You are on page 1of 48

100 Ultimate Limit Strength (ULS) of

Fixed Offshore Platforms


Abstract
This guide provides the approach to be used on Chevron projects to determine the
ultimate capacity, also called the ultimate limit strength (ULS), of fixed steel jacket
platforms. The guide is for use in worldwide metocean environments.

October 2007

Contents

Page

110

Purpose

100-3

120

The Ultimate Strength Method

100-6

121

Background

122

When to Use Ultimate Strength Methods

123

Ultimate Strength in Offshore Codes

124

ULS Software

130

Information Requirements

131

Basic Information Needed

132

Availability of Data

140

Modeling

141

Jacket Modeling

142

Foundation Modeling

143

Deck Modeling

150

Load Definition

151

Gravity and Buoyancy (Not Ramped)

152

Wave and Current (Ramped)

153

Wind (Ramped)

154

Wave Load on the Deck (Ramped)

155

P-Delta Effect

2007 Chevron USA Inc. All rights reserved.

100-13

100-17

100-24

100-1

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

100-2

160

Ultimate Strength Analysis

161

Pushover Load Approach

162

General Process for Pushover Analysis

163

Dynamic Effects

164

Collapse Definition

165

Quality Assurance (QA) of Results

170

Results

171

General

172

Pushover Curve

173

Platform ULS

174

Reference Level Loads

175

RSR

176

Identification of Collapse Mechanism

177

Graphic Sequence of Platform Collapse

178

Identification of Critical Platform Members

179

Damaged and Repaired Capacities

1710

Change in Capacity due to Additional Gravity Loading

1711

Change in Capacity due to Additional Lateral Loading

1712

Specific Member Information

180

Select References

181

Codes

182

Assessment of Existing Platforms

183

Overall Platform Modeling for ULS

190

Terminology

191

Acronyms

192

Definitions

2007 Chevron USA Inc. All rights reserved.

[BookTitle]

100-28

100-39

100-45

100-47

October 2007

[BookTitle]

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

110 Purpose
This document provides an accurate and consistent procedure for the determination
of the ULS of fixed offshore platforms. The document should be used for all ULS
computations performed either by Chevron or by consultants. The methodology is
applicable worldwide.
The procedure is applicable to new and existing platforms but, in this document, is
mostly focused on existing platforms. The focus of this document is ULS for
metocean environmental conditions. ULS for seismic conditions is similar but is not
covered in this document.
The determination of ULS is essentially a generic procedure for most structures and
is the process described in this document. The approaches described are based upon
many years of experience performing ultimate strength analysis. The descriptions
are intentionally straightforward and concise. They provide sufficient depth to
understand the approaches, including their background, without getting bogged
down in too much detail, making the document user friendly.
Figure 100-1 provides a brief description of the sections of this document, including
purpose and some details of the associated discussion. This can be used as a
roadmap when using this guide.
The document is part of the Chevron Engineering Standards (CES) developed and
maintained by the Floating and Fixed Systems Unit of the Facility Engineering
Department.

October 2007

2007 Chevron USA Inc. All rights reserved.

100-3

Summary of ULS Pushover Guide Discussions (1 of 3)

Section
120

130

Ultimate
Strength
Method

Purpose

Details

Definition of ULS and


how it is used for
offshore platform
studies.

a. The ULS of the platform is based upon a pushover analysis that loads the platform laterally until collapse.

Information
Data necessary for an
Requirements accurate ULS analysis.

b. ULS is used in the assessment of existing platforms or design of new platforms.

a. Structural Drawings:
Jacket, pile, and deck. Ideally, for the existing configuration.

2007 Chevron USA Inc. All rights reserved.

b. Geotechnical Report:
Site specific report containing data on soil strength and pile capacity.
c. Recent Inspection Reports:
Outlines existing condition of platform. Identifies damage and other data, such as marine growth, number of conductors and
risers, and boat landings and bumpers.
d. Recent Photos:
Visual confirmation of structure configuration above water. Used to confirm deck framing, deck condition, major equipment, etc.
Can be used to estimate deck elevation.
e. Equipment Weights and Layout:
Allows the accurate input of equipment loads in the correct locations.
140

Modeling

Process for building a


a. Jacket Modeling:
ULS structural computer Aspects of the model to consider are braces, legs, joint capacity, joint flexibility, conductors, risers, and appurtenances. The
model of the platform.
braces and legs are modeled with nonlinear elements.

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

100-4
Fig. 100-1

b. Foundation Modeling:
Nonlinear pile-soil springs from the geotechnical report are used to model the pile-soil interaction. Piles are modeled with
nonlinear properties.
c. Deck Modeling:
Simplified in ULS analysis, since the jacket is most critical. Consists of all primary deck members. Neglect secondary
structural members (but include their weight). The deck is typically modeled using linear elements, since it is not likely to
control platform collapse.

[BookTitle]

October 2007

Summary of ULS Pushover Guide Discussions (2 of 3)

Section
150

Load
Definition

Purpose

Details

The types of loads


applied to the ULS
model for pushover
analysis.

a. Gravity, Dead and Live Loads, and Buoyancy Load:


Constant for ULS analysis.
b. Wave and Current Load:
The dominant pushover load. The wave load recipe should comply with the most recent edition of API RP2A.
c. Wind Load:
Secondary pushover load that accounts for 5% to 15% of total base shear at ULS for most platforms. Wind loads may be larger
than wave/current loads for shallow water platforms.
d. Wave in Deck Load:
Used when the wave crest height is above the height of the bottom of steel of the cellar deck. The load should be computed in
accordance with the most recent API approach.

2007 Chevron USA Inc. All rights reserved.

e. P-Delta Load:
The effect of gravity pulling the structure over as the structure moves laterally during the pushover. This should be included in
all pushover analyses.
160

Ultimate
Strength
Analysis

Determine the
platform's ability to
resist loading.

a. 11 Step Procedure:
See Figure 100-8 for a summary of the 11 steps to determine ULS by static pushover analysis.
b. Static Pushover Analysis:
The step wise application of increasing static lateral load until the platform collapses. The ULS is maximum load that the
platform can resist in a pushover.
c. Dynamic Effects:
Considered for platforms in water depth greater than 300 ft. Uses an estimated dynamic amplification factor (DAF) and adds
that to the static base shear for the reference loads.

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

100-5

Fig. 100-1

d. Quality Assurance (QA):


An accurate understanding of the platform collapse mechanism is the most critical part of the QA process.

[BookTitle]

October 2007

Summary of ULS Pushover Guide Discussions (3 of 3)

Section
170

Results

Purpose

Details

Description of the types


of results and the types
of conclusions from the
ULS analysis.

a. Pushover Curve:
A plot of the pushover load versus the deck lateral displacement that shows the nonlinear response of the platform. The curve
should identify the key platform member failures and the ULS.
b. Platform ULS:
The maximum base shear that the platform can sustain prior to collapse.
c. Reference Level Loads:
Reference values used to determine platform acceptability. Examples include the reserve strength ratio (RSR), minimum wave
heights, and minimum base shears.
d. RSR:
Defined as the ratio of the platform ULS to the 100 yr metocean condition base shear.

2007 Chevron USA Inc. All rights reserved.

e. Identification of a Collapse Mechanism:


By examining the pushover output, the collapse mechanism can be identified. This reveals the critical platform members.
f. Damaged and Repaired Capacities:
The effects of damage can be determined by comparing the ULS with and without the damage. The effects of a repair can also
be measured similarly.
g. Changes in Capacities due to Additional Loading:
Additions to the topside weight and/or additions, such as conductors or risers, increase the loads on the platform. Their
acceptability can be determined by comparing the ULS before and after the additions. Removal of equipment, risers,
conductors, boat landings, etc., reduces load and can be similarly used to improve platform performance.
180

190

Select
References

Provides several key


references related to
ULS.

a. These references can be used to obtain further information and details about ULS analysis.

Terminology

Provides brief
definitions of ULS
terminology.

Further information on the terminology can be found in the references provided in Section 180.

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

100-6

Fig. 100-1

b. The references are organized by topic.

[BookTitle]

October 2007

[BookTitle]

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

120 The Ultimate Strength Method


121 Background
Design Load Levels
Platforms are designed to comply with design codes that use design load levels,
such as a 100 yr return period (RP) storm (wave, wind, and current), and factors of
safety that allow the structure to perform in a linear elastic manner. This is shown
conceptually in Figure 100-2 by the force displacement relationship.

Design Load
A load is applied laterally to the structure in an increasing manner, and the
deflections are measured at the top of the structure. The plot of force versus
displacement is linear up to and beyond the code based design load shown as
Fdesign. This is the desired performance of a structure to loading, in that after
loading has stopped (e.g., storm), the structure will return to its initial condition in
an elastic manner without any permanent displacement or damage. For offshore
fixed platforms, this is the design load level. For loadings up to the design load, the
platform should remain linear with no significant damage such that it can operate
safely immediately following the storm.
Fig. 100-2

Platform Force Displacement Relationship

Fmax = ULS
Fserviceability

Ductile
Collapse

Fyield

Brittle
Collapse

Force (F)

Fdesign

Linear Response

Non-Linear Response

Displacement (Delta)

October 2007

2007 Chevron USA Inc. All rights reserved.

100-7

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

[BookTitle]

Yield Load
Because design codes contain factors of safety, the structure will continue to
perform elastically, beyond the code design load Fdesign, until the factors of safety
are used up. At that time some, but not all, of the structures components first
begin to yield, shown in the figure as Fyield. However, the structural system does
not collapse, because there is still resistance provided by these members (they have
only partially failed) and because, even if the members completely failed, there are
other members that provide redundancy and continue to resist the load. Instead, the
structure begins to perform in a nonlinear manner, with a larger lateral
displacement in accordance with increment of load than in the linear range. This is
shown in Figure 100-2 by the gradual flattening of the force deformation curve. The
amount of nonlinearity is a function of material nonlinearity due to the inelastic
performance of the members and joints as they yield and fail. The nonlinearity is
also a function of geometric nonlinearity, often called P-delta, which tends to pull
the platform over as it moves laterally, as defined in Section 155.

Service Ability Limit


As the force increases in the nonlinear response range, the structure begins to incur
a permanent deflection if the storm subsides. At some point, this permanent
deformation becomes large enough that the structure may no longer be useful,
since the lateral displacement will be so large that the wells can no longer be
serviced, or the structure may be tilted and no longer capable of operations. This
is shown as Fserviceability in Figure 100-2. However, this is not always the case,
and many platforms are serviceable at displacements up to their collapse state.

Maximum Force
At some point of loading, the structure reaches the maximum load that it can resist,
shown as Fmax, and this is called the structures ULS. As described later, this is the
critical parameter in an ultimate strength analysis and is used to determine
acceptability of the platform.

CollapseGeneral
At a displacement equal to displacement associated with Fmax or at a displacement
greater than that of Fmax, the structure collapses. There are two general types of
collapse modesbrittle and ductile.

Brittle Collapse
A brittle collapse occurs at a small additional displacement after Fmax, when the
structure quickly loses the ability to sustain load. A brittle collapse mode is also
sometimes characterized by a small amount of displacement, if any, between Fyield
and Fmax. There is no Fserviceability due to the rapid collapse; in other words, the
structure collapses quickly once the load is equal to Fyield. Certain types of
platform bracing configurations, such as K bracing, tend to have brittle failure
modes, since there is no alternative load path once several of the K braces fail.

100-8

2007 Chevron USA Inc. All rights reserved.

October 2007

[BookTitle]

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

Ductile Collapse
In contrast to brittle collapse, a ductile collapse occurs when there is a larger amount
of additional displacement after Fmax and a slower reduction in load carrying
capacity. A ductile collapse mode is typically characterized by a larger amount of
displacement between Fyield and Fmax in a gradual manner and is the preferred
mode of failure for a platform design. A platform with X bracing tends to have a
ductile collapse mode, since the compression tension pairs of the braces fail at
different load levels, providing a redundant load path.

Static Pushover Analysis


A special type of engineering approach, called a nonlinear static pushover
analysis, is commonly used to determine the ULS of structures. The approach is
called a static pushover, since dynamic effects are either excluded from the analysis
or included as a multiplier of the static forces, and the forces are applied as
incremental static loads until the platform collapses. The procedures and associated
technical background to perform static pushover for a fixed jacket on Chevron
projects worldwide are contained in this document.
Static pushover is a common methodology used to design new structures, as well as
assess existing structures for extreme loads, such as storms and earthquakes. Static
pushover is used in the offshore structure industry (API and ISO), as well as the
onshore building industry (FEMA 356 for earthquakes).

122 When to Use Ultimate Strength Methods


ULS is one of the key metrics to evaluate the structural performance of a platform.
ULS can be used to help make decisions for the assessment of existing platforms or
design of new platforms, although the more typical use is for existing platforms.
ULS is also used to design platforms in earthquake regions, but the focus of this
document is for regions where the platform design is dominated by waves.
Several specific reasons to determine a platforms ULS that are common in the
industry are summarized in the following seven subsections.

Determine ULS (Platform Capacity)


ULS provides a specific measure of the platforms maximum strength that can be
used to determine its performance against some measure of adequacy. The measure
can be in terms of a specific load value (kips) or normalized parameter, such as RSR
(defined in the next subsection). These values can then be compared to minimum
values to meet Chevron or regulatory requirements.

Determine RSR
RSR is a formal definition that is consistent with API and ISO and is the ratio of a
platforms ULS to the load acting on the platform for a 100 yr RP storm condition
(in accordance with API standards). If the 100 yr storm is the design criteria for
the platform, RSR is defined by the ratio of ULS/Fdesign in Figure 100-2. RSR
therefore provides a measure of the factor of safety inherent in the platform, above

October 2007

2007 Chevron USA Inc. All rights reserved.

100-9

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

[BookTitle]

the design point. Hence, an RSR of 2.0 means that the platform can take twice the
loading of the 100 yr storm before it fails.
RSR for a new Gulf of Mexico platform is typically in the range of 1.6 to 2.5. RSR
for North Sea platforms is typically in the range of 2.0 to 3.0, due to the flatter slope
of the North Sea metocean hazard curve. Existing older platforms may have RSRs
as low as 0.5, due to the inadequacies of the earlier generation fixed platforms, such
as the use of lower RP waves (25 yr), lack of joint cans, and lower deck elevations.
Platforms with an RSR less than 1.0 are at risk of complete failure in a 100 yr storm.

Damage Assessment
The effect of structural damage to a platform can be determined by comparing the
ULS of the platform in the intact condition to the ULS of the platform in the
damaged condition. Examples of common damage include dented or missing
members, corrosion, holes, and cracks. Since offshore steel jacket platforms have
redundant framing, ULS does not always decrease significantly when a member is
damaged. It usually requires several damaged members or damage to a key member
such as a leg, to reduce the ULS significantly. As an example, API specifies that a
10 percent ULS reduction should be considered as significant damage (dented, bent,
or missing members), such that further evaluation, including repair, may be
required.

Platform Additions
ULS analysis can be used to determine if additional weight (e.g., new drill package)
or storm loading (e.g., extra conductors or risers) can be added to the platform.
Platform additions can also occur if there is a change of use for a platform, such as
the addition of a quarters building. Much like a damage assessment (described in the
previous subsection), ULS analysis can be performed with and without the new
equipment to determine the change in ULS. The addition of topsides equipment
tends to not significantly change the ULS, since ULS is more greatly influenced by
lateral load than by vertical load. However, changes in vertical load can influence a
platform that has a ULS controlled by pile axial loading. The addition of wells and
risers may affect the ULS dramatically, since added wells and/or risers directly
increase the load acting on the platform. Similar to platform damage, API considers
a change in vertical or lateral loading as significant if the change is more than
10 percent. Significant increases in loading require a structural assessment to
determine if the platform can adequately sustain the increased loads.

Survival Event Check


A platform may be required to survive storm conditions of a specific RP (e.g.,
1,000 yr storm). This can be determined by comparing the ULS capacity with the
load acting on the platform (measured as base shear) for a range of storm
conditions. The storm with the base shear matching the ULS capacity is the storm
that would cause the platform to fail. Based upon metocean studies for the platform
region, this storm condition can be associated with a specific RP or metocean
condition, such as wave height. If the platform does not collapse for a load equal to
or less than this RP (or wave height), the platform passes the assessment. API

100-10

2007 Chevron USA Inc. All rights reserved.

October 2007

[BookTitle]

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

requires that an existing platform be shown to be able to survive a storm with


specific conditions, primarily based upon wave height.

Reliability Based Studies


ULS can be used to determine the reliability of the platform, which is often used in
risk and economic studies. The reliability can be determined based upon the
probability that the platform will experience the survival event (defined in the
previous subsection). Once the survival event and its associated RP are known, the
RP can be converted to a failure probability, which can be used for platform
reliability studies. Several examples of approaches to determine platform reliability
are contained in the references (refer to Section 180).

123 Ultimate Strength in Offshore Codes


Several worldwide offshore codes use ULS as a basis for decision making for
existing and new platforms. These are summarized in the remainder of this section.
Code requirements and additional description of ULS approaches and guidance can
be found in these documents.

API RP2A
API RP2A, Section 17, provides guidance for assessment of exiting platforms. The
API RP2A, Section 17, approach for waves that allows a ULS check was first
published in 1993 as API RP2A, 20th Edition. ULS can be used to check the
adequacy of the platform. API provides specific ultimate strength wave heights
that the platform must be shown to be able to survive. The user needs to
demonstrate that the platform will survive a base shear equal to or larger than the
ultimate strength wave conditions (based upon the platforms water depth).

API RP2SIM
A pushover analysis is typically used to determine the platform's ULS to compare to
the ultimate strength wave base shear. Prior to performing a ULS check, the user
can perform a design level check, which checks the platform using a linear
method, much like new design. However, if the ULS wave has a crest elevation that
reflects the platforms cellar deck, a ULS analysis is always required. API RP2SIM
is an emerging API document that will replace API RP2A, Section 17, in 2007 or
2008. The API RP2SIM ULS technical approach will be similar to API RP2A,
Section 17.

ISO 19902
ISO 19902 provides guidelines for the design of offshore steel platforms. ISO 19902
has a section related to assessment of existing platforms that is similar to
API RP2A, Section 17, and allows ULS approaches. ISO 19902 also has a
convenient section for computer modeling of damaged members (refer to
Section 141).

October 2007

2007 Chevron USA Inc. All rights reserved.

100-11

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

[BookTitle]

Other Codes
DOE, NPD, and others address ULS for platform design, but the details are not
covered here. Some of these codes also use a long return period metocean criteria,
i.e., a multiple of the elastic design criteria, such as 1,000 yr or 10,000 yr to
demonstrate the platform has adequate ULS characteristics. This process is
essentially a collapse check for design of new platforms, i.e., in addition to
elastic design (for example, 100 yr conditions). The platform will also not collapse
for these long return period conditions, although it may be damaged.

Seismic Design
API RP2A and ISO 19901-2 provide guidance for design and assessment of
platforms for earthquakes. A two level approach is used to design platforms for
earthquakes. The first is a design level approach (DLE) that is much the same as the
100 yr storm for metocean design. The platform is designed using factors of safety,
using an earthquake with an RP of 200 yr (Fdesign in Figure 100-2). However, the
platform is also designed using a ULS approach so that it does not collapse (ULS
in Figure 100-2) for a larger earthquake, typically on the order of 1,000 yr to 3,000
yr RP. ULS for earthquakes is typically computed using a nonlinear time-history
analysis or a pushover analysis, similar to that described in this document. The key
difference in an earthquake pushover is the development and shape of the pushover
load profile, which is based on a distribution of mass in the platform and associated
earthquake accelerations. Refer to ISO 19901-2 for further information.

124 ULS Software


There are several software programs that perform ULS pushover analysis for
offshore steel jacket platforms in a semi automated manner. In almost all cases, the
user needs to carefully develop the associated computer models, apply the pushover
load, and interpret results in order to ensure an accurate answer. Examples of
programs are:

SACS. EDI, Kenner, Louisiana. The COLLAPSE module of SACS is


commonly used worldwide for pushover analysis. http://www.sacs-edi.com

USFOS. http://www.usfos.no

CAP. This is the PC version of the SEASTAR/INTRA code originally


developed by PMB Engineering. http://www.capfos.com

EDP. Digital Structures, Berkeley, California. This is similar to CAP and is a


variation of the INTRA code. Digital Structures Inc., 2855 Telegraph Ave.
#300, Berkeley, California 94704.

MICROSAS. http://www.jraymcdermott.com/jrme

General Purpose Software. These programs are less commonly used for
offshore ULS analysis but can do the job.

100-12

ABAQUS. http://www.abaqus.com
ANSYS. http://www.ansys.com

2007 Chevron USA Inc. All rights reserved.

October 2007

[BookTitle]

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

Other software can perform ULS analyses that are not listed here; however, they
may not have the special features of the software listed here. If using another code,
be sure that it uses a documented approach to perform the pushover. The code
should be tested or calibrated to perform the specific type of ULS analysis for fixed
platforms as described in this document. The user should ask if there has been prior
ULS work that can be reviewed or whether there are any benchmark problems
to demonstrate the softwares ULS capability. Of particular interest is how the
software handles wave loading, member buckling (tubular members), tubular joint
capacity, large deformation (P-delta), and pile capacity (lateral and axial), as these
are some of the controlling and unique factors for offshore platforms ULS analysis.
The use of software that does not have the automated wave loading and nonlinear
pile-soil spring features found in offshore platform codes, such as SACS, USFOS,
and CAP, can be very time consuming because these loads have to be entered
manually into the program.

130 Information Requirements


131 Basic Information Needed
The information needed for a ULS analysis is similar to that needed for new design
or any type of design check for an existing platform. This is primarily data related to
the primary structural framing, the site soils, and the platform equipment loads. For
existing platforms, it is important that this data reflect the platform in its current
condition.

Main frame jacket, deck, and pile drawings


As built drawings should be used, if possible. The drawings should also reflect the
current configuration of the platform, since structural changes may have occurred
since the platform was installed.

Geotechnical report
Site specific data, including shear strength profile and pile axial compression and
tension capacity curves, are updated to modern API recommendations. Pile driving
records may be available to determine actual pile penetration.

Topside weights
Report the actual topsides currently on the platform. Area loads (e.g., 200 psf) can
be used in lieu of equipment, often conservatively modeled as a large load on the
structure. However, actual loads are preferred, since too large or too small a load
influences the pile foundation's ultimate capacity and can lead to inaccuracies in
the actual ULS if collapse is controlled by pile axial failure. Area loads, if used,
tend to be too heavy and provide a conservative (low) ULS. Area loads should only
be used for screening purposes to determine the approximate ULS adequacy of a
platform. In some cases, a topside survey is performed to confirm the location and
sizes of major topside equipment. The analysis should also consider whether there

October 2007

2007 Chevron USA Inc. All rights reserved.

100-13

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

[BookTitle]

are any planned future changes to the topsides that should be accounted for, such as
new drilling or production equipment.

Latest inspection results


Inspection reports provide information about the current state of the platform,
including damage, if any, such as dents, cracks, holes, or corrosion. The inspection
report should also be used to establish actual marine growth (versus code based
marine growth that may be used for new design). In some parts of the world, marine
growth is cleaned periodically, and this needs to be considered. In most cases, but
not all, the inspection report also contains information, such as the number and
location of risers and caissons, location of boat landings, platform orientation,
verification of the platform underwater elevations (which can be checked against
the drawings to see if the platform has subsided), and other useful information.

Photos
Above-water photos of the platform are critical for the engineer to provide a feel
for the platform, such as overall configuration and size, but they also provide visual
confirmation of the amount of deck equipment, orientation, number of boat
landings, number of risers and conductors, and deck elevation and overall platform
condition, such as corrosion. These and other items that can be seen in the photos
should match what is in the drawings. If there is no match, these items need to be
field verified.

Appurtenances
Report the actual number and location of conductors, risers, boat landings, bumpers,
and other appurtenances. These are usually found on the drawings but are best
confirmed via the inspection reports and photos. The number of conductors actually
installed on the platform tends to routinely vary from the number of slots and should
be independently verified. Sometimes there are fewer conductors, sometimes there
are more conductors that have been added over the years but have not been well
documented.
Future use of the platform should be considered in the ULS analysis. Drill rig or
additional conductors that are to be added to the platform should be considered for
the analysis. Otherwise, the analysis will have to be redone when these new loading
elements are added.

132 Availability of Data


General
The required data is not always available for all platforms, especially older
platforms installed prior to 1980. The major structural drawings are typically
available for most platforms, either in whole or in part. What tends to be missing
are some of the details. If the drawings are completely missing, the member sizes
and thickness will have to be determined by underwater inspection. This is very
costly and time consuming and is only warranted in very special cases. A partial set
of drawings can typically be used to infer the missing data. For example, if only one

100-14

2007 Chevron USA Inc. All rights reserved.

October 2007

[BookTitle]

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

or two rows of the vertical elevation drawings are available, the other rows can be
estimated, since most platforms are symmetric in design. Likewise, the horizontal
elevations are typically of common configuration. Drawings from other nearby
similar platforms can sometimes be used to replace missing information, especially
items, such as pile size and pile penetration.

Original Sources
The original platform owner (if the platform was acquired) or the original design
firm for the platform may have copies of the platform drawings, specifications, and
installation records and should be contacted. The local regulator may also have this
information.

Geotechnical Criteria
The most common missing data is the geotechnical criteria (borings) for the site.
These are usually in a separate report that tends not to be stored with the drawings.
Note that the pile size and penetrations are usually with the structural drawings. In
such cases, Chevron geotechnical experts should be contacted to assist in the
development of the most appropriate criteria for the site. For example, soils from
a nearby platform can be used if site specific soils are not available. In areas where
the soils are known to be generally uniform, nearby borings can be used to
interpolate reasonable soil strength, as determined by a geotechnical engineer.
Sensitivity studies may be performed to determine the effect of the soil strength on
the ULS. For example, if the platform failure mode is controlled by the strength of
the jacket, the soil capacity is not critical. In some regions, the local regulator may
specify the foundation information requirements. In the Gulf of Mexico, the MMS
requires the use of soil reports that are within 500 ft of the platform, although
studies by a geotechnical engineer or sensitivity studies can be used in lieu of a site
boring. However, in some cases, the soils information may just not be adequate, and
a new boring should be taken, particularly for platforms with planned upgrades.

Discrepancies
There can often be a discrepancy between the drawings and the current known
platform configuration. For example, the number of conductors or risers may
differ or there may be deck extensions (including subcellar decks) that were added
at a later date. These can typically be easily surveyed in the field, and it is
recommended that this be done (versus assuming which of the data is correct) in
order to ensure an accurate ULS analysis.

Marine Growth
If the marine growth is not available in the inspection report, the code based criteria
can be used. For example, 1.5 in. of marine growth on the radius from the waterline
to 150-foot water depth in the Gulf of Mexico is a typical code specification.

Steel Strength
The nominal yield strength of the steel used to fabricate the platform is not always
provided on the drawings or associated specifications. Mill certificates are typically
not available for older platforms. Nominal steel yield strength, if not available,

October 2007

2007 Chevron USA Inc. All rights reserved.

100-15

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

[BookTitle]

should be taken as 33 ksi for platforms installed in 1975 or earlier and 36 ksi for
platforms installed after 1975. The platforms should first be analyzed with the
nominal steel yield value throughout the entire structure, including the joint cans
and piles. If there are members or joints or piles that have high stresses and the
platform is not meeting the required acceptance criteria (design level or ULS), a
15 percent increase in the nominal yield strength can be included for those
locations to account for adjustment of nominal to mean yield strength. If the
nominal steel yield strength is 50 ksi, only a 10 percent increase can be applied. For
high consequence and/or platforms deemed critical to operations, a material sample
and appropriate testing should be considered to confirm the steel yield strength.

Grouted PilesGeneral
A grouted leg-pile annulus increases the ULS. The combined leg-pile cross section
results in a stiffer leg and also serves for better performance of the brace-leg joints
in terms of strength, as well as fatigue. It is not always clear whether the leg-pile
annulus was designed to be grouted, and if it was designed for grouting, whether the
grout was installed. There have been numerous examples of platforms that were
supposed to have a grouted leg-pile annulus but did not. A grouted leg-pile is
evidenced by notes on the drawings, drawings of grout piping running along the
legs to near the mudline, or by installation records. Installation records are the only
sure method to confirm whether the legs are grouted based upon existing data. A
grouted leg-pile can be verified during routine underwater inspections by using
ultrasonic testing (UT) to determine the leg wall thickness.

Grouted PilesVerification
UT will show a thicker leg if the leg-pile is grouted. Each leg should be verified that
it is grouted, as sometimes only some of the legs are grouted (due to weather,
mechanical breakdowns, etc., during installation). The grout verification needs to be
done only once in the platforms life. It is recommended that the presence of grout
be verified in this manner, especially where the presence of the grout controls the
platform collapse mode or results in a significant increase in the ULS. If the grout is
not verified via installation records or UT inspection, the results of such analysis
should be used with caution. Confirmation of grout should also be performed for
high consequence platforms.

Cellar Deck Elevation


The elevation to the bottom of steel (BOS) of the cellar deck is a critical value and
should always be field verified as part of the ULS assessment, if it is not available
from prior work. If the crest of the ULS wave is above the deck BOS, the platform
loading is very different and needs to be handled separately. Many platforms
experience subsidence over their life, and this can often be seen in photos of the
platform by examining the elevation of the first elevation of the waterline, typically
approximately 10 to 15 ft above the waterline. If this looks off by approximately
two or more feet, field verification should be performed. The elevation can be
scaled in a photo using common member sizes as a reference, such as a 24 in.
diameter brace visible in the photo.

100-16

2007 Chevron USA Inc. All rights reserved.

October 2007

[BookTitle]

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

Missing Data
If data is missing and an estimate needs to be made because field verification is
not possible, the estimate should be conservative. For example, if it is unclear
whether a conductor is 26 or 28 in., use 28 in. Generally, estimates for one or two
items do not have a significant impact on the ULS results, although estimates for
critical structural members, such as leg wall thickness, should always be done with
caution.

140 Modeling
Most ULS software will automatically develop the nonlinear ULS model from a
typical linear model used for design. While this is convenient, the user needs to be
aware of the different types of elements used and their shortcomings.
Figure 100-3 shows an overall view of an offshore platform that has been specially
modeled for ULS analysis. The figure shows some of the key features of the model
including the typical nonlinear member force displacement response for the primary
platform components. Further discussion is contained in the following sections
divided by Jacket, Foundation, and Deck.
Fig. 100-3

Specialized Nonlinear Modeling for ULS Analysis and Typical Member Response Mode

Linear
Elastic

s
e

DECK

Bending & Axial


Yield or Local
Buckling

Buckling

BRACES

LEGS

Full
Nonlinear
Response

s
e

SOIL-PILE

October 2007

Bending & Axial


Yield or Local
Buckling

s
e

PILES &
CONDUCTORS

2007 Chevron USA Inc. All rights reserved.

100-17

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

[BookTitle]

141 Jacket Modeling


The ULS jacket model is generally the same as a typical model for new platform
design in terms of the overall platform geometry, member sizes, and gravity loading.
The key difference is the use of nonlinear elements in the ULS model, versus linear
elements in a new platform design model.
The following sections identify the key platform components and how they should
be modeled:

braces,
legs,
joint capacity,
joint flexibility,
conductors,
conductor guide framing,
risers,
appurtenances, and
damaged members.

Braces
Braces are usually the controlling factor for lateral jacket capacity. If the diagonal
braces fail, the jacket bay loses the ability to transfer lateral load. The jacket bay is
defined as the associated jacket structure from one horizontal elevation to the next
(e.g., 100 to 130 ft). The braces are primarily long and slender and are therefore
prone to buckling at failure, and this must be properly captured in ULS analysis.
Buckling is a quick and sudden failure in a brittle manner with the brace quickly
losing its ability to carry additional load. This is in contrast to shorter members or
heavier braces that do not buckle but instead fail in axial yielding or bending in a
more relaxed ductile manner. Figure 100-4 shows a typical force deformation plot
for a brace that buckles and one that fails by yielding.
Fig. 100-4

Brace Failure by Buckling and by Yielding

FORCE

Yielding (Ductile):
short heavy braces

Brittle:
long slender braces

DEFORMATION

100-18

2007 Chevron USA Inc. All rights reserved.

October 2007

[BookTitle]

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

The buckling type of brace is often called a strut, since it carries predominately
axial loads. There is bending in the member, but it tends to be small, since the
member is flexible. The brace is, therefore, often modeled using this strut approach
and either neglects the moment in the member or adjusts the buckling capacity to
account for the moment. The buckling capacity is determined from empirical or
experimental data. The most common and thought to be the best method for
buckling capacity was developed by the Structural Stability Research Council
(SSRC), which publishes a specific formula for axial capacity of tubular members.
This formula is also used by API in the LRFD version of API RP2A (refer to
Equation D2.2.1). This formula provides a convenient quality assurance (QA)
check of the expected buckling load of a member by removing the API LRFD
factors of safety from the equation.
Some software uses an alternative approach that subdivides the member into a
super element that divides the member into approximately 10 segments and uses a
special nonlinear analysis subroutine to mimic the member nonlinear performance.
Figure 100-5 shows a super element in a platform model. The software uses a
sensing routine that initially models all members as linear and then monitors their
load during the analysis. If a member approaches nonlinearity, the member is
replaced by the super element to perform the nonlinear calculations. The advantage
of using super elements is that it is an automated and user friendly approach.
However, the user needs to pay special attention to the results, as the super elements
can end up with some unusual member failure modes, that are theoretically correct
but do not make sense (although this is typically not the case). The other problem is
that, in order to properly mimic buckling, the super element requires an initial out of
alignment (called eccentricity) in a brace model. The segmented member then uses
the axial load eccentricity combination to continually displace the member out of
plane until it eventually buckles. The user can define the eccentricity directly, or
some codes will estimate it automatically. A typical value of one to five percent
eccentricity of the member diameter should be used. Examples of software that use
this approach are SACS and USFOS.

Legs
Platform legs are heavier with thicker walls and have lower slenderness ratios than
braces and therefore generally fail in bending. These types of members are not
typically prone to global buckling, although local buckling (in the form of a bulge)
has been observed to be a problem in some older platforms if the wall thickness is
thin. Global buckling of the leg is also seldom a problem, since the enclosed pile
prevents the leg from buckling. The legs are commonly defined using nonlinear
beam columns, although the automated super element previously discussed for
braces is also used for legs. If the leg-pile annulus is grouted, the leg properties
should be a combination of the leg and inner pile. Most software has a routine to
account for this, which increases the leg wall thickness to account for the pile,
often limited to 1.75 to 2 times the leg thickness. A grouted leg-pile annulus also
provides substantially stronger joint capacities, especially in compression.
Platforms with grouted leg-pile annuluses have been observed to perform better in
hurricanes than similar ungrouted platforms. Refer to Section 182 (Puskar, F.J. and
Spong, R.E., Energo, 2006). In situ grouting of a leg-pile annulus is a common

October 2007

2007 Chevron USA Inc. All rights reserved.

100-19

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

Fig. 100-5

[BookTitle]

Super Element Representing a Brace in a Platform ModelThe brace is divided into segments as shown, each
with its own nonlinear finite element properties. The resulting super element can mimic nonlinear behavior in
bending, compression (including buckling), and tension.

*Reference: SACS at www.sacs-edi.com

technique to improve the performance of existing platforms (that were not grouted
at installation).

Joint Capacity
Joints should be included in the model using specific empirical formulation. The
most common available options are the API RP2A joint equations (without the
factors of safety) and joint ultimate capacities determined from the Joint Capacity
JIP (refer to Section 181). The SACS software, as well as others, has the equations
and capacities built in with an automated checking process to determine if the joint
or the brace fails first. The software monitors the load on the joint from the brace
using one of the empirical formulations, and when the brace load exceeds the joint
capacity, the brace is disconnected and eliminated from the ULS computer model.

Joint Flexibility
Most structural programs consider joints as rigid when computing the forces in a
member, particularly moments at the member end. In reality, the joints are
flexible and the members at a joint rotate relative to one another based upon
relative stiffness. Generally, joint flexibility is neglected due to the added
numerical complexity, and this is conservative, since the loads in the members are
higher without the joint flexibility. Since the loads in the members are higher, they
fail sooner, and the global ULS of the platform is conservatively lower. Joint
flexibility reduces the loads in members and increases the ULS on order of a few
percent up to 10 or 15 percent, depending upon the platform geometry and relative
stiffness of the legs and braces. Simple formulations to compute joint flexibility by
hand calculations can be found in Section 183 (Bouwkamp, J.G., OTC Paper 3901,
1980). Joint flexibility is also a built in option in some structural software programs.

100-20

2007 Chevron USA Inc. All rights reserved.

October 2007

[BookTitle]

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

Conductors
Conductors should be included in the ULS platform strength model, particularly in
terms of their contribution to foundation resistance. This is contrary to a new
platform design model, where the conductor hydrodynamic loading is included, but
the foundation resistance is typically neglected. Instead, the conductor load goes
though the jacket into the soils, resulting in a conservative platform design. For
ULS analysis, the inclusion of the conductor foundation resistance is a closer
representation to the actual load transfer and is typically included. The amount of
platform resistance and effect on the ULS of the conductors is partly a function of
the presence of the conductor framing at the mudline. If the conductor framing is
present, the conductor also helps to restrain the jacket at the mudline and, hence,
contributes to the overall platform ULS. If the conductor framing is not present, the
conductors move independently of the jacket at the mudline, and they provide less
resistance to global platform loads, other than their own loading. The conductors
should be modeled as linear members, since they typically include internal drilling
pipes that are grouted to the conductor. Hence, their element capacity is difficult to
determine and, it is, therefore, best to treat them linearly. The conductors should be
modeled to move independently in the vertical direction through the conductor
guides as is done for a design model. SACS uses a special element called a
wishbone for this. The conductor foundation should be the same as for piles with
full nonlinear pile-soil springs (refer to Section 142).

Conductor Guide Framing


Conductor guide framing should be modeled in detail at each elevation in order to
provide proper load transfer from the conductors back into the primary jacket
framing. The small guide tubulars are modeled individually to form the conductor
guide grid pattern. The plated area and the cone guide can be replaced by stiff
tubular X brace members (e.g., 24 by 1 in.), with the center of the X the location of
the center of the conductor. The conductor wishbone type of element then
connects to the center of the X and allows the conductor to move freely (vertically),
but lateral metocean loads go into the center of the X and are then distributed to the
guide grid framing. The steel plating and guide cones found in many conductor
trays strengthen the guide framing and generally prevent failure of these members
but are difficult to model accurately. Therefore, for ULS analysis, this grid work of
guide framing is usually assigned to be elastic in order to prevent early failure of
these members during the pushover analysis.

Risers
In most cases, risers are excluded from the ULS (or design) model in terms of
strength and are only included in terms of their contribution to wave and current
loading. Most programs have the ability to perform this type of modeling by putting
the member in the model but identifying it as a dummy member, which accepts
hydrodynamic load and allows it to be carried into the jacket but does not contribute
to the strength of the platform. In some cases, the riser is an integral part of the
structural framing, for example, a J tube or curved conductor. In these situations,
they should be modeled as a load carrying structural component.

October 2007

2007 Chevron USA Inc. All rights reserved.

100-21

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

[BookTitle]

Appurtenances
Similar to risers, the boat landings, barge bumpers, and other appurtenances are not
included in the platform strength model. They are also modeled as non load carrying
dummy members (refer to the Risers subsection). In some cases, appurtenances
may be considered as part of the structure and contribute to its strength. For
example, a launch runner along a leg adds to the strength of the leg. This type of
appurtenance strength modeling needs to be considered on a case by case basis,
since it adds to the complexity of the structural model (especially during the
modeling effort), and the added strength may be marginal, having little effect on the
ULS and not worth the effort.

Damaged Members
Damaged members should be modeled where appropriate. As an initial check that is
conservative, the member can be completely removed from the computer model in
order to determine if the platform performs adequately. Because of the redundant
framing in a typical offshore platform, the damage of one or even a few braces
typically has only a small effect on the ULS. If there is a need to model the member
in its damaged state, there are several formulations provided in ISO 19902 that
provide a partial strength approach to mimic the damage.

142 Foundation Modeling


Pile Steel
The foundation should be modeled using structural elements to model the pile steel
and nonlinear springs to model the pile-soil interface. The pile-soil springs are
commonly called p-y springs for lateral pile loading, t-z for vertical pile loading,
and q-z for pile tip loading.
The pile steel must be modeled using nonlinear beam type members that can mimic
the bending of the pile below the mudline. One of the key platform failure modes is
pile bending below the mudline, typically due to a combination of small pile (30 in.,
for example), with little bending resistance and weak soils that provide little lateral
restraint. The maximum pile bending moment is typically located a distance of 5 to
10 times the pile diameter below the mudline, and this is often the location of the
first pile yield. Most platforms have a thicker pile wall or sometimes higher yield
strength for this length to prevent this type of bending failure, but the thicker section
is not always present or is insufficient in older platforms. Therefore, it is important
to model the pile accurately below the mudline in order to capture these effects.

Pile-Soil Springs
The pile-soil springs (p-y, t-z, and q-z) are either provided in a geotechnical report
for the site or are computed automatically by the software using API or another
code. They are called pile-soil springs, since they represent the specific type of
behavior of the pile and the nearby soils. Hence, they are a function of the pile and
the soils. Larger piles result in stronger pile-soil springs. Similarly, stronger soils
result in stronger pile-soil springs.

100-22

2007 Chevron USA Inc. All rights reserved.

October 2007

[BookTitle]

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

The pile-soil springs are provided in a static (sometimes also called virgin) or
dynamic (sometimes also called degraded) format. The degraded value should
be used for pushover analysis. This represents the pile-soil capacity after a large
number of waves have passed the structure, causing the jacket-pile system to move
back and forth, effectively degrading the soil capacity. Since the largest wave is
expected at the peak of a storm, after numerous other smaller waves have passed the
platform, it is reasonable to use the degraded strength. The amount of degradation
depends upon the type of soils but is on the order of 20 to 35 percent.
Soil-springs that are based upon the geotechnical report are input using explicit p
and y values that are a function of the depth below the mudline (similarly for t and z
and q and z). If they are to be determined by the ULS program, the typical input is
the soil strength as a function of depth (shear strength for clays and friction angle
for sand), combined with the soil unit weight. These are typically provided in the
geotechnical report. The geotechnical report often includes pile compression and
tension axial capacity as a function of depth. Most software will also output these
values. It is good practice to compare the ULS model axial capacities with the
geotechnical report capacities, as these should be reasonably close. Chevron
geotechnical engineers should be consulted if there are any concerns with pile-soil
modeling.

Sensitivity Studies
Investigation of platform failures following hurricanes in the Gulf of Mexico shows
that there are few, if any, documented pile failures. However, results of ULS
analyses on some actual platform failures, as well as platforms that survived,
indicates that pile failure, either lateral bending or axial (or a combination), should
have occurred. This indicates that there is conservatism in the pile-soil springs or
in the overall ULS method related to pile capacity, particularly for the soft clays
typical in the Gulf of Mexico. Therefore, if the platform failure mode is controlled
by the foundation, some sensitivity studies should be performed, including
variation of the pile-soil strengths used in the ULS analysis.

143 Deck Modeling


General
The deck structure can be simplified in most ULS analysis, since the jacket
structure is the most critical component to resist lateral metocean loading. The deck
structure model serves as a method to interconnect the deck legs, apply gravity,
wind, and topside loads, and input wave in deck loading (if any), for which detailed
modeling of every structural member is not required. Simplification of the deck
structure is also often a requirement in order to reduce the size of the structural
computer model.

Simplified Deck Structure


A simplified deck structure consists of all of the primary deck members, including
the deck legs, truss rows, and major lateral bracing. Secondary bracing related to
smaller deck beams and bracing to support floor loads, as well as deck plating, can

October 2007

2007 Chevron USA Inc. All rights reserved.

100-23

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

[BookTitle]

be neglected (although, the weight of this secondary bracing needs to be included).


Such floor bracing sometimes provides the lateral stability to the deck, and removal
of the secondary bracing may make the deck unstable to torsion loads. In this
situation, the deck framing can be replaced with an X brace to adequately stiffen
the deck in plan.

Structure Element Type


The deck is almost always modeled using linear structural elements, since failure of
the deck structure itself is not an issue. The only time that the deck may include
nonlinear elements is if there is a special case of deck loading, and local failure of
the deck members is a concern. This is seldom the case. The use of linear deck
elements simplifies the deck modeling and structural analysis process. Note that the
deck legs must be nonlinear, since they can be the cause of failure of the platform,
especially for wave in deck cases.

Subcellar Decks
Subcellar decks that are located below the cellar deck, such as a scaffold deck,
spider deck, or sump deck, should be modeled in sufficient detail in order to capture
the wave loads that may be acting on them and to transfer these loads back to the
cellar deck. This can be accomplished by modeling of the main members of the
deck, including their hydrodynamic area. If the subcellar deck includes more than a
few pieces of equipment or if the equipment is large, the additional hydrodynamic
area of this equipment should also be included.

API Simplified Procedure


The API simplified procedure for wave in deck loads is typically only used for the
cellar deck and above. This procedure uses an equivalent projected area combined
with the kinematics of the pushover wave crest to determine lateral wave loads.
Refer to Section 154 for more details.

150 Load Definition


151 Gravity and Buoyancy (Not Ramped)
Gravity and buoyancy loads are applied in the same manner as for new platform
design. These loads are constant for the pushover analysis and are not ramped.
Gravity loads are also discussed in Section 155, describing P-delta.

152 Wave and Current (Ramped)


Wave loading is the dominant pushover load. The specific wave height used should
be combined with the associated storm surge, current, and wind. Current loading
is not as significant as wave loading. The methods for computing these loads on a
platform are in accordance with API RP2A 20th edition or later. The metocean kinematics factor, drag coefficient (Cd), inertia coefficient (Cm), and other variables are
all the same as for new platform design as defined in API.

100-24

2007 Chevron USA Inc. All rights reserved.

October 2007

[BookTitle]

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

The wave height may vary by direction, although this has little effect on the ULS,
since ULS is predominantly a function of the platform strength. The wave height
does influence the reference base shear acting on the platform, for example, the
RSR. Hence, a platform may have a different RSR for different directions,
depending upon the platform orientation.
Wave and current loads are ramped during pushover analyses. Additional
discussion of wave and current pushover loads is provided in Section 161.

153 Wind (Ramped)


General
Wind is usually a secondary pushover load and accounts for 5 to 15 percent of the
total base shear at ULS for most platforms. The wind acts primarily on the topsides,
especially for a pushover, where the wave crest is almost to the lower deck
elevation. Hence, there is no need to compute wind loads on the deck legs,
boatlandings, etc., as these will all be underwater. In some cases, the wind load
can be comparable or higher than the wave load, for example, quarters platforms
that have large wind areas or shallow water platforms, where most of the platform
is above water. In these cases, the user should pay greater attention to the computed
wind loads.

Computation
Wind loads should be computed in accordance with API RP2A, in the same manner
as new platform design, where a combination of wind loading area and wind speed
are used to determine the wind force. The wind speed should be approximately
consistent with the wave being used for the pushover. For example, if the pushover
wave is a 100 yr condition, the wind speed associated with the 100 yr wave should
be used. If a 50 yr wave is used, the wind speed should be the associated wind of
that event. The wind area used in the model should consider any future planned
topsides equipment and structure.

Source of Wind Data


In practice, such detailed wind information may not be available (but check with
Chevron metocean experts). Usually only one or two wind speeds are available, for
example, just the 100 yr and, in some cases, the 50 yr, as well. Since the wind is a
small percentage of the pushover load, it may be easier to use the conservative
100 yr (or other value) for all of the pushover cases. This reduces the time required
to determine the wind speed for each pushover wave, as well as the time required to
recompute the wind loads.

Wind Load Ramping


The wind load will be ramped in parallel with the wave and current loads during
the pushover analysis. Several programs, such as SACS, will automatically
determine the wind loads based upon the user defined wind speed and the exposed
wind loading area of deck members and user defined equipment. While this is a
convenient feature, the user must make sure that these loads are adequately ramped

October 2007

2007 Chevron USA Inc. All rights reserved.

100-25

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

[BookTitle]

with the hydrodynamic loads on the platform during the pushover. An alternative is
to compute the wind loads by hand and apply them to the deck as a static load that is
ramped (i.e., increased incrementally) with the wave and current loads.

Shallow Water Structures


For some shallow water structures (less than 50 ft WD) with large topsides, the
wind load can be a greater percentage and may even control design. Additional care
should be taken to compute wind loads more accurately for these platforms.

154 Wave Load on the Deck (Ramped)


API RP2A, Section 17, provides a simplified approach to determine the
hydrodynamic loads acting on a platform deck when hit by a wave. The approach
determines the wave kinematics velocity at the crest and uses this velocity along
with the expected wetted area (termed silhouette area) for the topsides and a drag
factor to compute the wave load on the deck (refer to Figure 100-6). The drag factor
is adjusted to account for a heavily, moderately, or bare equipped deck. The wave in
deck loads should be computed separately for each pushover direction studied, since
the deck structure may have a different profile for different directions (including
diagonal). Details are provided in API RP2A, Section 17.
Fig. 100-6

API Silhouette Area for Computing Wave in Deck LoadsScaffold deck can be modeled directly to incur
hydrodynamic loads, or alternatively, it can be part of the silhouette area as shown.

Silhouette Area
Shown in Gray

Crest of Wave

Main Deck
Deck legs and
braces are part
of deck area

Cellar Deck
Scaffold Deck

Deck legs and


braces are part
of jacket

The API RP2A, Section 17, wave in deck approach is conservative. Alternative
wave in deck loading algorithms are available, as provided in Section 180, and may
predict lower wave in deck loads. These rely on explicit calculation of the deck
wave load areas on a member by member basis, combined with application of wave
kinematics near the crest and drag factors. However, these can be very time

100-26

2007 Chevron USA Inc. All rights reserved.

October 2007

[BookTitle]

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

consuming to implement, and it is recommended to start with the API approach and
only use these more elaborate methods if necessary.

155 P-Delta Effect


ULS analysis should consider a phenomenon called P-delta, where P is defined
as the resultant of the gravity load, and delta is the lateral displacement (refer to
Figure 100-7). In simple terms, P-delta is the effect of gravity pulling the structure
over as it moves laterally during the pushover. As a structure moves laterally, the
center of gravity (CG) is also moved laterally due to the displaced shape of the
platform. This causes an eccentricity between the original CG location and the
displaced CG location. This eccentricity, combined with the resultant gravity load,
results in an overturning moment at the base of the platform. The larger the lateral
displacement, the larger the overturning moment will be. In other words, as a
structure is displaced laterally, it tends to pull itself over and at some point will
topple due to the moment caused by P-delta. In structural analysis terms, P-delta is
called large deformation theory, is available in most ULS programs, and is
applied on a member by member basis in order to determine the global effect on the
platform. P-delta can, however, lead to numerical instability, especially as the
displacements in the pushover become larger. P-delta is not normally used for the
design of new platforms.
Fig. 100-7

P-Delta EffectNote that displacements are exaggerated to more clearly show the
P-delta effect.

P
Delta

October 2007

2007 Chevron USA Inc. All rights reserved.

100-27

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

[BookTitle]

The P-delta effect is created as the platform moves laterally, creating an


eccentricity (delta) between the platform's original center of gravity and the CG of
the displaced shape. The result is an overturning moment that increases with
displacement (hence, the nonlinear effect) and tends to pull the platform over.
A platform analyzed with P-delta has a lower capacity (lower ULS) than a platform
without P-delta. Effects of P-delta begin to be significant for platforms in water
depths greater than approximately 150 ft and should always be used for platforms in
water depths greater than 300 ft. Shallow water platforms have little P-delta effect
(approximately less than five percent total lateral load), since they are stiff laterally
and fail at relatively small lateral displacements (inches), therefore having a small
P-delta overturning moment. Deeper water platforms have a large P-delta effect (for
example, 10 to 20 percent total lateral load), since they are less stiff laterally and fail
at much larger displacements (tens of inches). Likewise, platforms with heavy decks
have a large P-delta effect.

160 Ultimate Strength Analysis


161 Pushover Load Approach
This section provides a general description of the pushover approach to determine
the platform ULS. The process takes a static snapshot of wind/wave/current acting
on the platform and monotonically increases this snapshot until the platform
collapses. Vertical gravity and buoyancy loads are held constant through the
process. Most ULS programs have an automated process for developing the
pushover load profile and for performing the analysis. This is often called load
ramping or wave ramping. The wave ramping term comes from the fact that the
wave size is modified (typically ramped up) until the collapse load matches the base
shear of the wave.
The process assumes that the ULS computer model is prepared, a quality assurance
check has been completed (to be sure that the model matches the drawings and other
information like material strengths and topsides loads), and the model is ready for
analysis. The discussion also assumes that a single direction is considered, although
a pushover is typically run in several directions (as described later). Figure 100-8
provides a convenient summary of the ULS analysis with the key items for each
step.

100-28

2007 Chevron USA Inc. All rights reserved.

October 2007

Summary of the Eleven Steps of the ULS Analysis Approach

Step

Description

Comments

Structural ModelSelect an
existing model or develop a
new model for analysis.

a. Typically, the first step in a platform assessment may be a design level check (in accordance with API RP2A, Section 17) using a
linear model. This model can often be converted for ULS analysis.
b. An existing model should be verified to ensure that it accurately reflects the drawings, specifications, topsides weight, and current
condition of the platform.
c. A model may also be developed from scratch specifically for ULS analysis.

2007 Chevron USA Inc. All rights reserved.

Gravity and Buoyancy


LoadsApply gravity, dead,
live, and buoyancy loads.

a. Gravity dead and live loads should reflect the existing platform configuration, including adjustments for future plans (e.g., new
compression skid).

Select Metocean
ConditionsInitial estimate of
the metocean conditions that
will cause platform failure.

a. The design wave for the region (e.g., 100 yr) is a good starting point. The associated wind, current, and storm surge need to be
included.

Develop Pushover Load


ProfileDevelop the load
profile of jacket lateral loads.

a. Using the initial metocean conditions from Step 1, determine the metocean loads acting on the platform. The load profile is typically
automatically generated by the software, based upon the maximum base shear as the wave passes the platform.

Wave in DeckIf necessary, a. Determine whether the wave used in the pushover has a crest elevation higher than the bottom of steel of the cellar deck. If so,
determine Wave in Deck loads. continue with this step. If not, go to Step 6.

b. These loads are held constant during the pushover analysis and are input separately from the metocean loads.

b. In the USA, API provides specific criteria (A-1, A-2, and A-3) for assessment of existing platforms that are convenient for the initial
conditions.

b. The API wave load recipe should be used to determine the metocean loads.

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

100-29

Fig. 100-8

b. Wave in deck load should be computed in accordance with the API simplified (but conservative) procedure and is a function of the
crest elevation, crest kinematics, silhouette area of the deck, and the amount of equipment on the deck that blocks the wave from
passing.
c. Wave in deck load is computed separately for each pushover direction.
d. For simplicity, apply the wave in deck load by equally dividing the total load by the number of legs and adding this amount to each
leg at the cellar deck nodes. If the deck is braced with vertical diagonals from below, the wave in deck load should be applied to the
cellar deck nodes where the wave would first impact.
6

Initial Pushover Loading


a. A small percentage (e.g., 10%) of the pushover load profile is applied to the platform.
Apply the initial pushover load. b. The platform displaces accordingly, and loads are developed in platform members.
c. The platform response is linear during the initial application of load (in some cases there may be a small nonlinearity due to the
pile-soil p-y nonlinear response).

[BookTitle]

October 2007

Summary of the Eleven Steps of the ULS Analysis Approach

Step

Description

Comments

Increment Pushover Loads


Increment the pushover load
profile until the platform
collapsesthis is called load
ramping.

a. The pushover load profile is increased from the initial load in a stepwise manner until the platform collapses (see definition of
collapse in Section 164).
b. The load steps can be large at first, since the platform performs in linear manner (other than some small pile-soil nonlinearity) at
loads below the first member failures.
c. The load steps should be decreased after the first member failures to provide a more stable solution for the software and
adequately capture the member failure sequence.
d. Most software uses an automated load stepping process, but the user should intervene as necessary to improve results and
provide a better understanding of the platform ULS.

2007 Chevron USA Inc. All rights reserved.

Preliminary ULSDetermine a. Defined as the maximum base shear acting on the platform during the analysis. The ULS may or may not be at the point of
the preliminary platform ULS by collapse.
checking the Load Factor.
b. The user should determine if the pushover load profile used in the analysis is appropriate by determining the load factor, defined as
the ratio of the ULS to the base shear of the load profile. Ideally, the load factor is equal to 1.0 in a pushover analysis.
c. The load factor should be within 20% of 1.0 (e.g., 0.8 to 1.2). Otherwise, the initial metocean conditions used to develop the load
profile are not an accurate representation of the ULS condition. Another concern is if there is wave-in-deck loading.
d. If the load factor is in this range, the pushover load condition is reasonably accurate, and there is no wave in deck, proceed to
Step 10. If it is not in this range, proceed to Step 9 to select a new pushover wave profile.
e. If there is wave -in -deck and the load factor is less than 1.0, also proceed to Step 9.

Iterate Process (as


necessary)Iterate the
metocean conditions of the load
profile as necessarythis is
called wave ramping.

a. The load profile needs to be adjusted down if the load factor is less than 0.8, since the pushover wave is too big. This provides an
overly conservative result.
b. The load profile needs to be adjusted up if the load factor is more than 1.2, since the pushover wave is too small. This provides a
result that is not conservative.

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

100-30

Fig. 100-8

c. To save time, use a wave height that creates a base shear that is close to the ULS. This can be found by running a few sample
waves past the platform to determine their base shear and then selecting the wave with a base shear that is close to the ULS base
shear.

10

Final ULSDetermine the final a. The final ULS capacity is determined once the iterative process has reached a point when the load factor is in the range of 0.8 to
ULS.
1.2.

11

Other DirectionsRerun
pushover for other directions.

b. For cases with wave in deck, the load factor needs to be in the range of 1.0 to 1.2.
a. The ULS is typically determined for three platform directions: broadside, end-on, and diagonal, depending upon the platform
geometry and symmetry.
b. The user must carefully consider the direction of the platform bracing when deciding which directions need to be run. Consider
whether bracing will be placed in tension and compression as these may give a different ULS.

[BookTitle]

October 2007

[BookTitle]

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

162 General Process for Pushover Analysis


The general process for a pushover analysis is described by the following 11 steps.

Step 1: Structural Model


The first step in any analysis is to select an existing model or develop a new model
to be used for the analysis. An existing model is sometimes available from prior
work on the platform or as part of the design level linear assessment check in
accordance with API RP2A, Section 17. Such models are almost always linear and
will need to be converted to a nonlinear model for ULS analysis. If using a program,
such as SACS, the conversion is rather straightforward (refer to SACS users guide).
For other software, the conversion may be more cumbersome.
In any case, an existing model should be verified as accurate by ensuring that the:

model reflects the platform configuration in accordance with drawings and


specifications;

associated topsides loads (weight and wind area) are correct;

model represents the specific configuration of the platform under study (i.e.,
current configuration, including any damage, or a future configuration being
studied).

Other similar quality assurance checks should also be made as necessary. A


new ULS model can also be developed from scratch if there is no existing model
available.

Step 2: Apply Gravity (Dead and Live Loads) and Buoyancy Loads
These tests are computed in accordance with Section 151 and applied to the
platform at 100 percent in accordance with the first load step, which is separate
from the later pushover load steps. These loads are not ramped during the pushover,
since they are constant.

Step 3: Select Metocean Conditions for ULS Check


The user needs to make an initial estimate of the metocean conditions that will fail
the platform. Since the wave height is the most dominant factor, this is usually the
control parameter (wind may control in shallow water less than 50 ft deep). The
wave height used to develop the pushover load profile should create a base shear
that is reasonably close to the ULS (within 20 percent of the ULS). If not, the
pushover profile is not a realistic representation of metocean loads acting on the
platform. For example, a 25 ft wave should not be used to develop the pushover
load provided, when, in fact, the platform collapse is equal to a 75 ft wave, since the
load distribution of forces acting on the platform will be dramatically different for
the two waves. The centroid of load for the 25 ft wave would act much lower on the
platform than the centroid for the 75 ft wave (therefore putting less overturning
moment on the platform), and the ULS will be different. Hence, a 70 ft or 80 ft
wave would be a more realistic and acceptable representation in this case. As
indicated in Step 9, the user may have to iterate on the pushover wave height, but
a reasonable first estimate (as explained in the beginning of this step) can help. It

October 2007

2007 Chevron USA Inc. All rights reserved.

100-31

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

[BookTitle]

usually converges in three attempts. This is called wave ramping, since the user is
using different wave heights for each pushover.
One convenient method to select the initial load profile is to use the new platform
design wave height and associated metocean conditions for a region, such as the
100 yr, since most platforms have an RSR of 1.0 or greater. API RP2A, Section 17,
wave heights for the USA are also a convenient starting point. If the platform is
designated as A-1, start with the A-1 ultimate strength wave height. The same
applies for A-2 and A-3 platforms. In fact, if the platform does not collapse at the
point where the pushover profile has a load factor of 1.0 (i.e., it has been applied
100 percent), technically the platform passes the API RP2A, Section 17, ultimate
strength check. Once the wave height is selected, the associated wind, current, and
storm surge/tide need to be defined.

Step 4: Develop Load Profile of Jacket Lateral Loads


Using the initial metocean condition from Step 1, determine the metocean loads
acting on the platform as described in Section 152. Usually the point of the
maximum base shear acting on the jacket is used to define the load profile. The
maximum overturning moment can also be used as the defining value. This is
essentially a static snapshot of lateral metocean loads resolved back to the nodes
of the platform.

Step 5: Determine Wave in Deck Loads (if applicable)


The user needs to determine if the pushover wave of interest has a crest elevation
below or above the BOS of the cellar deck. If the crest elevation is below, there is
no need for wave in deck loading for this iteration of the wave height (it may be
needed later). If the crest elevation is above, the wave in deck load should be
computed separately in accordance with Section 154, in accordance with the amount
of crest that extends above the BOS cellar deck. A single, combined lateral load
profile should be developed using the Step 4 pushover load profile and the Step 5
wave in deck loads.
The wave in deck loads are most simply applied to jacket by equally dividing the
total wave in deck load by the number of legs and adding this amount to each leg at
the cellar deck nodes (because the deck is so stiff, it would equally divide the load
anyway). If the deck has vertical diagonal bracing below the BOS, the wave in deck
force should be applied only to the leg nodes at the cellar deck row first affected by
the wave, since the deck may not equally distribute the wave in deck load in this
case.

Step 6: Apply Initial Pushover Load


The pushover load profile developed in Step 4 (or Step 5 with wave in deck loads) is
applied to the platform on an initial small percentage basis, for example, 10 percent.
The platform then displaces accordingly, and loads are developed in the platform
members. This is the first step in the pushover analysis and is the initial incremental
ramping of the load profile acting on the platform.

100-32

2007 Chevron USA Inc. All rights reserved.

October 2007

[BookTitle]

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

Step 7: Increment Combined Loads until Platform Collapse


The software will apply more load in an increasing manner (20, 30, 40 percent, etc.)
up to and above a factor of 100 percent until platform collapse. As the load becomes
large, components of the platform begin to fail, and the platform incurs a larger
displacement for each load increment, since it is no longer acting in a linear manner
(it is now nonlinear). This is the reason that the pushover curve begins to flatten
out as shown in Figure 100-2, as well as in other example pushover curves in this
document.
Most software automates the load ramping process, and the user need only define
the number of increments to ramp the load. Some software, such as CAP, have an
automated load stopping process that increases or decreases the size of the load
steps, depending upon how easily the software is converging to an answer. If there
are no or few nonlinearities, for example, at the beginning of the analysis, when
small loads act on the platform, the software uses large load increments. If there are
a lot of nonlinearities, such as near collapse when there are numerous members
failing, the software uses small increments to more adequately seek a stable
numerical solution and determine collapse.

Step 8: Determine the Preliminary Platform ULS


The preliminary ULS is defined as the base shear acting on the platform at the point
of collapse. Collapse is further defined in Section 164. The load factor is defined
as the ratio of the base shear at the time of the ULS to the base shear from the
pushover wave (Step 2, Step 3 with wave in deck). This is considered the
preliminary ULS, since the pushover wave load profile may or may not match the
initial load profile used for the analysis. If the load factor is 0.8 to 1.2, the pushover
wave is approximately correct, and there is no need to revise the pushover wave and
rerun the analysis. If it is outside these bounds, the pushover wave should be rerun
as discussed in Step 9.
If the load factor is greater than 1.0, the ULS wave will have a higher wave crest
(than used for the pushover), and there is a chance that this wave crest may affect
the deck. This needs to be checked and the pushover rerun if the increased wave
height would affect the deck. As previously discussed, wave in deck loading results
in a dramatic change in the way the loads are applied to the platform, and this needs
to be included in the analysis, even if the load factor is between 1.0 and 1.2.
However, if wave in deck loading is already included in the pushover load profile,
the result will be acceptable.

Step 9: Iterate Process as Necessary


If the load factor is less than 0.8, the wave load profile should be adjusted to a
smaller wave. If the load factor is greater than 1.2, the wave load profile should be
adjusted to a larger wave. The next logical wave is one that has the same base shear
as the preliminary ULS. The user may have to run a few sample waves past the
platform to find a wave with a base shear equal to the ULS. The refinement need
not be precise, and whole increments of wave height in feet are adequate (i.e.,
fractional wave heights, such as 72.3 ft, are not necessary; 73 ft would be adequate).

October 2007

2007 Chevron USA Inc. All rights reserved.

100-33

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

[BookTitle]

Step 10: Determine the Final Platform ULS


The final ULS capacity is determined once the iterative process has reached a point
when the load factor is approximately 0.8 to 1.2, since further refinement of the
pushover wave does not result in significant changes in the ULS.

Step 11: Rerun Pushover for Other Directions


Usually, a pushover does not always have to be run in all directions for the
platform. The user should take advantage of platform symmetry, and only three or
four directions may need to be run. These are typically the two orthogonal and the
diagonal direction. An additional case may be needed to align the platform with the
maximum expected wave direction or some other platform geometry consideration.
The need to run a pushover only in one of each of the platform perpendicular
directions assumes forward-backward symmetry. In other words, no matter which
direction the pushover is run, the ULS is the same. This is not always the case,
and the user should look carefully at the direction of the platform bracing. If one
direction puts all of the braces in compression and the other in tension, the ULS
will be different, and a pushover needs to be run for each direction. Note that
directionality of waves does not influence the ULS of the platform. Rather, it
influences the RSR and other similar platform performance metrics (refer to
Section 170).

163 Dynamic Effects


Dynamic effects need to be included for certain types of platforms. This
includes deepwater structures in water depth greater than 300 ft and some
damaged structures.

Deepwater Structures
Deepwater structures are defined within API RP2A as requiring dynamics to be
considered. Simple single degree of freedom approaches can be used to estimate a
dynamic amplification factor (DAF), based upon the platforms natural period,
which should be applied to the static base shear for the reference loads, such as the
API RP2A, Section 17, A-1 or A-2 base shear requirement. The natural period of
these structures is typically greater than three seconds. For example, a typical DAF
may be 1.1 to 1.2 for a 400 ft water platform.
The DAF is not applied to the pushover load. It is instead used to adjust the
reference level base shear upward, such as that created when using the A-1, A-2, or
100 yr wave. This, in turn, lowers the RSR for the platform. More sophisticated
dynamic analysis approaches are available, but the simple DAF approach is
adequate for most cases.

Shallow Water Structures


A shallow water damaged platform can also become dynamic if it has broken
braces across several vertical bays. The net result can be a significant loss of lateral
resistance in one of the platforms major directions, with lateral resistance provided
primarily by portal action of the legs. If this occurs, the user should determine a

100-34

2007 Chevron USA Inc. All rights reserved.

October 2007

[BookTitle]

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

DAF for the platform in the damaged state, and this should be applied to the
reference load (e.g., A-1 or A-2 base shear). Dynamic effects should be considered
if the damaged platforms natural period is three seconds or greater. Damaged
platforms that may have a dynamic issue are often those that are reported as
moving, even in small seas after the storm has passed. Refer to Figure 100-14 for
an example of how the DAF can be incorporated into the ULS results.

164 Collapse Definition


The specific reason for the collapse of the platform should be defined and
rationalized for each pushover analysis. In most cases, the collapse can be traced to
failure mechanisms that occur in the jacket/deck or that occur in the foundation
piles. A good understanding of the collapse mechanism and the fact that it makes
sense for the loads applied provides an initial quality assurance of the overall
results. However, as discussed previously and shown in Figure 100-2, the collapse
load may not always be the ULS. The ULS is the maximum load that the platform is
able to withstand prior to collapse.
Collapse occurs when one or more locations in the platform are no longer able to
resist the applied vertical or lateral loads. For example, the deck legs fail due to
large bending loads causing plastic hinging at the top of jacket due to large wave in
deck loads. Another example is multiple members failing at the same leg joint,
which leaves the platform without the ability to carry lateral loads in that region.
However, collapse modes are not always apparent and are more of a mixed mode
type with multiple members and or piles failing all at once. The primary collapse
modes are defined in the following four subsections.

Jacket Failure
There can be failure of the bracing or legs (including deck legs) such that the jacket
can no longer support vertical loads. In most cases, it takes multiple failures of the
braces to collapse the jacket, due to the redundant framing. Refer to Figure 100-9 as
an example of K brace failure in the jacket controlling the platform ULS.

Pile Axial Failure


A pullout or plunge of piles due to inadequate pile axial capacity along the pile-soil
interface and at the pile tip results in pile axial failure. Refer to Figure 100-10 for an
example of pile lateral failure (double hinging) controlling the platform ULS.

Pile Lateral Failure


A bending failure of the piles below the mudline caused by weak piles or soft soils
(or a combination) results in pile lateral failure. This is sometimes in the form of a
single bending or hinging of the piles. In other cases, it is double hinging of a pile
or piles. Refer to Figure 100-10 for an example of pile axial failure controlling the
platform ULS.

October 2007

2007 Chevron USA Inc. All rights reserved.

100-35

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

[BookTitle]

Combination
In many cases, a combination of jacket failure, pile axial failure, and pipe lateral
failure occurs almost simultaneously near collapse, and it is difficult to determine
that one or the other is the specific cause of failure.
Fig. 100-9

Jacket Failure Controlling ULS During PushoverMetocean load is acting from left to right. The colored braces
indicate member nonlinearity, with red being the largest. Red arrows identify the buckled braces in the middle
figure. The sequence of buckled K braces in the transverse direction is the collapse mechanism.

A common problem in defining collapse is that the platform displaces a


considerable distance laterally (several feet) and has not yet failed. Such large
displacements are typically the result of analytical aspects of the analysis and are
not true representations of the likely field performance of the platform. For typical
four leg to eight leg fixed platforms in water depths up to 300 feet, any results that
have lateral deck displacements more than two ft (24 in.) should be viewed with
caution. Deeper water structures and tripods or other slender platforms may have
larger displacements, but care should be taken in QA of results.

165 Quality Assurance (QA) of Results


QA of nonlinear ULS type models is more difficult than for the more common
linear models. For a linear model, once the general computer modeling problems

100-36

2007 Chevron USA Inc. All rights reserved.

October 2007

[BookTitle]

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

Fig. 100-10 Pile Failure Controlling ULS During PushoverThe figure on the left shows a lateral pile failure when a double
hinge occurs below the mudline. The figure on the right shows an axial pile failure by a combination of piles
plunging and pullout.

(such as node fixities) have been worked out of linear platform models, as used for
new platform design, the element force and displacement relationships are generally correct. There may be problems due to inaccuracies in deck loadings or member
sizing, as compared to the platform drawings, but these are typical problems faced
by every developer of platform models.
With a linear model, an applied force in one direction generally results in
displacement in that direction and linear displacements of the platform. This can
be seen in Figure 100-2, whereas load is applied up to Fyield, the resulting
displacement is linear. There may be some slight curve to the response due to the
nonlinear pile-soil foundation, but it is generally linear and is more easily verified.
If the response is not linear, there is a problem with the platform model.
Nonlinear analysis is much more difficult to determine if the results are correct.
This is because it almost always gives some sort of answerbut is it correct? Users
will often just determine the platform capacity and not pay attention to the platform
collapse mechanism, when, in fact, understanding the collapse mechanism is the

October 2007

2007 Chevron USA Inc. All rights reserved.

100-37

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

[BookTitle]

most important part of the QA process. At a minimum, the user should QA the
following situations as defined in the next five subsections:
1.

Platform Collapse,

2.

Platform Member Failure Sequence Prior to Collapse,

3.

Platform Displacements,

4.

Stress Strain Relationships in the Failed Members, and

5.

Sensitivity Checks Confirm Results.

Platform Collapse
This is probably the most direct QA. If the failure mode is in the jacket, the
associated failed braces should be slender and thin walled and likewise for piles
that fail laterally. Piles that fail due to axial loads are generally in soft soils and/or
shallow penetration. Alternatively, the deck loads may be very heavy. Platforms that
fail at the brace joints should be ones with no or very thin joint cans. The user needs
to absolutely have a good understanding of the platform collapse mechanism and
the physical reason why.

Platform Member Failure Sequence Prior to Collapse


Proper pushover analysis results in initial member failures followed by load
redistribution to alternate load paths. The resulting member failure sequences should
make sense. The members should fail in the expected order. This cannot always be
determined in all cases, but an attempt should be made nonetheless.

Platform Displacements
The platform that is being pushed in the longitudinal direction should move in that
plane, and there should be very little motion out of plane in the transverse direction.
This seems simple, but a fixed node or load that is applied out of plane during the
ramping process can easily twist the platform. An out of plane motion should be
investigated and the cause rationalized. Out of plane and twisting motions of the
jacket typically only occur near collapse as multiple members have failed. The
platform displaced shape throughout the pushover and especially near collapse
should be scrutinized. Animations should be used where available. Review of the
displaced platform shape is often the most informative QA procedure but is often
the most overlooked.

Stress Strain Relationships in the Failed Members


Review of the loads and displacements of the individual members that become
nonlinear in the pushover should be reviewed to ensure that the members are acting
as assigned. For example, the axial loads at which the most critical braces buckle in
the pushover should be hand checked against simple formulas in API or AISC. The
hand checks should be reasonably close to the buckling loads in the computer
model. The maximum pile top axial load at failure (if the foundation fails) should
be checked against the pile compression and tension capacity curves typically
available in the associated geotechnical reports.

100-38

2007 Chevron USA Inc. All rights reserved.

October 2007

[BookTitle]

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

Sensitivity Checks Confirm Results


Additional sensitivity analyses should be run to assist in confirming results. One
method is to increase the wall thickness of the members controlling collapse. This
should result in an increase in ULS or a switch to an alternative collapse mode. If it
does not, the suspected collapse mechanism is not likely correct.

170 Results
171 General
There are several key types of results that should be shown, including the relevant
acceptance criteria. These are discussed in this section. Figure 100-11 shows the
eight leg, 150-ft water depth ULS platform model used to demonstrate the example
results.
Fig. 100-11 Example Platform Used to Show Pushover ResultsTypical ULS model of Gulf of
Mexico platform with eight legs in 150-ft water depth

October 2007

2007 Chevron USA Inc. All rights reserved.

100-39

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

[BookTitle]

172 Pushover Curve


In this document, a plot of the pushover load applied to the platform as a function
of the deck displacement, as previously shown in Figure 100-2 and as shown in
Figure 100-12, is for an actual platform. The associated pushover load can
alternatively be determined as the resistance base shear at the mudline. The form
of the curve is linear initially until member failures begin to occur (K braces fail
in this example), and the curve flattens until the platform collapses. Refer to
Section 121 for a detailed description of a pushover curve.
Fig. 100-12 Example Pushover Curve

5000

Platform
Collapse

ULS
4000

Base Shear (kip)

K-Brace Failures
3000

Ultimate Capacity Curve


from Pushover Analysis
2000

1000

0
0

10

15

20

25

30

35

40

45

50

Deck Displacement (in)

173 Platform ULS


This is the maximum base shear that the platform can sustain at collapse. This
specific ULS value is used to judge adequacy of the platform and to compute the
RSR.

174 Reference Level Loads


These loads are not specifically an output of the ULS analysis but instead are
computed to help evaluate the ULS results. These are the specific load acting in the
platform for different types of reference design conditions, such as the 100 yr storm
used for many platform designs. These values are used to measure the adequacy of
the platform as determined by the ULS capacity, either as a direct comparison (e.g.,
the ULS is higher than the API A-1 requirement) or as a ratio measure, such as the
RSR. They are computed separately from the ULS and are simply the base shear

100-40

2007 Chevron USA Inc. All rights reserved.

October 2007

[BookTitle]

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

acting on the platform for the associated metocean conditions. Typical reference
level loads include:

100 Year Storm Base Shear


This is the same as the API L-1 high consequence category and is used to compute
the RSR. There are also L-2 medium and L-3 low consequence categories. Caution
is suggested in direct comparison of the ULS to any of these conditions as they
represent design wave heights, applicable when the appropriate factors of safety are
used. There are no factors of safety considered in a ULS analysis.

API Existing Platform Ultimate Strength Conditions

A-1, High Consequence, Full Population Hurricane


A-2, Medium Consequence, Sudden Hurricane
A-3, Low Consequence, Winter Storm

Reference level loads can be presented in a table with numerical values or in the
pushover plots directly (or both). Figure 100-13 through Figure 100-15 show results
for the example platform where the API RP2A, Section 17, values are shown. In this
case, the platform is categorized as A-2 and passes the assessment, since its capacity
is higher than the A-2 ultimate strength requirement. It is also helpful to put the
associated wave heights on the reference level loads as noted in order to provide a
physical meaning to the load levels. This type of presentation works well, in that it
provides a sense of how close (or how far) a platform is from these reference levels.
Fig. 100-13 Pushover Curve with Reference Level Loads Indicated
A-1 Full Population Storm Base Shear (Wave Ht = 67 ft )

5000

Platform
Collapse

ULS
4000

Base Shear (kip)

K-Brace Failures
A-2 Sudden Hurricane Base Shear (Wave Ht = 56 ft )

3000

Ultimate Capacity Curve


from Pushover Analysis
A-3 Winter Storm Base Shear (Wave Ht = 47 ft )

2000

1000

0
0

10

15

20

25

30

35

40

45

50

Deck Displacement (in)

October 2007

2007 Chevron USA Inc. All rights reserved.

100-41

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

[BookTitle]

175 RSR
RSR has been previously defined as the ratio of the ULS to the 100 yr storm
condition base shear. A specific minimum target RSR (e.g., 1.5) is often the goal.
Similar to the reference level loads, the acceptable RSR may also be shown on the
pushover plot as a horizontal line.

176 Identification of Collapse Mechanism


General
The predicated collapse mechanism for the platform should be described.

Primary Collapse Mechanism


As noted in Section 164, there are three primary types of collapse mechanisms:
1.

Jacket FailureFailure of the bracing or legs in the jacket such that it can no
longer support vertical loads.

2.

Pile Axial FailurePullout or plunge of the piles due to inadequate pile axial
capacity of the soils.

3.

Pile Lateral FailureA bending failure of the piles below the mudline caused
by weak piles or soft soils (or a combination).

Combined Collapse Mechanism


A combination of the primary collapse mechanisms occurs almost simultaneously
near collapse.

177 Graphic Sequence of Platform Collapse


Figure 100-14 shows an example of the progressive collapse of a platform. These
images improve the understanding of the collapse mode and provide additional QA
of results, as the failure sequence should make sense. For example, unexplained
movement of the jacket, such as twisting as it collapses (assuming that symmetric
loads are applied and the jacket is symmetric), means that there may be an error in
the computer model. Some software programs also provide the collapse as an
animated file that is similarly useful.

178 Identification of Critical Platform Members


The initial failure of one or two braces or other platform members usually begins
the collapse sequence. These members should be identified, and an explanation
should be apparent that these members are the first to fail. For example, these
members may be undersized, or they may be the main members for load transfer,
such as the deck legs, if there is wave in deck. This assists in the QA of the results,
as there should be a reasonable explanation why these members failed first. This
type of information can be used to develop underwater inspection plans, as these
critical members can be checked during future inspections to be sure that they are

100-42

2007 Chevron USA Inc. All rights reserved.

October 2007

[BookTitle]

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

still in good condition. However, for some platforms, there are several
simultaneous jacket and or pile failures, and it can be difficult to determine the
exact member failure sequence, but even in these cases, the members or piles that
are failing should make sense.

179 Damaged and Repaired Capacities


The effect of damage to a platform can be determined by comparing the ULS
without the damage and with the damage. This can be done by direct comparison
of the ULS numerical values, including on a percent basis, such as a 25 percent
reduction in capacity. Alternatively, it can be done graphically by comparing the
pushover curve for the undamaged or intact as built condition with the damaged
condition, as shown in Figure 100-14. In this case, the damaged platform barely
meets the A-3 criteria. Note also that the controlling collapse mode changes from
K brace failures to a lower vertical diagonal (VD) brace failure, due to the damage,
which changes the load path in the platform. Conversely, a comparison of the
damaged and repaired platform ULS can be made, although, in this case, it is a gain
in strength due to the repair. The repair is seen to return the platform to its original
pushover curve and, hence, ULS. If the damage to the platform results in a dynamic
effect and resulting DAF, as described in Section 163, the reference level loads
should be increased by the DAF. For example, Figure 100-14 shows the adjusted
A-3 condition assuming a 15 percent increase for DAF. In this case, the damaged
platform no longer meets the A-3 requirement.

1710 Change in Capacity due to Additional Gravity Loading


This includes the addition of topside weight, e.g., additional process packages or a
new drill rig that is heavier than the original design. These additional loads need to
be checked for local structural support in the deck areas, such as local beams,
bracing, and stiffeners, but they should also be evaluated for the global load on the
platform. These additional weights often do not make a large change to the ULS,
since lateral loading generally controls the ULS. However, if the jacket failure mode
is pile axial plunging, these can reduce the ULS, as the additional vertical loads
reduce the amount of available pile axial resistance that can be used to resist lateral
loading.

1711 Change in Capacity due to Additional Lateral Loading


This includes primarily the addition of conductors or risers that increase the lateral
loading on the platform. Other items that can change the lateral load are new boat
landings, barge bumpers, or other items near the water line, but these generally have
a lesser effect than conductors and risers. In most cases, these ties do not change the
ULS but instead change the reference load acting on the platform that is used to
determine if the platform is acceptable. Figure 100-15 shows how the addition of
two conductors to the platform does not decrease the capacity of the platform but
instead increases the A-2 reference load acting on the platform. In this case, the
platform has gone from being acceptable to being unacceptable by the addition of
the conductors.

October 2007

2007 Chevron USA Inc. All rights reserved.

100-43

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

[BookTitle]

Fig. 100-14 Presentations of Pushover Results for Repair of a Damaged PlatformThe repaired platform is shown to
actually improve the platform ULS to a higher level than its original as built configuration. In this case, the repair
includes strengthening of several weak joints that caused premature failure of the jacket, and this increases the
overall platform ULS. The platform was damaged in hurricane Katrina in 2005.

A-1 Full Population Storm Base Shear (Wave Ht = 67 ft )

5000
A-2 with Two New Conductors

ULS
Repaired
4000

Platform
Collapse

Base Shear (kip)

K-Brace Failures
A-2 Sudden Hurricane Base Shear (Wave Ht = 56 ft )

3000

Ultimate Capacity Curve


from Pushover Analysis

A-3 base shear using D.A.F.

Damaged A-3 Winter Storm Base Shear (Wave Ht = 47 ft )


2000
Lower VD Brace
F il

1000

0
0

10

15

20

25

30

35

40

45

50

Deck Displacement (in)

Fig. 100-15 Increase in Reference Load by the Addition of Two ConductorsIn this example, the two additional
conductors increase the A-2 base shear above the ultimate capacity.
A-1 Full Population Storm Base Shear (Wave Ht = 67 ft )

5000
A-2 with Two New Conductors

ULS
Repaired
4000

Platform
Collapse

Base Shear (kip)

K-Brace Failures
A-2 Sudden Hurricane Base Shear (Wave Ht = 56 ft )

3000

Ultimate Capacity Curve


from Pushover Analysis

A-3 base shear using D.A.F.

Damaged A-3 Winter Storm Base Shear (Wave Ht = 47 ft )


2000
Lower VD Brace
F il

1000

0
0

10

15

20

25

30

35

40

45

50

Deck Displacement (in)

100-44

2007 Chevron USA Inc. All rights reserved.

October 2007

[BookTitle]

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

1712 Specific Member Information


The specific member force deformation relationships for critical members should be
as expected. For example, long slender members should buckle, and the legs should
fail in bending. Plotting and review of this information for critical members to
ensure that the failure mode is correct, as well as the member buckling at the proper
load, are critical steps in QA of the results. If the ULS analysis is being done for a
repair, the loads acting in the members at collapse can be used for the design of a
replacement member.

180 Select References


181 Codes

October 2007

1.

API RP2A-LRFD, Recommended Practice for Planning, Design and


Constructing Fixed Offshore PlatformsLoad and Resistance Factor Design,
1st Edition, July, 1993.

2.

API RP2A-WSD, Recommended Practice for Planning, Design and


Constructing Fixed Offshore Platforms, 21st Edition, Errata and
Supplement 2, October 2005.

3.

API RP2SIM. This API publication is pending and is under development by the
Joint Industry Project, Recommended Practice for Structural Integrity
Management (SIM) of Fixed Offshore Platforms, MSL Services Corporation,
Final Report, 2007.

4.

DTI, United Kingdom Department of Trade and IndustryCommercial Code


of Practice.

5.

FEMA 356, Pre-standard and Commentary for the Seismic Rehabilitation of


Buildings, November 2000. Federal Emergency Management Agency.

6.

ISO 19901-2. Specific requirements for offshore structuresPart 2: Seismic


design procedures and criteria.

7.

ISO/CD 19902, Draft E June 2004, International Standards Organization, Petroleum and Natural Gas IndustriesOffshore StructuresPart 2: Fixed Steel
Structures.

8.

NPD, Norwegian Petroleum DirectorateResource Management Regulations.

9.

Digre, K.A., Puskar, F.J., Aggarwal, R.K., Irick, J.T., Kreiger, W.F. and
Petrauskas, C.; Modification to and Applications of the Guidelines for
Assessment of Existing Platforms Contained in Section 17.0 of API RP2A.
Proceedings 27th Offshore Technology Conference, OTC No. 7779, May 1995.

2007 Chevron USA Inc. All rights reserved.

100-45

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

[BookTitle]

182 Assessment of Existing Platforms


1.

Bea, R.G., Puskar, F.J., Smith, C., and Spencer, J.S.; Development of AIM
(Assessment, Inspection, Maintenance) Programs for Fixed and Mobile
Platforms, Proceedings 20th Offshore Technology Conference, OTC
No. 5703, May 1988.

2.

Kreiger, W.F., Banon, H., Lloyd, J.R., De, R.S., Digre, K.A., Nair, Dl, Irick,
J.T., Guynes, S.J.; Process for Assessment of Existing Platforms to Determine
Their Fitness for Purpose, Proceedings 26th Offshore Technology
Conference, OTC No. 7482, May 1994.

3.

Petrauskas, C., Finnigan, T.D., Heideman, J.C., Vogel, M., Santala, M., and
Berek, G.P.; Metocean Criteria/Loads for Use in Assessment of Existing
Offshore Platforms, Proceedings 26th Offshore Technology Conference, OTC
No. 7484, May 1994.

4.

Puskar, F.J. and Spong, R.E., Assessment of Fixed Offshore Platform


Performance in Hurricanes Andrew, Lili and Ivan, Energo Engineering, Inc.,
Final Report, January 2006.

5.

Stewart, G., Moan, T., Amdahl, J., and Eide, O.I.; Nonlinear Re-Assessment of
Jacket Structures Under Extreme Cyclic Storm Loading - Part I: Philosophy
and Acceptance Criteria, Offshore Mechanics and Arctic Engineering
Conference, Glasgow, 1993.

6.

Stewart, G., and Troman, P.S. Nonlinear Re-Assessment of Jacket Structures


Under Extreme Cyclic Storm LoadingPart II: Representative Environmental
Loading Histories, Offshore Mechanics and Arctic Engineering Conference,
Glasgow, 1993.

183 Overall Platform Modeling for ULS

100-46

1.

API RP2A-LRFD for CRC formula for buckling capacity of tubular members.

2.

Bouwkamp, J.G., The Effects of Joint Flexibility on the Response of Offshore


Towers, Proceedings 12th Offshore Technology Conference, OTC No. 3901,
May 1980.

3.

Kallaby, J., Lee, G., Crawford, C., Light, L., Dolan, D., Chen, J.H.; Structural
Assessment of Existing Platforms, Proceedings 26th Offshore Technology
Conference, OTC No. 7483, May 1994.

4.

SSRC, Structural Stability Research Council, Guide to Stability Design


Criteria for Metal Structures, Latest Edition,
http://campus.umr.edu/ssrc/html/guide.htm.

2007 Chevron USA Inc. All rights reserved.

October 2007

[BookTitle]

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

190 Terminology
The following acronyms and definitions are used in this document. Most are
common industry terms. Some are used solely as a matter of convenience in this
document.

191 Acronyms
AISCAmerican Institute of Steel Construction
BOSBottom of Steel (Typically used in the definition of the deck elevation as the
lowest point of the cellar deck beams [bottom flange].)
CESChevron Engineering Standards
CGCenter of Gravity
DAFDynamic Amplification Factor
DLEDesign Level Approach
DOEDepartment of Energy
JIPJoint Industry Project
MMSMinerals Management Service (The USA regulator of offshore platforms.)
NPDNorwegian Petroleum Directorate
QAQuality Assurance
RPReturn Period (Used to define the recurrence interval of metocean conditions,
e.g., 100 yr RP waves.)
RSRReserve Strength Ratio (Computed as the ratio of the ULS capacity to the
load acting on the platform for the 100 yr storm.)
SSRCStructural Stability Research Council (Provides information on the
ultimate strength of structural members.)
ULSUltimate Limit Strength (The capacity of the platform at collapse.)
UTUltrasonic Testing
VDVertical Diagonal
WDWater Depth

192 Definitions
Cellar DeckThe lowest major deck on the platform that typically extends
between the platform legs. See API RP2A for an expanded detailed definition.
Change of UseA formal API definition related to a platform that is modified such
that its primary use has changed. For example, a drilling platform that becomes a
hub for deepwater pipelines.

October 2007

2007 Chevron USA Inc. All rights reserved.

100-47

100 Ultimate Limit Strength (ULS) of Fixed Offshore Platforms

[BookTitle]

ElasticThe displacement of a structure in a linear manner for each applied load


increment. For elastic behavior, the platform will return to its original position prior
to load application.
InelasticThe unequal displacement of a structure for each applied load
increment. For inelastic behavior, the platform will not return to its original
position prior to load application but will instead have a permanent deformation.
Also known as material nonlinearity.
Load FactorRatio of the ULS to the base shear of the wave used to generate the
pushover profile (including wave in deck loads).
NonlinearThe response of a structure to applied loads that results in unequal
displacement in accordance with each equal load increment. Caused by inelastic
(material nonlinearity) and/or P-delta effects (geometric nonlinearity).
P-DeltaOverturning effect created by gravity load as the platform is displaced
laterally, causing an overturning moment at the mudline, which tends to pull the
platform over. Also known as geometric nonlinearity.
Quality AssuranceThe verification of the computer model to ensure that it
accurately reflects the platform configuration, topsides loading, current condition
(such as damage and marine growth), and other important factors.
Subcellar DeckA small deck located beneath the cellar deck. Sometimes called a
spider deck.
Ultrasonic TestingThe use of acoustic waves to determine the thickness of
members. This can be used to determine the leg wall thickness. UT will show a
thicker leg if the leg-pile is grouted.
Wave in DeckA special case of extreme waves with a crest elevation that is
higher than the BOS of the cellar deck.

100-48

2007 Chevron USA Inc. All rights reserved.

October 2007

You might also like