You are on page 1of 81

Gas Hydrates-Not So Unconventional

For years, gas hydrates have been a resource that produced enormous inplace volume estimates. The U.S. Geological Survey (USGS) estimated as
much as 16.7Tm (590Tcf; 106Bboe) in-place on the North Slope of Alaska
in 1995 that covers an area of about 145,000 km (equivalent to
approximately 25 North Sea quadrants). World wide estimates also
focused on in-place gas volumes because, at that time, there was not any
methodology to assess what actually could be produced from these
deposits.
Since these first published reports, a number of tests, experiments, and
much research has been conducted on gas hydrates, allowing
assessments to focus on the technically recoverable resources for the first
time.
"Test results from the Mallik well and other research have pointed the way
to treating gas hydrate in a much more conventional manner," say Dr.
Timothy Collett, a leading gas hydrate researcher for the USGS. "Our
recently releasedAssessment of Gas Hydrates on the North Slope,
Alaska, 2008, is based on the same geologic elements to define a Total
Petroleum System (TPS) we use to assess conventional plays. We were
able, for the first time, to obtain the undiscovered, technically recoverable
volume of gas from hydrates for an area, in this case, northern Alaska. To
this end, these resources can be discovered, developed, and produced by
using current oil and gas technology."

Naturally occurring - huge volumes

The USGS estimates


about 2.4Tm (85Tcf; 15Bboe) of undiscovered, technically recoverable gas resources within gas
hydrates in Alaska. The area is mostly Federal, State, and Native lands covering 145,000 km (the
equivalent of some 25 North Sea quadrants). USGSA

gas hydrate is similar to ice, a

crystalline solid, where the gas molecule is surrounded by a cage of water


molecules. Many gases are suitable to form hydrates, however methane is
by far the most commonly found in marine and polar occurrences.
Experimenting with mixtures of chlorine and water, Humphrey Davy and
Michael Faraday may have been the first to discover hydrates. The 1823
paper from the Transactions of the Royal Society of London, by Faraday,
discusses experiments with the chlorine hydrate. Because a hydrate had
never been seen in nature, they remained an academic curiosity through
the 19th century. When hydrates were found to be forming in the gas
pipelines in the mid-1930s, the North American gas industry focused on
ways to predict and inhibit hydrate formation. In 1970, Soviet researchers
registered the possibility that gas hydrates could exist in large volumes in
the earth's crust, and, in 1972, naturally occurring gas hydrates were
eventually recovered from the Black Sea.
Since that discovery, marine hydrate-containing sediments covering huge
areas of the sea floor have been found in deep-sea drilling and coring in

more than 50 regions of the world's oceans and even in some large lakes
(Baykal in Russia is one). Hydrates are stable in marine sediments below
about 500 meters in water depth and stable in association with permafrost
in Polar Regions in both onshore and offshore sediments.
Gas is greatly concentrated in hydrates where a unit volume of methane
hydrate can produce about 160 unit volumes of gas at one atmosphere.
Worldwide, the methane contained (in-place volumes) in gas hydrates is
about twice the amount of carbon held in all the earth's fossil fuels.

Testing the production capacity

The presence of gas


hydrates is commonly identified by bottom simulating reflectors (BSRs) in oceanic sediments.

USGS

Sampled and

inferred gas hydrates occur world wide in oceanic sediment of continental margins and in
permafrost regions. Inferred gas hydrates are from bottom simulating reflectors (BSRs) on seismic
profiles. The Mt. Elbert and Mallik test sites are located at the top of the map. USGS Once

the

presence of gas hydrates was confirmed and it had been shown that the
volume of gas is huge, the next question was if this resource be produced
and how. Significant technical issues stood in the way before gas hydrates
could be considered a viable energy source. Hoping to find the answers,
Canada, Japan, India, and the United State launched ambitious research
projects.
"Past research has focused on three fundamental issues," according to Dr.
Collett. "Where do gas hydrates occur, how do they occur, and why do
they occur in a particular setting? However, there is a fourth issue. Until
recently, little had been done in evaluating the potential for viable
production of gas hydrates."
In 2002, the U.S. Department of Energy and BP Exploration (Alaska), Inc.
(BPXA), in cooperation with the USGS, initiated a research program on
Alaska North Slope (ANS) gas hydrates. The ultimate goal was a production
test to determine if the gas hydrates could be a viable energy resource.

Three ambitious research programs have also been undertaken in the last
decade next door in the Mackenzie Delta at the Mallik research site. Led by
Canada and Japan, the 1150m deep Mallik 2L-38 well was completed in
1998. For the first time, cores were brought to the surface from an Arctic
gas hydrate occurrence delineating approximately 120m of hydrate
bearing section within coarse grained clastic sands.
A second phase of this research program was undertaken in 2002 with a
broader five country partnership (Canada, Japan, Germany, USA, India)
that completed three additional wells and the first production test. This
would be the first time gas would be flared from hydrate.
While the 1998 and 2002 Mallik gas hydrate research projects enabled
many new ground breaking studies, the flow rates were modest. Japan
(JOGMEC) and Canada (NRCan) therefore decided to return to the Mallik
site in the winters of 2007 and 2008 to initiate a new testing program
focusing this time on a full scale production draw down test.

The world's first

The Mt. Elbert-01


stratigraphic test well, Milne Point Unit on the North Slope of Alaska, collected the first open-hole
formation pressure response data in a gas hydrate reservoir. This test and reservoir simulations are
some of the first steps that could lead to production of hydrate reservoirs. Photo: Tim Collett,
USGS"We

have now progressed way beyond just looking at the in-place

resources," says Dr. Collett. "Gas hydrate research is now looking at more
conventional prospects that can be produced by current technologies. The

February 2007 22-day field program at Mt. Elbert has taught us a lot about
how hydrates occur in porous media, why they occur where they do, and
the location and design of long-term testing."
The field program at the Mt. Elbert site was designed to collect as much
data as possible from the gas hydrate zones. Researchers were able to
obtain a full suite of open-hole well logs, over 152m of continuous core,
and open-hole formation pressure response tests. They used chilled oilbased drilling fluids to maintain excellent hole conditions and thus obtain
the best quality data.
"We found approximately 30m of gas hydrate saturated, fine-grained sand
reservoir," says Dr. Collett. "Gas saturations ranged from 60% to 75%
mostly due to differences in reservoir quality."
The culmination of the program was the open-hole tests. They were able to
obtain reservoir pressure response data in addition to gas and water
samples.
"This well clearly demonstrated that open-hole data can be safely and
efficiently obtained from shallow, sub-permafrost sediments," says Dr.
Collett.
Taking an even bigger step toward the production of methane from gas
hydrates was the Japanese and Canadian sponsored 2006-2008 Mallik 5L38 research well. Just completed in April 2008, the goal was to undertake a
longer production test than the 2002 Mallik project to advance new
research and development studies utilizing simple depressurization of the
reservoir. "The first tests were conducted during the winter of 2007," says
Dr. Scott Dallimore of the Canadian Geological Survey and lead researcher
for the Mallik Program. "Sand production prevented continuous pumping.
However, during a 12.5 hour interval, at least 830m of gas was produced.
It was the world's first gas production using large scale depressurization of
a gas hydrate."

"After some modifications to the pumping system, 2008 field operations


were geared for a long duration test of the same 1093 to 1105 m depth
interval tested in 2007," Dr. Dallimore adds. "They were able to
continuously pump for six days until down-hole pressures stabilized. Gas
flowed to the surface and was flared while pressure, temperature, gas, and
liquid flow rate data were measured."
The Japanese and Canadian team reported that the flow rates ranged from
2,000 to 4,000m per day. Cumulative gas production was over 13,000m
for the six-day test.

Eileen trend prospects

"What the Mt. Elbert and the Mallik research is telling us is that we should
be able to develop gas hydrate prospects," says Dr. Collett. "When we
started the Mt. Elbert project, we first mapped 14 gas hydrate prospects in
the Eileen trend (Milne Point area northwest of the Prudhoe Bay field).
From 3-D seismic data and rock physics relationships conditioned by well
data, we were able to predict gas hydrate pay thickness and saturation.
The entire Eileen gas hydrate trend, including the 14 prospects in the
Milne Point area, has approximately 0.93Tm (33Tcf; 5.9Bboe) gas in place
with initial reservoir modeling suggesting as much as 0.34 Tm (12Tcf;
2.2Bboe ) recoverable."
The Mt. Ebert prospect turned out to be the highest-ranked prospect in the
Milne Point area and was selected for the field data acquisition program.
This prospect had a strong and well organized seismic response. Amplitude
anomalies observed in two horizons were restricted within a well defined
three-way fault closure suggesting the accumulation was originally a freegas accumulation later converted to gas hydrate through depressed
thermal gradients associated with the development of permafrost.

Predictions were also made on individual reservoir sands that ranged in


thicknesses of 14 to 21m. The coring and logging program closely
confirmed these predictions.

Open-hole logs from the Mt. Elbert test clearly showing high resistivities in the gas saturated zones. The Unit D
sand had 68% gas hydrate saturation and Unit C 89%. The low resistivity sands below the Unit C, but still in the
hydrate stability zone, are water saturated. Tim Collett, USGSrnttttMap of the fault bounded amplitude feature
(yellow to magenta, yellow being the thickest and highest concentration of gas hydrate) defining the Mt. Elbert
prospect. Tim Collett, USGS

1.

2.

More exploration

The successful research approach demonstrated at the Mt. Elbert well and
prospect has opened the door to more exploration and testing of gas
hydrate prospects.
A 2-year program is underway to map, drill and log gas hydrate-bearing
sands in the deep water in the northern Gulf of Mexico. The project is the
latest phase in an ongoing Joint Industry project led by Chevron in
collaboration with the U.S. Department of Energy. They hope to further test
geological models and geophysical interpretations supporting the
existence of high gas hydrate saturations in reservoir-quality sandstone.
Another recently completed study off the Indian continental margin has

confirmed the presence of gas hydrates in four offshore basins. Not only
did they find gas hydrates, they found one of the richest gas hydrate
accumulations yet documented (Krishna-Godavari Basin, Bay of Bengal)
and found the thickest and deepest gas hydrate stability zone near the
Andaman Islands also in the Bay of Bengal.

Significant resource opportunity

Amplitude anomalies
associated with the gas hydrate drilling targets at the C Horizon. The anomalies are restricted
within a well defined three-way fault closure. Tim Collett, USGS "For

years, gas hydrate

research has been framed by extravagantly large numbers (see above)


that refer only to the in-place volumes of methane in gas hydrates," says
Dr. Ray Boswell, Manager-Methane Hydrate R&D Programs, U.S.
Department of Energy.

"These volumes are significant in an environmental context, and the


climate implications of gas hydrates are a major topic of study in the U.S.
effort. However, in the context of potential resources, what is most
interesting is what portion of that resource is a realistic future source of
gas supply. The array of recent field programs, combined with continuing
work in the lab, is now providing the first insight into those technicallyrecoverable volumes. Although these numbers lack eye-popping appeal,
they clearly indicate significant new resource opportunities that, given its
global distribution, have the potential to alter existing energy production
and supply paradigms," says Dr. Boswell.
"After a decade of R&D studies at Mallik, we have addressed many
scientific unknowns and established proof of concept that gas hydrate
production can be sustained with the pressure draw down technique, using
conventional oil and gas production techniques modified to accommodate
the unique physical properties of gas hydrates," Dr. Scott Dallimore adds.
"However, our testing at Mallik was only for six days duration. The next
milestone advances in production research are likely to come through
longer testing which will allow consideration of the field-scale response as
well as providing a basis to consider environmental issues associated with
production. It is my understanding that longer term testing programs are
actively being considered in northern Alaska, the Mackenzie Delta and
even in offshore settings like the Nankai Trough (offshore Japan)."

Long term endeavour

"Natural gas demand is expected to grow in the first decades of this


century as production from conventional sources decline," concludes Dr.
Tim Collett. "Vast deposits of marine and arctic gas hydrate are seen as a
possible source of natural gas, but there are many challenges to be
overcome before widespread production of natural gas from gas hydrate

are possible. The evaluation of gas hydrate as an energy source is clearly


a long-term endeavour that will require significant financial investment."

Hydrates and the petroleum system


"Recent research has helped address key issues of the formation, occurrence, and stability
of naturally occurring gas hydrates," says Dr. Collett. "The concept of a gas hydrate
petroleum system' is well within our reach. The source, migration, reservoir, trap, and
timing as in any petroleum system must be examined when developing exploration plays.
One further consideration, the gas hydrate pressure-temperature stability conditions must
also be accounted for." Indeed, most of the recent literature examines these factors when
dealing with gas hydrates.

Gas hydrate stability conditions:


The depth to the base of the gas hydrate stability zone (GHSZ) can be calculated from
down-hole temperature data and the pore pressure gradient. In most areas, the calculated
depth to the base of the GHSZ corresponds closely to the depth of the BSR inferred from
seismic data.

The Petroleum System


Source: Abundant methane gas sourced from either biogenic or thermogenic or both is
an important factor to the distribution of gas hydrates. Much of the offshore data points to
more biogenic origin for the methane gas; however both sources have been identified
through carbon isotope analyses.

Migration: Like any accumulation, there must be a way to get the hydrocarbons to the
reservoir. Migration pathways for both gas and water are a must for significant hydrate
volumes to accumulate. Many areas show gas hydrates to be closely associated with
structural features where gas and water can migrate through a fracture or fault system.
This can also be confirmed on a macroscopic to microscopic scale as most hydrates
recovered are either pore-filling in coarser grained sediments or fractured-filling in more
fine grained sediments.

Reservoir: High saturations of gas hydrates are consistently associated with high
porosity and permeability reservoirs, either good quality sand-rich sediments or fractured
mudstones.

Trap: Most high concentrations of gas hydrates in permafrost environments are


associated with some type of trapping mechanism. The traps being considered for early
exploration have been mostly structural as was the case with early conventional
hydrocarbon plays.

Timing: Finally, as with any accumulation, the geologic history has to be such as to trap
and preserve the migrated hydrocarbons.

Gas hydrates concentrate along porous and permeable zones such as fractures (core from India on top) to the good quality
conglomeratic sandstones found in the Mallik well. Photos: Tim Collette, USGS and 2006-2008 Canada-Japan Mallik Project

Gas Hydrates - Part I: Burning Ice

In a series of articles we will explain in a simple way what gas hydrates


are, where they can be found in nature, and what their physical properties
are. Natural gas hydrates are a potential source of energy and may play a
role in climate change and geological hazards. A number of countries,
including Japan, USA , India, China, Korea, Germany, and others, have
national programmes for studying and industrial production of natural gas
from hydrates.

Doubted by Experts

Yakutia, or the Sakha Republic, is a vast unexplored region of Siberia in the


north-east of Russia. It is one of the rare places left on Earth where large
expanses of wild nature mountains, rivers, lakes, forests, tundra are
saved untouched by civilisation. Yakutia is located in a permafrost zone,
although the climate is continental, with summer temperatures in July in
the city of Yakutsk reaching 40C, as hot as Tokyo, while in winter it goes
down to -50C.

The wild mountain landscape of Siberian taiga along the Lena River in Lenskie Stolby
National Nature Park, Yakutia, Russia. Source: Tatiana Grozetskaya / Dreamstime.com

Yakutia is where we find the Cold Pole, the coldest place in the northern
hemisphere, where in 1924 the lowest ever temperature for our
hemisphere, -71.2C, was recorded. Here, beyond the Polar Circle, the
night lasts all winter, and the day all summer.
It is 1963, and Yuri Makogon, then at the Moscow Oil-Gas Gubkin Institute
with a fresh MSc degree in petroleum engineering (now recently retired

from Texas A&M University), is staying in the north-western part of Yakutia.


He participates in the drilling of the Markhinskaya well, down to 1,800m
depth where the temperature is 3.8C. The well reveals a section of rock at
0C temperature at a depth of 1,450m, where permafrost ends at a depth
of around 1,200m. Yuri recognises that this section of the rock matches
hydrate formation conditions. He hypothesises the possibility that gas
hydrates can exist and accumulate in such cold layers. His hydrate
hypothesis is seriously doubted by the experts, and the idea needs
experimental verification, so in 1965 Yuri experimentally proves that gas
hydrates may accumulate as large natural deposits in porous rock. In 1969
this discovery is formally recognised and registered in the USSR. Yuri is
today recognised as the first to discover that hydrates of natural gas can
accumulate as deposits in nature.

Easy Rapid Process

Thermogram of the Markhinskaya No. 1 well. Source: Makogon,


1966Monterey

Canyon is a submarine canyon in Monterey Bay, California. It

begins at the middle of the Monterey Bay, and extends 153 km into the

Pacific Ocean where it terminates at the Monterey Canyon submarine fan,


reaching depths of up to 3,600m below surface level at its deepest. The
canyons depth and nutrient availability due to the regular influx of
nutrient-rich sediment provide a habitat suitable for many marine life
forms.
In 1996, the remotely operated vehicle (ROV) Ventana comes to rest in the
Monterey Bay Canyon at 910m depth where the temperature is 4C.
Operated by scientists from Monterey Bay Aquarium Research Institute,
Stanford University, and the US Geological Survey, an amount of methane
is injected into the water and bottom sediments. Within minutes, this
mixture of gas and water forms into a solidified block, bright white and
fluffy. The experiment shows not only that in natural seawater but that the
process is extremely easy and rapid, given that the pressure and
temperature conditions are right. Prior calculations had shown that the
local hydrographic conditions gave an upper limit of 525m for the
pressure-temperature (P-T) boundary defining methane hydrate formation
at this site, and thus the experiment takes place well within the stability
range for this reaction to occur.

A massive hydrate layer obtained from fine-grained sediment recovered in a marine setting in the Gulf of Mexico.
Bill Winters, USGS

Compared to hydrate recovered in a coarse-grained gravel layer from an Arctic setting, the Mallik 2L-38 well,
drilled in the Northwest Territories of northern Canada. Bill Winters, USGS

1.

2.

Gas Hydrates

A model of methane hydrates cage-like


structure in which methane (the grey/green molecule) is enclosed inside a lattice of water (the red
molecules). This unit cell joins together with other unit cells by sharing faces to build the hydrate
solid, even at temperatures well above the melting point of water ice. When gas hydrates
dissociate (melt), the crystalline lattice breaks down into liquid water (or ice) and the gas is
released. Jarle Huseb, StatoilA

model of methane hydrates cage-like structure in

which methane (the grey/green molecule) is enclosed inside a lattice of


water (the red molecules). This unit cell joins together with other unit cells
by sharing faces to build the hydrate solid, even at temperatures well
above the melting point of water ice. When gas hydrates dissociate (melt),
the crystalline lattice breaks down into liquid water (or ice) and the gas is
released. Jarle Huseb, Statoil
You are familiar with ice: it is cold, hard, slippery, and buoyant. You can find
icebergs and ice floes on the surface of the ocean in polar regions. But
there is another, higher-pressure form of ice trapped at and beneath the
seafloor that is far less familiar. This ice, which is known as gas hydrate, is
created by the reaction of gas predominantly methane with water at low
temperature and high pressure to form a crystalline solid. Such compounds
are a common occurrence buried in permafrost and continental slope
sediments around the world ocean. For methane hydrate, the water and
gas combine in a ratio of six molecules of water to one of gas. The gas

draws in the water molecules to form a cage within which the methane
molecule flutters in a strange molecular dance, feeling the attraction and
pull of the hard ice walls of its prison, but without ever touching. Most
natural gas hydrate appears to be in this structure, with methane as the
trapped guest molecule, although alternative structures have also been
identified, with guest molecules such as isobutane and propane, as well as
lighter hydrocarbons. Gas hydrates provide an extremely effective way of
storing natural gas or methane (CH4). At standard atmospheric
temperature (20C) and pressure (1 atm) conditions, 1 m of solid methane
hydrate is equivalent to 160 m of free gas. Although global estimates
range widely by more than two orders of magnitude, the most cited value
is that of Kvenvolden in 1988, at 2x1016 m of gas, or 10,000 gigatons of
carbon. In comparison, estimates for the known combined reserves of
conventional hydrocarbons (natural gas, oil, coal) are about half of that
value.

An Unusual Catch

Barkley Canyon is located off Vancouver Island, Canada, on the northern


typically sites of underwater landslides as well as movement of sediment,
organic material and nutrients to the deep ocean. The upwelling of
nutrients driving the regions rich biodiversity makes the upper slope a
crucial study area for scientific and environmental policy purposes. It also
makes it a popular region for commercial fishing.
One November day in 2000 the vessel Ocean Selector is trawling over the
Barkley Canyon in an area of 800m water depth. (Trawls are fishing nets
that are pulled along the bottom of the sea or in midwater at a specified
depth). When the captain pulls up the net, the crew see the most unusual
catch: over a ton of strange white solid that fizzes and crackles on the
deck. They shovel it overboard, fortunately without harm. The crew has by
accident discovered a large seafloor deposit of methane hydrate. If one of
the seamen had lit his smoke it could had resulted in a disaster as the
crackling actually released significant amounts of methane gas.

Methane is a fuel, and despite being trapped in an ice, it will readily burn.
For this reason, methane hydrates are known as burning ice.
That day in November introduced the scientific community to a new,
massive seafloor outcrop of gas hydrate on a 500m wide, 1 km long
plateau perched 150m above the canyon floor. In 2004, Canadian
scientists explored and verified the site, and the ROV revealed pinnacles
cutting up through the seafloor. Massive outcrops are exposed in 12m
high mounds covered by a thin veneer of sediment. A light yellow
condensate fluid is present in the surrounding sediment and associated
with the hydrate, causing yellow staining. With buoyant hydrate slabs on
the seafloor reaching 7m in length and 3m in height and with only a thin
sediment cover, the hydrate mass must continue deeper and be anchored
below the surface.
Although methane hydrate is known to be stable at the sea floor for water
depths greater than 500600m at temperate latitudes, observed outcrops
of hydrate at the sea floor are rare and somewhat poorly understood.

Gas Hydrate Stability Zone (GHSZ)

Permafrost - Methane gas hydrate stability zone (GHSZ) forpermafrost and marine settings. The pressuretemperature phase boundary (idealised) is shown as a green curve, and the local thermal (assumed) gradients are
in red. In permafrost, the GHSZ can begin at 100300m depth and extend for hundreds of meters beneath the base
of the permafrost (typically occurring at 150600m depth). Source: Courtesy SEG


Marine - In marine sediments, the GHSZ typically begins below 300600m and extends for hundreds of meters. The
thickness of the marine GHSZ depends on sea floor water temperature (typically 34C), salinity, geothermal
gradient and depth. Source: Courtesy SEG

1.

2.

Gas hydrates are stable only under high-pressure and low-temperature


conditions. The region where gas hydrates are stable the gas hydrate
stability zone (GHZS) is defined by the intersection of the pressuretemperature phase boundary and the local geothermal gradient. Because
of the nature of the conditions for stability, gas hydrates are usually found

only in permafrost regions and on continental slopes where water depth


exceeds 300500m.
As shown in the figure for the marine setting, the top of the GHSZ occurs
above the sea floor. However, the ocean water does not contain enough
gas to stabilise hydrate and the top of the GHSZ is normally defined at the
sea floor. Here, the temperature is normally 34C. Going down into the
sediment, the temperature slowly increases; the global average of the
geothermal gradient is 0.02C/m. While the pressure increases with depth,
after 5001,000m depth the temperature becomes too high for hydrates to
remain stable. This is the base of the GHSZ.
In permafrost, the situation is similar. The top of the GHSZ is where the
temperature line crosses the hydrate stability line, often beginning at 100
300m depth. The GHSZ typically extends for hundreds of meters. The depth
of the GHSZ is also related to the base of the permafrost, which is at 0C.
The deeper the base of the permafrost the deeper the GHSZ becomes.

Gas Hydrates - Part II: Rock Physics, an Introduction

A major and obvious difference between ice and hydrate is that the
hydrate is highly flammable (burning ice).
It is of course the methane content in hydrate that makes it of commercial
interest. In fact the difference between ice and methane hydrate is not
huge: one could consider hydrate as one dash of methane gas and six
dashes of pure ice. Ice has a hexagonal crystalline structure, with a density
of roughly 0.92 g/cm at 0C and increases to 0.93 for a temperature of
3

-180C. The density of methane hydrate is slightly less (0.90 g/ cm ), since


3

there is only 1 mole of methane per 5.75 moles of ice. Another way to
formulate this is to say that whereas ice has the formula H O, the empirical
2

chemical formula for methane hydrate is (CH )8(H O)46.


4

So, if the density difference between ice and hydrate is negligible, what
about other crucial seismic parameters, such as velocities?

Rock Physics Experiments

In 2001 Michael Helgerud presented a comprehensive PhD thesis at


Stanford where he analyzed wave velocities in gas hydrates and sediments
containing gas hydrates.
For pure gas hydrate samples made in the laboratory he measured P-wave
(compressional) velocities around 3700 m/s and S-wave (shear) velocities
around 1950 m/s. Only small variations with temperature and confining
pressure were observed.

The ratio between P-wave velocity of pure hydrate versus ice is 0.98. For
the S-wave velocity the corresponding ratio is close to 1.0. This means
that, acoustically, pure ice and pure methane hydrate are very similar. This
is maybe as expected since the dominant crystalline structures of the two
are similar.

Compressional and shear velocity versus time. The figure shows the response to a warming of the sample from 5
to 20C followed by a cooling back to 5C again. Confining pressure: 9,000 psi. Source: Helgerud (2001)

Compressional and shear wave velocities versus confining pressure for methane hydrate. The horizontal scale is
from 4,000 to 9,000 psi. Source: Helgerud (2001)

Compressional and shear velocity versus time. The figure shows the response to a warming of the sample from 5
to 20C followed by a cooling back to 5C again. Confining pressure: 9,000 psi. Source: Helgerud (2001)
Compressional and shear wave velocities versus confining pressure for methane hydrate. The horizontal scale is
from 4,000 to 9,000 psi. Source: Helgerud (2001)

1.

2.

Rock Physics
Rock physics provides the connections between geophysical (elastic and
electromagnetic) properties of the rock measured at the surface of the
earth, within the borehole or in the laboratory, with the intrinsic properties
of rocks, such as mineralogy, porosity, pore shapes, pore fluids, pore
pressures, permeability, electrical resistivity, viscosity, stresses and overall
architecture such as laminations and fractures. These parameters affect
how seismic and electromagnetic waves/fields physically travel through the
rocks. Establishing relationships between geophysical expression and
physical rock properties therefore requires knowledge about the
elastic/electromagnetic properties of the pore fluid and rock frame, and

models for rock-fluid interactions.


Equations that attempt to describe the relationships between seismic
velocities and lithology, porosity, pore fluid, etc. are either theoretical or
empirical.
Zhijing Wang summarizes the tension between theoretical and empirical
approaches this way: most direct measurements are carried out either in
the laboratory or inside a borehole, whereas most theoretical calculations
are based on the Gassmann equation (Gassmann, 1951) because of its
simplicity and ease of use. Direct laboratory measurements are carried out
in controlled, simulated reservoir environments and provide accurate
effects of pore fluids on seismic properties. Direct borehole measurements,
however, are often affected by uncontrollable factors such as stress
concentration, hole washout, mud invasion/filtration, and saturation
conditions. In both laboratory and borehole measurements, the wave
frequencies are higher than seismic frequencies.
Jan Dewar summarizes that the crux of all this is that there are a great
number of relationships between seismic velocities (and constituent elastic
properties) and rock parameters, all valid to some degree but not always
valid, and many that do not illuminate the physical principles involved. The
trick is to try to gain a fundamental understanding so that, on a practical
basis, the different relationships can be evaluated for their applicability to
solving specific problems.
Theoretical rock physics modeling can also be applied to understand the
physical properties of natural gas hydrate systems and accumulations. A
major goal is to establish linkages between gas hydrate
concentration/saturation in sediments and the measureable physical
properties, like P- and S-wave velocities and electrical resistivity. Modeling
the elastic properties of sediments as a function of gas hydrate content can
be achieved in various ways, such as effective medium modeling or threephase Biot theory. The effective medium theory can incorporate the effects

of cement and grains. In the case of gas hydrate, the hydrate can form in
various ways. The effective seismic velocities vary quite strongly
depending on which formation scenario is used.
To model gas hydrate systems via rock physics one needs to define the
elastic/electromagnetic properties of the system in terms of (i)
elastic/electromagnetic properties of the (unconsolidated) sediments that
host the hydrates, (ii) elastic/electromagnetic properties of the embedded
gas hydrates, (iii) the concentration of hydrates in the sediments, and (iv)
geometrical details of the distribution of hydrates within their host
sediments. The inverse modeling problem is to infer hydrate concentration
from geophysical measurements.
Acoustic well-logs with full waveform show a pronounced decrease of wave
amplitude in hydrate-bearing zones.

Compressional and shear velocity versus porosity for ice. Source: Helgerud (2001) P-wave velocity versus
porosity assuming that the methane hydrate is a part of the rock frame (black line), of the pore fluid (green line).
The red line shows a rock which is 100% water saturated (no hydrate), the blue solid line represents 1% patch
gas saturation, and the dark blue line represents 1% homogenous gas saturation. Source: Helgerud (2001)

Compressional and shear velocity versus porosity for ice. Source: Helgerud (2001) P-wave velocity versus
porosity assuming that the methane hydrate is a part of the rock frame (black line), of the pore fluid (green line).
The red line shows a rock which is 100% water saturated (no hydrate), the blue solid line represents 1% patch
gas saturation, and the dark blue line represents 1% homogenous gas saturation. Source: Helgerud (2001)

1.

2.

Fluid, Frame or Cement?

Is hydrate a part of the fluid, the frame or is it cement?


We have seen that pure hydrate and ice are very similar with respect to
traditional seismic parameters, such as velocities and densities. However,
in nature, the hydrate is found as a part of a rock, making the rock physics

relations more complex. What happens when methane hydrate enters into
a sedimentary rock? Hydrate is found both in clay rich sediments and in
sands. Several models have been proposed for this, varying from
regarding the hydrate as a part of pore fluid fill, via being a part of the rock
frame or acting as cement between sand grains. Helgerud found that the
P-wave velocity is slightly higher when assuming that the hydrate is a part
of the rock frame.
There are two ways of addressing this problem: either to perform
measurements in wells drilled into hydrate-bearing rocks, or to inject
hydrate in a controlled way into a rock in the laboratory. We will discuss
both methods in the next issue of GEO ExPro.

1.

2.

In this box we show two examples of some of the most frequently used rock
physics relations. The most famous model in rock physics is the Gassmann
(1951) equation, which is used to describe the effect of gradually changing
the pore fluid in, for instance, a sandstone. It needs calibration prior to use,
for which we often use well logs. The figure on the right shows how the
relative P-wave velocity varies as a function of water saturation, assuming
that the pores are filled with either oil or water. Another key reservoir
parameter is the pore pressure. For wave velocities it turns out that it is the
effective pressure (which is approximately equal to the overburden

pressure minus the pore pressure) that controls the velocity. The simplest
version for a theoretical model that includes changes in effective pressure
is the Hertz-Mindlin model. This is a contact model, where it is assumed
that the rock grains are identical spheres with a given contact area that
increases with the effective pressure. As this contact area increases, the
velocity increases, as shown below. The black line corresponds to a Mindlincoefficient of 1/6 and the red line to 1/10, demonstrating that also in this
case we need a calibration procedure prior to practical application of this
simple model. In most cases we assume that the density does not change
as the pore pressure changes. However, in some cases the rock might
compact or undergo stretching, and in such cases the porosity will change
and hence also the density. Typical examples are chalk reservoirs that may
compact by several meters, due to production. Rock physics plays a crucial
part both in exploration and production geophysics: advanced and
quantitative use of, for example, pre-stack seismic data require rock
physics input to link the seismic parameters to key reservoir parameters
such as saturation, pressure and porosity.

Gas Hydrates - Part III: Rock Physics Hydrate Experiments

In the previous issue of Geo ExPro we discussed whether gas hydrate is a


part of the pore fluid or a part of the rock itself. It was shown that this
matters if we want to estimate geophysical parameters for a gas hydrate
rock. There are two ways (at least!) to investigate this further: either to
observe geophysical parameters from wells being drilled through a
hydrate-bearing rock, or to inject methane hydrate into a rock sample in
the laboratory. We will discuss examples of both and put particular
emphasis on a recent lab experiment performed in China.

Seismic Attenuation

Like the presence of free gas, the presence of gas hydrate affects seismic
attenuation. Attenuation has the potential to map hydrate concentrations
through the effect of local blanking of sediment stratigraphic reflectivity.
However, there are few studies related to seismic attenuation in hydratebearing sediments and it remains an open topic for future studies. The
attenuation of seismic energy by gas hydrates is likely to depend on the
concentration of hydrate, the thickness of hydrate, the mechanism of
hydrate formation, and the dominant frequency of the seismic
measurements, in addition to the lithology changes.
VSP data in the Mackenzie Delta indicate that quite thick hydrate-bearing

zones have significant attenuation at seismic frequencies of 10200 Hz (Qvalues, which describe inverse attenuation, of around 10).

Observations from Wells

Methane hydrate
concentration in the sediments at ODP site 995 from P-wave sonic and resistivity logs. Column 1:
Comparison of P-wave velocity (red line) with model results assuming methane hydrate (solid black
lines) or homogeneously distributed methane gas (dashed lines) are part of the pore fluid. Column
2: Comparison with model results assuming methane hydrate is a sediment frame component (solid
black lines) or methane gas is patchily distributed in the pore space (dashed lines). Column 3:
Comparison of methane hydrate concentration estimates derived from the resistivity log to
estimates derived from the compressional wave sonic log using the gas hydrate as sediment frame
component model. Source: Helgerud, 2001In

a field example from a well drilled at the

Blake Ridge (ODP site 995) offshore South Carolina, USA, Helgerud found
that the rock physics model that assumes hydrate is part of the rock frame

gave a reasonable fit between hydrate concentrations estimated from Pwave well log measurements and those obtained from the resistivity log.
The deviation between the two models is not huge, but based on the well
log observation it is clear that the model which assumes hydrate is part of
the rock itself and acts as a kind of cement explains the well log data best.
A relative good fit between observed saturations of hydrate and those
estimated using the Archie equation (discussed in more detail below) is
achieved.

Qingdao Experiment

This year, a very interesting experiment on hydrate formation and


dissolution was presented by Hu and Ye of the Qingdao Institute of Marine
Geology in China. They measured P- and S-wave velocities as the hydrate
concentration in two rock samples was gradually increased from zero to
70%. The sediment samples were first immersed by pure water, then
loaded into a high-pressure vessel, and injected with methane. The
temperature was kept at 2C for hydrate formation. To simulate the effect
of hydrate dissolution, the temperature was gradually increased to room
temperature.

Microstructural Models

Existing
microstructural models of gas hydrate-bearing sediments (adapted from Dai et al., 2004).
Observations from various fields show that gas hydrate mainly forms as a supporting matrix-grain
(scenario 3). Source: Lasse AmundsenHigher gas hydrate concentrations yield an increase in

the elastic properties. There are a number of rock physics models in the literature that
attempt to quantify this effect (Dai et al., 2004). The cementation models treat the grains
as randomly packed spheres where the gas hydrates occur at the contact point (model 1)
or grow around the grains (model 2). However, these models predict large increases in the
elastic properties with only a small amount of gas hydrate but stay relatively flat as the
concentration of gas hydrate increases further.

Models 3 and 4 are variations of the cementation models, but consider the gas hydrate as
either a component of the load-bearing matrix or filling the pores. These use the HertzMindlin contact theory to calculate dry rock moduli at critical porosity (3540%). A
modified lower Hashin-Shtrikman (HS) bound is used for porosity smaller than critical
porosity, and a modified upper HS bound is used for porosities larger than critical porosity.
The Gassmann equation is then used to derive the composite rock velocities.

Model 5 is an inclusion-type model that treats gas hydrate and grains as the matrix and
inclusions respectively, solving for elastic moduli of the system by iteratively solving either
the inclusion-type or self-consistent type equations.

Models 15 all consider gas hydrate as homogeneously distributed in the sediments.


However, evidence of gas hydrate coring reveals that hydrates often exist as nodules and
fracture fillings in the shallow shaly sediments. This geometry is illustrated in model 6. No
quantitative treatment of this geometric model exists in the literature.

P- and S-wave
velocities versus hydrate concentration for a consolidated sand sample. Source: Hu and Ye, 2012

P- and S-wave
velocities versus hydrate concentration for an unconsolidated sand sample. Source: Hu and Ye,
2012

For an unconsolidated sand sample, they found that the P-wave velocity
increased from approximately 1,600 m/s for zero hydrate concentration to
approximately 3,600 m/s for 70% hydrate concentration. The
corresponding values for the S-wave velocity were 600 and 1,600 m/s,
respectively. This means that the Vp/Vs-ratio decreases from 2.7 for no
hydrate to 2.25 at 70% hydrate concentration. For comparison, it is
interesting to note that Helgerud (2001) measured a Vp/Vs-ratio of
approximately 1.9 for pure hydrate. A linear extrapolation of the measured
Vp/Vs-ratio from Hu and Yes experiment yields a Vp/Vs-ratio of
approximately 2.0.
For the consolidated sample, the Qingdao experiment showed, as
expected, higher acoustic velocities: a P-wave velocity increase from 4,250
m/s to 4,700 m/s as the hydrate concentration is increased from 0 to 70%.
The corresponding numbers for the S-wave velocity are 2,500 m/s and
2,750 m/s. These numbers correspond to a constant Vp/Vs-ratio of
approximately 1.7, practically independent on hydrate concentration.
Hu and Ye also noticed a hysteresis effect for the P-wave velocities. For the

unconsolidated sand sample they found that for a hydrate concentration of


50% the P-wave velocity was 2,700 m/s as hydrate was formed and only
2,000 m/s as hydrate was dissolved. The S-wave velocity showed a similar,
but weaker hysteresis effect. For the consolidated sample they found that
this hysteresis effect was opposite: the P-wave velocity during formation of
hydrate was lower than the corresponding value for dissolution. Hu and Ye
suggest that this hysteresis effect is caused by two very different
mechanisms for unconsolidated and consolidated rocks. For
unconsolidated rocks they suggest that the high velocity during formation
is caused by hydrate cementation, and that this cementation process
behaves differently during formation and dissolution. For consolidated
rocks they discuss a two-stage formation process of methane hydrate: first
a water-hydrate slurry is formed, followed by a slow solidification process.
This two-stage process might explain that P-wave velocities are lower
during formation compared to dissolution for the same hydrate
concentration. These results are interesting, and might be of importance
for geophysical analysis of data from hydrate-bearing sediments. A
velocity difference of 700 m/s related to whether hydrate is slowly
dissolving or being formed should be possible to detect. For instance, if a
hydrate-bearing rock is being produced by heating the rock, hydrate will
start to dissolve, and a significant change in P-wave velocity should be
detectable on conventional time lapse seismic data. This depends of
course on the initial concentration of hydrate within the porous rock; for
low concentrations the changes in P-wave velocity will be less pronounced,
as demonstrated by Hu and Yes experiment.

Resistivity Variations

Nigel Edwards is one of the pioneers in investigating the


mapping of submarine hydrates using electromagnetic surveying methods.
In 1997 he published a paper describing how seafloor transient electric
dipole-dipole methods can be used to detect hydrates. He uses a
simplified version of the Archie equation (see box above) to assess the

effect of methane hydrate concentration (Sh) on the formation resistivity


():
Here, is the resistivity of the sea water (0.3 Ohm-meter) and denotes
porosity. We clearly see from this equation that the formation resistivity
increases rapidly as the hydrate concentration increases, a fact that makes
methane hydrates a candidate for electromagnetic surveying and
complementary to seismic surveying. Spangenberg and Kulenkampff
(2006) use artificial samples (glass beads) to measure the resistivity
versus hydrate concentration. They found that the resistivity increased
from 5.1 Ohmm at zero percent hydrate saturation to 265 Ohmm at 95%
hydrate saturation. They also investigated the accuracy of the Archieequation given above, and found that the exponent (equal to 2 in the
above equation) deviated significantly from 2, especially for hydrate
saturations above 50%, where an exponent above 4 was found.

Archies Equation

Rock physicists talk about velocities and elastic parameters, because


these are what link physical rock properties to seismic expressions. Petrophysicists are
generally less concerned with seismic, and more concerned with using wellbore
measurements to contribute to reservoir description.

In the field of petrophysics, Archies equation relates the in-situ electrical conductivity of a
sedimentary rock to its porosity and brine saturation. Named after Gus Archie (1907
1978), his empirical relationship laid the foundation for modern well log interpretation as it
relates borehole electrical conductivity measurements to hydrocarbon saturations.

Archies relation for gas hydratebearing sediments reads:

where is the true or measured bulk formation resistivity, is the pore fluid resistivity, is the

sediment porosity, 0.5 < a < 2.5 is a constant, and 1.5 < m < 3 is the cementation factor
that increases as the grains becomes less spherical with depth. Sh is the hydrate
concentration. The value of n depends on the grain-hydrate-fluid structure. If n is relatively
large, gas hydrate forms in a way that strongly impedes current flow and increases bulk
sediment resistivity (e.g., gas hydrate located in the pore throats), whereas if n is relatively
small, gas hydrate forms in a way that has a lesser effect on sediment resistivity (e.g., gas
hydrate occurrence in the pore space, making minimal contact with sediment grains).
Pearson et al., (1983) calculated an estimate for n of 1.94; however, modelling by
Spangenberg (2001) has shown that n depends somewhat on grain size distribution and
the gas hydrate saturation itself.

In practice, for marine sediments, the pore fluid resistivity usually can be adequately
estimated from the equation of state of seawater, if in-situ pressure, temperature and
salinity are known. The gas hydrate saturation now can be estimated directly from the
equation above, given that the empirical Archie parameters a, m and n are known.

Gas Hydrates Part IV: Where Are Gas Hydrates Found?

While research on methane hydrates is still in the early stages, these


research efforts could potentially yield significant new supplies of natural
gas and further expand US energy supplies.
On August 31, 2012 US Secretary of Energy, Dr. Steven Chu, announced an
investment of $5.6 million in research on methane hydrates.
From Part I (GEO ExPro Vol. 9, No. 3) of our series on gas, the key learning

is that four magic ingredients must be present for gas hydrates to exist.
They form when there is a sufficient supply of water and gas,
predominantly methane (99%), at relatively low temperatures and high
pressures, with temperature and pressure in the so-called Gas Hydrate
Stability Zone (GHSZ), (see box, p34). Favourable hydrate formation
conditions exist off the coasts on the continental margins and below the
permafrost.
In marine settings, temperature is controlled by the ocean bottom water
temperature and the geothermal gradient at any given location, while
pressure is controlled by sea level. In aquatic sediment where water
depths exceed about 300m and bottom water temperatures approach 0C,
gas hydrate is found at the seafloor to sediment depths of about 1,100m.
The general temperature range is from 2 to 20C.
In a permafrost setting, however, temperature gradients are considerably
lower than in the ocean. The ambient temperature and the thickness of
the permafrost layer therefore are of significant importance for the
stability of gas hydrate. In polar continental regions, methane hydrate can
occur at depths ranging from 150 to 2,000m, with a general temperature
range from -10 to 20C.

Bottom Simulating Reflector (BSR)

Distribution of known
methane hydrate accumulations. The yellow dots show where actual samples of gas hydrate have
been recovered whereas the red dots show where gas hydrate occurrences have been inferred
based on BSRs and well logs. It is evident that gas hydrates are found along most continental shelf
and slope regions and in many permafrost areas. Hydrates have also been found in inland seas
(e.g., Black Sea and Caspian Sea) and in fresh water lakes (Lake Baikal). (Courtesy of Council of
Canadian Academies (2008), based on data from Kvenvolden and Rogers, 2005.)

Seismic example of
Bottom Simulating Reflector (BSR), from Indonesia. The water depth is 1,8002,000m. Observe that
the BSR has opposite polarity to that of the seabed. The average thermal gradient, about 35C,
permits the gas hydrate to exist down to around 300m sediment depth below that temperature is
too high and methane can exist only as free gas. Source: Statoil Indonesia The

geothermal

gradient is important. At a certain depth in ocean sediment the


geothermal gradient makes the sediment too warm to support the solid
gas hydrates, so any methane produced below this depth will be trapped
as a layer of free gas in the pore space beneath the solid gas hydrate
layer. Often, but not always, the interface between the gas hydrate and
the free gas is an anomalous seismic reflector called a Bottom Simulating

Reflector (BSR), as this reflector necessarily is roughly parallel to the


seafloor morphology along isotherms. BSRs therefore need not follow the
trend of stratigraphic horizons, but may intersect them.
In seismic sections, BSRs are usually characterized by large amplitudes but
exhibit reversed polarity compared with the sea-bottom reflection.
The BSR indicates the lower boundary of gas hydrate stability.
Consequently, gas hydrate is often assumed to exist above the BSR;
otherwise, the free gas below the BSR would have migrated upwards. But,
while a BSR does illustrate the volume of sediment inside the stability zone
it does not provide information on the actual hydrate saturation in-place.
BSRs can be observed even when very little hydrate is present, and BSRs
need not always be observed in hydrate-bearing sediments.

Exploration for Gas Hydrates

Gas hydrate saturation


of the pore space. Source: modified from Klauda & Sandler (2005)

Predicted thickness of
the global GHSZ. The thickest zones (600800m) are mainly situated in high-latitude regions (Arctic
and Antarctic) due to low bottom water temperatures which maintain conditions required for
hydrate formation. Extended GHSZs are also observed along continental margins (>500m) where
thick sedimentary sequences are deposited. In these settings the extent of the GHSZ is not limited
by sediment thickness so that free gas can accumulate in sediments below it. Note that this
definition of the global GHSZ gives an upper limit to possible gas hydrate occurrences. Source:
modified from Burwicz et al 2011To

date, around 100 sites have been identified as

containing gas hydrate deposits. Samples have been taken at


approximately 20 different sites, while at another 80 sites the existence of
gas hydrate has been suggested by seismic evidence, in the form of BSRs.
Exploration for gas hydrates is not much different from exploration for
conventional hydrocarbons: important factors to recognize are source,
migration, reservoir, and seal.
If there is not sufficient gas supply, there will be no gas hydrates. Two
distinct processes produce hydrocarbon gas: biogenic and thermogenic
degradation of organic matter. Biogenic gas is formed at shallow depths
and low temperatures, up to 7580C, by anaerobic bacterial
decomposition of sedimentary organic matter. It is very dry and consists
almost entirely of methane. In contrast, thermogenic gas is formed at
deeper depths, much deeper than the GHSZ, in the temperature range 50

200C by thermal cracking of sedimentary organic matter into


hydrocarbon liquids and gas. This type of gas, which is common in
conventional gas reservoirs, can be dry, or can contain significant
concentrations of wet gas components (ethane, propane, butanes) and
condensate.
Fluid migration from the source through faults, folds, and fractures into the
GHSZ plays a critical role in the formation of a gas hydrate accumulation.
Rapid gas transport is required to concentrate gas in permeable reservoir
sediments where gas hydrate crystallizes. Water transport is usually
thought to be less important because water is virtually omnipresent in
sediments, although it may be a limiting factor for gas hydrate
crystallization in some areas. Sand-rich reservoir environments are better
than clay-dominated systems. As far as seals are concerned, gas hydrates
themselves are the seals.
The possibility of production from hydrates is highly dependent on the
particular reservoir characteristics. Many of the known marine deposits are
probably unfeasible for hydrate production. The candidates that are
currently being explored are high concentration accumulations in coarsegrained sand environments with high porosity and permeability.

Global GHSZ Thicknesses

Burwicz et al (2011) have calculated GHSZ thicknesses based on the


global bathymetry, salinity, bottom water temperature, and heat flow (as a
proxy to geothermal gradients as they are not globally available).
GHSZ thicknesses can be considered a proxy for potential hydrate deposits
distribution but not necessarily for the real volume of hydrate-bearing
sediments. The formation of hydrates is mainly controlled by methane
supply either through the direct degradation of organic matter within the
GHSZ or through an upward flux of deeper biogenic and thermogenic

methane. Global estimates of methane fluxes from deep sediments are


poorly constrained.

Gas Hydrate Stability Zone

Methane
Hydrate in Permafrost Soils

Methane
Hydrate in Seafloor SedimentsThe most common type of gas hydrate is methane hydrate and

the conditions required for its stability can occur in marine sediments and in permafrost
soil. The phase diagrams redrawn from Kvenvolden and Lorenson (2001) show the physical
conditions (temperature and pressure) required for the stability of methane hydrate in the
marine environment (top) and the permafrost environment (bottom).

First, we discuss the marine setting. Salty oceanic water can be no colder than about ?
1.8C before freezing. Assume that you are in a polar region, where the sea bottom
temperature is 0C. Furthermore, assume that the average temperature increase is 3C
per 100m sediment depth. The figure then shows that methane hydrate cannot be stable
at a water depth of 100m. But it may occur in a seafloor that is 400m below sea level.
When drilling at a water depth of 400m, you can expect or hope to find a 370m thick
hydrate layer. Beneath this depth the temperatures get too high for a formation of gas
hydrate, so that free gas and water is found. For a case of 1,000m water depth, the
hydrate layer will be 600m thick. Obviously, the thickness of the hydrate zone will depend
on the temperature gradient. In sediments that display a stronger increase in temperature,
which can be the case, for example, at active continental margins (46C per 100m
depth), the hydrate zone will generally be thinner.

Next, we look at the permafrost setting, where temperature gradients are considerably
lower than in the ocean. Typically, the temperature can be expected to change by 1.3C

per 100m within the permafrost zone, and with 2C per 100m in layers below the
permafrost zone. The ambient temperature and the thickness of the frozen layer are
therefore of significant importance for the stability of gas hydrate.

Consider the case where the base of the permafrost is at a depth of 100m or less. The
figure shows that the physical conditions will not be adequate for the formation of gas
hydrate. If the permafrost base is, say, at 750m, the thickness of the gas hydrate zone is
900m.

Since the stability of gas hydrates is related to relatively low temperatures and high
pressure, any change in these two parameters can increase or decrease the stability of the
gas hydrate. For example, if either the temperature is increased or the pressure is
reduced, the gas hydrate will change phase from a solid to a gas and liquid.

Gas hydrates are not chemical compounds since the sequestered molecules are never
bonded to the lattice. The formation and decomposition of hydrates are first-order phase
transitions. However, the detailed formation and decomposition mechanisms are still not
well understood on a molecular level.

PART V: Gas Hydrates - The Resource Potential

Natural gas hydrate occurs worldwide: in oceanic sediments of continental


slopes; in deepwater sediments of inland lakes and seas; and in both
continental and continental shelf polar sediments. In oceanic sediments,
where water depths exceed about 300m and bottom water temperatures
approach 0C, gas hydrate is found at the seafloor and down to sediment

depths of about 1,100m. The typical depth range for hydrate stability lies
100500m beneath the seafloor. In polar continental regions, gas
hydrate can occur in sediments at depths ranging from 150 to 2,000m.
Occurrences of hydrates within the gas hydrate stability zone (GHSZ) are
affected by numerous additional factors, including availability of gas,
water, and geological controls. About 98% of the gas hydrates are believed
to be concentrated in oceanic sediments, while the other 2% are in polar
landmasses.

How Much Gas Hydrate Exists?

Gas hydrate resource pyramid: Gas hydrates exist in a variety of forms that pose different
opportunities and challenges for energy resource exploration and production. The left axis displays

lithology of the host sediment. The right axis shows associated estimates of natural gas resources.
Gas hydrate-bearing sands are the most feasible initial targets for energy recovery. Other
occurrences, such as gas hydrate-filled fractures in clay-dominated reservoirs, may become
potential energy production targets in the long-term future (Courtesy: Ray Boswell). Gas

volumes are often cited in units of trillion cubic feet (Tcf), and there
are approximately 35.3 cubic feet in a cubic metre. It is estimated that
resources of methane in natural hydrate reservoirs range anywhere from
10 to 2.8 x 10 Tcf, or around 2.8 x 10 to 8 x 10 m , indicating that more
5

15

15

carbon is contained in methane hydrate than in all other organic carbon


reservoirs on earth combined.
These estimates, however, include hydrate in low-grade shale deposits
as well as in high-grade sand deposits. Only a fraction of the methane
sequestered in global gas hydrate deposits is likely to be both
concentrated and accessible enough to ever be considered a potential
target for energy resource exploitation.
The relative amounts of gas hydrate in the global system can be
illustrated by the hydrate resource pyramid, which captures the
distribution of sequestered methane among the major types of global gas
hydrate deposits. Only the hydrates at the top of the pyramid a small
subset of the hydrate deposits are likely to be considered viable as a
source of commercial quantities of natural gas.

Occurrences in Muds and Coarse Silt

At the top of the pyramid lie high permeability sediments in


permafrost areas. The amount of gas hydrate in these settings globally is
relatively small, but permafrost-associated gas hydrates might be the
easiest to commercialise, particularly in areas with well-developed
infrastructure from conventional hydrocarbon production, such as the
Alaskan North Slope.
Gas hydrate resources housed in marine sand reservoirs are also

obvious major targets for any longer-term development of gas hydrates as


a resource. Highly permeable marine sands with moderate to high gas
hydrate saturations are considered the best targets for resource
development. Recent logging-while-drilling in the Gulf of Mexico
has identified geologic units with inferred hydrate saturations as high as
80%.
Reservoir quality is expected to increase with increasing grain
size. However, the primary control of importance may be intrinsic
permeability. Sediments of high intrinsic permeability may have the
capability to host hydrate at high saturations (5090% of pore space).

Occurrences in Muds and Fine Silt

Below marine sands in the gas hydrate resource pyramid is the category
for muds and fine silt. Fractured muds are less permeable, usually smallergrained sediments that may host gas hydrates in fracture-related
permeability. Drilling on the Indian and Korean margins and in the Gulf of
Mexico has found gas hydrate filling pervasive fractures within low
permeability sediments (e.g., silts and clays). Such sediments may not
have a high average saturation of gas hydrate, maybe around 20%,
but targeted production from gas hydrates within the fractures could
theoretically yield significant gas.
At the base of the resource pyramid lie gas hydrates in low
permeability, undeformed fine-grained muds. Such sediments host most of
the global gas in place in methane hydrates and are unlikely to become a
target for commercial production of gas from methane hydrates. The
saturation typically is only 5%.
Sea-floor mound deposits are small size and ephemeral. They
are environmentally sensitive due to associated unique biological
communities and thus unattractive as a resource target.

Potential Worldwide

Hydrate Energy International (HEI) recently released estimates of the gas hydrate resource
potential, utilising a petroleum systems approach (Source: Johnson, 2011). In

conventional

petroleum systems analysis, the geological components and processes


necessary to generate and store hydrocarbons are well established:
source, migration, reservoir, seal, and timing. To apply this petroleum
system model to a methane hydrate resource system, one needs also to
incorporate the parameters that determine methane hydrate stability
conditions: formation temperature and pressure, pore water salinity, water
availability, gas source, gas transport, gas concentration, and the time
over which the system evolves.
Recently, Hydrate Energy International (HEI), as part of the Global Energy
Assessment being conducted by the International Institute for Applied
Systems Analysis (IIASA), released the results of a new evaluation of the
gas hydrate resource potential, utilising a petroleum systems approach.
Their median assessment is around 43,000 Tcf.

Geological Settings of Gas Hydrate

Gas hydrates occur in a wide variety of geologic settings and modes of


occurrence. These include gas hydrate concentration, host lithology,
distribution within the sediment matrix, burial depth, water depth, and
many others. The major controlling factor on where gas hydrate forms is
lithology and availability of methane.
The HOTSPOT IMAGE (below) from Boswell (2011) gives a schematic
depiction of the components of various methane hydrate systems.
Examples A and B represent massive forms in hydrate-bearing marine
clays. Example C shows a hydrate-bearing marine sand. Examples D and E
represent sea-floor mounds (outcrops) and hydrate-bearing clays (finely
dispersed).
Three dominant types of gas hydrate accumulations can be defined and
distinguished based on the mode of fluid migration and gas hydrate
concentration within the GHSZ (Milkov and Sassen, 2002). The end-

members are structural and stratigraphic accumulations, but combination


accumulations controlled both by structures and stratigraphy may occur.

A, B

F
HOTSPOT IMAGE: A schematic depiction of the components of various
methane hydrate systems.
Typical methane hydrate reservoir morphologies include (A) networks of hydratefilled veins; (B) massive hydrate lenses; (C) grain-filling methane hydrate in
marine sands; (D) massive sea-floor mounds; (E) grain-filling methane hydrate in
marine clays; (F) grain-filling methane hydrate in onshore
arctic sands/conglomerates (Click on map for photographic examples).
The general location of the most resource-relevant (blue circles) and
most climate-relevant (green circles) methane hydrate occurrences are
also shown. Other parts of the methane hydrate system as depicted include
the relationship between microbial and thermogenic gas sources and
gas migration controls.
SOURCE: R. Boswell, 2011.

Structural Accumulations

Structural gas hydrate accumulations occur in advective high fluid flux


settings, where highly permeable fractured conduits like fault systems,
mud volcanoes and other geological structures facilitate rapid
fluid transport from depth into the GHSZ. The gas hydrate concentration in
the sediments is relatively high. Gas hydrate deposits associated with
active faults and craters of deepwater mud volcanoes usually present high
gas hydrate concentrations, with 3050% of the pore space filled by
hydrates.

The shallow seafloor consists typically of non-consolidated silts and clays.


Various types of gas hydrates may occur: layers of hydrates of
thicknesses from millimetres to tens of centimetres, massive hydrate
deposits, or hydrate outcrops (mounds) on the seafloor.
Bottom-simulating reflectors (BSRs) are not common in
structural accumulations as they do not typically seal much gas below the
gas hydrate layer. If present, they are patchy and displaced and they do
not parallel the seafloor.

Stratigraphic Accumulations

Stratigraphic gas hydrate accumulations generally occur in advective low


fluid flux settings within passive margins in relatively coarse-grained
sediments, from biogenic methane gas generated in situ, or gas which is
slowly supplied from deeper in the subsurface.
In stratigraphic accumulations, gas hydrate tends to be highly
dispersed through the GHSZ, and low hydrate concentrations are
commonly measured; 112% of the pore space is filled by hydrates. The
low hydrate concentration can be explained by the low permeability and
porosity in clay-rich sediments, which hinder the mobility of both water
and gas, necessary for hydrate formation. Most of the hydrate in claydominated sediments is present in a network of tiny fractures.
However, there are significant exceptions. Both the lithostatic
pressure (depth) and the sediment type influence how the gas hydrate will
occupy the sediment pore space. Deeper in the sediment column below
the seafloor, the hydrate cannot overcome the lithostatic pressure
between the sediment grains and must reside in the pore space or
in fractures. For coarse-grained sediments, such as sands, hydrates can
become highly saturated.

A well-known example is the Nankai Trough, where gas hydrate occupies


up to 82% of pores in thin but very permeable sand units. The Nankai
Trough is located beneath the Pacific Ocean off the south-east coast of
Japan, and is known as an active subduction and earthquake zone.
BSRs are commonly observed on the eastern Nankai margin. This is to
date the only place where a successful gas hydrate production test has
been performed.

Seismic attribute co-blend map (RMS amplitude/coherence) showing sand channels in excess of 150m thick. The
bright yellow and orange colours highlight zones with high seismic amplitudes characteristic of sand channels.
The displayed interval shows several generations of sand deposits within the gas hydrate stability zone. If
charged with gas they could form prospective targets for gas hydrate exploration. (Reichel and Gallagher, 2014)

Mapping of Gas Hydrate

Interpretation of seismic data provides the most important means for


mapping and characterising the distribution of gas hydrates and possible
underlying free gas. Shallow high-amplitude events can be generated by
features other than gas hydrate for example, carbonate cemented zones,
layered clays, the bases of mass transport complexes, and unconformities.
However, the presence of seismic bottom-simulating reflectors is the most
common indicator of the presence of gas hydrate. The BSR is often a
strong, coherent reflector that lies at the base of the gas hydrate stability
zone and is overlain by sediments containing gas hydrate and underlain by
sediments containing free gas. The BSR has negative reflection amplitude
caused by the difference in elastic impedance. The base of the free-gas
zone is rarely evident in the seismic section and the concentration of free
gas is thought to decrease gradually downward to water-saturated
sediments. In rare cases we may observe a flat spot in exceptional cases,
as in the illustration below, even two beneath the BSR.

Seismic example of marine gas hydrates above a double flat spot. The contrast between the high-velocity hydrate-bearing strata and the

In this example, the gas hydrates act as seals for underlying hydrocarbon reservoirs (Courtesy: Statoil, Sonangol E.P. and Schlumberger


Seismic feature enhancement of BSR and two underlying flat spots (Courtesy: Statoil, Sonangol E.P. and Schlumberger Multiclient).

You might also like