Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

OTC 19535

Modern Deepwater Site Investigation: Getting It Right The First Time


Kerry J. Campbell, Gary D. Humphrey, and Robert L. Little, Fugro GeoConsulting, Inc.

Copyright 2008, Offshore Technology Conference

This paper was prepared for presentation at the 2008 Offshore Technology Conference held in Houston, Texas, U.S.A., 5–8 May 2008.

This paper was selected for presentation by an OTC program committee following review of information contained in an abstract submitted by the author(s). Contents of the paper have not been
reviewed by the Offshore Technology Conference and are subject to correction by the author(s). The material does not necessarily reflect any position of the Offshore Technology Conference, its
officers, or members. Electronic reproduction, distribution, or storage of any part of this paper without the written consent of the Offshore Technology Conference is prohibited. Permission to
reproduce in print is restricted to an abstract of not more than 300 words; illustrations may not be copied. The abstract must contain conspicuous acknowledgment of OTC copyright.

Abstract

Deepwater site investigation for petroleum production facilities requires much more than the simple one-boring
approach that has been common at many continental shelf sites. Instead, because foundation-zone soil and
geologic conditions found at many deepwater sites can be complex, a multi-phased approach that integrates and
analyzes various geoscience and geotechnical data is required to optimize deepwater siting, geohazard risk
assessment, and foundation design. This paper summarizes best practices for planning and executing a modern
deepwater site investigation. The paper is based on our experience in carrying out various aspects of more than
100 deepwater site investigations over some 20 years. Although others have documented case histories
illustrating specific deepwater site investigations, described various data acquisition tools, and discussed
geohazards risk assessment in some detail, to our knowledge no generic, public summary exists that: 1) explains
the numerous considerations and significance of each in planning and executing a deepwater site investigation; 2)
describes the range and significance of geoscience and geotechnical components that can be required; 3) gives
the generic phases and sequence of execution required to optimize the process; and 4) provides guidance on the
cost and schedule time required to carry out a deepwater site investigation.

The paper outlines a methodology for efficient, effective deepwater site investigation that is a useful explanation
and guide for anyone responsible for this activity. The methodology will help operators optimize the planning and
execution of deepwater site investigations, including geohazard risk assessments, and deliver results to project
design teams in a timely manner. Unlike major operators, most independent operators and facilities designers/
constructors do not have site investigation and geohazard specialists on staff who are experts in the methodology
presented. This paper is written principally for these organizations and provides them with a practical
methodology for planning and executing effective deepwater site investigations.

Introduction

Purpose and Scope. The purpose of this paper is to describe what we believe to be the best-practice
approach to a modern deepwater site investigation required for various petroleum exploration and development
activities. Specifically, the paper outlines a methodology for efficient, effective deepwater site investigation and is
intended as a useful explanation and guide for anyone responsible for this activity. The paper briefly describes
what needs to be done and the various tools used, why it is important to do the various tasks involved, and when
and in what sequence they should be done. The paper generally does not describe how to do the individual tasks
as many other previously published papers have discussed details of deepwater site investigations (see
References and Other Selected Literature at end).

Definitions. For the purposes of this paper, the following definitions apply:

• Site investigation: the sequence of geophysical survey and soil data acquisition programs, soil
sample testing, and data analysis, integration, and synthesis needed to characterize and assess
seafloor and foundation zone/tophole conditions, including geohazard risk assessment, in
connection with deepwater drilling operations and facilities siting and foundation design.
2 OTC 19535

• BML: Below mudline.

• Deepwater: water depths of more than 650 ft (~200 m).

• Foundation zone: extending from the seafloor downward to as much as 650 ft (~200 m) BML.

• Tophole zone: extending from the seafloor downward to as much as 5,000 ft (~1525 m) BML.

Target Audience. The target audience for this paper is principally those who have little or no deepwater
experience but who are a “stakeholder” in deepwater site-investigation work. This includes not only those who
are directly responsible for planning, specifying, and managing deepwater site investigations, but those in upper
management, asset development and budget planners, those responsible for permitting, and the principal end
users of the technical information including facility designers and risk analysts. As a practical matter, this target
audience is largely comprised of those associated with independent and small energy companies and various
third-party facilities designers who do not have geohazard and geotechnical site-investigation specialists on staff.
In sharp contrast, virtually all major energy companies now recognize the importance, the optimum approach,
typical costs, and extended schedule required for deepwater site investigations. Today, the largest energy
companies have one or more in-house teams of site-investigation specialists who are experts in the various
aspects of deepwater site investigation.

Background. In the early days of the offshore oil business, a site investigation invariably meant a soil boring
investigation. Because geologic conditions in the foundation zone are typically relatively simple on the continental
shelf (although there are important exceptions), this “soil-boring-only” approach was generally adequate and held
sway for several decades. Although MMS (and predecessor)-required geohazard surveys have been done since
about 1970, these were initially largely used to satisfy permitting requirements in connection with well drilling.
Only when the industry moved into deep water beyond the edge of the continental shelf (>650 ft water depth) did
shallow geologic conditions become a common concern, and the need for more robust site investigation became
apparent.

Today, we know that deepwater site conditions can be complex and difficult, and can cause siting, foundation
design, and construction problems. Indeed, some consider assessment of deepwater geohazards to be “mission
critical” to successful deepwater development. Deepwater conditions that can cause serious engineering
difficulties for facility siting and foundation design, installation, or performance, and thus increase development or
maintenance costs, include: 1) irregular, sometimes rocky, topography with sharp, local relief ranging from a few
feet to several tens of feet or more; 2) steep and potentially unstable slopes of as much as 45° or more and up to
several hundred or more feet high; 3) shallow over-pressured gas or water sands (with potential for shallow-
water-flow); 4) mud volcanoes or other fluid vents and associated flows; 5) active faults with seafloor scarps
ranging up to several hundred feet high and which could damage or sever production-well casing if fault offset
occurs; 6) soils containing gas hydrates (solid, ice-like mixtures of gas and water) that can compromise
foundation and well integrity when heated by proximity to production wells; 7) both modern slides and slumps and
ancient (buried) slide deposits; and 8) materials ranging from weak, under-consolidated soils to rock. In some
areas, seismic shaking or salt tectonics, as examples, can exacerbate and further complicate conditions.
Campbell (1999) discusses deepwater geohazards and their potential engineering significance. Hill (2001)
discusses the business impact of deepwater geohazards. Papers by Hogan and others (2008), Niedoroda and
others (2003), Nowacki and others (2003), Orange and others (2003), and Shipp and others (2004) discuss
different aspects of offshore geohazards.

To reliably characterize and assess deepwater site conditions, a modern site investigation requires both
remote sensing (geophysical survey) and site-specific testing (geotechnical and geological sampling and testing),
and integration and synthesis of all results. Special acquisition and testing tools and more extensive analyses are
required, significantly increasing both cost and time required to complete a typical deepwater site investigation
when compared with most investigations on the continental shelf. From our long experience, however, many
deepwater site-investigation stakeholders still have (unrealistic) expectations of schedule and budget apparently
rooted in their past continental shelf experience. This ongoing circumstance has prompted us to write this paper,
which is based on our experience in carrying out various aspects of more than 100 deepwater site investigations
over some 20 years. Our hope is that this paper: 1) will be a helpful explanation and guide for those needing an
introduction to deepwater site investigation; 2) will give stakeholders an appreciation for the relative complexity of
deepwater site investigation; and 3) will give those responsible for commissioning investigations a methodology
for selecting appropriate technology and developing realistic site investigation approaches, schedules, and
OTC 19535 3

budgets. In short, we hope that this paper will help those inexperienced with deepwater site investigation to “get it
right” the first time.

Conceptual Approach to Deepwater Site Investigation

Philosophy. A modern deepwater site investigation: 1) begins before the first wildcat well is drilled; 2)
incorporates geophysical, geological, geotechnical, and met-ocean components; 3) uses the appropriate data
acquisition and analysis tools, with tasks executed in optimal sequence; 4) allows adequate time to analyze data
between phases; 5) uses results of each phase interactively to plan and guide subsequent phases, and to help
interpret results of previous phases; 6) assesses the significance of geological conditions and geohazards along
with soils and other earth materials such as gas hydrates; 7) requires special analyses, takes longer, and costs
more than a typical site investigation on the continental shelf; and 8) provides optimal siting and design solutions,
reduces risk, facilitates project fast-tracking, and saves operators money overall if carried out appropriately.

Ideally, a deepwater site investigation should begin before drilling the first exploration well – using 3-D
exploration seismic (3DX) data during the pre-drilling geohazards assessment – and continue in phases through
completion of final siting and foundation design. The reason for the very early start is to help avoid drilling
(expensive, potential production) wells at sites that could be unfavorable for long-term development. Because of
their shorter term, exploration-drilling activities generally are less sensitive to many active geologic processes
when compared to the longer-term activities associated with development and production. Thus, conditions that
generally would be of little or no consequence to exploratory drilling activities (typically carried out over a few
months at a given wellsite), could pose a major risk to production facilities that will be operating onsite for years.
Although it is not (yet) common, we are encouraged by an increasing number of cases where geotechnical
engineers or others normally concerned with development attend (exploration) drilling planning meetings to offer
perspective and advice about site suitability for long-term development.

Principal Components and Investigation Sequence. – The principal components of a modern deepwater
site investigation include both geophysical survey and geotechnical data acquisition programs and the associated
integrated data analysis (“geotechnical” program includes some combination of shallow gravity coring or rotary
drilling, down-hole logging, in situ testing, soil sampling, and laboratory testing). Geophysical data defines
geologic conditions and the 3-dimensional stratigraphic framework on which the geologic model is based,
whereas “geotechnical” data provides, as examples, quantitative geotechnical properties of soils and allows both
materials and age-date “calibration” of the geologic model. The site investigation is done in phases, with
knowledge of site conditions being increasingly refined as a result of each phase, and with subsequent phases
guided and modified by results of previous phases.

The site investigation should always begin with seafloor and tophole interpretation of the 3DX data. The
preliminary geologic/geohazards model developed from the 3DX data is subsequently refined using high-
resolution site-survey data. Ideally, the high-resolution site survey should always be completed and the data
evaluated before initiating the geotechnical program, with the results of the site survey being used to help plan the
geotechnical program. Specifically, the knowledge of geologic conditions is used to optimize the numbers, types,
and locations of geotechnical cores/boreholes, and to help design the logging, testing, and sampling programs at
each location. Conversely, there should always be engineering input when designing the site survey to ensure
that the data to be collected will be adequate to meet all engineering objectives. Specifically, the production
design concept(s) being considered, all critical issues relating to siting, field layout, and foundations, the depth of
the foundation zone, and any other engineering issues must be defined and taken into account when designing
the survey.

Data Integration and Site Characterization. A modern deepwater site investigation thus requires that
geophysical and geotechnical data be interactively analyzed, interpreted, synthesized, and effectively integrated
to provide a comprehensive, coherent characterization across the site. Site characterization includes defining:

• Water depth and seafloor topography (as illustrated by bathymetric maps, DTM files, and Fledermaus
visualization files, as examples).
• Geotechnical properties of materials (soil, rock, gas hydrate) and 3-dimensional relationships among
soil strata (as shown by various maps and cross sections, soil sampling and in situ testing logs,
geophysical down-hole logs, lab test results, soil province maps, and 3-D predictive soil models, as
examples).
4 OTC 19535

• Geologic conditions, including structure, stratigraphy, past events/processes, shallow overpressures,


identifiable features, etc. (as shown on geologic features and other maps, and by geologic test results,
as examples).
• Active geologic processes such as faulting, seafloor instability/mass transport, turbid flow, fluid
venting, seafloor scouring by currents, etc. (as determined by time-lapse surveys, direct measurement,
or visual observation, as examples).
• Rates, frequencies, recurrence intervals, and magnitudes of active geologic processes as a basis for
quantitative risk assessment (as derived in part from interpretation of seismic data integrated with age
dates of key strata, for example).

Results typically are delivered in both digital and paper format. Increasingly, clients are requiring that site
characterization results be delivered in GIS (Geographic Information System) and 3-D visualization format
(Fledermaus format, for example). GIS delivery allows for rapid customization of maps and more robust analysis
of site conditions. Quantitative seismic geomorphology, an analysis approach increasingly being applied offshore
as part of site investigations, is facilitated by GIS tools. Three-dimensional visualization helps users readily
visualize seafloor and subsurface conditions and helps them make better decisions faster. Doyle (1998),
Jeanjean and others (two papers, each 2003), Kolk and others (1997 and 2005), Liedtke and others (2002), and
Moore and others (2007) discuss various aspects of marine integrated site investigations and study results.

Geohazard Risk Assessment. Quantitative (probabilistic) or semi-quantitative geohazards risk assessment


is required to help identify and prioritize potential engineering issues and, ultimately, to quantify the impact of
geohazards on project economics. Fenton and others (2002) discuss a quantitative geohazards risk assessment
methodology used for an offshore development. Power and Clayton (2003) discuss site investigation and risk
assessment methodologies designed specifically for application to deepwater development. Several other papers
have been written in recent years that address marine geohazard risk assessment. Among them are papers by
Bryn and others (2007), Evans and others (2007), Galavazi and others (2006), Jeanjean and others (2005),
Kvalstad (2007), and Power and others (2005).

Idealized Site Investigation Phases

The following is a 15-phase idealized sequence of phases undertaken to complete a deepwater site
investigation. In practice, for numerous reasons, virtually no investigations actually include all phases exactly as
outlined here. However, this is one generic approach. There are other valid approaches, and the phases given
here are intended to be tailored as necessary to the circumstances of each project. Because of the offshore
leasing approach in the U.S., and MMS regulations that are not specifically addressed here, some modification of
the details would be appropriate for deepwater developments off the U.S. For similar reasons, some modification
might also be appropriate for site investigations off other countries. However, this general, multi-phase approach
is still valid overall, regardless of location.

Phase 1 - Regional Geohazards Screening (for non-U.S. sites, typically thousands of square kilometers
using 3-D exploration seismic (3DX) data; done before exploration drilling begins)

• Determine type, general distribution, frequency-of-occurrence, and severity of geohazards, for


both drilling and development
• Determine adequacy of 3DX data for pre-drilling geohazards assessment
• If not adequate, decide on 3-D short-offset re-processing, or acquisition of HR2-D data

Phase 2 - Pre-drilling Geohazards Assessment of Prospect Area (typically 10 to 100 square kilometers
using 3DX, 3DX short-offset reprocessed, and/or HR2-D data)
• Carry out detailed drilling hazards assessment of prospect area

Phase 3 - Detailed, Wellsite-Specific Assessment


• Do detailed drilling hazards assessment of each proposed wellsite, and develop Tophole
Prognosis Charts

Phase 4 – Preliminary Engineering Assessment (Typically based on 3DX exploration seismic data after
discovery made at deepwater sites)
• Do preliminary site characterization (refine geologic and soil models for engineering application)
• Do preliminary siting and foundation design assessment
• Develop geohazards register to determine engineering implications and likely degree of risk
OTC 19535 5

• Identify foundation design considerations/assess concept options


• Develop (for example) preliminary pile sizing/capacity based on inferred soils

Phase 5 – High-Resolution Geophysical Survey-Program Planning


• Define engineering design concept(s)/objectives/critical issues/foundation zone
• Define survey objectives
• Specify survey tools
• Define survey grid(s)
• Develop survey strategy

Phase 6 – High-Resolution Geophysical Survey Program


• Carry out AUV, HR2D, or other high-resolution geophysical survey work
• Do preliminary interpretation as survey progresses
• Revise survey plan as necessary during survey based on preliminary results
• Survey supplemental lines, if necessary
• Carry out micro 3-D or other special survey, if appropriate
• Carry out shallow sampling program and core logging if part of geophysical survey program

Phase 7 – Geophysical Survey Data Processing/Preparation


• Multi-beam data QC review, datum check, and gridding and contouring
• Prepare side-scan sonar mosaic
• Complete 2-D or 3-D seismic data processing
• Load all data onto interpretation workstation

Phase 8 - Preliminary Site Characterization (Interpretation and mapping, interactive geoscience and
geotechnical effort)
• Map water depth and seafloor topography
• Refine preliminary predictive soil model
• Map geologic conditions in foundation zone, identify potentially active geologic processes, and
refine preliminary geologic model
• Refine preliminary assessment of soils and geohazards

Phase 9 - Geotechnical-Program Planning (note that the term “geotechnical program” is used here to
include all down-hole logging, geotechnical and geological sampling, and in situ testing)
• Use geotechnical objectives in conjunction with the preliminary soil model to develop strategy for
the geotechnical investigation (select numbers, locations, and depths of soil borings and sampling
points, and in situ tests)
• Use geologic model and preliminary geohazards assessment to develop geological sampling
strategy

Phase 10 - Geotechnical Program


• Carry out shallow coring and rotary boring work, down-hole wireline logging, geotechnical and
geological sampling, and insitu testing
• Do on-board sample logging and testing, including multi-sensor core logging (density, resistivity,
magnetic susceptibility, p-wave velocity, etc.)
• As the program progresses, compare preliminary results to preliminary geologic and soil models
• Determine if revisions to the geotechnical/geological boring/sampling/in situ testing plan are
needed (interactive geoscience and geotechnical effort)

Phase 11 – Sample Packaging and Shipping to Onshore Lab


• Special climate-controlled shipping (frozen, unfrozen, special cryogenic containers for gas
hydrate samples, etc., depending on program particulars)

Phase 12A - Geological Laboratory Testing


• Special geological testing (radiometric or other age dating, detailed stratigraphic analysis using
thin-slab x-ray radiography or digital tomography, paleontological analysis for environment of
deposition, etc.)
• Multi-sensor core logging (density, resistivity, magnetic susceptibility, p-wave velocity, etc.), if not
done onboard
• Revise geological testing program as necessary based on initial results
6 OTC 19535

Phase 12B - Geotechnical Laboratory Testing


• Standard and advanced (for example, dynamic) geotechnical testing
• Revise geotechnical testing program as necessary based on initial results

Phase 13 – Seismic Inversion and Development of Final Geotechnical Criteria (Intensive geoscience/
geotechnical interaction required)
• Finalize predictive soil model (use seismic inversion, if appropriate)
• Use geophysical data/soil model to help define soil strata and finalize boring logs
• Assign static soil parameters for foundation design
• Determine soil parameters for advanced analytical studies (for example, dynamic loading, slope
stability, etc.), if required

Phase 14 – Geohazard Assessment, Special Engineering Analyses, and Risk Assessment (Intensive
geoscience/geotechnical interaction required)
• Analyze results of special geological testing
• Analyze active/inactive geologic processes (quantitative site/geologic process characterization):
• Sliding and slumping (seafloor stability)
• Venting
• Faulting
• Scouring
• Gas hydrate formation
• Seismicity, etc.
• Quantitative or semi-quantitative risk assessment (probability of occurrence x potential effect on
facilities; development of f/n curves, etc.)

Phase 15 – Finalize Integrated Site-Characterization Model and Prepare Integrated Report


• Joint geoscience/geotechnical effort
• Package and deliver results ready to be used for economic risk assessment, final siting, and
foundation design.

Data Acquisition Tools

Geophysical Survey Tools. There are three principal types of geophysical survey tools that are most useful
for deepwater site characterization and geohazards assessment. These are: 1) 3-D exploration seismic (3DX)
systems that provide low-resolution data for preliminary site characterization and assessment; 2) Autonomous
Underwater Vehicle (AUV) or DeepTow (DT) systems that provide design-level high-resolution swath bathymetric,
seafloor image, and shallow subbottom profile data in the upper part of the foundation zone; and 3) ultra-high-
resolution or high-resolution 2-D or 3-D (UHR2D, HR2D, or HR3D) multi-channel seismic systems that provide
deeper-penetrating subbottom profile data throughout the foundation and tophole zones, respectively.

3DX Data. Conventional or reprocessed 3DX data is now commonly, if not quite universally, available for
deepwater sites. This type of data is readily available on energy company workstations and is routinely being
used by explorationists for prospect generation and wellsite selection, and by geohazards specialists for pre-
drilling geohazards assessments of the tophole section. Although this “free”, essentially-immediately-available,
data has tremendous value to facility engineering aspects of early field development planning, it is often
overlooked for these applications. 3DX data is used to provide a preliminary site characterization and give an
early indication of the types, severity, and general distribution of geohazards. It can also be used to infer overall
soil conditions in the foundation zone and tophole section.

The chief limitation of 3DX data is relatively low (horizontal and vertical) resolution when compared to high-
resolution site-survey data. 3DX data is designed to principally image conditions at reservoir level, and low
frequency sound sources must be used to acquire the data from these relatively great depths, resulting in low-
resolution data [areally, data points are typically on 40 to 80 ft (~12.5 to 25 m) centers; vertical resolution is
typically 25 to 30 ft (~8 to 10 m)]. Nonetheless, 3DX data should always be used early in the exploration
sequence to derive as much site information as possible, and can often boost one’s knowledge of site conditions
to the notional 50 to 80 percent level. Obviously, the longer the use of 3DX data is delayed, the less value it will
have to the development project as more decisions about production concept feasibility and field layout already
will have been made.
OTC 19535 7

In some cases 3DX data is “short-offset” re-processed to improve both horizontal and vertical resolution.
However, as more people come to appreciate the engineering and other value of seafloor and tophole information
derived from 3DX data, more attention is being paid to optimizing the shallow part of the section during acquisition
and/or the original processing sequence, obviating the need for supplemental short-offset re-processing.

AUV Data. Data collected with an AUV system typically includes: 1) Multi-Beam Echo Sounder (MBES) swath
bathymetric data that defines details of seafloor topography and water depth over the entire seafloor with gridded
data points nominally spaced on 6 ft (~2 m) or so centers; 2) side-scan sonar data that is closely analogous to an
aerial photo of the seafloor with resolution typically better than 3 ft (~1 m); and 3) subbottom profiler data that
shows a vertical section of acoustic (as proxy for soil) stratigraphy to penetrations ranging from a few to 225 ft
(~70 m) or so, depending on geologic conditions. The profiles give fine details of shallow stratigraphy and
geologic structure, with vertical resolution typically being on the order of 1 to 1.5 ft (~30 to 50 cm). Backscatter
(strength-of-seafloor-reflection) data is also recorded from the MBES data and is useful for mapping changes in
seafloor materials. George and others (2002) discuss AUV site characterization surveys done in support of
deepwater field development.

AUVs are tetherless, and are thus capable of executing complex or compound (“designer”) survey patterns
comprised of curved, very short, or closely spaced lines. This capability allows AUV data to be collected, for
example, along circular lines connecting proposed anchor locations in a spread mooring pattern (allowing
alternative sites to easily be evaluated should the pattern later need to be rotated), and allows micro 3D surveys
to be done where local details need to be defined in 3-dimensional detail. An example of this specialized
application is given in Campbell and others (2005).

Older-generation, tethered DeepTow (DT) survey systems are used for some deepwater surveys instead of
the more modern, free-flying AUV systems. Although DT systems generally provide the same types of data an
AUV does, DT survey efficiency is typically much lower because of the wide, slow line turns required due to the
long tether by which DT sensors are deployed. The long tether also prevents DT systems from executing the
circular or other complex survey patterns that AUVs are easily capable of. Finally, depending on the method of
bottom tracking used, DT data quality can be unduly compromised in areas of rugged topography.

Because AUV and DT data provide only the details of the seafloor and shallow foundation-zone, they are
typically not collected until after development has been sanctioned. AUV or DT data typically would not be
collected for engineering purposes prior to exploration drilling because of its limited additional utility for this
application vs. 3DX data. However, AUV data are routinely collected over deepwater Gulf of Mexico blocks for
which the MMS requires pre-drilling archeological surveys.

UHR2D, HR2D, and HR3D Data. Depending on the foundation concept and depth of the foundation zone
below the seafloor, AUV data alone may not image the full depth of interest. In these cases, UHR2D (ultra-high
resolution 2D) or HR2D (high-resolution 2D) multi-channel seismic data typically would be needed to supplement
the AUV data. Also, UHR2D or HR2D data may be needed to investigate somewhat deeper conditions that may
affect the foundation zone, or to characterize the tophole zone to help assess the long-term integrity/stability of
production wells. As examples, UHR2D or HR2D data could be needed to 1) characterize faults at depth that
extend upwards into the foundation zone, especially if they may be acting as conduits for upward migration of gas
or other fluids, or 2) to investigate the presence of gas hydrate below the foundation zone.

UHR2D data typically provide penetration from several hundred to 1,600 ft (~200 to 500 m) BML, depending
on tool configuration and geologic conditions. Vertical resolution is on the order of 3 to 5 ft (~1 to 1.5 m). HR2D
data is similar to UHR2D data, but typically images the entire tophole zone (to 5,000 ft or ~1525 m) and has
somewhat lower vertical resolution of about 10 ft (~3 m). UHR2D data would be collected to give the best detail in
the foundation zone. HR2D data typically would be collected if both the foundation zone and the entire tophole
zone both need to be investigated. In rare cases where the AUV data does not quite image the entire foundation
zone, a combination of AUV and 3DX data may instead be adequate as the geophysical survey component of a
deepwater design-level site investigation. Regardless, a UHR2D or HR2D survey spread should always be
available for deployment from the AUV survey vessel in case the AUV data not adequately image the zone of
interest.

Because of the relatively shallow penetration of UHR2D data, like AUV data, UHR2D data are generally not
collected until field development is sanctioned. Increasingly, the same is true for HR2D data. Even though
penetration of HR2D data typically extends through the entire tophole zone, most operators now use only 3DX
data for deepwater pre-drilling geohazards assessments during the exploration drilling phase. This approach has
8 OTC 19535

been used successfully for some 15 years by an increasing number of operators, and eliminates the time and
expense of a dedicated HR2D survey during the exploration drilling phase (although a high-resolution site survey
is always eventually needed as the survey component of a deepwater design-level site investigation).

HR3D data is superior to HR2D data because a 3-dimensional “volume” of data is available rather than data
being available only along widely spaced individual lines. Data points (bins) typically would be spaced on 20 ft
(~6.25 m) by 25 ft (~7.5 m) centers. As a result, HR3D data allows more robust interpretations to be made faster
and with higher confidence than HR2D data does, especially in areas of complex geology. However, acquisition
of HR3D data requires close hydrophone spacing of, say, 80 ft (~25 m) (when using a multi-hydrophone array)
compared to HR2D data, for which line spacing would typically be 320 ft (~100 m) or more. Thus, acquisition of
HR3D data either requires numerous passes by the survey vessel (pseudo 3-D survey) or a special multi-
hydrophone spread with precision positioning to collect up to 6 lines of data simultaneously. In either case, the
cost of HR3D surveys is considerably more than for conventional HR2D surveys. Consequently, HR3D surveys
are uncommon, and today typically are carried out only in special cases where complex geology is unavoidable
and risks are high.

ROVs and Manned Submersibles. Increasingly, geophysical survey tools mounted on Remotely-Operated
Vehicles (ROVs) are being used to investigate critical details of small areas or where survey data would otherwise
be difficult to collect by other means (such as on very steep slopes). ROVs are also used where visual inspection
is required to, as examples, investigate fine-scale evidence for venting, faulting, or other active processes, or to
control or confirm and document careful, undisturbed collection of soil samples for precision dating. Hewitt and
others (2008) discuss the related application of using AUV and ROV data together to predict environmentally
sensitive biological communities in deepwater. Although uncommon with respect to industry application, manned
submersibles have been used for visual investigation of seafloor conditions as a component of some deepwater
site investigations in the Gulf of Mexico.

Geophysical Down-hole Logging Tools. Down-hole logging tools are used to provide continuous records of
several properties of the seabed formations. The tools are often configured to record measurements of a number
of different properties concurrently in one deployment and can log at rates of 5 m/min (~16 ft/min). Some of the
types of down-hole tools commonly used in deepwater site investigations and applications for the data they
provide are listed below:

• Natural gamma ray emission: Lithological log showing mineralogical variations in the stratigraphy;
• Gamma density: Bulk formation density, derived porosity, and chemical composition of formation rock;
• Neutron porosity: Formation porosity;
• Electrical resistivity: Resistivity is dependent primarily on the chemistry of the pore fluids, formation
porosity and how interconnected the pore spaces are. Consequently, resistivity is a key indicator for gas
hydrates and helps in distinguishing soil type and mineralogical changes;
• Full sonic waveform: Velocity and waveform data used to correlate the various logging data with seismic
records and to assist with stratigraphic interpretation;
• Caliper: Borehole diameter. Awareness of variations in the diameter allow better interpretation of other
logging data and indicate changes in the formation competency; and
• Vertical seismic profiling: Provides seismic velocities used to correlate the various logging data with
seismic records.

Other logging tools are available for recording borehole fluid temperature and pressure, borehole orientation,
spontaneous potential, and high-resolution color-coded images showing electrical resistivity.

Geotechnical/Geological Sampling and In situ Testing Tools. The two approaches that are typical for
deepwater sampling and testing are the seabed mode and the down-hole mode. Seabed-mode methods provide
soil data to adequately characterize the shallower strata, but do not require the use of drill pipe or rotary drilling
techniques. However, the penetration capability varies with soil type and consistency, and is usually limited to
penetrations of less than 160 ft (~50 m). Seabed mode methods typically include use of large-diameter piston
core samplers together with in situ testing performed from a seabed frame deployed from the investigation vessel.
The large-diameter corers typically retrieve continuous sediment samples of approximately 4-in (~100-mm)-
diameter from the seabed to about 60 to 80 ft (~20 to 25 m) BML. The larger-diameter corers are favored as they
induce less disturbance than smaller-diameter corers and thus better preserve sedimentological details. Seabed-
mode in situ testing equipment often includes piezocone penetrometer testing (PCPT), vane shear tests (VST), T-
OTC 19535 9

bar, and ball penetrometer testing. These tests can be performed from the mudline to penetrations of up to 160 ft
(~50 m) BML. Peuchen and Rapp (2007) discuss some of these tools in more detail.

Seabed mode sampling can also include use of smaller-diameter gravity corers and box corers. These corers
are easier to deploy from smaller vessels and can often be deployed during the geophysical survey. However,
these corers only achieve shallow penetration, with box corers limited to the upper 1.5 to 3 ft (~0.5 to 1 m) BML,
and the small-diameter gravity corers usually limited to less than 15 ft (~4.5 m) penetration. Newly developed
methods of testing surficial sediments include the “Deck Scout” for “in situ” testing of the seafloor sediment
recovered in box cores. The Deck Scout characterizes the response of the soil to cyclic loading, which is
particularly important for seabed-founded structures and pipelines.

Soil characterization for deep driven-pile foundations associated with tension-leg platforms (TLPs), compliant
platforms, or driven-pile solutions for mooring of SPARS, semi-submersibles, FPSOs, etc., require the down-hole
mode of sampling and in situ testing. The down-hole mode is required to reach sufficient penetrations to
adequately characterize the geotechnical properties for deep-foundation solutions. The down-hole mode involves
advancing a borehole into the seabed using rotary-boring techniques. Coring tools are deployed and retrieved
through the drill pipe to sample sediment below the bit in advance of the borehole. Coring tools include: pushed-
in tubes and piston tube samplers for normal foundation-zone sediment; specialty rotary corers and percussion
corers that can sample more competent hard and dense soils and rock; and pressurized corers to retrieve
sediments at their ambient pressure (especially important in assessing gas-hydrate saturation). In situ testing in
the down-hole mode includes most tests available in the seabed mode (PCPT, VST, and ball penetrometer), plus
specialized tools for in situ pore water sampling, pore-pressure dissipation testing, and soil conductivity testing.

Laboratory Testing

Geological Laboratory Testing. The principal reasons for doing geological testing of seafloor and
foundation-zone soils are to determine age dates and environments or mechanisms of deposition. Age dates
indicate how long ago geologic events occurred (seafloor mass-transport failures and mud volcano eruptions, as
examples), provide a basis for estimating the recurrence interval of events (repeated episodes of mass-transport
failures, intermittent fault offset, as examples), and allow average rates of sedimentation or fault offset to be
determined. This information is important for defining and understanding the geologic model, which becomes the
basis for quantitative geohazards risk assessment. Some age-dating methods include: carbon-14, cesium-137,
oxygen-isotope, and micropaleontology. Each has specific applications and limitations, and each is useful in
different situations. Some aspects of marine-sediment age dating done in connection with a deepwater
development in the Gulf of Mexico are discussed by Slowey and others (2003).

Knowledge of environments or mechanisms of deposition can help confirm origins of sedimentary units
interpreted from seismic data and are commonly used to differentiate among mass-transport deposits, turbid-flow
deposits, and “normal” sediments deposited slowly and uniformly from suspension. Thin-slab x-ray radiography
or digital CAT scanning (tomography) of samples, coupled with detailed analysis by expert sedimentologists, and
micropaleontology, are commonly used for these purposes. Other special geological testing, such as x-ray
diffraction or heavy minerals analysis, may also be useful for some applications. Another useful geologic testing
tool is the Multi-Sensor Core Logger (MSCL). This device provides non-destructive readings of P-wave velocity,
bulk density, electrical resistivity, magnetic susceptibility, color reflectance, and natural-gamma-ray emission for
either whole or split cores.

Geotechnical Laboratory Testing. Results of the geotechnical laboratory testing program conducted for
deepwater development investigations may be used for three different purposes: 1) to characterize the soil
stratigraphy and properties at specific locations where structures are to be installed so that appropriate design
parameters may be assigned; 2) to define soil conditions within particular geologic units, allowing inferences to be
made about soil conditions located away from actual sampling/in situ testing sites; and 3) to determine the
engineering properties of sediments susceptible to geohazards (sliding/slumping or erosion by seafloor currents,
as examples) that may pose risk to the development. Typically, during a geotechnical site investigation, much of
the standard geotechnical testing for soil classification and static strength is performed offshore on the
geotechnical vessel. These tests typically include the physical description, water content, unit weight, and
measures of undrained shear strength of soil samples. These test results are typically available early and are
helpful in calibration of the well logging and geophysical survey data.

Tests that require measurement of very small weights, such as particle-size distribution tests and Atterberg
limit tests (used in soil classification) as well as more advanced tests, require a stable work platform and special
10 OTC 19535

equipment not available in the offshore laboratory. Advanced tests are designed to investigate the stress history
and compressibility relations of the sediments, normalized strength relations, and the response of the soils to
dynamic and cyclic loading. When designing the laboratory testing program, the goal should be testing-for-
purpose: test assignments should be formulated to provide, as necessary, the information for site-specific
foundation design, generalized site characterization, and sediment susceptibility to geohazards (sliding/slumping
or erosion by seafloor currents, as examples).

Integrated Data Analysis

The primary goals of integrated data analysis are to provide a characterization of the site in terms of:

• the geologic setting,


• the geohazards,
• the quantification, evaluation, and management of geohazard risks, and
• foundation soils.

To best achieve these goals, the data analysis process needs to be one that integrates all the data into a
coherent geologic model – one that plausibly unites the current geologic setting with past and present
environmental conditions, identified geological processes, and rational soil and rock mechanical principles.

Geologic Model. Data reduction and analysis to achieve an integrated characterization of the site first
requires loading of all geophysical and soil data sets onto networked seismic interpretation workstations. This
then allows mapping of the topography and sub-surface horizons, identification of significant geologic features,
and evaluation of geologic processes, either relict or active, which explain the development of the observed
present conditions. To the extent feasible, rates, frequencies, and magnitudes of geologic processes are
quantified. Results of this analysis allow the development of a conceptual geologic model that can be used in the
evaluation of potential geohazards and in the evaluation of soil conditions across the development area. As
applied to the seafloor and deeper foundation-zone sediments, the geologic model typically includes a spatial
organization of the major geologic features, location and effects of faults and other structural elements,
consideration of tectonic events, stratigraphic evaluation by geologic units, and descriptions of the depositional
and erosional environments.

Petrophysical data from well logs, geologic data, and geotechnical data from laboratory and in situ testing
programs are all integrated with the geophysical data on the interpretation workstation. Correlation of these four
data sets provides specific information that can be used to cross-calibrate the data. This integration expands the
applications of the geologic model, allowing inferences to be made about the subsurface conditions and
engineering properties of the soils at sites away from individual well, borehole, and core locations.

Interdisciplinary Team. Data analysis in deepwater investigations demands evaluation of conditions from a
variety of disciplines. Critical aspects of a site’s condition revealed during data analysis by one discipline may
have implications for issues customarily addressed by other disciplines. For example, a pipeline engineer might
conclude that certain seafloor currents wouldn’t impart excessive loads directly on the pipe, but a geotechnical
engineer with knowledge of the seafloor soils might conclude that severe erosion resulting in spanning would
occur. Thus, compartmentalizing a project entirely into individual specialties, without an established process for
an exchange of information between disciplines, can result in critical conditions being temporarily overlooked.
These overlooked conditions, once discovered, could change the emphasis and objectives of later investigation
phases, the overall site characterization, or even the development plan itself. Such events not only affect the
schedule to complete the investigation, but may leave data gaps that result in unnecessary development costs.
To reduce the chances of such mis-steps, an interdisciplinary team should be tasked with overseeing the data
analysis and interpretation. The overall team goal, of course, is to get the right information to the right decision
makers as early in the development process as possible.

Typically, the team should ideally include at least three primary investigators that are working in coordination
to ensure an interdisciplinary perspective is brought to the evaluation of the combined data sets:

• an operator’s representative to identify the current concepts for development, and the operational,
reservoir, and construction concerns and scheduling needs;
• a geotechnical engineering lead to coordinate the development of foundation design parameters
and the geo-mechanical interpretation and analyses of geohazards, and
OTC 19535 11

• a geoscience lead to guide the geophysical survey data interpretation, characterize the seabed
morphology and subsurface geologic features and conditions, define stratigraphic units, and develop a
quantitative geologic processes model for evaluation of geohazards.

As the analyses proceed, the team can identify other technical specialties that may need to be incorporated,
either on an interim basis or for advanced studies. Examples include: oceanographic specialists to help
investigate the effects of deepwater currents on the facility structures and the foundation response;
paleontological and sedimentological specialists to provide depositional and geochronologic (age-dating)
information necessary in establishing the extent and frequency of past geologic events; geochemical specialists to
help determine the recency of activity of geologic processes, such as venting and pockmark formation; and
facilities designers to help ensure that the deliverables identified by the analysis team are directly applicable to
end-users’ needs.

Advanced Analytical Studies. Advanced analytical studies may be warranted to provide geohazard
quantitative risk assessment (Galavazi and others, 2006) and for deepwater foundation design. Geohazards in
deepwater developments are often complex, such as slope failures which can initiate submarine slides, debris
flows, and turbidity currents. To evaluate these risks quantitatively, the probability of an event occurrence must
be established, the consequences of an event impacting facilities must be quantified (inspection costs,
replacement/repair costs, loss of production, etc.), and mitigation plans assessed. Computer modeling is often
used in evaluation of consequences and for assessment of mitigation measures. The modeling required is not
necessarily ready for use, but must often be designed specifically for the conditions unique to the development
site. An example is the evaluation of potential impact of a slope instability upslope of a seabed structure. The
evaluation may require modeling of the non-linear seismic ground motion propagation through the soil column
above the base rock to the seabed, application of the seabed accelerations from these studies into slope stability
analyses, incorporation of the slope stability results into mass gravity flow models, and conversion of the
hydrodynamic flow characteristics into forces and loads impacting seabed structures.

Deepwater foundation designs may also require more rigorous analytical studies than those for shallow water
environments. Special analytical studies may be necessary to address geohazards effects on soil foundation
response (for example, the dissociation of biogenic gas-hydrates within the foundation-zone sediments), or the
effects of environmental loads (loop current effects), or simply to optimize the foundation design (pad-eye
placement and internal bracing for anchor piles). In these situations, finite element analyses are often used to
model foundation response.

Commercial Considerations

Schedule. Schedule time required to complete a deepwater site investigation varies widely from project-to-
project, and depends on several factors. However, schedule time is lengthened, and sometimes greatly, for
deepwater site investigations compared to typical shallow-water (continental shelf) site investigations. Availability
of specialized vessels and equipment; in some cases, the need for multiple (preliminary and final) field data
acquisition campaigns; the need for special lab testing; the need to integrate and synthesize large amounts of
diverse data; and the need for specialized analyses to characterize and assess complex conditions, commonly
contribute to an extended schedule. Results sometimes cause development plans to be re-evaluated, and in
some cases scenarios need to be iterated several times, causing further delay. In our experience, minimum
schedule time for a deepwater site investigation after development has been sanctioned is on the order of six
months, and typical schedule time is 8 to 12 months. Some site investigations for large deepwater developments
in areas of complex or problematic geology have extended over periods of 2 to 3 years.

Cost. For many of the same reasons deepwater site investigations require much longer to complete,
deepwater site investigations are also much more costly. Costs can be one to two orders of magnitude more
than the cost of a routine shallow-water site investigation (combined geophysical survey and geotechnical
programs).

Summary Case Histories

Case History #1 (an example of how not to do it). A pre-drilling geohazards assessment was carried out for
a proposed exploration-drilling site in more than 5,000 ft of water. The geohazards assessment was done using
3DX data and showed that the site was clear of drilling hazards. The well was a discovery and a spar-type
production facility was proposed for installation. Three groups of three suction-anchor piles each, arrayed around
the spar site at a radius about equal to the water depth, would be used to hold the spar on station. Contrary to
12 OTC 19535

recommendations, the operator initially decided that the 3DX pre-drilling hazards assessment was also adequate
for final facilities siting. Accordingly, the operator proceeded with the geotechnical program and production facility
design, and ordered materials, including the anchor cables. Eventually, the operator became convinced that both
MMS requirements and good engineering practice required a high-resolution site survey. This survey was carried
out and revealed complex faulting at one of the anchor clusters (these faults were too small to be detectable on
the 3DX data used for the drilling hazards work).

The operator decided to relocate the anchor piles in the affected cluster away from the faults because of: 1)
the potential for ongoing fault movement; 2) concern that, although unlikely, the faulting process may have
significantly altered the geotechnical properties of the surrounding soils for an unknown distance from the faults;
and 3) to reduce regulatory and certification concerns. However, by this time, options for moving anchor piles
further away from the structure to avoid the faults were limited because of the maximum length of the available
cable (which had been ordered earlier). Further, options for either moving the anchor piles closer-in or rotating
the entire array were limited because of design/hurricane loading/anchor-pile performance considerations. This
resulted in unexpected meetings to develop a solution. The eventual solution included carrying out an additional,
unplanned, detailed AUV micro 3-D seismic survey to allow precise definition of sites where anchors could be
placed in proximity to the faults and yet avoid intersecting individual fault planes (“thread-the-needle” approach).
Moving the anchors in this cluster also resulted in re-design of the entire mooring spread. The overall
consequence was project delay and additional cost. Although a relatively “easy” solution was found in this case,
both the project delay and the added cost could have been avoided by carrying out a high-resolution site survey
before beginning final design and ordering anchor cables and other materials.

Case History #2 (an example of how to do it “right”). In mid-July, an operator requested that a pre-drilling
geohazards/shallow-water-flow assessment be carried out at a deepwater GOM site using 3DX data
(conventional pre-drilling geohazards data of marginal quality that had been acquired earlier was also available,
and was used to the extent it could be). A discovery was made in October of the same year and the operator
immediately requested that a preliminary engineering-geologic assessment be made to facilitate rapid
development planning. The 3DX data, already loaded on the workstation as a result of the drilling hazards work
done in July/August, was used to quickly confirm and then further refine information about site conditions that
could adversely affect foundation concepts, field layout, and other aspects of development. The principal
geohazards to development were (active) faults and associated fault scarps, some more than 200 ft high, and the
potential for slope failure on the steep scarps. Results of this 3DX-based study were also used to plan the high-
resolution geophysical site/route survey, which was carried out the following April. Results of this design-level
geophysical survey were, in turn, used to plan the geotechnical program for foundation design, which was
subsequently carried out in May. Ultimately, all geophysical and geotechnical data were integrated, and all
geohazards and geotechnical assessments were completed in mid-September. Schedule time elapsed from start
of initial drilling hazards work to finish was about 16 months. A total of 8 geophysical and geotechnical reports,
covering various aspects of the integrated site investigation, were completed during this time.

Other case histories are given in Campbell (1999).

Costly Pitfalls

Based on our experience, some site-investigation pitfalls that can be costly to deepwater development
projects include:

• Being unaware of the potential complexity of deepwater site conditions and geohazards, and the
engineering difficulties they can pose – sometimes referred to as the “shelf mentality”: that is,
assuming a flat, smooth seafloor with stable, uniform soils as is typical of most sites on the continental
shelf. Consequently, there is no incentive to approach deepwater site investigations any differently than
usual.

• Failing to use available 3-D exploration seismic (3DX) data to develop a preliminary
characterization and engineering-geologic assessment of a site – costly early planning errors or
inefficiencies can result from not taking advantage of the considerable preliminary site information that
can be derived from 3DX data – data that is now commonly available very early in the exploration-
appraisal-development sequence and on which exploration and subsequent drilling is based.

• Delaying customized site-investigation work until development plans are well underway – believing
that soil borings are needed only to confirm design assumptions about soils, and thinking that delaying
OTC 19535 13

the site investigation until the last possible minute will result in overall savings, when in fact this approach
can result in large unexpected costs and project delays. Early information on site conditions gives
operators and designers a rational basis for avoiding, designing for, or accepting risks posed by site
conditions, and thus allows realistic development planning, budgeting, and scheduling. The late
availability of site information decreases its value to the project – the later the information becomes
available to the project team, the less value it has for making decisions. And of course the late delivery of
adverse site-investigation results can cause an unexpected consequential project delay.

• Being unrealistic about how much time it takes to plan and execute a deepwater site investigation
- which includes both geophysical and geotechnical data acquisition programs (and sometimes multiple
programs), and integrated analysis and synthesis of the resulting data to turn it into useful information. A
corollary is assuming that the special tools and resources needed for a deepwater site investigation can
be immediately available, and that work can be completed as quickly as for a site on the continental shelf.
AUVs and dedicated, deepwater-capable, geotechnical vessels are in limited supply and availability often
needs to be arranged months in advance.

• Not having an adequate budget available for deepwater site investigation work – which results in
“sticker shock” caused by assuming that deepwater site investigation costs will be about the same as for
site investigations on the continental shelf.

• Using the wrong geophysical survey or geotechnical data acquisition equipment - using (or
attempting to use) conventional geophysical survey or geotechnical equipment and not the specialized
tools designed to acquire quality data at deepwater sites. This commonly results from those who write
site investigation specifications not being knowledgeable about deepwater tools and techniques. An
extreme case of this is when someone decides that 3DX data alone is adequate for design-level site
characterization and geohazards assessment. More commonly, tools designed for shallow (continental
shelf) water depths are specified for deepwater surveys, resulting in poor quality data not fit for purpose.
This would then result in: 1) commissioning a second survey with appropriate tools; 2) assuming higher
risk by not identifying some hazardous feature or condition; or 3) in an overly conservative, more costly,
approach if a poorly imaged “anomaly” can not be assessed with confidence.

• Collecting geophysical survey data over too small an area – also known as the “postage-stamp”
approach: trying to save money by unreasonably minimizing the amount of survey work even in the face
of 1) expensive mobilization/de-mobilization charges; 2) complex geologic conditions that may need to be
avoided or otherwise mitigated; and 3) field-development plans that are still in flux (and thus potentially
leading to the need for a second survey campaign should the target location move). Having survey data
over too small an area also inhibits geoscientists’ ability to understand and assess complex geologic
conditions with confidence.

• Carrying out the geotechnical program without benefit of geophysical survey data – believing that:
1) only soil borings are needed for a site investigation; 2) doing the geotechnical work before results of
the survey work become available; or 3) not considering the geophysical survey results when planning
the geotechnical program. The result can be not enough or too many borings/in situ test sites; borings/in
situ test sites poorly located; borings/in situ tests that do not go deep enough; difficulty interpreting the
meaning or validity of geotechnical results; and, in extreme cases, the need to mobilize a second,
unplanned geotechnical program.

• Not recognizing the need to use specialist expertise to characterize and assess deepwater soil
and geologic conditions – assuming that, for example, documents prepared for pre-drilling hazards
assessment would also be adequate for facility siting and foundation design without further scrutiny,
analysis, data integration, synthesis, or geohazard risk assessment.

Concluding Comments

Although there is still much to be learned by all involved in deepwater site investigations, much has been
learned through varied deepwater experience over the past 30 years or so. Given adequate lead time, the
opportunity to select appropriate tools, plan and execute campaigns in the optimum sequence, use results of one
phase to plan the next phase, and apply the right expertise to data analysis, site investigations will be most
effective and will provide optimum value. Key lessons learned include:
14 OTC 19535

• Begin preliminary site investigation work early by using 3DX data.

• Establish an interdisciplinary geohazards and engineering team to oversee data analysis and integration,
and to disseminate information efficiently.

• Carry out a design-level site investigation using appropriate geophysical survey and geotechnical tools
before beginning final field layout, facilities design, and placing material orders.

• Execute the high-resolution geophysical site survey program first and use these results to plan and better
interpret the results of the geotechnical program.

• Integrate results of the geophysical survey and geotechnical programs to provide robust site
characterization and basis for geohazards risk assessment.

• Allow adequate schedule time and budget for an integrated site investigation.

Acknowledgements

The authors thank our many deepwater clients over the past decades for whom we have worked and together
have developed what we believe to be an optimum approach to deepwater site investigation. We thank our many
Fugro colleagues with whom we have worked over the years and with whom we have developed and shared
ideas regarding deepwater site investigation. We also thank Luisa Botelho for final paper formatting, and Fugro
GeoConsulting, Inc., for providing support and permission to publish this paper.

References and Other Selected Literature

Al-Khafaji, Z.A., A.G Young, W. DeGroff, and G.D. Humphrey, 2003, “Geotechnical Properties of the Sigsbee
Escarpment from Deep Soil Borings”, Offshore Technology Conference paper 15158, 26 pages.

Bryn, Petter, Espen S. Andersen, and Reidar Lien, 2007, “The Ormen Lange geohazard experience: Best practice
for geohazard evaluations of passive continental margins”, Offshore Technology Conference paper 18712,
9 pages.

Campbell, Kerry J., 1999, “Deepwater geohazards: How significant are they?”, The Leading Edge, April 1999,
vol. 18, issue 4, pages 514-519.

Campbell, Kerry J., Richard Burrell, Michael S. Kucera, and Jean Audibert, 2005, “Defining fault exclusion zones
at proposed suction-anchor sites using an AUV micro 3D seismic survey”, Offshore Technology Conference
paper 17669, 9 pages.

Doyle, Earl H., 1998, “The integration of deepwater geohazard evaluations and geotechnical studies”, Offshore
Technology Conference paper 8590, 8 pages.

Evans, Trevor, Nathaniel Usher, and Roger Moore, 2007, “Management of geotechnical and geohazard risks in
the West Nile Delta”, in Proceedings, SUT Offshore Site Investigation and Geotechnics Conference, 11-13
September, London, pages 263-270.

Fenton, Clark, Daniel Jayson, Murdo Gillies, and Anthony Parkin, 2002, “Integrated geohazards evaluation and
risk assessment for subsea facilities”, Offshore Technology Conference paper 14271, 7 pages.

Galavazi, Martin, Roger Moore, M. Lee, Denys Brunsden, and Bryn Austin, 2006, “Quantifying the Impact of
Deepwater Geohazards”, Offshore Technology Conference paper 18083, 5 pages.

George, Robert A., Lindsay Gee, Andrew W. Hill, James A. Thomson, and Philippe Jeanjean, 2002, “High-
resolution AUV surveys of the eastern Sigsbee Escarpment”, Offshore Technology Conference paper 14139,
12 pages.
OTC 19535 15

Hewitt, Antony T., Jason L. Smith, and Richard J. Weiland, 2008, “AUV and ROV data integration to predict
environmentally sensitive biological communities in deepwater”, Offshore Technology Conference paper 19358,
14 pages.

Hill, Andrew W., 2001, “The impact of deepwater geohazards on business delivery”, in proceedings of the
Deepwater Geohazards Workshop sponsored by the Energy Research Clearing House, Drilling Engineering
Association, and the U. S. Minerals Management Service, and held at Del Lago Conference Center, Montgomery,
TX, 3 – 4 April. Proceedings were published by the Energy Research Clearing House, The Woodlands, TX.

Hogan, Phil, Andy Lane, Jim Hooper, Aaron Broughton, and B. Romans, 2008, “Geohazard Challenges of the
Woodside OceanWay LNG Development, Offshore Southern California”, Offshore Technology Conference paper
19563, 10 pages.

Jeanjean, Philippe, Andrew W. Hill, and Stephen Taylor, 2003, “The challenges of siting facilities along the
Sigsbee Escarpment in the southern Green Canyon area of the Gulf of Mexico: Framework for integrated
studies”, Offshore Technology Conference paper 15156, 12 pages.

Jeanjean, Philippe, Andrew W. Hill, and James A. Thomson, 2003, “The case for confidently siting facilities along
the Sigsbee Escarpment in the southern Green Canyon Area of the Gulf of Mexico: Summary and conclusions
from integrated studies”, Offshore Technology Conference paper 15269, 10 pages.

Jeanjean, Philippe, K. Hampson, T. Evans, E. Liedtke, and E.C. Clukey, 2005, “An operator’s perspective on
offshore risk assessment and geotechnical design in geohazard-prone areas”, p. 115 – 143, in Proceedings,
Frontiers in Offshore Geotechnics, ISFOG 2005, Perth, Australia, 19 – 21 September 2005, Susan Gourvenec
and Mark Cassidy, editors. Published by Taylor & Francis Group, 1110 p.

Kolk, Harry J., and Kerry J. Campbell, 1997, “Significant developments in offshore geosciences”, p. 3 – 40, in
Proceedings, Volume 1, Behaviour of Offshore Structures, BOSS ’97, Delft University of Technology, 7 – 10 July
1997, J.H. Vugts, editor. Published by Pergamon, 383 p.

Kolk, Harry J., and Juliette Wegerif, 2005, “Offshore site investigations: new frontiers”, p. 145 – 161, in
Proceedings, Frontiers in Offshore Geotechnics, ISFOG 2005, Perth, Australia, 19 – 21 September 2005, Susan
Gourvenec and Mark Cassidy, editors. Published by Taylor & Francis Group, 1110 p.

Kvalstad, T.J., 2007, “What is the current “Best Practice” in offshore geohazard investigations? A state-of-the-art
review”, Offshore Technology Conference paper 18545, 14 pages.

Lacasse, Suzanne, Farrokh Nadim, Amir Rahim, and Tom R. Guttormsen, 2007, “Statistical description of
characteristic soil properties”, Offshore Technology Conference paper 19117, 8 pages.

Liedtke, E., Ed Clukey, Horkowitz, Kathleen O., Andrew W. Hill, Gary D. Humphrey, and Jeremy Dean, 2002, “An
integrated deepwater site investigation: southern Green Canyon, Gulf of Mexico”, in Proceedings, SUT Offshore
Site Investigation and Geotechnics Conference, 26-28 November, London, pages 267-283.

Moore, Roger, Nathaniel Usher, and Trevor Evans, 2007, “Integrated multidisciplinary assessment and mitigation
of West Nile Delta geohazards”, in Proceedings, SUT Offshore Site Investigation and Geotechnics Conference,
11-13 September, London, pages 33-42.

Morgan, E., B. McAdoo, L.G. Baise, and D.J. DeGroot, 2007, “Quantitative seafloor geomorphology and offshore
geohazards”, Offshore Technology Conference paper 18736, 11 pages.

Niedoroda, Alan W., Christopher W. Reed, Lyle Hatchett, Alan Young, Dan Lanier, Vernon Kasch, Philippe
Jeanjean, Dan Orange, and William Bryant, 2003, “Analysis of past and future debris flows and turbidity currents
generated by slope failures along the Sigsbee Escarpment in the deep Gulf of Mexico”, Offshore Technology
Conference paper 15162, 7 pages.

Nowacki, Fritz, Egil Solhjell, Farrokh Nadim, Eric Liedtke, Knut H. Andersen, and Lars Andresen, 2003,
“Deterministic slope stability analyses of the Sigsbee Escarpment”, Offshore Technology Conference paper
15160, 12 pages.
16 OTC 19535

Orange, Daniel L., Michael M. Angell, John R. Brand, Jim Thomson, Tim Buddin, Mark Williams, William Hart, and
William J. Berger III, 2003, “Geological and shallow salt tectonic setting of the Mad Dog and Atlantis Fields:
Relationship between salt, faults, and seafloor geomorphology”, Offshore Technology Conference paper 15157,
16 pages.

Peuchen, Joek, and C. Rapp, 2007, “Logging, sampling, and testing for offshore geohazards”, Offshore
Technology Conference paper 18664, 6 pages.

Power, Pat T., and C. R. I. Clayton, 2003, “Managing geotechnical risk in deepwater off West Africa”, in
proceedings, Pennwell’s 7th Annual Offshore West Africa Conference and Exhibition, March 11-13, Windhoek,
Namibia.

Power, Pat T., Martin Galavazi, and Gareth Wood, 2005, “Geohazards need not be – redefining project risk”,
Offshore Technology Conference paper 17634, 7 pages.

Quiros, Gerardo W., and Robert L. Little, 2003, “Deepwater soil properties and their impact on the geotechnical
program”, Offshore Technology Conference paper 15262, 11 pages.

Shipp, Craig, Jerry Nott, and Jason Newlin, 2004, “Physical characteristics and impact of mass transport
complexes on deepwater-jetted conductors and suction anchor piles”, Offshore Technology Conference paper
16751, 11 pages.

Slowey, Dr. Niall, Dr. William Bryant, Daniel A. Bean, Alan G. Young, and Dr. Stefan Gartner, 2003,
“Sedimentation in the vicinity of the Sigsbee Escarpment during the Last 25,000 yrs”, Offshore Technology
Conference paper 15159, 15 pages.

Tootill, Nigel P., M.P Vandenbossche, and M.L. Morrison, 2004, “Advances in deepwater pipeline route selection
– A Gulf of Mexico case study”, Offshore Technology Conference paper 16633, 5 pages.

You might also like