Sei sulla pagina 1di 10

OTC 19601

Borehole Pressure Coring and Laboratory Pressure Core Analysis for Gas
Hydrate Investigations
P.J. Schultheiss and M.E. Holland, Geotek Ltd.; and G.D. Humphrey, 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

Marine gas hydrate investigations, whether they be conducted as resource evaluations for national governments, as geohazard
assessments for major oil companies, or as climate-related scientific endeavors funded through international funding
agencies, have begun to develop common strategies. Gas hydrate can occur up to several hundred meters below the
sediment-water interface, depending on the water depth and thermal gradients; consequently, for a full assessment of gas
hydrate occurrence, nature, distribution, and concentration, samples must be obtained through drilling and coring. Samples
retrieved and analyzed at full in-situ pressures using pressure coring techniques are the only way to ground truth the findings
from other techniques, including physical and chemical analysis of non-pressure cores or the interpretation of seismic and
borehole log data.

Thermal imaging of non-pressure cores provides an immediate and valuable method for observing the endothermic behavior
of dissociating gas hydrate but does little in itself to characterize the gas hydrate morphology or concentration. Porewater
analysis techniques rely on the freshening of porewaters when gas hydrate dissociates, generally measured via chlorinity;
these analyses provide excellent spot measurements of the concentration of gas hydrate if the background chlorinity profile is
well established, but tells the investigator little about the nature of the gas hydrate. Furthermore, in lakes or nearshore
environments where the porewaters may have very low or highly variable chlorinities, this technique will be far less reliable
or accurate. Only methane mass balance calculations from depressurization of cores recovered at full pressure provides the
benchmark or “gold standard” for gas hydrate concentration assessment.

Measurements of the density and volume of pressure cores through X-ray imaging and gamma densitometry have removed
errors in estimation of pore volume, making this methane mass balance technique accurate and robust. Non-destructive
testing and analysis of gas-hydrate-bearing cores at in-situ pressures and temperatures also provides detailed information on
the in-situ nature and morphology of gas hydrate in sediments. The detailed profiles of density and VP, together with spot
measurements of Vs, electrical resistivity, and hardness, provide background data essential for modeling the behavior of the
formation on a larger scale. X-ray images show the hydrate morphology in relation to the sediment, which will control the
kinetics of methane release. Gas-hydrate-bearing pressure cores subjected to X-ray tomographic reconstruction provide
compelling evidence that gas hydrate morphology in many natural sedimentary environments is particularly complex and
impossible to replicate in the laboratory.

Review of strategies and techniques for gas hydrate expeditions

Gas hydrate is studied in oceans around the world for a variety of reasons, including resource assessment, geohazard
evaluation, and general scientific interest including climate change. Recent expeditions that have focused on gas hydrate
studies include Ocean Drilling Program (ODP) Leg 204 to Hydrate Ridge, Oregon Margin (Tréhu et al., 2003); the
Chevron/DOE (US Department of Energy) Naturally-Occurring Hydrates JIP (Joint Industry Project), Gulf of Mexico
(Claypool et al., 2005); Integrated Ocean Drilling Program (IODP) Expedition 311, Cascadia Margin (Riedel et al., 2006);
the Indian National Gas Hydrate Program Expedition 1 (Collett et al., 2006), Bay of Bengal; the Chinese Guangzhou Marine
Geological Survey Expedition 1 (Zhang et al., 2007), South China Sea; and the Korean Ulleung Basin Gas Hydrate
2 OTC 19601

Expedition 1, East (Japan) Sea. All these expeditions had, at their essence, the same goals: characterization of the
sedimentary environment, definition of the probable gas hydrate zone, determination of gas hydrate distribution,
quantification of gas hydrate concentration and composition, and measurement of gas hydrate morphology within the
sediment and the in-situ properties of the matrix. Together, these deliverables describe the gas hydrate in relation to the
surrounding sediment.

The common set of techniques used to achieve these objectives includes pre-expeditionary seismic measurements, sometimes
three-dimensional; bottom surveys with conventional cameras and sidescan sonar; and piston coring. Expeditions themselves
are often in two phases: a “logging” phase, consisting of borehole measurements, and a “coring” phase, consisting of wireline
sampling and in-situ testing. The logging phase consists of wireline or Logging-While-Drilling (LWD) technologies to
investigate downhole density, electrical resistivity, sonic velocity, lithology via natural gamma radiation, and other borehole
parameters. The coring phase consists of sampling with wireline coring tools that bring back non-pressurized cores, which
are then examined for thermal anomalies via infrared camera and subjected to porewater analysis before being
sedimentologically described. It also includes sampling with wireline pressure coring tools that bring back cores under full
in-situ pressure (preserving any gas hydrate). These cores are immediately imaged using X-rays, profiled to determine their
detailed velocity and density structures, and may be depressurized to collect quantitative gas samples. The coring phase
always includes in-situ temperature measurements (to help determine the thermodynamic zone of gas hydrate stability) and
may be augmented with in-situ pore water sampling, strength, and pressure/permeability measurements.

The data gained from pressure cores taken during the gas hydrate expedition are required to meet many of the primary goals
of the expedition. Because gas-hydrate-bearing sediments sampled by conventional (non-pressure-retaining) coring
techniques are disturbed by gas hydrate dissociation and gas exsolution, the use of a pressure-retaining coring system is the
only way to ensure the recovery of gas-hydrate-bearing sediment with minimal disturbance. Gas-hydrate-bearing sediment
sampled by pressure core can be used immediately to determine the distribution and morphology of gas hydrate relative to
the surrounding sediment and to measure the properties of the gas-hydrate/sediment matrix. If the gases trapped in the core
are released and collected, the pressure core provides the best quantitative assessment of gas hydrate concentration possible
today; critically, it is also the only technique that can confirm the absence of gas hydrate. If the core remains under pressure,
samples of gas-hydrate-bearing sediment can also be transferred under pressure to shore-based institutions for other detailed
laboratory investigations. This paper describes the analyses that are performed on pressure cores in current practice and how
they fit into a gas hydrate expedition data set.

Development of wireline pressure coring system specifically for core analysis

Pressure cores have been used in scientific ocean drilling gas hydrate research since Kvenvolden et al. (1983) attempted the
first depressurization and gas collection experiment to quantify gas hydrate within cores from the Deep Sea Drilling Project’s
Pressure Coring Barrel (Deep Sea Drilling Project, 1984). The Ocean Drilling Program (ODP) later developed the Pressure
Core Sampler (PCS; Pettigrew, 1992; Graber et al., 2002) and the Pressure-Temperature Coring System (PTCS) was
developed for Japan Oil, Gas and Metals National Corporation (JOGMEC, formerly Japanese National Oil Company, JNOC;
Takahashi and Tsuji, 2005). However, the analysis of pressure cores did not significantly progress beyond their initial use of
core collection and depressurization until the HYACE/HYACINTH program developed the HYACINTH suite of coring and
core analysis tools.

The HYACINTH pressure coring and analysis systems (Schultheiss et al., 2006) were initially developed by a consortium of
industry and academia funded by the European Union; development and operation of the HYACINTH system has now been
handed over to a partnership between Geotek Ltd. and Fugro. This integrated system includes two coring tools and an array
of downstream core processing equipment and capabilities, currently designed to operate at up to 250 bar (equivalent to
approximately 2,500 meters water depth). The combined suite of equipment enables the cores to be acquired and transferred
in their core liners from the pressure coring autoclaves into chambers for non-destructive testing, sub-sampling and storage as
might be required for different investigations. The two coring tools, the Fugro Pressure Corer (FPC) and the Fugro Rotary
Pressure Corer (FRPC), were designed to recover high-quality cores in a complete range of sedimentary formations. Testing
and use of the FPC and FRPC have been made in close cooperation with ODP and its successor program, the Integrated
Ocean Drilling Program (IODP).

The two HYACINTH coring tools have been used on three dedicated scientific gas hydrate expeditions on the drilling vessel
JOIDES Resolution alongside the PCS: ODP Leg 204 in 2002 (Hydrate Ridge, offshore Oregon; Tréhu et al., 2003); IODP
Expedition 311 in 2005 (Cascadia Margin, offshore Vancouver Island, Canada; Riedel et al., 2005); and the Indian National
Gas Hydrate Project (NGHP) Expedition 1. The HYACINTH tools have also been deployed off geotechnical drilling vessels
and platforms on five other dedicated commercial gas hydrate expeditions for national governments and industry in 2003,
2005, 2006, and 2007. The overall performance of the coring tools has steadily improved during their development with
success rates for both tools in the last two years typically 70–80%. Most importantly, the performance of the tools coupled
OTC 19601 3

with fast operations ensures that gas hydrate within the core remains permanently inside the thermodynamic stability field as
demonstrated repeatedly by the internally recorded pressure-temperature histories (Fig. 1).

The HYACINTH coring tools differ in a number of significant respects from the early PCS and PTCS designs. The
HYACINTH tools penetrate the sediment using downhole driving mechanisms powered by fluid circulation rather than by
top-driven rotation with the drillstring, which can significantly improve core quality. Both HYACINTH tools use “flapper
valve” sealing mechanisms in the autoclave rather than a ball valve, which maximizes the diameter of the recovered core
relative to the drill pipe. The HYACINTH pressure core autoclaves are attached to gas-charged pressure accumulators,
which partially compensate for drops in core pressure due to tool volume expansion during core retrieval (Fig. 1). The
recovered HYACINTH cores are contained in an inner plastic liner and can be manipulated and transferred into other
chambers for analysis, storage and transportation under full pressure.

The Fugro Pressure Corer (FPC) is a HYACINTH-compatible percussion corer developed by Fugro Engineers BV. A water
hammer, driven by the circulating fluid pumped down the drill pipe, is used to drive the FPC core barrel into the sediment up
to one meter ahead of the drill bit. In soft sediments it simply acts as a push corer and no hammering is required. The core
liner (63 mm outside, 57 mm inside diameter) is retracted into the autoclave chamber past a flapper valve which seals the
bottom end of the core. In its current configuration, the FPC is suitable for use with unlithified sediments ranging from soft
through stiff and sandy clays.

The Fugro Rotary Pressure Corer (FRPC, previously known as the HYACE Rotary Corer, HRC) is a HYACINTH-
compatible rotary corer developed by the Technical University of Berlin and the Technical University of Clausthal. An
Inverse Moineau Motor, driven by the circulating fluid pumped down the drill pipe, is used to rotate the cutting shoe
independently of the drill pipe up to one meter ahead of the roller cone bit. A narrow kerf, dry auger design cutting shoe with
polycrystalline diamond cutting elements, designed to core into lithified sediments, was the original bit used with the FRPC.
This design allows the core to enter into the inner barrel before any flushing fluid can contaminate the material being cored.
The lined core is 51 mm in diameter (liner outer diameter is 56 mm). On completion of coring, the tool is lifted off the
bottom using the drillstring and then the core is retracted into the autoclave by pulling in on the wireline in a similar manner
to the FPC, with the pressure again being sealed by a flapper valve. Most gas hydrate expeditions do not encounter the very
hard lithified material for which the FRPC bit was originally designed, so on the three most recent expeditions, a new “auger
style” helical bit (the “Viking”) was used successfully to enhance penetration in much softer materials and has proved very
successful, making the FRPC a very versatile tool for a wide range of lithologies simply by changing the cutting shoe.

The breakthrough for gas hydrate core analysis using the HYACINTH system was the ability to transfer cores out of the
autoclave and into other pieces of pressurized equipment. The Pressurized Core Analysis and Transfer System (PCATS),
illustrated in Fig. 2, contains an automated core manipulation device which latches onto HYACINTH cores and can move
them (repeatedly if necessary) in and out of specialized chambers for analysis, sub-sampling and storage as required. The
automation in the PCATS has enabled pressure core manipulation and measurements under pressure to become routine.
When a core is first recovered in the corer autoclave, the autoclave is removed from the coring tool and connected to the
water-filled PCATS where the pressures are carefully balanced before the connecting ball valve is opened. The long
manipulator is moved from the PCATS into the corer autoclave to “catch” the core and the core is retracted into the PCATS.
All the core movement operations are performed under computer control to ensure accurate positioning and torque. The
autoclave is then removed and connected to the analytical equipment where the computer-controlled manipulator in the
PCATS allows the core to be translated and rotated accurately (±1 mm and ±0.5 degrees of accuracy, respectively) inside the
analysis chambers. Once data is collected, the core may be depressurized while obtaining more measurements (see “Analysis
of pressure cores”) or can be stored under pressure for future experiments.

Core stored under pressure may be subjected to other analyses through the storage vessel walls or transferred into further
pressurized analytical chambers. The specifications for the HYACINTH mating flange and clamps used in the HYACINTH
PCATS (Fig. 3) are publicly available and any party may design pressurized equipment that connects to the PCATS.
Independent research scientists have already developed pressurized equipment that has been used with the PCATS and stored
HYACINTH cores, including the Instrumented Pressure Testing Chamber (IPTC; Yun et. al., 2006; Fig. 3) and the
DeepIsoBug (Schultheiss et al., 2006). To improve the flexibility of pressure core utilization, the PCATS is currently
undergoing modifications to allow each core to be cut, under pressure, into sections of varied lengths. In this way parts of a
core could be depressurized on board ship, while other parts of the same core could be stored under pressure for shore-based
studies. This increased versatility will enable each core to be more fully assessed and further increase the value of every
pressure core recovered.
4 OTC 19601

Analysis of pressure cores

While cores remain under near in-situ pressures and temperatures, measurements of morphology and some in-situ properties
can be made. Many of these measurements are non-destructive and non-invasive, allowing multiple data sets to be acquired
per core. All measurements made under pressure involve specialized apparatus. When the measurements that can be made
under pressure are completed, further information can be gained by careful observation of the cores and their gaseous
products while the pressure is released slowly. Alternatively, cores can depressurized rapidly for storage in liquid nitrogen
and further analysis at atmospheric pressure, providing some of the least-disturbed samples of gas hydrate at atmospheric
pressure.

Non-destructive measurements on pressure cores are important as an immediate survey of the core, both to determine if a
successful core has been retrieved and to look for obvious signs of the presence of gas hydrate. They also provide primary
data on sediment-hydrate properties to ground-truth larger-scale measurements. The analytical portion of the PCATS, the
MSCL-P (Pressurized Multi-Sensor Core Logger), was designed as part of the HYACINTH suite to collect routine non-
destructive data on HYACINTH-compatible pressure cores, including gamma density, P-wave velocity, and X-ray images.
Gamma density is measured using a 137Cs source and NaI detector. Ultrasonic P-wave velocity (VP) is measured using two
500 kHz acoustic transducers mounted inside the pressure chamber, perpendicular to the core and the gamma ray beam. X-
ray images are obtained through an aluminum pressure chamber using a linear X-ray device consisting of a lead-shielded
microfocal X-ray source and phosphor image intensifier; the intrinsic spatial resolution of the images is approximately
150 µm.

Fig. 4 shows an example of non-destructive data collected on HYACINTH pressure cores, which highlights the surprising
observations made on pressure cores regarding the nature of gas hydrate in clay sediments. This core is composed of fine-
grained sediment. The X-ray image of this core revealed gas hydrate nodules, horizontal lenses, and sub-vertical veins (Fig.
4a). Though the core was rotated (Fig. 4b) to discern the differences in the velocity and density anisotropy (Fig. 4c), the true
complexity of the grain-displacing hydrate was not evident until the core underwent X-ray CT (computed tomography)
analysis (Fig. 4d). This was the first observation of complex gas hydrate veins in clay sediments and has provided the
impetus for improvement of the MSCL-P to include routine X-ray CT analysis of pressure cores. A detailed assessment of
the nature and causes of the complex vein structures is beyond the scope of this paper. However, these morphological details
are extremely important, not only for explaining the mechanisms of gas hydrate growth in fine-grained sediments but also for
predicting the sediment behavior during gas hydrate dissociation. Models predicting the behavior of such gas-hydrate-
bearing sediments during dissociation, whether for well-bore stability, geohazard assessment, or potential methane gas
production, must acknowledge the complex nature of these morphologies.

The slow, isothermal release of pressure from a pressure core allows gases to exsolve from pore fluids and gas hydrate to
dissociate. Measuring the quantity of gas, its composition, and its evolution relative to time and pressure provides
information on the quantity, composition, and surface area of gas hydrate (Kvenvolden et al., 1983; Dickens et al., 2000;
Milkov et al., 2004; Fig. 5). The fundamental number obtained through these experiments is the nominal concentration of
methane in the pore fluids, assuming all methane is in solution. If this nominal concentration is greater than the calculated
methane saturation, gas hydrate (or free gas, depending on the thermodynamic conditions) is assumed to be present and the
amount can be quantitatively calculated. Data that shows the sediment is under-saturated in methane is also extremely
important, as this is the only technique that can confirm the absence of gas hydrate. Plotting methane concentration, both
above and below saturation, relative to the changing thermodynamic boundaries downhole allows visualization of the
methane gradient and is a requirement for definition of the probable zone of gas hydrate occurrence (Fig. 6).

These depressurization experiments can be augmented by testing of physical properties during depressurization.
Measurements of the changing physical properties of the sediment-hydrate matrix during gas hydrate dissociation will be
critical for any study of gas hydrate production as a resource or gas hydrate geohazard assessment. The addition of
information regarding the physical properties of the core also enhances the methane mass balance information gained during
a depressurization experiment. Repeated density and X-ray scans during depressurization have allowed observation of gas
evolution from and assignment of gas hydrate to specific core layers, as well as tracking movement of sediment during
depressurization to aid in quantitative mass balance analyses.

Pressure core data in the context of a gas hydrate expedition

The gas-hydrate-related data collected on pressure cores is important for meeting the goals of defining the probable gas
hydrate zone, quantifying hydrate concentration, and measuring hydrate morphology and in-situ properties of the sediment-
hydrate matrix. It is also key ground-truth data for other, more indirect measures of gas hydrate that are used to quantify gas
hydrate and determine the gas hydrate distribution. Pressure cores typically comprise less than 10% of a cored section in a
OTC 19601 5

borehole; 30–90% of the remaining sediment is generally recovered by conventional coring. The properties of any sediment
not recovered in cores must be inferred from downhole measurements. By applying the pressure core data to the data sets
collected by conventional coring or downhole measurements, a more quantitative gas hydrate distribution can be obtained.

Quantitative point measurements of gas hydrate in conventional cores are made using porewater freshening analysis, a
measurement of the fresh water left behind by gas hydrate dissociation. Estimates of gas hydrate from porewater freshening
analysis can benefit from comparison with quantitative gas hydrate measurements made on pressure cores. Pressure cores
can help establish the baseline porewater chemistry that must be assumed when calculating the amount of freshening;
fluctuations in this baseline from fresh water input can mimic gas hydrate presence. Gas hydrate morphology from pressure
cores can also be useful for developing a strategy for sampling porewater from conventional cores.

The distribution of gas hydrate in conventional cores is estimated from infrared thermal images, which are routinely acquired
along the entire length of the recovered core. The energy removed from the cooled regions of the core can be used in a semi-
quantitative manner, as a minimum estimate of gas hydrate dissociated in the core. Again, if these semi-quantitative
estimates are compared to the quantitative estimates from pressure cores, the data may be adjusted to better represent the total
distribution of gas hydrate in the cored interval.

Downhole electrical resistivity measurements have long been used to estimate the distribution of gas hydrate in sediments;
however, though electrical resistivity measurements combined with Archie’s relation for pore volume saturation have been
used quantitatively in gas-hydrate-bearing sands, this approach does not work for marine hydrate-bearing clays. Quantitative
gas hydrate estimates from electrical resistivity values must be ground-truthed with pressure cores or porewater freshening
analysis, as the electrical resistivity response will vary strongly with gas hydrate morphology. However, anecdotal evidence
suggests that if ground-truthed, electrical resistivity measurements for a local area can be used in a semi-quantitative or even
quantitative fashion.

All the data collected on a gas hydrate expedition combines in a synergistic fashion to produce a picture of the gas-hydrate-
bearing formation that could never be formed from one single technique, but pressure cores provide key pieces of
information to ground-truth more continuous data sets. Pressure coring and pressure core analysis have become standard
components of gas-hydrate-related research and expeditions for academia and industry. The technologies have matured and
the success rates improved to levels where the operations have become routine, allowing pressure core analysis to produce
the scientific surprises that every new technique brings to its field.

Acknowledgements

The authors would like to thank the people and organizations who have been involved in the development and use of pressure
coring and pressure core analysis systems. We thank all the partners within the EU-funded HYACE and HYACINTH
programs, in particular, our colleagues at Fugro BV and the Technical University of Clausthal; the engineers who have had
the most recent input to the design and operation of the coring tools at sea are Floris Tuynder, Roeland Bass and Martin
Rothfuss. All the scientists and crew on the operations that have taken place on the all the drilling vessels especially the
JOIDES Resolution and the Bavenit deserve our gratitude during the development of the HYACINTH systems. We are
particularly indebted to Hans Amann, Tim Francis, Frank Rack, Tim Collett, Michael Riedel, and Mike Storms for their long-
term visions of the importance of pressure coring. We are grateful to the US Department of Energy for practical and
financial support on a number of projects. We thank Barry Freifeld at LBNL who contributed significantly during the
development of the X-ray system in the MSCL-P. Special thanks go to John Roberts, Matthew Druce, Sally Marine and all
the staff at Geotek.

References

Claypool, G. 2005. Joint Industry Project (JIP) Gulf of Mexico Gas Hydrate Coring Update, Fire In the Ice (US DOE–NETL newsletter)
Summer 2005, pp. 9–13. Available online at: www.netl.doe.gov/technologies/oil-gas/publications/Hydrates/Newsletter/
HMNewsSummer05.pdf.
Collett, T.S., Riedel, M., Boswell, R., Cochran, J.R., Kumar, P., Sethi, A.K., Sathe, A.V. and NGHP Expedition-01 Scientific Party. 2006.
International Team Completes Landmark Gas Hydrate Expedition In the Offshore of India, Fire In the Ice (US DOE–NETL
newsletter) Fall 2006, pp. 1–4. Available online at: www.netl.doe.gov/technologies/oil-gas/publications/Hydrates/Newsletter/
HMNewsFall06.pdf.
Deep Sea Drilling Project. 1984. Design and operation of a wireline pressure core barrel. DSDP Tech Note 16, Washington (U.S. Govt.
Printing Office).
Dickens, G.R., Wallace, P.J., Paull, C.K. and Borowski, W.S. 2000. Detection of methane gas hydrate in the pressure core sampler (PCS):
volume-pressure-time relations during controlled degassing experiments. In Paull, C.K., Matsumoto, R., Wallace, P.J., and Dillon,
W.P. (Eds.), Proc. ODP, Sci. Results, 164: College Station, TX (Ocean Drilling Program), 113–126.
6 OTC 19601

Graber, K.K., Pollard, E., Jonasson, B. and Schulte, E. eds. 2002. Overview of ODP Engineering Tools and Hardware. ODP Tech. Note,
31. Available online at: www-odp.tamu.edu/publications/tnotes/ tn31/INDEX.HTM.
Kvenvolden, K.A., Barnard, L.A. and Cameron, D.H. 1983. Pressure core barrel: application to the study of gas hydrates, Deeps Sea
Drilling Project Site 533, Leg 76. In Sheridan, R.E., Gradstein, F.M., et al., Init. Repts. DSDP, 76, Washington (U.S. Govt. Printing
Office), 367–375.
Milkov, A.V., Dickens, G.R., Claypool, G.E., Lee, Y-J., Borowski, W.S., Torres, M.E., Xu, W., Tomaru, H., Tréhu, A.M., and Schultheiss,
P. 2004. Coexistence of gas hydrate, free gas, and brine within the regional gas hydrate stability zone at Hydrate Ridge (Oregon
margin): evidence from prolonged degassing of a pressurized core, Earth Planet. Sci. Lett. 222, 829–843.
Pettigrew, T.L. 1992. Design and operation of a wireline pressure core sampler. ODP Tech Note, 17 College Station, TX (Ocean Drilling
Program).
Riedel, M., Collett, T.S., Malone, M.J. and the Expedition 311 Scientists. 2006. Proc. IODP, Exp. Repts., 311: College Station, TX (Ocean
Drilling Program).
Schultheiss, P. J., Francis, T J G., Holland, M., Roberts, J.A., Amann, H., Thjunjoto, Parkes, R.J., Martin, D., Rothfuss, M., Tuynder, F.
and Jackson, P.D. 2006. Pressure Coring, Logging and Sub-Sampling with the HYACINTH system, In Rothwell, G. (Ed.), New
Techniques in Sediment Core Analysis, Geol. Soc. London, Spec. Pub. 267. pp. 151–163.
Takahashi, H. and Tsuji, Y. 2005. Multi-well exploration program in 2004 for natural hydrate in the Nankai trough, offshore Japan, OTC
17162. Proc. Offshore Technology Conference, Houston TX, 2–5 May, 2005.
Tréhu, A.M, Bohrmann, G., Rack, F.R., Torres, M.E., et al. 2003. Proc. ODP, Init. Repts., 204 [CD-ROM]. Available from: Ocean Drilling
Program, Texas A&M University, College Station TX 77845-9547, USA.
Xu, W. 2002. Phase balance and dynamic equilibrium during formation and dissociation of methane gas hydrate, in Proc. 4th International
Conference on Gas Hydrates, pp. 195–200, Yokohama, Japan.
Xu, W. 2004. Modeling dynamic marine gas hydrate systems, American Mineralogist, 89, 1271–1279.
Yun, T.S., Narsilio, G., Santamarina, J.C. and Ruppel, C. 2006. Instrumented pressure testing chamber for characterizing sediment cores
recovered at in-situ hydrostatic pressure. Marine Geology 229: 285–293.
Zhang, H., Yang, S., Wu, N., Schultheiss, P. and GMGS-1 Science Team, 2007 China’s First Gas Hydrate Expedition Successful, Fire In
the Ice (US DOE–NETL newsletter) Spring/Summer 2007, p.1. Available online at: www.netl.doe.gov/technologies/oil-gas/
publications/Hydrates/Newsletter/HMNewsSpringSummer07.pdf.

Figures & captions

Fig. 1—Annotated plot of pressure vs. temperature vs. time from an internal pressure-temperature data logger for a typical FPC
deployment, showing trajectory relative to gas hydrate stability (35 ppt salinity) as calculated after Xu (2002, 2004). Note that the
core remains well inside gas hydrate stability field at all times. Temperature is passively controlled by winch speed and, once on
deck, an ice bath. Final point (circle) is pressure and temperature of core inside cold analysis laboratory.
OTC 19601 7

Fig. 2—Cartoon of pressurized HYACINTH core manipulation using the PCATS in a 20-foot refrigerated container. a) “start”
position, b) the “catch” position after the autoclave has been attached with core under pressure, c) “retract” position showing the
core removed from the autoclave, d) “core log” position with the MSCL-P in place. (In a 40-foot container, the MSCL-P is in-line at
all times and step “d” is integrated with the core removal operation from “b” to “c”). Core can then proceed to e-I) “degas” position
with the core catcher under the gas escape port (core is removed when pressure drops to atmospheric pressure), or e-II) “cut”
position in the transfer chamber where the core liner is cut under full pressure, f) “push” position where the core is pushed into the
storage chamber, and g) “store” position where the manipulator rod is retracted, the ball valve closed, and the core free to be
transported in the storage chamber at full in-situ pressure.
8 OTC 19601

Fig. 3—Mating new equipment to the HYACINTH PCATS. Top left: quick-clamp. Bottom left: ball valve (65 mm internal diameter) and
mating flange. Right: The Instrumented Pressure Testing Chamber (Yun et. al., 2006) connected to the HYACINTH PCATS using
quick-clamps, showing ports with protruding sensors for P-wave velocity, shear wave velocity, electrical resistivity, and sediment
strength.
OTC 19601 9

Fig. 4—Data from a HYACINTH FPC core obtained on NGHP Exp. 1 using the PCATS (4a–4c) and a medical CT scanner (4d). All data
was collected at in-situ pressure. In the X-ray images, denser features (carbonate nodules) are dark; less dense features (gas
hydrate veins) are light. a) X-ray images with enlargements showing different gas hydrate morphologies in fine-grained sediment.
b) X-ray images of same pressure core, rotated every 15 degrees. The lighter patches (less dense) in (a) are revealed to be dipping
veins of gas hydrate seen from different angles. c) Two sets of data (X-ray images, P-wave and gamma density profiles) collected at
right angles to each other on the same pressure core. In the first data set, the profiles are taken perpendicular to (through) the
major gas hydrate veins, and a slight lowering of density and a smooth increase in P-wave velocity is seen in the area of greatest
gas hydrate concentration. In the second data set, the profiles are taken parallel to the major gas hydrate veins, showing low-
density zones and a complex (and sometimes inaccurate) P-wave velocity profile caused by pulse interference effects from hydrate
structures.) d) Horizontal X-ray computed tomographic slices of the same pressure core showing the complexity of cross-cutting
gas hydrate vein features present in this clay core.
10 OTC 19601

Fig. 5—Pressure vs. released gas volume, released water volume, and calculated total volume of gas during depressurization of a
pressure core from NGHP Exp. 1. Calculated total volume of gas includes gas released from PCATS and gas remaining within
PCATS, estimated from liquid released from PCATS. Gas samples were taken for gas composition analysis.

Fig. 6—Logarithm of methane concentration in millimoles per liter (mM) vs. depth in meters below seafloor (mbsf), plotted over
phase boundaries for Structure I methane hydrate calculated after Xu (2002, 2004). Methane concentrations that exceed the phase
boundaries for dissolved methane indicate either gas hydrate or free methane gas; one pressure core at this site was oversaturated
in methane and contained gas hydrate. This plot shows methane was undersaturated at depths 150 mbsf and higher, providing no
opportunity for gas hydrate formation at these depths. Dashed line is estimated depth of seismic Bottom-Simulating Reflector
(BSR) which agrees closely with the calculated thermodynamic Base of Gas Hydrate Stability (BGHS) as indicated by the solid
horizontal line of the phase diagram.

Potrebbero piacerti anche