You are on page 1of 18

OTC 24211

Fundamental and Specific Marine Geohazard Risk Assessment


Categorisation
A.W. Hill, G.A. Wood, and C.A. Scherschel, BP PLC

Copyright 2013, Offshore Technology Conference

This paper was prepared for presentation at the Offshore Technology Conference held in Houston, Texas, USA, 6–9 May 2013.

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
Until around 1990 the main marine geohazard risk addressed solely from geophysical data was the presence of shallow gas.
Presence had been described using a number of non-standard approaches for its potential presence – distinct from the actual
risk it presented to operations. In the 1992 UKOOA Guidelines for the delivery of marine rig site surveys a standardised
nomenclature was suggested that was largely taken up as a standard by the industry.

Ironically, at exactly this point in time, as the industry had begun to move off the shelf and into deep water a number of other
issues had started to present themselves: shallow water flow, chemosynthetic communities, hydrates etc. for which the
industry has yet to publish a set of standard guide points on the risk of presence.

In contrast with shelf settings where the risks to operations were generally around shallow gas, anchoring and footing
assessment and clearance, the industry is now faced with complex slope and deep-water settings where the sources of risk can
be multiple, over printed, and inter-related. In such settings early effort in the exploration and production cycle is required to
ensure appropriate data are gathered and assessments delivered in proper time for drilling and facilities engineers to either
avoid the presence of hazards identified or properly mitigate their presence through appropriate engineering design.

This paper will outline a suggested methodology both to assess fundamental geohazard complexity and manageability, at the
basin or fairway scale, and thence suggest specific risking criteria for the presence of individual geohazard elements at a
specific location such that engineers can clearly, and consistently, understand potential hazard presence and thence the need
to adopt appropriate risk mitigation methods in engineering design.

Introduction
An unrecognised geohazard translates into risk being carried forward into operations with the possibility of a major accident
through use of an operational plan that fails to avoid, or mitigate, the unidentified risk faced by the activity. The avoidance of
this situation has been the underlying driver of marine geophysical site investigation in delivery of hazard or site surveys, as
they are often referred to, over the last fifty years.

To properly assess and describe hazards across an offshore site requires appropriate front end loading (time) for the project to
understand the relative level of complexity faced across a site, allowing design and delivery of a suitable set of geophysical
data to assess the hazards present in reflection of the operations planned, and finally by the application of clear and consistent
methodologies to identify and describe hazards in a manner that is unambiguous to the end users of the work: engineers.
For the majority of the first thirty years of the marine geophysical site investigation industry (1960-1980s) however,
techniques developed originally on the shelf, encouraged both by industry or regulatory practices, did not always permit this
to be achieved in a manner that was always consistent with the risks faced at an individual site or location.

In the United States the concept of the “Hazards Survey” developed around the Block Clearance survey with a standard line
plan of 300 x 900m to cover a standard 3x3 mile OCS lease block. In Europe the “Site Survey” developed that was designed
to clear a standalone well site with a coverage of approximately 1x1km, for a jack-up rig, to 3x3Km or more for a semi-
submersible anchored rig. This approach was perfectly acceptable for shelf operations, with a fairly standard set of site
2 OTC 24211

investigation concerns that needed to be addressed. The problem comes when license blocks are many hundreds of square
miles in size, and the settings are far more complex than had ever been considered in the early years of exploration and
production in shelf seas.

In deep water settings, on the continental slope and rise, the industry has discovered a far greater range of geohazard
complexity than could possibly be imagined on the shelf and this, in turn, has required the industry to consider adopting
different practices to review prospective drilling or development sites.

In addition to this, the identification and description of individual geohazard risk items is an increasingly non-trivial issue
requiring the analysis of different types of geophysical data, and thence the integration of the results with other geoscience, or
engineering, data sets to allow proper identification of the presence of a geohazard and its level of impact on the operation
that is planned. To ensure that new practitioners entering into the industry are quickly and effectively able to reliably deliver
results consideration is needed of a way to ensure delivery of consistent and safe results.

This paper sets out to describe an approach that has been developed to define the relative complexity of an individual frontier
license area prior to, or upon reward, to allow its setting to be compared on a uniform basis to other settings such that
comparative complexity is clearly communicated to management and thereafter the selection of an appropriate scope, and
scale, of geohazard program is placed in its right context.

Similarly, the paper will describe specimen standardized risking rationales for the presence of individual source risks using
the example of shallow gas, shallow water flow, and chemosynthetic communities. In so doing the value of check lists to
trigger the thoughts of both the experienced interpreter and the novice will be highlighted.

New Frontiers
Through the 1960s to 1980s geological and geophysical marine site investigation around the world was dominated by activity
in Shelf seas - and with it a fairly linear set of problems dominated by the issues of shallow gas, the risk of blowout, or the
stability of bottom founded rigs or platforms.

Those problems lead to the development of a standard series of geophysical and geotechnical engineering tools to assess the
safety, or suitability, of any particular location that was being considered for drilling or development (Ardus et al, 1993).
However since the late 1980s, and the move off the shelf down the continental slope, or even onto the continental rise, a
much broader set of geohazard problems have come into play. Table 1 drawn from the OGP Guidelines for the Conduct of
Offshore Drilling Hazard Site Surveys (OGP, 2011) gives a good image of the diversity of issues that now need to be
considered. With this change has come a parallel change in scope, and thus cost and duration, for delivery of an integrated
site investigation campaign that has grown out of all proportion to what might have been expected on the shelf in the past
(Prior et al 1988, Jeanjean et al 2003, Hill, 2004, Solheim et al, 2005, and Hill et al, 2011).

For Operators focused on the avoidance of major accident risk as they enter new unexplored frontier areas, the question is
how can the level of geohazard complexity relative from one area to another be gauged, simply, on a consistent basis, in a
manner easily understood by the non-expert, and yet that clearly articulates the level of risk, and thus the pace of investment
required to safely evaluate geohazard risk and, therefore, allow risks to be properly mitigated in initial exploratory drilling
and yet be ready, promptly, to support development activities in a success case.

What has become clear is that to simply continue to apply the old “block clearance” or “site survey” approach in these
settings fails to fully address site specific risks of the area in question. In short a “horses for courses” approach needs to be
adopted (Hill, 1996). To effectively use this approach, however, requires early appropriate guidance on the complexity faced.

BP in the last four years has licensed new frontier exploration in a range of new settings (Figure 1). As can be seen, the areas
lie in a large range of differing geological settings. In doing so, the issues described above have been regularly confronted.
Therefore, an approach has been adopted to gauge and effectively communicate complexity immediately upon entry and then
start an appropriate focused set of studies to address complexity.

Outward Complexity
Figures 2 and 3 provide two pairings of outwardly similar geological settings. In each case however the pairs are of two
locations located on opposite sides of the world from each other, yet demonstrate a remarkably similar setting and scale of
feature. Each, however, shows some indication of “recent” geological activity – but how recent and how active? Looking just
at these images, which area, from the standpoint of an Operator, might provide the most significant risk of Geohazard issues
and major accident risk? Therefore which needs to be worked most proactively immediately upon entry to ensure risks are
understood, and thence mitigated, to ensure delivery of safe operations?
OTC 24211 3

Yet the imagery, shown in these figures, of the current day seafloor conditions, is better than an Operator would normally
expect to be available upon, or before, entry to an unexplored license, or play, area. Normally one might expect a sparse grid
of low resolution exploratory 2D seismic data and, if lucky, one might be able to lever off some additional academic research
of the shallow geology in the area that, again, may only be limited to some sparse shallow profiler lines, small areas of
swathe bathymetry seafloor imagery, and allied published papers. Figure 4 shows the database of 2D exploratory data that
was present over a recent new entry license area, a block of approximately 25,000 square kilometers – the equivalent of 1000
OCS lease blocks. There were no wells in the area, no ODP boreholes, no recent academic geological research cruises, or
published papers and, as can be seen a 2D exploratory seismic database with a line spacing of 10 - 20Km. Water depth across
the area varied by in excess of 1000m.

In such a setting where does an Operator start?

To fully address every single possible source of Geohazard risk, or concern, at the outset through over analysis of sparse
geophysical imagery of conditions, and extrapolation thereof, might result in gross under, or over, estimation of the actual
risks faced.

So, given a sparse dataset, yet a multiplicity of potential sources of risk, how can one go about defining comparability of risk
reliably, on a consistent basis, and quickly direct upon entry, or, even better, before entry, to ensure that potential
comparative hazard complexity is understood by the Project Team from the outset.

It is proposed that the key is not to study each individual risk but to be able to effectively evaluate relatively few drivers of
geohazard risk that, in different mixes, when convolved together, drive the majority of sources of major accident risk that are
currently considered to be a concern. The question therefore is how many drivers to consider?

The Geohazard Risk Pentogram


Figure 5 shows a proposed approach for providing an integrated ranking scheme of geohazard complexity upon entry. The
approach is based upon an evaluation of five individual subject axes:

 Shallow Structural Complexity


 Seismicity (Sa,map(1,0))
 Recent Deposition Rate (Over last 1Ma)
 Shallow Hydrocarbon Presence
 Database Quality

that represent a basic set of fundamental Geohazard risk drivers that, considered in separate groupings, fundamentally
underlie, and therefore, drive different sources of Geohazard risk.

The applicability of the five axes and their interrelationship is well illustrated in considering the issue of slope instability.

Slope Instability Drivers


Figure 6 drawn from an area of known high fundamental geohazard risk demonstrates a setting where the generally accepted
drivers for marine slope instability are “in play” driving a potentially high risk of future slope instability.

Gas saturated sediments are identified below the slump toe potentially weakening and, therefore removing the support of the
toe. The anomalously high sedimentation rate across the area of study presents the possibility that in addition to gas
saturation, pore water pressures are elevated in the shallow soils - potentially further reducing soil strength. Structurally the
setting is on the flank of a major anticlinal uplift, defined by extensional faulting that reaches beyond the shallowest
competent horizon. Indications of faults offsetting the seabed (outside this example) suggest continued uplift, tilting and thus
increased gravitational forces on the failed block, with one major extensional fault sitting directly below, and offset by, the
back wall of the failure. Lastly the area of study is in an area of active seismicity with significant, known recent tremors on
record.

The risk drivers in this case could be “deconvolved” as shown in the cartoon in Figure 7:

Shallow Hydrocarbon: Sills and Wheeler (1992) demonstrated that the presence of gas in shallow soils can result in
lowering of the compressibility and undrained shear strength of a fine-grained soil, with a similar effect on undrained
strength of a gassy sand, each increasing the susceptibility of the soil to failure.
4 OTC 24211

Seismicity: ground failures are well documented in the aftermath of seismic events. Seismic induced accelerations induce
both ground and direct structural failure. ISO (2004) have defined seismic risk categories linked directly to acceleration
levels, Sa,map(1,0). Thus an area’s risk can be categorized on expected exposure level to severity of ground motions.

Deposition Rate: elevated pore water pressures, in the same way as gas, can reduce the undrained shear strength of a fine
grained soil. Thomson et al (1999) demonstrated that excess shallow pore pressure could be directly related, through
modeling, to anomalously high recent deposition rates that prevent sediments from dewatering adequately.

Structural Complexity: seabed uplift, regardless of its cause, may result in over steepening of sediments and increased risk
of gravitational failure. In extreme areas of structural complexity, faulting may provide locally increased fluid flux to the
shallow section, or, in the extreme, the emplacement of mud volcanoes that all may increase risk of ground rupture and
ensuing failure.

Lack of Understanding: Database Quality: While all the above elements can be properly defined with adequate seismic,
well, borehole, or imagery data, on entry to an area these may well not be available. A poor, or non-existent, database reduces
understanding and increases risk on entry as significant potential problems may remain unidentified.

Commonality and Interplay of Fundamental Risk Drivers


The four fundamental risk drivers for slope instability outlined above, although not an absolutely comprehensive list, can be
considered to capture the most significant drivers for marine slope stability. However what is interesting is that these drivers,
grouped in alternate ways, can be looked at as driving the majority, if not all, of significant marine geohazard risks faced in
frontier settings.

Furthermore in the case of background Seismicity and Deposition Rate (as a driver for shallow overpressure) publications
already offer a simple manner in which to categorize likely risk.

For Seismicity, as already mentioned, ISO have published a five level categorization system using expected Sa,map(1,0)
accelerations for an area (Table 2). ISO have also published reference maps for the entire world an example of which is
shown in Figure 8, which allow entry point guidance of exposure level.

Guidance however has also been published on the risk of onset of Shallow Water Flow (SWF), as related to the recent
depositional rate which, per se, is also indicative of elevated shallow pore pressures. Alberty et al (1997) originally stated that
depositional rates below 500 feet per million years were unlikely to develop overpressure. Thomson et al (1999) took this a
stage further by modeling the effects in detail and then refining the risk levels and portraying them as they applied to the Gulf
of Mexico (Figure 9). The map well illustrates the high risk areas for shallow water flow as reflected in industry experience
over the past twenty years. While the north east corner of the Mississippi Canyon protraction area is shown as low risk for
SWF the western half and particularly South West corner, around Mars and Ursa basins, are shown as High risk areas.
Similarly Green Canyon and Eastern Garden Banks are High Risk. The map correlates very closely to published industry
experiences (BOEM, website) and the suggested risk boundaries drawn in other works such as Ostermeier et al (2000).

An Integrated Ranking Scheme


It is therefore clear that it is relatively simple, using published work to categorize, or quantify if one so desires, two of the
major underlying geohazard risk driver categories chosen using published work for seismicity (ISO, 2004) or shallow
overpressure (Thomson et al, 1999).

It is thence relatively easy to develop similar schema, albeit qualitative, for Shallow Hydrocarbon Presence or Structural
Complexity. Indeed for structural complexity one may wish to adopt a quantitative schema based on, for example, rate of
uplift or compression.

Categorizing the direct presence of Shallow Hydrocarbons is straightforward given modern day data quality (Hill, 1996) and
fundamental ground physics (Domenico, 1976). How one then decides to draw a classification scheme that reflects evidence
for the total spectrum of shallow gas, oil or gas hydrate is for the individual Operator or Regulator to define. It is suggested
that this can be done robustly, consistently, and simply upon entry.

Shallow structural complexity can also be graded, taking into account both the setting (on the basis of published or internal
evaluation) and direct appearance of the acreage under study (regional seismic oversight) ranging from: horizontal layer cake
to highly contorted structural settings.
OTC 24211 5

Knowledge however underpins all: absence of, or very sparse, available data (seismic, well or borehole) or published works
will increase risk through fundamental absence of either knowledge or data by which to assess the acreage.

It is proposed, therefore, that an area’s fundamental geohazard risk can be judged by the evaluation of five fundamentals:
Shallow Hydrocarbon Presence, Recent Deposition Rate, Structural Complexity, Earthquake Risk and Data Base Quality.
The results can be shown on a pentogram plot where “scores” for each axis can be plotted and the underlying complexity
shown via the gross area that is defined (Figure 5).

How an Operator chooses to score, or combine the ensuing scores, to define an area’s “Fundamental Geohazard Risk” is for
the individual Operator to decide: a straight summation, multiplication, or even a weighted sum/product could all be used.
Whichever approach is chosen, however, it provides the opportunity to simply and consistently gauge fundamental risk and
then present results to the Project Team for guidance on technical programme content. The approach also allows changes in
understanding to be tracked through the exploration cycle.

This result can then be taken further and placed onto a Fundamental Risk vs. Manageability cross plot to allow direct
comparison of one area against another. Figure 10 shows such a cross plot where the relative complexity and manageability
of eleven different areas is shown. While there might be a certain element of linearity to the results: low complexity being
highly manageable and vice versa, it can be seen this is not always the case – Areas 4, 5, 7 and 8 scoring similarly on
complexity but plot differently in terms of manageability.

Manageability itself can be gauged in a number of ways. At its simplest, with no apparent significant issues of concern,
manageability may be straightforward – as suggested by the linearity of some entries on the cross plot. However management
complexity may then develop through simple avoidance via selective choice of surface well location, then, with increasing
engineering effort, through use of: pump and dump mud, special cementing solutions, commitment of extra casing strings,
etc. Relevant offset experiences, if available, may bring guidance on ease, or difficulty, of manageability. It is exactly this
effect that is seen on the movement of Area 4 to the right on Figure 10.

In the development case a different set of manageability considerations would be applied.

From such methods it is possible to quickly identify the relative necessity of an appropriate set of geohazard front end
loading works, be it regional screening on 3D data or for the requirement for early specific, focused data acquisition, to allow
the Operator to be well prepared ahead of a drilling location being brought forward by the Exploration Team.

Location, Location, Location


An approach has been described for effective gauging and description of the fundamental geohazard risk faced in a new
exploration area. Eventually however it is always hoped that an Exploration Team will identify an exploration lead in the
acreage that is believed to be worth drilling.

A proposed drilling location will eventually be brought forward for analysis and, dependent on setting and local regulatory
requirement, a focused dataset that appropriately addresses site specific risks will be made available to the interpreter to
assess specific risks at the location itself.

In the same way that a method has been described to efficiently and consistently evaluate fundamental risk on a semi-regional
basis, it is even more important ahead of drilling to accurately assess specific risks from seabed down through the top-hole
section to the setting depth of the first pressure containment string – where the blow out preventer is set atop the well.
This is a non-trivial matter and, of course, has a direct safety implication.

The SINTEF Marine Blowout Database (SINTEF, 2011) demonstrates that the industry, has delivered huge improvements in
shallow gas prediction since the 1980’s but has yet to eliminate shallow gas incidents (Hill et al, 2013). Similarly, twenty
years of improved understanding of shallow water flows has yet to prevent incidents.

Gawande (2009) describes how in the medical industry despite individuals training longer, specializing more, and using ever-
advancing technologies there are still failures at the point of delivery. With site investigation the industry also has available
vastly improved data and technologies with which to integrate and interpret them, yet still the industry has not made shallow
hazard prediction infallible (Lundquist, 2012).

The underlying issue however is two-fold: firstly in consistent and effective interpretation and assessment of hazard
presence, and secondly in consistent communication of hazard presence to the end user: the Drilling of Facilities Engineer.
6 OTC 24211

Gawande makes a very convincing argument, with examples of striking improvements in the medical sphere, that the way to
drive improvement in delivery is through the use of the checklist.

To a certain degree this approach has been in place for Site Investigation, albeit hidden away, since the 1980s for shallow
gas. Brown (1986) in his first edition of AAPG Memoir 42 on 3D seismic interpretation included what can be considered a
checklist of seventeen “Questions an interpreter should ask in an attempt to validate hydrocarbon indicators.” In 1990
UKOOA thought the questions of such value that they took them and modified them specifically to address shallow gas and
included a similar list in their Guidelines for the Conduct of Marine Rig Site Surveys (UKOOA, 1990). Six editions of the
AAPG Memoir, and twenty-five years, later Brown’s checklist remains in the Memoir – with only a few minor edits and
additions to its content despite the huge step changes in technology and data available (Brown, 2011). Following his list
Brown highlights that “Hydrocarbon validation on seismic data necessarily involves accumulation of evidence.” The more
questions on the checklist that can be ticked off as a positive, therefore, improves confidence. Repeating this approach in a
rigorous manner will therefore improve confidence not only in the interpretation – but also in the end user in the consistency
of the output they are being given. Table 3, therefore, sets out to build on this approach and suggests a checklist to be
followed for the identification of shallow gas. It builds on both the latest Brown checklist and the ammendments made to it
by UKOOA.

One might suggest that this reduces the scope of the interpreter to interpret. However it is quite the opposite. The checklist
ensures that the interpreter knows exactly what evidence needs to be present to build a case for the presence of shallow gas.
Figure 11 shows a shallow anomaly on HR2D data from a recent survey in the North Sea: on top un-interpreted (with an
insert of the seabed event for illustration of phase/polarity of the data) and at the bottom the Table 3 checklist numbered off.
In this case the evidence is conclusive - if not providing a complete clean-sweep of gas indicators.

Through use of the checklist approach the interpreter is bringing sustainable rigour into their analysis of the potential
presence of shallow gas. Even if the interpreter is not working with the checklist directly alongside them, training them in its
existence and content, should stimulate their conscious recognition on seeing an anomaly, that may be shallow gas, to allow
their automatic cognitive recognition to provide the pattern matching skills from memory and, if in doubt, pull out the list
itself to check it over again. Continuing application of the technique will lead to rapidly increased experience of the analyst
and, through that improved pattern matching skills – ending with safer results.

Presence of the checklist for the “Checker” of the analysis provides the opportunity to invoke a forced check of the
assessment: is the evidence there as the interpreter sees it?

Given the suggested checklist approach, and its apparent value, can it be applied elsewhere, to other known potential sources
of Hazard, especially in cases where the interpretation will require more complex integration of far more varied sources of
evidence than for a seismic bright spot on a workstation screen.

Shallow Water Flow


SWF is a far more complex issue to evaluate than shallow gas. SWF’s possible presence does not jump out at the interpreter
from the workstation screen. However, again, while the contributory attributes may not be available in a single place, the
interpretation still “necessarily involves accumulation of evidence” of likely presence - in a similar manner to shallow gas.
The various papers published on the subject, however, are varied and spread wide in their contribution. Table 4, therefore
attempts to provide an equivalent checklist for SWF zone interpretation.

In this case it can be seen that the questions can be split into different categories: Aquifer/Reservoir, Geo-Pressure Presence,
Seal, Offset Histories and Leakage. The fundamental approach remains the same: the list aids the building of a compelling,
and thorough, raft of evidence to deliver a reliable conclusion on the presence of a SWF interval.

Seafloor Communities
The approach can be repeated in any particular category that one may wish.

The presence of sensitive environmental communities on the seafloor is an area of ever increasing focus be it for Cold Water
Corals (Freiwald and Roberts, 2005) or Chemosynthetic Communities (Reilly et al, 1996) being two similar areas of concern
but with differing interpretive needs – the latter for example being tied to the need for seabed seepage flux to provide
nutrients. Table 5, therefore, provides a suggested checklist for the interpretation of the presence of Chemosynthetic
Communities.

In this case the scope of content includes, as shown by the sub-headings, inferences to be drawn from: Seabed Morphology,
Seepage Plumbing, Seabed Sediments, Seepage Activity, Core Analysis, and Visual Evidence. The checklist can, for
example, be compared against the types of evidence presented by Weiland et al (2008).
OTC 24211 7

It can be seen that a similar set of questions might be developed for defining the presence of Cold Water Corals that goes far
beyond, and would provide far more confidence, than purely picking suspected high reflectivity areas on side-scan sonar
data.

Communication to the Engineer


The underlying tenet in each of these three cases above is to provide the interpreter with a firm basis to regularly, and
consistently, build evidence for the presence of a hazard, regardless of what it may be, to make results repeatable, and drive
sustainable rigor into the work and, thereby the safety of the results presented – upon which operational decisions need to be
made.

However providing a list of evidence to an Engineer who is not specialized in the analysis of any of these specific features,
let alone the contributory elements, will only lead to confusion and uncertainty. Therefore the final stage needs to a be a
standardized nomenclature for the risk, that is effectively the confidence level from the interpretation, that a particular hazard
is indeed present.

In doing this the words High, Moderate, Low, and Negligible can be applied to describe the confidence level that a particular
hazard is interpreted to be present. To the end user, however, what do these mean in terms of the likelihood of Hazard
presence.

Table 6 sets out standardized definitions for the meaning of High, Moderate, Low and Negligible that can be applied to any
particular hazard and that may finally allow clear and consistent communication of risk of hazard presence to the end user.
The definitions follow from one to another with, at both extremes, clear confidence in the assessment of hazard presence or
absence, whilst, the two central levels, demonstrate that interpretive doubt may exist. It can be seen that the definitions lead,
in a linear fashion, from the checklist approach and the delivery of consistent communication to the end user.

Conclusions
The offshore industry now operates in the most diverse settings that it has ever faced since first stepping offshore.
Understanding the shallow geohazards that might be faced is a significant contributory aspect to the safety and environmental
integrity of offshore operations.

To do this requires rapid and effective understanding of the setting upon entry into a license to immediately start an
appropriate set of project works to fully understand the diversity geohazard risk sources that might be present and ensure they
are not overlooked, or underestimated, in operational planning. Only in this manner can an Operator protect its operations
against major accident risk, loss of life, or environmental damage as a result of unrecognized geohazard issues.

A robust approach has been described that can be adopted to use Shallow Hydrocarbon Presence, Recent Shallow Deposition
Rate, Seismic Risk, Structural Complexity and Database quality to evaluate Fundamental Geohazard Risk on a consistent
basis across the portfolio of an Operator and to track understanding over the life of Project.

Practices have also been suggested to drive improved rigor in the definition of Location specific risks on a consistent basis
that can then be incorporated into Operational Planning.

The availability and use of checklists are recommended for use in initially supporting the training of interpretive staff, then
building it into their standard thought processes, and, finally, for use in the verification of their ongoing analyses to drive
sustainable rigor into the results delivered.

Use of a standardized risk nomenclature has been described to consistently communicate likelihood of hazard presence to the
end user: the engineer to be properly considered in engineering design and hazard mitigation ahead of the intended operation.

It is suggested that adoption of these processes will reduce uncertainty being delivered into operational planning or, in the
worst case, delivery of incorrect or uncertain prognoses that result in key risks being missed, or misunderstood leading to
increased risk of major accident or environmental damage.

Acknowledgements
The authors would like to thank BP PLC for permission to publish this work. The authors also acknowledge numerous
colleagues who have input to discussion on this subject over the years and also those who have reviewed and commented on
this manuscript.
8 OTC 24211

References
Alberty, M, Hafle, M., Minge, J.C., and Byrd, T., (1997): “Mechanisms of Shallow Water flows and Drilling Practices
for Intervention,” Paper 8301, in the proceedings of the Offshore Technology Conference, 1997, Houston, Texas
Ardus, D.A., Clare, D., Hill, A.(W). Hobbs, R., Jardine, R.J., and Squire, J.M., 1993, “Offshore Site Investigation and
Foundation Behaviour, Advances in Underwater Technology, Ocean Science, and Offshore Engineering,” 28, Kluwer,
Dordrecht.
Brown, A.R., 1986, “Interpretation of three-Dimensional Seismic Data,” AAPG Memoir 42, Edition 1, American
Association of Petroleum Geologists, Tulsa, Oklahoma, USA, 1986
Brown, A.R., 2011, “Interpretation of three-Dimensional Seismic Data,” AAPG Memoir 42, Edition 7, American
Association of Petroleum Geologists, Tulsa, Oklahoma, USA, 2011
Domenico, S.N., 1976, “Effect of Brine-Gas Mixture on Velocity in An Unconsolidated Sand Reservoir,” Geophysics,
vol. 41, issue 5, p. 882
Freiwald, A. and Roberts, J. M. (eds), 2005, “Cold-Water Corals and Ecosystems,” Springer, Berlin.
Gawande, A., 2009, “The Checklist Manifesto, how to get things right,” Picador.
Hill, A.J., Southgate, J.G., Fish, P.R. and Thomas, S., 2010 “Deepwater Angola Part I: Geohazard Mitigation,”
Proceedings 2nd International Symposium Frontiers in Offshore Geotechnics Perth, November 2010.
Hill, A. W., 1996. “The Use of Exploration 3-D Data in Geohazard Assessment: Where Does the Future Lie?”, Paper
7966, in the proceedings of the Offshore Technology Conference, 1996, Houston, Texas
Hill, A.W., 2004, “Geohazards and the Offshore Industry - What a Difference a Decade Makes,” Seabed and Shallow
Section Marine Geoscience Conference, The Geological Society, London, UK, February 2004.
Hill, A.W., Walls, A.H., Romo, L., Turnbull, J.B., and Rogers, E., 2013, “Past, Present and Future Marine Geohazard
Issues: Developing Processes to Address Developing Industry Concerns” Paper 24248, in proceedings of the Offshore
Technology Conference, 2013, Houston, Texas.
ISO, 2004, “Petroleum and natural gas industries - Specific requirements for offshore structures - Part 2: Seismic
design procedures and criteria,” International Standards Organisation, 19901-2:2004
Jeanjean, P., Hill, A., Thomson, J., 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,” Paper
15269, in proceedings of the Offshore Technology Conference, 2003, Houston, Texas.
Lundquist, D., 2012. “Geohazard Interpretation truth or myth?” Oceanology 2012, London
MMS, “Shallow Water Flow Maps for Gulf of Mexico,”
http://www.gomr.mms.gov/homepg/offshore/safety/waterflow/wtrflow.html
OGP, 2010, “Guidelines for the Conduct of Offshore Drilling Hazard Site Surveys,” International Association of Oil and
Gas Producers, Report No. 373-18-1, April 2011.
Ostermeier R.M., Pelletier J.H., Winker C.D., Nicholson J.W., Rambow F.H. and Cowan K.M., 2000, “Dealing with
Shallow-Water Flow in the Deepwater Gulf of Mexico” Paper 11972, in proceedings of the Offshore Technology
Conference, 2000, Houston, Texas.
Prior, D., Doyle, E., Kaluza, M., Woods, M., and Roth, J., 1988, “Technical Advances in High-Resolution Hazard
Surveying, Deep Water Gulf of Mexico,” Paper 5758, Offshore Technology Conference, Houston
Reilly, H.H., MacDonald, I.R., Biegert, E.K., and Brooks, J.M., 1996, “Geological controls on the distribution of
chemosynthetic communities in the Gulf of Mexico”, American Association of Petroleum Geologist Memoir 66, p. 39-62.
Sills, G. and Wheeler, S., 1992, “The significance of gas for offshore operations,” Continental Shelf Research Vol. 12,
No 10, pp1239-1250
SINTEF Offshore Blowout Database, http://www.sintef.no/home/Technology-and-Society/Safety-
Research/Projects/SINTEF-Offshore-Blowout-Database/ Trondheim, Norway, 2011
Solheim, A., Bryn, P., Berg, K.., Sejrup, H.P., and Mienert, J., eds, (2005): “Ormen Lange - An Integrated Study for Safe
Field Development in the Storegga Submarine Area,” pp 326 Elsevier
Thomson, J.A., Dodd, T. and Hill, A.(W)., 1999, “Formation of Shallow Water Flow,” 1999 International Forum on
Shallow Water Flows, Conference Proceedings, Pennwell Publishing.
UKOOA, 1990, “Technical Notes for the Conduct of Mobile Rig Site Surveys” UKOOA Surveying and Positionning
Committee, Seabed Survey Technical Sub-Committee, UKOOA, London, 1990
Weiland, R.J., Adams, G.P., McDonald, R.D., Rooney, T.C., and Wills, L.M., 2008, “Geological and Biological
Relationships in the Puma Appraisal Area: from Salt Diapirism to Chemosynthetic Communities” Paper 19360, in the
proceedings of the Offshore Technology Conference, 2008, Houston, Texas.
OTC 24211 9

Tables

Natural Seabed Subsurface Geological


Features Features
 Seabed topography and relief  Sedimentary sequence
 Seafloor sediments and  Stratigraphy
textures  Shallow gas charged intervals
 Sand: banks, waves, and and gas chimneys
mega-ripples  Shallow water flow zones
 Fault escarpments  Abnormal pressure zones
 Diapiric structures  Buried infilled channels
 Gas vents  Boulder beds
 Unstable slopes  Palaeo-slumps and mass
 Mud: flows, gullies, volcanoes, transport complexes
lumps, lobes  Hydrate zones
 Slumps  Faults
 Collapse features  Erosion and truncation
 Fluid expulsion features surfaces
 Chemosynthetic communities  Salt or mud diapirs and
 Hydrate mounds diatremes
 Rock outcrops and pinnacles
 Reefs
 Hardgrounds
 Seabed channels

Table 1: Natural Geohazard features that are now considered by


OGP to require consideration. It can be seen that the list is
extensive. (OGP, 2010)

Table 2: ISO proposed Seismicity Zonation Scheme to categorize the area being studied
(Excerpted from ISO 19901-2:2004, Table 1 with the permission of ANSI on behalf of ISO.
© ISO 2004 – All rights reserved.)
10 OTC 24211

Geophysical Evidence
1. Is the reflection from the suspected gas pocket anomalous, in amplitude?
2. Is the amplitude anomaly structurally consistent?
3. Is the amplitude of the anomaly equivalent to five times, or more than, the background
(non-bright) value for the same reflector ?
4. If bright, is there one reflection from the top of the reservoir and one from the base?
5. Do the amplitudes of the top and base reflections vary in unison, dimming at the same
point at the limit of the reservoir?
6. Is a flat spot visible?
7. Is the flat spot flat, or dipping consistent with gas velocity sag?
8. If present, is the flat spot unconformable with the structure but consistent with it?
9. Have the seismic data being used been converted to zero phase?
10. Does the flat spot have the correct zero-phase character?
11. Do the bright spot, dim spot, or phase change show the appropriate zero-phase character?
12. Is the flat spot located at the down-dip limit of brightness (or dimness)?
13. Is a phase change (polarity reversal) visible at the edge of the anomaly?
14. Is the phase change structurally consistent and at the same level as the flat spot?
15. Is there a pull down effect of underlying reflectors below the suspected reservoir indicative
of gas velocity sag?
16. Is there an anomaly in stacking velocity derived interval velocity across the interval?
17. Is there a low frequency shadow below the suspected reservoir?
18. Are there direct indications of vertical gas migration pathways towards the suspected
reservoir?
19. Did a study of amplitude versus offset on the unstacked data support the presence of gas?
20. Does a near offset range stack show a lower amplitude response than a far offset range
stack for the same event?
21. Are any comparative P and S wave sections available to aid in clarification of gas
presence?
22. Do statistical cross-plotting techniques support gas presence?
Offset Well Evidence
23. Do seismic data allow the anomaly be tied to an offset well where gas was present in the
same interval?
24. If so at the well location are the same geophysical properties present?
Table 3: Proposed checklist for interpreting the presence of Shallow Gas, developed from
Brown, 1986, UKOOA, 1990 and the experiences of the authors.
OTC 24211 11

Aquifer/Reservoir
1. Does the interval contain a known aquifer e.g. The GoM Blue Interval?
2. Has seismic sequence analysis identified sedimentary deposits likely to contain a SWF
interval, e.g. a sand prone layer?
3. Is there a localized amplitude event consisting of an anomalously bright reflection that
may suggest sand presence? If so, can tuning effects be ruled out as the cause?
4. Is there a sand-prone layer contained within a structural trap?
5. Is there a stratigraphic trap formed by dipping sand-prone layer(s) truncated by faulting,
erosional downcutting or depositional pinch out?
6. Is there an isolated sand body capable of absorbing excess pressures caused by
compaction disequilibrium?
7. Are there high-amplitude, discontinuous reflectors within an expanded stratigraphic
sequence?
Geo-Pressure Presence
8. Is there evidence of high sedimentation rates (>1500 ft/my) and rapid burial leading to
pressure disequilibrium?
9. Is there evidence of differential compaction/loading resulting in excess pressures being
transferred from thicker overburden areas?
10. Is there geophysical evidence supporting the presence of a geopressured zone, i.e.
stratigraphic layer(s) containing pore pressure greater than hydrostatic pressure?
Seal
11. Is there a competent regional or sub-regional seal above the potential flow zone?
12. Is the zone buried deeply enough (>500 ft) for development of a sufficiently strong seal?
Offset Histories
13. Can a known shallow water flow zone from a nearby well be correlated to the interval?
If so, is there consistency of seismic character?
14. Has a nearby well proven that SWF can be ruled out? If so, is there consistency of
seismic character? [A negative indicator which significantly reduces SWF risk.]
Leakage
15. Are there buried expulsion features or chimneys recognized on subsurface data that
plumb back to the suspected SWF interval suggesting pressure is at or slightly above
fracture?
16. Are mud volcanoes, pockmarks or other expulsion features present at the seafloor?
17. Does the seafloor backscatter/amplitude map indicate areas of anomalous amplitudes
potentially indicating authigenic carbonate hardgrounds associated with seafloor flow?
Table 4: Proposed checklist for interpreting the presence of Shallow Water Flow Sands,
developed from various sources (Thomson et al (1999), Alberty et al, (1997), and
Ostermeier et al (2000)) and the experiences of the authors.
12 OTC 24211

Seabed Morphology
1. Do swath bathymetry or 3D seismic data show localized seabed mounding?
2. Is the mounding aligned to or at the confluence of a series of seabed faults?
3. Does a study of the swath bathymetry data indicate increased backscatter that is consistent
with the extent of the
seabed mounding?
4. Does a study of the variance in the amplitude of the 3D, and or 2D, seabed reflector
indicate an increase in amplitude that is
consistent with the extent of the seabed mounding?
5. If near bottom acquired side scan sonar exist across the feature of interest does an area of
seabed disturbance exist on sonar data that mimics the feature seen on swath bathymetry
data, 3D seismic and profiler data?
Seepage Plumbing
6. Do 2D or 3D HR multi-channel data or 3D exploration seismic data indicate the presence
of a deep routed fault below the suspected community?
7. Do 2D or 3D HR multi-channel data or 3D exploration seismic data indicate the presence
of a chimney below the suspected community?
8. Do 2D or 3D HR multi-channel data or 3D exploration seismic data indicate the presence
of [stacked] brights in the shallow section suggesting the presence of hydrocarbons and a
vertical migration pathway to seabed?
Seabed Sediments
9. If near bottom deployed profiler data exist across the suspected community does the data
corroborate the existence of a change in seabed reflectivity across the community?
Seepage Activity
10. Does the profiler data provide any indication of the presence of plumes in the water
column above the community?
11. Do the sonar data indicate the presence of plumes in the water column emanating from the
community?
12. Does SAR data coverage of the area suggest the existence of naturally occurring slicks
over the area?
13. Do SAR imaged slick locations integrated with sea current information suggest seabed
emission points that overlie the suspected seepage areas defined from other data types?
Core Analysis
14. Are seabed core data available?
15. Did a core, taken within the bounds of the suspected community, smell of gas and contain
shell or carbonate debris?
16. Does analysis of the core reveal anomalous quantities of worm or crustacean activity
relative to any offset “control” core?
Visual Evidence
17. Does direct visual, video, or photographic evidence provide categorical proof of the
presence of a chemosynthetic community?
18. Is the suspected community within, or close to, the location of a recorded (published)
chemosynthetic community?
Table 5: Proposed checklist for interpreting the presence of Chemosynthetic Communities
using various data types, developed from various sources.
OTC 24211 13

HIGH: Possesses all, or the majority, of the characteristics expected to be seen


from the Hazard under study if present, or that can be verified directly by
offset well or direct visual evidence.

MODERATE: Possesses a large number of the characteristics expected to be seen from


the Hazard under study if present, but an element of interpretive doubt
remains.

LOW: Shows only a small number of the characteristics expected from the Hazard
under study if present, on balance the hazard is believed to be absent
although some interpretive doubt exists for the total absence of the Hazard.

NEGLIGIBLE: Either shows none of the characteristics expected to be seen from the
Hazard under study, can be clearly interpreted as not being the hazard
under study, or can be proved directly, by offset well or direct visual
evidence, to be absent.

Table 6: Proposed Standardised risk levels to be applied to the interpreted presence of any
particular type of Geohazard. Note the “risk” level stated here appertains to the level of
confidance that the hazard is present – not to the level of risk to the operations being
planned after mitigation through direct avoidance or engineering design.

Figures

Figure 1: New Marine Frontier Exploration entry areas made


by BP over the last four years. Note the diversity of
geological settings under consideration.
14 OTC 24211

Figure 2: NW European Continental Slope (left) and North American Arctic Continental Slope (right), in
both cases the shelf break is at the bottom of the image, the upper slope is affected by iceberg scouring,
above a mid slope section characterized by evidence of multiple debris flows shown on 3D seismic seabed
pick. Scales are approximately the same.

Figure 3: US GoM Continental Slope (left) and Eastern Mediterranean Continental Slope
(right), in both cases the shelf break is at the top of the image, the upper slope is affected
by various phases of incised, partially infilled, canyons. Oustide the channel axes are
indications of various modes of slope failure Scales are approximately the same.
OTC 24211 15

Figure 4: Seismic Inventory covering a Frontier Licence Pre-Entry. Note the size of the
block is approximately 25000sq Km – equivalent to in excess of 1000 GoM lease blocks.
In this case a relatively uniform 2D exploratory data coverage exists with 10-20Km
spacings. There are no wells in the area. Water depth varied ~1000m across the area.

Geohazard Risk Pentogram

Database
5

2
Sh Hydrocarbon Structural
1

Deposition Rate Seismicity

Figure 5: Proposed Geohazard Risk Pentogram, the five axes, as described in the text,
are scored from the centre outwards, the larger the background fundamental risk the
larger area enclosed.
16 OTC 24211

Figure 6: South Caspian Basin, HR 2D data: a remarkably well defined “simple” rotational failure on a
large scale. The failure is 1km in length from headwall (right) to toe (left) and approximately 200m in
thickness (timing lines shown are 100ms TWT). Brightening of the shallow reflectors below the toe
suggests the presence of gas, recent deposition rate is anomalously high, a significant fault that
offsets the first competent (non soil) horizon lies directly beneath, and is offset by, the slide plane,
there is continued structural uplift which is tilting the seafloor seawards, while the license setting is
known to be affected by earthquakes.

Figure 7: Fundamental Geohazard Risk Drivers as applied to


the issue of Slope Instability. Compare these to the situation in
Figure 6.

Figure 8: ISO published Sa,map(1,0) data for North America (ISO, 2004) reference back to the
Table 2, above to understand the perceived severity or zonation of a potential operational area.
Note the differences between Southern Alaska seaboard and the GoM (Excerpted from ISO
19901-2:2004, Figure B.2.a with the permission of ANSI on behalf of ISO. © ISO 2004 – All rights
reserved.)
OTC 24211 17

Figure 9: Thomson et al (1999) Shallow Water Flow Risk Zonation map of the Gulf of
Mexico based solely upon the deposition rate over the last ~1 million years. The
darker the red the higher the risk. The red arrow on the scale is a modification of the
Alberty Criterion (Alberty et al, 1997) after verification by Basin Modeling.

Figure 10: Entry Point Fundamental Geohazard Risk vs Perceived manageability


cross plot. Fundamnetal Risk score is drawn from the Pentogram. Manageability is
assessed dependant on the ease of avoidance or engineering effort that may be
required to mitigate a hazard’s presence. The different circles represent different
areas assessed at point of licence entry. Areas falling into “red” area will require
careful front end assessment.
18 OTC 24211

Figure 11: Central North Sea Suspected Shallow Gas anomaly: Uninterpreted (top) and Interpreted
(bottom). Seafloor reflector in top Left corner for phase and polarity reference. The numbers on the
interpreted section refer directly to the Shallow Gas Checklist in Table 3. The anomaly would be given
a “High” risk of gas presence on the basis of Table 6.

You might also like