Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Excess pore pressure resulting from methane hydrate dissociation

in marine sediments: A theoretical approach


Wenyue Xu
School of Earth and Atmospheric Sciences, Georgia Institute of Technology, Atlanta, Georgia, USA
Leonid N. Germanovich
School of Civil and Environmental Engineering, Georgia Institute of Technology, Atlanta, Georgia, USA
Received 26 December 2004; revised 14 September 2005; accepted 12 October 2005; published 27 January 2006.
[1] This study quantifies the excess pore pressure resulting from gas hydrate dissociation
in marine sediments. The excess pore pressure in confined pore spaces can be up to
several tens of megapascals due to the tendency for volume expansion associated with gas
hydrate dissociation. On the other hand, the magnitude of excess pore pressure in well-
connected sediment pores is generally smaller, depending primarily on the
hydrate dissociation rate and the sediment permeability. Volume expansion due to gas
hydrate dissociation in well-connected pore spaces is related via Darcys law to an increase
in pore pressure and its gradient in sediment, which drives an additional upward fluid
flow through the sediment layer overlying the gas hydrate dissociation area. The
magnitude of this excess pore pressure is found to be proportional to the rate of gas
hydrate dissociation and the depth below seafloor and inversely proportional to sediment
permeability and the depth below sea level. The excess pore pressure is the greatest at low
initial pressures and decreases rapidly with increasing initial pressure. Excess pore
pressure may be the result of gas hydrate dissociation due to continuous sedimentation,
tectonic uplift, sea level fall, heating or inhibitor injection. The excess pore pressure is
found to be potentially able (1) to facilitate or trigger submarine landslides in shallow
water environments, (2) to result in the formation of vertical columns of focused
fluid flow and gas migration, and (3) to cause the failure of a sediment layer confined by
low-permeability barriers in relatively deep water environments.
Citation: Xu, W., and L. N. Germanovich (2006), Excess pore pressure resulting from methane hydrate dissociation in marine
sediments: A theoretical approach, J. Geophys. Res., 111, B01104, doi:10.1029/2004JB003600.
1. Introduction
[2] It has long been suggested that gas hydrate dissocia-
tion in marine sediment may lead to excess pore pressure
resulting in sediment deformation or failure, such as sub-
marine landslides [Field, 1990; Kayen, 1988; Kayen and
Lee, 1991; McIver, 1982; Mienert et al., 2005; Paull et al.,
1996; Rothwell et al., 1998], sediment slumping [Dillon et
al., 2001; Vogt et al., 1994], pockmarks and mud volcanoes
[Van Rensbergen et al., 2002; Vogt et al., 1994, 1999;
Zuhlsdorff and Spiess, 2004], soft-sediment deformation
[Kennett and Fackler-Adams, 2000] and giant hummocks
[Davies et al., 1999]. Rapid release of a large amount of
methane from dissociating submarine gas hydrates to the
atmosphere may drastically impact the global climate
[Dickens et al., 1995; Kennett et al., 2003; MacDonald,
1990; Nisbet, 1990]. However, it is shown [Xu et al., 2001]
that without invoking additional mechanisms, an enhanced
transport of methane through marine sediment due to gas
hydrate dissociation resulting from a falling sea level or an
increase in seafloor temperature is probably not sufficient
to cause the drastic increase in greenhouse gas released from
dissociating gas hydrates suggested by Dickens and cow-
orkers [Dickens et al., 1997a, 1995]. On the other hand, the
creation or reactivation of fault zones and fluid flow channels
or other types of sediment deformation and failure caused by
excess pore pressure as the result of gas hydrate dissociation
may provide pathways that are highly efficient in transporting
gas [Paull et al., 2003; Zuhlsdorff and Spiess, 2004]. A
careful quantification of expected excess pore pressure levels
related to various environmental changes is needed.
[3] Efforts to quantify excess pore pressures related to gas
hydrates in marine sediments have been made using several
different approaches. Flemings et al. [2003] tried to estimate
the magnitude of excess pore pressure beneath the gas
hydrate layer at Ocean Drilling Program Site 997, Blake
Ridge offshore North Carolina. They used an empirical
interpretation of measured porosity data. Therefore it is
not guaranteed that the dynamic feedback between sediment
compaction and pore pressure is adequately accounted for.
[4] Hornbach et al. [2004] suggested that a critically
pressured layer of free gas is present underneath most gas
hydrate provinces. Aside from the general lack of evidence
JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 111, B01104, doi:10.1029/2004JB003600, 2006
Copyright 2006 by the American Geophysical Union.
0148-0227/06/2004JB003600$09.00
B01104 1 of 12
supporting their suggestion, how such a critical pressure
may be maintained is questionable. First, the volume
fraction of gas bubbles in sediment pore space is estimated
to be less than 5% in most cases with a higher value of up to
12% at Blake Ridge [e.g., Dickens et al., 1997b; Holbrook
et al., 1996; Singh and Minshull, 1994; Yuan et al., 1996].
At these volume fractions, the formation of an
interconnected gas column can still be difficult because of
the capillary effect [Henry et al., 1999]. Moreover, the fact
that bottom-simulating reflectors (BSR) locations of many
gas hydrate provinces are consistent with the gas hydrate
stability boundary predicted by assuming a normal temper-
ature gradient and hydrostatic pressure is in contradiction
with the suggested critical pressure at the top of such a gas
column.
[5] Sultan et al. [2004] are among the first efforts to more
accurately quantifying the excess pore pressure related to
gas hydrate dissociation. However, their work considers
only the maximum excess pore pressure when gas hydrate
dissociation occurs in confined pore spaces. It does not
distinguish gas hydrate dissolution from dissociation and
inappropriately uses experimental data of CO
2
hydrate
dissolution as a proxy for methane hydrate dissolution.
These simplifications lead to the conclusion that excess pore
pressure can result from methane hydrate dissolution near
the top of the gas hydrate layer and contribute to the large-
scale Storegga submarine landslide offshore mid-Norway.
[6] Here we address this problem by first quantifying the
volume expansion associated with gas hydrate dissociation.
The resultant excess pore pressure is then calculated either
based on the amount of dissociated gas hydrate in a
confined pore space or according to the rate of gas hydrate
dissociation in the interconnected pore space of an incom-
pressible sediment by relating dissociation-released fluids to
an enhanced upward fluid flow away from the horizon
where gas hydrate dissociation takes place. The cases of
sea level fall, tectonic uplift, heating and inhibitor injection
in a horizontal layer of gas hydrate-bearing sediment are
also analyzed to obtain the associated rate of gas hydrate
dissociation. Finally, implications of the excess pore pres-
sure caused by gas hydrate dissociation to marine sediment
deformation or failure are discussed.
[7] Previous studies [Clennell et al., 1999; Xu and
Ruppel, 1999; Zatsepina and Buffett, 1998] suggest that a
three-phase zone in which gas hydrate, water and free gas
coexist may occur in marine gas hydrate systems. Further
investigations [Xu, 2002, 2004] indicate that at conditions
close to thermodynamic equilibrium, gas hydrate dissocia-
tion takes place within the three-phase zone until all gas
hydrates are dissociated. This equilibrium can be dynamic
in the sense that the equilibrium conditions may change
with time. This idea is fundamentally different from earlier
concepts of gas hydrate dissociation along a three-phase
interface separating an overlying zone of coexisting water
and gas hydrate from an underlying zone of coexisting
water and free gas. Unless explicitly stated otherwise, gas
hydrate dissociation and related processes considered
through out this study are assumed to occur at local
conditions close to thermodynamic equilibrium. All calcu-
lations of phase equilibrium and thermodynamic properties
of free gas, liquid solution and gas hydrate are done
according to the methods described by [Xu, 2002, 2004;
Xu and Ruppel, 1999]. All calculations are done for
structure I methane hydrate, the dominant component in
natural gas hydrate systems.
[8] For simplicity, calculations are done assuming a
constant salinity of 3.5% and effects of a varying salinity
on phase equilibrium and transition are neglected. In gen-
eral, salinity of the liquid phase decreases as gas hydrates
dissociate. This change raises the equilibrium temperature
and, consequently, stabilizes gas hydrates. It can also lower
the solubility of methane in liquid water. Dissociation of a
small amount, say less then 10% of pore volume, of gas
hydrates in a mixture of gas hydrate and liquid solution
results in a rather small, in the order of 10% or less, change
in liquid salinity and thus has little effect on the processes
studied in this work. Parameters and properties used in this
study are listed in Table 1.
2. Volume Expansion Due to Gas Hydrate
Dissociation
[9] To avoid misconceptions, we preface our gas hydrate
dissociation discussion whit a clarification of the distinction
between gas hydrate dissolution and dissociation. The
former can take place within the whole gas hydrate stability
zone as gas solubility of the coexistent liquid phase
increases. In general, no free gas is produced during the
course of gas hydrate dissolution and, since the density of
methane hydrates is lower than that of the coexistent liquid
water-gas solution, gas hydrate dissolution does not result in
excess pore pressure in natural environments. Sultan et al.
[2004] suggest that gas hydrate dissolution near the top of
the gas hydrate occurrence zone might have caused consid-
erable excess pore pressure and thus contributed to the
occurrence of the Storegga Slide offshore mid-Norway.
Their speculation is based on laboratory experiments done
on CO
2
hydrate [e.g., Zhang, 2003], which has a higher
density than the coexistent liquid water-gas solution, and
may not be applicable to natural environments where
methane is the most abundant gas component of gas
hydrates. In contrast, gas hydrate dissociation takes place
near its stability boundary, produces water plus free gas and
therefore results in excess pore pressure due to the volume
expansion associated with the dissociation. Though there
are certain exceptions [Xu, 2004], in natural environments,
gas hydrate dissociation occurs much more frequently near
the base of a gas hydrate stability zone within a dynamically
evolving natural gas hydrate system in a changing environ-
ment. Here we analyze this volume change resulting from
gas hydrate dissociation.
[10] Suppose that a mixture of liquid water, free gas and
gas hydrate resides in pore volume V
p
, which is
V
p
fADD 1
for a porous sediment column of cross sectional area A,
length DD and porosity f. The volume of gas hydrates
inside the pore space is
V
h
V
p
S
h
; 2
where S
h
is the volume fraction of gas hydrate averaged
over V
p
. The total volume change dV resulting from the
B01104 XU AND GERMANOVICH: HYDRATE DISSOCIATION AND PORE PRESSURE
2 of 12
B01104
dissociation of a small dV
h
gas hydrates, which releases dV
w
water and dV
g
free gas, includes the volume change due to
the fact that the densities of released water and free gas are
different from that of the dissociated gas hydrate and the
volume change due to the compressibility of extant free gas,
liquid and gas hydrate.
[11] The first volume change, (dV)
1
, related to density
differences is
dV
1
dV
w
dV
g
dV
h
: 3
Compared to that of gas released by dissociating gas
hydrate, the amount of gas dissolved in the released water is
small. As long as the gas solubility of water does not change
rapidly, the dissolved gas to the volume change is usually
negligible. Therefore, for a fixed gas mass fraction r
g
of the
gas hydrate,
dV
g
dM
g
=r
g
r
g
dM
h
=r
g
r
g
r
h
=r
g
_ _
dV
h
4
dV
w
dM
w
=r
w
1 r
g
_ _
dM
h
=r
w
1 r
g
_ _
r
h
=r
w
dV
h
;
5
where r
w
, r
g
, r
h
and M
w
, M
g
, M
h
denote the densities and the
masses of the liquid water, the free gas and the gas hydrate,
respectively. Thus the volume change relative to pore
volume V
p
is
dV
1
V
p
R
v
dV
h
V
p
R
v
dS
h
; 6
where
R
v
1 r
g
_ _
r
h
=r
w
r
g
r
h
=r
g
1 7
is the factor of volume expansion resulting from the gas
hydrate dissociation.
[12] Variations in r
w
, r
h
and r
g
are usually negligible. R
v
is
only weakly dependent on temperature, controlled instead
by density changes of the gas, which is usually highly
compressible. Gas hydrate dissociation leads to volume
expansion when R
v
> 0, whereas a volume contraction
results if R
v
< 0. This fact is depicted in Figure 1 for
structure I methane hydrate, which is calculated using
Setzmann and Wagners [1991] equation of state and shows
a dramatic increase in the tendency of expansion as pressure
Table 1. Parameters and Properties Used in This Study
Parameter Definition
A section area of sediment column (m
2
)
A
s
total area of dissociation interfaces within a sediment column with a unit
section area (m
2
)
c heat capacity of hydrate-bearing sediment (J kg
1
K
1
)
c
w
, c
h
, c
g
, c
s
specific heat of liquid water, gas hydrate, free gas or sediment (J kg
1
K
1
)
D depth to the layer of gas hydrate dissociation below seafloor (m)
DD thickness of the layer of gas hydrate dissociation (m)
g gravitational acceleration (m s
2
)
DH
dis
enthalpy of gas hydrate dissociation (4.3 10
5
J kg
1
)
k permeability (m
2
)
M
w
, M
h
, M
g
mass of liquid water, gas hydrate or free gas (kg)
P pressure (Pa)
P
ex
excess pore pressure (Pa)
P
ex
0
excess pore pressure t = t
0
(Pa)
P
exmax
maximum excess pore pressure (Pa)
P
0
initial pore pressure (Pa)
Q
e
rate of heating (W m
2
)
Dq
f
change in mass flux of fluid flow (kg m
2
s
1
)
R rate coefficient of gas hydrate dissociation (2.227 10
4
kg (m
2
K
2
s)
1
)
r
g
mass fraction of gas in hydrate form, 0.1292 for methane hydrate with
100% filling of the hydrate cages
R
v
volume expansion factor
S
w
, S
h
, S
g
volume fraction of liquid water, gas hydrate or free gas in pores
S
w
+ S
h
+ S
g
= 1
t time (s)
T temperature (K)
T
0
temperature at t = 0 (K)
T
0
temperature at t = t
0
(K)
T
e
temperature along the stability boundary (K)
T
e0
temperature along the stability boundary at t = 0 (K)
V
w
, V
h
, V
g
, V volume of liquid water, gas hydrate, free gas or their mixture (m
3
)
V = V
w
+ V
h
+ V
g
v
sd
rate of sea level drop or tectonic uplift (m s
1
)
V
p
= fADD volume of pore space (m
3
)
z space (z)
a ratio of the vertical extent of resultant hydrate dissociation zone to the
magnitude of sea level fall or tectonic uplift
f sediment porosity
k effective volume compressibility of the mixture of liquid water, gas
hydrate and free gas (Pa
1
)
l effective thermal conductivity of sediment column (W m
1
K
1
)
m
f
viscosity of fluids (Pa s)
r
f
, r
w
, r
h
, r
g
, r
s
density of fluids, liquid water, gas hydrate, free gas or sediment (kg m
3
)
B01104 XU AND GERMANOVICH: HYDRATE DISSOCIATION AND PORE PRESSURE
3 of 12
B01104
decreases. For example, the dissociation of methane hydrate
equilibrating with pure water or seawater at a pressure near
34 MPa results in a volume expansion of 30% compared
to a 210% volume expansion if the dissociation takes
place at 6.5 MPa, and 500% for dissociation at 2.5 MPa.
Because of higher pressures in marine sediment, the volume
expansion due to gas hydrate dissociation in marine sedi-
ment in general is much smaller than that at standard
pressure and temperature conditions. Therefore referring
to a much larger volume expansion at standard conditions
when discussing the dissociation of gas hydrates in marine
sediments can be misleading.
[13] The second part of the total volume change is related
to the compressibility of existing free gas, gas hydrate and
liquid solution. When volume compression resulting from
the increasing pore pressure is considered, the magnitude of
volume expansion becomes even smaller than what would
be when pore pressure is kept constant. Assuming thermo-
dynamic equilibrium, the effective compressibility, k, of the
mixture of free gas, gas hydrate and liquid solution along
the gas hydrate stability P-T boundary may be calculated as
k
1
V
@V
@P

@V
@T
dT
e
dP
_ _

S
g
r
g
@r
g
@P

@r
g
@T
dT
e
dP
_ _

S
w
r
w
@r
w
@P

@r
w
@T
dT
e
dP
_ _

S
h
r
h
@r
h
@P

@r
h
@T
dT
e
dP
_ _
; 8
where T
e
is the stability temperature of structure I methane
hydrate corresponding to the given pressure and S
g
, S
h
and
S
w
are the pore space volume fractions of free gas, gas
hydrate and liquid solution, respectively. Compared to that
of free gas, the compressibility of liquid solution or gas
hydrate can usually be neglected. However, the latter
becomes significant when the volume fraction of free gas is
small. The effective compressibility is larger when the
pressure is lower and the volume fraction of free gas is
higher (Figure 2). When there is no pressure change other
than the excess pore pressure P
ex
resulting from gas hydrate
dissociation, the relevant volume change is
dV
2
V
p
kdP
ex
: 9
Therefore the total volume change is
dV
V
p

dV
1
V
p

dV
2
V
p
R
v
dS
h
kdP
ex
: 10
Note that dV 6 0 when the pore space is not completely
closed.
3. Excess Pore Pressure Resulting From Gas
Hydrate Dissociation
3.1. In Confined Pore Space
[14] When host sediment permeability is sufficiently low
or gas hydrates dissociate rapidly, the dissociation process
may be treated as taking place in a constant pore volume.
This may lead to a large excess pore pressure because fluids
released by dissociating gas hydrates do not have time to
escape the pore space. Since the pore space is confined, the
total volume change is zero and equation (10) becomes
R
v
dS
h
kdP
ex
0: 11
The magnitude of excess pore pressure can be calculated
by integrating equation (11) over the whole dissociation
process, which gives
P
ex

_
DSh
0
R
v
k
dS
h
; 12
Figure 1. Volume expansion factor R
v
calculated as a
function of pressure and temperature. The stability bound-
aries of methane hydrate in pure water and seawater
environments are also shown. Assuming thermodynamic
equilibrium, the dissociation of gas hydrate takes place
along the stability boundary.
Figure 2. Effective compressibility of the mixture of
liquid, free gas and gas hydrate along the stability boundary.
Calculations are for fixed 20% volume fraction of gas
hydrate. The sum of volume fractions of liquid and free gas
phases is 80%.
B01104 XU AND GERMANOVICH: HYDRATE DISSOCIATION AND PORE PRESSURE
4 of 12
B01104
where DS
h
is the change in volume fraction of gas hydrate
in pore space, which is a negative number for gas hydrate
dissociation. The excess pore pressure, as calculated using
equation (12), is plotted in Figure 3 as a function of DS
h
and the initial pore pressure before gas hydrate dissocia-
tion starts. The excess pore pressure can be tens of
megapascals even when only a small amount of gas
hydrate dissociates.
3.2. In Interconnected Pore Space
[15] In the following studies, fluid flow is assumed to
take place in rigid or incompressible porous sediments and
obey Darcys law with a relatively constant effective per-
meability. To avoid the possibility of nonlinear fluid flow
behavior as pointed out by Clennell et al. [2000], the
following analyses do not apply to conditions close to
capillary leakage pressures or pressures sufficiently high
to cause sediment fracturing. In addition, downward flow is
assumed negligible compared to upward fluid flow associ-
ated with excess pore pressure resulting from gas hydrate
dissociation. This is usually true when fluids outside of the
dissociation layer are nearly incompressible and because the
seafloor is the only place for fluid discharge. The downward
flow may be significant under certain circumstances, par-
ticularly when an underlain sediment layer containing large
amount of free gas is present, as can occur at some BSRs.
Then the compressibility of the free gas contained in
sediment under the dissociation layer will need to be
accounted for.
[16] Considering a vertical one-dimensional system with
a closed lower end, the volume increase resulting from the
dissociating gas hydrates is mostly accounted for by an
increase in the rate of upward fluid flow Dq
f
away from the
gas hydrate dissociation area
ADq
f
r
f
dV
dt
r
f
fADD R
v
dS
h
dt
k
dP
ex
dt
_ _
; 13
where r
f
is the density of the upward flowing fluid. The
flow rate increase can be directly related to a corresponding
pressure gradient increase D(@P/@z) across the overlain
sediment layer according to Darcys law
ADq
f

Akr
f
m
f
D
@P
@z
_ _

Akr
f
P
ex
m
f
D
; 14
where k denotes the permeability of the overlain sediment,
m
f
is the fluid viscosity and D is the depth of the top of the
dissociation area below the seafloor. Note that the
simplification for the right-hand side of equation (14) is
made based on the fact that in most circumstances, pressure
equilibrium reestablishes itself in accordance to any
introduced change in a relatively short time period. See
Figures 5 and 6 of Xu [2004] for examples. The pressure
reestablishment process may be slower for very low
permeability sediments. In that case, gas hydrate dissocia-
tion and pore pressure increase can be viewed approxi-
mately as occurring in confined space.
[17] Combining equations (13) and (14) yields
kfDD
dP
ex
dt

k
Dm
f
P
ex
R
v
fDD
dS
h
dt
; 15
which can be rewritten as
dP
ex
dt
aP
ex
b 0;
a
k
km
f
DfDD
;
b
R
v
k
dS
h
dt
:
16
By assuming a constant dissociation rate, the solution of
equation (16) is
P
ex

b
a
1 e
at
: 17
Solution (17) indicates that P
ex
increases with a character-
istic time 1/a and its maximum is
P
ex max

b
a

m
f
R
v
DfDD
k
dS
h
dt
: 18
[18] The magnitude of excess pore pressure P
ex
resulting
from gas hydrate dissociation is proportional to the factor of
volume increase R
v
, the rate of dissociation fDDdS
h
/dt and
the depth below seafloor D, and inversely proportional to
the permeability k. Figure 4 plots the excess pore pressure
P
exmax
as a function of initial pore pressure and a grouped
parameter (fDDDdS
h
/dt) calculated for k = 10
16
m
2
and
m
f
= 10
3
kg m
1
s
1
. Effect of permeability on P
ex max
is
the most significant (Figures 5 and 6). Unless explicitly
specified otherwise, these values of k and m
f
for Figure 4 are
used for the calculations thereafter. Note that r
f
and m
f
are in
general the density and viscosity, respectively, of a fluid
mixture of free gas and liquid solution. However, if the
volume fraction of free gas is sufficiently small, say a few
Figure 3. Excess pore pressure caused by dissociation of
gas hydrates (initially 20% of pore space) is in confined,
initially gas-free pore space plotted as a function of the
initial pore pressure and the amount of dissociated gas
hydrate expressed as the volume fraction of pore space.
B01104 XU AND GERMANOVICH: HYDRATE DISSOCIATION AND PORE PRESSURE
5 of 12
B01104
percent of the pore space, which is likely true within deeper
parts of most natural gas hydrate systems, then it is probably
safe to assume that free gas released from the dissociating
gas hydrates tends to stay in the pores as individual gas
bubbles [Henry et al., 1999] and the only mobile phase is
liquid solution. Consequently, r
f
and m
f
can be approximated
as those of liquid water, which do not vary much within
natural gas hydrate systems. In this case, only the methane
dissolved in liquid water is transported into the overlain
sediments. Since methane hydrate formation via this type of
processes is usually slow [Rempel and Buffett, 1997; Xu and
Ruppel, 1999] compared to the timescales of the processes
considered here, the effect of methane hydrate formation in
overlain sediments may be neglected.
[19] If either a or b varies with time, then equation (16)
usually does not have a simple analytical solution and the
excess pore pressure at t = t
0
+ Dt may be approximately
calculated for a sufficiently small time step Dt
P
ex

2 aDt P
0
ex
2bDt
2 aDt
; 19
where P
ex
0
is the excess pore pressure at time t = t
0
. Solution
(19) reveals that P
ex
decays with time when the dissociation
rate decreases so that b becomes small or even reduces to
zero after the dissociation ends.
[20] Above analyses are done for excess pore pressure
resulting from a given rate of gas hydrate dissociation. The
rate of dissociation is further investigated for gas hydrate
dissociation in marine sediments due to various geological
or operational processes.
4. Examples of the Excess Pore Pressure
4.1. Due to Continuous Sedimentation At Seafloor
[21] Continuous sedimentation at the seafloor causes gas
hydrate near the base of hydrate stability zone to move
downward with the host sediment and start dissociating
when the stability boundary is reached. We want to know
the magnitude of excess pore pressure caused by hydrate
dissociation after the whole process has reached a steady
state. Assuming that DS
h
amount of gas hydrates dissociates
completely in time period Dt while getting buried an
additional depth DD, which is the thickness of the dissoci-
ation layer with an average gas hydrate volume fraction of
approximately DS
h
/2, and according to equation (18), the
excess pore pressure is
P
ex

m
f
R
v
Dfv
s
DS
h
k
; 20
where v
s
= DD/Dt is the rate of sedimentation. For example,
at a sedimentation rate of v
s
= 1 m kyr
1
, f = 0.5, k =
Figure 4. Excess pore pressure caused by dissociation of
gas hydrates (initially 20% of pore space) sediments with
interconnected pores. (left) Maximum excess pore pressure
plotted as a function of the initial pore pressure and a
grouped parameter including the rate of dissociation. (right)
Time-dependent excess pore pressure calculated for initial
pore pressure of 20 MPa, (fDDDdS
h
/dt) = 2 10
8
m s
1
,
and dS
h
/dt = 4 10
10
s
1
is also plotted. Calculations are
done for permeability k = 10
16
m
2
.
Figure 5. Same as Figure 4, but for k = 10
17
m
2
.
Figure 6. Same as Figure 4, but for k = 10
18
m
2
.
B01104 XU AND GERMANOVICH: HYDRATE DISSOCIATION AND PORE PRESSURE
6 of 12
B01104
10
16
m
2
, D = 100 m and DS
h
= 1, (fDDDdS
h
/dt) <
10
9
m
2
s
1
, and P
ex
can be found approximately on
Figure 4 to be <0.01 MPa, except for shallow water
depths.
4.2. Due to Sea Level Dropping or Tectonic Uplifting
[22] In cases of sea level fall or tectonic uplift, the pore
pressure in sediment in general decreases and the excess
pore pressure is thus calculated relative to the decreasing
pore pressure as if no gas hydrate dissociation is involved
P
ex
P P
0
r
f
gDD
_ _
; 21
where P
0
is the pore pressure before the sea level dropping
or tectonic uplifting at t = 0, DD is the magnitude of the sea
level dropping or tectonic uplifting at time t and g is the
gravitational acceleration.
[23] The dissociation of gas hydrates near the base of the
hydrate stability zone takes certain amount of heat. Except
for situations where fluid flow is relatively focused and fast,
heat transport within natural gas hydrate systems is usually
dominated by thermal conduction. Consequently, the
amount of heat consumed by gas hydrate dissociation is
supplied via the cooling of the dissociation area and thermal
conduction to the dissociation front. As the dissociation
region cools, heat flow from surrounding sediments pro-
vides the heat consumed by further hydrate dissociation.
This heat balance can be described as
r
h
DH
dis
faDD
dS
h
dt
l
DT
aDD
caDD
dT
dt
; 22
where DH
dis
is the heat of gas hydrate dissociation, l and c
are the effective thermal conductivity and the effective heat
capacity, respectively, both consisting of contributions from
the individual phases and the host sediment, averaged over
the dissociation layer. For example, the c can be expressed
as
c f S
h
r
h
c
h
S
g
r
g
c
g
S
l
r
l
c
l
_ _
1 f r
s
c
s
; 23
where r
s
and c
s
are the density and the specific heat,
respectively, of the sediment host, and r
h
, c
h
, S
h
, r
g
, c
g
, S
g
,
and r
l
, c
l
and S
l
are the density, the specific heat and the
volume fraction of gas hydrate, free gas and liquid solution,
respectively. The thickness of the dissociation layer is aDD,
where a can be calculated in accordance with the pressure
and temperature regimes of the dissociation area. Inserting
equation (15) into equation (22) results in
r
h
DH
dis
R
v
kfaDD
dP
ex
dt

k
Dm
f
P
ex
_ _
l
DT
aDD
caDD
dT
dt
:
24
DD is proportional to the time since the geologic event
began
DD v
sd
t; 25
where v
sd
is the rate of sea level drop or tectonic uplift.
Since the phase transition occurs along the phase boundary,
the changes in temperature can be expressed approximately
as
DT
dT
e
dP
P P
0

dT
e
dP
P
ex
r
f
gDD
_ _
dT
dt

dT
e
dP
dP
dt

dT
e
dP
dP
ex
dt
r
f
gv
sd
_ _
:
26
Substituting equations (26) into equation (24) gives
dP
ex
dt

C
1
t

C
2
t
2
_ _
P
ex
C
3

C
4
t
_ _
0;
C
1
r
h
DH
dis
k
_
aDm
f
v
sd
r
h
DH
dis
kf R
v
c
dT
e
dP
_ _ _ _
;
C
2
R
v
l
dT
e
dP
_
a
2
v
2
sd
r
h
DH
dis
kf R
v
c
dT
e
dP
_ _ _ _
;
C
3
R
v
r
f
gv
sd
c
dT
e
dP
_
r
h
DH
dis
kf R
v
c
dT
e
dP
_ _
;
C
4
R
v
lr
f
g
dT
e
dP
_
av
sd
r
h
DH
dis
kf R
v
c
dT
e
dP
_ _ _ _
:
27
The solution to equation (27) is
P
ex

_
t
0
C
3

C
4
x
_ _
x
t
_ _
C1
exp
C
2
t

C
2
x
_ _
dx: 28
This excess pore pressure is plotted as a function of time
and initial pore pressure in Figure 7. Its magnitude is
negligible when the initial pressure and hence the water
depth is large compared to that in shallower water
environments. In the context of paleoclimate, a sea level
fall is usually accompanied with a cooling at seafloor.
However, such a cooling would tend to stabilize the
methane hydrate at a different timescale. The characteristic
timescale of heat conduction is D
2
/l/(rc), with l/(rc)
10
6
m
2
s
1
for typical marine sediments. For instance,
if the undisturbed base of methane hydrate stability zone is
D 100 m below the seafloor, a thermal signal would need
300 years to have a considerable effect on hydrate
stability at that depth. Therefore, when the base of methane
hydrate stability is located at depths on the order of hundreds
of meters or greater, the excess pore pressure due to the sea
level fall would stabilize well before the cooling signal
would reach the same depth. On the other hand, for a much
shallower depth of the base of methane hydrate stability, the
thermal effect can be significant. A quantitative analysis of
the effect of a seafloor cooling or warming is mathematically
more complicated and is not dealt with in this study.
4.3. Caused by Heating
[24] Heating is one of the primary methods proposed for
extracting natural gas from natural gas hydrate reservoirs.
The basic idea is to induce gas hydrate dissociation by
heating and collect the natural gas released by the dissoci-
ating gas hydrates. To a certain degree, gas hydrate disso-
ciation caused by a rapid increase in seafloor temperature
may also be similarly considered. However, it requires a
much more complicated analysis and is beyond the scope of
this study.
B01104 XU AND GERMANOVICH: HYDRATE DISSOCIATION AND PORE PRESSURE
7 of 12
B01104
[25] For simplicity, here we consider a horizontal heating
plane within or immediately below a gas hydrate layer such
that the whole system can be viewed as one-dimensional
vertically. A heating rate of Q
e
is provided to dissociate gas
hydrates in a layer of DD thick while heating up the
hydrate-bearing sediment within and near the dissociation
area, namely,
Q
e
cDD
dT
dt

lDT
DD
r
h
DH
dis
fDD
dS
h
dt
: 29
Assuming thermodynamic equilibrium and neglecting
salinity changes, temperature T is a function of pore
pressure within the hydrate dissociation area. Therefore
Q
e
cDD
dT
e
dP
dP
ex
dt

l
DD
dT
e
dP
P
ex
r
h
DH
dis
fDD
dS
h
dt
: 30
Combining equation (30) with equation (15) leads to a first-
order ordinary differential equation of P
ex
in the same form
of equation (16) but with differing coefficients
dP
ex
dt
a
H
P
ex
b
H
0;
a
H

1
DD
R
v
l
DD
dT
e
dP

kr
h
DH
dis
Dm
f
_ __
R
v
c
dT
e
dP
fkr
h
DH
dis
_ _
;
b
H

Q
e
R
v
DD
_
R
v
c
dT
e
dP
fkr
h
DH
dis
_ _
:
31
The solution of equation (31) is of the same form of solution
(17) and the maximum excess pore pressure is
P
ex max

b
H
a
H
Q
e
_
l
DD
dT
e
dP

kr
h
DH
dis
DR
v
m
f
_ _
: 32
[26] Figure 8 plots the maximum excess pore pressure
estimated using equation (32) as a function of the heating
rate Q
e
. For an order of magnitude estimation, using k
10
16
m
2
, R
v
1, D 10
2
m, l 1 W m
1
K
1
, DD 10
m, c 10
6
J (m
3
K)
1
and dT
e
/dP 1 K MPa
1
, the
characteristic time 1/a
H
is 10
6
s. Therefore a couple of
weeks after the start of heating, the increase in pore pressure
slows down as the excess pore pressure approaching its final
maximum value. For relatively high permeabilities (k >
10
16
m
2
), the second term of the denominator in equation
(32) is usually larger than the first term, and therefore
P
ex max

R
v
m
f
DQ
e
kr
h
DH
dis
: 33
Compared to the maximum excess pore pressure (18), it is
apparent that the rate of gas hydrate dissociation is now
fDD
dS
h
dt

Q
e
r
h
DH
dis
: 34
Thus the magnitude of excess pore pressure may also be
estimated from solution (18) using the rate of gas hydrate
dissociation calculated according to equation (34). When
the permeability k is low (k < 10
18
m
2
) or DD is small
during the initial stage of gas hydrate dissociation, the first
term of the denominator may be larger than the second one
and hence
P
ex max
Q
e
_
l
DD
dT
e
dP
_ _
: 35
Again, this maximum excess pore pressure can be found
using solution (18), but with
fDD
dS
h
dt

Q
e
kDD
DR
v
m
f
l dT
e
=dP
: 36
4.4. Caused by Inhibitor Injection
[27] Gas hydrate inhibition is another method proposed
for introducing gas hydrate dissociation via injection of
alcohols or glycols to destabilize of gas hydrate. For
example, addition of 10 wt% methanol can lead to a
5C decrease in stability temperature of pure methane
hydrate at any given pressure, or an increase of several
megapascals in stability pressure at a given temperature
[Sloan, 1998]. Considering an inhibitor injection within a
layer of thickness DD near the base of gas hydrate stability,
as gas hydrates in contact with the inhibitor start to
dissociate spontaneously, local temperature decreases while
pressure increases as the result of dissociation. Assume that
the rate of gas hydrate dissociation may be described by
[Kamath, 1984]
r
h
fDD
dS
h
dt
rA
s
T T
e

2
r 2:227 10
4
kg= m
2
K
2
s
_ _
;
37
where A
s
is the total surface area of dissociation interfaces
between hydrates and water within the vertical column with
a unit section area and T
e
is the lowered stability
Figure 7. Excess pore pressure, as defined by equation
(21), caused by dissociation of gas hydrates (initially 20%
of pore space) resulting from a sea level dropping or a
tectonic uplifting of 30 m kyr
1
plotted (left) as a
function of the time and the initial pore pressure and
(right) as a function of time for an initial pore pressure of
20 MPa.
B01104 XU AND GERMANOVICH: HYDRATE DISSOCIATION AND PORE PRESSURE
8 of 12
B01104
temperature due to the addition of inhibitor. Although
equation (37) could be overly simplified and may or may
not be applied to the field, it is still worth to using it as an
example to demonstrate how the method developed in this
study may be applied to problems in industrial operations.
[28] Since the dissociation of gas hydrates takes place
away from the three-phase equilibrium, it is difficult to
relate the change in temperature to the excess pore pressure.
However, we can directly solve equation (15) for P
ex
,
assuming k may be calculated approximately using (8),
and equation (22) for T separately and numerically (similar
to equation (19)) using equation (37) for the rate of hydrate
dissociation. At time t = t
0
+ Dt the solutions are
P
ex

2 a
IP
Dt P
0
ex
2b
IP
Dt
2 a
IP
Dt
;
a
IP

k
km
f
DfDD
;
b
IP

R
v
rA
s
T T
e

2
r
h
kfDD
;
38
T
2 a
IT
Dt T
0
2b
IT
Dt
2 a
IT
Dt
;
a
IT

l
c DD
2
;
b
IT
a
IT
T
0

DH
dis
rA
s
T T
e

2
cDD
;
39
where T
0
and T
0
are temperatures at t = 0 and t = t
0
,
respectively. Figure 9 shows the excess pore pressure and the
temperature calculated for an initial pressure of P
0
= 20 MPa
and a T
e0
T
0
= 1C shifting of gas hydrate stability
temperature due to the injection. Initially the dissociation of
gas hydrates causes a rapid increase in pore pressure and a
rapid decrease in temperature. Consequently, the tempera-
ture decrease and an increase in gas hydrate stability
temperature resulting from the rising pore pressure lead to a
decrease in the rate of dissociation as defined by equation
(37). At some point in time, the dissociation rate becomes so
low that the pore pressure starts to decrease. Eventually, the
excess pore pressure approaches its steady state value,
which is small since the heat needed for gas hydrate
dissociation is solely supplied by a slow conductive heating
by the surrounding sediment.
[29] Heating may be applied together with inhibitor
injection to enhance the recovery of natural gas from gas
hydrates. In this case, instead of equation (22) as used
before, equation (29) is used to solve for the temperature
and the solution becomes
T
2 a
IT
Dt T
0
2b
IT
Dt
2 a
IT
Dt
;
a
IT

l
c DD
2
;
b
IT
a
IT
T
0

Q
e
DH
dis
rA
s
T T
e

2
cDD
:
40
Solution (40) reduces to equation (39) when Q
e
= 0. The
effect of an additional heating of Q
e
= 5 W/m
2
is
demonstrated in Figure 10. Because of the heating, the
initial decrease in temperature due to a rapid dissociation is
eventually reversed and the pore pressure continues to
increase until reaches its final steady state.
5. Implications to Marine Sediment
Deformation or Failure
[30] The calculations so far have shown that gas hydrate
dissociation induced by various geological or operational
Figure 8. Excess pore pressure caused by dissociation of
gas hydrates (initially 20% of pore space) resulting from
heating from a horizontal plane plotted (left) as a function of
the rate of heating and the initial pore pressure and (right) as
a function of time for a heating of 0.5 W m
2
and initial
pore pressure of 20 MPa.
Figure 9. (left) Excess pore pressure and (right) tempera-
ture variation caused by dissociation of gas hydrates
(initially 20% of pore space) resulting from inhibitor
injection plotted as a function of time. The dashed lines
are the temperatures of gas hydrate stability, and the dotted
lines are the temperatures and the excess pore pressures at
the corresponding steady states.
B01104 XU AND GERMANOVICH: HYDRATE DISSOCIATION AND PORE PRESSURE
9 of 12
B01104
processes may lead to excess pore pressure in marine
environments. The magnitude of the excess pore pressure
varies for different processes that cause the dissociation of
gas hydrate and depends on the properties of the marine
sediment, particularly the permeability (Figures 4 to 6). In
general, the excess pore pressure is high in sediments with
low permeability (lower than, say, k = 10
18
m
2
) and
reaches it maximum in the extreme case of a confined pore
space.
[31] Sufficiently elevated excess pore pressure can lead
to growth of cracks or decementation and even liquefac-
tion, which may form a weak or unstable zone within the
host sediment. Still higher excess pore pressure, such as
that in nearly confined pore spaces, may cause rapid and
extensive mechanical failure of the sediment. These effects
may be related to events of marine sediment deformation
or failure observed over geologic history. Below we
provide a brief discussion of potential geologic consequen-
ces of this type of excess pore pressure. Detailed analysis
and quantification of failure mechanisms of marine sedi-
ment related to gas hydrate dissociation are beyond the
scope of this study and will be dealt with in another
publication.
5.1. Submarine Landslides
[32] Gas hydrate dissociation tends to take place near the
base of hydrate stability zone within natural gas hydrate
systems. The excess pore pressure due to gas hydrate
dissociation can lower the effective stress and hence the
strength of sediment along the base of hydrate stability
zone. If hydrate-bearing pores are well connected and the
permeability of the host sediment is not too low (k 10
16
m
2
or higher), the magnitude of excess pore pressure is
usually moderate and by itself is unlikely to trigger, but
could facilitate, a submarine landslide. When sediment
permeability is sufficiently low (k < 10
18
m
2
) or if pores
in sediment with k > 10
18
m
2
are already overpressured to
within several megapascals from lithostatic before gas
hydrate dissociation takes place due to, say, underconsoli-
dation of the sediment, excess pore pressure resulting
from gas hydrate dissociation may be able to trigger a
landslide.
[33] Submarine landslides can reach lengths of 100 km,
with a length-to-thickness ratio as large as 1000. Puzrin
and Germanovich [2005] reasoned that gas hydrate disso-
ciation may form an initial weakened zone approximately
parallel to the seafloor extending subhorizontally from tens
of meters up to 1 km. They explained the evolution of a
landslide on submarine slope by a catastrophic shear band
propagation of this flaw. Our calculations suggest that the
growth of fractures and sediment decementation or lique-
faction resulting from the excess pore pressure due to gas
hydrate dissociation may indeed have lead to submarine
landslide events.
[34] Figures 3 to 10 indicate that the magnitude of excess
pore pressure decreases rapidly with increasing initial pore
pressure and hence the seafloor depth. Therefore this type of
sediment weakening is more relevant to marine gas hydrate
systems with a shallow seafloor depth of hundreds of
meters. The lower boundary of hydrate-bearing sediment
in these systems is usually no deeper than several tens of
meters. Consequently, this type of gas-hydrate-dissociation-
related submarine landslides should most frequently take
place in shallow water environments (less then 1000 meters
of water depth) and involve a relatively thin sediment layer
(usually 100 m or less). The thickness of the landslide
sediment layer tends to be proportional to the depth of
seafloor.
5.2. Soft Sediment Deformation and Vertical
Columns of Focused Flow and Transport
[35] As excess pore pressure builds up due to continuous
gas hydrate dissociation, the host sediment within the
dissociation area can be liquefied. In some circumstances
the required pore pressure may be significantly lower than
the lithostatic pressure. Since the horizontal stress sus-
tained by sediment is usually smaller than the vertical
stress due to overburden pressure, a partial liquefaction of
sediment occurs first when the increasing pore pressure
reduces the effective horizontal stress to zero. On the other
hand, the dissociation of gas hydrates residing in pore
spaces can also lead to an interpore shear stress because
the excess pore pressure is essentially nonuniformly distrib-
uted. This may result in localized deformation or failure of
the host sediment. Sediment that has undergone such an
internal disturbance has consequently a much lower shear
strength compared to the undisturbed sediment, and the result
can be very similar to sediment liquefaction. The soft-
sediment deformation observed by Kennett and Fackler-
Adams [2000] might be related to this process.
[36] The same excess pore pressure may also initiate
hydraulic fractures above the dissociation area [Zuhlsdorff
and Spiess, 2004] and, since the area of gas hydrate
dissociation often has a lower density than the surrounding
sediment, the liquefied sediment material tends to fill and
further pressurize the fracture. In this scenario, the pressure
decrease due to fracturing further enhances gas hydrate
dissociation and fluid supply. Furthermore, such fractures
are likely to migrate along closely spaced trajectories.
Eventually, a quasi-vertical elongated region of disturbed
sediment forms and appears on seismic profiles as a vertical
column and as a pockmark at the seafloor. The gas-charged
Figure 10. Same as Figure 9, but with an additional
heating of Q
e
= 5 W m
2
.
B01104 XU AND GERMANOVICH: HYDRATE DISSOCIATION AND PORE PRESSURE
10 of 12
B01104
soft sediment might provide a mechanism of gravitational
instability that leads to the formation of giant hummocks
[Davies et al., 1999]. Note that a temporary liquefaction of
sediment caused by traveling seismic waves is probably
irrelevant in this case since the duration of liquefaction is
probably too short.
5.3. Seafloor Collapse
[37] Compared to that in the case of well-connected
sediment pores, excess pore pressure resulting from disso-
ciation of gas hydrates in confined space can be as high as
several tens of megapascals (Figure 3). For instance, such a
large excess pore pressure may build up underneath a
horizontally extended massive gas hydrate layer, which
serves as a seal or caprock of the dissociation zone. The
dissociation of gas hydrate taking place underneath such a
layer may be viewed as occurring in a more or less confined
environment and, in certain circumstances, may result in a
very high excess pore pressure that is sufficient to overcome
the load of the overlying sediment. For instance, the range
of pressures near the BSR at the Blake Ridge collapse is
33 MPa hydrostatic to 41 MPa lithostatic. A dissocia-
tion of gas hydrates by just 1% of the pore space would
lead to a 10 MPa excess pore pressure and a pore pressure
that exceeds the lithostatic pressure.
[38] Acknowledgments. This study is funded by NSF (USA) grant
OCE-0242163. W.X. acknowledges support of the Chinese Academy of
Sciences (project KZCX3-SW-224) and NSF (China) grants 40074033
and 40274026. L.N.G.s work is also supported by NSF (USA) grant
CMC-0421090. Constructive comments by Ben Clennell, an anonymous
reviewer, and the Associate Editor William Waite helped improve the
manuscript significantly.
References
Clennell, M. B., M. Hovland, J. S. Booth, P. Henry, and W. J. Winters
(1999), Formation of natural gas hydrates in marine sediments: 1. Con-
ceptual model of gas hydrate growth conditioned by host sediment prop-
erties, J. Geophys. Res., 104, 22,98523,003.
Clennell, M. B., A. G. Judd, and M. Hovland (2000), Movement and
accumulation of methane in marine sediments: Relation to gas hydrate
systems, in Natural Gas Hydrate in Oceanic and Permafrost Environ-
ments, edited by M. D. Max, pp. 105122, Springer, New York.
Davies, R., J. Cartwright, and J. Rana (1999), Giant hummocks in deep-
water marine sediments: Evidence for large-scale differential compaction
and density inversion during early burial, Geology, 27, 907910.
Dickens, G. R., J. R. ONeil, D. K. Rea, and R. M. Owen (1995), Dis-
sociation of oceanic methane hydrates as a cause of the carbon isotope
excursion at the end of the Paleocene, Paleoceanography, 10(6), 965
971.
Dickens, G. R., M. M. Castillo, and J. C. Wlakern (1997a), A blast of gas in
the latest Paleocene: Simulating first order effect of massive dissociation
of oceanic methane hydrate, Geology, 25, 259262.
Dickens, G. R., C. Paull, and P. Wallace (1997b), Direct measurement of in
situ methane quantities in a large gas-hydrate reservoir, Nature, 385,
426428.
Dillon, W., J. Nealon, M. Taylor, M. Lee, R. Drury, and C. Anton (2001),
Seafloor collapse and methane venting associated with gas hydrate on the
Blake Ridge: Causes and implications to seafloor stability and methane
release, in Natural Gas Hydrates, Occurrence, Distribution and Detec-
tion, Geophys. Monogr. Ser., vol. 124, edited by C. K. Paull and W. P.
Dillon, pp. 211233, AGU, Washington, D. C.
Field, M. E. (1990), Submarine landslides associated with shallow seafloor
gas and gas hydrates off northern California, AAPG Bull., 74(6), 971
972.
Flemings, P. B., X. L. Liu, and W. J. Winters (2003), Critical pressure
and multiphase flow in Blake Ridge gas hydrates, Geology, 31, 1057
1060.
Henry, P., M. Thomas, and M. B. Clennell (1999), Formation of natural
gas hydrates in marine sediments 2: Thermodynamic calculations of
stability conditions in porous sediments, J. Geophys. Res., 104,
23,00523,022.
Holbrook, W. S., H. Hoskins, W. T. Wood, R. A. Stephen, and D. Lizarralde
(1996), Methane hydrate and free gas on the Blake Ridge from vertical
seismic profiling, Science, 273, 18401843.
Hornbach, M. J., D. A. Saffer, and W. S. Holbrook (2004), Critically
pressured free-gas reservoirs below gas-hydrate provinces, Nature, 427,
142144.
Kamath, V. A. (1984), Study of heat transfer characteristics during disso-
ciation of gas hydrates in porous media, Ph.D. thesis, Univ. of Pittsburgh,
Pittsburgh, Pa.
Kayen, R. (1988), Arctic ocean landslides by sea level fall-induced gas
hydrate decomposition, M.S. thesis, Calif. State Univ., Hayward.
Kayen, R. E., and H. J. Lee (1991), Pleistocene slope instability of gas
hydrate-laden sediment on the Beaufort Sea margin, Mar. Geotechnol.,
10, 125141.
Kennett, J. P., and B. N. Fackler-Adams (2000), Relationship of clathrate
instability to sediment deformation in the upper Neogene of California,
Geology, 28, 215218.
Kennett, J. P., K. G. Cannariato, I. L. Hendy, and R. J. Behl (2003),
Methane Hydrates in Quaternary Climate Change: The Clathrate Gun
Hypothesis, 224 pp., AGU, Washington, D. C.
MacDonald, G. J. (1990), Role of methane clathrates in past and future
climates, Clim. Change, 16, 247281.
McIver, R. D. (1982), Role of naturally occurring gas hydrates in sediment
transport, AAPG Bull., 66(6), 789792.
Mienert, J., M. Vanneste, S. Bunz, K. Andreassen, H. Haflidason, and
H. P. Sejrup (2005), Ocean warming and gas hydrate stability on the
mid-Norwegian margin at the Storegga Slide, Mar. Pet. Geol., 22, 233
244.
Nisbet, E. (1990), The end of the ice-age, Can. J. Earth Sci., 27, 148157.
Paull, C., W. J. Buelow, W. Ussler, and W. S. Borowski (1996), Increased
continental-margin slumping frequency during sea-level lowstands above
gas hydrate- bearing sediments, Geology, 24, 143146.
Paull, C. K., P. G. Brewer, W. Ussler, E. T. Peltzer, G. Rehder, and
D. Clague (2003), An experiment demonstrating that marine slumping
is a mechanism to transfer methane from seafloor gas-hydrate deposits
into the upper ocean and atmosphere, Geo Mar. Lett., 22(4), 198203.
Puzrin, A. M., and L. N. Germanovich (2005), The growth of shear bands
in the catastrophic failure of soils, Proc. R. Soc. Math. Phys. Eng. Sci.,
461(2056), 11991228.
Rempel, A., and B. A. Buffett (1997), Formation and accumulation of gas
hydrate in porous media, J. Geophys. Res., 102, 10,15110,164.
Rothwell, R. G., J. Thomson, and G. Kahler (1998), Low-sea-level empla-
cement of a very large Late Pleistocene megaturbidite in the western
Mediterranean Sea, Nature, 392, 377380.
Setzmann, U., and W. Wagner (1991), A new equation of state and tables of
thermodynamic properties for methane covering the range from the melt-
ing line to 625 K at pressures up to 1000 MPa, J. Phys. Chem. Ref. Data,
20, 10611151.
Singh, S. C., and T. A. Minshull (1994), Velocity structure of a gas hydrate
reflector at Ocean Drilling Program site 889 from a global seismic wave-
form inversion, J. Geophys. Res., 99, 24,22124,234.
Sloan, E. D., Jr. (1998), Clathrate Hydrates of Natural Gases, 2nd ed., rev.
and expanded, 754 pp., CRC Press, Boca Raton, Fla.
Sultan, N., P. Cochonat, J. P. Foucher, and J. Mienert (2004), Effect of gas
hydrates melting on seafloor slope instability, Mar. Geol., 213(14),
379401.
Van Rensbergen, P., M. De Batist, J. Klerkx, R. Hus, J. Poort, M. Vanneste,
N. Granin, O. Khlystov, and P. Krinitsky (2002), Sublacustrine mud
volcanoes and methane seeps caused by dissociation of gas hydrates in
Lake Baikal, Geology, 30, 631634.
Vogt, P. R., K. Crane, E. Sundvor, M. D. Max, and S. L. Pfirman (1994),
Methane-generated(?) pockmarks on young, thickly sedimented oceanic
crust in the Arctic: Vestnesa ridge, Fram strait, Geology, 22, 255258.
Vogt, P. R., J. Gardner, and K. Crane (1999), The Norwegian-Barents-
Svalbard (NBS) continental margin: Introducing a natural laboratory of
mass wasting, hydrates, and ascent of sediment, pore water, and methane,
Geo Mar. Lett., 19(1/2), 221.
Xu, W. (2002), Phase balance and dynamic equilibrium during formation
and dissociation of methane gas hydrate, paper presented at 4th Interna-
tional Conference on Gas Hydrates, Cent. for Gas Hydrate Res., Yoko-
hama, Japan.
Xu, W. (2004), Modeling dynamic marine gas hydrate systems, Am. Miner-
al., 89, 12711279.
Xu, W., and C. Ruppel (1999), Predicting the occurrence, distribution, and
evolution of methane gas hydrate in porous marine sediments, J. Geo-
phys. Res., 104, 50815096.
Xu, W., R. P. Lowell, and E. T. Peltzer (2001), Effect of seafloor tempera-
ture and pressure variations on methane flux from a gas hydrate layer:
Comparison between current and late Paleocene climate conditions,
J. Geophys. Res., 106, 26,41326,423.
B01104 XU AND GERMANOVICH: HYDRATE DISSOCIATION AND PORE PRESSURE
11 of 12
B01104
Yuan, T., R. D. Hyndman, G. D. Spence, and B. Desmons (1996), Seismic
velocity increase and deep-sea hydrate concentration above a bottom
simulating reflector on the northern Cascadia continental slope, J. Geo-
phys. Res., 101, 13,65513,671.
Zatsepina, O. Y., and B. A. Buffett (1998), Thermodynamic conditions for
the stability of gas hydrate in the seafloor, J. Geophys. Res., 103,
24,12724,139.
Zhang, Y. (2003), Formation of hydrate from single-phase aqueous
solutions, internal report, 77 pp., Univ. of Pittsburgh, Pittsburgh,
Pa.
Zuhlsdorff, L., and V. Spiess (2004), Three-dimensional seismic character-
ization of a venting site reveals compelling indications of natural hydrau-
lic fracturing, Geology, 32, 101104.

L. N. Germanovich, School of Civil and Environmental Engineering,


Georgia Institute of Technology, Atlanta, GA 30332, USA.
W. Xu, School of Earth and Atmospheric Sciences, Georgia Institute of
Technology, 311 Ferst Drive, Atlanta, GA 30332, USA. (wenyue.xu@eas.
gatech.edu)
B01104 XU AND GERMANOVICH: HYDRATE DISSOCIATION AND PORE PRESSURE
12 of 12
B01104

You might also like