Download as pdf or txt
Download as pdf or txt
You are on page 1of 24

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/276068771

Free Span Design according to the DNV-RP-F105 for Free Spanning Pipelines

Conference Paper · January 2002

CITATIONS READS

4 3,189

3 authors, including:

Olav Fyrileiv Knut O. Ronold


DVN GL D NV GL
45 PUBLICATIONS   248 CITATIONS    74 PUBLICATIONS   589 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Fracture toughness View project

Pile Reliability JIP View project

All content following this page was uploaded by Olav Fyrileiv on 29 January 2016.

The user has requested enhancement of the downloaded file.


OPT’02
Offshore Pipeline Technology
Amsterdam, 2002

FREE SPAN DESIGN ACCORDING TO THE DNV RP-F105


FOR FREE SPANNING PIPELINES

Olav Fyrileiv, Kim J. Mørk & Knut O. Ronold


Det Norske Veritas

ABSTRACT

An official Recommended Practice DNV RP-F105 for free spanning pipelines is to be issued early in
2002. The main objective of the Recommended Practice is to provide rational design criteria and
guidance on design of free spans subjected to wave and/or current loading.
The Recommended Practice is based on an update of DNV Guideline no 14. The update incorporates
the technical development within pipeline free span technology in recent design and R&D projects
performed by Danish Hydraulic Institute, Norsk Hydro, Statoil and Det Norske Veritas.
This paper will give a brief introduction to the changes in the new Recommended Practice compared
to Guideline no 14. Emphasis will be given on the major benefits obtained by using the
Recommended Practice compared to older methods/codes. A real free span case will be presented for
comparison and illustration.

1 INTRODUCTION

Free spans become a problem in pipeline design and operation due to uneven seabed or seabed
scouring effects. The costs related to seabed correction and span intervention is often considerable.
On the other hand the potential cost related to a fatigue failure of a pipeline are enormous. Despite of
these aspects, free spans are normally designed applying unduly conservative concepts and often
very simple analytical tools.
The DNV guideline no 14 (GL 14) for free spanning pipelines was issued in 1998 and has been used
on a large number of projects and gained general acceptance in the industry. This guideline has been

Olav Fyrileiv 30/01/02


1
reviewed and updated lately to account for technical development and research within this field, to
account for experience in applying the guideline and to improve the user friendliness. The new
document is issued as a Recommended Practice (RP).
The major benefits in using the new Recommended Practice comprise:
• A simplified screening criterion
• Complete fatigue criteria including wave loading
• An ultimate limit criterion
• Updated VIV models for pipe in trench
• Improved structural response quantities and
• Updated pipe-soil modelling

Standard industry practice with respect to free span design has up till recently been to apply some
sort of on-set criterion against VIV. This is a simple but often over-conservative approach. The
potential savings in engineering costs will, however, normally be more than exceeded by the costs of
unnecessarily seabed correction and span intervention works. The major objection is, however, that
no safety factor is given in the codes/standards for such a criterion. The consequence is that this
essential aspect is left to engineering judgement. This further implies that the safety level varies
substantially from one pipeline design to another. DNV has realised that there is a need for a simple
criterion to be used in early phase design and screening of survey results in order to emphasis when
more detailed analysis is required. Therefore the new RP-F105 contains such a criterion.
In shallow waters, the direct wave loading may be significant. The new RP-F105 gives guidance on
how to account for such loads in fatigue calculations and in the check of the ultimate limit state for a
free spanning pipeline.
In many cases local scouring or erosion underneath the pipeline causes the free spans. There has
been significant uncertainty on how to assess free spans with a low gap and/or located in a trench.
The new RP-F105 contains updated models for flow and pipeline response in the vicinity of the
seabed based on comprehensive CFD analyses.
In order to make the Recommended Practice more accurate and user-friendly, the structural response
models and pipe-soil modelling have been reviewed and significantly updated based on a large
number of finite element analyses.

2 DESIGN CRITERIA

2.1 General
The fatigue criterion as given in GL 14, has been kept without any changes except for a change in
the safety factors. If vibrations of a free span are allowed, the ultimate limit state (ULS) has to be
checked in addition to fatigue. This is to ensure that the span is not failing due to excessive loading,
e.g. by local buckling. The ULS criterion was also relevant for designs according to GL14, but it was
not explicitly stated how the extreme load effects should be calculated.
In addition to the fatigue and ULS criteria, a new screening criterion has been provided. This is a
simple but conservative criterion to be used for early design checks, first screening of survey results
et cetera.

Olav Fyrileiv 30/01/02


2
2.2 Fatigue criterion
The fatigue damage assessment is to be based on the accumulation law by Palmgren-Miner:
ni (2.1)
Dfat = ∑ ≤η
Ni
Dfat is the accumulated life time fatigue damage from environmental wave and current loading. η is
the allowable damage ratio and Ni is the number of cycles to failure at stress range Si defined by the
SN (fatigue) curve on the form:
N i = a ⋅ Si
−m (2.2)

m is a fatigue exponent (the inverse slope of the S-N curve) and a is the characteristic fatigue
strength constant defined as the mean-minus-two-standard-deviation curve.
The number of cycles, ni corresponding to the stress range block Si is given by:
n i = P(•)f v Texp (2.3)

P(•) is the probability of a (combined) wave or current induced flow “event”. fv is the dominating
vibration frequency of the considered pipe response and Texp is the time of exposure to fatigue load
effects (i.e. design lifetime).
The following safety factor format are used:
fv (2.4)
D fat = Texp ∑ (γ s S( γ f , γ k , γ on .) )m P(•) ≤ η
a
γf, γk, γs and γon denote partial safety factors for the natural frequency, damping (stability parameter),
stress range and on-set of VIV, respectively.

2.3 ULS criterion


The ultimate limit state (ULS) shall be checked according to the relevant criteria in DNV-OS-F101.
The most relevant one for ULS is the load-controlled local buckling criterion. Then static and
dynamic bending moments, axial force and pressure effects shall be accounted for.
The maximum dynamic bending moment due to VIV and/or direct wave action may be found from
the dynamic stresses:
2⋅I (2.5)
M E = σ dyn
Ds − t
where I is the moment of inertia, Ds the outer diameter of the steel pipe and t the wall thickness. σdyn
is the dynamic stress due to VIV and/or direct wave action and may be calculated as:
1  A 
in − line σdyn = max Sin ; 0.5 ⋅ Scr in ; SFM, max 
2  A cr  (2.6)
1
cross − flow σdyn = Scr
2
where Sin and Scr is the stress ranges due to in-line and cross-flow VIV, SFM,max is the maximum
stress due to wave action and Ain and Acr is the in-line and cross-flow stress amplitudes for a
deflection equal to a diameter.

Olav Fyrileiv 30/01/02


3
It should be noted that the dynamic stresses and bending moments are to be calculated without any
safety factors. The load effect, the bending moment, is to be factorised according to the criterion in
DNV-OS-F101.
In cases where the direct wave action is significant, detailed modelling of the soil response at the
shoulders may be required as the extreme conditions may cause large deformations/local sliding at
the shoulders. As a simplification, the boundary conditions for the free span may be assumed as
pinned-pinned for the calculation of wave actions and ULS check.

2.4 Screening criterion


The screening criterion is similar to the commonly used on-set criteria. The difference is that while
the on-set criteria is based on no vibrations at all, the screening criterion will allow for some
vibrations. However, due to its simplicity and therefore missing ability to capture the complexity of a
vibrating free span it must be more restrictive than the fatigue criterion.
The in-line screening criterion requires that the in-line natural frequency, f0,in, must fulfil:
f 0,in U c,100 year  L / D  γ in (2.7)
> ⋅ 1 − ⋅
γf VRin,onset ⋅D  250  α

where Uc,100 year is the 100 year return period current flow at pipe level, VRin,onset is the on-set value
for in-line VIV, D is the outer diameter of the pipe, L the span length, α is the extreme current flow
over total flow ratio and γf and γin are safety factors. The safety factors have been calibrated against
full fatigue analyses to provide a fatigue life in excess of 50 years. A lot of cases with different pipe
diameter, span lengths and environmental conditions have been analysed to establish the safety
factors.
A similar criterion is given for the cross-flow frequency.
In order to ensure that fatigue analysis due to direct wave action is not required the following
condition has to be fulfilled:
U c,100 year (2.8)
> 0 .5
U w ,1year + U c,100 year

If this is not the case, then a full fatigue analysis, due to in-line VIV and direct wave action, is
required.

2.5 Safety Factors


The safety factors to be used with the in-line and cross-flow screening criteria are listed below.
Table 2-1 Safety factors for screening criteria.
γin 1.15
γcf 1.3

Pipeline reliability against fatigue uses the safety class concept, which takes account of the failure
consequences, see DNV-OS-F101, Section 2. The set of partial safety factor to be applied with the
fatigue criterion is specified in Table 2-2 for the individual safety classes.

Olav Fyrileiv 30/01/02


4
Table 2-2 Safety factors for fatigue.
Safety Safety Class
Factor Low Normal High
η 1.0 0.5 0.25
1)
γs 1.05 (1.0)
γf 1.201) (1.15)
γk 1.30
γon 1.10
1) This safety factor is intended to be used in design when detailed data about span length, gap etc is not known. If a span is assessed in-service with
updated and measured span data, the safety factor in brackets may be used.

Comments that applies to the safety factors:


• η apply to both the Response Model and the Force Model
• γs is to be multiplied to the stress (S γS)
• γf applies to the natural frequency (fo/γf)
• γon applies to onset values for in-line and cross-flow VIV (VR,on/γon)
• γk applies to the stability parameter (KS/γk)
• For ULS the calculation of the load effects are to be performed without safety factors (γS = γf = γk
= γon = 1.0). The relevant ULS criteria and associated safety factors are given in DNV-OS-F101.

3 FATIGUE ANALYSES FOR DIRECT WAVE LOAD

The model used for fatigue analysis against direct wave is denoted a Force Model. It applies to the
in-line direction in conditions where the wave-induced flow is governing (i.e. substantially larger
that the current flow velocity) and in-line vibrations due to VIV are mitigated by the presence of
waves.
The approach is similar to standard fatigue analysis with Morison type of loading and may be
handled by Time Domain (TD) simulations or Frequency Domain (FD) analysis
Time Domain analyses with Rain Flow counting technique is tractable in case of large non-
linearity’s in the loading and structural response but may be extremely time consuming for most
practical purposes. The recommended approach in RP-F105 is a FD solution for the short term-
fatigue damage due to combined current and direct wave actions in a single sea-state based on:
• Palmgren-Miner approach using SN-curves;
• A linearized drag term in the Morison equation based on conservation of damage;
• The effect of co-linear mean current included in linearisation term;
• A narrow banded fatigue damage with semi-empirical correction to account for wide-band
characteristic;
The formulation presented in this document has been successfully verified against comprehensive
time domain simulations using Rain flow Counting techniques and applies to the vast majority of
free span scenarios with well defined boundary conditions, see e.g. Mørk and Fyrileiv (1998). The
formulation is based on the following assumptions:
• the main damage contribution comes from lowest natural mode, i.e. the excitation frequency is
far from the natural frequency for the higher order modes;

Olav Fyrileiv 30/01/02


5
• the effective mass, me, and standard deviation of the flow velocity, σU, are invariant over the free
span length, i.e. for span length less than the dominant wavelength.
The short term fatigue capacity against direct wave actions in a single sea-state characterised by (Hs,
Tp, θ) is given in the following form:
η ⋅ a 1 ⋅ S− m 1
THFM
S , TP , θ
= ×
f v ⋅ κ RFC ( m1 )
−1
  m   S  2   m   S  2 
 G1 1 + 1 ;  sw   + χ ⋅ G 2 1 + 2 ;  sw   (3.1)
  2   S    2   S  

κ RFC (m 2 ) a1 ( m 2 − m1 )
χ = S
κ RFC (m1 ) a 2
S = 2 2 σs ⋅ γ s

where σS and fv is the standard deviation of stress amplitude and vibration frequency derived from
the stress spectrum, respectively. The rain flow counting correction factor, κRFC, accounts for the
“exact” wide-banded damage, i.e. correcting the implicit narrow-banded Rayleigh assumption for the
stress amplitudes to provide results similar to those arising from a state-of-the-art rain flow counting
technique. Further, a1; a 2 and m1;m2 are SN curve constants, Ssw the stress range, for which change in
SN curve slope occurs and γs and η safety factor on stress range and usage factor, respectively.
Further:

G 1 (ϕ, x ) = ∫ e − t t ϕ−1 dt
x (3.2)
x
G 2 (ϕ, x ) = ∫ e − t t ϕ−1 dt
0

The FD solution is easily programmed and provides results within a few percent of the TD solution
at a very small fraction of the CPU time required for TD analyses.
In situations where quasi-static stress response can be assumed, a simplified fatigue assessment may
however be tractable rather than a complete TD or FD approach.
In such cases, the short term fatigue capacity against direct wave actions in a single sea-state
characterised by (Hs, Tp, θ) may be estimated as follows:
THFM
S ,TP ,θ
= η ⋅ a ⋅ S − m Tu (3.3)

where S is the quasi-static stress range response from a direct regular wave load at pipe level using
Morison’s equation and Tu is the mean zero up-crossing period.

3.1 Force Coefficients


The Force Model is based on the classical Morison’s equation:
& − C π ρD 2 &z&
P( x, t ) = g D (U − z& ) U − z& +g I U
(3.4)
a
4

Olav Fyrileiv 30/01/02


6
where ρ is the water density, D the outer pipe diameter, U the instantaneous (time dependent) flow
velocity, z the pipe in-line displacement, gD = 0.5ρDCD is the drag force term and gI = π4 ρD 2 C M is the
inertia force term.
A fundamental assumption is that the force coefficients are representative for the free span scenario
considered. In the new RP-F105, the drag coefficient, CD, and inertia coefficient, CM, to be used in
Morison’s equation is a function of :
• the Keulegan Carpenter number, KC;
• the current flow ratio, α = Uc/ Uc + Uw ;
• the gap ratio, (e/D);
• the trench depth, (∆/D);
• Reynolds number, Re;
• the pipe roughness, (k/D);
In addition also the cross-flow vibration level, (Az/D) influences the drag coefficient. The “base
case” force coefficients for a pipe (rough and smooth) located far from the seabed is taken in
compliance with recognised references, see Sumer and Fredsøe (1997), Blevins (1994), DNV
Classification Note CN30.5 (1991), DNV’81 (1981). It is noted that available sources provide
somewhat different recommendations in particular as a function of Reynolds number. In RP-F105,
the dependency of the Reynolds number is embedded in the cylinder roughness effect. The effect of
gap ratio, trench depth and current flow ratio have been assessed using comprehensive CFD
analyses, see Hansen et al, (2001).
The recommendations in RP-F105 follow GL14 to a large extent but gentle modifications are
introduced when deemed appropriate.

2.0
1.8
1.6
Drag Coefficient CD

1.4
1.2
α
1.0 0.0
0.1
0.8 0.2
0.3
ψ k ,CD = 1.0 0.4
0.6
ψ proxi,CD = 1.0 0.5

0.4
ψ trench,CD = 1.0
0.2 ψ A,CD = 1.0
0.0
0 5 10 15 20 25 30 35 40
KC

Figure 3-1 Drag coefficient CD versus KC and α

Olav Fyrileiv 30/01/02


7
The drag coefficient CD is to be taken as:
C D = C D,0 ⋅ ψ k ,CD ⋅ ψ proxi,CD ⋅ ψ trench ,CD ⋅ ψ A,CD (3.5)

where CD,0 is the basic drag coefficient in oscillatory flow for a free concrete coated pipe (KC>5),
see Figure 3-1.
Further, ψproxi,CD is a correction factor accounting for the seabed proximity. ψ k , CD is a correction
factor for the pipe roughness. ψtrench,CD is a correction factor accounting for the effect of a pipe in a
trench and ψA,CD is an amplification factor due to cross-flow vibrations, i.e., proportional to
increased projected area to mean flow.
The inertia coefficient, CM, is to be taken as:
C M = C M ,0 ⋅ ψ k , CM ⋅ ψ proxi, CM ⋅ ψ trench , CM (3.6)

where CM,0 is the basic inertia coefficient for a free concrete coated pipe taken as, see Figure 3-2:

2.0
α
1.8
0.0
1.6
0.1
Inertia Coefficient CM

1.4 0.2
1.2 0.3
1.0 0.4
0.5
0.8
0.6 ψ k,CM = 1.0
0.4 ψ proxi, CM = 1.0
0.2 ψ trench, CM = 1.0

0.0
0 5 10 15 20 25 30 35 40
KC

Figure 3-2 Inertia coefficient CM versus KC and α


ψ k , CM is a correction factor accounting for the pipe roughness. ψproxi,M is a correction factor
accounting for the seabed proximity. ψtrench,M is a correction factor accounting for the effect of a pipe
in a trench.
Explicit expressions for the correction functions are given in RP-F105.

4 VIV RESPONSE MODELS

Vortex Induced Vibration (VIV) is based on so-called Response Models providing a link between
the hydrodynamic parameters such as reduced velocity, Keulegan-Carpenter number, current flow
ratio etc and the vibration amplitude.

Olav Fyrileiv 30/01/02


8
The in-line VIV response model in RP-F105 is similar to the one in GL14 but with minor
modifications as follows:
• The transformation factor λmax has been removed since it assumes equal to 1.0 in most
applications.
• The mode shape factor ψmod assumes a value between 1.16-1.17 for the first mode and between
1.16-1.20 for higher modes. For free spans the variability is hence of minor importance compared
to the prevailing uncertainties and ψmod have been implemented in the maximum amplitudes
directly.
• The documentation from the MULTISPAN project has been reviewed, and the reduction
functions for turbulence intensity and flow angle have been revised slightly.
• The auxiliary factor, depending on the free span scenario, ψR is removed from the Response
Model part and is now considered in connection with the selection of safety factors.
• A reduction function, ψα,in, is introduced to account for reduced in-line VIV in wave dominated
conditions. The actual model is based on engineering judgement and provides a smooth transition
from the current dominated to the wave-dominated region.

The cross-flow model recommended in GL14 was deliberately taken to be on the conservative side.
The basic principle has been to approach the problem from the conservative side. Cross-Flow VIV is
still not fully understood, however, the RP-F105 represent the current state-of-the understanding in
case of single modal response.
The following changes have been implemented:
• The effect of a pipe in the vicinity of a trench has been based on CFD analyses on a relative
basis, see Hansen et al, (2001).
• A correction function for effective mass is based on a review of the MULTISPAN database.
• ψmod has been implemented in the maximum amplitudes directly, see comment for in-line above.
• The auxiliary factor depending on the free span scenario, ψR, is removed from the Response
Model part.
• The response for low KC numbers in the cross-flow response model is not in a narrow sense
related to the VIV phenomena. Compared to the GL14, model the case of low KC numbers has
been removed from the model but potential vibrations at low KC numbers must still be accounted
for if relevant, see RP-F105.
• Based on the high quality laboratory test performed at Marintek for the Ormen Lange pipeline,
the onset value has been modified slightly. On-set is now defined in connection with a vibration
amplitude of 10% of the diameter, and small vibrations may occur for a reduced velocity as low
as 2.0. This is in compliance with observations from full-scale test.

4.1 RP-F105 formulation for in-line


The In-line VIV induced stress range, Sin, is calculated by the response model:
Sin = 2 ⋅ A in ⋅ (A Y / D) ⋅ ψ α ,in ⋅ γ S (4.1)

where Ain is the unit stress amplitude (stress due to unit diameter in-line mode shape deflection),
(AY/D) the normalised in-line VIV response amplitude, ψα,in the correction factor for current flow
ratio α and γs the safety factor to be multiplied on the stress range.

Olav Fyrileiv 30/01/02


9
0.20
0.18
Riθ,1 = 1.0 Ksd=0.00
0.16
Inline VIV Amplitude (Ay/D)

Riθ,2 = 1.0
0.14 γon = 1.0
Ksd=0.25
0.12
0.10 Ksd=0.50
0.08
Ksd=0.75
0.06
Ksd=1.00
0.04 Ksd=1.25

0.02 Ksd=1.50

0.00
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0
Reduced Velocity VRd (=V R/γf)

Figure 4-1 Illustration of the in-line VIV Response Amplitude versus VR and KS.

0.9
Riθ,1 θ=60ο
0.8

0.7

0.6

0.5 Riθ,1 θ=45ο

0.4

0.3 Riθ,1 θ=30ο

0.2
Riθ,2
0.1 ο all angles
Riθ,1 θ=0
0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2
Turbulence Intensity, Ic

Figure 4-2 Reduction function wrt turbulence intensity and flow angle

RIθ,1(Ic,θ)∈ [0;1] and RIθ,2(Ic)∈ [0;1] are reduction functions accounting for the effect of the
turbulence intensity and angle of attack (in radians) for the flow, see Figure 4-2.
ψα,in is a reduction function to account for reduced in-line VIV in wave dominated conditions:
 0 .0 for α < 0.4 (4.2)

ψ α ,in = 2.5α − 1.0 for 0.4 < α < 0.8
 1 .0 for α > 0.8

Olav Fyrileiv 30/01/02


10
4.2 RP-F105 formulation for cross-flow
The cross-flow VIV induced stress range, Scr, due to a combined current and wave flow is assessed
using the following response model:
S cr = 2 ⋅ A cr ⋅ (A Z / D) ⋅ R k ⋅ γ S (4.3)
where Acr is the unit stress amplitude (stress due to unit diameter cross-flow mode shape
deflection), (AZ/D) the normalised cross-flow VIV response amplitude, Rk an amplitude reduction
factor due to damping similar to GL14 and γs the safety factor to be multiplied on the stress range.
1.5 2
α> ; all KC
1.4 3 ψ proxi,onset = 1.0
1.3
Cross-Flow VIVAmplitude (AZ/D)

1.2
ψ trench,onset = 1.0
1.1 2 ψ mass, onset = 1.0
1 α≤ ; KC > 30
3
0.9 ψ α,onset = 1.0
2
0.8 α ≤ ; KC < 10
γ on = 1.0
3
0.7
0.6
0.5
0.4
0.3
0.2
0.1 VRcr,onset
0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16
Reduced Velocity VR,d (=V

Figure 4-3 Basic Cross-Flow response model (∆/D=0; e/D>0.8)

The reduced on-set velocity for cross-flow VIV, VRcr,onset depends on the seabed proximity, trench
geometry and current flow ratio α while the maximum amplitude is a function of α and KC.
3 ⋅ ψ proxi , onset ⋅ ψ mass, onset ⋅ ψ α , onset ⋅ ψ trench , onset (4.4)
VRcr, onset =
γ on

ψproxi,onset is a correction factor accounting for the seabed proximity:


 1 e e (4.5)
(3 + 1.25 ) for < 0.8
ψ proxi ,onset =  4 D D
 1 else
ψmass,onset is a correction factor accounting for the specific mass (gravity) of the pipe:

Olav Fyrileiv 30/01/02


11
 1 1 ρs ρs (4.6)
 + for < 1 .5
ψ mass,onset = 2 3 ρ ρ
 1 else
ψα,onset is a correction factor accounting for the current wave ratio:
 α (4.7)
1 + for α < 0.5
ψ α ,onset = 2

1.25 else
ψtrench,onset is a correction factor accounting for the effect of a pipe in a trench:
∆ (4.8)
ψ trench ,onset = 1 + 0.5
D
∆/D denotes a relative trench depth given by:
∆ 1.25d − e
=
D D (4.9)

∈ [0;1]
D
The trench depth d is to be taken at a width equal to 3 outer diameters. ∆/D = 0 corresponds to a flat
seabed or a pipe located in excess of D/4 above the trench, i.e. the pipe is not affected by the
presence of the trench.

e d

Figure 4-4 Definition of Trench Factor

The restriction ∆/D < 1.0 is applied in order to limit the relative trench depth.

5 STRUCTURAL RESPONSE OF FREE SPAN

The new RP-F105 has also been updated with respect to structural response quantities such as natural
frequencies and associated stress ranges. The most essential changes are:
• Stiffening effect of concrete coating is given explicit
• Updated and simplified soil stiffnesses
• Improved accuracy of simplified response quantities (beam theory)

Olav Fyrileiv 30/01/02


12
5.1 Concrete coating effects
Standard industry practice, with respect to all types of coating, is to account for the additional loads
caused by the coating, e.g. increased weight and drag forces. On the other hand, any increased load
carrying capacity is normally to be disregarded. This is the case for several pipeline design codes,
see for example DNV OS-F101. However, a study made by Statoil revealed that the stiffening effect
of a thick concrete coating is significant and should be accounted for to obtain good estimates for the
structural response of a free span.
A comparison was done between natural frequencies measured on real free spans and corresponding
frequencies calculated by non-linear finite element (FE) methods. After refining the FE model with
respect to span geometry, support conditions including the soil stiffness and the effective axial force,
it was concluded that the reason for the remaining difference between measured and calculated
frequencies had to be the stiffening effect of concrete coating.
This effect was studied with basis in the analytical model of concrete coated pipes developed by
Ness and Verley (1996, 1995) to study strain concentrations during pipe lay. This model accounts for
cracking in the tensile zone and sliding between concrete coating and the steel pipeline if the shear
capacity of the corrosion coating is exceeded. In addition the increased, localised bending at the field
joints and the crushing of concrete are considered in the model. The last effect is, of course, not
relevant for a free spanning pipeline during vortex induced vibrations since the bending strain level
is low. The following simple expression for the stiffening effect of concrete coating was found:
0.75
 EI 
CSF = k ⋅  conc 
 EI steel  (5.1)
SCF = 1 + CSF
where CSF denotes the stiffness of concrete coating relative to the steel pipe stiffness and SCF is the
stress concentration factor due to the concrete coating and localised bending. k is an empirical
constant accounting for the deformation/slippage in the corrosion coating and the cracking of the
concrete coating. k may be taken as 0.33 for asphalt and 0.25 for PP/PE corrosion coating.
The cross-sectional bending stiffness of the concrete coating, EIconc, is the initial, uncracked stiffness.
The Young’s modulus for concrete may be taken as:
E conc = 10000 ⋅ f cn 0.3 (5.2)

where fcn is the construction strength of the concrete. Both Econc and fcn are to be in N/mm2.
It is important to note that in case the increased stiffness effect is utilised, the increased bending
stresses due to field joints also have to be accounted for.

5.2 Pipe-soil interaction

The new RP-F105 gives recommendations for static and dynamic stiffness to represent the pipe-soil
interaction on the span shoulders. The expressions for these stiffnesses as given in the old GL14 have
been used for this purpose, however, the expressions and their use are explained and described in
more detail in the new RP.
The static stiffness is based on bearing capacity formulas and refers to the virgin penetration of the
pipe on the shoulders. The expressions for dynamic stiffness are based on elastic halfspace theory

Olav Fyrileiv 30/01/02


13
formulas for rectangular foundations and provide lower-bound values for the dynamic stiffness at
small-amplitude deformations.
Some adjustments of the formulas have been made relative to the presentation in GL14. In particular,
they have been expanded to account for the effects of the weight of the free span where applicable,
and for the extent of the support length on the shoulder. The support length on the shoulder depends
much on the soil type, and recommendations for this length in lieu of data are given, calibrated by
FEM analysis.
The stiffness formulas also depend on pipe diameter, contact width, void ratio of soil and submerged
unit weight of soil and are thus rather elaborate in use. To simplify the stiffness calculations when
normal conditions prevail, when insufficient data are available and/or when no detailed stiffness
analysis is carried out, the new RP-F105 provides a set of simplified formulas for dynamic stiffness.
The dynamic vertical stiffness KV and the dynamic lateral stiffness KL are considered. The simplified
formulas are based on the detailed formulas and read:
2 ρs 1 (5.3)
K V = CV ( + ) D
3 ρ 3
2 ρs 1 (5.4)
K L = CL ( + ) D
3 ρ 3
where D is the pipe diameter in units of m, and where KV and KL come out in units of kN/m/m when
the coefficients CV and CL are taken according to Table 5-1 and Table 5-2. The soil type, which is
used as entry to these tables, is identified by the value of the friction angle ϕ for sand and by the
value of the undrained shear strength su for clay. The specific mass ratio ρs/ρ is defined as the ratio
between the mass of the pipe and the mass of the displaced water volume and is dimensionless. The
simplified formulas are valid for dynamic conditions with small-amplitude deformations. In addition
to the coefficients CV and CL, Table 5-1 and Table 5-2 also give simplified values for the vertical
static penetration stiffness KV,S for use when no detailed bearing analysis is carried out.

Table 5-1 Dynamic stiffness factor and static stiffness for pipe-soil interaction in sand
Sand ϕ(°) CV CL KV,S
type (kN/m5/2) (kN/m5/2) (kN/m/m)
Loose 28-30 16000 12000 250
Medium 30-36 22000 16500 530
Dense 36-41 32000 24000 1350

Table 5-2 Dynamic stiffness factor and static stiffness for pipe-soil interaction in normally
consolidated clay
Clay type su (kPa) CV CL KV,S
(kN/m5/2) (kN/m5/2) (kN/m/m)
Very soft <12.5 1800 1200 50-100
Soft 12.5-25 4200 2800 160-260
Firm 25-50 9000 6000 500-800
Stiff 50-100 15000 10000 1000-1600
Very stiff 100-200 33000 22000 2000-3000
Hard >200 36000 24000 2600-4200

Olav Fyrileiv 30/01/02


14
5.3 Simplified span response quantities
The most essential structural response parameter for free span calculations is the natural frequency.
In addition other response quantities such as dynamic stress amplitude, static bending moment and
static displacement may be needed for fatigue calculations and check of the ULS criterion. The most
accurate and general method to obtain these response quantities is obviously use of non-linear finite
element methods. However, these require comprehensive and detailed input and modelling and may
be cumbersome in case of preliminary free span design, when establishing conservative values for
maximum allowable span lengths and screening free span survey results. In these cases, response
quantities based on classical beam theory may be found appropriate.
In case classical beam theory solutions are used to estimate the natural frequencies, the main
uncertainties are related to the effect of the effective axial force and the support conditions provided
by the soil at the span shoulders. Restricting the use of beam theory solutions to tensile and moderate
compressive effective axial forces may solve the first aspect. By doing so, the sagging of the pipe
and associated non-linear effects are limited. Assuming a theoretical pinned-fixed boundary
condition is the normal approach towards the second aspect. Such a boundary condition is clearly not
physically correct for a free span and may be both overly conservative and non-conservative
depending on the soil stiffness and the span length.
The classical solution of the dynamic differential equation for a beam with axial force spanning
between two supports reads:

EI  N  (5.5)
f 0 = C1 ⋅ 1 + C 2 ⋅
4 

mL  PE 

where EI is the bending stiffness of the steel pipe, m the mass, L is the span length and N is the axial
force (positive in tension). PE is the Euler buckling force = π2EI/L2. C1 and C2 are boundary
condition coefficients with the values 1.57 and 1.0 for the pinned-pinned case and 3.56 and 0.25 for
the fixed-fixed case.
The expression above may provide good estimates for the frequency of a spanning pipe with some
minor modifications:

EI  S  (5.6)
f 0 = C1 ⋅ 1 + CSF ⋅ 1 + C 2 ⋅ eff 
m e L4  PE 

Here, CSF is the concrete stiffness factor, me the effective mass including hydrodynamical added
mass and content and Seff is the effective axial force (positive in tension). PE is now to be based the
so-called effective span length. The problem of estimating the frequency is now reduced to have
realistic values for the boundary condition coefficients, which should reflect the continuos pipe
supported on elastic foundation.
In the same way, the differential equation for a beam may provide solutions for the unit diameter,
dynamic stress amplitude:

Olav Fyrileiv 30/01/02


15
D ⋅ (D s − t ) ⋅ E (5.7)
A = C 4 (1 + CSF ) 2
L

where D and Ds is the outer (total and steel) pipe diameters, t is the steel pipe wall thickness and
C4 is a boundary condition coefficient.
A large number of non-linear FE analyses with varying span length, soil stiffness and axial force
have been performed to obtain a database for span response quantities. With basis in these results,
fitting of the boundary conditions coefficients have been made.
An essential parameter in the fitting was found to be the so-called effective span length, Leff. The
effective span length is the length of a virtual span with fixed-fixed boundary conditions having
the same frequency as the real span with flexible supports. A fit to the FE results gave the
following relations for the effective span length:
L eff 4.03 (5.8)
= 1+ 1.1
L  
 KL4 
1.08 + 4
 (1 + CSF)EI 
 
where K is the relevant soil stiffness (vertical/horizontal, static/dynamic). It can be seen that Leff/L
decreases as span length and soil stiffness, K, increase. The expression for effective span length
relates well to the results from the studies of Hetenyi (1946) and Hobbs (1986). However, their
works are based on iterative solutions of the governing differential equation.
The boundary condition coefficients based on the effective length factor are listed in Table 5-3. Here
both values for pinned-pinned, fixed-fixed and single span on a flexible seabed are given.
Table 5-3 Coefficients for idealised boundary conditions
Pinned- Fixed- Single span on seabed
Pinned Fixed
C1 1.57 3.56 3.56 / (Leff/L)2
C2 1.00 0.25 0.25
C4 4.93 14.1 Shoulder: 1.11 C12
Mid-span: 2.42 C1

As an example the values for the C1 coefficient are shown in Figure 5-1 for a typical 40” pipeline.
The FE results are shown with symbols, while the corresponding, fitted values as given by the
expressions above are shown with lines. As seen, the fit is very good giving values close to the FE
results. It should also be noted that the C1 coefficient is strongly dependent on the soil stiffness and
the span length. It is seen that a fixed-pinned value of 2.5 would provide conservative estimates of
the frequency for long spans, while for short spans and especially on very soft soils it will be non-
conservative. For the sake of curiosity it can be mentioned that in the extreme case, having a L/Ds of
20, even the pinned-pinned case will be non-conservative.

Olav Fyrileiv 30/01/02


16
Horizontal frequency Neff/Pe=0, low sag,
RP fit
3.6

3.4

3.2

3.0

2.8

2.6

2.4
C1

2.2
clay-very soft
2.0 clay-soft
clay-firm
1.8 sand-loose
sand-medium
1.6 sand-dense
clay-very soft
1.4
clay-soft
clay-firm
1.2
sand
1.0
0 20 40 60 80 100 120 140 160
L/Ds

Figure 5-1 C1 for different span lengths and soil types, horizontal direction, 40” pipe. FE
results shown by symbols and estimated values by lines.

6 CASE STUDY

To demonstrate the effects of the main updates in the new RP-F105, a simple case have been
analysed using a 40” gas export pipeline with a D/t of 42. The pipeline is coated with 100 mm
concrete and is located on 80m water depth in a typical North Sea environment. As the considered
stretch is far away from the inlet, the temperature of the gas is equal to the ambient temperature. The
operating pressure is 130 bar. The seabed consists of loose sand, which causes free spans due to
scouring. A design life of 50 years is assumed in the calculations.
First different codes, DNV Class Note 30.5 (CN 30.5), DNV Guideline no 14 (GL 14) and the new
RP-F105, have been compared. The results in terms of fatigue lives (minimum of in-line and cross-
flow) are shown in Figure 6-1. Here the fatigue lives have been calculated using the same
frequencies and associated stresses for the spans in order to compare directly the changes in response
models and associated sets of safety factors.
From the comparison it is seen that the old CN 30.5 allows longer spans than the newer codes. This
is mainly caused by the non-conservatively high onset value for cross-flow and the use of safety
factors on frequencies in the newer codes. The high onset value for cross-flow in CN 30.5 implies
that cross-flow is not governing for any of the span lengths shown, while for GL14 and the new RP,
cross-flow is governing for fatigue lives equal to the design life. From Figure 6-1 it is also seen that
GL14 yields more conservative results than the RP. This is as expected and mainly due to the slight
relaxation of safety factors in RP-F105 in order to make it more consistent with the fatigue criterion
in DNV-OS-F101.

Olav Fyrileiv 30/01/02


17
It may be concluded that the new RP-F105 allows longer spans than GL 14, while using the old CN
30.5 is not recommended as it does not reflect the updated knowledge and data about response model
in general and onset of cross-flow VIV in particular.

Fatigue lives using different codes


(different response model and safety factors, same frequencies)
10000
RP-F105
GL 14
CN30.5
50 years

1000
Fatigue life (years)

100

10

1
40 45 50 55 60 65 70 75 80
Span Length (m)

Figure 6-1 Comparison between fatigue lives using different codes, i.e. different response
models with associated sets of safety factors.

When scouring causes free spans, a trench will develop underneath the span. The effect of such a
trench is accounted for in RP-F105. Its effect on the fatigue lives is clearly illustrated in Figure 6-2
where the same spans have been analysed with and without a trench. The trench will influence on the
cross-flow response, the cross-flow induced in-line and the in-line direct wave action. It is seen that
for the design life of 50 years, the allowable span length increases from 60m to 67m as the trench
depth reaches 1.5m.

Olav Fyrileiv 30/01/02


18
Effect of trench

10000
no trench
trench - 0.7m
trench - 1.5m
50 years

1000
Fatigue life (years)

100

10

1
40 45 50 55 60 65 70 75 80
Span Length (m)

Figure 6-2 Effect of pipe in trench.

Another effect accounted for in the new RP-F105 is the stiffening effect of concrete coating. In this
case, with a relatively thick coating and a thin steel wall, the relative increase in total bending
stiffness of the pipe cross-section is estimated to 30%. As seen from Figure 6-3, this increases the
allowable span length from 60m to 64m.
The concrete coating will have a beneficial effect on the natural frequencies as the total bending
stiffness increases. However, since most field joint coatings do not provide the same stiffening
effect, they will give rise to bending concentration in case of vibrations caused by VIV. Therefore, a
stress concentration of the same magnitude as the stiffening effect is conservatively assumed for the
fatigue calculations caused by VIV. In case of direct wave loading, the load and therefore the
bending moment, will be the same in the girth welds irrespective of the stiffening effect of the
coating. These aspects is clearly shown in Figure 6-3 where the force model (direct wave action)
dominates the fatigue life for span lengths shorter than 50m while the VIV is dominating for longer
spans.

Olav Fyrileiv 30/01/02


19
Fatigue lives using RP-F105

10000
no concrete stiffnening
concrete stiffnening
50 years

1000
Fatigue life (years)

100

10

1
40 45 50 55 60 65 70 75 80
Span Length (m)

Figure 6-3 Concrete coating stiffening effect on fatigue life.

One significant update in the new RP-F105 is the concept of effective span length and the associated
boundary condition coefficients. This update will of course have no effect if the structural response
quantities are established by use of finite element analysis. However, in many projects, the simplified
beam expressions are used, and standard practice until now has been to apply some sort of pinned-
fixed boundary condition.
A comparison between the natural frequencies found by FE analyses and the new RP-F105 estimates
using simplified boundary conditions, is shown in Figure 6-4. The FE analyses were performed at a
real 40” pipeline case and observed free spans using the measured seabed configuration and span
lengths. The following comments to the FE analyses applies:
• No concrete coating stiffening effect was accounted for.
• A low added mass coefficient, Ca = 1.1-1.2, was applied.
• An unrealistic low dynamic soil stiffness (300 kN/m/m) was applied.
In order to compare the results, the same added mass and soil stiffness were used in the RP-F105
expressions for the horizontal frequency as in the FE analyses.
As seen from the figure, the boundary condition for a free span may be considered as close to
pinned-pinned for short spans. When the span length increases, the effective boundary condition
approaches asymptotically the fixed-fixed case. This tendency is also shown in Figure 5-1 and is
confirmed by the FE analyses of the real spans. The reason for the small scatter shown in the FE
results is partly different gaps and thereby different added masses, and partly the fact that all the real
spans are not idealistic single spans on a flat seabed but may be affected by small neighbourhood
spans. Anyway it may be concluded from this comparison that the updated expressions in RP-F105
provides, for this case, natural frequencies of the same accuracy as a non-linear finite element
analysis.

Olav Fyrileiv 30/01/02


20
Horisontal frequencies

0.7

0.6

0.5

0.4 fixed-fixed
f0

0.3

RP-F105 span

0.2

0.1 pinned-pinned

0.0
30 40 50 60 70 80 90 100

Figure 6-4 – Natural frequencies from FE analyses compared to estimates based on simplified
boundary conditions.
As most free span designs and assessments are performed using some sort of simplified, beam theory
expressions, a final code comparison has been performed using the structural response quantities
associated with the different codes. For CN 30.5, the classical pinned-fixed boundary condition is
used.
The maximum allowable span lengths for a design life of 50 years are listed in Table 6-1. It is seen
that the new Recommended Practice allows significantly longer spans than Guideline no 14 due to
the updates discussed previously in this paper.
Table 6-1 Maximum allowable span lengths according to different codes.
Code Max span length
(m)
CN 30.5 54
GL 14 47
RP-F105 no trench, no concrete 60
RP-F105 in trench, concrete stiffening 69

Olav Fyrileiv 30/01/02


21
7 SUMMARY AND CONCLUSIONS

A short introduction to the main updates in this new DNV Recommended Practice F105 for free
spanning pipelines has been given. The following observations and comments apply:
• The updated response model and associated safety factors represent a slight relaxation compared
to Guideline 14.
• The new RP-F105 accounts for the effect of a trench underneath the spanning pipe. This effect is
found to be significant and should be accounted for when free spans develop due to scouring.
• For pipelines with a thick concrete coating, the effect on the estimated fatigue life due to the
stiffening effect of the coating is found to be significant.
• A new set of boundary condition coefficients has been established which provides good estimates
for the structural response of single free spans on a relatively flat seabed.
• A comparison between different codes revealed that the new RP-F105 allows significantly longer
spans than other older codes.
The main conclusion from the case study presented in this paper is that the new RP-F105 is
recommended to be used for free span assessment as it represents state-of-art in free span design and
will minimise the costs related to seabed correction and span intervention work.

REFERENCES

Blevins, R.D., “Flow-Induced Vibrations”, Krieger Publishing Company, Florida, 1994.

DNV, “Rules for Submarine Pipeline Systems, Det Norske Veritas”, 1981.

DNV-OS-F101, “Submarine Pipeline Systems, Det Norske Veritas”, 2000.

DNV Guidelines No. 14, “Free Spanning Pipelines”, 1998.

DNV RP F105, “Free Spanning Pipelines”, Draft for industry hearing, October 2001 (to be issued
early 2002).

DNV CN 30.5, “Environmental Conditions and Environmental Loads”, 1991 (updated in 2000).

Hansen, E.A., Bryndum, M., Mørk, K., Verley, R., Sortland, L. and Nes, H., “Vibrations of a Free
Spanning Pipeline Located in the Vicinity of a Trench”, OMAE’2001.

Hetenyi, M., “Beams on Elastic Foundation”, University of Michigan Press, Ann Arbor, 1946.

Hobbs, R.E., “Influence of Structural Boundary Conditions on Pipeline Free Span Dynamics”,
OMAE’86, 1986.

Olav Fyrileiv 30/01/02


22
Mørk, K.J., Fyrileiv, O., Verley, R., Bryndum, M., Bruschi, R. “Introduction to the DNV Guideline
for Free Spanning Pipelines”, OMAE’98, Lisboa, July 6-9, 1998.

Ness, O. and Verley, R.L.P, “Strain Concentrations in Pipelines with Concrete Coating: An
Analytical Model”, Proc. of the 14th Int. OMAE Conf., 1996.

Sumer B.M. & Fredsøe, J. “Hydrodynamics around Cylindrical Structures”, Advanced Series on
Ocean Engineering – Volume 12, World Scientific, London, 1997.

Verley, R.L.P and Ness, O., “Strain Concentrations in Pipelines with Concrete Coating: Full Scale
Bending Tests and Analytical Calculations”, Proc. of the 13th Int. OMAE Conf, 1995.

- o0o -

Olav Fyrileiv 30/01/02


23
View publication stats

You might also like