Download as pdf or txt
Download as pdf or txt
You are on page 1of 94

D 3.

4
DELIVERABLE

PROJECT INFORMATION

Systemic Seismic Vulnerability and Risk Analysis for


Project Title:
Buildings, Lifeline Networks and Infrastructures Safety Gain

Acronym: SYNER-G

Project N°: 244061

Call N°: FP7-ENV-2009-1

Project start: 01 November 2009

Duration: 36 months

DELIVERABLE INFORMATION

Deliverable Title: D3.4 – Fragility functions for gas and oil system networks

Date of issue: 31 October 2010

Work Package: WP3 – Fragility functions of elements at risk

Deliverable/Task Leader: Bureau de Recherches Geologiques et Minieres (BRGM)

Reviewer: University of Rome La Sapienza (UROMA)

REVISION: Final

Project Coordinator: Prof. Kyriazis Pitilakis


Institution: Aristotle University of Thessaloniki
e-mail: kpitilak@civil.auth.gr
fax: + 30 2310 995619
telephone: + 30 2310 995693
Abstract

In the frame of SYNER-G Work Package 3 – Fragility functions of elements at risk –


and Task 3.2 (Fragility functions for utility system networks), the present deliverable
aims at presenting fragility curves for components of gas and oil system networks
(Sub-Task 3.2.2). These fragility functions need to be applicable to the specific
European context and they are intended to be integrated to the general evaluation of
the systemic vulnerability. The following network components are considered: buried
pipelines, storage tanks and processing facilities (including compression stations,
pumping stations). Based on a literature review, it is found that the available fragility
functions are mostly empirical and should be applied to the European context, given
the current lack of data needed to validate potential analytical methods of
vulnerability assessment. For buried pipelines, fragility relations from (ALA, 2001)
are selected, with respect to both wave propagation and ground failure. The curves
from the HAZUS methodology (NIBS, 2004) may be used for storage tanks and
compression stations, as these facilities are decomposed into a fault-tree analysis.

Gas network, Oil network, Fragility curves, Empirical relations

i
Acknowledgments

The research leading to these results has received funding from the European
Community's Seventh Framework Programme [FP7/2007-2013] under grant
agreement n° 244061

iii
Deliverable Contributors

BRGM Pierre Gehl


Arnaud Reveillere
Nicolas Desramaut
Hormoz Modaressi
AUTH Kalliopi Kakderi
Sotiris Argyroudis
Kyriazis Pitilakis
Maria Alexoudi

v
Table of Contents

1 Intr oduction........................................................................................................................................ 1
2 Oil and gas systems in Eur ope .......................................................................................................... 1
2.1 NATURAL GAS SYSTEM ............................................................................................ 1
2.1.1 Production Facilities and Extra European supply sources............................... 2
2.1.2 Tank Farms...................................................................................................... 2
2.1.3 Natural Gas Pipelines ...................................................................................... 3
2.1.4 Stations............................................................................................................ 4
2.2 OIL SYSTEM ............................................................................................................... 4
2.2.1 Refineries......................................................................................................... 4
2.2.2 Oil Pipelines..................................................................................................... 5
2.2.3 Pumping Plants................................................................................................ 5
2.2.4 Tank Farms...................................................................................................... 5
3 Past ear thquake damages on system elements ................................................................................ 7
3.1 PHYSICAL DAMAGES / MAIN CAUSES OF DAMAGE .............................................. 7
3.1.1 Buried pipelines damages ............................................................................... 7
3.1.2 Storage tanks damages................................................................................... 8
3.1.3 Processing facilities damages ......................................................................... 9
3.2 CLASSIFICATION OF FAILURE MODES / DIRECT LOSSES ................................. 10
3.2.1 Pipeline failure modes ................................................................................... 10
3.2.2 Tanks failure modes ...................................................................................... 11
3.2.3 Support facilities failure modes ...................................................................... 13
4 Methodology for the vulner ability assessment of system elements ............................................. 15
4.1 IDENTIFICATION OF THE MAIN TYPOLOGIES ...................................................... 15
4.1.1 Pipelines ........................................................................................................ 15
4.1.2 Tank farms..................................................................................................... 17
4.1.3 Stations.......................................................................................................... 17
4.2 GENERAL DESCRIPTION OF EXISTING METHODOLOGIES................................ 20
4.2.1 Empirical relations ......................................................................................... 20
4.2.2 Bayesian approach ........................................................................................ 21
4.2.3 Analytical approach ....................................................................................... 21
4.2.4 Fault-tree analysis (for support facilities) ....................................................... 22

vii
4.3 DAMAGE STATES .................................................................................................... 24
4.3.1 Pipeline components ..................................................................................... 24
4.3.2 Storage tanks................................................................................................. 25
4.3.3 Processing facilities (pumping / compressor stations) ................................... 28
4.4 INTENSITY INDEXES ............................................................................................... 30
4.4.1 Pipeline components ..................................................................................... 30
4.4.2 Storage tanks................................................................................................. 33
4.4.3 Processing facilities ....................................................................................... 33
5 Fr agility functions for system elements ......................................................................................... 35
5.1 STATE-OF-THE-ART FRAGILITY CURVES PER COMPONENT ............................ 35
5.1.1 Pipeline components ..................................................................................... 35
5.1.2 Storage tanks................................................................................................. 47
5.1.3 Processing facilities ....................................................................................... 54
5.2 VALIDATION / ADAPTATION / IMPROVEMENT ...................................................... 57
5.2.1 Pipeline components ..................................................................................... 57
5.2.2 Storage tanks................................................................................................. 58
5.2.3 Processing facilities ....................................................................................... 58
5.3 FINAL PROPOSAL .................................................................................................... 58
5.3.1 Pipeline components ..................................................................................... 58
5.3.2 Storage tanks................................................................................................. 63
5.3.3 Processing facilities ....................................................................................... 65
6 Analytical expr essions of fr agility functions.................................................................................. 69
6.1 PIPELINE COMPONENTS ........................................................................................ 69
6.1.1 Wave propagation.......................................................................................... 69
6.1.2 Permanent ground deformation ..................................................................... 70
6.2 STORAGE TANKS .................................................................................................... 71
6.3 PROCESSING FACILITIES ....................................................................................... 72
6.3.1 Pumping / compressor stations ..................................................................... 72
6.3.2 Re.Mi cabins .................................................................................................. 73
6.3.3 GRF reduction groups ................................................................................... 73

viii
List of Figures

Figure 1 Schema of a gas system. ........................................................................................ 1


Figure 2: Overview. Gas in Europe (Nies, 2008). ................................................................... 2
Figure 3 Pipeline damage in (a) perpendicular and (b) parallel crossings of a lateral spread
(Rauch, 1997)...................................................................................................... 8
Figure 4 Continuous steel pipeline (source: FULCRUM DYNAMICS, LLC) ........................ 10
Figure 5 Segmented concrete pipeline and rubber gasket joint (Kim et al., 2010) .............. 11
Figure 6 Common failure modes for steel tanks (a) sloshing damage to the upper shell (b)
pipe connection failure (c) elephant's foot buckling (Berahman & Behnamfar,
2007) ................................................................................................................. 12
Figure 7 RE.MI cabin in the L'Aquila area: outside view (courtesy of Enel Rete Gas) ........ 19
Figure 8 RE.MI cabin in the L'Aquila area: inside view (courtesy of Enel rete Gas) ............ 19
Figure 9 View of a reduction group as used in the L'Aquila area (courtesy of Enel Rete Gas)
........................................................................................................................... 20
Figure 10 Decomposition of a compressor station into a fault-tree...................................... 23
Figure 11 Example of a fault-tree for an anchored steel tank failure modes (ALA, 2001) ... 27
Figure 12 Empirical correlation between PGS (strain) anf maximum horizontal PGV
(Paolucci & Pitilakis, 2007) ................................................................................ 31
Figure 13 Strain vs Repair rate correlation, for both wave propagation and permanent
ground displacement (O’Rourke & Deyoe, 2004).............................................. 32
Figure 14 Pipeline fragility data of Katayama et al. (1975) as presented by O'Rourke & Liu
(1999) ................................................................................................................ 37
Figure 15 Bilinear pipeline fragility relations by (Eguchi, 1991) ........................................... 38
Figure 16 Fragility relations of Barenberg (1988) and O'Rourke & Ayala (1993)................. 39
Figure 17 Fragility function by Eidinger et al. (1995, 1998), for all data............................... 40
Figure 18 Fragility relations by Isoyama et al. (2000), for both PGA (a) and PGV (b)
parameters, without corrective factors............................................................... 42
Figure 19 "Backbone curve" proposed by ALA (2001), representing the median repair rate
of all data points, and the corresponding 16th and 84th quantiles. ..................... 43
Figure 20 Fragility relations proposed by (Eguchi, 1983) for fault ruptures ......................... 46
Figure 21 Fragility curves of Berahman & Behnamfar (2007), for DS2................................ 49
Figure 22 Fragility curves of Berahman & Behnamfar (2007), for DS3................................ 50
Figure 23 Fragility curves of Berahman & Behnamfar (2007), for DS4................................ 50

ix
Figure 24 Fitted surface for fragility median µ...................................................................... 51
Figure 25 Fitted surface for fragility standard deviation .................................................... 52
Figure 26 Fault-tree analysis proposed by HAZUS (NIBS, 2004) to assess the vulnerabiltiy
of tank farms...................................................................................................... 53
Figure 27 Fragility curves for gas pumping / compression stations (anchored components)
proposed by Risk-UE (Alexoudi & Pitilakis, 2003)............................................. 56
Figure 28 Fragility curves for gas pumping / compression stations (unanchored
components) proposed by Risk-UE (Alexoudi & Pitilakis, 2003) ....................... 57
Figure 29 Comparison of the pipeline fragility relations for PGV. Arrows refer to the range of
applicability of a given relation, approximated from knowledge of the dataset
from which it wa derived (Tromans, 2004) ........................................................ 59
Figure 30 Proposed fragility curves for the most common gas & oil pipeline typologies (ALA,
2001), for wave propagation.............................................................................. 61
Figure 31 Comparison of three fragility curves, with respect to PGD .................................. 62
Figure 32 Proposed fragility curves for the most common gas & oil pipeline typologies (ALA,
2001), for permanent ground deformation......................................................... 63
Figure 33 Fragility curves for steel tank farms (NIBS, 2004) ............................................... 64
Figure 34 Fragility curves for Greek pumping / compressor plants (SRMLIFE, 2003-2007) 65
Figure 35 Fragility curves for generic pumping / compressor plants (NIBS, 2004).............. 66
Figure 36 Fault-tree analysis of a Re.Mi cabin according to (Esposito et al., 2011) ............ 67
Figure 37 Fault-tree analysis of a Reduction Group (GR / GRM) according to (Esposito et
al., 2011)............................................................................................................ 68

x
List of Tables

Table 1 Two main types of sollicitations ................................................................................. 7


Table 2 Main features of the studied pipeline networks........................................................ 16
Table 3 Proposed damage states for pipeline components .................................................. 25
Table 4 Damage states defined in HAZUS (NIBS, 2004) and (O’Roure & So, 2000).......... 25
Table 5 Correlation between damage modes and repair costs (ALA, 2001) ....................... 26
Table 6 Damage states adapted from (ALA, 2001) ............................................................. 27
Table 7 Damage states defined by (Kappos et al., 2006) for buildings ............................... 28
Table 8 Damage scale proposed by (LESSLOSS, 2007) and (SRMLIFE, 2003 - 2007) for
pumping / compressor stations.......................................................................... 28
Table 9 Description of damage states and functionality indicators for gas stations
(SRMLIFE, 2003-2007) ..................................................................................... 29
Table 10 Maximum longitudinal strains induced by seismic waves propagation along a
pipeline (St John & Zahrah, 1987)..................................................................... 30
Table 11 Summary of the fragility functions from the literature............................................ 35
Table 12 Values of corrective factor K1, according to Eidinger et al. (1995, 1998).............. 40
Table 13 Values of corrective factors according to (Isoyama et al., 2000). Bracketed values
are less reliable due to small sample size. ........................................................ 41
Table 14 Values of corrective factor K1, according to (ALA, 2001) ..................................... 44
Table 15 Summary of the past studies on fragility of pipeline components subjected to
permanent ground deformation ......................................................................... 45
Table 16 Values of the corrective factor K2 according to (ALA, 2001) ................................. 46
Table 17 Proposed vulnerability indices by Ballantyne, for various pipe materials and joint
types (B&S: bell and spigot, RG: rubber gasket, R: restrained, UR: unrestrained)
........................................................................................................................... 47
Table 18 Median and standard deviation parameters for the fragility curves proposed by
O'Rourke & So (2000) ....................................................................................... 48
Table 19 Median and standard deviation parameters for the fragility curves proposed by
ALA (2001) ........................................................................................................ 49
Table 20 Fragility curves of components (anchored or unanchored) of tank farms
components, according to HAZUS (NIBS, 2004) .............................................. 53
Table 21 Fragility parameters (lognormal distribution) for steel tank farms, according to
HAZUS (NIBS, 2004)......................................................................................... 54
Table 22 Median and standard deviation parameters of the fragility curves for pumping
plants (LESSLOSS, 2007)................................................................................. 54

xi
Table 23 Parameters of fragility curves for stations components taken from (NIBS, 2004). 55
Table 24 Median and standard deviation parameters of the fragility curves (lognormal
distribution) for pumping plants (SRMLIFE, 2003-2007) ................................... 55
Table 25 Median and standard deviation parameters of the fragility curves (lognormal
distribution) for pumping plants, proposed by the HAZUS methodology (NIBS,
2004) ................................................................................................................. 56
Table 26 Pipe material and pipe diameter categories included in the dataset for the (ALA,
2001) fragility relation, according to (Tromans, 2004) ....................................... 59
Table 27 Fragility parameters for steel tank farms, according to HAZUS (NIBS, 2004) ...... 64
Table 28 Fragility parameters for Greek pumping plants, according to (SRM-LIFE, 2003-
2007) ................................................................................................................. 65
Table 29 Fragility parameters of the fragility curves for pumping plants, proposed by the
HAZUS methodology (NIBS, 2004) ................................................................... 66
Table 30 Damage states for the Re.Mi cabin....................................................................... 67
Table 31 Damage states of the Reduction Group (GR / GRM) ........................................... 68
Table 32 Values of corrective factor K1 (ALA, 2001)............................................................ 69
Table 33 Values of corrective factor K2 (ALA, 2001)............................................................ 70
Table 34 Fragility parameters for steel tank farms (HAZUS, 2004) ..................................... 71
Table 35 Damage states definitions for tank farms (HAZUS, 2004) .................................... 71
Table 36 Fragility parameters for pumping / compressor stations (HAZUS, 2004) and
(SRMLIFE, 2003-2007) ..................................................................................... 72
Table 37 Damage states definitions for pumping / compressor stations (HAZUS, 2004) and
(SRMLIFE, 2003-2007) ..................................................................................... 73

xii
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

1 Introduction

The SYNER-G project intends to quantify the impact of the disruption of various urban
systems (building areas, health-care centers, transportation networks, utilities) after an
earthquake event. Among the studied utility systems, the gas & oil network has proven to be
prone to some damages during past earthquakes, often with significant consequences:
pollution of waterways, triggering of fires, disruption of gas services.
In the frame of Work Package 3 – Fragility functions of elements at risk – and Task 3.2
(Fragility functions for utility system networks), the present deliverable aims at presenting
fragility curves for components of gas and oil system networks. These fragility functions
need to be applicable to the specific European context and they are intended to be
integrated into the general evaluation of the systemic vulnerability.
The present report reviews at first the different damages that occurred to gas and oil network
components in the past seismic events, along with the most common damage mechanisms.
The following components are proposed to be studied in the scope of SYNER-G:
- pipeline components, which can be damaged by both wave propagation and
permanent ground deformation;
- atmospheric storage tanks;
- processing facilities (e.g. gas pumping / compressor stations);
Then, the second part deals with the description of the European typologies of each of the
gas and oil network components, with an emphasis on the elements located in the selected
case studies of Thessaloniki, Vienna and the L’Aquila area. A review of existing
methodologies (empirical relations, numerical models, fault-tree analyses…) is followed by
the definition, for each component, of some key parameters:
- a damage scale;
- an intensity index (e.g. the selected intensity measure);
- performance indicators that can help to specify what is the link between the damage
state of the component and its serviceability / functionality;
Finally, based on a review of state-of-the-art fragility curves for each component, we
examine whether there is a need for further development or not. For those elements that
have already suitable fragility curves, a discussion is needed to select the most appropriate
fragility function.
At last, the report ends with a short section that specifies the following points, needed for the
coding of the fragility curves into the software tool:
- typology classification of each component;
- damage scale definition;
- intensity index used;
- fragility curve parameters, for each damage state and each typology.

1
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

2 Oil and gas systems in Europe

2.1 NATURAL GAS SYSTEM

Natural Gas System transfers gas. Gas is composed mainly of methane (CH4) and other
gases in smaller percentages. Five different types of Natural Gas exist: Dry, Liquefied
Natural Gas (LNG), Sour, Sweet and Wet.
Gas system consists of Figure 1:
- Production wells, offshore platforms and gathering facilities
- Terminals
- Compressor station or pressure reduction station
- Storage tanks
- Transmission and distribution pipelines
- Communication and control facilities
- Isolation valves
- Maintenance support facilities (these include maintenance centres and facilities for the
storage of the spare equipment and parts)

Figure 1 Schema of a gas system.

1
Natural Gas System

2.1.1 Production Facilities and Extra European supply sources

Production Facilities serve as the source of the gas supply and consist either of onshore
facilities (production field) or offshore platforms.
Gas supply in Europe essentially comes from four sources outside of domestic production;
production within the EU accounts for around a third, and imports come from the following
four countries: Russia (46% of imports), Norway (27%), and Algeria (20%), and to a lesser
extent Nigeria (less than 8%). Proportions of supply sources vary from member state to
member state for obvious geographic reasons. The dominance of Algerian gas in the mix of
Mediterranean states (Italy, France, and also Portugal) contrasts with Russia’s dominance in
Central Europe, notably in the new member states and Germany. The rest comes from
internal production, which rose to 33% in 2005 (Figure 2).

Figure 2: Overview. Gas in Europe (Nies, 2008).

2.1.2 Tank Farms

Tank farms are facilities that store fuel products (it is suitable for oil system). They include
tanks, pipes and electric components. There are two types of storage facilities: Underground
storage facilities and Storage Tanks.

Underground Storage Facilities

The use of sub-surface facilities for storing gas usually is used to balance seasonal
variations in demand. They may be classified as:

2
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

- seasonal supply reservoirs, designed to be filled during the 214 day non-heating season
(mostly gas/oil fields and aquifers);
- high-deliverability sites for 151-day heating season (mostly salt cavern reservoirs).
These facilities are located hundred meters below the surface. They are usually natural
geological reservoirs, such as depleted oil or gas fields or water-bearing sands on the top
and impermeable cap rock.

Above ground Storage Tank

LNG tanks are used to store Liquefied Natural Gas. They differ from Water/ Wastewater
tanks as they are designed to minimise any heat ingress. The insulation of the tanks will not
keep the temperature of LNG cold by itself. LNG will stay at near constant temperature if
kept at constant pressure. As long as the steam (LNG vapour boil off) is allowed to leave the
tank, in a safe and controlled manner, the temperature will remain constant. This
vaporisation loss is collected from the tank and either reabsorbed as a liquid, sent to the gas
output line connecting to the national gas grid, or used as fuel on the site. The LNG tanks
would be of a full containment design. In a full containment system two tanks are employed,
an inner tank which contains the stored liquid, and an outer tank which provides security in
the event of any loss of containment or leak from the inner tank. Sophisticated automatic
protection systems are employed to monitor the tank levels, pressures, temperatures and
any potential leakage from the inner tank.

2.1.3 Natural Gas Pipelines

Natural gas networks are operating at different pressures: supra-regional transmission


pipelines operate at very high pressures. These pipelines have a maximum diameter of
1.40 m and are operating at pressure higher than 100 bars. Such gas pipelines can cover
distances of up to 6 000 km (e.g. from west Siberia to Europe).
Supra-regional transmission pipelines are then separated to several branches of high-
pressure pipelines (<70 bars). The high-pressure pipelines distribute natural gas to several
regions. Distribution pipelines then are used to serve the needs of communities. The
pressure for these regional networks range between from 1 to 70 bars, while local
distribution systems are usually operates in the medium (0.1 - 4 bars) or low-pressure
(<0.1 bas) range.
Offshore pipelines have smaller diameters (maximum 1.05 m) although they are designed
for much higher pressures (up to 200 bar).

3
Natural Gas System

2.1.4 Stations

Compressor Stations

Compressor station is a facility, which supplies gas with energy to move in transmission
lines. Otherwise compressor stations are operated at underground storage facilities to raise
the pressure of the gas injected into storage or to compress the natural gas as it leaves
storage to be fed into the pipeline.
The distance between compressor stations along a transmission trunk line is usually
between 100 and 250 km. Each station contains one or more centrifugal or reciprocating
compressor units, and auxiliary equipment for purposes such as generating electricity,
cooling discharge gas and SCADA system that controls the station with all the equipments.

Metering/ Pressure Reduction or compression Stations

Metering stations are used for controlling the flow of gas and its quality in the pipelines
network.
Pressure reduction or compression stations are used in order to set the gas pressure at the
required level for its industrial or commercial use. In each station is using SCADA systems.

LNG Terminal Stations

The terminal stations include moorings for tankers, used for the transportation of liquefied
natural gas, storage tanks for the liquefied gas and re-liquefaction installations. Piping
system in the station or in terminal stations should be designed in order to prevent rupture
during the worst case earthquake. No specific guide exists.

2.2 OIL SYSTEM

An oil system typically consists of refineries, pumping plants, tank farms, and pipelines.

2.2.1 Refineries

Refineries are an important part of an oil system. They are used for processing crude oil
before it can be used. Although supply of water is critical to the functioning of refinery, it is
assumed in the methodology that an uninterrupted supply of water is available to the
refinery.

4
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

2.2.2 Oil Pipelines

Oil pipelines are used for the transportation of oil over long distances. A large segment of
industry and millions of people could be severely affected by disruption of crude oil supplies.
Rupture of crude oil pipelines could lead to pollution of land and rivers. Pipelines are typically
made of mild steel with submerged arc welded joints, although older gas welded steel pipe
may be present in some systems.

2.2.3 Pumping Plants

Pumping plants serve to maintain the flow of oil in cross-country pipelines. Pumping plants
usually use two or more pumps. Pumps can be of either centrifugal or reciprocating type.
However, no differentiation is made between these two types of pumps in the analysis of oil
systems. Pumping plants are classified as having either anchored or unanchored
subcomponents.

2.2.4 Tank Farms

Tank farms are facilities that store fuel products. They include tanks, pipes and electric
components. Tank farms are classified as having either anchored or unanchored
subcomponents.

5
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

3 Past earthquake damages on system


elements

3.1 PHYSICAL DAMAGES / MAIN CAUSES OF DAMAGE

3.1.1 Buried pipelines damages

Like many other underground components, buried pipelines are very sensitive to permanent
ground deformation (resulting from various ground failures), in addition to transient ground
deformation due to seismic wave propagation: the characteristics of these physical
phenomena are summed up in Table 1.

Table 1 Two main types of sollicitations

Ground failure Transient ground


deformation
Hazard surface faulting, liquefaction, R-waves, S-waves
landslides
Usual descriptor PGD PGV, PGA, strain
Spatial impact local and very site-specific large and distributed

The first sign of damages to buried pipelines is the 1906 San Francisco earthquake, which
resulted in significant fires through the city, due to the rupture of water lines needed by fire-
hydrants. Regarding the causes of damage, according to O’Rourke & Liu (1999), the zones
of lateral spreading accounted for only 5% of the built-up area affected by strong ground
shaking, yet approximately 50% of all pipeline breaks occurred within one city block of these
zones: this fact demonstrate the high impact of ground failure on pipelines damage.
Indeed, according to Eguchi (1987), past earthquakes have caused significant damages to
underground pipelines throughout the world: yet inertia forces are not the main issue for
buried components, whereas faulting, landslides or liquefaction pose the most problems
(Hall, 1987).
(O’Rourke & Ayala, 1990) report that a few earthquakes have induced damages to pipelines
only by the effect of seismic wave propagation:
- 1985 Michoacan earthquake, which damaged a large corrosion-free modern
continuous steel pipeline;
- 1964 Puget Sound, 1969 Santa Rosa, 1983 Coalinga and 1989 Loma Prieta
earthquakes;
Yet, in most cases, it appears that seismic wave propagation damaged mainly pipelines that
were previously weakened either by corrosion or welds of poor quality (EERI, 1986).

7
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

On the contrary, there are numerous examples of damages induced by permanent ground
deformation: 1906 San Francisco, 1952 Kern County, 1964 Niigata, 1964 Alaska, 1971 San
Fernando, 1978 Miyagi-ken-oki and 1983 Nihonkai-Chuba earthquakes. During the 1971
San Fernando earthquake, the steel pipeline system withstood significant ground shaking,
yet it was damaged by abrupt vertical or lateral dislocations or ground ruptures: lateral
spreading (Figure 3) induced severe damages during that earthquake (EERI, 1986,
O’Rourke & Trautmann, 1981, O’Rourke, 1988), and one of the most severe damage was
observed in a in pipeline that was deformed by a differential lateral movement up to 1.7 m.
Regarding liquefaction, a good example is the 1964 Niigata earthquake, where the average
failure ratio for a pipeline system was as high as 0.97 per km, with all kinds of failure types
(pipe body breaks, weld breaks, joint separations).

Figure 3 Pipeline damage in (a) perpendicular and (b) parallel crossings of a lateral
spread (Rauch, 1997)

More recent events, like the 1994 Northridge, 1995 Kobe, 1999 Kocaeli or 1999 Chi-Chi
earthquakes, confirmed the relative vulnerability of piping systems to strong ground motions
and the somewhat good performance of welded-steel pipes with respect to seismic wave
propagation.
As a result, the emphasis is put on the ductility of pipes and the quality of welds, when
building earthquake resistant piping systems: still, pipe welds or joints seem to be the most
vulnerable parts of this component. Also, corrosion is an aggravating factor of the pipeline
vulnerability (Young & Pardon, 1983, Ogawa, 1983).

3.1.2 Storage tanks damages

According to (EERI, 1986), damage to tanks is quite common throughout past earthquakes:
- 1952 Kern County earthquake: damage occurred near the top, due to sloshing oil,
with more severe damages on floating roof tanks (ASCE, 1987);
- 1964 Niigata earthquake: oil leaked from ruptured tanks and caught fire, damaging
two refineries (ASCE, 1984);
- 1971 San Fernando earthquake: many incidents with storage tanks were reported,
most of them were half-full to full;

8
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

- 1975 Imperial Valley earthquake: out of six unanchored full tanks in the area, the
largest one was damaged and the failure of the fixed roof allowed for the sloshing oil
to spill (EERI, 1986);
- 1978 Miyagi-ken-oki earthquake: a large oil refinery with 90 tanks had three of them
failed and three others damaged without failure. Tank failure induced oil spillage and
pollution of waterways (ASCE, 1984);
- 1985 Chile earthquake: a refinery with 120 storage tanks had 12 full or nearly full
tanks damaged, with elephant’s foot buckling, roof damage, base plate failure, or
damage to the exiting piping (due to rocking of the tanks and differential settlement)
(EERI, 1986);
- 1989 Loma Prieta earthquake: tank damages could be observed up to 120 km from
the epicentre location. Unanchored tanks underwent an uplift of the walls (up to 200
mm displacement between the foundation and the shell), which resulted in elephant’s
foot buckling, vertical splits in tank walls, ruptures of elephant’s foot buckles,
puncture of tanks by restrained pipes, and damage to restrained pipes anchored to
both tank and foundation (EERI, 1990).
As a result, most tanks show a significant vulnerability to ground shaking, and due to the
potential gravity of indirect damages (fire risk, liquid spillage and pollution…), the seismic
vulnerability assessment of such component should be looked at thoroughly, in the scope of
a systemic vulnerability analysis.
All these damage reports from past earthquakes seem to indicate that unanchored tanks
seemed the most vulnerable, along with vertical cylinders tanks with a large height-to-
diameter ratio (EERI, 1990). Finally, the amount of liquid stored within the tank has a
significant impact, as full tanks are subject to larger lateral forces and overturning moments
due to liquid sloshing, which can also damage the tank roof. Other damage mechanisms
include elephant’s foot buckling (generated by compressive forces against the tank wall) and
leakage or breakage of pipe connections (due to tank rocking and sliding).

3.1.3 Processing facilities damages

Reports from past earthquakes show a little information on above ground facilities, yet the
small number of incidents associated to these components tends to indicate a good
behaviour of these support facilities during an earthquake:
- 1971 San Fernando earthquake: underground facilities buried in vaults or manholes
did not suffer from any structural damage (ASCE, 1974);
- 1985 Chile earthquake: some industrial facilities endured minor damage, yet they
were still functional and there was no shut-down (EERI, 1986);
- 1987 Ecuador earthquake: a pumping station was damaged, with a jammed control
valve and the outage of electrical power and back-up generators (Crespo et al.,
1987);
- 1989 Loma Prieta earthquake: no serious damage was reported for industrial
facilities
Thus, this limited experience shows that modern facilities (compressor stations, pumping
stations, control stations) that are built according to seismic codes with anchored equipment

9
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

exhibit good resistance to ground shaking. (Bettinger, 1980). The anchorage of


subcomponents is especially a crucial point, as unanchored equipment can lead to the
rupture of electrical connections (Nyman, 1987) or the tipping and sliding of mechanical
parts.

3.2 CLASSIFICATION OF FAILURE MODES / DIRECT LOSSES

3.2.1 Pipeline failure modes

Continuous pipelines (e.g. welded-steel pipes, Figure 4) are built with rigid joints (welds) and
they have a tendency to show good performance during past earthquakes. They usually fail
due to compression strains, that induce buckling of the pipe body (like a beam), or warping
and wrinkling of the pipe wall (ALA, 2001). This deformation may not generate leakage, yet
the modification of the pipe shape may produce disruption of the gas / oil flow. A crucial
factor for the resistance of continuous pipelines is the quality of the welds, as past studies
have shown that pipes constructed before the 1930s with poor quality welds experienced
damages mostly at the joint locations.

Figure 4 Continuous steel pipeline (source: FULCRUM DYNAMICS, LLC)

Segmented or jointed pipelines usually consist of rigid pipe segments (e.g. cast-iron or
concrete, Figure 5) connected through loose or flexible joints. Three main failure modes
have been identified for this typology (ALA, 2001): tensile and bending deformations of the
pipe barrel, excessive rotation of a joint, and pullout of a joint (Singhal, 1984). This pipeline
type is much less frequent in oil / gas piping networks.

10
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

Figure 5 Segmented concrete pipeline and rubber gasket joint (Kim et al., 2010)

Aside from these usual failure modes, a piping system is more vulnerable at discontinuities
like pipe elbows, tees, in-lines valves or connections to adjacent structures (storage tanks,
racks, facilities…): high stresses are especially concentrating at these anchor points and
rigid locations (ALA, 2001). Also, corrosion has a strong influence on the pipes vulnerability,
as it decreases the wall thickness and creates heterogeneous zones that may lead to stress
concentrations.
It is very difficult to translate the described damage mechanisms in terms of direct losses
(intensity of leakage, breakage): indeed, damages such as joint failure or pipe buckling may
result in various outcomes, from the slight leakage to the complete spillage of the content.
Empirical data on pipeline failures from past earthquakes has been developed based on a
simple indicator: the number of repairs per unit length performed by the exploiting company.
Thus, no distinction is made between the types of repairs: those can include the complete
fracture of the pipe, a leak in the pipe, or damage to an appurtenance of the pipe (ALA,
2001). All this collected data is unfortunately not differentiated and these different types of
repairs have of course various impacts at the scale of the piping system evaluation. The
more recent post-earthquake review of damages carried out after the L’Aquila event also
makes no distinction between the different failure types.

3.2.2 Tanks failure modes

According to NZNSEE (1986), Kennedy & Kassawara (1989) and ALA (2001), the following
classification of failure modes is proposed:
- shell buckling: it is one of the most common forms of damage in steel tanks, and it is
expressed via an outward buckling of the bottom shell courses (“elephant foot”,
Figure 6c), that can sometimes occur over the full circumference of the tank. This
phenomenon may lead to the loss of the content due to weld or piping fractures, and
less frequently to the total collapse of the tank.
- roof damage (e.g. Figure 6a) : ground shaking may induce oil sloshing inside the
tank. When tanks are full or nearly full, this sloshing motion generates an upward
pressure distribution against the tank roof. This may cause a rupture of the joints
between the wail and the roof, leading to a spillage of tank contents over the tank
walls. Observations from past earthquakes show that floating roofs have generally

11
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

endured more severe damage than fixed steel roofs. It is yet important to note that
roof damage, although expensive to repair, usually does not lead to more than a third
of total content loss (ALA, 2001).
- anchorage failure: many tanks are anchored with steel braces or bolts, but it is still
possible that these anchors may be pulled out or stretched by the seismic load.
However, the failure of anchoring components does not necessary imply the loss of
the tank contents.
- tank support system failure: this failure mode is specific for above-grade tanks,
elevated by steel columns or frames. Even if the failure of the supporting system
often leads to complete loss of contents, this issue is of lesser concern to large oil
storage tanks, which are usually built at grade.
- foundation failure: this phenomenon can be common in the case of poor foundation
conditions prone to liquefaction, resulting in base rotation and important settlements.
In the case of unanchored tanks, tensile stress can also generate uplift displacement
of the tank base, separating it from the baseplate.
- hydrodynamic pressure failure: ground shaking generates pressures between the
fluid and the tank walls, thus resulting in tensile hoop stresses. The induced loads
may then lead to splitting of the wall and leakage, especially in the case of steel
tanks with riveted joints.
- connecting pipe failure (e.g. Figure 6b): this is one the most common failure modes
that can induce a total loss of the tank contents. The fracture of the pipes at the
connections to the tank results from differential displacement between the piping and
the tank (uplift displacements, foundation failure).
- manhole failure: because of significant stresses against the manhole cover, the latter
can fail which results in loss of content through the opening.

( a )

( b )

( c )

Figure 6 Common failure modes for steel tanks (a) sloshing damage to the upper
shell (b) pipe connection failure (c) elephant's foot buckling (Berahman & Behnamfar,
2007)

The variety of failure mechanisms makes it very difficult to assess the direct losses. One
approach is to convert everything in terms of economic losses, based of the replacement
value of the damaged tank or the cost of repairs. Yet, these direct economic losses are not
well correlated to the functionality of a tank: for example, elephant’s foot buckling might
deform the tank wall (inducing heavy repair cost) without generating any loss of contents,

12
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

whereas a pipe connection failure (which represents a small fraction of the tank value) can
result in the total spillage of contents and the loss of the tank function.

3.2.3 Support facilities failure modes

Because of the little information available on support facilities damages, which is a


consequence of their good behaviour in past earthquakes, main failure modes are not well
identified. However, a fault-tree analysis of a gas compressor station or a pumping station
can help identifying the following key components:
- building: the collapse of the structure sheltering the facility may damage the
equipment with falling debris;
- pump / compressor: this key element is connected to the piping system and its
failure, due to sliding or rocking if unanchored, can generate leakage or breakage of
the pipe;
- electrical / mechanical components: these miscellaneous components, necessary for
the compressor to operate, can also be damaged if not anchored;
- electric power supply: external power can be shut down because of the electric
power network disruption, or the connection failure between the power lines and the
facility building. However, most facilities are equipped with backup power generators.
Again, these different causes of damage are very disparate in terms of direct economic
losses, but the smallest component failure can have the same impact as the building
collapse, in terms of the facility functionality.
Regarding in-line valves, many types are found along the piping network (gate valves,
butterfly valves, check valves, ball valves…) and they can be either buried with the pipeline
or located in underground concrete vaults. Following the recommendations of (ALA, 2001), it
is proposed to treat pipeline valves in the same framework as pipe components. In other
words, they are affected only if the pipe segment they are fixed to is affected. It has to be
noted that this is different from not considering the components at all, as they require the
functionality of other networks (e.g. power supply and communication links) to assume their
role in the system. Also, their functionality is related to the connection with the pipeline: if the
pipe-valve joint fails, the flow continuity id not ensured (break).
Finally, SCADA equipment includes many components (instrumentation, power supply,
communication components, vaults). For hardware located in metal cabinets, the main
observed damages comprise batteries falling over, circuit boards dislodging and gross
movement of the cabinet enclosure (ALA, 2001). Regarding pressure / flow measuring
instruments, ground shaking is likely to induce air bubbles that can provoke false reading.

13
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

4 Methodology for the vulnerability assessment


of system elements

4.1 IDENTIFICATION OF THE MAIN TYPOLOGIES

4.1.1 Pipelines

Natural gas networks are operating at various pressures, depending on their scale:
- supra-regional transmission pipelines: these pipelines operate at very high pressures
(~100 bar) and present large diameters (up to 1.40 m). Such pipelines can cover large
area (e.g. from west Siberia to Europe, from Norway to France…);
- regional transmission/distribution pipelines: these pipes still operate at high pressure
(from 1 to 70 bar) and are used to connect local distribution systems;
- local distribution pipelines: these smaller pipelines usually operate in the medium (0.1
– 4 bar) or low-pressure (< 0.1 bar) range;
In addition to this classification, the pipeline typologies mainly rely on the following
parameters:
- material type;
- material strength;
- diameter;
- wall thickness;
- smoothness of coating;
- type of connection;
- design flow;
Focusing mainly on the typologies inherent to the three cases studies (Thessaloniki, Vienna
and L’Aquila), pipeline components from Greece present the following characteristics:
- transmission pipelines (19 bar): welded-steel, diameters ranging between 100 –
250 mm and wall thickness from 4.37 mm to 5.56 mm;
- distribution pipelines (4 bar): made of PVC (with electro-fusion connections), with
diameters between 125 and 160 mm and wall thickness ranging from 11.4 mm to 14.6
mm;
These natural gas pipelines are located at a conventional depth, 1.10 m + pipeline diameter
+ 0.15 m.
In Austria, there are several long distance transmission pipelines going through (TAG,
WAG, HAG…). They consist of welded-steel and have diameters ranging from 200 to 1,400
mm. They are operated at 84 bar and are buried at an average depth of 1 m.

15
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

The regional transmission/distribution pipelines operate at a pressure of about 16 bar, and


get down to 1 bar locally: these pipelines are made of PVC.
In L’Aquila, the transmission network (operated by SNAM at a national level) is made of
welded –steel pipes, with an internal diameter of 103.9 mm and wall thickness of 5 mm. The
transmission and delivery pressure for the L’Aquila area is 64 bar. Locally, the gas is
distributed via a 621 km pipeline network: 234 km of pipes operating at medium pressure
(2.5 – 3 bar), and the remaining 387 km with gas flowing at low pressure (0.025 – 0.035 bar):
these pipelines are either made of steel or HDPE (High Density Polyethylene). HDPE pipes
have a nominal diameter ranging from 32 to 400 mm, whereas diameter of steel pipes is
usually between 25 and 300 mm.
As a result, it is reasonable to find the pipelines typologies based on the following known
features:
- material type: welded-steel, PVC or HDPE;
- operating pressure;
- pipe diameter;
- connection type (if known);
On Table 2, the main characteristics of the studied pipelines for each area are summed up.

Table 2 Main features of the studied pipeline networks

Type Pressure Material Diameter (mm) Thickness (mm)


Greece transmission 19 bar welded-steel 100 – 250 4.37 – 5.56
distribution 4 bar PVC 125 – 160 11.4 – 14.6
Austria supra- 84 bar welded-steel 200 – 1,400 -
regional
transmission
transmission 16 bar PVC - -
distribution 1 bar PVC - -
L’Aquila transmission 64 bar welded-steel 103.9 -
distribution 2.5 – 3 bar welded-steel 25 – 300 / 3.2 – 5.6 /
/ HDPE 32 – 400 2.3 – 36.4
distribution 0.025 –
(local) 0.035 bar

Regarding oil pipelines, these components are less frequent in the studied areas and they
may present features similar to supra-regional gas transmission pipelines: indeed, most oil
pipelines crossing Europe are made of welded-steel with arc-welded joints, with large
diameters (250 – 1,000 mm), and operate at high pressure/flow.

16
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

4.1.2 Tank farms

Natural gas

Aside from underground storage facilities, natural gas is usually stored while in its liquefied
state (LNG) in specific LNG tanks: these facilities are designed to insulate the gas from any
heat ingress, using “auto-refrigeration” techniques (Vaporization of part of the stored
liquefied gas counteracts the unavoidable heat flow coming from the outside, keeping the
two phases system at the equilibrium, which implies a constant temperature for a given
pressure). This technique requires an inner tank (which contains the stored liquid) and an
outer tank (which provides security in the event of any loss of containment from the inner
tank).
Inner shells are usually made of a nickel-steel alloy, whereas the outer shell is a pre-
stressed concrete construction. The LNG tanks are usually vertical cylinders and can
represent huge facilities, which make them too specific to study in the scope of this project.

Oil and fuel

Liquid products (oil and fuel) are stored in atmospheric storage facilities, which include tanks
(vertical cylinders), pipes and electric components. The tank typologies are usually classified
according to the following characteristics:
- material: steel or reinforced-concrete;
- construction type: at-grade or elevated;
- anchored or unanchored;
- roof type;
- capacity;
- shape factor: height vs diameter ratio;
- amount of content in the tank: empty, half-full, full;
There is a wide variety of storage tanks in Europe, and main typologies should be extracted,
mainly based on the material and construction types, and whether components are anchored
or not.

4.1.3 Stations

Compressor stations

This type of station is a facility which supplies gas with energy to move along the
transmission lines. Compressor stations are also used in storage facilities to compress the
gas when it is fed into the pipeline. The distance between compressor stations along a
transmission trunk line is usually between 100 and 250 km. Each station contains one or
more centrifugal or reciprocating compressor units, and auxiliary equipment for purposes

17
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

such as generating electricity or cooling discharge gas and SCADA system that controls the
station with all the equipments. Two or more compressors at a station can be used either in
parallel or in series (FEMA 233, 1992): however, no differentiation is made between these
two types of compressors in the analysis of natural gas systems.
In Greece, it seems that most compressor stations are housed in RC low-rise buildings, with
anchored components. Those we propose to consider the following typologies for
compressor stations:
- with anchored or unanchored components;
- within low-rise buildings, made of masonry or RC;
We also propose to include oil pumping stations in this category of stations, as they are very
similar to compressor stations (presence of a pump instead of a compressor).

Metering / Pressure reduction stations (M/R stations)

These stations are used to reduce the gas pressure for industrial or pressure use. In each
station, there is a SCADA system that controls the equipment. The natural gas system of
Thessaloniki comprises two central M/R stations. In the L’Aquila area, the medium-pressure
network is connected to the high-pressure transmission lines through three M/R stations
(referred to as RE.MI “REgolazione e MIsura” in Italian). RE.MI stations are one-story
masonry buildings with steel roofs (Figure 7).
Inside the M/R stations (Figure 8) the gas undergoes the following operations and
processes:
- gas preheating;
- gas-pressure reduction and regulation;
- gas odorizing;
- gas-pressure measure;
- data transmission through SCADA system (no SCADA for the Re.Mi cabins in the
L’Aquila area);
The specificities of the operations performed within these stations prevent them from being
included in the compressor stations typologies.

18
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

Figure 7 RE.MI cabin in the L'Aquila area: outside view (courtesy of Enel Rete Gas)

Figure 8 RE.MI cabin in the L'Aquila area: inside view (courtesy of Enel rete Gas)

Reduction groups

In L’Aquila, about 300 Final Reduction Groups (referred to as GRF “Gruppi di Riduzione
Finale” in Italian, Figure 9) allow for the transformation of the medium distribution pressure
into the low distribution pressure. These facilities can be either buried, sheltered in a kiosk or
housed within a building.

19
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

Figure 9 View of a reduction group as used in the L'Aquila area (courtesy of Enel Rete
Gas)

4.2 GENERAL DESCRIPTION OF EXISTING METHODOLOGIES

4.2.1 Empirical relations

The most used and straightforward approach is based on empirical data collected
throughout past earthquakes.
In the case of pipeline components, the usual practice is to evaluate the repair rate per unit
length of pipe, with respect to a parameter representative of ground shaking (e.g. PGV or
PGA) or ground failure (e.g. permanent ground deformation, PGD). Empirical data is
collected from gas / oil companies operating the pipelines and consists of the following:
length of pipes subjected to a given level of ground shaking, and the number of repairs
carried out for that segment. This means that this data is very generic and no distinction is
made between the different kinds of repairs: complete fracture of the pipe, leak in the pipe or
damage to an appurtenance of the pipe (ALA, 2001).
Usually, some adjustments to the raw data are performed: for instance, in the ALA (2001)
methodology, only the damage to the main pipe is used to assess the relative vulnerability of
different pipe materials. Also, data points assumed to contain permanent ground
displacement effects can be eliminated when studying only the effects of ground shaking.
Then, based on the data points, a correlation procedure is performed in order to fit a
predefined functional form with the empirical data. For example, ALA (2001) explored a
linear model (RR = a.IM) and a power model (RR = b.IMc). Depending on the consistency of
the available data, it is possible to build specific models based on various factor such as pipe
material, pipe diameter or pipe connections. Many empirical studies have been carried out,
like the ones from HAZUS (NIBS, 2004), Eguchi et al. (1983), Eidinger (1998), Isoyamaet al.
(2000) or Toprak (1998).
Regarding storage tanks, empirical relations are also quite common, such as (O’Rourke &
So, 2000), HAZUS (NIBS, 2004), (ALA, 2001) or (Fabbrocino et al., 2005). During
earthquakes, at each refinery or storage facility subjected to a given level of ground shaking,

20
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

the proportion of damaged tanks is evaluated. Observations may give some details about
the type of failure (roof damage, pipe connection failure, elephant’s foot buckling…), which is
then translated into a damage state (whose definition changes depending on the study). The
occurrence of a damage state is then fitted into a probability function, taking the form of a
lognormal distribution with two parameters (median and standard deviation).
The accuracy of empirical relations is heavily relying on the quality and quantity of available
data: for example, the distinction of vulnerability into different typologies is limited to what
has been observed and recorded in a statistically reliable way. Also, most of these empirical
studies have been based on non-European earthquakes and damages that occurred on
fairly ancient components (e.g. pipelines built in the early 30s). However, it has to be
stressed that these empirical studies have come a long way, from the first relation (Eguchi et
al., 1983) to the work by O’Rourke & Deyoe (2004), which introduced the ground strain
(based on wave velocity) as an intensity measure. In this scope, the 1944 Northridge
earthquake constituted an important breakthrough, as it allowed for the collection of valuable
damage data on fairly modern infrastructures.

4.2.2 Bayesian approach

In the case of storage tanks, one study (Berahman & Behnamfar, 2007) proposes to use a
Bayesian approach to improve the empirical procedure. The authors use field observations
of unanchored on-grade steel tanks, previously reported by ALA (2001), and they aim at
accounting for both aleatory and epistemic (model bias, small data sample, measurement
errors …) uncertainties. Fragility models are developed using a probabilistic limit state
function and a reliability integral, solved with Monte-Carlo simulation. It was found that the
fragility curves were less conservative than purely empirical models from ALA (2001) or
NIBS (2004), suggesting a better tank performance than expected. Also, one important
result is that commonly used lognormal distributions do not seem to be the best fit to the
available empirical data.
However, this study was only conducted for a specific typology of tanks (unanchored on-
grade steel tanks) and other sets of fragility curves should be built to cover all typologies.
Finally, the proposed fragility curves are based on an integral formulation and are not
associated with an analytical function (like the lognormal distribution, which can be easily
described with two parameters).

4.2.3 Analytical approach

In the case of buried pipelines, there are not many examples of analytical studies. Terzi et al.
(2007) developed fragility curves for the case of segmented pipelines subjected to
permanent ground deformation, using a FEM model and accounting for pipe-soil interaction.
The results were confronted with the case of a PVC pipe that suffered damage from the
2003 Lefkas earthquake.
Regarding storage tanks, a study by Iervolino et al. (2004) performs numeric analyses on
dynamic models of unanchored steel tanks. Using a limit state function (e.g. axial stress,
governing the failure by elephant’s foot buckling), the authors propose a decomposition of
the random structural variable into those affecting the capacity (e.g. mechanical or material
properties) and those affecting the demand (set of parameters defining the structure, shape,
dimensions…). A design of experiments is set up, with two axis (fluid level-over-radius ratio,

21
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

and friction coefficient between the baseplate and the unanchored tank), and for each of the
studied structure, a reliability analysis is performed in order to obtain a fragility curve. Then,
through a second order polynomial model, it is possible to obtain the response surface of the
fragility parameters (median and standard deviation of the lognormal distribution). As
expected, a low friction coefficient and a nearly full tank were found to have the highest
probability of failure. This method is interesting, as it allows for the vulnerability assessment
of all structures belonging to the “tank” typology. Yet, this study was only applied to one
specific damage state (failure by elephant’s foot buckling), and other sets of response
surfaces should be needed for an exhaustive evaluation of the vulnerability of storage tanks.

4.2.4 Fault-tree analysis (for support facilities)

In-line components or processing facilities such as gas compressor stations include many
subcomponents, which make a quantitative vulnerability assessment quite difficult.
In the configuration where support equipment (e.g. pumps, compressor, electric cabinets…)
are sheltered within a building, a solution is to treat these facilities as a common building.
Thus, one can use the fragility curves for low-rise RC or masonry structures to assess the
vulnerability of the compressor or pumping stations. (Kappos et al., 2006) proposed such
fragility curves for pumping plants in low-rise RC buildings, with anchored components.
Another approach is to consider these facilities as systems and to aggregate the fragility of
each component into a global systemic vulnerability. Such a work has been carried out in the
SRMLIFE project (Greek project, 2003 – 2007, where a gas compressor stations is
decomposed into the following components:
- building;
- electrical / mechanical equipment;
- pump;
- commercial power;
- power backup;
Then, using the HAZUS (NIBS, 2004) fragility curves for these individual components and
the curves from Kappos et al. (2006) for the building, it is possible to compute the global
fragility curve of the plant, based one the following fault-tree (this fault-tree represents the
global functionality analysis of the station, particular damage states may be represented with
a different logic):

22
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

Figure 10 Decomposition of a compressor station into a fault-tree

The fault-tree decomposition follows a logic structure, with AND / OR operators that indicate
how to aggregate the fragilities of two connected components. For two components A and B
assembled in series (e.g. AND operator), like a pump and the building where it is stored into,
the probability of failure of the system A-B is given by ( Equation 1

Pf ( AB) ? Pf ( A) © Pf ( B) ( Equation 1 )

On the other hand, when components A and B are mounted in parallel (e.g. OR operator),
like commercial power connection and backup power, the probability of failure is expressed
as ( Equation 2:

] _]
Pf ( AB) ? 1 / 1 / Pf ( A) © 1 / Pf ( B) _ ( Equation 2 )

Using these basic rules, it is indeed possible to build up the global probability of failure of the
compressor stations and to account for the fragility of both the building and the components
within.

23
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

4.3 DAMAGE STATES

4.3.1 Pipeline components

Damage states definition

As stated before, empirical relations for the fragility of pipelines are based on the recorded
number of repairs during past earthquakes, and no distinction is really made between leaks
and breaks of the pipes. As a result, all fragility relations for pipelines are given for a single
“damage state”, e.g. the repair rate per unit length of pipe. However, according to HAZUS
(NIBS, 2004), the type of repair or damage depends on the type of hazard: a damaged pipe
because of ground failure is likely to present a break (it is assumed 80% breaks and 20%
leaks), whereas ground shaking may induce more leak related damages (e.g. 20% breaks
and 80% leaks). These percentages are accepted in the reviewed publications that are
interested in making such distinction even if these values do not result from a specific
argumentation. However, they provide a rare link between what can be seen as the repair
cost (almost directly linked to the number or repairs, which is the recorded data) and the
functionality (whether a pipe is leaking and assumed to continue being operational, or
broken and unable to assume its function, as discussed further in 3.5.1).

Then, using a Poisson probability distribution and the repair rate RR, one can assess the
probability of having n pipe breaks / leaks in a pipe segment of length L l g! g i
ヾ i j k g gl hi i.):

P * N ? n + ? e / RR© L ©
*RR © L +n ( Equation 3 )
n!

Finally, assuming that a pipe segment fails (flow rupture) when it has a least one break along
its length, the probability of failure is given by ( Equation 4:

Pf ? 1 / P* N ? 0+ ? 1 / e / RR©L ( Equation 4 )

Finally, using the HAZUS assumption and considering the type of hazard, it is possible to
assess the probability to have a pipe break or a pipe leak along the length of the segment.

Performance indicators

The performance indicators of a pipeline in terms of serviceability or functionality may take


various definitions, depending on the point of view. For the end-user, it can be for example
the availability of the desired amount of product (at the requested flow / pressure / quality);
for the operating companies, this can be measured at the network level by the ratio of
satisfied demand on the total demand, or the total loss of contents; also, for environment
agencies, it can represent the amount of spilled contents in the nature.

24
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

Table 3 Proposed damage states for pipeline components

Damage state Damage description


DS0 no damage no break / leak
DS1 leakage at least one leak along the pipe length
DS2 failure at least one break along the pipe length

Regarding the pipelines, as two damage states are considered, it is possible to propose the
following correspondence between pipe damage and serviceability/ functionality (Table 3):
- DS1: leakage – we assume that the pipe may still be functional, yet with a reduced
flow (a given % of the flow is assumed to go out of the system);
- DS2: failure – the flow is disrupted, and we can consider the pipe segment as
disconnected from the rest of the network;

4.3.2 Storage tanks

Damage states definition

Fragility curves from the literature, whether they are empirical or analytical, usually propose
the same number of damage states (e.g. 5, including “no damage”), and very similar
definitions (O’Rourke & So, 2000), (ALA, 2001), (NIBS, 2004), (Berahman & Behnamfar,
2007):
- DS1: no damage
- DS2: slight / minor
- DS3: moderate
- DS4: extensive
- DS5: complete / collapse

Table 4 Damage states defined in HAZUS (NIBS, 2004) and (O’Roure & So, 2000)

Damage state Damage definition


DS1 None No damage to tank or I/O pipes
DS2 Slight / minor Damage to roof, minor loss of contents, minor damage to piping,
but no elephant’s foot buckling
DS3 Moderate Elephant’s foot buckling with minor loss of content
DS4 Extensive Elephant’s foot buckling with major loss of content, severe damage
DS5 Complete Total failure, tank collapse

25
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

The detailed damage states used by HAZUS (NIBS, 2004) and O’Rourke & So (2000) are
given in Table 4. HAZUS damage states were initially developed for building-type structure
for which it is reasonable to relate increasing direct damage and decreasing functionality.
This has been adopted for tanks, whereas there is no direct correlation between repair cost
and functionality for tanks. Damage states are consequently based on the physical damage
of the tank (direct economic losses) and their classification does originate from the
serviceability but from the repair cost (expressed as a percentage of the whole tank
replacement cost).

Performance indicators

The authors of ALA (2001) identified the need of relating serviceability to a given aggression
level. They propose the following Table 5 to relate a damage state (which corresponds to a
repair cost) to a level of functionality.

Table 5 Correlation between damage modes and repair costs (ALA, 2001)

Most common damage modes Repair costs (% of Functionality as


the replacement percentage of
cost) content lost
immediately
after the
earthquake
Rupture of drain pipe 1% to 2% 50% to 100%
Rupture of overflow pipe 1% to 2% 0% to 2%
Rupture of Inlet/Outlet pipe 1% to 5% 100%
Rupture of bottom plate from bottom course 2% to 20% 100%
Roof system partial damage 2% to 20% 0% to 10%
Roof system collapse 5% to 30% 0% to 20%
Upper shell buckling 10% to 40% 0% to 20%
Elephant’s foot buckling with no leak 30% to 80% 0%
Elephant’s foot buckling with leak 40% to 100% 100%

The damage states definitions from ALA (2001) are very similar to the table 4 as it has been
built in order to closely match the fragility curves used in HAZUS. In addition, they were
modified to account for the potential large loss of functionality due to inexpensive damages.
For instance, a tank with “broken I/O pipes” is classified in DS4 (cf. Table 6).

26
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

Table 6 Damage states adapted from (ALA, 2001)

Damage Damage definition Functionality as the


state content lost immediately
after the earthquake
DS1 None No damage to tank or I/O pipes 0%
DS2 Slight / Damage to roof other than buckling, 1% to 20%
minor minor damage to piping
DS3 Moderate Elephant’s foot buckling with minor loss 20% to 40%
of content
DS4 Extensive Elephant’s foot buckling with major loss 40% to 100%
of content, severe damage, broken I/O
pipes
DS5 Complete Total failure, tank collapse 100%

The authors are still careful with the construction of the corresponding fragility curves:
“Because of incomplete descriptions of the actual damage to some tanks, the definition of
damage state between DS = 2, DS = 3 and DS = 4 is sometimes left to judgment”.
One approach for obtaining the probability of occurrence of these damage states is to
decompose each state into a fault-tree, down to the basic failure modes of the tank (Figure
11).

Figure 11 Example of a fault-tree for an anchored steel tank failure modes (ALA, 2001)

27
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

4.3.3 Processing facilities (pumping / compressor stations)

Damage states definition

Following the approach by (Kappos et al., 2006) a pumping / compressor stations may have
the same damage states as a usual building, the loss index being defined by the percentage
of failed structural elements (criterion also used in HAZUS methodology, Table 7):

Table 7 Damage states defined by (Kappos et al., 2006) for buildings

Damage state Loss index


DS0 None 0%
DS1 Slight 0 – 1%
DS2 Moderate 1 – 10%
DS3 Substantial to heavy 10 – 30%
DS4 Very heavy 30 – 60%
DS5 Collapse 60 – 100%

In the case of a fault-tree analysis of the compressor station, the global damage state is
based on the individual damage state of its components. For example, a slight / minor
damage (e.g. short-time malfunction of the plant) to the station may be induced by the loss
of electrical power and backup generators, or a slight damage to the building. Such an
approach was used in the LESSLOSS (2007) and SRMLIFE (2003-2007) projects, which
resulted in the damage scale presented in Table 8.

Table 8 Damage scale proposed by (LESSLOSS, 2007) and (SRMLIFE, 2003 - 2007) for
pumping / compressor stations

Damage state
DS0 None
DS1 Slight / minor
DS2 Moderate
DS3 Extensive
DS4 Complete

The damage states above are not directly related to physical damage (percentage of loss),
and they integrate also functional aspects: therefore a straightforward description is not
available. A more detailed analysis of this case is presented is the next sections
(performance indicators).

28
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

Finally, regarding in-line valves or SCADA equipment, as it was explained before, no


quantitative study of their vulnerability is available, and there is no point in defining damage
states for now.

Performance indicators

Performance indicators associated with processing facilities such as pumping / compressor


stations may include measures of loss of contents or percentage of decrease of output
pressure / flow.
The damage states defined for pumping / compression stations (see Table 8) in projects
LESSLOSS (2003-2007) and SRMLIFE (2003-2007) can be considered as performance
indicators, as they already account for the functionality loss (aside from the physical
damage). The belowTable 9 proposes indeed a correlation between the damage states
(repair costs in % of the replacement value) and some functionality / serviceability indicators:

Table 9 Description of damage states and functionality indicators for gas stations
(SRMLIFE, 2003-2007)

Repair
Serviceability/ Functionality Damage state
cost (%)
Disability of Complete
75 – 100 Building collapsed
boosting gas in damage
compression Full loss of
function Building being
station
(un repairable extensively damaged,
Extensive
damage) 50 – 75 or the pumps badly
Damage
damaged beyond
repair.
Considerable damage
to mechanical and
electrical equipment
Malfunction.
Moderate or considerable
(Full function 30 – 50
Damage damage to building or
after repairs)
loss of electric power
and of backup for 7
days.
Several stops and Slight damage to
reduced flow of gas building or full loss of
10 – 30 Slight/Minor
in the transmission commercial power
gas pipelines Full function Damage
and backup power for
1 - 10 few days (< 3 days)
Normal function
- No -

29
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

4.4 INTENSITY INDEXES

4.4.1 Pipeline components

Wave propagation

Among the numerous studies proposing empirical relations for pipelines subjected to
transient ground deformation, several intensity measure parameters have been proposed:
- Macro-seismic intensity: (Eguchi, 1983), (Eguchi, 1991);
- PGA: (Katayama et al., 1975), (Isoyama & Katayama, 1982), (Hamada, 1991),
(Isoyama et al., 2000), (O’Rourke et al., 1998);
- PGV: (Barenberg, 1988), (O’Rourke & Ayala, 1993), (Eidinger et al., 1995), (Eidinger,
1998), (Isoyama, 1998), (O’Rourke et al., 1998), (Toprak, 1998), (Eidinger & Avila,
1999), (O’Rourke & Jeon, 1999), (ALA, 2001), (Pineda & Ordaz, 2003), (HAZUS,
2004), (O’Rourke et Deyoe, 2004);
- PGV²/PGA: (Pineda & Ordaz, 2007);
- PGS (transient strain): (O’Rourke & Deyoe, 2004), (Paolucci & Pitilakis, 2007);
We can notice that PGV is by far the most used intensity-measure parameter, from the first
empirical relations to the more recent ones. However, the seismic response of buried pipes
is mostly controlled by the amplitude of transient strain, induced in the ground by the wave
propagation. The peak horizontal ground strain can indeed be related to the peak horizontal
velocity V (PGV), by the following equation ( Equation 5 ):
V
g ? ( Equation 5 )
C

where C is the wave propagation velocity in the surrounding soil, relative to the ground
surface. In the case of body waves, S-waves only are considered, since they tend to
generate larger ground motion that P-waves. Concerning surface waves, the most significant
motions are those caused by Rayleigh waves. According to LESSLOSS (2007), longitudinal
strains are the most damaging factor, compared to shear strains: Therefore it is possible to
determine the incident angle of the waves that can generate maximum longitudinal strains
(St John & Zahrah, 1987). Their results are summed up in Table 10.

Table 10 Maximum longitudinal strains induced by seismic waves propagation along


a pipeline (St John & Zahrah, 1987)

Ground motion direction Maximum strain


Incident angle Value
R waves Horizontal component parallel to 0° PGV R
the wave direction g?
CR
S waves Perpendicular to the wave 45° PGV S
propagation direction g?
2C S

30
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

O’Rourke & Deyoe (2004) proposed a relation between peak ground strain and repair rate of
pipes, obtaining very promising results. Also, they introduced a distinction between
earthquakes depending on which type of wave is responsible for the damage: the authors
assumed that surface R waves are dominant for a shallow distant earthquake (ratio of
epicentral distance to focal depth > 5), whereas body waves S-waves control for deep and
close earthquakes. This study resulted in two statistically reliable expressions between PGV
and repair rate, based on the type of waves.
Also, Paolucci & Pitilakis (2007) established an empirical relation between PGV and PGS
(strain), based on a selection of a few earthquake records (see Figure 12). It was
established( Equation 6 ):

PGV ]m / s _
PGS ? ( Equation 6 )
1300

Figure 12 Empirical correlation between PGS (strain) anf maximum horizontal PGV
(Paolucci & Pitilakis, 2007)

Although this approach of empirical relations using transient ground strain as an intensity-
measure parameter seems really powerful, the studies carried out so far do not develop
fragility functions for the typology directly concerned with gas pipelines: these are usually
made of ductile materials, whereas the results of (O’Rourke & Deyoe, 2004) are mainly
relying on brittle segmented pipes, like cast-iron. Therefore we propose to adopt PGV as the
intensity-measure parameter: this choice allows us to benefit indeed from a large number of
empirical relations, based on a wide range of pipe typologies.
We may notice that all the presented empirical fragility curves do not consider the direction
of the pipe whereas it is acknowledged that the longitudinal strain is responsible for the
failures. This is justified because once used on the distribution network of a study case,
these relations are applied to a large number of pipelines that can assume to be randomly

31
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

oriented. Consequently, the additional effort of considering the angle between the pipe and
the direction of the transient displacement does not seem particularly interesting as long as
we are looking at the large number of pipes and failures.

Ground failure

Concerning ground failure, all studies found in the literature rely on the permanent ground
displacement (PGD) to describe the fragility of buried pipes: therefore this parameter is
selected as the intensity-measure index for pipeline damage due to ground failure.
However, it is worth noticing that the study by O’Rourke & Deyoe (2004) has established a
good correlation between ground strain and repair rate (Figure 13): this relationship is valid
for both transient and permanent ground deformations, thus such an intensity index could
also be used for ground failure.

Figure 13 Strain vs Repair rate correlation, for both wave propagation and permanent
ground displacement (O’Rourke & Deyoe, 2004)

32
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

4.4.2 Storage tanks

Past studies on the vulnerability of storage tanks usually propose PGA as the earthquake
descriptor used to define the fragility curves. This seems to be a good choice as this
acceleration-driven parameter is appropriate to account for the inertia forces inherent to
these large and usually tall structures and the liquid contents within.

4.4.3 Processing facilities

As we consider here mostly facilities sheltered in a building, a classic parameter is PGA, as


it is widely used to describe the fragility of RC or masonry buildings. Also, the behaviour of
anchored or unanchored components within the facility seems acceleration-driven and its
fragility is indeed with respect to PGA in HAZUS (NIBS, 2004).

33
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

5 Fragility functions for system elements

5.1 STATE-OF-THE-ART FRAGILITY CURVES PER COMPONENT

5.1.1 Pipeline components

Ground shaking

The literature review has resulted in a total of 20 empirical studies (the list, in Table 11, may
not be exhaustive) that addressed the issue of fragility relations for pipeline components
subjected to transient ground shaking. We display these studies in the table below, along
with the earthquake descriptor used, the typology of pipes, and the quality of the empirical
data (i.e. number of earthquakes used):

Table 11 Summary of the fragility functions from the literature

Study Typology Intensity index Nb of earthquakes


studied
Katayama et al., - mainly cast-iron pipes PGA 6
1975 - poor, average or good
conditions
Isoyama & - mainly cast-iron pipes PGA 1
Katayama, 1982
Eguchi, 1983 - WSGWJ (welded-steel MMI 4
gas-welded joints),
WSAWJ (welded-steel
arc-welded joints), AC
(asbestos cement),
WSCJ (welded-steel
caulked joints), CI (cast
iron)
Barenberg, 1988 - mainly cast-iron pipes PGV 3
Eguchi, 1991 - WSGWJ (welded-steel MMI 4
gas-welded joints),
WSAWJ (welded-steel
arc-welded joints), AC
(asbestos cement),
WSCJ (welded-steel
caulked joints), CI (cast
iron), DI (ductile iron),
PVC, PE (polyethylene)

35
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

Study Typology Intensity index Nb of earthquakes


studied
O’Rourke et al., - MMI 7
1991
Hamada, 1991 - PGA 2
O’Rourke & Ayala, - brittle or flexible pipes PGV 6
1993
HAZUS (NIBS,
2004)
Eidinger et al., 1995 - material type PGV 7
Eidinger, 1998 - joint type
- diameter
- soil type
O’Rourke et al., - mainly cast-iron pipes PGV, PGA, MMI 4
1998
Isoyama, 1998 - material type PGV 1
- diameter
Toprak, 1998 - no distinction PGV 1
O’Rourke & Jeon, - mainly cast-iron pipes PGV 1
1999 - diameter
Eidinger et Avila, - material type PGV -
1999 - joint type
- diameter
- soil type
Isoyama et al., 2000 - DI, CI, PVC, steel, AC PGA, PGV 1
- diameter
- soil type
ALA, 2001 - material PGV 18
- joint type
- soil type
- diameter
Pineda & Ordaz, - mainly brittle pipes (CI, PGV 1
2003 AC)
O’Rourke & Deyoe, - mainly cast-iron pipes PGV, PGS 5
2004
Pineda & Ordaz, - mainly brittle pipes (CI, PGV²/PGA 1
2007 AC)

36
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

The work of Tromans (2004) resulted in a very thorough review of existing pipeline
vulnerability studies. The following subsections propose a summary of some of the most
consistent empirical relations.

Katayama et al., 1975


This is one of the very first studies trying to establish a correlation between observe seismic
damage and a strong-motion parameter (Figure 14). It is based on pipe failure rates
obtained for six earthquakes (4 of them Japanese): 1923 Kanto, 1948 Fukui, 1964 Niigata,
1968 Tokachi-oki, 1971 San Fernando and 1972 Managua earthquakes. A large scatter in
the data is observed, probably due to larger damage rates induced in certain cases by
permanent ground deformation. Most of the data used concerns cast-iron pipes, although
the 1968 Tokachi-oki earthquake includes also damage to asbestos-cement pipes. No
distinction is made on pipe diameter, joint types or pipe material. However the authors
introduce a parameter b ( Equation 7), which depends on several factors like soil condition or
pipe age. Depending on the “poor”, “average” or “good” conditions, this constant can take
the respective values of 4.75, 3.65 or 2.0 (Ayala & O’Rourke, 1989).

RR ? 10b-6.39 log PGA ( Equation 7 )

Figure 14 Pipeline fragility data of Katayama et al. (1975) as presented by O'Rourke &
Liu (1999)

37
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

Eguchi, 1991
This study is an update of the earlier work of Eguchi (1983), based on damage date from
four earthquakes: 1969 Santa Rosa, 1971 San Fernando, 1972 Managua and 1979 Imperial
Valley earthquakes. This study explicitly separates wave propagation damages from those
induced by permanent ground deformation, and a distinction is made between different pipe
materials and joint types. Bilinear relations with respect to macroseismic intensity (Modified
Mercalli Intensity) are proposed for each of the pipe types.

Figure 15 Bilinear pipeline fragility relations by (Eguchi, 1991)

O’Rourke & Ayala, 1993


This study uses the original data from Barenberg (1988), plus three additional earthquakes:
1983 Coalinga, 1985 Michoacan and 1989 Tlahuac earthquakes. A total of 11 data points
are used to plot the trend line with respect to PGV (Figure 16), as it was found earlier by
Katayama et al. (1975) and Barenberg (1988) that PGA is not the best earthquake descriptor
for pipeline damage. The fragility equation is given by( Equation 8:

RR ? 0.0001PGV 2.25 ( Equation 8 )

It appears to be only valid for brittle pipes, as the damage data is based on asbestos-
cement, concrete and cast-iron pipes: for more ductile pipe material, it is recommended to
multiply the repair rate by a corrective factor of 0.3.

38
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

Figure 16 Fragility relations of Barenberg (1988) and O'Rourke & Ayala (1993)

Eidinger et al, 1995, 1998


The work of Eidinger et al. (1995) and Eidinger (1998) is based on the same data as
O’Rourke & Ayala (1993), plus the 1989 Loma Prieta earthquake (i.e. seven US and
Mexican earthquakes in total). Detailed pipeline data have allowed the authors to estimate
different fragility relations based on various factors such as pipe material, diameter, joint type
or soil corrosion. The “best-fit” relation (in this case regression with all data) is defined by the
equation below ( Equation 9 ) and displayed on Figure 17:

RR ? K 1 0.0001658PGV 1.98 ( Equation 9 )

Without any distinction on the pipe features, K1=1. Otherwise, the values of this corrective
factor are given inTable 12, for different pipe configurations.

39
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

Figure 17 Fragility function by Eidinger et al. (1995, 1998), for all data

Table 12 Values of corrective factor K1, according to Eidinger et al. (1995, 1998)

pipe material joint type soil Diameter K1 quality index


CI cement unknown Small 0.8 B
cement corrosive Small 1.1 C
cement non corrosive Small 0.5 B
rubber gasket unknown Small 0.5 D
WS arc welded unknown Small 0.5 C
arc welded corrosive Small 0.8 D
arc welded non corrosive Small 0.3 B
arc welded all Large 0.15 B
rubber gasket unknown Small 0.7 D
AC rubber gasket all Small 0.5 C
cement all Small 1.0 B
cement all Large 2.0 D
C welded all Large 1.0 D
cement all Large 2.0 D
PVC rubber gasket all Small 0.5 C
DI rubber gasket non corrosive All 0.3 C

40
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

Each of these K1 values is linked to a quality index that gives the degree of confidence in the
empirical data used:
- B: “there is a reasonable amount of background empirical data and study”;
- C: “limited empirical data and study”;
- D: “based largely on extrapolation and judgment, with very limited empirical data”;

Isoyama et al., 2000


The work by Isoyama et al. (2000) is based on earlier studies of the 1995 Kobe earthquake.
The damage data is concentrated on distribution pipes located in Kobe and two others cities
nearby. The following functional forms ( Equation 10 )and( Equation 11 ) is adopted to
represent the pipeline repair rate:

RR( IM ) ? Bp Bd Bg BL R0 ( IM ) ( Equation 10 )

R0 ( IM ) ? a * IM / IM min + ( Equation 11 )
b

The intensity measure parameter, IM, is either PGA or PGV. R0 represents the “standard”
repair rate, for cast-iron pipes with medium diameter located in alluvial soil (Figure 18). The
authors account for various typologies (pipe material, diameter, ground topography,
liquefaction) by introducing corrective factors Bp, Bd, Bg and BL (see Table 13).

Table 13 Values of corrective factors according to (Isoyama et al., 2000). Bracketed


values are less reliable due to small sample size.

pipe material, Bp pipe diameter (mm), ground topography, liquefaction, BL


Bd Bg
DI 0.3 75 1.6 disturbed hill 1.1 no liquefaction 1.0
CI 1.0 100-150 1.0 Terrace 1.5 partial 2.0
liquefaction
PVC 1.0 200-400 0.8 narrow valley 3.2 total 2.4
liquefaction
Steel (0.3) > 500 (0.5) Alluvial 1.0
AC (1.2) stiff alluvial 0.4

For PGA, 19 data points were used and a relation was established with a=2.88x10-6 and
b=1.97. In the case of PGV, the values a=3.11-3 and b=1.6 were found to best fit the 16 data
points. The corresponding lines are displayed in the figure below:

41
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

Figure 18 Fragility relations by Isoyama et al. (2000), for both PGA (a) and PGV (b)
parameters, without corrective factors

ALA, 2001
The work carried out in (ALA, 2001) is a compilation of several past studies, including a total
of 18 earthquakes:
- Eidinger et al., 1995;
- Katayama et al., 1975;
- O’Rourke & Ayala, 1993;
- Shirozu et al., 1996;
- Toprak, 1998;
The study consists mainly in a homogenization of all available data and some data cleaning
(some points were excluded due to an excessive influence of permanent ground deformation
effects). This compilation gathers also a good sample of different material types, including
ductile ones. A total of 81 data points are extracted and used to build a “backbone” curve,
based on a single linear model, defining the median slope of all data points:

42
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

Figure 19 "Backbone curve" proposed by ALA (2001), representing the median repair
rate of all data points, and the corresponding 16th and 84th quantiles.

The median curve is given by the following equation (with K1=1):

RR ? K 1 0.002416 PGV ( Equation 12 )

Like the work of Eidinger et al. (1995, 1998), a corrective factor K1 is introduced in order to
account for various configurations such as pipe material, diameter, joint type, soil corrosion.
The different values of this factor are given in the table below:

43
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

Table 14 Values of corrective factor K1, according to (ALA, 2001)

pipe material joint type soil diameter K1


CI cement unknown small 1.0
cement corrosive small 1.4
cement non corrosive small 0.7
rubber gasket unknown small 0.8
WS arc welded unknown small 0.6
arc welded corrosive small 0.9
arc welded non corrosive small 0.3
arc welded all large 0.15
rubber gasket unknown small 0.7
screwed all small 1.3
riveted all small 1.3
AC rubber gasket all small 0.5
cement all small 1.0
C welded all large 0.7
cement all large 1.0
rubber gasket all large 0.8
PVC rubber gasket all small 0.5
DI rubber gasket all small 0.5

O’Rourke & Deyoe, 2004


The work of (O’Rourke & Deyoe, 2004) includes data from three US and two Mexican
events: 1965 Puget Sound, 1971 San Fernando, 1983 Coalinga, 1985 Michoacan and 1994
Northridge earthquakes. They introduce a criterion to select only statistically reliable data (
Equation 13 )where n is the minimum number of km of pipe data for a given repair rate):

1 / RR
n 15.36 ( Equation 13 )
RR

This led to the selection of 14 data points. Then, using the PGV value, the authors back-
calculate the transient ground strain (based on the apparent wave propagation velocity, see
equations in Table 10). A distinction is made between earthquakes generating surface
waves (“shallow” earthquake and “distant“ basin) and events where body S-waves are
dominant: the velocity of R-waves is assumed to be CR=500 m/s, whereas CS=3000 m/s.
These assumptions are used to develop the following fragility relations Equations 14, 15 and
16, based on both transient ground strain and PGV:

44
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

RR ? 724g 0.92 (for all cases) ( Equation 14 )

RRR ? 0.064PGV 0.92 (if Rayleigh waves are dominant) ( Equation 15 )

RRS ? 0.0035PGV 0.92 (if shear waves are dominant) ( Equation 16 )

This analysis has been performed mainly on segmented cast-iron pipelines.

Ground failure

For the damages related to permanent ground deformation effects, the literature review
resulted in 7 studies, which are presented in the table below:

Table 15 Summary of the past studies on fragility of pipeline components subjected


to permanent ground deformation

Study Typology Proposed relation


Eguchi, 1983 - WSGWJ (welded-steel gas- see Figure 20
welded joints), WSAWJ
(welded-steel arc-welded
joints), AC (asbestos
cement), WSCJ (welded-
steel caulked joints), CI (cast
iron)
Honegger & Eguchi, 1992 - ductile (steel, DI, PVC) or RR ? K 7.821PGD 0.56
HAZUS (NIBS, 2004) brittle (AC, concrete, CI)
K=0.3 for flexible pipes
Ballantyne & Heubach, 1996 - material : WS, old steel and relative vulnerability (see
CI, locked converse, AC, CI l g! g i
ヾ i j k g gl
hi i.)
Eidinger & Avila, 1999 - ductile of brittle pipes RR ? K 2 23.674PGD 0.53
K2=0.5 for flexible pipes
ALA, 2001 - material RR ? K 2 11.223PGD 0.319
- joint type
See Table 16for K2 values
Terzi et al., 2006, 2007 - segmented pipes (PVC) RR ? 3.2103 PGD 0.4103

The empirical fragility curves proposed by Eguchi (1983) are shown in the figure below. The
author has focused his work on localised abrupt PGD induced by fault ruptures or landslides.
These hazards are indeed likely to induce high strains, as opposed to spatially distributed
deformations induced by ground shaking.

45
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

Figure 20 Fragility relations proposed by (Eguchi, 1983) for fault ruptures

The study by ALA (2001) is based on empirical judgement, along with engineering
judgement. Like for the wave propagation related damage, the proposed “backbone curve”
can be modified by a corrective factor K2 according to the pipe’s material and joint type. The
different values of the factor K2 are presented in the table below:

Table 16 Values of the corrective factor K2 according to (ALA, 2001)

Pipe material Joint type K2


Unknown Unknown 1.0
CI cement 1.0
rubber gasket 0.8
mechanical restrained 0.7
WS arc-welded, lap welds 0.15
rubber gasket 0.7
AC rubber gasket 0.8
cement 1.0
Concrete welded 0.6
cement 1.0
rubber gasket 0.7
PVC rubber gasket 0.8
DI rubber gasket 0.5

46
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

A study by Ballantyne & Heubach (1996) proposes also to rate various pipe typologies:
ruggedness (strength and ductility of the pipe barrel), resistance to bending failure, joint
flexibility, and joint restraint. A grade from 1 to 5 is attributed to each of these criteria and a
global evaluation of the relative vulnerability of each typology is established (Table 17).

Table 17 Proposed vulnerability indices by Ballantyne, for various pipe materials and
joint types (B&S: bell and spigot, RG: rubber gasket, R: restrained, UR: unrestrained)

material joint type Total

restraint
bending

flexibilit
rugged
ness

joint

joint
y
HDPE Fusion 4 5 5 5 19
Steel Arc-welded 5 5 4 5 19
Steel Riveted 5 5 4 4 18
Steel B&S, RG, R 5 5 4 4 18
DI B&S, RG, R 5 5 4 4 18
Steel B&S, RG, UR 5 5 4 1 15
DI B&S, RG, UR 5 5 4 1 15
Concrete B&S, R 3 4 4 3 14
PVC B&S, R 3 3 4 3 13
Concrete B&S, UR 3 - 4 1 12
AC (>8” d.) Coupled 2 4 5 1 12
CI (>8” d.) B&S, RG 2 4 4 1 11
PVC B&S, UR 3 3 4 1 11
Steel Gas welded 3 3 1 2 9
AC (<8” d.) Coupled 2 1 5 1 9
CI (<8” d.) B&S, RG 2 1 4 1 8
CI B&S, rigid 2 2 1 1 6

5.1.2 Storage tanks

We propose here a brief description of each of the fragility curves found in the literature
review for storage tanks.

O’Rourke & So, 2000

This study is based on the investigation of the fragility of on-grade steel tanks: more than
400 tanks damages are selected from 9 earthquake events: 1933 Long Beach, 1952 Kern
County, 1964 Alaska, 1971 San Fernando, 1979 Imperial Valley, 1983 Coalinga, 1989 Loma

47
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

Prieta, 1992 Landers and 1994 Northridge earthquakes. The size of the available data
allowed the authors to investigate the effects of two parameters:
- the tank’s height to diameter ratio, H/D;
- the relative amount of stored contents, % Full;
No distinction is made between anchored and unanchored tanks, as this condition was only
specified for 40% of the data: however, for the tanks, where the anchorage is known, it is
found that a vast majority of them were unanchored. The damage states are described in
Table 4, and were chosen very similar to the ones of the HAZUS methodology.
Using a logistic regression to fit the data, the authors convert it into the following functional
form (lognormal distribution shown in ( Equation 17 )):

Ç1 Ã xÔ
P*ds x + ? h È lnÄÄ ÕÕÙ ( Equation 17 )
É d Å µ ÖÚ

The parameters µ and are specified in the table below, for each tank configuration:

Table 18 Median and standard deviation parameters for the fragility curves proposed
by O'Rourke & So (2000)

All tanks H/D < 0.70 H/D > 0.70 % Full < 50% % Full > 50%
µ(g) µ(g) µ(g) µ(g) µ(g)
DS2 0.70 0.48 0.67 0.50 0.45 0.47 0.64 0.41 0.49 0.55
DS3 1.10 0.35 1.18 0.34 0.69 0.32 - - 0.86 0.39
DS4 1.29 0.28 1.56 0.35 0.89 0.21 - - 0.99 0.27
DS5 1.35 0.22 1.79 0.29 1..07 0.15- - - 1.17 0.21

In the case of tanks that are less than 50% full, only the fragility curve of DS2 is defined: this
is due to the lack of consistent data for higher damages.

ALA, 2001

The ALA study is based on a damage inventory of 424 on-grade steel tanks carried out by
Cooper (1997), plus additional data from other earthquake events. During the 1989 Loma
Prieta earthquake, around 1670 tanks were exposed to relatively low levels of ground motion
(between 0.03 and 0.1g): hardly any of these tanks suffered some damage, therefore they
were not included in the analysis. As a result, 532 tanks, which experienced strong ground
motions of 0.1g or higher, were used to develop empirical fragility curves. For each data
point, the damage description in Table 6 is used to set the corresponding damage state.
A typology distinction is made depending on the percentage of stored contents and the
anchorage of the tank to the baseplate. The parameters µ and (of a lognormal distribution,
( Equation 17 )) of the corresponding fragility curves are given in Table 19.

48
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

Table 19 Median and standard deviation parameters for the fragility curves proposed
by ALA (2001)

µ(g) µ(g) µ(g) µ(g)


DS2 0.56 0.80 0.18 0.80 0.71 0.80 0.15 0.12
DS3 >2.00 0.40 0.73 0.80 2.36 0.80 0.62 0.80
DS4 - - 1.14 0.80 3.72 0.80 1.06 0.80
DS5 - - 1.16 0.80 4.26 0.80 1.13 0.10
% Full < 50% % Full > 50% % Full > 50% % Full > 50%
All All Anchored Unanchored
N=95 N=251 N=46 N=201

Like O’Rourke et So (2000), the ALA study concludes that the tanks that are less than half-
full did not experience enough damage to compute fragility curves for DS4 and DS5. Thus,
only the tanks with a fill percentage higher than 50% were considered to estimate additional
curves, based on the anchorage of tanks.

Berahman & Behnamfar, 2007

This study has already been introduced in sub-section 3.2.2: the authors develop fragility
curves for unanchored on-grade steel tanks, using a Bayesian approach. Using the same
damage states definitions and the same tanks database as ALA (2001), fragility curves for
DS2, DS3 and DS4 are developed: no functional form is available, and the curves are
displayed in the figures below. Due to a too small amount of data for DS5, this curve has not
been developed by the authors.

(g)

Figure 21 Fragility curves of Berahman & Behnamfar (2007), for DS2

49
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

(g)

Figure 22 Fragility curves of Berahman & Behnamfar (2007), for DS3

(g)

Figure 23 Fragility curves of Berahman & Behnamfar (2007), for DS4

The “predictive” fragility curve is the one to consider, as it accounts for epistemic
uncertainties (model parameters are then treated as random variables, in the same manner
as aleatory variables). Also, the authors discuss on whether to exclude the 1994 Northridge
earthquake data: indeed, this earthquake generated very few damages given the ground
motion level, and it is found that the data without Northridge provides a better fit. However, in
order to be coherent with other studies like O’Rourke & So (2000) or ALA (2001), the curves
presented above finally include the Northridge earthquake.

50
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

Iervolino et al., 2004

This analytical approach, already presented in sub-section 3.2.3, focuses on unanchored


steel tanks failure by elephant’s foot buckling. The results (µ and parameters of the
lognormal distribution) are displayed as a response surface, accounting for two effects:
- fluid height-over-radius ratio;
- friction coefficient between the tank and the baseplate;
It is of course found that the tank’s vulnerability increases with the level of stored liquid. The
response surfaces for both µ and are presented below (respectively Figure 24 and Figure
25).

Figure 24 Fitted surface for fragility median µ

51
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

Figure 25 Fitted surface for fragility standard deviation

HAZUS, 2004

The HAZUS methodology proposes fragility curves for “tanks farms”, accounting also for the
fragility of the equipment needed for the tank facilities to function properly:
- electric power (commercial or backup generator);
- tank;
- elevated pipes;
- electrical/mechanical components;
These components are then assembled in the logic fault-tree below (Figure 26):

52
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

Figure 26 Fault-tree analysis proposed by HAZUS (NIBS, 2004) to assess the


vulnerabiltiy of tank farms

The fragility curves of each component are defined by the median and standard deviation
parameters (lognormal distribution) in Table 20:

Table 20 Fragility curves of components (anchored or unanchored) of tank farms


components, according to HAZUS (NIBS, 2004)

Components Damage Anchored Unanchored


state
µ(g) µ(g)
Electric Power (Backup) minor 0.80 0.60 0.20 0.60
moderate 1.00 0.80 0.40 0.80
Loss of commercial Power minor 0.15 0.40 0.15 0.40
moderate 0.30 0.40 0.30 0.40
Electrical/ Mechanical Equipment moderate 1.00 0.60 0.60 0.60
Steel tank minor 0.30 0.60 0.15 0.70
moderate 0.70 0.60 0.35 0.75
extensive 1.25 0.65 0.68 0.75
complete 1.60 0.60 0.95 0.70
Elevated pipes extensive 0.53 0.60 0.53 0.60
complete 1.00 0.60 1.00 0.60

Finally, using the fault-tree analysis and the fragility parameters of the sub-components, it is
possible to propose fragility curves for steel tank farms, with anchored or unanchored
components. Table 21 shows the parameters of the corresponding lognormal distributions:

53
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

Table 21 Fragility parameters (lognormal distribution) for steel tank farms, according
to HAZUS (NIBS, 2004)

Typology Damage state µ(g)


Tank farm with slight / minor 0.29 0.55
anchored
moderate
components 0.50 0.55
extensive
complete 0.87 0.50
Tank farm with slight / minor 0.12 0.55
unanchored
moderate 0.23 0.55
components
extensive 0.41 0.55
complete 0.68 0.55

5.1.3 Processing facilities

LESSLOSS, 2007

Concerning pumping / compressor stations, a study by LESSLOSS (2007) and Kappos et al.
(2006) has proposed an hybrid approach (building fragility curve and Boolean logic tree for
the components) to develop fragility curves for these facilities (in R.C. low rise buildings,
anchored components). The parameters of the lognormal distribution are given in the table
below:

Table 22 Median and standard deviation parameters of the fragility curves for
pumping plants (LESSLOSS, 2007)

Typology Damage state µ(g)


Pumping plants with anchored slight / minor 0.10 0.55
subcomponents and RC1.1.X.X building
moderate 0.15 0.55
(1-3 floors, with low level seismic codes)
extensive 0.30 0.70
complete 0.40 0.75
Pumping plants with anchored slight / minor 0.15 0.30
subcomponents and RC1.1.X.Y building
moderate 0.30 0.35
(1-3 floors, with advanced seismic codes)
extensive 1.10 0.55
complete 2.10 0.70

54
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

SRMLIFE, 2003-2007

More recently, the SRMLIFE (2003-2007) Greek project resulted in a new set of fragility
curves: also based on a fault-tree analysis of various components (commercial power,
backup generators, pump, electrical/mechanical equipment…), the fragility curves have been
developed for pumping / compressor stations (RC low-rise building, anchored components).
The fragility parameters (median and standard deviation) of the components have been
taken from the HAZUS methodology, as shown in the table below:

Table 23 Parameters of fragility curves for stations components taken from (NIBS,
2004)

Components Damage State µ(g)


Electric Power (Backup) minor 0.80 0.60
moderate 1.00 0.80
Loss of commercial Power minor 0.15 0.40
moderate 0.30 0.40
Vertical/ extensive 1.25/1.60 0.60
Horiz. Pump
Electrical/ Mechanical Equipment moderate 1.00 0.60
Building ( 1.1. . ) minor 0.28 0.733
moderate 0.72 0.733
extensive 1.66 0.733
complete 2.17 0.733

Then, using the fragility functions of all these components and the fault-tree analysis
described in section 3.2.4 (Figure 10), it is possible to establish global fragility curves for the
gas station (Table 24):

Table 24 Median and standard deviation parameters of the fragility curves (lognormal
distribution) for pumping plants (SRMLIFE, 2003-2007)

Typology Damage state µ(g)


Anchored minor 0.30 0.70
components, RC
moderate 0.55 0.45
low-rise building
(advanced code) extensive 0.80 0.50
complete 2.20 0.70

55
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

Risk-UE and HAZUS, 2004

Finally, the Risk-UE project (Alexoudi & Pitilakis, 2003) has used vulnerability models for the
components from the HAZUS methodology (NIBS, 2004), (parameters given in Table 20),
resulting in the following fragility curves parameters values for pumping plants with anchored
or unanchored components (Table 25), based on a fault-tree analysis of the structure of the
facility.

Table 25 Median and standard deviation parameters of the fragility curves (lognormal
distribution) for pumping plants, proposed by the HAZUS methodology (NIBS, 2004)

Typology Damage state µ(g)


Anchored Minor 0.15 0.75
components
Moderate 0.34 0.65
Extensive 0.77 0.65
Complete 1.50 0.80
Unanchored Minor 0.12 0.60
components
Moderate 0.24 0.60
Extensive 0.77 0.65
Complete 1.50 0.80

The Risk-UE fragility curves (taken from NIBS, 2004) are represented below, for both
anchored and unanchored components:

Figure 27 Fragility curves for gas pumping / compression stations (anchored


components) proposed by Risk-UE (Alexoudi & Pitilakis, 2003)

56
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

Figure 28 Fragility curves for gas pumping / compression stations (unanchored


components) proposed by Risk-UE (Alexoudi & Pitilakis, 2003)

Other components

Regarding pressure reduction groups, in-line valves and SCADA systems, no quantitative
fragility curves are available at present.

5.2 VALIDATION / ADAPTATION / IMPROVEMENT

5.2.1 Pipeline components

According to the available typologies for gas & oil pipelines in Europe (mostly welded-steel,
PVC and HDPE continuous ductile pipes), it seems that the empirical relations found in the
literature should be satisfying in the scope of the SYNER-G project.
With the criterion of ductile pipes and the use of PGV and PGD as respective intensity
indexes for ground shaking and ground failures, the following relations constitute good
candidates for the final proposal:
- For wave propagation:
o O’Rourke & Ayala, 1993 (HAZUS);
o Eidinger et al., 1995, 1998;
o Isoyama et al., 2000;
o ALA, 2001;
- For permanent ground deformation:
o Eguchi, 1983;

57
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

o Honegger & Eguchi, 1992 (HAZUS);


o Eidinger & Avila, 1999;
o ALA, 2001;
Some of these relations have been tested and confronted to a European case study (2003
Lefkas earthquake), however only for water distribution pipelines: therefore these results
may not apply to the specific case of gas & oil pipelines.

5.2.2 Storage tanks

The typology of European atmospheric storage tanks may be mostly on-grade steel tanks
with anchored or unanchored components. This type should be well covered by the existing
fragility curves. Thus, we will consider the following studies, which apply to the specific
typologies and propose relevant damage states:
- O’Rourke & So, 2000;
- ALA, 2001;
- HAZUS (NIBS, 2004);

5.2.3 Processing facilities

For pumping / compressor stations, the SRMLIFE (2003-2007) study covers well the Greek
typology of gas stations. Yet, some possible typologies seem to be missing, like the pumping
plants in masonry buildings, and the case of unanchored components (for both RC and
masonry buildings). Thus, we need to develop these additional curves, based on the similar
fault-tree analysis and the use of HAZUS (NIBS, 2004) fragility parameters for the station’s
components.
Also, the Re.Mi cabins and the GRF reduction groups located in the L’Aquila area appear to
be very specific and an on-going project between AMRA and ENEL gas network may
provide for the adequate fragility curves.

5.3 FINAL PROPOSAL

5.3.1 Pipeline components

Ground shaking

The ALA (2001) study is the most recent one, as the HAZUS curves are still based on the
O’Rourke & Ayala (1993) study. Thus, the ALA (2001) are based on the largest set of
empirical data, including the 1994 Northridge earthquake: 18 events are used, as opposed to
6 in the O’Rourke & Ayala (1993) study. Moreover, the data from ALA (2001) comprises the
study from O’Rourke & Ayala (1993) and is enriched with other datasets.
In the ALA (2001) study, a balanced sample of U.S., Central American and Japanese
earthquake is used, accounting for the variability of pipeline codes among various countries.
Also, the consequent amount of data points (81, as opposed to 11 in O’Rourke & Ayala

58
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

(1993)) allows for a more balanced distribution of pipeline typologies, as shown in the table
below:

Table 26 Pipe material and pipe diameter categories included in the dataset for the
(ALA, 2001) fragility relation, according to (Tromans, 2004)

Pipeline Category Description % of total


characteristic database
Material type AC asbestos-cement 12.3
CI cast-iron 46.9
CP concrete 2.5
DI ductile iron 11.1
MX mixed (CI & DI) 11.1
S steel 16.0
Diameter DS distribution system (small diameter) 70.3
LG large diameter (> 30.48 cm) 9.9
SM small diameter (< 30.48 cm) 19.8

Moreover, the study by Tromans (2004) offers to compare some of the existing empirical
relations: these curves are plotted on the same figure below, by assuming a corrective factor
K=1 (use of the “backbone curves”).

Figure 29 Comparison of the pipeline fragility relations for PGV. Arrows refer to the
range of applicability of a given relation, approximated from knowledge of the dataset
from which it wa derived (Tromans, 2004)

As stated by O’Rourke (1999), the fragility relation by O’Rourke & Ayala (1993) seems to be
over-conservative, with pipeline repair rates being unduly affected by the long durations of

59
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

ground shaking experienced during the 1985 Michoacan earthquake (Tromans, 2004). The
relations by ALA (2001) and Isoyama et al. (2000) offer the longest applicability range, as
opposed to the O’Rourke & Ayala (1993) and Eidinger et al. (1995) relations, which should
not be extrapolated to large values of PGV.
The use of ALA (2001) and Isoyama et al. (1998, 2000) is also advocated by some validation
studies carried out on the 1999 Düzce and 2003 Lefkas earthquakes: “Based on these
validations it was found that the ALA (2001a,b) relationships for water/waste-water system
and Isoyama et al. (1998) for gas system are most suitable for the European distinctive
features.” (Pitilakis et al., 2006).
To conclude, we propose to sum up some comments from Tromans (2004) on the relevance
of existing fragility curves:
- the relation of O’Rourke et al. (1998) are to be used specifically in the U.S., as data
from other locations have not been included: moreover, this relation should only be
applied to cast-iron pipes;
- the Isoyama et al. (1998, 2000) relations are suggested for Japan. Yet, application to
other locations is difficult, due to the specific topographic classification scheme, which
is not normally used outside of Japan (see Table 13);
- for general application, the relation of ALA (2001) is recommended, as it is derived
from a global database;
The study of Tromans (2004) includes also an estimation of the correlation coefficient r² for
each empirical relation: it appears that r² is very low for the ALA (2001) curve, stressing a
huge scatter. However, such a scatter in the data can be expected, as the data points are
selected from various configurations: the proposed relation allows then many of the
contributors of this scatter to be accounted for, using various corrective factors (see Table
16).
Based on this discussion, we finally propose to adopt the following fragility curve from ALA
(2001):

RR ? K1 0.002416 PGV ( Equation 18 )

Where RR is the repair rate per km, and PGV is given in cm/s.
The following figure represents the adequate fragility curves to be used for each pipe
typology considered in the scope of SYNER-G (see Table 2):
- Greek transmission lines: welded-steel, small diameter => K1=0.6;
- Greek distribution lines, Austrian transmission and distribution lines, L’Aquila
distribution lines: PVC, small diameter => K1=0.5;
- Austrian supra-regional lines: welded-steel, large diameter => K1=0.15;

60
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

Figure 30 Proposed fragility curves for the most common gas & oil pipeline
typologies (ALA, 2001), for wave propagation

It is important to note that L’Aquila distribution lines are made of HDPE: this material is not
well studied in the current literature and we make the assumption that these pipelines can be
associated with the PVC fragility curves, although some reports show that HPDE pipes tend
to behave better than PVC.

Ground failure

Like for wave propagation-related damage, the relation of ALA (2001) is the most recent
one, as the one proposed by the HAZUS methodology is taken from Honegger & Eguchi
(1992). The dataset from ALA (2001) comprises 41 data points from 4 earthquakes (one
Japanese and three U.S.), with liquefaction as the main failure mechanism.
Thus, the ALA (2001) curve is based on the most complete empirical data and we propose
to compare some fragility curves (use of the “backbone curve”, without any corrective
factors), in the figure below:

61
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

Figure 31 Comparison of three fragility curves, with respect to PGD

The figure above shows important discrepancies between the different studies: the curve of
ALA (2001) lies in between the relations from Honegger & Eguchi (1992) and Eidinger &
Avila (1999). Based on this discussion and in order to be coherent with the fragility curve
selected for transient ground motion, we finally propose to adopt the following fragility curve
from ALA (2001):

RR ? K 211.2238PGD 0.319 ( Equation 19 )

Where RR is the repair rate per km, and PGD is given in m.


The following figure represents the adequate fragility curves to be used for each pipe
typology considered in the scope of SYNER-G (see Table 2):
- Greek transmission lines, L’Aquila transmission and distribution lines: welded-steel,
small diameter => K2=0.7;
- Greek distribution lines, Austrian transmission and distribution lines, L’Aquila
distribution lines: PVC, small diameter => K2=0.8;
- Austrian supra-regional lines: welded-steel, large diameter => K2=0.15;

62
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

Figure 32 Proposed fragility curves for the most common gas & oil pipeline
typologies (ALA, 2001), for permanent ground deformation

5.3.2 Storage tanks

The studies by O’Rourke & So (2000) and ALA (2001) are the most thorough, as they allow
for distinction between many characteristics such as:
- % of contents stored;
- anchored or unanchored components;
- height-over-radius ratio;
However, some of the proposed fragility curves are based on really scarce empirical data,
and this may raise issues on the reliability of the data. Also, the damage states proposed by
these two studies are mostly defined by physical damage mechanisms that prove difficult to
link to any loss of functionality.
Besides, oil storage tanks are located in very complex facilities (e.g. refineries, storage
facilities…) and consider only the damage to the tank body seems to be a quite simplistic
and rather unconservative approach: indeed, the whole system of the “tank farm” should be
accounted for, including elevated pipes, power sources, mechanical equipment…
We propose then to adopt the fragility curves for “tank farms” developed in the HAZUS
methodology (NIBS, 2004) (Table 27 and Figure 33). These curves can be applied to on-
grade steel tank, with a distinction on whether components are anchored or not.

63
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

Table 27 Fragility parameters for steel tank farms, according to HAZUS (NIBS, 2004)

Typology Damage state µ(g)


Tank farm with slight / minor 0.29 0.55
anchored
moderate
components 0.50 0.55
extensive
complete 0.87 0.50
Tank farm with slight / minor 0.12 0.55
unanchored
moderate 0.23 0.55
components
extensive 0.41 0.55
complete 0.68 0.55

The parameters µ and are respectively the median and standard deviation of the
lognormal distribution of probability, with respect to PGA.

Figure 33 Fragility curves for steel tank farms (NIBS, 2004)

64
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

5.3.3 Processing facilities

Greek pumping /compressor plants

The project SRMLIFE (2003-2007) has developed fragility curves specifically applicable to
Greek gas stations, which consist of low-rise masonry buildings with anchored components
(Figure 34 and Table 28).

Table 28 Fragility parameters for Greek pumping plants, according to (SRM-LIFE,


2003-2007)

Typology Damage state µ(g)


Anchored minor 0.30 0.70
components, RC
moderate 0.55 0.45
low-rise building
(advanced code) extensive 0.80 0.50
complete 2.20 0.70

The parameters µ and are respectively the median and standard deviation of the
lognormal distribution of probability, with respect to PGA.

Figure 34 Fragility curves for Greek pumping / compressor plants (SRMLIFE, 2003-
2007)

65
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

Generic pumping / compressor plants

Apart from the Greek context, the typology of European gas stations is not well known, and
one solution could be to use the generic fragility curves of the HAZUS methodology (NIBS,
2004), which are based only on the distinction between anchored and unanchored
components (Table 29 and Figure 35).

Table 29 Fragility parameters of the fragility curves for pumping plants, proposed by
the HAZUS methodology (NIBS, 2004)

Typology Damage state µ(g)


Anchored Minor 0.15 0.75
components
Moderate 0.34 0.65
Extensive 0.77 0.65
Complete 1.50 0.80
Unanchored Minor 0.12 0.60
components
Moderate 0.24 0.60
Extensive 0.77 0.65
Complete 1.50 0.80

The parameters µ and are respectively the median and standard deviation of the
lognormal distribution of probability, with respect to PGA.

Figure 35 Fragility curves for generic pumping / compressor plants (NIBS, 2004)

66
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

Re.Mi cabins (L’Aquila)

For this specific component, we can use preliminary results from the ongoing project
between AMRA and Enel Rete Gas on the l’Aquila gas system.
The fault-tree decomposition has been applied to Re.Mi cabins in order to identify which sub-
component is critical with respect to seismic fragility. All sub-components are considered to
be not anchored and simply supported on the ground (with the exception of bowls located in
a seperated area that are ceiling-mounted). The Re.Mi station is decomposed into the
following sub-components:
- Building;
- Regulators;
- Mechanical equipment (heat exchangers, boilers and bowls);
To compute the global fragility curve of the whole plant, the following faut-tree may be
formulated:

Figure 36 Fault-tree analysis of a Re.Mi cabin according to (Esposito et al., 2011)

Since gas supply has to be maintained at all times, two installations are mounted in parallel
where each installation is characterized by a regulator and a monitor. The monitor is a safety
device that has to be able to prevent the outlet pressure from exceeding safe thresholds in
the case of complete failure of the regulator, taking over the function of the primary, normally
active regulator. During normal operation the monitor is fully open and if the pressure
becomes equal to the setpoint of the monitor, the monitor will close to constrain the
pressure.
According to experts’ experience, two damage states have been identified: complete and
extensive which correspond to a complete loss of functionality of the system (no gas supply).
The following table (Table 30) summarizes the damage states and their description.

Table 30 Damage states for the Re.Mi cabin

Damage state Description

Complete Building collapse


Extensively damaged building, or both
Extensive
regulators damaged, or badly damaged
boilers/heat alimentation

67
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

In particular, when boilers break down the gas flow is not ensured, since the freezing stops
the system.

GRF Reductions groups (L’Aquila)

Reduction groups are decomposed in the following sub-components:


- Regulators;
- Masonry housing (when it is present, otherwise within a kiosk);
To compute the global fragility curve of the whole plant, the following fault-tree has been
formulated:

Figure 37 Fault-tree analysis of a Reduction Group (GR / GRM) according to (Esposito


et al., 2011)

In most cases the safety device is ensured by the presence of shut-off valves that are able to
block the gas flow. When the pressure exceeds a maximum value, the valves close.
Also in this case two damage states have been identified and are summarized in the
following table:

Table 31 Damage states of the Reduction Group (GR / GRM)

Damage state Description

Complete Building collapse (mansory housing)


Extensively damaged housing, or both
Extensive
regulators damaged.

However, some Reduction Groups do not have the second regulator and this characteristic
implies a higher vulnerability.

68
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

6 Analytical expressions of fragility functions

6.1 PIPELINE COMPONENTS

6.1.1 Wave propagation

Intensity index: PGV


Fragility curve (ALA, 2001):

RR ? K1 0.002416 PGV ( Equation 20 )

Where RR is the repair rate per km, and PGV is given in cm/s.

Table 32 Values of corrective factor K1 (ALA, 2001)

pipe material joint type soil diameter K1


CI cement unknown small (<30cm) 1.0
cast-iron cement corrosive small 1.4
cement non corrosive small 0.7
rubber gasket unknown small 0.8
WS arc welded unknown small 0.6
welded-steel arc welded corrosive small 0.9
arc welded non corrosive small 0.3
arc welded all large (>30cm) 0.15
rubber gasket unknown small 0.7
screwed all small 1.3
riveted all small 1.3
AC rubber gasket all small 0.5
asbestos- cement all small 1.0
cement
C welded all large 0.7
concrete cement all large 1.0
rubber gasket all large 0.8
PVC rubber gasket all small 0.5
DI rubber gasket all small 0.5
ductile iron

69
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

6.1.2 Permanent ground deformation

Intensity index: PGD


Fragility curve (ALA, 2001):

RR ? K 2 2.5831PGD 0.319 ( Equation 21 )

Where RR is the repair rate per km, and PGD is given in cm.

Table 33 Values of corrective factor K2 (ALA, 2001)

pipe material joint type K2


Unknown Unknown 1.0
CI cement 1.0
rubber gasket 0.8
mechanical restrained 0.7
WS arc-welded, lap welds 0.15
rubber gasket 0.7
AC rubber gasket 0.8
cement 1.0
Concrete welded 0.6
cement 1.0
rubber gasket 0.7
PVC rubber gasket 0.8
DI rubber gasket 0.5

70
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

6.2 STORAGE TANKS

Intensity index: PGA


Fragility curve (HAZUS, 2004):

1Ç Ã ln PGA / ln µ Ô
P*D DSk + ? È1 - erf Ä ÕÙ ( Equation 22 )
2 ÈÉ Å d 2 Ö ÙÚ

Table 34 Fragility parameters for steel tank farms (HAZUS, 2004)

Typology Damage state µ(g)


Tank farm with slight / minor 0.29 0.55
anchored
moderate
components 0.50 0.55
extensive
complete 0.87 0.50
Tank farm with slight / minor 0.12 0.55
unanchored
moderate 0.23 0.55
components
extensive 0.41 0.55
complete 0.68 0.55

Table 35 Damage states definitions for tank farms (HAZUS, 2004)

Damage state Description


DS1 none Fully functional
DS2 slight / minor Malfunction of tank farm for a short time (less than thee days)
due to loss of backup power or light damage to tanks
DS3 moderate Malfunction of tank farm for a week or so due to loss of
backup power, extensive damage to various equipment , or
considerable damage to tanks
DS4 extensive Extensive damage to tanks or elevated pipes
DS5 complete Complete failure of all elevated pipes, or collapse of tanks

71
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

6.3 PROCESSING FACILITIES

6.3.1 Pumping / compressor stations

Intensity index: PGA


Fragility curve (HAZUS, 2004) and (SRMLIFE, 2003-2007):

1Ç Ã ln PGA / ln µ Ô
P*D DSk + ? È1 - erf Ä ÕÙ ( Equation 23 )
2 ÉÈ Å d 2 Ö ÚÙ

Table 36 Fragility parameters for pumping / compressor stations (HAZUS, 2004) and
(SRMLIFE, 2003-2007)

Typology Damage state µ(g)


Greek typology: minor 0.30 0.70
Anchored
moderate 0.55 0.45
components, RC
low-rise building extensive 0.80 0.50
(advanced code)
complete 2.20 0.70
“Generic stations”, Minor 0.15 0.75
Anchored
Moderate 0.34 0.65
components
Extensive 0.77 0.65
Complete 1.50 0.80
“Generic stations”, Minor 0.12 0.60
Unanchored
Moderate 0.24 0.60
components
Extensive 0.77 0.65
Complete 1.50 0.80

72
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

Table 37 Damage states definitions for pumping / compressor stations (HAZUS, 2004)
and (SRMLIFE, 2003-2007)

Repair cost Serviceability/


Damage state
(%) Functionality
D1 No - - Normal
D2 1 - 10 function

Several
Slight damage to building stops and
or full loss of commercial Full function
Slight/Minor reduced flow
power and backup power 10 – 30 of gas in the
for few days (< 3 days) transmission
gas
pipelines
D3 Considerable damage to
mechanical and electrical Malfunction.
equipment or (Full
Moderate considerable damage to 30 – 50 function
building or loss of electric after
Disability of
power and of backup for repairs)
boosting gas
7 days.
in
D4 Building being compression
extensively damaged, or station Full loss of
Extensive 50 – 75 function
the pumps badly
damaged beyond repair. (un
repairable
D5 Complete damage)
Building collapsed 75 – 100
damage

6.3.2 Re.Mi cabins

As a first approximation, we are compelled to use the fragility curves from HAZUS (NIBS,
2004) for pumping / compressor stations (“Generic stations” in Table 39), as the fragility of
those specific subcomponents (e.g. regulators) is not well known.

6.3.3 GRF reduction groups

As a first approximation, we are compelled to use the fragility curves from HAZUS (NIBS,
2004) for pumping / compressor stations (“Generic stations” in Table 39), as the fragility of
those specific subcomponents (e.g. regulators) is not well known.

73
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

References

Adachi, T. 2007. Impact of cascading failures on performance assessment of civil


infrastructure systems. PhD Thesis, School of Civil and Environmental Engineering,
Georgia Institute of Technology.
ALA. 2001. Seismic fragility formulations for water systems. American Lifeline Alliance,
ASCE.
Alexoudi, M., and K. Pitilakis. 2003. Vulnerability assessment of lifelines and essential
facilities (WP06): methodological handbook – Appendix 7: gas utility system. Risk-UE
Report n° GTR-RSK 0101-152av7.
ASCE. 1984. Guidelines for the seismic design of oil and gas pipeline systems. Technical
Council on Lifeline Earthquake Engineering, prepared by the Committee on Gas and
Liquid Fuel Lines, ASCE, 473 pp.
ASCE. 1987. The effects of earthquakes on power and industrial facilities and implications
for nuclear power plant design. Committee on Dynamic Analysis of the Committee on
Nuclear Structures and materials of the Structural Division, ASCE.
Ayala, G., and M. J. O'Rourke. 1989. Effects of the 1985 Michoacan Earthquake on Water
Systems in Mexico. Technical Report NCEER-89-0009, National Center for Earthquake
Engineering Research, State University of New York at Buffalo, Buffalo, NY.
Ballantyne, D. B., and W; Heubach. 1996. Earthquake loss estimation for the city of Everett,
Washington, lifelines. K/J/C 906014.00, Federal Way, Washington.
Barenberg, M.E. 1988. Correlation of pipe damage with ground motion. Journal of
Geotechnical Engineering 114(6): 706-711.
Berahman, F., and F. Behnamfar. 2007. Seismic fragility curves for unanchored on-grade
steel storage tanks: bayesian approach. Journal of Earthquake Engineering 11(2): 1-31.
Bettinger, R.V. 1980. Economic and seismic mitigation for gas and electric utility. In
Conference on Social and Economic Impact of Earthquakes on Utility Lifelines. Edited by
J. Isenberg, ASCE, pp. 98-106.
Cooper, T. W. 1997. A study of the performance of petroleum storage tanks during
earthquakes, 1933-1995. Prepared for the National Institute of Standards and
Technology, Gaithersburg, MD.
Crespo, E., K. J. Nyman, T. D. O’Rourke. 1987. Ecuador earthquakes of March 5, 1987.
Earthquake Engineering Research Institute, EERI Newsletter 21(7):1-4.
EERI. 1986. Reducing earthquake hazards: lessons learned from earthquakes. Publication
n°86-02, Earthquake Engineering Research Institute, El Cerrito, CA.
EERI. 1990. Loma Prieta earthquake reconnaissance report. Supplement to vol. 6, Technical
Editor, Lee Benuska, Earthquake Spectra, Journal of the Earthquake Engineering
Research Institute.
Eguchi, R. T., M. R. Legg, C. E. Taylor, L. L. Philipson, J. H. Wiggins. 1983. Earthquake
Performance of Water and Natural Gas Supply System. J. H. Wiggins Company, NSF
Grant PFR-8005083, Report 83-1396-5.
Eguchi, R. T. 1987. Seismic risk to natural gas and oil systems. FEMA 139, Earthquake
Hazard Reduction Series 30, pp. 15-33.
Eguchi, R. T. 1991. Early post-earthquake damage detection for underground lifelines. Final

75
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

Report to the national Science Foundation, Dames and Moore PC, Los Angeles, CA.
Eidinger J., B. Maison, D; Lee, B. B. Lau. 1995. East Bay municipality utility district water
distribution damage in 1989 Loma Prieta earthquake. In Fourth US Conference on
Lifeline Earthquake Engineering, Monograph 6, ASCE, pp. 240-247.
Eidinger, J. 1998. Lifelines, Water Distribution System. in The Loma Prieta, California,
Earthquake of October 17, 1989, Performance of the Built Environment –Lifelines. US
Geological Survey Professional Paper 1552-A, pp A63-A80, A. Schiff Ed.
Eidinger, J., and E. Avila. 1999. Guidelines for the seismic upgrade of water transmission
facilities. ASCE, TCLEE, Monograph 15.
Esposito, S., S. Giovinazzi, I. Iervolino. 2011. Assessing the post-earthquake residual
functionality of gas networks and planning for restoration. In 2011 Pacific Conference on
Earthquake Engineering,14-16 April, New Zealand.
Fabroccino, G., I. Iervolino, F. Orlando, E. Salzano. 2005. Quantitave risk analysis of oil
storage facilities in seismic areas. Journal of Hazardous Materials 23: 61-69.
Hall, W. J. 1987. Earthquake engineering research needs concerning gas and liquid fuel
lifelines. FEMA 139, Earthquake Hazard Reduction Series 30, pp. 35-49.
Hamada, M. 1991. Estimation of earthquake damage to lifeline systems in Japan. In Third
Japan-US Workshop on earthquake resistant design of lifeline facilities and
countermeasures for soil liquefaction, San Francisco, CA.
Honegger, D. G., and R. T. Eguchi. 1992. Determination of the relative vulnerabilities to
seismic damage for San Diego County Water Authority: water transmission pipelines.
Iervolino, I., G. Fabbrocino, G. Manfredi. 2004. Fragility of standard industrial structures by a
response surface based method. Journal of Earthquake Engineering 8(6): 927-945.
Isoyama, R., and T. Katayama. 1982. Reliability evaluation of water supply systems during
earthquake. Report of the Institute of Industrial Science, University of Tokyo, vol. 30(1).
Isoyama, R., E. Ishida, K. Yune, T. Shirozu. 2000. Seismic damage estimation procedure for
water supply pipelines. In Twelfth World Conference on Earthquake Engineering, CD-
ROM paper n°1762, 8 pp.
Kappos, A., G. Panagopoulos, Ch. Panagiotopoulos, Gr. Penelis. 2006. A hybrid method for
the vulnerability assessment of RC and URM buildings. Bulletin of Earthquake
Engineering 44: 391-413.
Katayama, T., K. Kubo, N. Sato. 1975. Earthquake damage to water and gas distribution
systems. In US National Conference on Earthquake Engineering, Oakland, CA, EERI,
pp. 396-405.
Kennedy, R. P., and R. P. Kassawara. 1989. Seismic evaluation of large flat-bottomed tanks.
In Second Symposium on Current Issues Related to Nuclear Power Plant Structures,
Equipment, and Piping with Emphasis on Resolution of Seismic Issues in Low-Seismicity
Regions, EPRI NP-6437-D.
Kim, J., S. O’Connor, S. Nadukuru, J. P. Lynch, R. Michalowski, R. A. Green, M. P. Ghaz,
W. J. Weiss, A. Bradshaw. 2010. Behavior of full-scale concrete segmented pipelines
under permanent ground displacements. In Conference on Health Monitoring of
Structural and Biological Systems, San Diego, CA.
LESSLOSS. 2007. Damage scenarios for selected urban areas (for water and gas systems,
sewage mains, tunnels and waterfront structures: Thessaloniki, Istanbul (European side),
Düzce. LESSLOSS Project Deliverable n°117.
NIBS. 2004. Earthquake Loss Estimation Methodology HAZUS. National Institute of Building
Sciences, FEMA, Washington D.C.H
NZNSEE. 1986. Seismic design of storage tanks. Recommendations of a Study Group of the
New Zealand National Society for Earthquake Engineering.

76
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

Ogawa, N. 1983. A vibration test of earthquake induced hydraulic transients of liquid-filled


pipelines. In Symposium on Lifeline Engineering, Earthquake Behavior and Safety of Oil
and Gas Storage Facilities, Buried Pipelines and Equipment. PVP 77, ASME.
O’Rourke, M. J. 1988. Mitigation of seismic effects on water systems. In Seismic Design and
Construction of Complex Civil Engineering Systems. Symposium sponsored by
TCLEE/ASCE, National Convention, St Louis, MO, pp. 65-78.
O’Rourke, M. J., and G. Ayala. 1990. Seismic damage to pipeline: case study. Journal of
Transportation Engineering 116(2): 123-134.
O’Rourke, M. J., and G. Ayala. 1993. Pipeline damage due to wave propagation. Journal of
Geotechnical Engineering 119(9):1490-1498.
O'Rourke, M. J., and X. Liu. 1999. Response of Buried Pipelines Subject to Earthquake
Effects. MCEER Monograph No. 3.
O’Rourke, M. J. 1999. Estimation of post-earthquake water system serviceability. In Seventh
Japan-US Workshop on Earthquake Resistant Design of Lifeline Facilities and
Countermeasures for Soil Liquefaction, MCEER, State University of New York at Buffalo,
Buffalo, NY, 391-403.
O’Rourke, M. J., and P. So. 2000. Seismic fragility curves for on-grade steel tanks.
Earthquake Spectra 16(4).
O’Rourke, M. J., and E. Deyoe. 2004. Seismic damage to segment buried pipe. Earthquake
Spectra 20(4): 1167 – 1183.
O’Rourke, T. D., and C. H. Trautmann. 1981. Earthquake ground rupture effects on jointed
pipe. In Second Specialty Conference of the Technical Council on Lifeline Earthquake
Engineering. D. J. Smith, Editor, pp. 65-80.
O’Rourke, T. D., S. Toprak, Y. Sano. 1998. Factors affecting water supply damage caused
by the Northridge earthquake. In Sixth US national Conference on Earthquake
Engineering.
O’Rourke, T. D., and S. S. Jeon. 1999. Factors affecting water supply damage caused by
the Northridge earthquake. In Fifth US Conference of Lifeline Earthquake Engineering,
Seattle, WA.
Paolucci, R., and K. Pitilakis. 2007. Seismic risk assessment of underground structures
under transient ground deformations . In Fourth International Conference on Earthquake
Geotechnical Engineering, Thessaloniki, Greece.
Pineda, O., and M. Ordaz. 2003. Seismic vulnerability function for high diameter buried
pipelines: Mexico City’s primary water system case. In International Conference on
Pipeline Engineering Construction, vol. 2, pp. 1145-1154.
Pineda, O., and M. Ordaz. 2007. A new seismic parameter to estimate damage in buried
pipelines due to seismic wave propagation. Journal of Earthquake Engineering 11(5):
773-786.
Pitilakis, K., M. Alexoudi, S. Argyroudis, O. Monge, C. Martin. 2006. Earthquake risk
assessment of lifelines. Bulletin of Earthquake Engineering 4: 365-390.
Rauch, A. F. 1997. EPOLLS: An empirical method for predicting surface displacements due
to liquefaction-induced lateral spreading in earthquakes. PhD Thesis, Virginia
Polytechnic Institute and State Univ., Blacksburg, VA.
Singhal, A. C. 1984. Nonlinear behavior of ductile iron pipeline joints. Journal of Technical
Topics in Civil Engineering 110(1).
SRMLIFE. 2003-2007. Development of a global methodology for the vulnerability
assessment and risk management of lifelines, infrastructures and critical facilities.
Application to the metropolitan area of Thessaloniki. Research Project, General
Secretariat for Research and Technology, Greece.

77
SYNER-G – D3.4 – Fragility functions for gas and oil system networks

St John, C. M., and T. F. Zahrah. 1987. Aseismic design of underground structures. TUST
2(2): 165-197.
Terzi, V. G., M. N. Alexoudi, T. N. Hatzigogos. 2007. Numerical assessment of damage state
of segmented pipelines due to permanent ground deformation. In Tenth International
Conference on Applications of Statistics and Probability in Civil Engineering, Tokyo,
Japan, Paper n°202.
Toprak, S. 1998. Earthquake effects on buried lifeline systems. PhD Thesis, Cornell
University, Ithaca, NY.
Tromans, I. 2004. Behaviour of buried water supply pipelines in earthquake zones. PhD
Thesis, Imperial College of Science, Technology and Medicine, University of London.
Young, F. M., and J. M. Pardon. 1983. Hydraulic transients in liquid-filled piping networks
due to seismic excitation. In Symposium on Lifeline Engineering, Earthquake Behavior
and Safety of Oil and Gas Storage Facilities, Buried Pipelines and Equipment. PVP 77,
ASME.

78

You might also like