You are on page 1of 20

Frontiers in Offshore Geotechnics: ISFOG 2005 Gourvenec & Cassidy (eds)

2005 Taylor & Francis Group, London, ISBN 0 415 39063 X

Pipeline geotechnics state-of-the-art


D.N. Cathie & C. Jaeck
Cathie Associates SA/NV, Brussels, Belgium

J.-C. Ballard & J.-F. Wintgens


Fugro Engineers SA/NV, Brussels, Belgium

ABSTRACT: Pipeline geotechnics deals with soil-pipeline interaction. This covers installation issues (pipeline
penetration and short-term lateral stability), axial and lateral response to loads. It then encompasses pipeline
trenching, backfill engineering and pipeline stability when buried. This review provides an overview of all aspects
of pipeline geotechnics except trenching. The focus of the paper has been on the mechanics of each problem,
explaining the issues with a view to developing understanding, rather than providing ready made solutions. The
interested reader can make use of the references for going deeper into particular aspects of the subject.

geotechnical industry has put together a guidance


document to assist non-specialists to understand the
basic data required from a survey (OSIF 1999). Characterisation of the pipeline route with emphasis on
trenching issues is discussed in Cathie (2001).
Pipeline design and engineering will require geotechnical data if major assumptions are to be avoided.
For design, all soils require classification tests (particle size distribution, index tests and basic shear
strength parameters). The in situ density of granular
soils and the undrained shear strength of cohesive
soils should be determined.
If the soil is very soft it is not sufficient to collect
drop cores as the disturbance can lead to underestimating the soil strength. This could affect axial and lateral
friction assessments, and trenching engineering issues.
The box corer coupled with in situ vane testing (OSIF
1999) is recommended for obtaining good quality data
in very soft soils. A substantial quantity of soil can be
collected using a box corer and this may be necessary if
riser-soil interaction needs to be investigated in specific
tests, or if backfill properties need to be investigated.
In granular material, the cone penetration test (CPT)
remains the most attractive tool to determine the in situ
density (OSIF 1999). CPTs should always be combined with sampling techniques in order to determine
the physical properties of the sand, particularly the
grain size, which may be important for trenchability.

1 INTRODUCTION
Subsea pipelines are laid on the seabed and may or
may not be buried. Pipeline design needs to account
for the ways in which the pipeline interacts with the
soil. This paper is about pipeline-soil interaction and, in
particular, the geotechnical issues that must be faced
by pipeline designers if soil-pipe interaction is to be
captured adequately in the design process. Pipeline
geotechnics is an emerging specialty that involves
applications of geotechnical theory and practice unique
to the construction of underwater pipelines.
For this State-of-the-Art review, the material has
been organized broadly in the order in which the phenomena arise embedment and lateral friction mobilisation during pipelay, stability against current and
wave action when on the seabed, development of axial
friction as a result of thermal and pressure loading etc.
Since embedment during pipeline affects axial and lateral friction this appears to be a logical progression.
Some areas that are not covered include self-burial
of pipelines as a result of wave/current action and
sediment transport, liquefaction around pipelines, and
the whole area of ploughing and trenching (Cathie &
Wintgens 2001) which would make the paper too
extensive.
2 GEOTECHNICAL PARAMETERS

3 PIPELINE EMBEDMENT

If soil-pipeline interaction is to be understood and


modelled then it is clear that geotechnical data along
the pipeline route is required. This basis point is not
always recognized and a group involved in the offshore

Pipeline embedment begins during pipe lay and may


increase with time due to hydrodynamic forces, pipe

95

content load (e.g. hydrotest of the line) and creep.


Embedment is primarily an issue of penetration bearing capacity failure until the soil resistance is sufficient to resist the load applied by the pipeline. The
loads applied by a pipeline during installation will be
discussed first followed by the bearing capacity
issues.

4.0
Load Concentration in Soil (-)

3.1

4.5

Soil loading during pipeline installation

Consider a pipeline being laid in perfectly still conditions. As the pipeline is laid on the seabed the force
applied by the pipe near the touchdown point is much
greater than the submerged weight of the pipe as a
result of the seabed interrupting the normal catenary of the pipe. An analysis of soil reaction loads for
a 12 pipeline in very soft and firm soil conditions is
shown on Figure 1 computed using a finite element
analysis which models the pipe lay process and an
elastic-plastic seabed.
In the example shown the load concentration factors (applied to the pipe submerged weight) are 2.0
and 4.2 for the very soft and firm soil respectively.
The load concentration is also a function of the catenary shape (including the tension in the pipeline) as
well as the soil stiffness.
During real pipelaying operations, vessel heave
occurs which induces an additional vertical motion at
the vessel end of the pipe and cyclic touchdown and
lift off cycles at the seabed end. This vessel motion
will inevitably induce some changes in pipeline tension on the seabed and therefore additional soil pressures. Changes in tension are also introduced as new
sections of pipe are welded and then over-boarded in
a stepwise manner. This change in tension causes sections of pipe near the seabed to touchdown and lift off
more than once.
Wave and current action on the pipeline in the
water column will provide another source of dynamic
loading and this could be in any direction causing the
pipeline to move horizontally as well as vertically as
it touches down. Pipeline motion near the seabed
could also produce a scouring effect from below the
line in granular soils.
The purpose of describing pipeline behaviour during laying is to provide the right framework for understanding pipeline embedment. Any analysis that is
performed should consider (a) the static pipeline weight,
(b) the amplification of that load at touchdown, and
(c) the likely magnitude of cyclic loading (both vertical
and lateral) that will accompany the laying process. All
of these processes will result in the pipe being embedded deeper than a simplified analysis based on the
weight alone would suggest. This is extremely important because lateral resistance that is needed during
pipelaying to negotiate a turn, or that is needed to resist
hydrodynamic forces during operations (on-bottom

Very soft
Firm

3.5
3.0
2.5
2.0
1.5
1.0
0.5
0.0

10
20
30
40
Distance from Touch Down Point (m)

50

Figure 1. Example of load concentration at touchdown point.

stability), is critically dependent on embedment, particularly for soft clay soils.


3.2

Pipe penetration analysis

Pipe embedment is fundamentally a large deformation penetration problem rather than a bearing capacity problem. As the pipeline touches down, plastic
deformation of the soil occurs until the penetration is
sufficient (i.e. the bearing area is large enough) to
provide the resistance necessary. If the soil loading
during touchdown is greater than the pipeline weight
then the soil experiences unloading as the pipe lay
process moves on. Therefore, most pipelines are
believed to be overpenetrated (terminology of
Zhang et al. 1999), i.e. they have experienced a vertical load greater than they are experiencing at present,
a state that is not unlike overconsolidation in some
respects. Of course the degree of overpenetration may
vary, for example if the pipe is filled with water for
hydrotesting, or with oil for production.
Various approaches have been proposed to analyse
the static penetration problem (Fig. 2a). Early attempts
were based on considering the pipe as a strip footing
where the width of the footing is taken as the chord
length of the embedded section of the pipe (Small et al.
1971). Recent work on riser-soil interaction has continued to use this method (Bridge et al. 2004). Verley
& Lund (1995) proposed an empirical equation for
penetration in clay, based on an extensive laboratory
testing programme, expressed as

(1)

96

Wagner et al. (1987) and is applicable to remoulded


clay with an undrained shear strength around 1 kPa.
Some of the scatter may arise from the difficulty of
measuring strength and its variation with depth.
The plasticity solutions do not account for buoyancy effects, soil heave, or any increase in strength
with depth. Murff et al. have shown the potential impact
of heave and increase in strength. The effect of buoyancy would typically be about 510% depending on
the soil strength but would be more important in fluid
muds. Soil heave is potentially even more important.
Soil heave around the pipe increases the bearing area
by about 20% at z/D  0.2 and increases the contact
area at the same depth by 35%. At a penetration of
about z/D  0.25 almost the full diameter of the pipe
is bearing on the soil when heave is accounted for. It
is difficult to see how laboratory and field data can be
properly interpreted without taking account of the
geometric changes that occur with penetration. There
is scope for further theoretical work in this respect.
Since pipe penetration involves remoulding the
soil locally to the pipe wall, and since repeated loading due to hydrodynamic effects would only accentuate this effect, it seems reasonable to assume that
penetration assessment should consider a zone of
remoulded soil below the pipeline during laying. This
would suggest that using a low soil-pipe adhesion for
embedment assessment would be appropriate.
In the view of the authors, the issue of initial penetration resistance is relatively well understood and
established. Focus should now be on the nature and
magnitude of the loading applied by a pipeline during
laying, and the possible magnitude of the cyclic effects.

heave

z
a)

b)

Figure 2. Pipeline embedment (a) initial penetration, (b)


lateral movement.
6

Normalised resistance (F/SuD)

0
0.0

0.2
0.4
0.6
Normalised penetration (z/r)

0.8

Murff et al - rough

Murff et al - smooth

Verley linear

Verley and Lund

1.0

3.3

Data (Murff et al)

Penetration due to repeated loads

As discussed above, as a pipe is laid on the seabed the


movement of the pipe during the laydown process
will increase the penetration.
Cyclic vertical loading has been investigated by
Dunlap et al. (1990), Fontaine et al. (2004) and in the
STRIDE and CARISIMA joint industry projects
aimed at understanding riser-soil interaction (Bridge
et al. 2004). The latter authors provide a clear explanation of repeated penetration and pullout, particularly
focusing on the suction that develops during pullout
in soft cohesive soils. Dunlap et al. indicate that
limited cyclic loading without break out resulted in
little additional burial while larger cyclic loads which
broke suction and pulled the pipe free resulted in further penetration. This may be due to the additional
remoulding experienced in the full breakout case.
Repeated lateral loading induced by current loading and dynamic response of the suspended section of
pipeline can result in a considerable increase in penetration (Morris et al. 1988), particularly if the pipe is

Figure 3. Lower bound plasticity solutions and empirical


approaches for pipeline embedment for cohesive soil.

where z  penetration, D  pipeline diameter,


su  undrained shear strength,   soil unit weight,
Fz  vertical load per unit length on soil.
Verley & Sotberg (1992) propose equivalent equations for penetration in sand but for cyclic loading.
Murff et al. (1989) have presented upper and lower
bound plasticity solutions for rough and smooth pipes
(full adhesion and no adhesion) in cohesive soils. The
lower bound results are shown on Figure 3 and compared with experimental data and the Verley & Lund
approaches. Upper bound solutions were only slightly
higher.
These solutions are of most importance in very
soft clays where embedment may be significant. Much
of the experimental data shown in Figure 3 is from

97

properties also take on more importance for lateral


and axial resistance.

not overpenetrated. Consider a pipe penetrated with a


certain vertical load and let us assume that the load is
kept constant. Any horizontal load will then result in
both vertical and horizontal movements since the pipe
is on the failure envelope. This is discussed further in
the next section.
Morris et al. (1988) carried out laboratory tests to
assess the embedment due to horizontal cyclic loading in very soft clay. The additional penetration was
largely dependant on the magnitude of the force, or
displacement, and the duration during for which it
was applied. Although the rate of penetration was
found to decrease over the duration of each test, this
rate did not appear to reach a constant value or the
pipeline to reach a limiting burial depth. As the pipe
continued to sink, the mounds grew in size, attaining
heights of up to 0.8 D. Morris et al. also considered
the accumulated burial effect of variable cyclic loading and found that the effect could be modelled with
sets of uniform cycles.
Verley & Lund (1995) proposed the empirical
equation

4.1

Lateral resistance

Three different approaches have evolved for assessing lateral resistance:


1 a single friction factor approach where the lateral resistance is related to the submerged weight
of the pipeline and the soil type;
2 a two component model consisting of a sliding
resistance component and a lateral passive pressure
component frictional model supplemented with
passive resistance of the wedge of soil (Nyman
1984, Wagner et al. 1987, Lieng et al. 1988, Verley &
Sotberg 1992, Verley & Lund 1995);
3 a plasticity model approach (Zhang et al. 1999,
2002, Cassidy 2004).
Generally, the two component models are based on
empirically fitting laboratory test data. A summary of
some of the proposed equations is given in Table 1.
While being practically useful, these models do not
enlighten the user with the actual mechanics of the problem. Moreover, they become more and more empirical
when cyclic behaviour needs to be introduced.
The plasticity framework outlined by Zhang et al.
(1999, 2002) provides a much more fundamental way
of understanding the mechanisms involved as well as
being much more general. Zhang et al. have developed
the plasticity model for calcareous sands but the
application to clays should be straightforward and is
in progress by the authors. The concepts are now well
developed for surface footings. As described by
Cassidy (2004) the models contain a yield surface, a
strain-hardening expression, elastic behaviour inside
the yield surface, and a flow rule to define the direction of movement during yield.
Considering pipe behaviour under combined vertical (V) and horizontal (H) loads in calcareous sand in
centrifuge tests, Zhang et al. have confirmed what
was known for surface footings that the shape of the
yield surface is almost parabolic. A simplified representation was proposed, given by

(2)

where a is the amplitude of horizontal movement, to


assess the maximum penetration that can be achieved
for a given amplitude of motion and to assess the
development of the penetration as a function of the
work done by the pipe on the soil. The authors suggest
that the applicability should be limited to z/D of 0.3
due to range investigated in their study.
Lund (2000) performed an investigation on the penetration depth achieved for a large diameter pipeline
(1.2 m OD). While the calculated embedment was
about 0.05 m, the actual embedment varied between
0.250.4 m. Much of the route was sand. Lund concluded that much of the additional embedment was
due to lateral oscillation at the touchdown point.
4 LATERAL AND AXIAL RESISTANCE OF
PARTIALLY EMBEDDED PIPELINES

(3)

Lateral and axial resistance of an unburied pipeline


needs to be assessed at the design stage. Lateral resistance is important for the design of the pipeline onbottom weight and any weight coating that may be
required. Axial resistance controls pipeline expansion
and affects end connections, spool pieces, and upheaval
buckling. Since environmental conditions (wave/current) leading to lateral forces and thermal/pressure
effects leading to axial forces are variable and cyclic in
nature, the subject must be considered for both monotonic and cyclic loading. The pipeline outer coating

where  is a parameter associated with the friction


between pipe and soil for low vertical load, Vmax and
Vmin are the positive and negative intercepts (H  0)
on the vertical load axis, and
(4)
where  is determined from calibration tests. This
expression gives an intercept on the horizontal load

98

Table 1. Examples of lateral resistance models.


Reference

Equations

Details

Wagner et al. (1987)

Sands:
Fy   (W FL)    A

  8.6 kN/m3
Monotonic
  0.6
  38
 9.6 kN/m3
Monotonic
  0.6
  79
Cyclic load
( static failure)
Embedment x 2
 reduced by 50%
Cyclic load
( 5% D)
Embedment x 3
 reduced by 8090%
Monotonic
  0.2
  39.7
Cyclic
( static failure)
Embedment x 2
  31.7
All clays
cyclic load
( static failure)
Embed x 2.5
  15.7
  0.6 (sands)
  0.2 (clays)
FR calculated considering
accumulated energy
All sands
  0.6

Fy  horizontal resistance
A  0.5xEmbedded area
W  submerged pipe wt
FL  hydrodynamic lift

Clays:
Fy   (W FL)   Su A/D

Lieng et al. (1988)

Fy   (W FL)  FR

Verley & Sotberg (1992)

Fy  Fc  FR
Fy   (W FL)  FR
FR  D2 (4.5 0.11D2/Fc) (z/D)1.25
Fy  Fc  FR
Fy  (W FL)  FR

Verley & Lund (1995)

FR  4.13 DSu(Su/(D))0.392 (z/D)1.31

axis when vertical load is zero, and represents a passive soil resistance with a magnitude Vmin. Vmax represents the maximum vertical load for a given
penetration (the preload). Figure 4 shows the normalised form of the yield surface for different values
of . The peak horizontal resistance is achieved at
about 40% of the maximum vertical load.
The proposed hardening function is based on the
monotonic vertical penetration resistance of the pipe
(plastic stiffness) and the rebound response gives the
elastic stiffness. This enables the increment of vertical
plastic strain to be defined in terms of the elastic and
plastic stiffnesses and the increment in Vmax.

Clays
(Su  70 kPa)
  0.2

The plastic potential takes a different but similar form


to the yield surface, and was defined by considering the
displacement increment vectors in different tests:
(5)
where t is the shape parameter. The exponent m has
the effect of adjusting the value of the vertical load at
which the normal to the plastic potential becomes
parallel to the H axis. Figure 5 shows the plastic
potential for different values of m. The model predicts
upwards pipe movement and strain softening when

99

-0.2

Normalised lateral resistance H/ 'z2

=0.7

50

100

150

200

0.1

0.4
Embedment (z/D)

V/Vmax

0.2

=0

0.6

=0.1
0.8

=0.2

1
0.0

0.1

0.2

0.3

0.4

0.5

0.2

0.3

0.6

H/Vmax

0.4

Figure 4. Yield envelope for sand (Zhang et al. 1999).


0.5

-0.2

=0.05

Plasticity model

Verley & Sotberg (1994)

Lieng et al (1988)
Brenodden et al, 1989
Verley and Sotberg (1994)

Palmer et al, 1988


Wagner et al, 1987
Zhang et al (2001)

Figure 6. Lateral resistance models for sands (monotonic


tests).

V/Vmax

0.2
0.4
0.6

m=0.1
m=0.2

0.8

m=0.3
m=0.4

0.1

0.2

0.3

0.4

0.5

0.1

1
0.0

Normalised lateral resistance H/ 'D2


1
2
3

0.6
Embedment (z/D)

H/Vmax

Figure 5. Plastic potential for sand (Zhang et al. 1999).

V/Vmax is between 0 and 0.25 depending on the


exponent m.
For the calcareous sand tested by Zhang et al.
Figure 5 makes it clear that unless the pipe is highly
overpenetrated (i.e. V/Vmax is very low) further penetration of the pipe can be expected under horizontal
loading.
The model has been calibrated for the specific calcareous sand. However, it is of interest to see how the
model compares with other methods for calculating
lateral resistance. Care must be taken to correctly
interpret laboratory tests in terms of whether they are
sideswipe tests (horizontal displacement control at
constant vertical displacement) or probe tests (horizontal displacement at constant vertical load).
Figure 6 shows experimental data for monotonic
tests with an overpenetration ratio of 1 (current vertical load is the maximum experienced). Normalisation
with the penetration depth is logical since vertical
penetration is a key variable but the plot tends to put

0.2

0.3

0.4

0.5
Plasticity model
Lieng et al (1988)
Brenodden et al, 1989
Verley and Sotberg (1994)

Verley & Sotberg (1994)


Palmer et al, 1988
Wagner et al, 1987
Zhang et al (2001)

Figure 7. Lateral resistance model for sands normalised


with pipe diameter.

the data scatter in a good light. Presenting data in


terms of D (Fig. 7) shows the wide range of data.
Both plots demonstrate that the Zhang et al. plasticity
model is a reasonable representation of all the data
despite the fact that it has not been calibrated for
other sands.

100

0.4
0

0.6

0.8

Friction Factor
1
1.2

1.4

1.6

Table 2. Resistance factor for pipeline axial friction coefficient under fully drained conditions (after Finch et al.
2000).

1.8

Embedment (z/D)

0.1

0.2

0.3

0.4

0.5
Plasticity model
Lieng et al (1988)
Brenodden et al, 1989
Verley and Sotberg (1994)

Verley & Sotberg (1994)


Palmer et al, 1988
Wagner et al, 1987
Zhang et al (2001)

Figure 8. Lateral resistance models for sands in terms of


friction factor.

In order to relate the models back to the traditional


friction factor, the horizontal resistance has been normalised with the applied vertical load to give the
friction factor (Fig. 8). The low prediction of the
plasticity model at greater depth reflects the Zhang
data but not the other data.
The plasticity model is likely to require development as the model described above is a single surface
strain hardening model. Other approaches are likely to
be required for modeling cyclic loading. In particular,
geometric changes will need to be considered where
large lateral movements of several diameters occur.
4.2

Axial resistance

Axial loads normally apply some time after installation.


The implications are different for sands and clays. In
sands, wave and current action may have induced
minor lateral loading which has in turn induced some
further embedment or densification of the soil. There
could be some build up of sediment against the pipe.
In clays, set-up following remoulding will have had
time to develop. This could include components of
strength increase arising from thixotropy and consolidation. It is also likely that good adhesion between the
pipe and the soil will have developed in soft clays.
Although the axial resistance is influenced by
embedment and time dependent factors such as those
described above, simple Coulomb friction models are
often adopted to evaluate the axial resistance of partially embedded pipelines in all soils, given by
(6)

Condition

fr

Granular cohesionless soil and


D50  pipe roughness
Granular cohesionless soil and
D50  pipe roughness
Fine-grained cohesive soil and
D50  pipe roughness
Clay and D50  pipe roughness
Silt and D50  pipe roughness

fr  1
0.75  fr  0.9
fr  1
fr  0.6
fr  0.4

Strictly, this is only valid for drained conditions,


but this may be a reasonable assumption for both
sands and clays if the loading rate is slow enough. For
thermal expansion, the temperature increase is likely
to take several hours and this could be taken as justification for a drained analysis.
Axial friction assessment then reduces to evaluating the friction coefficient .  depends on the internal
friction angle of the soil and on the properties of the
soil-pipeline interface. There are various guidelines
used in the offshore industry for both pipelines and
piles:
  tan( 5) (API RP2A WSD 2000)
  2/3 tan() (Bureau Vritas)
  fr tan() (Finch et al. 2000)
The first two formulations assume that the interface is soil-steel. For pipelines this is rarely the case
since the outer coating is generally a corrosion protection such as polypropylene (PP) or a concrete
weight coating. PP coatings may be smooth but special materials can be ribbed to improve friction for
transport and handling. Based on research performed
on a range of coatings (Finch 1999), Finch et al.
(2000) recommend values of fr shown in Table 2 as a
function of coating roughness and soil grain size:
For fine-grained sediments, and where loading
may be rapid enough to elicit an undrained response
from the soil, the axial resistance would be a function
of the contact area, and of the undrained shear
strength of the soil, expressed as
(7)
where   adhesion factor and L  arc length in
embedded soil (including heave).
Appropriate values of the shear strength and adhesion factor will depend on whether the peak or residual
axial resistance is required, and how long the pipeline
has been installed without load. Laboratory shear
tests are recommended for the specific soil and coating under consideration. In very soft clays, in the

101

absence of specific data, and subject to the roughness


of the interface, the adhesion factor may be taken as 1
for the peak resistance and related to the soil sensitivity, St, for the residual strength (  1/St ).
The research programmes discussed by Finch
(1999) have shown that the peak resistance in sands is
typically achieved within an axial displacement of
2 mm. This value is consistent with the 0.1 (2.54 mm)
generally considered for the axial friction mobilization of tz curves of driven piles (API RP2A WSD
2000). A linear elastic-perfectly plastic axial resistance
mobilization curve (with a peak resistance reached at
around 2 mm) is therefore a reasonable assumption in
granular soils.
4.3

5 PROPERTIES OF TRENCH BACKFILLS


The properties of a trench backfill are necessarily a
function of how the backfill is placed. Four broad categories can be considered:
natural infill (wave/current induced);
mechanical backfilling;
backfill following jetting;
active jet cutting and collapse.

5.1

Nature of original material

Characteristics of fill

Fairly clean sand


(15% passing No.200 sieve)
Silty or clayey sand

Reasonably uniform fill


of moderate density
Very heterogeneous fill
of large void ratio
Skeleton of clay balls,
with matrix of sand and
clay
Laminated normally
consolidated or
underconsolidated clay

Stiff cohesive soil


Soft cohesive soil

5.2

Axial creep or walking

Observations and analysis have shown that pipelines


can walk or creep axially (Tornes et al. 2000, Carr et al.
2003) due to internal heating and cooling. The driving
mechanism is the expansion and contraction of the
pipeline and whether there is an effective anchor
point where no movement occurs. The rate of creep
will depend not only on the temperature profiles but
also on the magnitude of the axial resistance, the
mobilization distance and the degradation to residual
conditions.

Table 3. Classification of hydraulic fills (Whitman 1970).

Natural infill

Wave or current induced natural infill is applicable


mainly in sands and in relatively shallow water where
seabed currents are sufficient to induce transport.
Soil particles are deposited under a relatively high
energy environment which results in a structure that is
typically loose to medium dense. Rates of trench
infill can be estimated using methods such as Schapp
(1982) and Niedoroda & Palmer (1986), or by
directly modeling the flow regime accounting for
spoil heaps using computational fluid dynamics simulations. Other relevant information is given in Van
Rijn (1993) and Fredsoe (1978). Note that natural
densification can also occur with time as a result of
wave action (Clukey et al. 1989).

Mechanical backfilling

Mechanical backfilling involves scraping the spoil


(previously removed from the trench) back into the
trench. It is applicable to all types of soil conditions.
The backfilling process is discussed in detail in Cathie
et al. (1998). Soil in the spoil heaps, and sometimes
some of the in situ seabed soil, is mixed and deposited
into the trench very rapidly. Water is believed to be
entrained with the backfill and the resulting mass is
expected to have a higher macro water content than in
the spoil heaps, particularly if the soil contains a cohesive component.
Mechanical backfills are not unlike hydraulic fills
and a starting point for considering the properties of
backfills is to use this work. Whitman (1970) proposed a classification system, which provides a starting point.
There is no specific published data on the properties of mechanically backfilled trenches as far as the
authors are aware.
For sand backfills, some information about in-situ
densities after hydraulic filling and slumping is provided by Stoutjesdijk et al. (1998). They considered
the formation of submarine slopes with hydraulic
sand fill and found that very low densities could
develop during hydraulic filling (1030%) but that
liquefaction and flow could increase the relative density (to 2050%). The rate of filling was apparently not
important (Bezuijen & Mastbergen 1988). It seems
reasonable to assume that similar considerations apply
to mechanical backfilling of sands. Therefore, sand
backfills are expected to be in a loose state after backfilling. It is not difficult to demonstrate that drainage
of a 1.5 m trench should be largely complete in a few
minutes for most sands.
Clay backfills or mixed sand/clay materials are
believed to be heterogeneous after backfilling. Stiff
clay is ploughed out of the trench into the spoil heaps
and then left exposed to free water until backfilling
takes place. The surfaces of the lumps will take in water
and soften. During backfilling, further disturbance

102

and deformation of lumps will occur and the voids


between lumps will be filled with water, slurry and
more mobile sand components. Further details and
discussion can be found in Karthikeyan et al. (2001,
2002), Hartlen & Ingers (1981) and Mendoza &
Hartlen (1985). Lumpy clay backfills consisting of
stiff clay will generally consolidate much more
quickly than would be expected for homogeneous insitu soil due to the voids and channels available for
free water. A model for consolidation of lumpy clay
fills has been proposed by Yang et al. (2002).
Soft clay backfills created by mechanical trenching and backfilling are also believed to be heterogeneous and consist of softened and remoulded material
close to the in-situ water content in a slurry of much
higher water content soil. Their properties are strongly
time dependent as consolidation and thixotropic regain
takes place. Bruton et al. (1998) studied the consolidation of clay slurry in the laboratory and showed that
cavities and channels formed within the mud as it settled. Such channels tend to form even in a homogeneous slurry as it sediments but are further encouraged
if it is heterogeneous, if there is sufficient adhesion in
the clay to permit cracks to remain open, and if there
are silt or sand inclusions. This heterogeneity was
found to result in much more rapid consolidation than
would have been measured in an oedometer test on
homogeneous soil. Mechanical backfilling probably
does not destructure the soil as completely as jet
trenching and therefore the quality of the backfill soil
should also be better. This is an area of active work at
the present time and much is as yet unpublished.
5.3

Backfilling following jetting

Lowering or burial of a pipeline by jetting results in


destruction of the in situ soil, local movement of the
soil behind the trencher and deposition of the soil particles or lumps back into the trench or just outside.
While clays are cut and broken before being transported, sands and silts are eroded and transported in
suspension.
Sand backfills. There is almost no data related to
actual soil densities in a real backfill post-jetting.
However, Kvalstad (1999) does indicate that low cone
resistances have been measured even some time after
trenching indicating the sand to be loose to very
loose. These loose sands had apparently been stable for
a long period. We have also seen very loose sand offshore that have cone resistances equivalent to soft clays.
As discussed for mechanical backfilling, the
mechanisms described in Bezuijen and Masterbergen
(1988) for example for hydraulic filling are likely to
be relevant. However, the mechanism of sedimentation of sand in the trench following trenching is quite
rapid but may occur in a low energy environment.
Rapid sedimentation of sand after trenching is quite

similar to the method of water pluviation to achieve a


minimum density soil (Kolbuszewski 1948) and it is
known that slower rates of deposition/sedimentation
lead to higher densities. Resedimentation tests by
Kvalstad on fine sand have shown that relative densities in the range 1020% have developed. The density of the backfill is likely to be affected by:
rate of trenching and sedimentation, use of backflow jets (rate of deposition and water flow through
the sand affects final density (Kolbuszewski
1948);
particle angularity, grain size distribution (uniformity) and grain size, mineralogy (Youd 1973, Hight
et al. 1999);
depth of backfill (density of very loose sand will
probably increase with stress level due to the high
contractive potential at very loose states);
time, Pipeline movements, wave action, and ageing
will all tend to cause the loose sand to densify
with time.
Some experiences with uplift of pipelines in very
loose backfills appears to confirm that the sand can
be very loose and susceptible to structural collapse
and static liquefaction. Laboratory testing on soils at
very low densities requires special techniques (Lade
1992). Logically, samples prepared by water pluviation would appear to have the best potential for representing the fabric of a post-jetted sand.
Clays. A jetted backfill consisting of clay is
believed to consist of lumps of semi-intact material in
a matrix of unconsolidated slurry. While not unlike
the mechanical backfill of the same material, the
water content is probably higher initially and the
remaining intact soil reduced to smaller more disturbed lumps. Again, very little field data has been
published but Newson et al. (2004) have demonstrated by in situ CPT and T-bar tests in a high plasticity clay in the Nile delta that about 50% of the in
situ strength was regained within 3 months and little
additional strength gain occurred over the next
5 months. This appears to confirm the assumption of
heterogeneous conditions postulated for mechanical
backfilling which allows drainage and consolidation
to occur relatively quickly. Unpublished laboratory
testing on soft blocky clays does confirm the same
hypothesis but no testing that really simulates the jet
cutting and deposition process has been attempted to
date. One important conclusion however from this
testing is that shear strength of blocky soil for uplift
resistance may be characterised by a frictional model
as used for sands (Bruton et al. 1998, Bolton &
Barefoot 1997) although this should be confirmed by
laboratory testing on project specific soils.
Finally, the strength regain of a soft clay backfill is
made up of both reconsolidation and thixotropic
regain. Laboratory testing is beginning to be used to

103

characterise both thixotropic regain (without change


in water content) and consolidation strength regain.
Reference is made to Burland (1990), Leroueil &
Hight (2003), Leroueil (2003), Locat et al. (2003),
Schmertmann (1991), Sills (1995), Silva (1974),
Skempton & Northey (1952) and Skempton (1970) for
further details related to the behaviour and strength of
soft clay backfills.

6 PIPELINE STABILITY DURING


TRENCHING AND BACKFILLING
6.1

Ploughing

Pipeline stability is generally not an issue during


trenching by ploughing since the pipe is lowered into
the cut trench as the plough moves forward. However,
backfill ploughing can lead to pipeline uplift in certain conditions (Cathie et al. 1996, 1998). A pipeline
lying in a V-shaped trench may be susceptible to uplift
caused by:
transverse flow of the soil down the slopes and high
transient hydraulic pressure below the pipe;
turbulence and in-line water flow driven by the
backfill plough;
pipe spans and out-of-straightness;
low pipe weight;
soil liquefaction and slow drainage.
Pipeline uplift is known to have occurred in
pipelines with a specific gravity of between 1.2 and 1.6.
As suggested by Cathie et al. (1998), there are usually
several factors that combine to cause uplift during
backfilling. A backfill plough creates considerable turbulence and the effect of the plough (particularly the
mould boards) and the soil mass advancing rapidly
can cause uplift of a lightweight pipeline in a trench.
Flow of soil down the slopes of the trench is an
important cause of uplift since high transient uplift
forces are developed when the soil impacts the
pipeline (Powell et al. 2002). Both transient hydraulic
pressures and possibly wedging effects if the soil has
a shearing resistance would act to destabilize the pipe.
The kinetic energy of the soil flow is likely to be
greater if the trench is deep and the backfilling is
rapid. High backfilling speeds, even if short lived,
may create conditions in with the uplift initiates. After
initiation of an uplift feature, the gap between the
trench base and the pipe is ahead of the plough. This
in turn permits soil to flow under the pipe and propagate the feature in a progressive manner.
According to Powell et al. (2002), and based on
both model testing and experience, uplift does not
occur if the pipe weight is sufficient. They suggest
that the minimum specific gravity for a mechaniccally backfilled pipeline should be 1.8. Pipelines of

lower density may need to be held in position, for


example, by stitch rockdumping.
6.2

Trenching

Pipeline lowering and burial by jet trenching requires


cutting, erosion and fluidization of the soil by the jets
and lowering of the pipeline into the area cut or fluidized by the jets. Clays are cut and broken by the jets
and transported behind the trencher. Sometimes eductors are used to clean out the trench since the clay
walls remain stable for some time. In the view of the
author, the problem of pipeline lowering and stability
is mainly dependent on the amount of slurry or lumps
left in the base of the trench. However, Powell et al.
(2002) suggest that the product specific gravity
should be greater than the specific gravity of a liquefied clay at the onset of effective stress. This may not
be a useful criterion since it is typically around 1.2 for
most muds.
For sands, the trench walls are not stable and the
pipe lowering is dependent on the pipeline being
heavier than the fluidized soil. The sand minimum
density could be used as a guide to the minimum pipe
weight but the author is unaware of work specifically
aimed at defining what this minimum pipe weight
would be. Nevertheless, it is clear that heavier, more
flexible, pipes will be easier to lower than light stiff
pipes. Flexure of the pipe below the trencher coupled
with maximizing the length of fluidised soil will
result in the most efficient lowering process.
To minimize the risk of floatation in sands (or to
maximize the lowering efficiency), Powell et al.
(2002) suggest that for rapid draining soil a specific
gravity of 1.5 is sufficient but that if silt or clay are
also present (lowering the permeability) then the risk
of floatation is significant if the product specific
gravity is less than 1.7.
The subject of pipeline stability during trenching
and backfilling is still wide open for further insights
and research. We caution against focusing too much
on the 2 D aspects of the problem (cross section) and
only on soil mechanics. Uplift initiation and propagation during trenching/backfilling is a 3 D problem and
involves pipeline structural response to bending and
hydrodynamic loads, soil transport and deposition,
and consolidation. These aspects should be investigated together.
7 AXIAL AND TRANSVERSE RESPONSE
OF BURIED PIPELINES
Axial and transverse response of a buried pipeline is
important for assessing the behaviour of the pipeline
to hydrotest or operational load conditions. Axial
loads induced by internal pressures or temperatures

104

render a pipeline susceptible to transverse loads as a


result of its out-of-straightness i.e. imperfections.
Generally the most serious imperfections are in the
vertical plane and make the pipeline susceptible to
upheaval buckling. However, horizontal or inclined
imperfections can also occur due to laying around a
bend or to the pipeline lying on one side of a sloping
trench.
Therefore, the pipeline response to both axial and
lateral loads is important for assessing buckling
potential. An early case history presented by Nielsen
et al. (1990) focused industry attention on the problem and Pedersen & Jensen (1988) and Nielsen et al.
(1988) suggested solution methods. Since then a lot
of experimental work has been performed to characterize uplift resistance in various soils. We provide a
summary in Section 7.2.
Transverse forces in a pipeline are a result of the
internal compressive forces which are unable to be
released due to the axial restraint of the soil surrounding the pipe. A model for the axial response is
therefore necessary and this is discussed first.
A buried pipeline may be partially in contact with
relatively undisturbed soil as well as being surrounded by backfill. Axial response is reasonably
assessed assuming the pipeline is surrounded by
backfill but considering the effect of some contact
with firmer/denser soil as part of a parametric analysis. Lateral response of a buried pipeline should consider in which direction the lateral movement is
taking places and select soil properties accordingly.
For both axial and lateral resistance assessment,
consideration should be given to whether the backfill
will be in a drained or undrained condition.
7.1

Axial behaviour

7.1.1 Ultimate axial resistance


The ultimate axial resistance of a buried pipeline in
freely draining soil may be determined by considering the mean normal (lateral) pressure,  on the
pipeline and the axial friction coefficient, fr (see
Section 4).
Considering the normal stresses on the top, bottom
and sides of an equivalent square leads to (Schaminee
et al. 1990, Finch et al. 2000):
(8)

mon practice to assume that operational loads


develop relatively slowly (e.g. over a period of hours
for temperature increases) and that both sands and
clays can be treated as drained for axial loading
(Finch et al. 2000). However, this may not necessarily
be the case and it is prudent to consider both drained
and undrained response in the design analyses unless
it can be demonstrated otherwise.
For undrained conditions, the undrained shear
strength is the basic parameter, with the resistance
given by
(10)
Finch (1999) suggests that for clays with low shear
strength, values of  should be 1.0 for peak resistance
and about 1/St where St is the sensitivity of the soil for
a residual strength.
Using axial resistance based on shear strength
implies a uniform strength along the length in areas of
uniform backfill. This is not the reality. Seabed features and out-of-straightness have the effect of increasing the apparent frictional resistance due to high
pressures on supporting areas. Finch recommends to
consider  as 1.0 but use of equation 10 can result in
values of  greater than 1.
7.1.2 Axial response
The response of buried pipelines to axial loads is generally assumed to be similar to partially buried
pipelines as discussed in Section 4. It is difficult to
escape the analogy with axial loading of piles and
similar approaches could be considered (e.g. Kraft et al.
1981, Randolph & Wroth 1978). Differences arise
because of the infinite length of the pipeline and
the non-axisymmetric nature of the problem. Nevertheless, it can be reliably speculated that the soil shear
modulus would fundamentally govern the initial axial
response. Since the modulus is likely to be related to
the stress level, the axial stiffness of the material
around the pipe would vary, particularly for shallow
burial. In practice, the traditional elastic-perfectly
plastic model is generally used in finite element
analyses of buried pipes using a mobilization distance
of 23 mm to mobilize ultimate resistance, values that
are justified by some experimental work.
7.2

Transverse behaviour

Since the most critical form of transverse response is


uplift, this is addressed first. A brief treatment of
transverse resistance in other directions is given later.

so that the axial resistance of a buried pipe is:


(9)
For cohesive soils, a decision has to be made
regarding drained or undrained behaviour. It is com-

7.2.1 Uplift resistance in uniform soil


Uplift resistance models were developed initially for
the pull-out capacity of anchors. Rowe & Davis (1982)
made theoretical and experimental investigations and

105

z
W

P
Figure 10. Circulation mechanism in very loose sand
(Vanden Berghe 2005).

Figure 9. Uplift wedge failure mechanism.


Table 4. Uplift factors (Schamine et al).

provide a good summary of earlier work covering


sands and clays. Subsequently, Murray & Geddes
(1987), Dickin (1988) and others developed the understanding mostly concentrating on uplift of anchor
plates in sand. Recently, Merifield et al. (2001) have
provided a rigorous theoretical solution for the
undrained anchor problem.
Work on buried pipes includes Matyas & Davis
(1983), Trautmann et al. (1985) and others. As the
offshore industry recognized the upheaval buckling
risk of buried pipelines, Schamine et al. (1990) published a fundamental set of experimental data applicable to subsea pipelines. Further experimental data
has been collected (Ng & Springman 1994, White et al.
2001) but much is proprietary and only partially published (Finch 1999, Finch et al. 2000, Fisher et al.
2002).
Uplift resistance models for sand have been developed from the work on anchors in which clearly
defined slip planes are apparent up to relatively large
depths and the focus was on peak uplift resistance. As
shown by White et al. (op cit), a peak resistance may
be accompanied by an upward sliding block mechanism (Fig. 9) but as a gap is formed below the pipe the
boundary conditions of the problem change and the
mechanism changes to a circulation or flow around
mechanism. Pedersen & Michelsen (1988) sketched
the same concept but without elaboration and
Schamine et al. (1990) identified it from their testing
programme.
In very loose sands which contract when sheared,
Vanden Berghe et al. (2005) have demonstrated this
circulation mechanism numerically (Fig. 10).
Compression of the very loose sand and upward
movement makes room for the flow mechanism to
develop. The dilatancy of the soil determines at which
H/D ratio the flow mechanism governs the peak
resistance as well as the residual.
The different mechanisms of failure go some way
to explain why the wedge failure model cannot capture
the whole range of soil density and H/D ratios very
effectively. It may also explain some of the scatter in

Soil

Uplift factor, f []

Very loose sand


Loose sand
Gravel/rockfill

0.15
0.4
0.6

test results (Schamine et al.) and particularly for the


loose soils.
Wedge failure models can broadly be classified into
a) simple vertical slip model (e.g. Schamine et al. see
Fig. 9 with   0) and b) inclined wedge models (e.g.
White et al).
Consider a vertical block mechanism (Fig. 9 with
  0) and assume only the soil above the crown of
the pipe is active (height, H). The uplift resistance, P,
per unit length for frictional (drained) behaviour is
given by:
(11)
where f  K tan  and is known as the uplift factor.
Uplift factors indicated by the Schamine tests are
shown in Table 4 for the limited range of H/D tested
(4 for sands).
A variation of the frictional model (Eqn (11)) is
attributed to Pedersen in which the whole volume of
soil above the pipe is involved:
(12)
where fp  K tan  (using the suffix p to differentiate
with the f of Schamine) and 0.1 (D/H) represents the
weight of the soil wedges between the pipe centreline
and the crown.
The equivalent vertical slip model for cohesive
(undrained) behaviour leads to:

106

(13)

Palmer et al. (1990) recommend more conservative factors (0.5 for dense material and 0.1 for loose)
and put forward a variation of this equation to address
the problem of deep or flow failure, expressed as:

uplift loading is often cyclic since it is associated


with production cycles; upward movement may
enable a gap to develop and particles to flow under
the pipe, leading to upward ratcheting or creep;
very soft backfills are heterogeneous and may
behave like a frictional material.

(14)

These topics are discussed below.


Discontinuous contact of the pipeline with the
trench is normal due to irregularities in the trench
level. This results in gaps (spans) under the pipe
which may not be fully filled with backfill soil.
Fortunately, this is typically associated with sag bends
which, when loaded axially, will tend to move down
into the soil. However, gaps could exist adjacent to a
supported area. A gap below the pipe increases the
tendency for a circulation or flow failure and therefore the uplift resistance could be lower than anticipated for a uniformly bearing pipeline. Design
conservatism must be introduced for incomplete contact with the seabed.
Thermal loading is associated with production and
when production is halted the pipeline cools down.
Thus pipelines experience many cycles of uplift load.
If the backfill resistance is low and the pipe is able to
move sufficiently to create a gap under the pipeline,
soil particles can flow into this gap during a hot
phase. On cooling the pipe cannot return to its original position. Cyclic loading can thus lead to the pipe
creeping or ratcheting upwards. With loss of cover and
resistance a strain-softening resistance is experienced
and full uplift failure can occur. Nielsen et al. (1990)
postulate that this was the mechanism that occurred in
their project. Finch et al. (2000) suggest, based on 1 g
laboratory testing that ratcheting only occurs in clean
sands where the grains are not held by adhesion. In
order to limit progressive uplift the magnitude of
uplift displacements must be controlled.
Very soft clay backfills are heterogeneous and are
believed to drain more rapidly than the in situ soil.
Bolton & Barefoot (1997) have shown that for the
Atlantic mud tested, both drainage is relatively rapid
and the clay dilates when sheared at very low stress
levels. This results in higher than anticipated uplift
resistance. They argue that adopting a frictional
model with an uplift factor f of 0.4 can be justified for
the mud tested. The authors have been involved in
projects involving jet trenching of very soft clays. An
uplift factor of 0.3 was used based on centrifuge tests
on the soil and no uplift problems have been reported.

For the cohesive model, the more rigorous solutions of Merifield et al. (2001) are preferred. Uplift
bearing capacity factors (P/suD) have been computed
using both upper and lower bound solutions accounting for shallow and deep failure mechanisms. Thorne
et al. (2004) has investigated in detail the issue of suction behind the pipe.
Bolton & Barefoot (1997) and White et al. (2001)
have justified the wedge mechanism shown in Figure
9 by demonstrating that the angle  corresponds to
the dilation angle of the soil and accounting for an
increase in the vertical stress (and shear resistance) in
the vicinity of the pipe. The shear stress along the
slip surface is expressed as:

(15)
and thus:

(16)
and K0 can be taken as 1 sin crit
The dilation angle can be assessed from the relative density of the sand, the stress level and the particle characteristics (Bolton 1986). For a range of sands
between loose and dense White et al. show that the
correlation is good. In the experience of the authors,
this approach may overestimate uplift resistance for
very loose sands where a negative dilation angle
would be applicable and where a circulation flow
mechanism occurs.
Recommended uplift factors are also given by
Finch et al. (2000) for different soil conditions.
In practice, some other issues must be addressed
when considering uplift resistance:
pipelines are not always in continuous contact with
the seabed;

7.2.2 Uplift response


Numerical methods used to design against upheaval
buckling require not only the ultimate capacity but
also the displacement response to mobilize this capacity. Results can be sensitive to this stiffness. Much
less interest has been shown on uplift displacements

107

to reach failure but Finch et al. (2000) propose guidelines based on their experimental program. Trautmann
et al. (1985) suggest that the displacement at peak
resistance is between 0.51.5% of the pipeline depth.
For a 16 (0.4 m OD) pipeline buried at a depth of 1 m
(H/D  2.5) the displacement at peak would be
between 26 mm which is in agreement with Finch
et al. at that depth.
7.2.3 Lateral resistance
The lateral resistance of buried pipelines is not generally important for buckling but becomes important
if ground movements occur, such as by faults or
mudslides.
Pipelines buried in sand have been studied by
Audibert & Nyman (1977), Nyman (1984) and
Trautman & ORourke (1985). The ultimate lateral
resistance can be written:
(17)
where the dimensionless lateral bearing capacity factor Ny depends on the relative density of the sand and
on the embedment of the pipeline. Trautman &
ORourke (1985) showed that for loose and medium
dense sands, Ny increases approximately linearly with
the embedment for H/D  8, whereupon Ny becomes
constant, indicative of the transition from shallow to
deep soil failure mechanism. For dense to very dense
sands, the transition was not reached at H/D of 11.
Trautmann & ORourke also demonstrated that the
values of Ny defined for the holding capacity of
anchor plates (Rowe & Davis 1982) were in good
agreement with their own data for pipes. Rowe and
Davis showed that Ny depends primarily on the friction angle and embedment ratio, and on the roughness
of the embedded structure.
For homogeneous cohesive soils, the ultimate lateral resistance of buried pipelines can be based on the
work of Merifield et al. 2001 for plate anchors :
(18)
where the dimensionless factor Nyu depends on the
embedment of the pipeline and to a lesser extent on its
surface roughness.
Considering conservatively the results of the lower
bound plasticity analysis quoted by Merifield et al.
(2001), the dimensionless factor Nyu can be written:
(19)

(20)

1.2
1

Py / Pyu

0.8
0.6
0.4

Audibert & Nyman (1977)

0.2

Trautman & ORourke


(1985)
Simplified bi-linear

0
0

0.2

0.4

0.6
y / yu

0.8

1.2

Figure 11. Lateral force-displacement curves for pipelines


embedded in sand.

The limiting value of 10.47 reflects the transition


from shallow to deep behaviour.
7.2.4 Lateral response
Moving on to the lateral force-displacement models
in sands, as described by Trautman & ORourke, a
hyperbolic relationship is proposed given by
(21)
where P*y  Py/(Ny  H D) is the normalised force and
y*  y/yu is the normalised displacement; a and b are
the model parameters. Proposed values for a and b by
Trautman & ORourke are 0.17 and 0.83, respectively.
Slightly different values are proposed by Audibert &
Nyman (1977). The displacement at the peak resistance yu depends on the embedment ratio (H/D) and
decreases with increasing relative density of the sand.
The hyperbolic force displacement curve can be
simplified into a bilinear representation. Trautman &
ORourke (1985) suggest an initial stiffness equal to
the secant stiffness at 70% of the ultimate resistance.
In that case, the maximum force is reached at a displacement of 0.4yu. Normalised force-displacement
curves are plotted on Figure 11. Note that the soils
modelled in this study have effective friction angles
between 2030 and thus would be in the relative
density range 020%.
7.2.5 Resistance to inclined transverse loads
In a study related to pipelines buried in very loose
sand, Vanden Berghe et al. (2005) have shown that there
is very little difference in uplift resistance when the load
direction is within about 30 of the vertical. Figure 12
shows the displacement patterns and Figure 13 depicts
the uplift factor Nz as a function of direction.
The implication of this finding for upheaval
buckling is that modes of deformation in an inclined

108

Table 5. Typical thermal conductivities.

a) vertical

b) 22.5 to the vertical

d) horizontal

Figure 12. Failure mechanisms for different displacement


directions (Vanden Berghe et al. 2005).
8
H/D=2

Phi=20- Psi= - 10

Nv [-]

Thermal conductivity (W/mK)

Quartz
Water
Clay (typical)

4.09.1
0.600.67
1.52.9

will depend on the thermal conductivity of the pipeline


and its surrounds. If transient solutions are required
(for example for the heating up or shutting down of the
system) the specific heat capacity will also be required.
Hence there is a need to know the thermal properties
of the soil and make use of the low thermal conductivity where possible to provide thermal insulation.
8.1

c) 45 to the vertical

Material

Phi=25 - Psi= - 5

Factors affecting thermal properties

Heat transfer can take place by conduction, convection and radiation. In saturated soils, however, heat
transfer is mainly due to conduction through the solid
framework and the pore water (Farouki 1986). Convection may be more important in coarse grained soils
and rockfill.
Heat conduction in soil can be described, for the
one dimensional case, by the Fourier equation:

Phi=30 - Psi = 0
4

(22)

0
0

22.5

45

67.5

90

Displacement Direction []

Figure 13. Comparison between vertical, oblique and lateral resistance in contractive soils (Vanden Berghe et al.
2005).

direction are very likely to occur since imperfections


are rarely 2-dimensional. This agrees with experience
of surveyed buckles.
8 THERMAL PROPERTIES OF SOILS
Efficient transportation of crude oil in a pipeline
requires a sufficiently high temperature to maintain
low product viscosity and avoid unwanted deposition
of wax. Since the temperature of the soil and water
around the pipeline is lower than the temperature of
the oil, heat flows from the oil to the environment.
Heat lost along the length of the line results in a temperature drop from inlet to outlet. Pipeline design
must ensure that this temperature drop is within
acceptable limits. The temperature loss along the line

where T is the temperature at time t and depth z; k is


the thermal conductivity (amount of heat that flows
through unit cross-sectional area under a unit temperature gradient (
T/
z) in unit time); C is the specific
heat of soil; and is the density. Since the specific
heat of particles of sand and clay is only about 20%
that of water (Geological Society of America 1942), it
is largely on account of the water content that soil is
capable of storing heat.
The main factors that affect the thermal conductivity of saturated soil are: mineral composition, particle
size distribution, density/water content, and temperature. In general, solids conduct heat better than liquids,
and liquids better than gases (Brandon & Mitchell
1989).
Table 5 indicates some typical values for soil components:
Mineral composition. Heat conduction through the
particles is an important mechanism of heat transfer.
All other factors being equal, sands containing a high
percentage of quartz will have a higher thermal conductivity than those containing a high percentage of
mica (Brandon & Mitchell 1989). Clays with a
relatively high content of kaolinite have relatively low
conductivities. Moreover, sands containing high percentages of silica may exhibit increased thermal conductivity with time, possibly due to the formation of

109

silica precipitants at the contact between grains


(Brandon & Mitchell 1989).
Particle size distribution. Well-graded soils conduct heat better than poorly graded soils because the
smaller grains can fit in the interstitial voids between
the larger grains, thus increasing the density and the
mineral-to-mineral contact area (Brandon & Mitchell
1989). Thermal conductivity in general varies with
the grain size of the soil. At a given density and moisture content, the conductivity is relatively high in
coarse grained soils such as gravel or sand, somewhat
lower in sandy loam soils, and lowest in fine grained
soils such as silty loam or clay (Kersten 1949).
Density and water content. Due to the relatively
high conductivity of minerals compared to water, the
conductivity increases with density (Kersten 1949,
Brandon & Mitchell 1989).
Temperature. Thermal conductivity may also be
influenced by temperature because each of the constituents has temperature dependent thermal conductivities (Brandon & Mitchell 1989). All crystalline
minerals in soils show a decrease in thermal conductivity with increasing temperature, except feldspar (clay).
The thermal conductivity of water increases with temperature because of the higher level of molecular movement at higher temperatures (heat transfer by collision).
8.2

Measuring thermal conductivity

Since thermal conductivity is dependent on the specific mineral constituents of the soil as well as its density/water content it is preferable to measure the
conductivity on soil samples in the laboratory or
in-situ. A common method is the thermal probe.
The thermal probe is a long needle that is inserted in
the soil containing both heating and temperature measuring elements. A known amount of current is passed
through the heater element and the resulting variation
of temperature is measured as a function of time. The
thermal conductivity of the soil can be deduced from
these measurements. The applicable procedure is
described by ASTM D5334 (2000). The thermal needle
probe has been presented by various authors (Hooper &
Lepper 1950, DeVries 1952, Woodside 1958, Falvey
1968, Mitchell & Kao 1978, among others).
8.3

Empirical methods for determining


conductivity

In the absence of specific laboratory data, various


empirical equations are available relating the thermal
conductivity of the soil to its water content, dry density
and type of soil. Some empirical approaches are given
in Table 6.
For saturated soils, the relationships presented in
Table 6 can be conveniently expressed in terms of the
water content rather than the dry density or porosity.
The three approaches presented in Table 6 have been

Table 6. Empirical equations for determining soil conductivity.


Reference

Equations (k in W/mK)

Kersten
(1949)

k  0.144  (0.9 log w0.2)  100.06364d


d  dry density (in kg/m3)
for silts and clayey soils
k  0.144  (0.7 log w  0.4)  100.06364d
for sands
k  ksat  ks(1n)kwn
ks  kqqk01q
kw  0.6 W/mK; kq  7.7 W/mK;
k0  2.0 W/mK
q  quartz content
Simplified equation for saturated soils.
k  (a log(w)  b)  10c
a  0.1424 0.000465 p
b  0.0419 0.000313 p
c  0.00062 d
p  weight percentage of soil finer
than 2 m

Johansen
(1975)

Makowski
& Mochlinski
(1956)

used to show variation that may be anticipated


(Fig. 14).
A detailed review is given by Farouki (1986) and
Rawat et al. 1979, Young et al. (2001) have also contributed. Rawat et al. suggested that maximum error
with the Kersten method was 25%. Farouki recommends the Johansen method for coarse to fine sand
(5% passing 2 microns). This method was found to
provide the best correlation because it takes into
account the mineralogy of the sand (which should be
determined by x-ray diffraction). Based on laboratory
data, Rawat considered that the Makowski and
Mochlinskis method overestimates the conductivity
unless the combined silt and clay fraction was used in
the equations.
Laboratory tests from high water content deepwater Gulf of Mexico clays, Indonesia and Nigeria
(remoulded and undisturbed) were reported by Young
et al. (2001). Remoulded data is shown in Figure 15.
This is of interest because it covers water contents
that would be typical for trench backfills.
The undisturbed soil samples had thermal conductivity values ranging from 0.65 to 1.25 W/mK while
remoulded values were in the range 0.8 1.05 W/mK.
8.4

Selection of thermal conductivity

As with all design parameter selection, the use of the


parameter should be considered. Finch et al. (2000)
suggest the following guidelines: upper bound values
should be adopted for thermal insulation design, as
high thermal conductivity represents high heat loss.
Conversely, lower bound values are applicable to
upheaval buckling assessments where heat retained in

110

the pipeline will tend to increase the uplift forces


experienced by a buried pipeline.
The thermal properties of jetted backfills may be an
important design component for a deep water system.
As discussed in connection with mechanical properties, the changes in water content during jet trenching
are not well known. This introduces significant uncertainty when considering the thermal properties of the
backfill. Deep water sites are commonly associated
with soft clays and are ideally suited to the use of jet
trenching. Even though the properties of the backfill
soil are significantly different from the virgin soil, the
backfill soil exhibits low values of thermal conductivities and sufficient strength and density to inhibit thermal convection currents (Young et al. 2001). In fact,
the very soft highly plastic clays encountered at most
deepwater locations have three characteristics that
make them a favourable medium for flowline insulation. First, the clays exhibit cohesion and low permeability making them strongly resistant to thermal
convection (water travelling freely through the soil to
and from the heat source). Second, saturated clays
with high water contents exhibit low values of thermal
conductivity. Third, the soils can be easily jetted to
produce a trench with steep but stable trench walls.

Thermal conductivity [W/mK]

Kersten
4
Sand
3

2
Clay
1

0
0

20

40

60

80

Thermal conductivity k [W/mK]

100

Johansen

50

0
1
0
0

Thermal conductivity [W/mK]

Quartz content, %

20

40

60

80

9 CONCLUSIONS
Makowski

Clay content, %

50

100

1
0
0

20

40
Water content [%]

60

80

Figure 14. Thermal conductivities by various methods for


saturated soils.

Pipeline geotechnics deals with soil-pipeline interaction. This covers installation issues (pipeline penetration and short-term lateral stability), axial and
lateral response to loads. It then encompasses pipeline
trenching, backfill engineering and pipeline stability
when buried. Particularly current topics such as risersoil interaction, upheaval buckling and lateral buckling/ snaking, all of which include many load cycles,
are all subject to ongoing investigations and joint
industry projects.
While performing this review, the authors have
been made aware again of the very wide range of
issues that must be faced in connection with pipeline
design. We have found no similar review papers covering the subject. Therefore, this paper claims uniqueness and we trust it will provide others with a starting
point in many of the specific subject areas. It was
written in the midst of a busy consulting schedule and
therefore does not treat all the subject matter as fully
as we would have liked. There are likely to be some
errors that have crept in. Nevertheless, we trust that
future reviews will find this a useful starting point.
REFERENCES

Figure 15. Remoulded conductivities (Young et al 2001).

ASTM D5334, 2000. Standard test method for determination of thermal conductivity of soil and soft rock by thermal needle probe procedure.

111

Audibert, J.M.E. & Nyman, K.J. 1977. Soil restraint against


horizontal motions of pipes. ASCE Journal of Geotechnical Engineering 103(GT10): 11191142.
Bezuijen, A. & Mastbergen, D.R. 1988. On the construction
of sand fill dams Part 2: soil mechanics aspects. Modelling Soil-Water-Structure Interactions. Kolkman et al.
(Eds), Balkema.
Bolton, M.D. 1986. The strength and dilatancy of sands.
Geotechnique 36(1): 6578.
Bolton, M.D. & Barefoot, A.J. 1997. The variation of critical
pipeline trench back-fill properties. Proceedings of IBC
Conference on Risk-Based and Limit State Design and
Operation of Pipelines. Aberdeen, May.
Brandon, T.L. & Mitchell, J.K. 1989. Factors influencing
thermal resistivity of sands. Journal of Geotechnical
Engineering, Vol.115, No.12: 16831698.
Brennodden, H., Lieng, J.T., Sotberg, T. & Verley, R.L.P.
1989. An energy-based pipe-soil interaction model. Proc.
Offshore Technology Conf., Houston, 14 May 1989 OTC
6057: 147158.
Bridge, C., Laver, K., Clukey, E. & Evans, T. 2004. Steel
catenary riser touchdown point vertical interaction models.
Proc.Offshore Technology Conf. Houston, OTC16628.
Bruton, D.A.S., Bolton, M.D. & Nicolson, C.T. 1998. Poseidon
Project Pipeline design for weak clays. Offshore Pipeline
Technology: 21.
Burland, J.B. 1990. On the compressibility and shear strength
of natural clays. Geotechnique 40, No. 3: 329378.
Carr, M. D., Bruton, D.A.S. & Leslie, D. 2003. Lateral buckling and pipeline walking, a challenge for hot pipelines.
Offshore Pipeline Technology Conf., Amsterdam.
Cathie, D., Machin J. & Overy R.F. 1996. Engineering
appraisal of pipeline floatation during backfilling. Offshore
Technology Conference, OTC8136, Houston: 197205.
Cathie, D., Barras S. & Machin, J. 1998. Backfilling Pipelines:
State of the Art. 21st Offshore Pipeline Technology Conference, Oslo, February.
Cathie, D. 2001. Towards Excellence in Pipeline Trenching
Engineering. Offshore Pipeline Technology Conference.
Cathie, D. & Wintgens, J-F. 2001. Pipeline Trenching Using
Plows: Performance and Geotechnical Hazards. Offshore
Technology Conf., OTC13145.
Cassidy, M.J. 2004. Use of force-resultant models of shallow
foundations in offshore applications. Computational
Mechanics, Sept, Beijing, Springer-Verlag.
Clukey, E.C., Jackson C.R., Vermersch J.A., Koch S.P. &
Lamb, W.C. 1989. Natural densification by wave action
of sand surrounding a buried offshore pipeline. 21st
Annual Offshore Technology Conference, OTC6151:
291300.
Dickin, E.A. 1988. Stress-displacement of buried plates and
pipes. Centrifuge 88, Cort (ed.). Rotterdam: Balkema.
De Groot, M.B., Heezen, F.T., Mastbergen, H., & Stefess
1988. Slopes and densities of hydraulically placed sands.
Hydraulic Fill Structures, ASCE: 3251.
DeVries, D.A. 1952. A nonstationary method for determining
thermal conductivity of soil in situ. Soil Sci. Soc. Am. J.
73(2): 8389.
Dunlap, W.A., Bhojanala, R.P. & Morris, D.V. 1990. Burial
of vertically loaded offshore pipelines. Offshore Technology Conference, OTC6375: 263270.
Falvey, D.M. 1968. Increase accuracy of soil measurements.
Electrical World, November.

Farouki, O.T. 1986. Thermal Properties of Soils. Series


on Rock and Soil Mechanic, Vol. 11, Trans tech
Publications: 136.
Finch, M 1999. Upheaval buckling and floatation of rigid
pipelines: The influence of recent geotechnical research
on the current state of the art. Offshore Technology
Conference, OTC10713.
Finch, M., Fisher, R., Palmer, A. & Baumgard, A. 2000. An
integrated approach to pipeline burial in the 21st century.
Proc. Deep Offshore Technology 2000.
Fisher, R., Powell, T., & Palmer, A.C. 2002. Full scale modeling of subsea pipeline uplift. Physical modeling in
geotechnics, Newfoundland.
Fontaine, E., Nauroy, J.F., Foray, P., Roux, A. & Gueveneux, H.
2004. Pipe-soil interaction in soft kaolinite: vertical stiffness and damping. Int. Offshore and Polar Eng. Conf.,
Toulon, France :517524.
Fredsoe, J 1978. Sedimentation of river navigation channels.
J. Hydr. Div. ASCE, Vol.104: 223236.
Geological Society of America 1942. Handbook of physical
constants.
Hartlen, J and Ingers, C. 1981. Land Reclamation using Fine
Grained Dredged Material. Proc. of the Tenth International
Conference on Soil Mechanics and Foundation Engineering, Stockholm, Vol. 1: 145148.
Hooper, F. C., & Lepper, F. R. 1950. Transient heat flow
apparatus for the determination of heating piping and air
conditioning. Trans., American Society of Heating and
Ventilation Engineers 56: 309324.
Hight, D.W., Georgiannou, V.N., Martin, P.L. &
Mundegar, A.K. 1999. Flow slides in micaceous sand.
Proc. Int. Symp. on Problematic Soils, IS-Tohoku 98,
Sendai, Japan, Balkema, Vol. 2: 945957.
Johansen, O. 1975. Thermal Conductivity of Soils. PhD
Thesis, Trondheim, Norway. (CRREL Draft Translation
637, 1977). ADA 044002.
Karthikeyan, M., Dasari, G.R. & Tan, T.S. 2001. Characterization of a reclaimed land site in Singapore, Soft Soil
Engineering, Eds Lee et al.: 587592.
Karthikeyan, M., Dasari, G.R. & Tan, T.S. 2002. In situ characterization of a land reclaimed using big clay lumps,
Submitted to Canadian Geotechnical Journal.
Kersten 1949. Thermal properties of soils. Engineering
Experiment Station Bulletin 28, University of Minnesota,
Minneapolis
Kolbuszewski, J.J. 1948. An experimental study of the maximum and minimum porosities of sands. Proc. 2nd Int.
Conf. Soil Mech. and Fdn Engng, Vol. 1:158165
Kraft, L.M., Ray R.P. and Kagawa T. 1981. Theoretical TZ
curves. J. Geotech.Eng.Div., ASCE, Vol.107, GT11:
15431561.
Kvalstad, T.J. 1999. Soil resistance against pipelines in
jetted trenches. Proceedings of the Twelfth European
Conference on Soil Mechanics and Geotechnical
Engineering. Amsterdam, The Netherlands, Vol. 2:
891898
Lade, P.V. 1992. Static instability and liquefaction of loose
fine sandy slopes. J. Geotech. Eng., ASCE, Vol.118,
No.1: 5171.
Leroueil, S. & Hight, D.W. 2003. Behaviour and properties
of natural soils and soft clays. Characterisation and
Engineering Properties of Natural Soils, Eds. Tan et al.
Swets & Zeitlinger, Lisse: 29254.

112

Leroueil, S. 2003. Linking field and laboratory soil behaviour. Int. Workshop on Geotechnics of Soft Soils-Theory
and Practice. Vermeer, Schweiger, Karstunen & Cudny:
7378.
Lieng, J.T. & Sotberg, T. & Brennodden, H. 1988. Energybased pipe-soil interaction models. SINTEF Report to the
American Gas Association.
Locat, J., Tanaka, H., Tan, T.S., Dasari, G.R. & Lee, H.
(2003). Natural soils: geotechnical behavior and geological knowledge. Characterisation and Engineering
Properties of Natural Soils, Eds. Tan et al. Swets &
Zeitlinger, Lisse: 328.
Loch, K. 2000. Flowline burial: an economic alternative to
pipe-in-pipe. Proc. Offshore Technology Conf., Houston,
14 May 2000 OTC 12034.
Lund, K.M. 2000. Effect of increase in pipe penetration
from installation, 19th International Conference On
Offshore Mechanics & Arctic Engineering, OMAE2000,
New Orleans.
Makowski, M. W. & Mochlinski, K. 1956. An Evaluation of
Two Rapid Methods of Assessing the Thermal Resistivity
of Soil. Proc. Inst. Elect. Engineers Vol. 103, Part A, No.
11: 453469.
Masterbergen, D.R., Winterwerp, J.C. & Bezuijen, A. 1988.
On the construction of sand fill dams Part 1:Hydraulic
aspects, Modelling soil-water interaction, Eds Kolkman
et al: 353362.
Matyas, E.L. & Davis, J.B. 1983. Prediction of vertical earth
loads on rigid pipes. Journal of Geotechnical Engineering, Vol. 109, No. 2: 190201.
Mendoza, M.J. & Hartlen, J. 1985. Compressibility of
clayey soil used in land reclamation, Proceedings of the
Eleventh International Conference on Soil Mechanics
and Foundation Engineering, San Francisco, Vol. 2:
583586.
Merifield, R.S., Sloan, S.W. & Yu, H.S. 2001. Stability of
plate anchors in undrained clay. Geotechnique 51(2):
141153.
Mitchell, J.K. & Kao, T.C. 1978. Measurement of soil thermal resistivity. ASCE Journal of Geotechnical Engineering
104(5): 13071320.
Morris, D.V., Webb, R.E. & Dunlap, W.A. 1988. Self-burial
of laterally loaded offshore pipelines in weak sediments.
Proc. Offshore Technology Conf., Houston, 25 May
1988 OTC 5855: 421428.
Murff, J.D., Wagner, MX.A. & Randolph, M.K. 1989. Pipe
penetration in cohesive soil. Geotechnique 39(2): 213229.
Murray, E.J. & Geddes, J.D. 1987. Uplift of anchor plates in
sand. Journal of Geotechnical Engineering, Vol. 113, No.
3: 202215.
Ng, C.W.W. & Springman, S.M. 1994. Uplift resistance of
buried pipelines in granular materials. Proceedings of
Centrifuge 94, Vol. 1: 753758. Rotterdam: Balkema.
Newson, T.A. & Bransby, M.F. 2004. Determination of
undrained shear strength parameters for buried pipeline
stability in deltaic soft clays. Int. Offshore and Polar Eng.
Conf., Toulon, France :3843.
Nielsen, N-J.R, Lyngberg, B. & Petersen P.T. 1990.
Upheaval buckling failures of insulated buried pipelines:
a case story, 22nd Annual Offshore Technology Conference,
OTC6488: 581591.
Nielsen, N.J.R, Petersen, P.T., Grundy A.K. & Lyngberg, B.
1988. New design criteria for upheaval creep of buried

sub-sea pipelines. Proc. 7th Int. Conf. on Offshore


Mechanics and Arctic Engineering, OMAE, Houston,
Vol V:243250.
Niedoroda, A.W. and Palmer A.C. 1986. Subsea trench
infill. Offshore Technology Conf. OTC5340: 445452.
Nyman, K.J. 1984. Soil resistance against oblique motions
of pipes. ASCE Journal of Transportation Engineering
110(2): 190202.
Offshore Soil Investigation Forum (OSIF) 1999. Guidance
notes on geotechnical investigation of marine pipelines
Rev 03.
Palmer, A.C. & Richards, D.M. 1990. Design of submarine
pipelines against upheaval buckling. Proc. Offshore
Technology Conf., Houston, 710 May 1990 OTC 6335:
551560.
Pedersen, P.T. & Michelsen, J. 1988. Large deflection
upheaval buckling of marine pipelines. Proc. Behaviour
of Off-Shore Structures (BOSS), Trondheim, Norway Vol.
III: 965980.
Powell, T.R., Fisher, R., Phillips, R. & Jee, T. 2002.
Reducing backfilling risks, Offshore Site Investigation
and Geotechnics Diversity and Sustainability, Society
for Underwater Technology.
Randolph, M.F. & Wroth, C.P. 1978. Analysis of deformation of vertically loaded piles, J. Geotech. Eng. Div.
ASCE, Vol.104, GT12:14651487.
Rawat, P.C., Agarwal, S.L., Malhotra, A.K., Gulhati, S.K. &
Rao, G.V. 1979. Determination of thermal conductivity
of soils: a need for computing heat loss through buried
submarine pipeline. Proc. Offshore Technology Conf.,
Houston, 30 April3 May 1979 OTC 3670: 27472753.
Rowe, R.K. & Davis, E.H. 1982. The behaviour of anchor
plates in sand. Geotechnique 32(1): 2541.
Schmertmann, J.H. 1991. The mechanical aging of soils,
Journal of Geotechnical Engineering, ASCE, Vol.117
(9): 12881330.
Schaminee, P.E.L., Zorn, N.F. & Schotman, G.J.M. 1990.
Soil response for pipeline upheaval buckling analyses:
full-scale laboratory tests and modelling. Proc. Offshore
Technology Conf., Houston, 710 May 1990 OTC 6486:
563572.
Schapp, D. 1982. Natural backfill of submarine pipeline
trenches, Offshore Oil and Gas Pipeline Technology,
Amsterdam.
Sills, G.C. 1995. Time dependent processes in soil consolidation, Compression and Consolidation of Clayey Soils,
Eds: Yoshikuni & Kusakabe, Balkema: 875890.
Silva, A.J. 1974. Marine Geomechanics: Overview and
Projections, In Deep Sea Sediments, Physical and
Mechanical Properties: 4576.
Skempton, A.W. & Northey, R.D. 1952. The sensitivity of
clays, Geotechnique 3: 3053.
Skempton, A.W. 1970. The consolidation of clays by gravitational compaction. Quarterly Journal of Geol. Soc.
London, Vol. 125: 373411
Small, S.W., Tamburello, R.D. & Piaseckyj, P.J. 1971.
Submarine pipeline support by marine sediments.
Offshore Technology Conf., OTC1357: 309318.
Stoutjesdijk, T.P., de Groot, M.B. & Lindenberg, J. 1998.
Flow side prediction method: influence of slope geometry, Canadian Geotechnical Journal, Vol. 35: 4354.
Tornes, K, Jury, J., Ose, B.A. & Thomson, P. 2000. Axial
creeping of high temperature flowlines caused by soil

113

ratcheting. Proc. 19th Int. Conf. Offshore Mechanics and


Arctic Engineering, OMAE2000, New Orleans.
Thorne, C.P., Wang, C.X. & Carter, J.P. 2004. Uplift capacity
of rapidly loaded strip anchors in uniform strength clay.
Geotechnique, Vol.54, No.8: 507517.
Trautmann, C.H. & ORourke, T.D. 1985. Lateral force-displacement response of buried pipe. ASCE Journal of
Geotechnical Engineering 111(9): 10771092.
Mendoza, M.J., and Hartlen, J. (1985). Compressibility of
clayey soil used in land reclamation, Proceedings of the
Eleventh International Conference on Soil Mechanics
and Foundation Engineering, San Francisco, Vol. 2:
583586.
Vanden Berghe, J-F, Cathie, D. & Ballard, J.-C. (2005)
Pipeline uplift mechanisms using finite element analysis,
Int. Conf. Soil Mech. Foundation Eng., Osaka.
Van Rijn, L.C. 1993. Principles of sediment transport in rivers,
estuaries and coastal seas, Aqua Publications, Amsterdam.
Verley, R.L.P. & Lund, K.M. 1995. A soil resistance model
for pipelines placed on clay soils. Proc. Offshore
Mechanics and Arctic Engineering Conf, Copenhagen,
1822 June 1995, Vol V: 225232.
Verley, R.L.P. & Sotberg, T. 1992. A soil resistance model
for pipelines placed on sandy soils. Proc. Offshore
Mechanics and Arctic Engineering Conf, Vol V-A pipeline
technology: 123131.
Wagner, D.A. Murff J.D. & Brennodden, H. 1987. Pipe-soil
interaction model. Proc. Offshore Technology Conf.,
Houston, 2730 April 1987 OTC 5504: 181190.

White, D.J., Barefoot, A.J. & Bolton, M.D. 2001. Centrifuge


modelling of upheaval buckling in sand. IJPMGInternational Journal of Physical Modelling in Geotechnics, 2: 1928.
Whitman, R. V. 1970. Hydraulic Fills to Support Structural
Loads. Journal of the Soil Mechanics and Foundations
Division, ASCE, Vol. 96, No. SM1: 2347.
Woodside, W. 1958. Probe for thermal conductivity measurement of dry and moist materials. Heat. Piping Air
Con, Sept: 163170.
Yang, L.A., Tan, T.S. & Leung, C.F. 2002. One-dimensional
self-weight consolidation of a lumpy clay fill,
Geotechnique, Vol 52, No 10: 713725.
Youd, T.L. 1973. Factors controlling maximum and minimum densities of sands, Am. Soc. for Testing and
Materials Spec. Tech. Pub. 523: 98112.
Young, A.G., Osborne, R.S. & Frazer, I. 2001. Utilizing
thermal properties of seabed soils at cost-effective insulation for subsea flowlines. Proc. Offshore Technology
Conf., Houston, 30 April3 May 2001 OTC 13137.
Zhang, J., Randolph, M.F. & Steward, D.P. 1999. An elastoplastic model for pipe-soil interaction of unburied
pipelines. Proc. Int. Offshore and Polar Engineering
Conf., Brest, 30 May4 June 1999: 185192.
Zhang, J., Stewart, D.P. & Randolph, M.F. 2002. Modeling
of shallowly embedded offshore pipelines in calcareous
sand. J. Geotechnical and Geoenvironmental Engineering,
ASCE, Vol.128, No.5: 363371

114

You might also like