shut-down cycles. In some cases, other forms of fatigue loading may be significant (e.g.
fatigue loading of spans associated with lateral buckles due to environmental loads or
slugging flow).
When considering the fatigue loading due to start-up and shut-down cycles it is
typically convenient to consider a number of different load cases, based on the
duration of a shut-down (e.g. less than 3, 6 or 12 hours, greater than 24 hours)
and/or the operating regime (e.g. normal operations, hot-oiling, cold displacement
etc.).
The guidance here assumes that the fatigue life will be governed by the fatigue
behaviour of the pipeline girth welds. This may not be the case if there are other
sources of significant stress concentrations in the pipeline. For example, the
presence of J-lay collars, buckle arrestors or anode attachment pads will lead to local
increases in stress. If possible, these design features should be avoided in pipelines
that are designed to laterally buckle. If they cannot be eliminated, then they should
be made as fatigue friendly as practical and the design must quantify the stress
concentration and address the influence of these features.
S-N curves and fatigue crack growth laws are specific to the environment, temperature and
loading frequency. Project specific testing may be required to establish suitable values,
particularly if the internal environment is sour or is expected to sour, see Section 6.7.
Souring of the reservoir can occur due to either bacterial or thermo-chemical
reduction of aqueous sulphates, or thermal decomposition of sulphides in sulphur
rich source rocks. Bacterial souring of the reservoir can occur because of bacteria
introduced into the reservoir with the injection water. Typically, bacterial souring of
the reservoir occurs a number of years after start-up. Following souring, typical H2S
production levels through field life are in the region 50 to 200 ppm (dependant upon
reservoir characteristics and injection water treatment).
At conceptual design, some of the information required to address the fatigue and fracture
limit states (e.g. fracture toughness, effect of the environment) may not be available.
Sensitivity calculations should be conducted to determine the significance of the various
assumptions, and hence whether project-specific testing and/or more detailed calculations
are likely to be required in detailed design.
also states that the S-N curves are applicable up to 100°C. The guidance in
PD 5500 is more conservative.
The recommended classifications for single sided pipeline girth welds are outlined in Table
6.2.
N A S m ..... 6.17
Classification A m
D 1.52x1012 3
12
E 1.04x10 3
11
F 6.33x10 3
Table 6.3 Constants in the in-air S-N Curves
The constant A in Table 6.3 is not dimensionless. The values defined are applicable when
the stress range is defined in N/mm2.
Miner‟s Rule can be used to combine the effect of different cycles of different stress ranges
(e.g. full and partial shut-downs) to give the total fatigue damage ratio.
n
D fat Ni fat ..... 6.18
i i
Suitable values for the allowable fatigue damage ratio are given in Table 6.4.
S R K t Km ..... 6.19
The nominal stress range must not exceed the elastic limit defined by the Bauschinger
effect; the appropriate limit is outlined in Section 6.6.
The applicability of S-N curves to the high stress-low cycle regime has been
established in SAFEBUCK. There is no limit on the local stress range, so long as the
nominal stress range meets the requirements of Section 6.6.
The calculation of nominal stress range should account for the stress concentration
associated with field joint geometry.
Where tref is 16 mm and the exponent k is defined in Table 6.2. For thicknesses less than
16 mm, Kt is unity.
e
Km 1 3 ..... 6.21
t
Where e is the eccentricity, defined as the axial misalignment between the wall thickness
centrelines of the two pipes.
There are a number of different expressions for calculating SCFs due to axial
misalignment. Several of these expressions, such as those proposed by Connelly
and Zettlemoyer[28], are known to under-predict SCFs for typical flowline dimensions
and misalignments. Equation 6.21 is known to be conservative, and should be used
at the early stage of design. As the design progresses, a flowline specific analyses
may be performed to develop a more accurate expression.
In determining an appropriate axial misalignment and hence SCF, due account should be
taken of the line pipe manufacturing tolerances and the welding tolerances. The simplest,
and most conservative, approach is to calculate it from the maximum permitted hi-lo, and the
wall thickness and diameter tolerances from the relevant line pipe specification. However,
this is likely to be over-conservative (and may be detrimental to the design). If line pipe and
welding dimensional data is available then statistical methods can be used to determine an
upper bound misalignment with a given probability of exceedance. In the absence of
project-specific data, data from previous projects can be used (if available) and its
appropriateness confirmed prior to fabrication using project-specific data.
6.4.3. Installation
The extent to which the fatigue loads during installation have implications for operation, and
vice versa, will vary from case to case. The largest cyclic loads during installation will occur
in a plane transverse to those during operation if the pipeline does not rotate and the lateral
buckle is confined to the horizontal plane. In this case, the interaction between the two will
be small (there is likely to be a small component of fatigue loading in the horizontal plane
during installation, and similarly in the vertical plane during operation). If the pipeline does
rotate, or if the lateral buckle is not confined to the horizontal plane, then the interaction
between the two will be more significant.
In conceptual design, it is recommended that 20% of the total allowable fatigue damage is
assumed to be consumed during installation (e.g. if the allowable fatigue damage is 0.20,
then the limit for fatigue loading during operation is 0.16).
In detailed design, the fatigue loading during installation should be determined by the
installation contractor.
A high integrity field-joint coating should demonstrably exclude sea-water from the
weld under all loads that the pipeline will experience from installation to the end of
the design life.
Lateral buckling imposes a very low frequency loading (less than 10 -5 Hz). Corrosion-fatigue
is sensitive to the loading frequency. In this context, per cycle, low frequency loading is
more severe than high frequency loading.
Knock-down factors (also referred to as fatigue life reduction factors) to be applied to the
in-air S-N curves are given in Table 6.5. In the absence of better data, these are suitable for
use in conceptual design. However, project-specific testing is recommended, see Section
6.7.2.
The guidance assumes low carbon-manganese line pipe steel. Refer to Section 6.4.5 for
other materials.
The tentative recommendations given here are based on fatigue tests undertaken by
SAFEBUCK and the limited number of tests in the published literature[29-49]. They
may be non-conservative. Environmental parameters such as pH, temperature, and
the partial pressures of H2S and CO2 are likely to have an effect on the observed
knockdown factor although quantitative guidance is not currently available. Project
specific testing is therefore recommended.
The range of values provided for sour environment reflect the range of environmental
conditions that may occur in different projects. For low pH conditions and high partial
pressures of H2S, factors towards the upper end of this range may be expected. For
higher pH conditions and/or lower partial pressures of H2S, factors nearer the lower
end of this range may be more likely.
During conceptual design, it is useful to compare the anticipated environmental
conditions within the pipeline with those used in previous testing programmes. It is
noted however that comparison of different test data is not straightforward, as
differences in cyclic loading frequency, stress range, and in-air fatigue performance,
can themselves have an influence on the apparent knockdown, and mask the true
environmental effect. In the latter instance it is recommended that knockdown factors
are determined with respect to an absolute baseline (e.g. a design S-N curve) rather
than the actual in-air performance.
6.5. ECA
A design ECA is conducted to determine the tolerable flaw size in all of the structural welds
subject to the loads imposed by the lateral buckling design. It is recommended that the ECA
is conducted in accordance with the guidance in BS 7910[50]. The use of other codes and
standards (e.g. API 579[51] and R6[52]) should give broadly similar results.
In this design ECA, it is necessary to consider all load cases, from installation through to the
end of the design life of the pipeline in order to identify the limiting case, and hence the
smallest tolerable flaw size. Typically, in a pipeline designed to buckle laterally, the limiting
condition will be defined by the maximum load in the buckle at the end of the design life, and
the fatigue loading due to the movement of the buckle.
Examples when there may be other limiting cases include: pipelines installed by
reeling and pipelines operating in sour service.
The ECA should consider both surface and embedded flaws. It should consider surface
flaws in both the weld cap and the weld root. In a PIP system it should consider both the
flowline and jacket welds.
The welds should be over-matched.
The tolerable flaw size determined from the ECA should be compared with flaw sizes that
would be acceptable to the relevant welding code, or could realistically escape detection
during production. If the tolerable flaw size is smaller then typical workmanship acceptance
levels, then the constructability of the pipeline may be compromised. The accuracy and
reliability of the inspection technique should be taken into account when determining the
acceptable flaw size.
Typical workmanship acceptance levels are: 3 mm deep by 25 mm long for a surface
flaw, and 3 mm deep by 50 mm long for an embedded flaw.
Project-specific testing may be required to determine the tensile properties, the fracture
toughness, and the fatigue crack growth law to be used in the ECA. Recommendations for
project specific testing are summarised in Section 6.7.
If the material is expected to exhibit a Lüder plateau and an actual or estimate stress-
strain curve is unavailable, then the approximate method given in BS 7910 should be
applied.
During detailed design, an ECA should be conducted to Level 2B, 3B or 3C, depending upon
the severity of the loading. A Level 3 assessment would be required if high plastic strains
are introduced during reeling or operation.
An assessment to Level 2B requires a material specific stress-strain curve. An
assessment to Level 3B or 3C requires a material specific stress-strain curve, and a
tearing resistance curve.
It has been demonstrated that the results of SENT tests can be used when the
pipeline is subject to longitudinal loading only (e.g. installation). It has not been
demonstrated that the results of SENT tests can be used when the pipeline is subject
to internal pressure and longitudinal loading (e.g. lateral buckling). The implications
of constraint and bi-axial loading at high levels of plastic strain on the crack driving
force and resistance are unclear. If it is proposed to use the results of SENT tests, or
if a constraint correction factor is to be applied to the results of SENB tests, then a
detailed, project-specific, justification is required.
Where Km is defined in Equation 6.21. The intersection of the Neuber curve (Equation 6.22)
with the engineering stress-strain curve for the parent pipe material defines the strain and
stress corresponding to the elastic SCF due to misalignment.
The Neuber approach is only applicable to static loads. The elastic SCF should be used in
fatigue calculations (both S-N and fracture mechanics).
6.5.6. Stresses
The stresses used in the assessment should be principal stresses.
For a circumferential flaw in a girth weld in a lateral buckle, it is normally appropriate
to assume that the axial stress is a principal stress.
The load cases that need to be considered will vary from case to case. The load history
from an FE analysis of several operation cycles is informative. The maximum tensile axial
stress and the maximum axial stress range should be identified. It would be appropriate to
take into consideration any reduction in the maximum stress and stress range over a number
of cycles, if the response of the lateral buckle is shown to stabilise. The maximum stress
and stress range may not be co-incident, and may not occur at the extrados of the crown of
the buckle. It may be necessary to consider flaws at several positions around the
circumference of the pipeline, in order to identify the limiting case. The variation of the
maximum stress and stress range around the circumference is informative. If the toughness
varies significantly over the design temperature range it will be necessary to consider the
variation of the axial stress with temperature; the maximum design temperature may not be
the limiting case.
The use of outer-fibre stresses will typically be conservative. If mid-wall stresses are used
then the variation of the stresses through the thickness should be considered.
Secondary stresses such as residual welding stresses should be considered. The residual
stresses should be assumed to be uniform and modified in accordance with the guidance in
BS 7910 on the relief of residual stresses under an applied load.
H 3 2
L Y2 H .... 6.23
2 4
Therefore, a simple modification would be to calculate the load ratio as the ratio of
the applied load to the axial stress at yield (rather than the ratio of the applied load to
the uni-axial yield strength).
An alternative modification would be to calculate the plastic collapse moment for the
given flaw geometry, pipeline geometry, tensile properties and applied loads. The
load ratio would then be calculated as the ratio of the maximum moment from the
lateral buckling analysis to the plastic collapse moment. This approach represents a
step towards a Level 3C analysis.
Neither of the above two modifications considers the effect of the bi-axial stress state
on the crack driving force or resistance. This is considered in a Level 3C analysis. In
the absence of a demonstration of the general applicability of either of the above
modifications, a detailed, project-specific, justification would be required if either were
to be applied.
The initial defect sizes shall account for any ductile tearing during the reeling process;
The analysis shall account for the impact of the reeling process on the residual stresses
at the weld;
Post-reeling tensile stress-strain properties shall be employed;
Post-reeling fracture toughness properties shall be employed.
The assessment shall evaluate both the case when the plane of reeling and plane of
operational loading are aligned and when they are rotated by 90°. The condition when the
original defect was on the neutral axis during reeling shall be evaluated using the virgin
material stress-strain response.
Sweet corrosion 3
Sour corrosion 10-40
1
Sea-water and cathodic protection 9
Table 6.6 Fatigue Crack Growth Law Acceleration Factors for Carbon Steels
Exposed to Corrosive Environments
1E-02
1E-03
da/dN, mm.cycle-1
1E-04
1E-05
1E-06
in-air
sweet
1E-07 sea-water and cathodic protection
sour
1E-08
10 100 1000 10000
ΔK, N.mm-3/2
Figure 6.2 The Fatigue Crack Growth Laws for Sweet and Sour Environments
There is little test data to support the tentative recommendations for sweet corrosion.
The fatigue crack growth law should be specific to the type and location of the flaw being
assessed. For an embedded flaw, not exposed to internal or external environment, in-air
behaviour can be assumed. For an internal surface breaking flaw, the crack growth law for
either sweet or sour environment (defined in Table 6.6) should be used, depending on the
nature of the product being carried within the pipeline. For an external surface breaking flaw
the crack growth law for seawater and cathodic protection (defined in Table 6.6) should be
used. However, if a demonstrable high integrity field joint coating is used, in-air behaviour
can be assumed for external surface breaking flaws.
Refer to Section 6.4.5 for other materials. In general, the fatigue crack growth rate
acceleration factor should again be consistent with the S-N life reduction factor. Project-
specific testing is recommended.
For clad or lined pipe, no flaws should be tolerated at the weld root or in the zone bounded
by the thickness of the CRA. The ECA should nevertheless consider hypothetical flaws for
comparison with the detection limit of the NDT techniques used. For an internal surface
breaking flaw, the end-of-life condition should be growth of the flaw through the CRA layer.
An embedded defect located within the carbon steel material, at the interface with the CRA
layer, should also be assessed.
6.5.11. Installation
An installation ECA may be conducted by the installation contractor. The installation ECA is
a sub-set of the design ECA, and the results will need to be considered in the design ECA.
The extent to which the loads during installation have implications for operation, and vice
versa, will vary from case to case. The largest static and cyclic loads during installation will
occur in a plane transverse to those during operation if the pipeline does not rotate and the
lateral buckle is confined to the horizontal plane. In this case, the interaction between the
two will be small (there is likely to be a small component of fatigue loading in the horizontal
plane during installation, and similarly in the vertical plane during operation). If the pipeline
does rotate or the lateral buckle is not confined to the horizontal plane then the interaction
between the two will be more significant.
The variation of the maximum stress and stress range around the circumference during
installation and operation is informative.
The allowance for fatigue loading during installation in the ECA calculations should be
consistent with that in the S-N calculations.
Where YT is the maximum specified yield to tensile ratio of the pipeline steel.
The limit ensures that excessive strains are not admitted in the design process. The
limit is based upon the test data base gathered as part of the SAFEBUCK JIP and
public sources[54, 55,56], Figure 6.3.
18 SAFEBUCK Denys
X80 Lillig et al
16 Stew art et al Design
14
Uniform Strain (%)
12
10
8
6
4
2
0
0.75 0.8 0.85 0.9 0.95 1
Y/T ratio
Equation 6.24 may be unduly conservative for CRA materials. If sufficient test data is
available, the limit can be defined as 1/3 of the uniform strain capacity of the material.
The equivalent strain can be calculated from the von Mises equivalent stress and
equivalent plastic strain as:-
eq
eq peq .... 6.25
E
The equivalent plastic strain is defined by:-
p
eq
2
3
2 2 2
Lp Hp Rp
.... 6.26
2
R 3 h
2 fB 1 .... 6.27
SMYS 4 SMYS
Pipe Type fB
Seamless carbon steel 0.8
UOE carbon steel 0.7
CRA 0.7
more severe than high frequency loading. The S-N curves and fatigue crack growth laws
quoted in codes and standards are based on tests conducted at high frequencies (typically
0.1 Hz and above). The focus of a project-specific test programme is therefore most likely to
be the effect of frequency on the fatigue behaviour.
A significant number of tests are required to produce a new S-N curve or a new fatigue crack
growth law. The extent of the required project-specific testing will depend upon the similarity
of the environment to that in previous projects and the sensitivity of the design to fatigue. It
may be sufficient to conduct a small number of tests to verify that the assumed S-N curve
and fatigue crack growth law is conservative.
A project-specific test programme should consist of tests on project-specific materials and
weld procedures qualified for the project in a simulated project environment.
Fatigue testing at low frequencies is time consuming. The implications of the time taken to
conduct tests on the project schedule should be considered.
SAFEBUCK has conducted low frequency fatigue tests of pipeline welds in the
following environments: sea-water with cathodic protection, sweet environment and
and sour environment[29,30]. A number of recent projects have conducted similar tests
in sour environments, some of which have been published.
Fatigue crack growth rate tests are typically conducted under conditions of increasing K.
Alternative testing methods (e.g. constant Kmax) may be required if the behaviour at low K is
to be investigated, to avoid problems with crack closure.
It may be more efficient to generate a set of test data at moderate frequency
(typically 0.2 Hz) and then conduct frequency scanning tests to establish the
influence of frequency. This is because the increasing K tests can take a very long
time at the appropriate test frequency, if this is low. If this approach is adopted,
frequency scanning tests should be undertaken at K values corresponding to
beginning and end of life.
It is noted that the presence of corrosion inhibitor may have a significant influence on
the rate of hydrogen uptake and therefore the resulting fracture toughness, but
experimental data quantifying the effect is limited.
KISCC is the appropriate measure of toughness when the crack tip is exposed to the sour
environment, e.g. surface flaws at the weld root.
KIH is the appropriate measure of toughness when the crack tip is not directly exposed to
sour environment, e.g. embedded flaws and surface flaws at the weld cap, and when the
sour environment has been displaced, as occurs during cold displacement.
7.1. PROCUREMENT
As a result of the severe design conditions it may be necessary to tighten up the basic
linepipe specification. Areas where enhanced project specific requirements should be
considered include:-
Dimensional tolerances (wall thickness, diameter and out-of-roundness);
Range of acceptable yield strength;
Reduced maximum value of YT ratio;
More frequent mechanical testing (e.g. longitudinal tensile, hardness and toughness);
Elevated temperature testing;
Increased NDT coverage (wall thickness, weld flaws);
More detailed and frequent dimensional measurements of pipe ends.
The feasibility of the requirements should be confirmed with the pipe mill as early in the
project as feasible.
It is important to ensure that the strength of the weld is greater than the strength of the pipe
material (weld overmatching) over the strain range of interest. Conventionally, procedures
aim to achieve this only for the ultimate tensile strength of welds. However, for HPHT
applications, the weld overmatching should be achieved over the strain range expected and
with respect to the pipe upper bound yield strength. The overmatching should be achieved
at ambient temperature and at the operating temperature.
Where reasonably achievable, an upper limit of SMYS + 100 MPa should be placed
on the yield stress of the linepipe.
7.1.1.3. Data
Project-specific dimensional data should be collected following manufacture, including
diameter and wall thickness around the circumference of each end of every pipe joint, or at a
reduced sampling rate that provides sufficient statistical data. The data shall be used to
assess the range of SCF likely to be achieved, incorporating the proposed line up and pipe
rotation practices of the contractor.
Detailed pipe data can be used in statistical analysis to determine a realistic SCF.
Detailed pipe measurements are routinely made at the pipe mill; however, they are
often not made available as a database to the client, hence the need to include this
requirement as part of the procurement contract. Providing there is sufficient
measurement data an SCF can be defined with respect to a probability of
exceedance. This approach has been successfully adopted on some recent projects.
7.2. INSTALLATION
There are a number of areas where the severe design conditions will place additional
burdens on the pipeline installation, these include:-
Tolerable weld flaws may be smaller than usual and AUT may be required to size flaws;
Increased testing requirements to confirm weld overmatching;
Additional restrictions on misalignment at girth welds may be specified (confirmatory
measurements may be required);
More accurate as-laid survey may be required (Section 7.3.1.1);
The buckle initiation strategy may complicate the lay (for example a more complex route
or more complex handling because of distributed buoyancy).
The installation contractor should be thoroughly briefed on the importance of the installation
to the design strategy. The feasibility of the requirements should be confirmed with the
installation contractor as early in the project as possible.
In deeper water, measurement of ROV position from surface support vessel is increasingly
inaccurate, so it is essential to use an overarching navigational system that integrates data
from USBL, DVL and INS.
The positional data should be recorded and reported at least one point per metre along the
pipeline. The survey contractor should process the data but no smoothing should be
undertaken.
The vertical position of the pipe should be measured using an ROV mounted digi-quartz
bathymetry unit with measurements reported at 1 m intervals or better.
Extensive calibrations should be undertaken prior to commencing the survey. Calibration of
accuracy and repeatability of local out-of-straightness measurements should be carried out
by surveying the same section of line, a number of times in different directions, preferably
along a section of line with a known OOS feature such as a buckle or a route-curve. This
should be repeated until the system algorithms and position calibration are acceptably
accurate.
Global positioning, while less critical, can be calibrated by holding the ROV at a fixed
location, ideally at known seabed datum point, e.g. a manifold. Over 10-minutes numerous
fixes (e.g. 100) should be taken. The mean position and standard deviation of global
positioning error can then be assessed.
Levels of pipeline embedment should also be assessed after installation or after flooding.
The best quality embedment and seabed data around the pipe comes from Digital Terrain
Mapping (DTM), utilising a dual head Multi-beam echo sounder (MBES) system. Continuous
five-point files (giving depth measurements at top of pipe and two locations – close and far –
to each side of the pipe) should be provided for assessment of embedment and local berm
features along the length of the pipeline. Cross profiles should be provided at 5 m intervals
or better over the region of each lateral buckle.
7.3.3.2. Buckles
The walking behaviour at the buckles should be monitored in the same way as for the
spoolpieces. It may be more difficult to identify axial displacement at the buckles, and
consideration should be given to installing datum marks along the pipe (field-joints are
spaced at least 12 m apart, which is too far for effective measurement), which will facilitate
monitoring. This is more practical at buckle sites employing a fixed structure such as a
sleeper.
7.3.4.2. Operation
The operational survey data must be reviewed to confirm the stability of the buckles. This
must quantify the size and shape of each buckle. In the event of behaviour outside design
predictions, an engineering assessment should be carried out to confirm the acceptability of
the pipeline configuration.
Operational pressure and temperature cycles should be monitored in operation to record
maximum operating load conditions and to ensure that cyclic loading is within the ranges
used in fatigue design. If the number of cycles is considered likely to exceed design
assumptions, a revised fatigue assessment should be considered based on measurement of
buckle shape and amplitude.
If walking is detected, the rate of walking should be established based upon the magnitude
of the overall walk and the number of shutdown cycles that have occurred. The long-term
integrity of the spoolpieces and buckles must be assessed using the observed walking rate.
If integrity cannot be guaranteed, then remedial measures, such as anchoring, must be
implemented.
8. REFERENCES
39. Maddox,S.J., Pargeter,R.J. and Woollin,P. Corrosion Fatigue of Welded C-Mn Steel
Risers for Deepwater Applications: A State of the Art Review. Paper No. OMAE2005-
67499, Proceedings of OMAE 2005: 24th International Conference on Offshore
Mechanics and Arctic Engineering, ASME, Halkidiki, Greece, 12-17 June, 2005.
40. Baxter,D.P., Maddox,S.J., and Pargeter,R.J. Corrosion Fatigue Behaviour of Welded
Risers and Pipelines. Paper No. OMAE 2007-29360, Proceedings of OMAE2007, 26th
International Conference on Offshore Mechanics and Arctic Engineering, San Diego,
California, USA, 2007.
41. Pargeter, R.J., Baxter, D and Holmes, B. Corrosion Fatigue of Steel Catenary Risers in
Sweet Production. Paper No. OMAE-57075, Proceedings of OMAE2008, 27th
International Conference on Offshore Mechanics and Arctic Engineering, Estoril,
Portugal, June 2008.
42. Baxter, D. SAFEBUCK JIP – Additional Corrosion Fatigue Data for Sweet Operating
Environment. Atkins Boreas Report BR09076/2, October 2009.
43. McMaster, F. et al. Sour Service Corrosion Fatigue Testing of Flowline Welds. Paper
No. OMAE 2007-29060, Proceedings of OMAE2007, 26th International Conference on
Offshore Mechanics and Arctic Engineering, San Diego, California, USA, 2007.
44. McMaster, F. et al. Sour Service Corrosion Fatigue Testing of Flowline and Riser Welds.
Paper No. OMAE-57059, Proceedings of OMAE2008, 27th International Conference on
Offshore Mechanics and Arctic Engineering, Estoril, Portugal, June 2008.
45. Buitrago, J., Hudak, S. and Baxter, D. High Cycle and Low Cycle Fatigue Resistance of
Girth Welds in Sour Service. Paper No. OMAE-57545, Proceedings of OMAE2008, 27th
International Conference on Offshore Mechanics and Arctic Engineering, Estoril,
Portugal, June 2008.
46. Gui, F. et al. Corrosion Fatigue Performance of Duplex 2507 For Riser Applications,
Paper No. OMAE-20609, Proceedings of OMAE2010, 29th International Conference on
Offshore Mechanics and Arctic Engineering, Shanghai, China, June 2010.
47. DNV JIP Lined and Clad Pipeline Materials Phase 2, Guideline For Design and
Construction of Clad and Lined Pipelines, DNV Technical Report 2007-0220.
48. Holtam, C.M., Baxter, D.P., Ashcroft, I.A. and Thomson, R.C. An Investigation into
Fatigue Crack Growth test Methods in a Sour Environment, International Journal of
Offshore and Polar Engineering, Vol. 20, No. 2, pp. 103-109, June 2010.
49. Holtam, C.M. and Baxter, D.P. Fatigue Crack Growth Performance of Sour Deepwater
Riser Welds in the Near Threshold Regime. Paper No. 21279, Proceedings of OTC
2010, Offshore Technology Conference, Houston, Texas, USA, 2–5 May 2011
50. Guide to Methods for Assessing the Acceptability of Flaws in Metallic Structures, BS
7910 : 2005, British Standards Institution, London, UK, July 2005.
51. API Recommended Practice 579. Fitness-For-Service. American Petroleum Institute,
First Edition, January 2000.
52. R/H/R6. Assessment of the Integrity of Structures Containing Defects. British Energy,
Barnwood, Gloucestershire, Revision 3, 2001.
53. Recommended Practice DNV-RP-F108. Fracture Control for Pipeline Installation
Methods Introducing Cyclic Plastic Strain. Det Norske Veritas, January 2006.
54. Denys, R., Lefevre, A and De Baets, P. Weld and pipe material requirements for a
strain-based pipeline design. The Journal of Pipeline Integrity, Q1 2003.
55. Stewart, G., Klever, F.J. and Ritchie, D. An analytical Model to Predict the Burst
Capacity of Pipelines. 13th International Conference on Offshore Mechanics and Arctic
Engineering. 1994.
56. Lillig, D., Hoyt, D., Hukle, M., Dwyer, J., Horn, A. and Manton, K. Materials and Welding
Engineering for ExxonMobil high Strain Pipelines. Sixteenth International Offshore and
Polar Engineering Conference. 2006.
SAFEBUCK III
Design Examples
TABLE of CONTENTS
A1 Introduction ................................................................................................. 3
A2 Design Example #1: Short Hot Pipeline ..................................................... 4
A2.1 BASIC DESIGN PARAMETERS ................................................................... 4
A2.2 IDENTIFY SUSCEPTIBILITY TO BUCKLING ............................................... 6
A2.3 ASSESSMENT OF UNCONTROLLED BUCKLING ...................................... 7
A2.4 ASSESSMENT OF CONTROLLED BUCKLING ......................................... 11
A2.5 PIPELINE WALKING .................................................................................. 11
A2.6 SUMMARY.................................................................................................. 12
A2.7 UNITY CHECK CALCULATIONS................................................................ 13
A3 Design Example #2: Long Hot Pipeline ................................................... 16
A3.1 BASIC DESIGN PARAMETERS ................................................................. 16
A3.2 IDENTIFY SUSCEPTIBILITY TO BUCKLING ............................................. 18
A3.3 ASSESSMENT OF UNCONTROLLED BUCKLING .................................... 19
A3.4 ASSESSMENT OF CONTROLLED BUCKLING ......................................... 23
A3.5 PIPELINE WALKING .................................................................................. 27
A3.6 SUMMARY.................................................................................................. 27
A4 Design Example #3: Modest Temperature Pipeline ................................ 29
A4.1 BASIC DESIGN PARAMETERS ................................................................. 29
A4.2 IDENTIFY SUSCEPTIBILITY TO BUCKLING ............................................. 31
A4.3 ASSESSMENT OF UNCONTROLLED BUCKLING .................................... 31
A4.4 INFLUENCE OF BATHYMETRY................................................................. 37
A4.5 ASSESSMENT OF CONTROLLED BUCKLING ......................................... 38
A4.6 PIPELINE WALKING .................................................................................. 38
A4.7 SUMMARY.................................................................................................. 39
A5 REFERENCES ........................................................................................... 40
A1 INTRODUCTION
To illustrate the application of the SAFEBUCK design guideline example calculations are
performed. Notation used in the design example is as shown in Section 1.8 of the
SAFEBUCK Design Guideline.
Three design examples are presented:-
Design example #1: short hot pipeline;
Design example #2: long hot pipeline;
Design example #3: modest temperature pipeline.
These are chosen to illustrate significantly different pipeline behaviour and how the design
approach addresses these very different problems.
The design approach is outlined in Figure 1.
Select basic design
parameters
Modify design parameters Modify design parameters
No
No Is pipeline
Is walking OK? susceptible to
buckling?
Yes
Yes
Is
Yes uncontrolled
buckling
acceptable?
No
Yes
Is Can
controlled No initiation No
buckling strategy be
acceptable? improved?
Yes
buckle
No
interaction and
walking OK?
Yes
Design Complete
Parameter Value
Design Temperature 120°C
Seawater Temperature 4°C
Design Pressure 400 bara
Operating Temperature, inlet 120°C
Operating Temperature, outlet 110°C
Operating Pressure 200 bara
Unload Pressure 10 MPa
Slope of start-up transient 20°C/km
Fatigue Environment 200 full shut downs
A2.1.2 Friction
The seabed comprises soft to very soft clay. The lateral pipe-soil frictional resistance
developed for the 10-inch pipeline is illustrated in Figure 2.
0.5
0
0 0.5 1 1.5 2 2.5 3 3.5
Dis placement/Diameter
Minimum
Median
Maximum
Resistance Force
Submerged Weight
The susceptibility of a pipeline to lateral buckling is assessed by working through Section 3.2
of the SAFEBUCK guideline. For the design example, only the steel yield stress has been
temperature de-rated.
The SAFEBUCK design guideline states that the pipeline is not susceptible to buckling if the
maximum force in the pipeline is less than the critical buckling force:-
Nmax Ncr Equation 3.1
where:-
Where:-
L
Nf max Nend fmax Equation 3.4
2
N0 40MPa 0.039m2 1 2 0.3 2.05 105 MPa 0.020m2 1.17 105 o C1 116o C 50kN
N0 6.14 MN
In this calculation the internal pressure during installation is assumed to be zero and the
temperature at installation is seabed ambient (4°C). The maximum available frictional force
is given by:-
kN 2000m
N f max 100kN 1.0 1.0149 1115 kN
m 2
The maximum force in the system is the minimum of these two values:-
Nmax 1115 kN
Since there are no route bends in the pipeline, the critical buckling force is based on the
Hobbs infinite mode:-
EI HLB
Ncr 0.65 N 2.51 Equation 3.6
D
The analysis performed does not account for the influence of hydrodynamic lift, or break-out
axial friction, thus the susceptibility to buckling is slightly underestimated. However, the
maximum force in the pipeline (1064 kN) is greater than the 65% of the critical buckling force
(627 kN), and the system is susceptible to buckling.
N NBH
X H 2.588 L BH 2 0 Equation 4.2
fLB
So
6140 119.4
X H 2.588 95.1 2 4,185 m
0.4721
N̂ N
X S CR BH 2.588 L
BH Equation 4.5
fLB
The upper bound critical buckling force for a straight laid pipeline can be calculated from:-
EI HUB EI HBE
N̂cr Max 4.4 ,7.1 Equation 4.6
D D
2243 119.4
XS 2.588 95.1 4,745 m
0.472
So,
X Char min 4745,4185,1000 1,000 m
In accordance with the recommendations of the design guideline, the OOS parameters X NH
is assumed to follow a log normal distribution with mean 1.18 and standard deviation 0.35.
The characteristic force, Nchar is given by:-
EI Ws 31.7MN / m2 1.049kN / m
Nchar 3.86 3.86 1.347 MN
D 0.273m
The critical force distribution is calculated using the characteristic force, the lateral friction
distributions outlined above and the distribution of the normalised force parameter, XNH. The
resulting distribution of critical buckling force is log normally distributed with a mean of
1.45 MN and standard deviation of 0.45 MN, Figure 3.
1
Pcrit
Mean
0.8
0.6
Pdf
0.4
0.2
0
1 2 3 4
P (MN)
6 0.03
Elastic Elastic
Plastic Plastic
Unload 0.02
4
Curvature (m-1)
0.01
y (m)
2
0
0
0.01
2 0.02
0 20 40 60 80 0 20 40 60 80
x (m) x (m)
Parameter Value
First Load compressive strain 0.28%
Maximum equivalent strain 0.28%
Maximum stress range 400 MPa
For local buckling the unity check is defined as the imposed strain divided by the allowable
strain. In calculating the imposed strain a strain concentration factor of 1.2 is employed to
account for the effect of the coating stiffness mismatches.
For fatigue the unity check is defined as the fatigue damage due to the 200 full shutdown
cycles divided by the allowable fatigue damage (0.2). The calculation is performed for
fatigue from the weld root (sweet environment assumed) and cap (seawater plus CP
environment).
The unity checks are summarised in Table 6 (detailed calculations are presented in Section
A2.7).
Check Imposed Load Value Unity Check
Local Buckling Maximum compressive strain 0.335% 0.14
Fatigue Local stress range 530 MPa 0.88
Plasticity Maximum equivalent strain 0.28% 0.16
Stress range Nominal stress range 400 MPa 0.61
The seabed has no significant slope and is not connected to a SCR. Consequently, the only
active driver for walking is the thermal transient; the maximum transient slope is 20°C/km.
Since the start-up shutdown cycles involve relatively slow axial movement, the drained axial
response is appropriate.
A2.5.1 Pipeline Walking – No Buckle
Pipeline walking is assessed using mathematical models outlined in Appendix D.
The analysis is performed using the median drained axial friction coefficient, 0.65 and
employs 100 increments (k=100). The predicted walk is 26 mm per cycle.
A2.5.2 Pipeline Walking with a Buckle
In the event of a buckle the effective length of the pipeline is reduced. The pipeline is split
up into sections between the buckle locations and the ends of the pipeline. In this case one
buckle near the centre of the pipeline is the most likely buckled configuration. This gives two
virtual pipelines of equal length for the walking analysis. The length is effectively half the
pipeline length.
The analysis is repeated and gives a walk of 6.5 mm per cycle.
A2.5.3 Assessment
If buckling occurs then the walking rate is 6.5 mm per cycle. If buckling does not occur then
the walking rate is 26 mm per cycle. The design number of shutdown cycles is 200.
However, this is the maximum expected number of cycles. The process group estimate that
the most likely number of shutdown cycles is 100; the walking assessment is based on this
estimate. Over 100 cycles this walking rate is equivalent to an end expansion of 2.5 m. A
displacement of this magnitude could lead to overloading of the end spools.
The phenomenon should be further evaluated in detailed design using FE analysis with
accurate transient profiles.
Given the uncertainty in the phenomenon, the recommendation is to implement monitoring
procedures rather than pre invest in expensive mitigation techniques. It may be also
sensible to incorporated anchor flanges into the design.
A2.6 SUMMARY
The pipeline is susceptible to lateral buckling, although the probability of buckling is low (the
probability of a buckle forming is approximately 1 in 100). However, the assessment
demonstrates that uncontrolled buckling is tolerable and no initiation strategy is required to
control the response.
Pipeline walking may be a problem, particularly if no buckling occurs.
1.0 Introduction
This file performs design code checks in accordance with the SAFEBUCK II Design Guideline.
Units
Operational design life (years) Life : 20 Number of full shutdowns per year nfull : 10
Ambient Temperature a : 4 C
2.4 SAFEBUCK Data
Service
Bauschinger Factor f B : 0.8
Strain Concentration Factor SNCFf : 1.2 Maximum yield to tensile ratio YT : 0.92
Peak Compressive Strain cmax: 0.142 % Peak Equivalent Strain eqmax : 0.280 %
3.0 Calculations
Derating
corroded WT t : t f tfcorr
t 22.4 mm
D fo
D over t (corroded) Dot : Dot 12.192
t
t
Critical Strain c : 0.5 c 4.101 %
D fo
pi D fi pe D fo
Hoop stress h : h 33.463 MPa
2 tf
2
h
Hoop stress correction f p : 1 4 fp 1.022
SMYSfamb
Peak Compressive
fme : cmax f f fme
a 0.279 %
Mechanical Strain
Design strain 2 : fme SNCFf
2 0.335 %
c fp
Allowable strain All :
f2 All 2.407 %
2
Local buckling unity check UCLB :
All
UCLB 0.139
3.3 Fatigue
ef
Misalignment SCF Km : 1 3 Km 1.236
tf
mm
otherwise
16
R
Local Stress range RL : Kt Km
MPa
RL 529.984
12.015 3 log( )
Root hot spot - 10
xE ( , ND) :
E Curve ND
12.182 3 log( )
Cap hot spot - 10
xD ( , ND) :
D Curve ND
Imposed Damage nf
D fat : D fat 0.176
min N fi , N fe
Fatigue unity check D fat
UCfat : UCfat 0.881
D fa
3 h
Maximum allowable Rall : SMYSf 2 f B 1 Rall 657.6 MPa
stress range 4 SMYS f
25 140
120
20
Temperature (°C)
100
Pressure (MPa)
15 80
60
10
40
5
20
0 0
0 5 10 KP 15 20 0 5 10 15 20
KP
A3.1.3 Friction
The seabed comprises soft to very soft clay. The range of frictional values to be employed
in the analysis is outlined in Table 9.
Direction Minimum Median Maximum
Axial 0.3 0.5 0.8
Lateral Break-out 0.53 0.82 1.1
Lateral Residual 0.39 0.59 0.9
The susceptibility of a pipeline to lateral buckling is assessed by working through Section 3.2
of the SAFEBUCK guideline. For the design example, only the steel yield stress has been
temperature de-rated.
The SAFEBUCK design guideline states that the pipeline is not susceptible to buckling if the
maximum force in the pipeline is less than the critical buckling force. The fully constrained
force is (equation 3.3):-
N0 40MPa 0.039m 2 1 2 0.3 2.07 10 5 MPa 0.020m 2 1.18 10 5 o C 1 116 o C 50kN
This gives N0 = 6.17 MN. In this calculation the internal pressure during installation is
assumed to be zero and the temperature at installation is seabed ambient (4°C). The
maximum available frictional force is given by equation 3.4:-
kN 20,000m
Nf max 100kN 0.8 0.826 6.708 MN
m 2
The maximum force in the system is the minimum of these two values:-
Nmax 6.17 MN
EI HLB
Ncr 0.65 N 2.51 Equation 3.6
D
The analysis performed does not account for the influence of hydrodynamic lift, or break-out
axial friction, thus the susceptibility to buckling is slightly underestimated. However, the
maximum force in the pipeline (6.17 MN) is greater than the 65% of the critical buckling force
(566 kN), and the system is susceptible to buckling.
N NBH
X H 2.588 L BH 2 0 Equation 4.2
fLB
So
6170 84
X H 2.588 113 2 49.4 km
0.2478
N̂ N
X S CR BH 2.588 L
BH Equation 4.5
fLB
The upper bound critical buckling force for a straight laid pipeline can be calculated from:-
EI HUB EI HBE
N̂cr Max 4.4 ,7.1 Equation 4.6
D D
1990 84
XS 2.588 113 7,985 m
0.2478
So,
X Char min 7985 ; 49,400 ; 10,000 7,985 m
In accordance with the recommendations of the design guideline, the OOS parameters X NH
is assumed to follow a log normal distribution with mean 1.18 and standard deviation 0.35.
The characteristic force, Nchar is given by:-
EI Ws 31.7 MN / m2 0.826 kN / m
Nchar 3.86 3.86 1.195 MN
D 0.2731 m
The critical force distribution is calculated using the characteristic force, the lateral friction
distributions outlined above and the distribution of the normalised force parameter, XNH. The
resulting distribution of critical buckling force is log normally distributed with a mean of
1.29 MN and standard deviation of 0.40 MN, Figure 6.
1.5
Pcrit
Mean
1
Pdf
0.5
0
1 2 3 4
P (MN)
8000
7000
Characteristic VAS (m)
6000
5000
4000
3000 Deterministic
0
0 5 10 15 KP 20
For local buckling the unity check is defined as the imposed strain divided by the allowable
strain. In calculating the imposed strain a strain concentration factor of 1.2 is employed to
account for the effect of the coating strength mismatches.
For fatigue the unity check is defined as the fatigue damage due to the 200 full shutdown
cycles divided by the allowable fatigue damage (0.2). The calculation is performed for
fatigue from the weld root (sweet environment assumed) and cap (seawater plus CP
environment).
The results are presented in Figure 8.
2
1.8
1.6
1.4
Unity Check
1.2
1
Local Buckling
0.8
SAFEBUCK Strain
0.6
Fatigue
0.4
0.2
0
0 5 10 15 KP 20
Ncr XNV N V
where,
The parameter XNV describes the OOS in the vicinity of the vertical imperfection and is
represented by a log-normal distribution with a mean of 0.725 and standard deviation of
0.14.
The critical force distribution is calculated using the distribution of XNv and the deterministic
value of Nv. The resulting distribution of critical buckling force is log normally distributed with
a mean of 0.495 MN and standard deviation of 0.096 MN, Figure 9.
5
Vertical Prop
Mean
4
3
Pdf
0
0.2 0.4 0.6 0.8 1
Pcrit (MN)
4 Sleepers (rogues)
6000
No Triggers
5000 Triggers
Characteristic VAS (m)
4000
3000
2000
1000
0
0 5 10 15 KP 20
0.8
Unity Check
0.6
0.4
0.2
0
0 5 10 15 KP 20
The seabed has no significant slope and is not connected to a SCR. Consequently, the only
active driver for walking is the thermal transient; the maximum transient slope is 10°C/km.
Since the start-up shutdown cycles involve relatively slow axial movement, the drained axial
response is appropriate.
A3.5.1 Pipeline Walking – No Buckling
Pipeline walking is assessed using mathematical models outlined in Appendix D of the
Design Guideline.
The analysis is performed using the median drained axial friction coefficient, 0.65, rather
than the undrained values defied in Table 9. In the absence of buckling the pipeline
becomes fully constrained. In this case the criteria for walking to stop is f/f θ > 1 (Section
3.1.2 of Appendix D).
f Ws 0.65 826 N / m
A 1.12 1
f EA q 4.091x10 N 1.17 x10 5 C 1 0.01 C / m
9
A3.5.3 Assessment
If buckling occurs then the walking rate is 11.7 mm per cycle. The pipeline is highly
susceptible to lateral buckling and this is the most likely prediction. The design number of
shutdown cycles is 200. However, this is the maximum expected number of cycles. The
process group estimate that the most likely number of shutdown cycles is 100; the walking
assessment is based on this estimate. Over 100 cycles this walking rate is equivalent to an
end expansion of 1.17 m. The end spools could be designed to accommodate an
incremental end expansion of this magnitude. Alternatively, a more sophisticated analysis
could be undertaken to evaluate the interaction of the lengths of pipe between buckles; this
is likely to reduce the overall walking rate.
A3.6 SUMMARY
The pipeline is highly susceptible to lateral buckling. If no buckle initiation strategy is
employed, then the most likely number of buckles along the pipeline is five. The assessment
predicts that uncontrolled buckling is not acceptable; the limit state unity checks are violated
between KP 1 and KP 16.
A buckle initiation strategy using vertical imperfections in the form of pipeline sleepers is
selected. The design employs four 0.9 m high sleepers, located at KP 2.5, 5.6, 8.7 and 12.0.
This successfully reduces the unity checks to an acceptable level, Figure 12.
1.6
Uncontrolled
1.4 Rogues
1.2 Sleepers
Unity Check
0.8
0.6
0.4
0.2
0
0 5 10 15 KP 20
Parameter Value
Design Temperature 60°C
Seawater Temperature 4°C
Design Pressure 250 bara
Operating Temperature, inlet 60°C
Operating Temperature, outlet 10°C
Operating Pressure 200 bara
Unload Pressure 2.5 MPa
Slope of start-up transient 10°C/km
Fatigue Environment 100 full shut downs
60 25
50
20
Temperature (°C)
Pressure (MPa)
40
15
30
10
20
5
10
0 0
0 5 10 KP 15 20 0 5 10 KP 15 20
The seabed comprises fine sand. The range of frictional values to be employed in the
analysis is outlined in Table 14.
The susceptibility of a pipeline to lateral buckling is assessed by working through Section 3.2
of the SAFEBUCK guideline. For the design example, no yield stress temperature de-rating
has been employed.
The SAFEBUCK design guideline states that the pipeline is not susceptible to buckling if the
maximum force in the pipeline is less than the critical buckling force. The fully constrained
force is (equation 3.3):-
N0 25MPa 0.110m2 1 2 0.3 2.07 105 MPa 0.0195m2 1.17 105 o C1 56o C 150kN
This gives N0=3.59 MN. In this calculation the internal pressure during installation is
assumed to be zero and the temperature at installation is seabed ambient (4°C). The
maximum available frictional force is given by equation 3.4:-
kN 20000m
Nf max 100kN 0.8 1.172 9.5 MN
m 2
The maximum force in the system is the minimum of these two values, 3.59 MN.
The critical buckling force is defined as (equation 3.6):-
The analysis performed does not account for the influence of hydrodynamic lift, or break-out
axial friction, thus the susceptibility to buckling is slightly underestimated. However, the
maximum force in the pipeline (3.59 MN) is greater than the 65% of the critical buckling force
(0.837 MN), and the system is susceptible to buckling.
To evaluate this it is necessary to calculate the Characteristic VAS and undertake a lateral
buckling analysis at the Characteristic VAS.
N̂ N
X S CR BH 2.588 L
BH Equation 4.5
fLB
The upper bound critical buckling force for a straight laid pipeline can be calculated from:-
EI HUB EI HBE
N̂cr Max 4.4 ,7.1 Equation 4.6
D D
2995 261
XS 2.588 100.3 6,091 m
0.4688
So,
X Char min 6,091 ; 13,553 ; 10,000 6,091 m
4 3
log normal
min
3 median
2 max
Occurence
Occurence
2
1
1
0 0
0.5 1 0.5 1 1.5
The design guideline recommends a log normal distribution with mean 1.18 and standard
deviation 0.35 for the OOS parameters XNH. However, more specific data available to the
project indicates a mean of 1.11 and standard deviation 0.15 is more appropriate; this is
used in the example.
The characteristic force, Nchar is given by:-
EI Ws 77 .1MN / m 2 1.172 N / m
Nchar 3.86 3.86 1.820 MN
D 0.4064 m
The critical force distribution is calculated using the lateral friction distributions outlined
above and the distribution of the normalised force parameter, XNH. The resulting distribution
of critical buckling force is log normally distributed with a mean of 1.95 MN and standard
deviation of 0.63 MN, Figure 15.
1.5
Pcrit
Mean
1
Pdf
0.5
0
0.5 1 1.5 2 2.5 3 3.5
P (M N)
7000
6000
Characteristic VAS (m)
5000
4000
3000
Deterministic
2000
Probabilistic
1000
0
0 5 10 15 KP 20
The figure also shows the deterministic Characteristic VAS. This is much greater than the
value derived from the probabilistic assessment; this illustrates the benefit of undertaking the
probabilistic assessment.
1.6
Local Buckling
1.4
SAFEBUCK Strain
1.2 Fatigue
Unity Check
0.8
0.6
0.4
0.2
0
0 2 4 6 8 KP 10
EI W 77 .1 MNm 2 1.172 kN / m
NV 4 4 1.901 MN
H 0.4 m
The parameter XNV describes the OOS in the vicinity of the vertical imperfection. For a
sleeper this is represented by a log-normal distribution with a mean of 0.725 and standard
deviation of 0.14. There is no reason for this distribution to apply to the bathymetric feature.
However, no better information is available at this stage in the design, and the critical force
distribution is calculated using the distribution of XNV and the deterministic value of Nv.
The resulting distribution of critical buckling force is log normally distributed with a mean of
1.378 MN and standard deviation of 0.266 MN.
5000
4500 with Bathymetry
4000 Flat
Characteristic VAS (m)
The figure shows the Characteristic VAS for a flat seabed (in blue) and incorporating the
bathymetry (in red). The bathymetry is modelled as discrete locations of vertical OOS (the
axial locations are shown as green squares in Figure 18.
Incorporating the bathymetry in the model reduces the Characteristic VAS over the first
9 km. This is as a result of the increased likelihood of buckles occurring. Between KP 9 and
12 the Characteristic VAS increases, this is also due to buckling from the bathymetric
features in a region where the force is too low to cause buckling from the lateral OOS.
1.6
1.2 Flat
Unity Check
0.8
0.6
0.4
0.2
0
0 2 4 6 8 KP 10
The seabed has no significant slope and is not connected to a SCR. Consequently, the only
active driver for walking is the thermal transient; the maximum transient slope is 10°C/km.
Since the start-up shutdown cycles involve relatively slow axial movement, the drained axial
response is appropriate. Pipeline walking is assessed using mathematical models outlined
Appendix D of the Design Guideline.
A4.6.3 Assessment
The predicted walking rate is low, between 5 mm and 12 mm per cycle. Over 100 cycles the
total axial displacement would be between 0.5 m and 1.2 m. The end spools should be
designed to accommodate an incremental end expansion of this magnitude.
A4.7 SUMMARY
The pipeline is susceptible to lateral buckling. The assessment illustrates that the seabed
bathymetry is sufficient to initiate regular buckling and uncontrolled buckling is acceptable.
However, the conceptual analysis is based upon a very approximate model of the seabed
bathymetry. Detailed FE analysis should be employed as early in the project schedule as
practical to confirm the reliability of buckle initiation from the inherent bathymetry.
A very low level of pipeline walking is predicted in the post buckle configuration. Over 100
cycles this is less than an end expansion of 1.2 m. The end spools should be designed to
accommodate an incremental end expansion of this magnitude. Alternatively, given the
uncertainty in the phenomenon and the low walking rate, it may be adequate to implement
monitoring procedures rather than pre invest in expensive mitigation techniques.
A5 REFERENCES
A1. Det Norske Veritas, “Submarine Pipeline Systems”, DNV-OS-F101, January 2000
(reprint with amendments and corrections as of January 2003).
A2. SAFEBUCK Phase II. Probabilistic Buckle Formation: BUCKFAST Program. Atkins
Boreas Report JIP-04/BR08054/A. December 2008.
SAFEBUCK III
Pipe-Soil Interaction
TABLE of CONTENTS
B1. INTRODUCTION
B1.1. BACKGROUND
Lateral buckling, pipeline walking, route-curve pullout and flowline anchoring are all
extremely sensitive to pipe-soil interaction forces and there is significant uncertainty
associated with the characterization of these forces in design. For design, it is important to
bound behaviour. Upper and lower bound values of soil resistance are both important.
Uncertainty in soil behaviour will lead to a large range between upper and lower bound
behaviour, which is to be avoided. It is also important that the designer does not assume an
unrealistically narrow band of response.
B1.2. SCOPE
This APPENDIX B to the SAFEBUCK Design Guideline provides guidance on the pipe-soil
interaction response and the derivation of axial and lateral friction parameters for use in
pipeline design, as summarised in Lateral and Axial Interaction.
Response Description
Embedment Initial embedment, defined by the soil conditions and the loads during and
following installation is a significant influence on the axial and lateral
response that follows:
Axial Friction Axial breakout resistance First load response, a peak generally
only occurs after some set-up time
Axial residual resistance Large displacement response as the
pipe expands and contracts
Cyclic axial resistance Long term cyclic response under
shutdown/restart loading
Lateral Friction Lateral break-out Initial resistance to pipe displacement
from as-installed position, usually as a
buckle is triggered on first load
Lateral residual resistance Large displacement response as a
lateral buckle forms
Cyclic lateral resistance Long term response with soil-berm
growth under shutdown/restart loading,
with potential for increasing resistance
with embedment after repeated cycles
Table 1.1 Characteristic Pipe-soil Responses
SAFEBUCK has invested significant research effort to evaluate, quantify and understand the
complex lateral and axial pipe-soil interaction mechanisms for the low shear strength, very
soft, highly plastic clays that dominate deepwater regions, where on-bottom pipeline
solutions are currently being employed.
An extensive review of lateral response data from several recent projects donated to the
SAFEBUCK JIP, combined with tests carried out by the JIP, has led to a radical over-haul of
established geotechnical practice in pipe-soil interaction. While significant uncertainty
remains over pipe-soil interaction in such soils, a rigorous theoretical understanding of the
lateral response has emerged. Meanwhile work continues on developing a similar level of
understanding for the axial response.
Current understanding is presented, with models based on the most recent research.
B1.3. SYMBOLS
B1.4. ABBREVIATIONS
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1.0
1.1
1.2
1.3
1.4
1.5
1.6
Good HE Poor HE
Good BE Poor BE
Good LE Poor LE
Figure 3.1 Comparison of good and poor shear strength profiles for similar soils
In this case the upper and lower bound values for good quality T-bar data are based on
± one standard deviation of measurements taken at over ten sites in one field development.
It is notable that the good data includes a sound datum to the seabed surface and a good
record of the weak near-surface soils.
B3.2.1. Clays
Key data for clays include:
Shear strength profiles in-situ to about 2.0 m depth, in soft clays this should
preferably be carried out by in-situ T-bar and/or box core T-bar measurements;
Sensitivity based on cyclic T-bar response or similar;
Submerged soil unit weight;
Project specific laboratory test data using surface soil samples may be carried out
using specialist testing methods.
The soil strength is characterised by the undrained strength (Su), the unit weight () and the
sensitivity (St), which is the ratio of undisturbed shear strength to the fully remoulded shear
strength.
As the soil strength usually increases with depth, it is necessary to select a representative
value of soil strength. Models presented in this guideline generally employ the soil strength
at the pipe invert (directly under the embedded pipe).
B3.2.2. Sands
Key data for sands include:
Particle size distribution
Relative density
Submerged soil unit weight;
Project specific laboratory test data, using surface soil samples may be carried out
using specialist testing methods.
Figure 3.3 Pipe-soil model testing in the UWA geotechnical beam centrifuge
Many lessons have been learned in developing tests and interpreting test data for projects
and for the SAFEBUCK JIP. These lessons are vital to the ongoing value of project-specific
pipe-soil tests and the development and improvement of test facilities. Such testing forms
the basis for the generic models for the lateral response proposed in this guideline and
ongoing tests are expected to provide the same level of confidence to the development of
improved axial response models. More recently, data from operational pipelines has also
provided good validation for existing test results. Further operational data will be reviewed in
SAFEBUCK Phase III, to validate theoretical models.
Where:
Wi is the pipe submerged weight at installation, usually the empty submerged weight
klay is the a multiplier to account for the touchdown reaction during pipelay
Wf is the flooded pipe submerged weight
Ws is the operating pipe submerged weight
Equation B4.1 is intended to capture the maximum vertical download that could occur during
installation, hydrotest or in operation. On clay soils, the remoulded soil strength should be
employed for assessing embedment in the installation case (as discussed in Section B4.4.2).
The appropriate submerged weight at installation and the touchdown factor will depend on
project specific installation procedures.
EI .Wi 1
2
z 1
m ln( 1 ) ... B4.2
2.D.SL m 1 m
2
D
Where
R
m 1 ... B4.3
Wi
z
b
'.Doc z z 2.z z z
sin 1 4
2
V a .Doc .SuR 1.5. 1 21 1 ...B4.5
Doc 4 Doc Doc Doc Doc Doc
z
Where
D 0.5
oc
b
z '.Doc
2
V a .Doc .SuR 1.5. ... B4.6
Doc 4 2
z
Where
D 0.5
oc
Modelling Uncertainty
• Uncertainty shall account for variation in each input parameter (Section B7)
• Tests tend to follow the trend of the theoretical solutions for static penetration
• Variation of 10% to 20% from the calculated value may be due to
• rates of pipe penetration
• roughness of model pipe
• uncertainty of soil strength in the near-surface zone
• soil strength ratios, leading to varying contributions from soil buoyancy
• rates of increase of soil strength with depth
• These effects are overshadowed by the dynamic lay effects
• A modelling uncertainty factor of 1.5 proposed
zinit
2/3
0.037 V ... B4.7
D ' D2
oc
Where:
Doc is the outside diameter (over coatings)
γ' is the soil submerged unit weight
V is the effective pipe weight
zinit is the initial prediction of pipe embedment
This equation does not account for static or dynamic lay effects, covered separately above.
No assessment of the applicability of this model or testing of embedment in sandy
soils has been undertaken as part of the SAFEBUCK JIP. The potential to carry out
such testing and assessment is under consideration.
2
Axial pipe-soil resistance
1.5
0.5
-0.5
-1
Undrained Brittle
-1.5 2nd Cycle
-2 Undrained
Drained
-2.5
Axial Displacement Design
i
The term „residual‟ is used by analogy with the residual friction angle that is mobilised within fine-
grained soils after continued shearing along a single plane.
Report No: 5087471/01/A
Issue Date: 30 August 2011
SAFEBUCK III / Safe Design of Pipelines with Lateral Buckling Page B19
Design Guideline - Appendix B
This response can vary significantly for different types of soil, different weights of pipe and
different coating roughness. In clay soils, the velocity of the pipe also has a significant effect
on the breakout and residual axial friction, because pipe displacement generates excess
pore pressure, leading to undrained conditions that can either: (i) increase the level of
resistance due to negative excess pore pressure; or (ii) reduce the level of resistance due to
positive pore pressure. High velocity can also introduce strain-rate effects that increase the
axial resistance.
In general, excess pore pressure tends to increase with increasing pipe velocity,
giving a low residual axial resistance. If pipe displacement occurs very slowly, such
that excess pore pressure is dissipated as fast as it is generated, this ‘fully drained’
condition exhibits a gradual increase to a significantly higher level of residual friction.
This is illustrated in Figure 5.1 by the ‘drained’ (shown green) force-displacement
curve, which provides a high residual axial resistance.
While dissipation of the excess pore pressure from pipe-lay is expected to occur
relatively quickly, that does not prevent undrained conditions being re-established
under conditions of rapid or large pipe displacements, indeed SAFEBUCK tests have
demonstrated this. The reduction in axial friction due to increasing velocity is not well
understood but has been demonstrated in a number of tests carried out for the
Safebuck JIP, as illustrated in Figure 5.2 and Figure 5.3 reviewed by White et al[B26].
These undrained (fast) and drained (slow) responses generally define the highest
and lowest values of axial resistance, as illustrated by the ‘Undrained’ and ‘Drained’
response curves in Figure 5.1.
In typical field conditions, the response is likely to lie between the fully drained and
undrained conditions.
Currently the important transition from undrained to drained conditions is not well
understood. So that the global average value of axial friction (Best Estimate) might
lie close to the upper bound or close to the lower bound friction value measured in
tests.
Since undrained (fast) and drained (slow) responses generally provide very different
values of axial resistance it is important to consider the interaction between the
available resistance and the resulting pipe velocity. A structural analysis based on
drained parameters that yields high pipe velocities is meaningless, and vice versa.
A significant peak – leading to a brittle breakout response – can occur when the pipe
moves axially for the first time, or after some time at rest, at a rate which is fast
enough to generate negative excess pore pressure. This brittle response appears to
be influenced more by excess pore pressure generation at the pipe-soil interface,
rather than by changes in mobilised friction angle.
The ductile response – with no obvious breakout – can occur when the pipe has not
been at rest for sufficient time to allow excess pore pressure to dissipate.
0.7
“drained” residual friction
0.6
Residual friction factor (H/W')
0.5
0.4
0.3
0.2 JIP2-AS1
JIP2-AS2
0.1 JIP2-AS3
“undrained”
JIP2-AS4 residual friction
0
0.001 0.01 0.1 1 10
Speed (mm/s)
Figure 5.2 Effect of axial pipe velocity on residual friction factors in short pipe tests
0.8
“drained” residual friction
0.7
0.6
Residual friction factor(H/W')
0.5
0.4
0.3
Test JIP2-AL1
0.2
Test JIP2-AL2
Figure 5.3 Effect of axial pipe velocity on residual friction factors in long pipe tests
In fully-drained conditions the axial friction factor A = tan , where is the pipe-soil interface
angle of friction, as measured on a flat shear plate or tilt-table device.
In undrained conditions, excess pore pressure, generated around the pipe-soil interface,
reduces the contact stress with the soil matrix to the „effective stress‟, which discounts the
load carried by excess pore pressure, based on the average excess pore pressure ratio:-
u
ru ... B5.3
n average
Where D‟oc is the width of the contact area defined by the pipe embedment.
This reaches a maximum value of V/Doc, for ≥ 50% embedment.
The friction angles ( and ) are significantly influenced by the level of effective
stress. The use of an interface friction angle measured at a more usual geotechnical
stress level would significantly under-predict the axial resistance. For an on-bottom
pipeline, the effective stress levels of 2 to 10 kPa generated by typical pipeline
weights is below that considered in conventional geotechnics. At these stress levels,
the drained friction angle of soft clays significantly exceeds that measured by
traditional testing apparatus operating at usual effective stress levels. The difference
can be critical to pipeline design. A consistent trend of residual friction is
demonstrated for different soils in Figure 5.4.
0.85
Carbonate Clayey SILT
Carbonate Clayey Silty SAND
0.80
Carbonate SAND
Drained Residual Friction Factor
(Tan of Secant Friction Angle)
0.70
0.65
0.60
0.55
0.50
0 2 4 6 8 10 12
Effective Stress (kPa)
Figure 5.4 Typical Measured Drained Friction Factors for Specific Soils & Coatings
1.4
Measured Residual Friction Factor
Test Data
1.2
UB
1.0
BE
0.8 LB
0.6
0.4
0.2
0.0
0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0
Note: 'Wedging' embedment Total Vertical Stress (kPa)
enhancement f actor eliminated
Figure 5.5 Axial resistances for moderately rough pipe coating in WA soil
These tests employed polypropylene (PP) coating that has been moderately
roughened by the application of sintered PP in the final stages of manufacture.
There is significant scatter in the data for a variety of reasons, including some
uncertainty in the test conditions. To reduce the spread in the data, it is normalised
by the embedment enhancement (or wedging) factor, which increases with
embedment. The data is presented as friction factor against total vertical stress, with
estimated curves of the most appropriate Upper Bound (UB), Best Estimate (BE) and
Lower Bound (LB) values, defined by the equation:-
Report No: 5087471/01/A
Issue Date: 30 August 2011
SAFEBUCK III / Safe Design of Pipelines with Lateral Buckling Page B27
Design Guideline - Appendix B
μς = p.σv-q ... B5.8
Where the parameters ‘p’ and ‘q’ are given in Table 5.1.
Parameters p q
UB 1.40 0.40
BE 0.88 0.40
LB 0.35 0.20
Table 5.1 Axial friction test curve parameters for fits to data
These curve fits are not applicable below 1kPa and the minimum value of friction at
high pipe weights should not be taken as less than 0.25. These correlations are
known not to encompass some of the data from another site for which SAFEBUCK
has generated undrained axial friction data, shown earlier in Figure 5.2 and Figure
5.3, so specific values should be used with caution.
Residual resistance
Horizontal
,
displacement
Accumulating passive
Horizontal
resistance
displacement
(a) Side scan at buckle crown on operating pipeline (b) centrifuge modelling of cyclic displacement
Figure 6.3 Typical cross-sectional profiles and soil berms
Surface soil, swept ahead of the pipe on each cycle, builds up into berms at the extremes of
the pipe displacement. These berms offer significant resistance to pipe movement and
define the shape of the buckle in operation. Subsequent cycles of lateral movement lead to
a steady increase in the restraint provided by the soil berms.
The reaction at the soil berm is a function of:-
Cumulative pipe displacement, defined by the number and size of previous sweeps, and
hence the volume of material added to the berm;
The rate of displacement into the berm and the drainage characteristics of the soil,
hence whether the response tends towards undrained or drained behaviour;
The strength of the berm, defined for clay by remoulding under repeated cyclic loading
and consolidation;
Encroachment into the berm, defined by the additional displacement into the berm
compared to the previous sweep;
The pipe weight relative to the strength of the berm.
The amplitude of the cyclic sweeps and the depth to which the pipe descends defines the
volume of soil added to the berms. After the initial cycles, some soil from the berms tends to
collapse back into the trench, creating a bowl shaped trench and additional cyclic constraint.
Figure 6.4 Comparison of ‘light’ and ‘heavy’ pipe cyclic lateral responses
The „light‟ pipe (Figure 6.4a,b) showed a lateral response that matches closely the schematic
shown in Figure 6.1, with the residual friction factor remained approximately constant (0.3)
throughout each lateral sweep until the static berms were reached. The embedment of the
pipe progressively increased with cycles, reaching approximately 0.5D after 8 cycles. The
trajectory within each sweep was approximately horizontal (noting that the spatial axes of
Figure 6.4a are distorted), although the pipe rose slightly as the static berms grew, with more
soil being added. The pipe therefore reached the berms earlier in each sweep, and a gradual
rise in the mobilised berm resistance was recorded, reaching a maximum of 1.5 at the end
of the test, after 21 cycles.
By contrast the „heavy‟ pipe (Figure 6.4c,d) showed a contrasting lateral response, which is
comparable to Figure 6.2. For comparison, the first sweep from the „heavy‟ pipe test is
shown superimposed on the „light‟ pipe data (Figure 6.4b,c). During the first lateral sweep of
only 2D amplitude the pipe embedded deeper than 1D. The increasing pipe depth caused a
sharp increase in the passive resistance and no steady residual value was reached. During
the return sweep the same behaviour was observed and video footage showed that pipe
was buried beneath remoulded soil which passed over the pipe crown during each sweep.
Within 6 cycles of only 2D amplitude the pipe had reached an embedment of 3D, mobilising
a friction factor of almost 5.
Figure 6.5 Model illustrating cyclic seabed penetration & contact pressure variation
[Note: This complex model was used to validate simplified ‘flat’ seabed
frictional response curves]
Height (m)
Height (m)
0.5 0.5
0.0 0.0
850 900 950 1000 1050 1100 1150 850 900 950 1000 1050 1100 1150
-0.5 -0.5
1.5 1.5
1.0 1.0
0.5 0.5
0.0 0.0
850 900 950 1000 1050 1100 1150 850 900 950 1000 1050 1100 1150
Figure 6.6 Influence of touchdown load and lateral resistance on buckle shape
This example in Figure 6.6 includes a reduction in pipe weight by the addition of
buoyancy between 900m and 1100m, thus combining the vertical upset and local
weight reduction techniques. It is demonstrated here that the highest vertical load
occurs within the lateral buckle and is a significant influence on the shape of the
lateral buckle.
A design approach that provides an increase in lateral resistance in these zones of high
contact stress is necessary, usually by defining the lateral resistance as a function of the
local contact force between the pipe and the soil (lateral friction factor). This approach
allows the FEA to take account of seabed stiffness and bending of the pipe in the vertical
plane.
The use of a non-linear lateral friction factor tends to overestimate the breakout
friction slightly (which is generally conservative) while residual lateral friction factor is
not much affected.
The actual seabed stiffness should be calculated for a specific pipeline using the
embedment calculation in Section B4.
2
1.8 Breakout rate: 4.5 mm/min
1.6
Breakout rate: 1.5 mm/min
1.4
Friction factor, H/V (-)
1.2
1
0.8
0.6 Breakout rate: 0.9 kN/m/day
0.4
0.2
0
0 0.5 1 1.5 2
Horizontal displacement, u/D (-)
0.1
uMU/D=0.025+0.10(z/D)
0.09
Mobilisation displacement [u/D]
0.08
0.07
0.06
0.05
uMB/D=0.016+0.04(z/D)
0.04
0.03
0.02 uML/D=0.004+0.02(z/D)
0.01
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Initial Embedment [z/D]
Figure 6.9 Examples of lateral breakout displacements for large-scale tests with
static penetration (these exclude heavy pipe tests with no peak breakout)
Values in (Figure 6.9) are based on a limited dataset of large-scale West African clay
and Ønsoy clay tests, excluding ‘heavy’ pipe response where there is no discernable
breakout. However, pipes that have been ‘dynamically’ embedded by applying small
amplitude cyclic lateral displacements during the embedment phase, can incur much
Breakout Displacement
This is based on a review of all tests, excluding heavy pipe response
1.20
A
Mobilisation Displacement [u/D]
1.00 B
0.80 C
D
0.60
E
0.40 F
0.20 G
H
0.00
0.00 0.20 0.40 0.60 0.80 1.00 1.20 1.40
Figure 6.10 Examples of lateral breakout displacements for large- and small-scale
tests that include simulated dynamic loading during penetration
The mobilisation displacements shown above may be also exceeded in soils that
differ from those employed in these tests.
i 0.32
j 0.8
k 0.8
Figure 6.11 Lateral sweeping mechanisms and the effective embedment concept[B22]
Figure 6.12 Increasing embedment of the pipe based on normalised vertical load
Figure 6.13 Mid-sweep cyclic resistance with mid-sweep embedment – shallow depths
Figure 6.14 Mid-sweep cyclic resistance based on mid-sweep embedment – all depths
0.15 '.Doc
2
V
0.05 Fp '.Doc (5 ) ( z / Doc )1.25
2
If B6.8
'.Doc
2
V
V
0.05 Fp 2 '.Doc (z / Doc )1.25
2
If B6.9
'.Doc
2
A value of 0.6 is adopted for the friction coefficient (μ) by Verley & Sotberg[B18].
Extensive data used in developing this model covers diameters from 0.3 to 1.0m,
embedments from 1% to 35% and V/γ‟.D2oc values from 1 to 0.04.
The level of uncertainty associated with these equations is not known.
A set of centrifuge model tests in two sandy carbonate soils was donated to the JIP
capturing embedment, breakout and large amplitude sweeping. A SAFEBUCK
review[B27] compared these results with the above breakout models and found that
they over-predict the breakout resistance by a factor of 1.5 on average.
ii
There is a typographic error in the second equation from this reference which is corrected here.
Report No: 5087471/01/A
Issue Date: 30 August 2011
SAFEBUCK III / Safe Design of Pipelines with Lateral Buckling Page B46
Design Guideline - Appendix B
These are not bounds in the strict sense of the word. The estimated probability that the
actual value falls outside the bounds is 2.3%iii on either side.
These results apply for typical sand deposits. Conditions that could result in exceptionally
loose deposits, such as shallow gas, or artesian water are excluded. Also excluded are any
conditions that would result in a value of 50% of less than 0.5. (For this case the upper
bound could still be calculated from Equation 6.8 using 50%=0.5.)
Whereas the relative density and the variation of relative density with depth seems to be
important, the effect of relative density is not well understood and the upper bound estimate
has been raised slightly based on engineering judgement rather than statistical theory to
account for this. Additional effective lateral friction data, coupled with careful measurements
of the relative density could be used in future to reduce the uncertainty, possibly enabling
narrower bounds to be justified.
In addition, coating roughness is an essential parameter to measure in any future tests.
This recommendation assumes that the residual resistance is independent of initial
embedment, which is in contrast to Verley and Sotberg[B18], who assume that the residual
embedment is half the embedment at breakout, and then use equations B6.7 to B6.9 to
assess residual lateral resistance. Based on the dataset review for carbonate sands, this
approach does not provide an improved basis for assessment of lateral residual friction over
the Peek model[B27].
Often conditions are encountered in which a veneer of sand overlies a stiffer substrate. If
the thickness of the sand layer is less than half a pipe diameter, then some reduction in the
friction coefficients may be appropriate. In any case the upper bound may be used as a
conservative value no matter how thin the sand layer may be.
It must be emphasized that these lateral friction coefficients only account for the berms that
form for monotonic steady-state lateral movements of the pipe. Cyclic displacements of the
pipe during shutdown-restart are likely to result in larger berms and increasing resistance to
growth in lateral displacements, this is addressed, to a limited extent for sandy soils in
Section B6.3.3.
Centrifuge model tests in two sandy carbonate soils were donated to the JIP. The results for
large amplitude sweeping were compared with the above model in a SAFEBUCK review[B27],
which showed that the response is fairly well represented, with values of measure/predicted
residual friction factor ranging from 0.6 to 1.1 with an average of 0.9.
A key conclusion from this SAFEBUCK review is that improved recommendations for
both siliceous and calcareous sands should be based on a re-evaluation of the
controlling mechanisms rather than a simple recalibration of the empirical parameters
in the current guidance.
For the „lower bound‟ (2.3%) or „upper bound‟ (97.7%) value, the probability of exceeding this value is
iii
the same as the probability that a normally distributed random variable exceeds the mean value plus
two standard deviations.
Report No: 5087471/01/A
Issue Date: 30 August 2011
SAFEBUCK III / Safe Design of Pipelines with Lateral Buckling Page B47
Design Guideline - Appendix B
Tests performed by SAFEBUCK[B1, Annex B] in 2004 using a 250mm diameter pipe in
siliceous sand (D60 = 250 µm, D10 = 110 µm) were most informative about the
general response, although the test rig had some issues, common to most early
tests, due to uncertain vertical load control. Nevertheless, cyclic frictional behaviour
was reasonably consistent; the equivalent friction rising from an initial value of about
0.4 to a final value of about 1.0. Fluctuations are also present, as the pipe appears
to climb the active berm and then accrete a new berm in a series of waves. These
tests observed extensive berm building. The resistance experienced by the pipe at
the berms was between 3 and 5 times the vertical load. Berm encroachment (berms
collapsing back into the trench) was observed as the berms become established over
a number of cycles, giving increased resistance as the pipe reaches maximum
amplitude on each cycle. Over a large number of cycles, this is likely to result in a
bowl-shaped lateral trajectory for the pipe; such behaviour has been observed in
practice.
The SAFEBUCK review[B27] of a set of centrifuge model tests donated to the JIP also
captured large amplitude sweeping in calcareous sand and calcareous silty sand.
These tests showed that the berm walls stabilised at 10° to 15° slope, with a flat
trench between them, when the cyclic amplitude was large enough. Over 50 cycles
the mid-sweep friction factor (Hmid/V) appeared to be stabilising at between 0.3 and
0.6 at mid-sweep. However, this is not a steady value across the sweep (as happens
on the first sweep); instead (after a few sweeps) the resistance is almost zero as the
pipe retreats from one berm (moving down the slope), increasing more rapidly with
every sweep as it advances (up the slope) towards the other berm. Indeed the mid-
sweep resistance rises to merge with the increasing berm resistance, reaching
resistance factors (Hberm/V) of 0.8 to 1.5 after 10 cycles, and 1.25 to 2.5 after 50
cycles. Pushing through the berms mobilised broadly similar levels of resistance,
with no sudden rise to a much higher resistance.
The range of responses observed here will be affected by the pipe diameter, pipe
weight and interface conditions as well as the soil properties. So, although these
cyclic results provide an indication of the cyclic behaviour and the governing
mechanisms, values of pipe-soil resistance that lie outside of these ranges are likely
to be observed in other conditions. In addition, the number of cycles in these tests is
likely to be significantly exceeded in practice and the trends may vary at higher
numbers of cycles.
B8. REFERENCES
B1. Bolton, M.D., Cheuk, C.Y., White, D.J., and Bonab, M.H. Modelling of soil-pipeline
interaction for large amplitude deflections for the SAFEBUCK JIP. Cambridge
University Technical Services. July 2004.
B2. Bruton, D., White, D., Cheuk, C., Bolton, M., Carr, M. “Pipe/Soil Interaction Behaviour
During Lateral Buckling”. SPE Journal of Projects, Facilities and Construction.
1(3):1-9 (2006).
B3. Bruton, D., White, D., Langford, T., Hill, A. “Pipe-Soil Interaction Testing For Design
of a Deepwater Project Offshore Angola”. Society for Underwater Technology, First
Annual Subsea Technical Conference, Perth. February 2009.
B4. Carr M., Matheson I. & Peek R. Design Specification for Clad Pipelines Subject to
Lateral Buckling on a Flat, Soft Clay Seabed. SIEP Report EP 2005-5154 (Restricted
Issue). (2005)
B5. Cheuk C.Y., Bolton M.D. “A technique for modelling the lateral stability of on-bottom
pipelines in a small drum centrifuge.” Proc. International Conference on Physical
Modelling in Geotechnics, Hong Kong (2006).
B6. Cheuk, C.Y. and White, D.J. “Centrifuge modeling of pipe penetration due to dynamic
lay effects.” Paper OMAE2008-57302 Proc. Int. Conf. on Offshore Mechanics and
Arctic Engineering, Estoril, Lisbon (2008)
B7. Dingle, H.R.C., White, D.J. and Gaudin, C. “Mechanisms of pipe embedment and
lateral breakout on soft clay.” Canadian Geotechnical Journal (accepted for
publication 2007).
B8. DNV-RP-C207 Recommended Practice: Statistical Representation Of Soil Data
October 2010.
B9. Hill, A.J. and Jacob, H “In-Situ Measurement of Pipe-Soil Interaction in Deep Water”
Proc. Offshore Technology Conference, Houston, USA. Paper OTC 19528 (2008).
B10. Langford, T.E., Dyvik, R. & Cleave, R., “Offshore pipeline and riser geotechnical
model testing: practice and interpretation.” OMAE2007-29458. Proc. of the
Conference on Offshore, Marine and Arctic Engineering (2007)
B11. Looijens, P. and Jacob, H.: “Development of a Deepwater Tool for In-Situ Pipe-Soil
Interaction Measurement and its Benefits in Pipeline Analysis”. Proc. 31st Offshore
Pipeline Technology Conf., Amsterdam, The Netherlands (2008).
B12. Merifield R.S, White D.J. & Randolph M.F. “The effect of pipe-soil interface conditions
on undrained breakout resistance of partially-embedded pipelines.” Int. Conf. on
Advances in Computer Meth. & Analysis in Geomechanics, Goa, India. (2008).
B13. Najjar, S.N., Gilbert, R.B., Liedtke, E.A., McCarron, W. “Tilt Table Test for Interface
Shear Resistance Between Flowlines and Soils.” Proc. Conf. on Offshore Mechanics
and Arctic Engineering (2003).
B14. Palmer, A. C., “Touchdown indentation of the seabed.” Applied Oceans Research
(Submitted for publication 2008). Based on Safebuck Report “Touchdown Indentation
of the Seabed During Pipelaying” Feb 2007.
B15. Peek, R., “Pipeline on Sand – Resistance to Lateral Movements” SIEP Internal
Report Jan 2006.
Report No: 5087471/01/A
Issue Date: 30 August 2011
SAFEBUCK III / Safe Design of Pipelines with Lateral Buckling Page B51
Design Guideline - Appendix B
B16. Randolph M.F. & White D.J.. “Upper bound yield envelopes for pipelines at shallow
embedment in clay.” Géotechnique, (in press 2008).
B17. Randolph, M.F. & White, D. J., “Pipeline Embedment in Deep Water: Process and
Quantitative Assessment”. Offshore Technology Conference Houston USA Paper
OTC 19128 (2008).
B18. Verley, R.L.P. and Sotberg, T. A Soil Resistance Model Pipelines Placed in Sandy
Soils. International Conference on Offshore Mechanics and Arctic Engineering.
OMAE, 1992
B19. Westgate, Z.J., White, D.J., Randolph, M.F., & Brunning, P. Pipeline Laying and
Embedment in Soft Fine-grained Soils: Field Observations and Numerical
Simulations. OTC 20407. Offshore Technology Conference, Houston (2010 )
B20. White D.J. & Randolph M.F. “Seabed characterisation and models for pipeline-soil
interaction.” Proc. 17th Int. Offshore & Polar Engng. Conference, Lisbon, Portugal
(2007). [Note: a typographic correction to the published equation was agreed with
the authors]
B21. White D.J., and Cathie D.N. “Geotechnics for subsea pipelines”. Proc. 2nd Int. Symp.
on Frontiers in Offshore Geotechnics. Perth. 87-123 (2010).
B22. White D.J., Dingle H.R.C. & Gaudin C.. SAFEBUCK JIP Phase II: “Centrifuge
modelling of pipe-soil interaction”: Factual Report for Boreas Consultants
(SAFEBUCK JIP), ref. UWA report GEO 07396. 57pp. (2007) (confidential)
B23. White D.J., Take W.A. & Bolton M.D. “Soil deformation measurement using Particle
Image Velocimetry (PIV) and photogrammetry.” Géotechnique 53 7:619-63 (2003).
B24. White, D.J., Cheuk, C.Y. Cambridge University Technical Services. Lateral pipe-
soil interaction: data review – Report prepared for the SAFEBUCK JIP.
October 2005. Report SC-CUTS-0502R02
B25. White, D.J. and Cheuk, C.Y. “SAFEBUCK JIP: Pipe-soil interaction models for lateral
buckling design [- Clay soils]: Phase IIA data review”. Report to Atkins Boreas and
the SAFEBUCK JIP, UWA report GEO 09497v3. 185pp. June 2010.
B26. White, D., Ganesan, S.A., Bolton, M.D., Bruton, D.A.S., Ballard, J-C., Langford, T.
SAFEBUCK JIP: Observations of axial pipe-soil interaction from testing on soft
natural clays. Offshore Technology Conference. OTC 21249 May 2011.
B27. White D.J. “SAFEBUCK JIP: Pipe-soil interaction models for lateral buckling design -
Sandy soils: Phase IIA data review”. Report to Atkins Boreas and the SAFEBUCK
JIP, UWA report GEO 10520v2. 56pp. October 2010
B28. Zhou H., White D.J. & Randolph M.F. “Physical and numerical simulation of shallow
penetration of a cylindrical object into soft clay.” Proc. ASCE GeoCongress, New
Orleans (2008).
SAFEBUCK III
TABLE of CONTENTS
C1 General ........................................................................................................ 3
C1.1 SAFETY CLASS ........................................................................................... 3
C1.2 LOAD AND CONDITION FACTORS ............................................................. 3
C2 Limit States.................................................................................................. 4
C2.1 LOCAL BUCKLING ....................................................................................... 4
C2.2 FATIGUE AND FRACTURE .......................................................................... 4
C2.3 PLASTICITY LIMITATIONS .......................................................................... 4
C3 REFERENCES ............................................................................................. 5
C1 GENERAL
An enormous amount of effort has been expended in developing the formulation outlined in
OS-F101[C1] and there is common acceptance within the pipeline industry that this code is
the most widely applicable set of design rules. The review performed for SAFEBUCK
confirmed that the formulations outlined in OS-F101 are suitable for application to the lateral
buckling problem. However, the requirements of OS-F101 should only be applied to a
pipeline that is wholly designed, specified and fabricated to OS-F101.
C2 LIMIT STATES
C2.1 LOCAL BUCKLING
The buckling limit state can be checked in accordance with the displacement-controlled
criteria of OS-F101, Section 5 D608 and 609i.
For a seamless linepipe the local buckling performance of the pipeline is
compromised by the plateau in the stress-strain response. For failure on the Lüder
plateau the design equations may result in unacceptably high failure probabilities.
The suitability of the local buckling design equations for use at a lateral buckle has
been evaluated in the SAFEBUCK JIP[C2]. This found that the basic equations are
not acceptable when failure is at strains below the Lüder strain. For internal over
pressure this is only likely to apply to for pipes with D/t>30 and σh/σy<0.5.
i Alternatively, the user is free to employ the moment controlled criteria of OS-F101.
C3 REFERENCES
C1. Offshore Standard DNV-OS-F101. Submarine Pipeline Systems 2010. Det Norske
Veritas. October 2010.
C2. SAFEBUCK Phase II. Lateral Buckling SRA. Atkins Boreas Report JIP-04/ BR06066/ B.
10 December 2007.
C3. Recommended Practice DNV-RP-C203. Fatigue Design of Offshore Steel Structures.
2008. Det Norske Veritas.
SAFEBUCK III
TABLE of CONTENTS
D1 First Load Elastic Plastic Lateral Buckling Model .................................... 3
D1.1 MODEL FORMULATION .............................................................................. 3
D1.2 MOMENT CURVATURE RESPONSE .......................................................... 7
D1.3 SOLUTION PROCEDURE .......................................................................... 11
D1.4 MODEL VALIDATION ................................................................................. 12
D2 Cyclic Elastic Lateral Buckling Model ..................................................... 27
D2.1 MODEL FORMULATION ............................................................................ 27
D2.2 VALIDATION............................................................................................... 32
D3 Pipeline Walking........................................................................................ 47
D3.1 THERMAL TRANSIENTS ........................................................................... 47
D3.2 SEABED SLOPES ...................................................................................... 48
D3.3 RISER TENSION ........................................................................................ 48
D3.4 LIQUID HOLD-UP ....................................................................................... 49
D3.5 APPLICATION TO PIPELINES WITH LATERAL BUCKLING...................... 49
D3.6 NOMENCLATURE ...................................................................................... 49
D4 REFERENCES ........................................................................................... 51
LB q
Mode 3
y y
μW μW
z x
P P
q
q
L2 L
L
LB L2 … D1.1
2
Initially the central lobe is considered. If the slope of the pipe is small, then the free-body
diagram developed is shown in Figure 2.
q
S
P
P M+dM
S+dS
y+dy
M
y
dx
y
dS
q … D1.2
dx
From moment equilibrium:-
dM dy
P S 0 … D1.3
dx dx
Differentiating equation D1.3 and combining with equation D1.2 yields:-
d2M d2 y
P q0 … D1.4a
dx 2 dx 2
This is the governing differential equation. A similar approach for the secondary lobe yields:-
d2M d2 y
P q0 … D1.5
dx 2 dx 2
Since the derivation is based entirely on statics, there is no assumption on how the moment
varies, i.e. there is no assumption of linearity.
y0 0 (Symmetry) …i
L
y 0 (Definition of lobe end) … iii
2
L L
y y2 (Continuity of slope) … iv
2 2
L L
2 (Continuity of curvature) …v
2 2
L L
M M2 (Continuity of shear force) … vi
2 2
where subscript 2 refers to the response in the secondary lobe. At x=-LB we get:-
dM dy
P q x 0 … D1.4b
dx dx
In which, the integration constant is zero as a result of boundary conditions (i) and (ii). Now
the moment is a general function of the curvature, κ, and curvature is defined as:-
y
… D1.6a
1 y
2 3/2
In this equation the sign arises since convex upwards (as drawn in Figure 2) is a negative
curvature. Since the model has been developed for small slopes this equation should be
simplified to:-
y … D1.6b
However, at this stage this simplification is not employed; this is mathematically inconsistent,
but has been found to provide superior results. Equation D1.4b and D1.6a can be recast in
terms of the pipeline slope, θ (which is simply y’):-
dM
P q x 0 … D1.4c
dx
… D1.6c
1
2 3/2
dM dM d
… D1.7
dx d dx
d 1 2 3 2
… D1.8
dx 1 2
5/2
So substituting equation D1.8 into equation D1.4c and re-arranging yields:-
P q x 1 2
3/2
3 2
dM
1
2
… D1.9
d
This is a second order differential equation, which can be integrated to obtain a solution.
However, to undertake this procedure a relationship between the moment and curvature is
required; this relationship is considered later. In addition, initial conditions for θ and θ’ are
required. These are provided by the compatibility conditions (iv) and (v). In order to
evaluate these it is necessary to consider the solution in the secondary lobe.
M EI y 2 … D1.10
Which is consistent with equation D1.6b for small slopes. Substituting this into equation
D1.5 yields:-
Which is the normal small slope form of the differential equation, where:-
P
… D1.12a
EI
q
… D1.12b
EI
The general solution of equation D1.11 is:-
z2
y 2 A cos z B sin z Cz D … D1.13a
2 2
Where the axial ordinate z is used rather than x (see Figure 1), since this simplifies the
mathematics. Applying boundary conditions (iii), (vii), (viii) and (ix) yields the solution:-
y2
cos z sin z z
z2 1 … D1.13b
4 2
where:-
cos L 2
L 2 2 1
2 … D1.13c
sin L 2 L 2
There is still one unknown, the length of the lobe, L2. To evaluate this we consider boundary
condition (vi). In the secondary lobe, equation D1.10 holds, so differentiating this yields:-
dM2
EI y2 … D1.14a
dx
In the primary lobe the moment derivative is defined by equation D1.4b. Noting the
compatibility condition (iv), the boundary condition (vi) yields:-
L
y2 L 2 2 y2 L 2 0 … D1.14b
2
Substituting the solution for y2, equation D1.13b, into this yields:-
sin L 2 L 2 L 2 L 1 L 2
2
cos L 2 … D1.15
2 2
dθ
y y
z
θ σL
x z0 z0
Axial
force
Fw L dA … D1.16a
A
Where σL is the axial stress. Since thin wall theory is employed, the equation can be re-cast
as:-
/2
Fw 2 R t L d … D1.16b
/ 2
Where R is the pipe mean radius and t is the wall thickness. Similarly, the moment in the
pipe is given by:-
/2
M L y dA 2 R t L y d … D1.17a
A
/ 2
L z … D1.18
Where κ is the imposed curvature and z is the distance from the neutral axis. Equation
D1.18 is true for both elastic and elastic-plastic deformation. The stress in the pipe is:-
L E z … D1.19a
Equation D1.19a is true only for elastic deformation. The distance z can be written in terms
of the angular coordinate, θ as:-
z z0 R sin … D1.20
Substituting equation D1.19a and D1.20 into D1.16b and D1.17a yields:-
Fw
z0 … D1.21
2 R t E
and
M EI … D1.22
dθ
z2
y y
z
θ σL
x z0 z0
z1
Axial
force
Y-
Figure 4 Elastic-Plastic Stress Distribution
The material yields in accordance with the von Mises yield criteria. Under the thin wall
assumption, the von Mises yield condition is:-
2 2 2
- Y
L L H H
… D1.23a
Where σH is the hoop stress and Y is the material yield stress. This can be rearranged to
provide the axial stress at yield as:-
H 3 2
Y Y 2 H … D1.23b
2 4
in tension and:-
H 3
Y Y 2 H2 … D1.23c
2 4
in compression.
The curvature to cause first yield can be evaluated by combining equations D1.19a to D1.21
with equations D1.23b and D1.23c to give:-
Fw Fw
Y Y
y min 2 R t , 2 R t … D1.24
E R E R
For a curvature in excess of the yield curvature, the distance from the neutral axis to the first
yield point is simply:-
Y Y
z1 ; z2 … D1.25
E E
Y z Y z
1 a sin 0 ; 2 a sin 0 … D1.26
E R R E R R
The strain distribution is unchanged (defined by equation D1.18), but the stress distribution
is then:-
Y 1
L E z 0 R sin 1 2 … D1.19b
Y 2
The lateral buckling model often fails to find a solution because the moment curvature
response goes flat when severe bending occurs (resulting in the denominator dM/dκ→0). To
counteract this, the true stress can be considered to provide a more robust solution and the
following approximation is used within the model:-
Y 1 L 1
L E z 0 R sin 1 2 … D1.19c
Y 1 L 2
The axial force in the pipe can then be found by combining equations D1.16b and D1.19c
as:-
Y 1 2 1 z 0 Y 2 2 1 z 0
Fw 2 R t
R Y cos2 Y cos1
… D1.16c
E z 0 2 1 R cos2 cos1
Equation D1.16c and D1.26 can be solved to yield the neutral axis offset, z0. Then
equations D1.17a, D1.19c and D1.26 can be combined to yield:-
Y
cos1 Y 1 z 0 E z0
E
cos 2 Y 1 z 0
Y E z0
MR t
2
… D1.17b
E
R E 2 1 Y 1 Y 2
2 2
L
y2 L 2
2
… D1.27a
L
y1 y2 L 2
2
… D1.27b
There are still two unknowns in the solution; the axial force in the buckle, P, and the length of
the central lobe, L. These must be evaluated by iteration.
0
L … D1.28
0
Iteration continues until the slope is less than the acceptable tolerance. The scheme
becomes inefficient as the solution is approached, and could be replaced by a more
sophisticated scheme, for example based on Newton’s method.
P P W L2s
0 Ls … D1.29
EA 2 EA
Where P0 is the fully constrained force, EA is the pipe axial stiffness, μ is the coefficient of
axial friction and Ls is the slip length, which is defined as:-
X
Ls LB … D1.30
2
Where X is the buckle spacing. An alternative form of Equation D1.29 is required for an
isolated buckle in a very long pipeline. This is trivial to implement, but is not considered
here.
The axial displacement within the buckle is:-
1 L / 2 2
L2
P0 P
uLB LB y dx y22 dz
… D1.31a
EA 2 0
0
The integration for the central lobe must be undertaken numerically, however, the integration
over the secondary lobe can be performed analytically:-
L
1 2 2
y2 dz 9.5535 10 5 2 L72 … D1.31b
2 0
du uLB … D1.32
Pj 1 Pj
du j Pj Pj 1 … D1.33
du j du j 1
Iteration continues until the error in the feed-in is less than an acceptable tolerance.
Case 1 2 3 4
CA (mm) 6 3 0 5
Pressure (MPa) 20 10 30 20
The flowline wall thicknesses are calculated in accordance with OS-F101. The jacket wall
thicknesses are calculated in accordance with the collapse requirements of OS-F101.
However, the minimum jacket wall thickness is then increased to provide some bending
resistancei. For example the jacket wall thickness of design case 3 has been increased from
the collapse requirement of 15.9 mm to 22.2 mm. The PIP systems are assumed to be fully
bonded systems.
The FEA models all employ the following assumptions:-
The seabed is modelled using a Coloumb friction representation;
The axial friction coefficient is 0.4 and the lateral friction coefficient is 0.7.
Pipe elements are employed;
The models are initially straight. Buckling is initiated employing a small trigger force.
The force is removed during initial heat-up, such that it is zero when 50% of the
temperature has been applied;
The pipe material is elastic-perfectly plastic, with constant temperature properties.
8 P=20 MPa
7 θ=100ºC
6
5
Mathcad
4
Abaqus
Offset (m)
3
2
1
0
-10.88 0.90 0.92 0.94 0.96 0.98 1.00
-2
Distance along pipe (km)
-3
i
The local buckling of the jacket pipe is a critical design issue within the buckle. This failure mode is
an interaction between the external pressure and the imposed bending. If the jacket is sized for
collapse, then there is no reserve left to accommodate significant bending. Consequently, an
increase in the basic thickness is required.
10 P=20 MPa
θ=141ºC
8
6
Mathcad
Offset (m)
4 Abaqus
0
0.88 0.90 0.92 0.94 0.96 0.98 1.00
-2
-4
Distance along pipe (km)
12 P=20 MPa
θ=200ºC
10
8
6 Mathcad
Offset (m)
4 Abaqus
2
0
-20.88 0.90 0.92 0.94 0.96 0.98 1.00
-4
-6
Distance along pipe (km)
D1.4.1.2 Curvature
The curvature developed in the model is presented in Figure 8 and Figure 9.
0.02
0.01
Curvature (1/m)
0
0.90 0.92 0.94 0.96 0.98 1.00
-0.01
Distance along pipe (km)
-0.02
-0.01
Curvature (1/m)
-0.015
-0.02
-0.025
-0.03
200 C MCAD 141 C MCAD 100 C MCAD
-0.035
200 C Abaqus 141 C Abaqus 100 C Abaqus
-0.04
D1.4.1.3 Moment
The moment developed in the model is presented in Figure 10 and Figure 11.
200
150
100
50
Moment (kNm)
-100
-150
-200
-250
-300
4
Analytic
3 Abaqus
Offset (m)
2 P=10 MPa
θ=100ºC
1
0
0.36 0.38 0.40 0.42 0.44 0.46 0.48 0.50
-1
Distance along pipe (km)
-2
0.006
0.002
Curvature (1/m)
-0.006
Distance along pipe (km)
Analytic
-0.01 Abaqus
-0.014
-0.018
1500
500
Moment (kNm)
-1500 Analytic
Distance along pipe (km)
Abaqus
-2500
-3500
4
Abaqus
3
Analytic
Offset (m)
2
P=30 MPa
1 θ=118ºC
0
0.90 0.92 0.94 0.96 0.98 1.00
-1
4
Abaqus
3
Analytic
Offset (m)
P=30 MPa
2
θ=141ºC
1
0
0.90 0.92 0.94 0.96 0.98 1.00
-1
Distance along pipe (km)
-2
5 Abaqus
4
Analytic
Offset (m)
3
P=30 MPa
2 θ=200ºC
0
0.90 0.92 0.94 0.96 0.98 1.00
-1
Distance along pipe (km)
-2
D1.4.3.2 Curvature
The curvature developed in the model is presented in Figure 18 and Figure 19.
0.015
0.01
0.005
Curvature (1/m)
0
0.90 0.92 0.94 0.96 0.98 1.00
-0.005
Distance along pipe (km)
-0.01
-0.015
200 C Analytic 141 C Analytic 118 C Analytic
-0.02 200 C Abaqus 141 C Abaqus 118 C Abaqus
-0.008
-0.01
-0.012
-0.014
-0.016
-0.018 200 C Analytic 141 C Analytic 118 C Analytic
-0.02 200 C Abaqus 141 C Abaqus 118 C Abaqus
D1.4.3.3 Moment
The moment developed in the model is presented Figure 20 to Figure 22.
600
400
200
Total Moment (kNm)
-600
100
50
Flowline Moment (kNm)
-100
118 C Abaqus 141 C Abaqus 200 C Abaqus
600
400
Jacket Moment (kNm)
200
-400
The analytic model is in good agreement with the Abaqus results; the difference in the
maximum moment in either pipe is below 6%.
50
-250
-350
P=30 MPa
-450 θ=200ºC
400
300
Jacket Axial Stress (MPa)
200
100
Distance (m)
8
7
Analytic P=20 MPa
6 θ=110ºC
5 Abaqus
Offset (m)
4
3
2
1
0
-10.84 0.86 0.88 0.90 0.92 0.94 0.96 0.98 1.00
-2 Distance along pipe (km)
-3
0.006
0.004 Analytic
0.002 Abaqus
Curvature (m-1)
0
0.84 0.86 0.88 0.90 0.92 0.94 0.96 0.98 1.00
-0.002
Distance along pipe (km)
-0.004
-0.01
2000 Analytic
1500
Total Moment (kNm)
1000 Abaqus
500
600
P=20 MPa
400 θ=110ºC
Flowline Moment (kNm)
200
1500
P=20 MPa
1000
θ=110ºC
Jacket Moment (kNm)
500
-1500 Abaqus
-2000
The analytic model is in good agreement with the Abaqus results; the difference in the
maximum moment in either pipe is below 3%. The axial stress developed in the buckle is
presented in Figure 30 and Figure 31.
250
150
50 Distance (m)
-50 `
900 920 940 960 980 1000
-150
-250
500
400
Jacket Axial Stress (MPa)
300
200
100
Distance (m)
y
Load
Unload
Stationary Point
Unload
Stationary Point
q
q
y L/2
μW
x
S
where:-
S
2 … D2.2a
EI
q
… D2.2b
EI
L L
y1 y 2 (Continuity of displacement) … iii
2 2
L L
y1 y2 (Continuity of slope) … iv
2 2
L L
y1 y2 (Continuity of curvature) …v
2 2
L L
y1 y2 (Continuity of shear force) … vi
2 2
At x=LBu we get:-
for LBu LB
YLu
y2 LBu (Slope at end of lobe) … viii
0 for LBu LB
for LBu LB
YLu
y2 LBu (Curvature at end of lobe) … ix
0 for LBu LB
L
y1 YLu (Displacement at stationary point) … vii
2
x2
y1 A1 cosh x A 2 sinh x A 3 x A 4 YLu … D2.3a
2 2
x2
y 2 A 5 cosh x A 6 sinh x A 7 x A 8 YLu … D2.3b
2 2
The boundary conditions (i) to (viii) can be used to solve for the eight unknown coefficients.
This yields the solution:-
2 L4 x 2
y1 B cosh x B YLu … D2.4a
L4 4
1 4
2 L4 x 2
y2 B cosh x B sinh x B
x
B YLu … D2.4b
L4 4
5 6 7 8
Lu
L 3
L LBu 2 coshLBu sinh
YLu
2
… D2.5a
2 sinhLBu
L
B1 cosh … D2.5b
2
L L L2 L2
B 4 coshLBu sinh sinhLBu LBu Bu 1 … D2.5c
2 2 4 8
B5 … D2.5d
L
B 6 sinh … D2.5e
2
L
B7 … D2.5f
2
L L L 2
B8 coshLBu sinh sinhLBu LBu Bu … D2.5g
2 2 4
The two remaining boundary conditions, ix and x, are used to obtain the relationship
between the overall buckle length, LB, the central length, L, and the unload buckle length,
LBu:-
L 2
coshLBu sinh sinhLBu
1
YLu … D2.6a
2 2 2
L L L
cosh cosh 2 L Bu
LBu 2 3 L 2 1
2 2 2 4 16 2
… D2.6b
4 Y
YL 2 YLu Lu
2 2
Load
Effective Force
Buckle Unload
P
Ls LB
W L2s S S0
L s … D2.7
2 EA EA
Where S0 is the fully constrained force, EA is the pipe axial stiffness, μ is the coefficient of
axial friction and Ls is the slip length, which is defined as:-
X
Ls LBu … D2.8
2
Where X is the buckle spacing. An alternative form of Equation D2.7 is required for an
isolated buckle in a very long pipeline. In addition, this model is only applicable if the change
in force at the anchor points (extreme left and right of Figure 34) is less than the change in
fully constrained force. This will be true unless the pipeline is lowly loaded or the buckle
spacing is very high. These modifications are trivial to implement within the model, but are
not considered here.
The axial displacement within the buckle is:-
1 L / 2 2
LB
S S0
uLBu LBu y1 dx y22 dx
… D2.9a
EA 2 0
L/2
S S0 2
uLBu LBu 7 f L, LB … D2.9b
EA
Where
Using equations D2.7 and D2.9b yields the compatibility of axial displacements
requirement:-
S j 1 S j
j S j1 j1 S j … D2.11
j 1 j
Two values of S are required to start the iteration; these are taken to be the magnitude of the
compressive force in the buckle on load and twice this value. Iteration continues until the
error in the feed-in is less than an acceptable tolerance.
D2.1.7 Limitations
The model has the following limitations:-
It is only applicable for a buckle between virtual anchors;
The force in the buckle on unload must be tensile (this may not be true for small cycles);
The change in force at the virtual anchors must be less than the change in fully
constrained force.
D2.2 VALIDATION
The model developed is validated against FEA models for the four design cases developed.
The FEA models all employ the following assumptions:-
The seabed is modelled using a Coulomb friction representation;
The axial friction coefficient is 0.4 and the lateral friction coefficient is 0.7.
Pipe elements are employed;
The models are initially straight. Buckling is initiated employing a small trigger force.
The force is removed during initial heat-up, such that it is zero when 50% of the
temperature has been applied;
Since the analytic model is wholly elastic, the validation is performed against an elastic FEA.
However, to illustrate the errors associated with plasticity, the stress range is also compared
with an elastic-perfectly plastic FEA.
12
10
Abaqus Load Abaqus Unload
8
Analytic Load Analytic Unload
6
Offset (m)
-4
Distance along pipe (km)
7
Offset Range (m)
5 Abaqus
Analytic
3
-1
0.86 0.88 0.90 0.92 0.94 0.96 0.98 1.00
-3
-5
Distance along pipe (km)
200
150
100
50
Moment (kNm)
600
400
Axial Stress Range (MPa)
200
0
860 880 900 920 940 960 980 1000
-200 Distance along pipe (m)
-400
Abaqus Extrados Abaqus Intrados
-600 Analytic Extrados Analytic Intrados
600
400
Axial Stress Range (MPa)
200
0
860 880 900 920 940 960 980 1000
-200 Distance along pipe (m)
-400
Abaqus Extrados Abaqus Intrados
-600 Analytic Extrados Analytic Intrados
14
12
10 Abaqus Load
8 Abaqus Unload
6
Offset (m)
Analytic Load
4 Analytic Unload
2
0
-20.86 0.88 0.90 0.92 0.94 0.96 0.98 1.00
-4
-6
Distance along pipe (km)
3 Abaqus
Analytic
2
Offset Range (m)
0
0.86 0.88 0.90 0.92 0.94 0.96 0.98 1.00
-1
-2
-3
Distance along pipe (km)
200
150
100
50
Moment (kNm)
400
300
Axial Stress Range (MPa)
200
100
0
-100 860 880 900 920 940 960 980 1000
-200
Distance along pipe (m)
-300
-400
-500 Abaqus Extrados Abaqus Intrados
-600 Analytic Extrados Analytic Intrados
7
6
5 Abaqus Load
4 Abaqus Unload
Offset (m)
3 Analytic Load
Analytic Unload
2
1
0
-10.38 0.40 0.42 0.44 0.46 0.48 0.50
-2
Distance along pipe (km)
2500
1500
Moment (kNm)
500
-2500
Analytic Load Analytic Unload
-3500 Abaqus load Abaqus Unload
500
400
Axial Stress Range (MPa)
300
200
100
0
-100 380 400 420 440 460 480 500
Distance along pipe (m)
-200
-300
-400 Abaqus Extrados Abaqus Intrados
-500 Analytic Extrados Analytic Intrados
500
400
Axial Stress Range (MPa)
300
200
100
0
-100 380 400 420 440 460 480 500
Distance along pipe (m)
-200
-300
-400 Abaqus Extrados Abaqus Intrados
-500 Analytic Extrados Analytic Intrados
7
6
5 Abaqus Load
4 Abaqus Unload
Offset (m)
3 Analytic Load
Analytic Unload
2
1
0
-10.38 0.40 0.42 0.44 0.46 0.48 0.50
-2
Distance along pipe (km)
3000
2000
1000
Moment (kNm)
-3000
Analytic Load Analytic Unload
-4000 Abaqus load Abaqus Unload
400
300
Axial Stress Range (MPa)
200
100
0
-100 380 400 420 440 460 480 500
-200
Distance along pipe (m)
-300
-400
-500 Abaqus Extrados Abaqus Intrados
-600 Analytic Extrados Analytic Intrados
7
6
5 Abaqus Load
4 Abaqus Unload
Offset (m)
3 Analytic Load
Analytic Unload
2
1
0
-10.88 0.90 0.92 0.94 0.96 0.98 1.00
-2
Distance along pipe (km)
600
400
200
Moment (kNm)
-400
Distance along pipe (m)
-600
-800 Analytic Load Analytic Unload
-1000 Abaqus load Abaqus Unload
50
Flowline Stress Range (MPa)
-50 860 880 900 920 940 960 980 1000
-150
-250
-450
500
400
Sleeve Stress Range (MPa)
300
200
100
0
-100 860 880 900 920 940 960 980 1000
Distance along pipe (m)
-200
-300
Abaqus Extrados Abaqus Intrados
-400
Analytic Extrados Analytic Intrados
-150
-250
-350
-450
Abaqus Extrados Abaqus Intrados
-550 Analytic Extrados Analytic Intrados
500
400
Sleeve Stress Range (MPa)
300
200
100 distance (m)
0
-100 880 900 920 940 960 980 1000
-200
-300
Abaqus Extrados Abaqus Intrados
-400 Analytic Extrados Analytic Intrados
8
7
6
Abaqus Load
5
Abaqus Unload
4
Offset (m)
Analytic Load
3
Analytic Unload
2
1
0
-10.88 0.90 0.92 0.94 0.96 0.98 1.00
-2
-3
Distance along pipe (km)
2000
1500
1000
500
Moment (kNm)
200
Flowline Stress Range (MPa)
100
0
860 880 900 920 940 960 980 1000
-100
-300
500
400
Sleeve Stress Range (MPa)
300
200
100
0
Distance along pipe (m)
-100 860 880 900 920 940 960 980 1000
-200
Abaqus Extrados Abaqus Intrados
-300
Analytic Extrados Analytic Intrados
200
Flowline Stress Range (MPa)
100
0
860 880 900 920 940 960 980 1000
-100 Distance along pipe (m)
-200
-300
500
400
Sleeve Stress Range (MPa)
300
200
100
D3 PIPELINE WALKING
Simple models to address the driving mechanisms have been developed within the
SAFEBUCK JIP[D1,D2]; these are outlined here.
These models are applicable to single pipe flowlines. While the tension and seabed
slope models are applicable to PIP, the transient model is not due to the influence of
the outer pipe which may also be heated to above ambient temperatures. FEA is
required to assess of walking of PIP under thermal transient loading.
S0
f f* .... D3. 1
L
In this equation L is the relevant length of pipe. In the absence of buckling it is the
overall pipe length of pipe between expansion spools. If buckling occurs it is the
distance between buckles.
The analysis involves a simple incremental calculation. To start the calculation an
appropriate number of increments, k, must be defined; a value of k=50 is typical. The
anchor point increment is then given by:-
L
xA .... D3. 2
k
The values of xθ are then defined by:-
xA k0
x k .... D3. 3
2 f x A L x A
k xA k x A x k 1
2
f
k0
Where, f EA q .
f x 2A
2 k 1
L
x k
EA 2
2
L
f x k
f x A 2 k 1
2
2 L
w k x k 1 x k .... D3. 4
EA 2 EA 2
2
f x 2A 2 k 1 f L x k 1 x k x k x k 1
2
x k 1
L
EA 2 EA 2
The total walk per thermal cycle is given by:-
The analysis presented here is based upon a constant heat-up gradient of qθ (ºC/m).
If a non-linear profile is known, an average slope (over the mobilised hot end) should
be used in these equations.
The model is valid for f/fθ<1, but will yield reasonable result for higher values of this
ratio.
This mechanism will cause the pipeline to walk towards the cold end of the pipeline.
S0 Ws L sin f L cos Ws L tan .... D3. 6
EA f
The equation is based upon a constant seabed slope. If the slope varies slowly then
the average slope over the pipeline can be employed. If the slope varies significantly
along the length, a more complex FE assessment is required.
The equation is independent of the sign of φ; for this mechanism walking will always occur
down hill.
If the pipeline reaches a position of full constraint on first load or cyclically then no walking
will occur for this mechanism.
( S 0 S R f L ) S R
S 0 S R f L
EA f
R .... D3. 7
0 S 0 S R f L
For this mechanism walking will always occur towards the riser.
If the pipeline reaches a position of full constraint on first load or cyclically then no walking
will occur for this mechanism.
D3.6 NOMENCLATURE
The following notation is employed in the analytic walking models.
ii
Reference D2 was re-issued to OTC correcting Equation 11, and the corrected version is available
on the SAFEBUCK website. The definition of SC was corrected with the addition of a minus sign at
the start of each expression. This correction brought the definition of SC in line with the theory
presented for other walking mechanisms, which require the change in force between load and unload
to be defined as a negative value.
Roman Symbols
EA Pipeline axial stiffness N
f Axial pipe soil resistance N/m
fθ Thermal axial resistance, N/m
f* Mobilisation axial resistance N/m
k Integer counter in walking calculation -
kmax Maximum value of k in walking calculation -
L Total pipeline length or pipeline length between buckles m
qθ Gradient of the transient heat-up profile °C/m
So Full constrained effective axial force N
SR Riser tension applied to pipeline N
Ws Pipeline submerged weight N/m
xA Location of virtual anchor point in transient heat up m
xθ Location of thermal anchor point in transient heat up m
Greek Symbols
Coefficient of thermal expansion -
ΔT Walk per cycle due to thermal transient loading m
Δφ Walk per cycle due to seabed slope m
ΔR Walk per cycle due to SCR loading m
ΔSo Change in fully constrained force due to change in operating conditions N
Δw Incremental walk m
φ Seabed slope radians
D4 REFERENCES
D1. M. Carr, F. Sinclair, D. Bruton. Pipeline Walking – Understanding the Field Layout
Challenges, and Analytical Solutions developed for the SAFEBUCK JIP. OTC 2006,
paper no. 17945.
D2. Bruton, D.A.S.; Sinclair, F.; Carr, M. Lessons Learned From Observing Walking of
Pipelines with Lateral Buckles, Including New Driving Mechanisms and Updated Analysis
Models. OTC 20750 (2010).
D3. Hobbs, R.E, In-service Buckling of Heated Pipelines. ASCE Journal of Transportation
Engineering, Vol. 110 No. 2, March 1984.
D4. Hobbs, R.E. Liang, F. Thermal Buckling of Pipelines Close to Restraints. OMAE 1989.
SAFEBUCK III
E1 Introduction ................................................................................................. 3
E2 Buckle Initiation Techniques ...................................................................... 4
E2.1 Snake Lay ..................................................................................................... 4
E2.2 Local Weight Reduction: Distributed Bouyancy ............................................. 9
E2.3 Local Weight Reduction: Concrete Removal ............................................... 11
E2.4 Vertical Upset.............................................................................................. 11
E2.5 Zero Radius Bend ....................................................................................... 15
E2.6 Influence of Bathymetry............................................................................... 16
E2.7 Combined approaches ................................................................................ 17
E2.8 Other Issues................................................................................................ 17
E3 Mitigation Techniques for Lateral Buckling ............................................ 20
E3.1 Increase Axial friction .................................................................................. 20
E3.2 Decreased Lateral Friction .......................................................................... 20
E3.3 Hot lay......................................................................................................... 20
E3.4 Reduced design conditions ......................................................................... 20
E3.5 Influence of Pipeline Configuration .............................................................. 21
E4 Alternatives to Lateral Buckling ............................................................... 22
E4.1 In-line expansion spools .............................................................................. 22
E4.2 Trench and Bury ......................................................................................... 22
E5 REFERENCES ........................................................................................... 23
E1 INTRODUCTION
The lateral buckling design strategy will often involve a buckle initiation technique, to
guarantee the formation of regular buckles. All of the buckle initiation techniques considered
to date attempt to impose relatively large OOS features at known locations. The design
methodology then relies on these “engineered” features (or triggers) being more reliable than
the inherent OOS.
There are a number of methods which have been employed or are currently proposed to
initiate buckling at a controlled spacing; each buckle initiation method is described in Section
E2 of this Appendix.
Although the various initiation techniques are designed to develop buckles consistently,
there is a practical limit as to how close together the buckles can be formed. As the distance
between intended buckle sites is reduced, the guarantee of formation reduces. If the buckle
initiation sites are too closely spaced, some sites will not develop buckles and the
Characteristic VAS will be greater than the distance between triggers. Nevertheless, placing
triggers at a closer spacing than that required to meet the tolerable VAS could reduce the
Characteristic VAS. For example, if triggers are spaced at half the Tolerable VAS spacing
and 50% (every other) of the triggers fail to initiate buckling, the design would still be
acceptable. The advantage of this approach is the potentially improved probability of
meeting the Tolerable VAS. If the Characteristic VAS is insufficient to ensure integrity of the
pipe, mitigation strategies could be adopted to reduce the severity of the design problem.
Some mitigation techniques are described in Section E3 of this Appendix and some design
alternatives are described in Section E4 of this Appendix.
E2.1.1 Background
In snake-lay, the pipeline is laid in a series of gentle curves, as illustrated in Figure 1.
Pipeline
Offset Plan
(typically view
100m)
Lay Centre-line
180
Offset (m)
175
170
165 Buckled
As-Laid
160
6.1 6.2 6.3 6.4 Lateral Buckling
Distance along pipe (km)
400 P1 P2 P3 P4 P5 P6 P7
350
300
250
200
y (m)
150
100
X1 X2 X3 X4 X5 X6 X7
50
0
-50 0 2000 4000 6000 8000 10000 12000
x (m)
-100
i
The as-laid curves have a relatively large radius. The range 1000 m to 2500 m is typical.
Report No: 5087471/01/A
Issue Date: 30 August 2011
SAFEBUCK III / Safe Design of Pipelines with Lateral Buckling Page E7
Design Guideline: Appendix E
40 P1 P2 P3 P4 P5 P6 P7
30
20
10
x (m)
0
y (m)
-20
-30
-40 X1 X2 X3 X4 X5 X6 X7
-50
E2.2.1 Background
In the distributed buoyancy method discrete lengths of typically 60m to 200m of the
pipeline are installed with additional buoyancy on them; these lengths are the intended
buckle initiation sites as illustrated in Figure 4.
Buckle spacing
Buoyancy added (typically 2 to 3km)
to reduce weight
(Typically < 100m)
E2.4.1 Background
Pipeline buckling can be initiated in either the horizontal or the vertical direction. If the
pipeline is laid on an uneven seabed, then an initial vertical movement is highly likely to
develop into a lateral buckle. This approach takes advantage of this fact. The technique
works by deliberately introducing significant vertical OOS at a number of points along the
pipeline. Two techniques have been employed:-
Sleepers[E6,E7];
Gravel dump berms[E8,E9].
In the gravel dump option, the sleeper is replaced by a gravel berm; otherwise, the concept
is identical, although the more gradual rise in slope towards a gravel berm can modify the
formation response, by encouraging upward displacement of the pipe before it falls over to
displace laterally. On a sleeper upheaval is immediately accompanied by lateral
displacement so that the pipe should not lift off the sleeper. Sleepers and gravel dump
options have already been adopted by specific projects.
The sleepers are installed prior to (or during) pipe lay at the appropriate spacing, as
illustrated in Figure 5.
Elevation
view
Sleeper spacing
(typically 2 to 3km)
Pipeline
Sleeper
Sleeper
Flowline
ii
The buckles are described by their mode shape, defined by the number of buckle lobes. When
pipes buckle on sleepers in a symmetric mode there is usually a large central lobe on the sleeper,
while the outer lobe displacements are quite small, as they embed into the soil. Such a shape is very
close to being mode-I and are therefore described as such.
Report No: 5087471/01/A
Issue Date: 30 August 2011
SAFEBUCK III / Safe Design of Pipelines with Lateral Buckling Page E15
Design Guideline: Appendix E
Height (m)
Height (m)
0.5 0.5
0.0 0.0
850 900 950 1000 1050 1100 1150 850 900 950 1000 1050 1100 1150
-0.5 -0.5
1.5 1.5
1.0 1.0
0.5 0.5
0.0 0.0
850 900 950 1000 1050 1100 1150 850 900 950 1000 1050 1100 1150
Pipeline length
E5 REFERENCES
E1. Carr, M., Matheson, I., Peek, R., Sanders, P. and George, N. Load and Resistance
Modelling of the Penguins Pipe-in-pipe Flowline Under Lateral Buckling. 23rd
International Conference on Offshore Mechanics and Arctic Engineering. Vancouver,
OMAE 2004.
E2. Peek, R., Carr, M., Matheson, I., Sanders, P. and George, N. Thermal Expansion by
Lateral Buckling: Structural Reliability for the Penguins Flowline. 23rd International
Conference on Offshore Mechanics and Arctic Engineering. Vancouver, OMAE 2004.
E3. Matheson, I., Carr, M., Peek, R., Sanders, P. and George, N. Penguins Flowline
Lateral Buckle Formation Analysis and Verification. 23rd International Conference on
Offshore Mechanics and Arctic Engineering. Vancouver, OMAE 2004.Peek, R., and
Yun, H. Flotation to Trigger Lateral Buckles in Pipelines on a Flat Seabed. Journal of
Engineering Mechanics. ASCE. April 2007.
E4. Peek, R., and Yun, H. Flotation to Trigger Lateral Buckles in Pipelines on a Flat
Seabed. Journal of Engineering Mechanics. ASCE. April 2007.
E5. Peek, R., Kristiansen, N. Zero Radius Bend Method to Trigger Lateral Buckles.
OMAE2008-58046. Proceedings 27th International Conference on Offshore Mechanics
and Arctic Engineering. June 2008.
E6. Harrison, G.E., Brunner, M.S, and Bruton, D.A.S. King Flowlines – Thermal Expansion
Design and Implementation. Proceedings of the Annual Offshore Technology
Conference. OTC 15310, 2003.
E7. Jayson, D., Delaporte,P., Albert, J., Prevost, M., Bruton, D, Sinclair, F. Greater
Plutonio Project – Subsea Flowline Design and Performance. Offshore Pipeline
Technology Conference, 2008.
E8. Nystrom, P., Tornes, K., Karlsen, J., Endal, G., Levold, E. Design of the Asgard
Transport Gas Trunkline for Thermal Buckling. ISOPE 2001. Proceedings 11th
International Offshore and Polar Engineering Conference.
E9. Slettbo, H., Seek, D., Jorgensen, GM. And Dertvik, S. Asgard production start up
confirms Dynamic flowline Design Approach. Offshore Pipeline Technology (OPT)
2001, Amsterdam.
E10. Anderson, M., Bruton, D., Carr, M. The Influence of Pipeline Insulation on Installation
Temperature, Effective Force and Pipeline Buckling. 26th International Conference on
Offshore Mechanics and Arctic Engineering. 2007.
SAFEBUCK III
SAFEBUCK Deliverables
Andrew Palmer
Third Party Comments on SAFEBUCK Phase I Graham Stewart I Aug 2006
Pipe-Soil Interaction Models for Lateral Buckling Design – D.J. White & C.Y.
Clay Soils Cheuk II Jun 2010