Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Soil Dynamics and Earthquake Engineering 31 (2011) 930–940

Contents lists available at ScienceDirect

Soil Dynamics and Earthquake Engineering


journal homepage: www.elsevier.com/locate/soildyn

Analysis of buried pipelines subjected to reverse fault motion


Shantanu Joshi a, Amit Prashant a, Arghya Deb b, Sudhir K. Jain c,n
a
Department of Civil Engineering, Indian Institute of Technology Kanpur, UP 208016, India
b
Department of Civil Engineering, Indian Institute of Technology Kharagpur, WB 721 302, India
c
Department of Civil Engineering, Indian Institute of Technology Gandhinagar, Ahmedabad 382424, India

a r t i c l e i n f o a b s t r a c t

Article history: Presently available simplified analytical methods and semi-empirical methods for the analysis of buried
Received 5 July 2010 pipelines subjected to fault motion are suitable only for the strike-slip and the normal-slip type fault
Received in revised form motions, and cannot be used for the reverse fault crossing case. A simple finite element model, which
7 February 2011
uses beam elements for the pipeline and discrete nonlinear springs for the soil, has been proposed to
Accepted 8 February 2011
analyse buried pipeline subjected to reverse fault motion. The material nonlinearities associated with
Available online 24 March 2011
pipe-material and soil, and geometric nonlinearity associated with large deformations were incorpo-
Keywords: rated in the analysis. Complex reverse fault motion was simulated using suitable constraints between
Buried pipeline pipe-nodes and ground ends of the soil spring. Results of the parametric study suggest that the
Finite element analysis
pipeline’s capacity to accommodate reverse fault offset can be increased significantly by choosing a
Nonlinear quasi-static analysis
near-parallel orientation in plan with respect to the fault line. Further improvement in the response of
Structural design of pipelines
Oil and gas pipelines the pipeline is possible by adopting loose backfill, smooth and hard surface coating, and shallow burial
Earthquake induced ground movements depth in the fault crossing region. For normal or near normal orientations, pipeline is expected to fail
Pipe–soil interaction due to beam buckling at very small fault offsets.
Pipeline buckling & 2011 Elsevier Ltd. All rights reserved.
Reverse fault motion

1. Introduction and large section deformations near the fault crossing point. Hence,
it is necessary to design the pipeline which can safely accommo-
Buried steel pipelines with continuous joints are commonly date large fault offsets without being ruptured or buckled.
used for transporting oil, gas and water over long distances. Such a Faults are most commonly classified based on the direction of
pipeline crossing an active fault zone may be subjected to large, relative slip. Portion of the ground, which remains stationary
abrupt differential ground movement due to the fault rupture. during the slip is referred to as foot wall, and the other portion
Several major pipeline systems have been identified with the that slips over the foot wall is referred to as hanging wall. The
pipelines passing through active fault regions [1]. Reverse faults hanging wall in normal-slip faults moves downward and in
result from compressional plate tectonic environment and are reverse-slip faults upward with respect to the foot wall. A low
abundantly present throughout the world. In India, major active dip angle (less than 451) reverse fault is called a thrust fault. In
faults are of reverse or thrust type and mainly distributed in strike-slip fault, the slip takes place in the horizontal direction.
Kachchh (Western India) and Himalayan frontal (North-western Response of buried pipeline is significantly influenced by the type
India) regions [2,3]. Some of these reverse faults can potentially of fault motion and orientation of the pipeline with respect to the
produce large fault offset, as high as several metres. Many cases of fault line [7]. In general, a steel pipeline strained in direct tension
pipeline damage due to fault rupture have been recorded during due to fault rupture can safely accommodate a larger fault offset
recent major earthquakes [4–6]. For example, a case of severe value compared to when it is strained in direct compression [8,9].
pipeline damage was reported due to the rupture of Chelungpu Pioneering work in the analysis of pipeline subjected to fault
fault during 1999, Chi-Chi (Taiwan) earthquake [4]. The fault was motion was done by Newmark and Hall [10]. They developed a
steep reverse type (total length of about 105 km), and fault offsets simplified method for analysis of pipeline subjected to fault
of 4–10 m were observed along its length during the earthquake. motion. This method assumed the pipeline to be subjected to
The damaged portion of this pipeline went through local buckling direct tension due to the fault motion and ignored lateral
resistance of the soil. Hence, the analysis of the pipeline was
performed by assuming it to be a cable deforming in straight line.
n Kennedy et al. [8] revised the Newmark–Hall method by incor-
Corresponding author. Tel.: +91 79 2397 2574; fax: + 91 792397 2586.
E-mail addresses: amit_prashant@yahoo.com (A. Prashant), porating bending of the pipeline near the fault crossing point and
skjain@iitk.ac.in (S.K. Jain). considering the soil lateral forces. However, the formulae for

0267-7261/$ - see front matter & 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.soildyn.2011.02.003
S. Joshi et al. / Soil Dynamics and Earthquake Engineering 31 (2011) 930–940 931

radius of curvature proposed in this method were suitable only other hand, for the case of buried pipeline crossing a reverse fault,
case of pipeline being in tension. Based on the Kennedy’s Method, two different crossing angles, the fault dip angle (c) and the pipeline
Wang and Yeh [11] developed a simplified analysis method, crossing angle (b) are involved (Fig. 1). The fault motion in this case
which could analyse the case of pipeline subjected to direct depends on both the crossing angles, b and c, which are present in
tension as well as direct compression. The analysis procedure the horizontal and the vertical plane, respectively. Hence, the analysis
was based on the assumption of uni-axial bending of pipeline, and would require simultaneous application of fault offset components in
it was only applicable to the pipeline subjected to strike-slip fault all three major pipeline-directions. Consequently, the pipeline may be
motion. Recently proposed semi-empirical method by Takada subjected to bi-axial bending. In other words, the pipeline bending in
et al. [12] and simplified analytical method by Karamitros this case cannot be assumed to take place in a predetermined plane
et al. [13] are also based on the assumption of uni-axial bending as can be assumed for the case of strike-slip fault. Besides, due to
of pipeline and are only applicable to the case of strike-slip fault. hanging wall moving towards the foot wall during reverse fault
Similarly, the semi-analytical method proposed by Trifonov movement, one can expect a significant amount of axial shortening in
et al. [14] is also suitable for only the cases of strike-slip fault the pipeline. Compressive stresses may cause buckling of the buried
and normal-slip fault crossings. pipeline either in the beam mode or in the shell mode. Thus, the
Literature review suggests that previous research in the analysis of various simplified methods mentioned earlier prove to be inadequate
pipeline subjected to fault motion has been mainly focused on the for the analysis of pipeline crossing a reverse fault, as none of those
case of strike-slip fault [8,11–16]. The simplified methods for deter- methods can analyse a pipeline subjected to bi-axial bending along
mination of total strain in a pipeline subjected to strike-slip fault with axial compression or predict its subsequent buckling. Hence, the
motion are based on the assumption that the fault motion would present research is focused on analysing the response of buried
cause uni-axial bending of the pipeline near the fault crossing point pipelines subjected reverse fault motion with due considerations to
accompanied by either axial elongation or axial shortening [11,13]. the three dimensional nature of reverse fault motion, bi-axial pipeline
The possibility of axial elongation or shortening depends on the bending and probable pipe buckling.
orientation of the pipeline with respect to the fault line in plan. The In this study, an attempt was made to analyse the response of
assumption of uni-axial pipeline bending in the case of strike-slip buried pipelines subjected to reverse fault motion by developing a
fault motion is justified by the fact that the problem involves a single simple finite element model using 3-D beam elements. Soil
crossing angle (b) between the fault-line and the pipeline. On the surrounding the pipeline was modelled using nonlinear springs,

Anchor ψ
point Δ Ground surface

Deformed profile Z
Δz Δ of pipeline

Original profile X’
Δ Cos (ψ) of pipeline Anchor
point

X
L2, or
Lo, if L2 is too large

β Y
Anchor
point
L1, or
Lo, if L1 is too large

Anchor
point Fault crossing point

Fault line

Δy

Δx

Fig. 1. Deformation of buried pipeline subjected to reverse fault motion: (a) section and (b) plan.
932 S. Joshi et al. / Soil Dynamics and Earthquake Engineering 31 (2011) 930–940

which support the pipeline at discrete points. Complex 3-D c are the angles of intersection of the pipeline with the fault, in
reverse fault motion was simulated by applying non-zero dis- plan and in section, respectively. X, Y, and Z axes represent
placements to the soil spring ends through suitable constraints longitudinal, transverse horizontal, and transverse vertical direc-
between pipe-nodes and corresponding soil spring ends. The tions of the pipeline, respectively. The fault offset (D) along the
major features incorporated in the model were pipe-material fault plane can be resolved in components along three pipeline
nonlinearity, nonlinear Winkler spring model of the soil, geo- directions as follows [17]:
metric nonlinearity associated with large deformations, and post- Dx ¼ DcosðcÞsinðbÞ ð1Þ
buckling behaviour of pipeline (in case of beam buckling). Using
this FEM model, buried steel pipelines were analysed for reverse Dy ¼ DcosðcÞcosðbÞ ð2Þ
fault motion to study the influence of design parameters viz.
crossing angle, backfill properties, burial depth, pipe surface Dz ¼ D sinðcÞ ð3Þ
property, pipe material and cross-section properties on maximum Movement of the soil mass along the fault plane results in
compressive strain, and buckling of the pipeline. pipeline–soil interaction in all the three pipeline directions. Dx is
resisted by frictional force developed at the pipeline–soil interface
over the unanchored length of the pipeline. For a reverse fault, Dx
2. Pipeline–soil interaction results in axial shortening of the pipeline. If the anchor point L1 or
L2 (Fig. 1) is sufficiently away from the fault crossing point, then
Fig. 1 shows a typical deformation profile of buried pipeline, in the pipeline can develop the effective unanchored length through
section and in plan, when subjected to reverse fault motion. b and unrestrained sliding in the soil. At the endpoint of this effective

Transverse spring

Axial spring

Fault crossing
point

Z
Vertical
X
spring

10 m

570 m

1m

Δy

30 m

Δx
30 m
Δz

570 m

Fig. 2. Geometry of proposed FEM model for buried pipeline.


S. Joshi et al. / Soil Dynamics and Earthquake Engineering 31 (2011) 930–940 933

unanchored length (say at L0 from the crossing point), the


Stress
maximum axial force developed in the pipeline near the fault
crossing point is completely removed by soil frictional force.
Hence, the pipeline can be considered to be effectively anchored
to the ground at distance L0 from the crossing point as shown
in Fig. 1b. For the case where L1 or L2 is smaller than L0, pipeline’s E2
movement will be restrained at the location of physical anchor
point itself. In this study, the anchor points were assumed to be σy
sufficiently away from the fault crossing point on either side of it,
which allowed for development of the effective unanchored
length. Dy induces lateral soil pressure on the pipeline in the
horizontal direction and results in bending of the pipeline in X–Y
plane. Similarly, Dz results in bending of the pipeline in X–Z plane. E1
However due to shallow burial depth of the pipeline, the limiting
uplift resistance of the soil is generally much smaller than the
ultimate bearing resistance. In this study, the above described
pipeline–soil interaction components in three directions were
equivalently represented by discrete nonlinear springs [18–20],
which supported the pipeline in the three directions. εy
Strain

Fig. 3. Pipeline material stress–strain curve.

3. Proposed finite element model


Soil surrounding the pipeline was modelled using discrete
3.1. Model details nonlinear springs. Similar nonlinear winkler foundation models to
represent the soil-structure interaction have been proposed by
FEM model for the buried pipeline was developed using the many researchers for buried pipes [11–13] and for piles [23,24]. A
commercial finite element software ABAQUS (Version 6.7). Fig. 2 pipe node was supported by a set of three mutually perpendicular
shows the geometry adopted for the proposed finite element soil springs viz. Axial, Transverse Horizontal, and Transverse
model. A 1200 m long straight pipeline segment was considered Vertical. These springs simulated the components of pipe–soil
for the analysis. The length of the pipeline was selected based on interaction in the respective pipeline directions. The soil-springs
the model lengths proposed by previous researchers [12,13]. The were modelled using connector elements (CONN3D2 in ABAQUS)
long length of the model justified the assumption that the anchor between the pipe-nodes and the corresponding ground nodes [21].
points were sufficiently away from the fault crossing point. The All the connector elements were associated with axial connector
fault was assumed to cross the pipeline segment at the centre of sections, which generate spring forces only in the local axial
its length. The pipeline segment was modelled using 2-noded direction of the element. For each soil spring, an elastic-perfectly
linear beam elements (B31 in ABAQUS) in space. The finite plastic force–deformation relationship was assumed and expres-
element formulation for this element is based on Timoshenko’s sions for the same were adopted from ALA-ASCE guidelines [18].
beam theory and takes into account transverse shear deforma- These expressions are based on field and laboratory experimental
tions [21]. This formulation can be considered appropriate for investigations on pipeline response, as well as general geotechni-
thin pipe-sections with large diameters, commonly used for steel cal approaches for related structures such as piles, anchor plates.
oil and gas pipelines. The entire 1200 m length of the model was Schematic representation of the force–deformation relationship
divided into three regions for the meshing. Element size was kept as adopted in this study for three springs is presented in Fig. 4.
uniform within each region. Region-1, closest to the fault crossing Force–deformation relationship for transverse horizontal springs
point had a total length of 30 m (15 m on either side of the fault of buried pipe (Fig. 4(b)) is similar to that adopted for lateral pile
crossing point). The smallest element size of 0.25 m was adopted springs [23,24]. The force–deformation relationship of soil springs
in Region-1 since the maximum strain in the pipeline is likely to for pile is considered the same in two lateral directions. However,
occur in this region. The element size was increased to 0.5 m in for a shallow buried pipeline the uplift resistance Pu provided by a
Region-2, which began at the ends of Region-1 on both sides and limited depth of top soil cover (backfill) is generally much smaller
extended up to 30 m. Region-3 represented the rest of the length than the bearing resistance Pb of soil below. This makes the
of the model on both sides and it had elements of size of 5 m. vertical spring force–deformation relationship unsymmetrical for
Convergence studies were performed on the model to arrive at the pipeline, as shown in Fig. 4(c). For the section of the pipe near
the above described mesh configuration [22]. For a pipeline the fault, the pipe–soil interaction could be more complex due to
subjected to large reverse fault offset (of the order of several large deformations of the pipeline, which required a smaller
metres), a certain length of the pipeline on either side of the fault spacing between the soil-spring sets in this section. In the
crossing point is expected to undergo material yielding due to proposed model, length of this region was assumed to be 30 m
development of large bending moments (in X–Y and X–Z planes) on either side from the fault crossing point and spring sets were
and axial compressive force. This expected material nonlinearity provided at a uniform spacing of 1 m. For the remaining length of
was considered in the model by modelling the pipeline as an the model (570 m on either side) where the pipeline was
elasto-plastic material with a bi-linear strain-hardening curve of expected to remain elastic, the soil-spring sets were spaced at
the type shown in Fig. 3. The von-Mises isotropic hardening 10 m.
model available in ABAQUS was used for the pipeline material.
The bending and axial compressive deformations of this near- 3.2. Simulation of fault motion
fault pipe section are often of the same order of magnitude as the
pipe itself. Hence, geometric nonlinearity was incorporated in the A nonlinear quasi-static solution is desirable for the problem
analysis. of a pipeline crossing an active fault, which assumes that the fault
934 S. Joshi et al. / Soil Dynamics and Earthquake Engineering 31 (2011) 930–940

Force Force Force


Tu Pl

Pu
Xu Disp. Yu Disp. Zb Disp.

Xu Yu Zu

Tu Pl Pb

Fig. 4. Force–deformation relationships for the soil-springs: (a) for axial spring, (b) for transverse horizontal spring, and (c) for transverse vertical spring.

displacement components are applied at a sufficiently slow rate simultaneous application of Dx and Dz. Spring forces would act
so as to ensure that the dynamic effects are negligibly small. The at Node-3 in the directions V0 and A0 , as shown in the figure, and
velocity of movement on one side of the fault with respect to the solver will have to achieve an equilibrated state in the deformed
opposite side is generally sufficiently low. Hence, the effects of configuration. Fig. 5(a) also suggests that the magnitude of the
fault rupture on the pipeline can be considered similar to those rotation of any spring will change if the original length of the
for statically applied relative movement [8]. In this study, an spring is changed, even if the magnitude of fault offset component
explicit dynamic analysis was performed using ABAQUS/Explicit. applied to this spring remains the same. This in turn will affect
The ABAQUS/Explicit solver was chosen for the nonlinear analysis the results of the analysis. Since the rotation of these springs can
due to its capabilities of analysing post-buckling behaviour more be significant during simulation of large fault offsets of the order
efficiently and to avoid the convergence problems typically of several metres, the dependence of the rotations on the original
encountered by implicit solvers in the post-buckling domain [21]. length of the spring (which may exhibit considerable variation)
ABAQUS/Explicit, being an explicit dynamic finite element code, will introduce a measure of arbitrariness in the results.
performs dynamic analysis on the model and requires an analysis To avoid this problem the springs were prevented from
time to be associated with non-zero displacement boundary rotating by constraining the ground node and pipe node of each
conditions. Hence a nonlinear static solution was achieved for spring to move identically in the directions in which the fault-
this problem by converting dynamic analysis into a quasi-static offset components of the spring were zero, as shown in Fig. 5(b).
mode. Quasi-static conditions were simulated in the analysis by Table 1 shows the boundary conditions assigned to ground ends
applying fault offset components to soil-spring ends through a of the soil-springs to simulate the reverse fault motion. For the
smooth loading function of time (Smooth Step Function in ABAQUS) sake of convenience, the ground ends of the soil-springs on the
which avoids any sudden load changes, and by keeping a left side of the fault crossing point associated with axial, trans-
sufficiently long loading duration for the fault offset components. verse, and vertical springs are referred to as axial loading points,
A small duration of loading would generate significant dynamic transverse loading points, and vertical loading points, respec-
effects in the simulation. Choosing a sufficiently long duration of tively. Similarly, the ground ends of soil-springs on the right side
loading ensures that these dynamic effects are negligibly small. of the fault crossing point are referred to as axial stationary
However, an excessively large loading duration increases the points, transverse stationary points, and vertical stationary points,
computational time substantially. Hence, numerical experiments respectively. Each ground end was constrained to move with
involving several trial runs were performed on the model to arrive its pipe node in two directions and it was either assigned a
at an optimum loading duration that satisfies quasi-static condi- fault offset component or a restraint in the third direction.
tions [22]. A constant magnitude reverse fault motion was Fig. 5(b) shows the same portion of the pipeline segment as
applied, and the satisfaction of the quasi-static condition was represented by Fig. 5(a) in X–Z plane, subjected to fault offset
checked by comparing time histories of the kinetic energy (KE) components Dx and Dz. For the vertical spring (V) in the figure,
and internal energy (IE) of the model for each trial run. The Node 1 is now constrained to have the same displacements
simulation was assumed to have run quasi-statically if the kinetic as Node 3 (pipe node) in X and Y directions. Similarly, Node 2
energy of the model was found to be negligibly small compared to (axial soil spring ground end) and the Node 3 are constrained in
the internal energy throughout the analysis. Based on these Y and Z directions. Springs in this case maintain their original
studies, a loading duration of 20 s was chosen for the analysis. orientation after the application of fault offset components with-
To simulate reverse fault motion, the fault offset components out any rotation about their respective ground node. Similar
Dx, Dy, and Dz (Eq. (1)–(3)) were assigned simultaneously to constraints were assigned between pipe nodes and corresponding
ground ends of the axial, transverse, and vertical springs, respec- ground ends of the transverse springs to avoid rotation. Because
tively, over half length of the pipeline segment. This simultaneous of these constraints between the pipe nodes and the soil ground
application of fault offset components resulted in rotation of nodes, the results of the analysis for a given fault motion were
springs if the ground ends of these springs were restrained in the observed to be independent of the assumed length of the soil
other two directions. This has been explained in Fig. 5(a), which springs.
shows a portion of the pipeline segment in the X–Z plane
subjected to fault offset components Dx and Dz. Node 1 (ground
end of vertical soil-spring) is subjected to Dz and restrained in X 4. Response of buried pipeline to reverse fault motion
and Y directions to avoid any free body motions during the
analysis. Similarly, Node 2 (ground end of axial soil-spring) is The factors influencing response of buried pipeline at reverse
subjected to Dx and restrained in Y and Z directions. These fault crossing include the fault offset (D), pipeline crossing angle
restraints would lead to rotation of the springs due to (b), fault dip angle (c), native and backfill soil type, burial depth
S. Joshi et al. / Soil Dynamics and Earthquake Engineering 31 (2011) 930–940 935

Node 3
A’

Node 2
Δx A

Pipeline

V V’

Δz V = Vertical spring in original orientation


V’ = Vertical spring after applying displacement
Node 1
A = Axial spring in original orientation
A’ = Axial spring after applying displacement

Δx
Node 3

A’
Node 2
A
Pipeline
V’
V
Z

Δz X
Node 1

Fig. 5. Schematic representation of portion of FEM model in X–Z plane: (a) rotation of soil-springs due to restrained spring ends and (b) Rotation avoided by assigning
suitable constraints between pipe-node and spring ground ends.

Table 1
Boundary conditions to soil spring ends with suitable constraints.

Soil reference point type Boundary conditions

X-direction Y-direction Z-direction

Axial loading points Dx Constrained to pipe node Constrained to pipe node


Transverse loading points Constrained to pipe node Dy Constrained to pipe node
Vertical loading points Constrained to pipe node Constrained to pipe node Dz
Axial stationary points Restrained Constrained to pipe node Constrained to pipe node
Transverse stationary points Constrained to pipe node Restrained Constrained to pipe node
Vertical stationary points Constrained to pipe node Constrained to pipe node Restrained

(H), pipe diameter (c) and thickness (t), pipe material, and pipe enable selection of design parameters that would enhance the
surface characteristics (f) [25]. Out of the above listed factors, D, pipeline’s capacity to accommodate the reverse fault offset.
c, and the type of native soil depend on geological and geotech-
nical characteristics of the site and their values, as obtained from
the site investigations, need to be considered in the analysis. 4.1. Influence of crossing angle
However, the designer has the choice to vary the factors b, the
backfill soil type, H, d, t, pipe material, and f so as to improve the A reverse fault motion was generated by choosing a particular
pipeline’s performance and optimise the design. Influence of combination of the magnitudes of D, b, and c. Pipe cross-section
these design parameters on the response of the buried steel pipe with outer diameter (D) 0.91 m and wall thickness (t) 12 mm was
to reverse fault motion was studied using the FEM model chosen for the analysis. Pipe material was assumed to be API5L-
described above. These analyses were performed with a view to X65 grade steel [26] with a bi-linear stress–strain curve (Fig. 3).
936 S. Joshi et al. / Soil Dynamics and Earthquake Engineering 31 (2011) 930–940

The top of the pipeline was assumed to be buried under 1.30 m of constant magnitude of D for various b angles, ecmax increases
medium density sand with friction angle, F ¼351 and unit weight, consistently as the crossing angle increased from 10–901. At each
g ¼18 kN/m3. Coefficient of friction of 0.8 was adopted for the value of D, the smallest value of ecmax was observed at b ¼101, and
interaction between the pipe surface and soil. The corresponding the largest at b ¼901. This significant strain reduction due to
spring yield forces and mobilising displacements were derived as change in the pipeline’s crossing angle was observed to be true for
per ALA-ASCE guidelines [18] and are listed in Table 2. The native both the dip angles investigated during this study. For example,
soil and the backfill were assumed to have the same mechanical for the case of c ¼ 401 (Fig. 6a), ecmax at D ¼2 m was observed to
properties. For each b–c combination, the pipeline was subjected reduce from 21% to 5% when b was changed from 901 to 101. In
to the maximum fault offset of 4 m in increments of 0.5 m. general, a consistent rise in ecmax was observed for a constant
The results of the analyses have been summarised in value of D, as b was successively increased. Results of this study
Fig. 6(a) and (b), which shows variation of the maximum total suggest that the pipeline compressive strain can be significantly
compressive strain (ecmax) in the pipeline with the fault offset reduced at reverse fault crossing by choosing a near-parallel
magnitude (D), for different pipeline crossing angles (b), for dip orientation of the pipeline with the fault line in plan. Such an
angles c ¼ 401 and 701. The total strain includes axial strain and orientation would reduce the direct axial shortening of the pipe-
bending strains at a pipe section. For each b of the pipeline, ecmax line and cause the horizontal component of fault offset (D Cos c)
can be seen to increase as the value of D gradually increased up to to be predominantly accommodated by transverse horizontal
4 m. Figure shows that, when the results are compared at a bending of the pipeline. On the other hand, for normal or near-
normal orientation, component D Cos c would need to be pre-
dominantly accommodated by axial shortening of the
pipeline. Fig. 6 also shows that for constant values of D and b,
Table 2 ecmax is considerably reduced when the fault dip angle is increased
Soil-spring properties used in the study to evaluate effect of crossing angle.
from 401 to 701. This was found to be true for all the crossing
Spring type Yield force (kN/m) Yield displacement (mm) angles (b) studied here. This suggested that the pipeline’s capacity
to safely accommodate reverse fault offset magnitudes would
Axial 40 3 reduce considerably at shallow dip angles.
Transverse 318 11
Vertical uplift 52 2.3
Vertical bearing 1360 100 4.2. Influence of backfill type

Fig. 7 shows the effect of backfill type on maximum total


compressive strain in the pipeline at various reverse fault offset
magnitudes. The same pipeline segment (with the same burial
0.30
depth and pipe surface properties), as adopted in the previous
β = 10° ψ = 40º section, was studied for three granular backfills: loose (F ¼301),
β = 30°
Maximum total compressive

0.25 medium dense (F ¼ 351), and dense (F ¼401). The native soil (soil
β = 45°
surrounding the trench) was assumed to be firm cohesive soil
0.20 β = 60°
having a cohesion value of c¼50 kPa. Fault motion parameters b
strain, εcmax

β = 90°
and c were chosen to be 301 and 401, respectively, and the
0.15 pipeline was subjected to maximum fault offset of 2 m in
increments of 0.5 m. For a constant magnitude of D, ecmax value
0.10 increased as the compactness of granular backfill increased. This
was true for all the considered values of D. Such increase in the
0.05

0.00 0.12
0 1 2 3 4 D = 0.9144m, Firm Native Soil, β = 30°, ψ = 40°
Fault displacement , Δ (m)
Maximum total compressive strain, εcmax

0.10

0.30
0.08
ψ = 70º
Maximum total compressive

0.25 β = 10°
β = 30° Loose Backfill
0.20 β = 45° 0.06 Medium Dense Backfill
strain, εcmax

β = 60° Dense Backfill


β = 90°
0.15 0.04

0.10
0.02
0.05

0.00 0.00
0 1 2 3 4 0.0 0.5 1.0 1.5 2.0
Fault displacement , Δ (m) Fault displacement, Δ (m)

Fig. 6. Effect of the crossing angle (b) on the pipeline performance: (a) for the dip Fig. 7. Effect of the type of backfill on the maximum total compressive strain in
angle c ¼ 401 and (b) for the dip angle c ¼701. the pipeline.
S. Joshi et al. / Soil Dynamics and Earthquake Engineering 31 (2011) 930–940 937

ecmax was expected due to the fact that increase in F of the backfill 0.14
would increase limiting frictional force (Tu) as well as limiting
uplift force (Pu) of the soil. Results of this study suggested that the D = 0.9144m, β = 30°, ψ = 40º, Loose backfill
maximum compressive strain in the pipeline could be reduced by 0.12

Maximum total compressive strain, εcmax


a fair amount by placing a very loose backfill in the fault crossing
region and by avoiding its unnecessary over-compaction. To
0.10
confirm the consistency of these results, the same pipeline was
analysed again for the three backfill types, for soft and stiff
cohesive native soils having c¼20 and 100 KPa, respectively [22]. H/D = 1.92
0.08
These analyses were then repeated for three other b–c combina- H/D = 3.78
tions; (b ¼301, c ¼151), (b ¼901, c ¼151), and (b ¼901, c ¼401). In
all these cases, ecmax values were the smallest for the loose backfill 0.06
(F ¼301), suggesting that the pipeline can be expected to perform
best under those conditions.
0.04

4.3. Influence of pipe surface characteristics


0.02
Further analyses were performed on the same pipeline section
to evaluate influence of soil–pipe interface friction on pipeline’s
0.00
performance. For this study the native soil was assumed to be
0 0.5 1 1.5 2
firm cohesive soil (c ¼50 kPa) and the backfill to be loose granular
(F ¼301). Fig. 8 shows response of the pipeline to given reverse Fault displacement , Δ (m)
fault motion for two different surface characteristics. Steel pipe- Fig. 9. Effect of burial depth on the maximum total compressive strain.
line with a smooth and hard surface coating was assumed to have
the coefficient of friction f to be 0.5, and pipeline without any 4.4. Influence of burial depth
special coating was assumed to have this coefficient to be 0.8 [19].
Considerable reduction in the value of ecmax can be seen here due
The same pipeline section, as above, was analysed for two
to use of smooth and hard surface coating in the fault crossing
burial depths. Fig. 9 shows plots of ecmax against the applied D for
zone. At D ¼2 m, the pipeline compressive strain can be seen to
two burial depths. ecmax was observed to increase as the soil cover
reduce from 12% for f¼0.8 to 0.9% for f¼0.5. The consistency of
(h) increased from 1.3 to 3 m. Similar effect of the burial depth
these results was also confirmed for medium dense (F ¼351) and
was observed when the same pipeline section was analysed for
dense (F ¼401) backfills; these results are available else-
medium dense and dense backfills. Results of the parametric
where [22]. Smooth surface coating reduces the value of coeffi-
study showed that the increase in burial depth increased the
cient of friction between pipe and the soil, and leads to reduction
values of limiting uplift soil force and limiting pipe–soil friction
in the value of limiting soil frictional force (Tu) acting over unit
force acting on unit length of the pipeline, and consequently led
length of the pipe. Consequently, for a constant magnitude of D,
to higher maximum compressive strain for a constant magnitude
the value of maximum axial force developed in the pipeline near
of D. Thus, burial depth as shallow as possible is preferable in the
the fault crossing would reduce and this force will be dissipated
fault crossing zone. For a granular backfill, minimum cover of D/6
over larger length of the pipeline. However, due consideration
(up to the top of pipe) is adequate, considering only the dead load
should be given to the effect of aging on properties of the surface
of soil [27]. However, the live loads and environmental factors
coating, which generally leads to increase in the friction
may also have an influence on the minimum soil cover.
coefficient.

4.5. Influence of pipe cross-section and material properties

0.10
A study was carried out to see effect of the pipe outer diameter
D = 0.9144m, β = 30°, ψ = 40°, Loose backfill
Maximum total compressive strain, εcmax

on the performance of pipeline. Two pipelines having outer


diameters of 0.91 and 1.3 m, respectively, but with the same wall
0.08
thickness of 12 mm were analysed for the case b ¼301 and
c ¼451, for a fault offset of 2 m. The maximum total compressive
f = 0.5
strain in the pipeline reduced from 7.35% to 5.80% when the outer
0.06
f = 0.8 diameter was increased from 0.91 to 1.3 m. One may also
consider replacing the pipeline section in the fault crossing zone
with the one having larger wall thickness and the same internal
0.04
diameter in order to reduce the maximum compressive strain in
the pipeline. The use of high yield strength material for the
pipeline could be more suitable in the fault crossing zone. For a
0.02
given pipe cross-section, pipe material with higher yield strength
can reduce the compressive strain in the pipeline compared to a
lower yield strength material.
0.00
0 0.5 1 1.5 2
4.6. Buckling of pipeline
Fault displacement , Δ (m)

Fig. 8. Effect of pipe surface characteristics on the maximum total compressive Buried pipeline subjected to large compressive force generally
strain. buckles either in the shell mode (local buckling) or in the beam
938 S. Joshi et al. / Soil Dynamics and Earthquake Engineering 31 (2011) 930–940

Deformed profile
of soil

Anchor
point Deformed profile
of pipeline
Anchor
Δz wmax
point

Anchor
point Pipeline profile
before deformation

Deformed profile
of pipeline
Anchor
point
wmax Anchor
Δz point

Anchor
point Deformed profile
Pipeline profile
before deformation of soil

Fig. 10. Deformed profile of the pipeline near the fault crossing point in X–Z plane (a) for the case b ¼ 901, c ¼151, and D ¼ 0.5 m; and (b) for the case b ¼901, c ¼ 151, and
D ¼1 m.

mode [28,29]. In the FEM model, as the entire pipeline segment


was modelled using beam elements, section deformations due to
local buckling could not be captured directly. Hence, local buck-
ling in the model was predicted based on the value of maximum
compressive strain in the pipeline, while the beam mode of
buckling was captured in the analysis. Theoretical wrinkling
strain (onset of wrinkling) for steel pipeline as given by ASCE-
TCLEE Guidelines [19] is given by the following expression:

ec ¼ 0:6t=R ð4Þ

where t is the pipeline wall thickness and R the pipeline radius.


Results of the experimental studies carried out on steel pipes
suggest that the pipes would normally begin to wrinkle at strains
of ½ to ¼ of the theoretical value [19]. IITK-GSDMA Guide-
lines [17] suggest suitable value for the allowable wrinkling
strain to be the mean of these experimental wrinkling strain Fig. 11. Time variation of internal energy (IE) and kinetic energy (KE) of the model
for the case b ¼901, c ¼151.
values. The corresponding expression for the wrinkling strain is

ec ¼ 0:175t=R ð5Þ

This expression for the allowable wrinkling strain was adopted the pipeline with the adjacent soil near the fault crossing
in the present research. It was assumed that local buckling would point. Fig. 10(b) shows the schematic representation of deforma-
initiate if the maximum compressive strain occurring in the tion profile of the same pipeline for the same b–c combination at
pipeline at any location exceeds this allowable value. However, D ¼1 m. Here, a distinct change in the deformed shape was
designers may choose the value for the allowable wrinkling strain observed. wmax (2.395 m) was observed to be about 10 times
as they consider appropriate based on experience. higher than the applied Dz (0.25 m). Displacement profile sug-
The model could successfully capture the beam mode of gested that the pipeline had undergone large lateral deformations
buckling. A certain length of pipeline on either side of the fault over a small stretch in the vicinity of the fault crossing point
crossing point showed abrupt and very large displacement in the consequent to beam buckling. A sharp rise observed in ecmax from
vertical direction. This displacement was several times higher 1% at D ¼0.5 m to 21% at D ¼1 m appeared to confirm this result.
than the actual fault offset component applied to the model. Further confirmation of the occurrence of beam buckling is
Fig. 10(a) shows schematic representation of the original provided by the time histories of the kinetic energy (KE) and
deformed profile (as observed in ABAQUS Postprocessor) of the the internal energy (IE) of the model during the analysis. Fig. 11
pipeline near fault crossing point for the case b ¼901 and c ¼151 shows these plots for the case discussed above. A sharp rise in KE
at D ¼0.5 m. At this fault offset value, the maximum vertical can be seen at the instance corresponding to the application of a
displacement of the pipeline (wmax) was nearly the same as Dz fault-offset of D ¼ 1 m. This rise confirmed the generation of large
(0.13 m). This suggested that the pipeline displaced with the soil velocities in the pipeline, resulting from the sudden movement of
in the vertical direction when D ¼0.5 m was applied to the model. a certain portion of the pipeline from one configuration to
Deformation profile of the pipeline suggests no loss of contact of the other.
S. Joshi et al. / Soil Dynamics and Earthquake Engineering 31 (2011) 930–940 939

Displacement profiles of the pipeline for all the cases shown of three mutually perpendicular nonlinear springs. Soil material
in Fig. 6 were similarly studied to cheque for the occurrence of nonlinearity was simulated using an elastic-perfectly plastic
beam buckling. For the cases having c ¼701, no beam buckling force–deformation curve for each type of spring. Complex 3-D
was observed up to the maximum fault offset of 4 m for all reverse fault motion was simulated by applying the fault offset
crossing angles. For c ¼401, the pipeline buckled in the beam- components to the respective free ends of soil-springs over the
mode well before reaching the 4 m fault offset for the crossing half length of the model. These displacements were applied quasi-
angles 451, 601, 751, and 901. However, for the same pipeline, statically. Suitable constraints were assigned between pipe-nodes
beam buckling was not observed at c ¼401 for the crossing angles and corresponding ground ends of soil springs to avoid rotations
101 and 301 up to the maximum applied fault offset. The pipeline of the springs during application of the fault motion. Explicit
with b ¼101 and 301 was further analysed for different conditions solver in ABAQUS was used for the nonlinear analysis due to its
of backfill, burial depth and pipe-surface properties. However, no capabilities of analysing post-buckling behaviour more efficiently
beam buckling was observed in any of the cases. On the other and to avoid the convergence problems, typically encountered by
hand, the pipeline with b ¼901 showed beam buckling in all the implicit solver in the post-buckling domain. Geometric nonlinear-
cases, for all combinations of backfill soil, burial depths and pipe- ity associated with the large deformations was incorporated in
surface properties. Thus, for the pipeline segment analysed in this this analysis. Local buckling of the pipeline was predicted based
study, it was observed that beam buckling of the pipeline would on the value of maximum compressive strain in the pipeline,
be primarily influenced by the pipeline’s crossing angle (b) and while the beam buckling mode was successfully captured in the
the fault dip angle (c). While no beam bucking occurred when simulation.
crossing angle was 301 or lower, it is possible that the pipeline The following conclusions could be drawn based on the
might fail due to local buckling in these cases. This local buckling studied response of pipeline subjected to reverse fault motion
would occur at a pipe-section where ecmax exceeded the allowable using the FEM model.
value of the wrinkling strain. Occurrence of beam buckling at
normal or near normal orientation of the pipeline with the fault (a) For the steel pipeline subjected to reverse fault motion,
line lead to failure of the pipeline at small fault offset magnitudes. maximum total compressive strain was always found to be
On the other hand, better performance of the pipeline was more critical than the maximum total tensile strain.
observed for the near-parallel orientations of the pipeline. For (b) The capacity of buried pipeline to accommodate the reverse
these orientations pipelines can be protected against local buck- fault offset could be increased substantially by adopting a
ling by limiting the maximum compressive strains in the pipe near-parallel orientation in plan with respect to the fault line,
walls. This reduction can be achieved by suitably changing the thereby accommodating the horizontal component (Dx) of the
backfill type, pipe-surface characteristics, burial depth, pipe sec- fault offset primarily by bending of the pipeline in the
tion dimensions and pipe material properties. horizontal plane (X–Y) plane. Out of all the investigated
design parameters, the pipeline crossing angle (b) was found
4.7. Comparison of design parameters to be the most influencing factor on the response of the
buried pipeline in terms of maximum compressive strain and
To study the relative influence of the above described design its buckling.
parameters on the response of pipeline, the results were com- (c) For low crossing angles, it was inferred that the buried
pared at D ¼2 m. Pipeline crossing angle (b) was found to be the pipeline buckled locally due to wall wrinkling.
most influencing design parameter. At D ¼2 m, the maximum (d) Pipeline crossing the fault line in normal or near-normal
compressive strain was reduced by 76% when b was changed orientation generally buckled in the beam mode at very small
from 901 to 101 (Fig. 6(a)). Influence of change in burial depth fault offset magnitudes, consequently leading to its early
from 3 to 1.3 m on the strain reduction was nearly the same as failure.
influence of change in friction coefficient (f) from 0.8 to 0.5 at (e) For a constant magnitude of D, the pipeline experienced
D ¼2 m. In both the cases, the maximum compressive strain was larger maximum compressive strain for a shallower dip angle.
reduced by about 20% (Figs. 8 and 9). However, at smaller fault Hence, the capacity of a pipeline to safely accommodate fault
offsets the influence of change in friction coefficient (f) was offset magnitude may be considerably smaller at a thrust fault
relatively insignificant in comparison to the influence of change than at a steeper reverse fault.
in the burial depth. The influence of backfill type was found to be (f) The capacity of the pipeline to safely accommodate the
relatively insignificant in comparison to the other design para- reverse fault offset can be further increased by choosing a
meters. At D ¼2 m, the maximum compressive strain reduced by loose granular backfill, adopting a shallower burial depth,
about 7% when the dense backfill was replaced with loose backfill using a smooth and hard surface coating, increasing pipe-wall
(Fig. 7). A reduction of 21% in the maximum compressive strain thickness or by using higher yield strength material.
was observed when the pipeline’s outer diameter was increased
from 0.91 to 1.3 m without changing the thickness. In another The present study ignored the effect of internal pipe pressure
parametric study the internal diameter of the pipeline was kept as the pipeline was modelled using beam elements. Also, precise
constant and the wall thickness was increased by 2 mm. As a simulation of the local buckling and large section deformations
result, about 16% reduction in the strain was observed [22]. were ignored due the use of beam elements.

5. Summary and conclusions


Acknowledgements
Buried steel pipeline subjected to reverse fault motion was
analysed using a simple FEM model which used 3D beam Financial support from Poonam and Prabhu Goel Foundation at
elements to model the entire pipeline segment. Pipe material Indian Institute of Technology Kanpur to carry out this research is
nonlinearity was considered in the analysis by associating a bi- gratefully acknowledged. Authors would like to express their
linear stress–strain curve to the beam elements. Pipeline–soil deep gratitude to Dr. Debasis Roy of Indian Institute of Technology
interaction was modelled by connecting each pipe-node to a set Kharagpur for his personal help during the research.
940 S. Joshi et al. / Soil Dynamics and Earthquake Engineering 31 (2011) 930–940

References [14] Trifonov OV, Cherniy VP. A semi-analytical approach to a nonlinear stress–
strain analysis of buried steel pipelines crossing active faults. Soil Dynamics
and Earthquake Engineering 2010;30(11):1298–308.
[1] Johansson J, Konagai K. Fault movement related damage examples and fault [15] Vazouras P, Karamanos SA, Dakoulas P. Finite element analysis of buried steel
provisioned design case histories. In: Bhattacharya S, editor. Design of pipelines under strike-slip fault crossings. Soil Dynamics and Earthquake
Foundations in Seismic Areas: Principles and Applications. Kanpur (India): Engineering 2010;30(11):1361–76.
NICEE, IIT; 2007. [16] Ariman T, Lee BJ. Tension/bending behavior of buried pipelines subjected to
[2] Malik JN, Nakata T. Active faults and related late quaternary deformation fault movements in earthquake. In: Proceedings of the 10th world conference
along the Northwestern Himalayan Frontal Zone, India. Annals of Geophysics on earthquake engineering. Madrid; 1992. p. 5423–26.
2003;46(5):917–36. [17] IITK-GSDMA. Guidelines for seismic design of buried pipelines. National
[3] McCalpin JP, Thakkar MG. Bhuj–Kachchh earthquake: surface faulting and its Information Center of Earthquake Engineering, Indian Institute of Technology
relation with neotectonics and regional structures, Gujarat, Western India. Kanpur (India), 2008.
Annals of Geophysics 2003 2001;46(5):937–56. [18] American Lifelines Alliance-ASCE. Guidelines for the design of buried steel
[4] Tsai JS, Jou LD, Lin SH. Damage to buried water supply pipelines in the Chi- pipe, July 2001 [with addenda through February 2005].
Chi (Taiwan) earthquake and a preliminary evaluation of seismic resistance [19] ASCE Technical Council on Lifeline Earthquake Engineering. Differential
of pipe joints. Journal of the Chinese Institute of Engineers 2000;23(4): ground movement effects on buried pipelines. guidelines for the seismic
395–408. design of oil and gas pipeline systems; 1984. p. 150–228.
[5] Liang J, Sun S. Site effects on seismic behavior of pipelines: a review. Journal [20] MCEER. Response of buried pipelines subject to earthquake effects. MCEER
of Pressure Vessel Technology 2000;122:469–75. Monograph no. 3; 1999.
[6] McCaffrey MA, O’Rourke TD. Buried pipeline response to reverse faulting [21] ABAQUS. ABAQUS user’s manual version 6.7. Simulia; 2007.
during the 1971 San Fernando earthquake. PVP, ASME 1983;77:151–9. [22] Joshi S.S. Analysis of buried pipelines subjected to reverse fault motion.
[7] Choo YW, Abdoun TH, O’ Rourke MJ, Ha D. Remediation for buried pipeline Mtech thesis, Indian Institute of Technology Kanpur, Kanpur, India; 2009.
systems under permanent ground deformation. Soil Dynamics and Earth- [23] Dash SR, Bhattacharya S, Blakeborough A. Bending–buckling interaction as a
quake Engineering 2007;27(12):1043–55. failure mechanism of piles in liquefiable soils. Soil Dynamics and Earthquake
[8] Kennedy RP, Williamson RA, Chow AM. Fault movement effects on buried oil Engineering 2010;30(1–2):32–9.
pipeline. Transportation Engineering Journal, ASCE 1977;103(5):617–33. [24] Dash SR, Govindaraju L, Bhattacharya S. A case study of damages of the
[9] Kennedy RP, Kincaid RH. Fault crossing design for buried gas and oil Kandla Port and customs office tower supported on a mat–pile foundation in
pipelines. PVP, ASME 1985;98-4:3–10. liquefied soils under the 2001 Bhuj earthquake. Soil Dynamics and Earth-
[10] Newmark NM, Hall WJ. Pipeline design to resist large fault displacement. In: quake Engineering 2009;29(2):333–46.
Proceedings of the US national conference on earthquake engineering. Ann [25] Abdoun TH, Ha D, O’Rourke MJ, Symans MD, O’Rourke TD, Palmer MC, et al.
Arbor: University of Michigan; 1975. p. 416–25. Factors influencing the behavior of buried pipelines subjected to earthquake
[11] Wang LRL, Yeh YA. A refined seismic analysis and design of buried pipeline faulting. Soil Dynamics and Earthquake Engineering 2009;29(3):415–27.
for fault movement. Earthquake Engineering and Structural Dynamics [26] American Petroleum Institute. API specification 5L. Specification for Line
1985;13(1):75–96. Pipe; 1990.
[12] Takada S, Hassani N, Fukuda K. A new proposal for simplified design of buried [27] Moser AP. Buried pipe design. 2nd ed. New York: McGraw-Hill; 2001.
steel pipes crossing active faults. Earthquake Engineering and Structural [28] Heedo Y, Kyriakides S. On the beam and shell modes of buckling of buried
Dynamics 2001;30(8):1243–57. pipelines. Soil Dynamics and Earthquake Engineering 1990;9(4):179–93.
[13] Karamitros DK, Bouckovalas GD, Kouretzis GP. Stress analysis of buried steel [29] Ariman T. Buckling and rupture failures of pipelines due to large ground
pipelines at strike-slip fault crossings. Soil Dynamics and Earthquake Engi- deformations. In: Proceedings of the 8th world conference on earthquake
neering 2007;27(3):200–11. engineering. San Francisco; 1984. p. 271–8.

You might also like