Sunteți pe pagina 1din 32

JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 113, B01201, doi:10.

1029/2006JB004768, 2008

A mathematical model for the formation and dissociation


of methane hydrates in the marine environment
Sabodh K. Garg,1 John W. Pritchett,1 Arata Katoh,2 Kei Baba,3 and Tetsuya Fujii4
Received 25 September 2006; revised 4 September 2007; accepted 16 October 2007; published 12 January 2008.

[1] To elucidate the geological processes associated with hydrate formation and

dissociation in the marine environment under a wide range of conditions, we have


developed a one-dimensional numerical computer model (simulator). The numerical
model can be used to simulate the following aspects of hydrate formation, decomposition,
reformation, and distribution: (1) burial history of deep marine sediments and
associated phenomena (e.g., sediment compaction and consequent reduction in sediment
porosity and permeability, fluid expulsion, time evolution of temperature and pressure,
heat flux), (2) in situ generation of biogenic methane from buried organic carbon and
methane solubility in formation brine, (3) methane hydrate formation, decomposition,
reformation, and distribution in response to changes in gas concentration, pressure,
temperature and fluid salinity (the hydrate formation and decomposition are treated as
equilibrium processes), (4) influence of sulfate reduction zone under the seafloor on
hydrate formation, (5) possible presence of a free gas zone beneath the gas hydrate
stability zone, and (6) multiphase (i.e., liquid brine with dissolved gas, free gas, gas
hydrate) flow through a deformable porous matrix. The model provides for a reduction/
increase in permeability due to the formation/decomposition of the gas hydrate. Initial
applications of the model to study hydrate distribution at the Blake Ridge (site 997) and
Hydrate Ridge (site 1249) are described. Model results are compared with chlorinity,
sulfate, and hydrate distribution data.
Citation: Garg, S. K., J. W. Pritchett, A. Katoh, K. Baba, and T. Fujii (2008), A mathematical model for the formation and
dissociation of methane hydrates in the marine environment, J. Geophys. Res., 113, B01201, doi:10.1029/2006JB004768.

1. Introduction
[2] Natural gas hydrates occur worldwide in oceanic
sediments and to a lesser degree in permafrost regions.
Kvenvolden [1993] placed the global methane content of gas
hydrates at 2  1016 m3, which suggests that more carbon is
tied up in gas hydrates than in the worlds fossil fuel
deposits. On the basis of laboratory studies of hydrate
stability, much of the earths continental slope and rise
where water depths exceed 530 m at low latitudes and
250 m at high latitudes may contain methane hydrate
accumulations. Apart from temperature and pressure, hydrate accumulation is controlled by methane supply. To
quantify hydrate formation in marine sediments, several
semianalytical models of varying complexity have been
developed in recent years. Hyndman and Davis [1992] were
among the first to suggest that hydrate layers are formed
through the removal of methane from upward moving fluids
as they enter the hydrate stability zone (HSZ). They
1
Science Applications International Corporation, San Diego, California,
USA.
2
JGI Inc, Tokyo, Japan.
3
Research Center, Japan Petroleum Exploration Co. Ltd., Chiba, Japan.
4
Japan Oil, Gas and Metals National Corporation, Chiba, Japan.

Copyright 2008 by the American Geophysical Union.


0148-0227/08/2006JB004768$09.00

indicated that although most of the methane is generated


below the HSZ, the methane is primarily biogenic in origin.
Rempel and Buffett [1997, 1998] and Xu and Ruppel [1999]
solved the coupled mass and energy balance equations to
compute the rate of hydrate formation when methane is
either transported into the hydrate stability zone or is
produced in situ by methane sources (e.g., decay of organic
carbon). These early models did not account for the source
of methane, and the deposition and subsequent burial of
oceanic sediments. Davie and Buffett [2001] formulated a
semianalytical model that includes sedimentation and in situ
production of biogenic methane. This model was subsequently applied by Davie and Buffett [2003a] to study
hydrate formation in both passive (Blake Ridge) and active
(Cascadia) margin settings. A steady state model for marine
hydrate formation, including the effects of a sulfate reduction zone, was published by Davie and Buffett [2003b] and
applied by Buffett and Archer [2004] to estimate the global
inventory of methane hydrates. Liu and Flemings [2006,
2007] modeled the three-phase hydrate stability zone present at the Hydrate Ridge (site 1249). Bhatnagar et al. [2007]
simulated basic modes of gas hydrate distribution in marine
sediments, and scaled the results using combinations of
dimensionless variables.
[3] The mathematical model presented herein builds on
and generalizes the models developed by Rempel and
Buffett [1997, 1998], Xu and Ruppel [1999], Davie and

B01201

1 of 32

B01201

GARG ET AL.: FORMATION AND DISSOCIATION OF HYDRATES

Buffett [2001], and Xu [2004]. Our model is most similar to


the work of Davie and Buffett [2001]. To formulate their
model, Davie and Buffett [2001] invoked a number of
assumptions: (1) sedimentation rate is constant, (2) porosity
is a monotonic function of depth z below the seafloor and is
independent of time, (3) temperature increases linearly with
depth below the seafloor and any perturbations due to
hydrate formation (or dissociation) or convective flow are
negligible, (4) the liquid moves with the sediments at great
depth, and (5) the free gas (i.e., gas bubbles) below the
hydrate stability zone (HSZ) is trapped in the sediments and
thus has the same velocity as the sediments. The formulation implies that the sediment velocity, and hence free gas
velocity below the HSZ, is positive for all z [see Davie and
Buffett, 2001, equation (11)]. Stated somewhat differently,
the free gas released by hydrate dissociation below the HSZ
is not allowed to migrate upward into the HSZ and increase
the hydrate concentration. Davie and Buffett [2001] report
that the dissolved methane concentration, hydrate volume,
chlorinity and bubble volume (i.e., free gas volume below
HSZ) all develop steady profiles after a period of sedimentation (about 5.9 Ma for the Blake Ridge case).
[4] Application of existing models [Xu and Ruppel, 1999;
Xu, 2004; Davie and Buffett, 2001; Liu and Flemings, 2006,
2007] to simulate hydrate distribution in the subsurface is
limited by the simplifying assumptions required to solve the
coupled system of mass and energy balance equations. To
better elucidate the geological processes associated with
hydrate formation and dissociation in the marine environment under a wide range of conditions, we have developed
a one-dimensional numerical computer model (simulator).
The numerical model can be used to simulate the following
aspects of hydrate formation, decomposition, reformation,
and distribution: (1) burial history of deep marine sediments
and associated phenomena (e.g., sediment compaction and
consequent reduction in sediment porosity and permeability,
fluid expulsion, time evolution of temperature and pressure,
changes in effective stress, heat flux), (2) in situ generation
of biogenic methane from buried organic carbon, and
methane solubility in formation brine, (3) methane hydrate
formation, decomposition, reformation, and distribution in
response to changes in gas concentration, pressure, temperature and fluid salinity (hydrate formation and decomposition are treated as equilibrium processes since kinetic
phenomena are not expected to be important on geologic
timescales), (4) influence of the sulfate reduction zone on
hydrate formation, (5) possible presence of a free gas region
beneath the hydrate stability zone, and (6) multiphase (i.e.,
liquid brine with dissolved gas, free gas, gas hydrate) flow
through a deformable porous matrix. The model provides
for a reduction/increase in permeability due to the formation/decomposition of the gas hydrate.
[5] Our mathematical model, including balance laws,
initial and boundary conditions and constitutive relations,
is described in section 2. Then, a brief outline of the
numerical computer model is provided (section 3). Finally,
in section 4 we describe initial applications of the model to
study hydrate distributions at Blake Ridge (site 997) and
Hydrate Ridge (site 1249). Significant model results for
these two sites are summarized in the next two paragraphs.
[6] Our calculations for site 997 at Blake Ridge indicate
that methane migration into the HSZ from a deep source (or

B01201

a mixed methane source, i.e., methane generation in situ and


deep methane influx) is the most likely mechanism for the
observed hydrate distribution at this site. If the upflowing
fluid is just saturated with methane, then the maximum
computed free gas saturation beneath the HSZ is less than 2
to 3%, in line with saturation values derived from seismic
observations [see, e.g., Holbrook et al., 1996]. To obtain a
larger free gas saturation beneath the HSZ as inferred from
well logs [Collett and Ladd, 2000; Flemings et al., 2003],
the upflow must contain some free gas. Model results also
indicate that the free gas saturation below the HSZ as well
as hydrate distribution near the bottom of the stability zone
are a strong function of the assumed critical gas saturation.
Once critical gas saturation has been attained, the free gas
(either from the upward fluid influx at the sediment-basement boundary or that released by the burial and dissociation of gas hydrates below the HSZ) moves up into the
hydrate stability zone and results in high hydrate concentrations. Unlike previous work [see, e.g., Xu and Ruppel,
1999; Davie and Buffett, 2001; Liu and Flemings, 2007],
our model provides for a direct computation of the effective
stress. The computed effective stress (equals vertical stress
plus gas pressure; negative in compression) at the base of
the HSZ is rather large and compressive even with a
continuous free gas column (gas saturation about 10%)
beneath the HSZ. Thus for realistic values of capillary
pressure, it is unlikely that the sediments at site 997 are
critically stressed as implied by some previous workers
[Flemings et al., 2003; Hornbach et al., 2004].
[7] Liu and Flemings [2006] were the first to discuss the
role of three-phase (hydrate, liquid and gas) equilibrium in
free gas migration to the seafloor at site 1249. Later, Liu and
Flemings [2007] presented generic calculations based on
observations at site 1249. Our calculations for Hydrate
Ridge site 1249 differ in several important respects from
the previous work of Liu and Flemings [2007]. We simulate
the observed distributions of porosity and in situ chlorinity
at site 1249. Unlike Liu and Flemings [2007], our model
includes the heat of hydrate formation. Formation of hydrate
at early times releases a large amount of heat, and results in
a shoaling of the base of the HSZ. Later on after most of the
hydrate has formed, the HSZ deepens. The computed
temperature profile within the HSZ is nonlinear; this model
result should be testable with precise temperature measurements in boreholes. Finally, we note that the computed
effective stress at the bottom of the HSZ is compressive but
small in magnitude. It is thus possible that capillary effects
at site 1249 may result in sufficiently high gas pressure to
cause sediment failure.

2. Mathematical Model
2.1. Balance Equations
[8] We consider a system of sediment, seawater, gas
hydrate and gas. The spatial coordinate (Eulerian coordinate) extends from the sediment-basement boundary at z =
z0 (t) to the seafloor at z = zmax (t), where t denotes time.
Because of sedimentation and burial, the spatial locations of
both boundaries, i.e., sediment-basement interface and seafloor, will vary with time. To treat the moving boundaries, it
is convenient to use a reference frame (Lagrangian coordinate system) attached to the sediment. The Lagrangian

2 of 32

B01201

GARG ET AL.: FORMATION AND DISSOCIATION OF HYDRATES

coordinate m denotes the mass of the sediment per unit area,


and extends from m = 0 at the sediment-basement interface
to mmax(t) at the seafloor. The following relations may be
used to transform the balance equations expressed in the
Eulerian framework to those in the Lagrangian framework:
dm 1  8rr dz

f z; t f m; t

B01201

Salt is excluded from the solid hydrate structure [see, e.g.,


Hesse and Harrison, 1981]; in addition, the amount of salt
carried over into the gas phase is likely to be exceedingly
small. Thus salt can be present only (1) within the liquid
brine and (2) as a solid precipitate. Our assumptions on
phase partitioning are similar to the ones adopted by Xu
[2004].
[12] Let Sb, Sg, Sh, and Sp denote the pore volume
fractions occupied by liquid brine, free gas, gas hydrate,
and solid precipitate (salt), respectively. Note that
Sh Sp Sb Sg 1

@f
@ f
1  8rr
@z
@m

df
@f
@f

u
dt
@t
@z
@ f
@ f

u  ur 1  8rr
@t
@m

[13] We next introduce the following mass fractions:


agm mass of methane in gas phase=mass of gas phase 1
8a

[9] In equations (1) (4), 8 is the porosity, rr is the grain


density of the sediments, f is a function of z (or alternately
of m) and t, u is a fluid particle velocity, and ur is the
sediment velocity.
[10] The rock grains are assumed to be incompressible;
the grain density rr can, however, vary with lithology. Thus

ahm mass of methane in hydrate phase=mass of hydrate


16=16 18n
8b

abm mass of methane dissolved in brine=mass of brine 8c

abw mass of liquid water in brine=mass of brine

8d

The available organic carbon is assumed to occupy the


same volume as the rock grains. The mass of available
organic carbon per unit volume, Cc, is given by

ahw mass of water in hydrate phase=mass of hydrate


1  ahm

8e

Cc ac 1  8rr

abs mass of salt dissolved in brine=mass of brine

8f

rr  rr z  rr m

where ac denotes the mass ratio of available organic carbon


to that of sediment. Production of methane gas from the
breakdown of organic carbon involves a complex sequence
of chemical processes [see, e.g., Claypool and Kaplan,
1974; Claypool and Kvenvolden, 1983]. For present
purposes, it will suffice to use the exponential decay
function of Davie and Buffett [2001] to represent methane
production from the decay of organic carbon. To ensure
mass conservation, it is expedient to replace Cc by C and ac
by a in equation (6), where C (= 4 Cc/3) is the mass per unit
volume of organic carbon and hydrogen that is available
for conversion to methane.
[11] The pore volume may contain up to four phases, i.e.,
two solid phases (gas hydrate, salt precipitate), one liquid
phase (brine), and a gas phase [see, e.g., Xu, 2004]. The
chemical species present in the pore volume are methane,
water, and salt. Methane may exist as (1) free gas, (2)
solution gas (i.e., gas dissolved in brine), and (3) gas
incorporated in the solid hydrate phase. The vapor pressure
for water is quite small (<0.02 MPa) at temperatures less
than 60C; it is therefore reasonable to assume that the gas
phase contains only a negligible amount of water vapor.
Stated somewhat differently, water is contained entirely
within (1) the liquid brine, and (2) the solid gas hydrate.

aps mass of salt in precipitate=mass of precipitate 1 8g

where n denotes the hydration number, and


abm abs abw  1

[14] Since hydrate formation and dissociation in the


marine environment occur over an extended period
(thousands to millions of years), it is permissible to assume
local thermodynamic equilibrium and neglect any chemical
kinetic effects. We will also assume that the two solid
phases in the pores (i.e., gas hydrate and salt precipitate)
and the available organic carbon are stationary relative to
the sediments. The assumption of local thermodynamic
equilibrium implies that we need to consider only a single
energy balance law for the entire system. Other balance
equations consist of a separate mass balance for each of the
five chemical species (sediment, available organic carbon,
water, salt, and methane), Darcys laws to establish brine
and gas velocities, and an overall momentum balance
relation. Darcys law is an approximation to the momentum
balance relation for slow fluid flow through deformable
sediments [see, e.g., Garg, 1987]; non-Darcian effects

3 of 32

GARG ET AL.: FORMATION AND DISSOCIATION OF HYDRATES

B01201

become important for only relatively rapid flows (e.g.,


radial gas flow near a wellbore) and should not be a concern
in hydrate formation. In the Eulerian reference frame, the
balance laws can be expressed as follows [see, e.g., Garg,
1987]:
Sediment mass balance
@
@
1  8rr 1  8rr ur 0
@t
@z

10

Organic Carbon mass balance


@
@
1  8rr a 1  8rr aur M_
@t
@z

11

Energy balance
@h
1  8rr Er CEr 8Sg rg Eg 8Sh rh Eh
@t
i @
8Sb rb Eb 8Sp rp Ep 1  8rr Er ur
@z
CEr ur 8Sh rh Eh ur 8Sp rp Ep ur
i
@
8Sg rg Eg ug 8Sb rb Eb ub 1  8sr ur
@z


 8Pg Sg ug  8Pb Sb ub  8Pb ur Sh Sp
h
1  8rr ur Cur 8Sg rg ug 8Sb rb ub



i
@
@T
8ur Sh rh Sp rp g
k
@z
@z

sr Pr

12

dC @C
@C

ur
lCo explt lC
dt
@t
@z

13

M_ lC C

o
@n
8Sb rb abs 8Sp rp
@t
o
@ n

8Sb rb abs ub 8Sp rp ur


@z


@
@

Ds 8Sb rb abs
@z
@z

20

where C0 is the available organic mass at the seafloor and l


is the rate constant for methanogenesis. Combining
equations (6), (11), and (20), there follows:

Salt mass balance

@ur
@z

21

[16] Balance laws (10), (21), and (12) (18) must be


adjoined by suitable constitutive relations, and boundary
and initial conditions.
14

Darcys law (gas phase)






kkrg @Pg
Mg 8Sg rg ug  ur 
 rg g
ng
@z

15

Darcys law (liquid phase)


Mb 8Sb rb ub  ur 



1
19
sz 8Pg Sg 8Pb Sb Sh Sp
1  8

[15] Following Davie and Buffett [2001], we assume the


following exponential decay law for the available organic
carbon C:

Water mass balance


@
f8Sb rb abw 8Sh rh ahw g
@t
@
f8Sb rb abw ub 8Sh rh ahw ur g 0
@z

18

where sr denotes the intrinsic sediment stress [see, e.g.,


Morland et al., 2004] and is given by

Methane mass balance


o
@n
8Sg rg 8Sb rb abm 8Sh rh ahm
@t
o
@ n

8Sg rg ug 8Sb rb abm ub 8Sh rh ahm ur


@z 

@
@
M_
Dm 8Sb rb abm
@z
@z

B01201

2.2. Boundary Conditions


[17] Boundary conditions need to be specified at m = 0
(bottom boundary) and at m = mmax (top boundary). At the
bottom boundary, we prescribe a heat flux and, if desired, an
upward mass flux. In addition, it is assumed that no
diffusion of methane (or salt) dissolved in the liquid phase
takes place across this boundary. Thus at m = 0, the
boundary conditions are
Heat flux

kkrb @Pb
 rb g
nb
@z

Q k

16

dT
Q0
dz

Mass flux
Solid momentum balance
i
@sz h
1  8rr C 8Sg rg 8Sh rh 8Sb rb 8Sp rp g 0
@z
17

M Mg Mb M0

Methane diffusive flux (liquid phase)


@abm
0
@z

4 of 32

B01201

GARG ET AL.: FORMATION AND DISSOCIATION OF HYDRATES

Salt diffusive flux (liquid phase)

the effective stress, se; more specifically, we will use the


following relationship:

@abs
0
@z

8 81 80  81 expse =B

[18] If a mass flux is prescribed, then it is necessary to


specify its composition (i.e., mass fractions of gas, water,
and salt) and enthalpy. If the basement is rigid, then the
vertical elevation of the sediment-basement interface will
always lie at
z0 t zmin 0:

However, in reality, as additional sediment accumulates, it is


likely that the basement rocks will deform to some degree,
so that the basement/sediment interface will tend to move
downward with time. To first order, this downward motion
can be modeled as linear elastic deformation, so that
z0 t zmin t Vbm m0  mmax t

t 0

23

In addition to the sedimentation rate (23), it is necessary to


specify the sediment density and available organic
content. Also at m = mmax, it is essential to prescribe the
temperature (equal to seafloor temperature), fluid pressure
(equal to the pressure exerted by the overlying column of
seawater), salinity abs (equal to the salinity of seawater) and
methane content abm (equal to the methane content of
seawater) of the brine.
2.3. Initial Conditions
[20] At t = 0, the fluid-saturated sediment column is
assumed to be in equilibrium under its own weight and that
of the overlying seawater column. Normally, the sediment
column is initially assumed to be devoid of available
organic carbon. If desired, an initial available organic
content can be prescribed as a function of the Lagrangian
variable m. Fluid pressures are hydrostatic and no fluid
upflow or downflow occurs. The temperature distribution is
determined by heat conduction with the prescribed heat flux
at m = 0 and seafloor temperature at m = mmax. The salt
content and dissolved methane content of the pore fluid
should also be specified. Ordinarily, the salt content of the
pore fluid may be taken as uniform and equal to that of the
overlying ocean; the mass fraction of dissolved methane can
be taken to be the same as that of seawater.

k 8 k0 8=80 a 1  80 =1  8 b expc8  80

25

where k0 is the permeability at 80, and a, b, and c are


material constants. Equation (25) is a rather general
expression. With a suitable choice of material constants a,
b, and c, equation (25) can be made to reproduce virtually
any measured variation of absolute permeability with
porosity.
[23] The relative permeabilities to brine and gas are
assumed to be functions of brine and gas saturations:

p  

krb Sb Sg frb Sb = Sb Sg

p  

krg Sb Sg frg Sg = Sb Sg

26

At present, little or no laboratory data are available on


relative permeabilities for hydrate-bearing sediments; such
data are essential for developing specific functional forms
for frb and frg.
[24] In the absence of solid phases (hydrate, precipitate),
equations (26) imply that krb and krg are functions of Sb and
Sg, respectively. The presence of solid phases results in a
reduction, represented by (Sb + Sg )p where p is a positive
constant, in the relative permeability. Kleinberg et al. [2003]
used a bundle of capillary tubes to model the decrease in
permeability due to the formation of gas hydrate (1) in the
center of the pores, and (2) along the walls of the pores. As
noted by Liu and Flemings [2007], permeability decreases
more rapidly with hydrate saturation for the pore-filling
model than for the pore wall model. The pore wall case is
represented by p = 2 in equation (26); larger values of p
correspond to a faster reduction in permeability with hydrate
saturation.
[25] The sediment (including available organic carbon)
enthalpy is given by
hr Er

2.4. Constitutive Relations


[21] Constitutive relations for the sediment and pore
fluids and solids are required to complement the balance
laws. The sediment porosity, 8, will in general depend upon

24

where B is a (positive) material constant, 80 is porosity at


zero effective stress, 81 is asymptotic porosity at very high
stress, and 80 > 81. The effective stress se (negative in
compression) is defined as the sum of the vertical stress and
gas pressure (equals liquid pressure plus capillary pressure).
We recognize that sediments behave differently during
loading (effective stress becoming more compressive) and
unloading (effective stress becoming more tensile). The
current version of our model assumes identical behavior
during loading and unloading. The numerical model can,
however, be extended to include hysteresis effects.
[22] Absolute sediment permeability, k, is taken to vary
with porosity according to

22

where m0 is the total mass of sediments at t = 0, mmax is the


total mass of sediments at time t, and Vbm is a material
constant related to the elasticity of the basement materials.
[19] At m = mmax (the seafloor), we prescribe the sediment mass deposition rate, i.e.,
dmmax
qt
dt

B01201

Pr
cr T
rr

27

where cr is the specific heat of the sediment.


[26] The pore volume may contain up to four phases:
liquid brine, free gas, solid hydrate, and salt precipitate [see

5 of 32

GARG ET AL.: FORMATION AND DISSOCIATION OF HYDRATES

B01201

also Xu, 2004]. To develop a constitutive description for the


pore contents, we introduce methane (am) and salt (as) mass
fractions as follows:am is mass of methane in pore space/
total mass of pore contents


Sg rg Sb rb abm Sh rh ahm

am

Sg rg Sb rb Sh rh Sp rp

28

B01201

[31] The gas hydrate and salt precipitate are treated as


thermoelastic solids:
ri r0i exp ci P  bi T ; i h; p

33

where r0i is the density at reference conditions (P = T = 0),


ci is the isothermal compressibility, and b i is the volumetric
thermal expansion coefficient. The salt enthalpy is given by

as is mass of salt in pore space/total mass of pore contents

hp cvp T P=rp


Sb rb abs Sp rp

29

where cvp is the specific heat for solid salt. The hydrate
enthalpy can be expressed as

[27] Note that the mass fraction of water, aw, is (1  am 


as). The enthalpy of the pore contents, h, is given by





hh ahw hw P; Teq ahm hg P; Teq  DH for T Teq






hh ahw hw P; Teq ahm hg P; Teq  DH cph T  Teq
for T < Teq
35


as

Sg rg Sb rb Sh rh Sp rp

Sg rg hg Sh rh hh Sb rb hb Sp rp hp


h
Sg rg Sb rb Sh rh Sp rp


30

where hi = Ei + Pi/ri, i = g, h, b, p.
[28] Given thermodynamic pressure P(= Ph = Pp = Pb),
enthalpy h, and mass fractions am and as, the constitutive
relations should provide values for the following 16
unknowns: saturations Sb, Sg, Sh, Sp, densities rb, rg, rh, rp,
enthalpies hb, hg, hh, hp, mass fractions abm, abs, abw, and
temperature T. Equations (7), (9), (28), (29), and (30) are five
of the sixteen relations needed to determine the sixteen
unknowns. Eight of the remaining equations relate densities
and enthalpies for brine, gas, hydrate, and salt precipitate to
pressure, temperature, salinity, and methane dissolved in
brine. Two equations provide constraints on salinity and
dissolved methane. The hydrate stability relation yields the
hydrate equilibrium temperature Teq as a function of pressure
and salinity, and constitutes the last relationship.
[29] The gas density rg is computed using the gas law:
rg Pg =Ng ZRTK

31

where Z is the gas compressibility factor, Ng is the number of


moles of gas per kilogram, R is the universal gas constant
(=8.3145 J mol1 K1), and TK is the absolute temperature
(kelvins). It is assumed that gas enthalpy hg is a unique
function of pressure and temperature. In this work, tabulated
values of Dins [1956] are used to represent Z and hg.
[30] The gas pressure Pg and the brine pressure Pb (=P)
are linked through the following capillary pressure relationship:
 

Pc Sg = Sb Sg Pg  Pb

32

Like the relative permeabilities (equation (26)), capillary


pressure Pc is taken to be a function of gas and brine
saturations. In the absence of solid phases (i.e., gas hydrate
and/or salt precipitate), capillary pressure Pc depends only
on gas saturation Sg.

34

where hw is the enthalpy of pure liquid water, hg is the gas


phase enthalpy (see above), DH is the heat of dissociation
(see section 2.5), and cph is the constant pressure specific
heat for gas hydrate.
[32] The brine density rb is approximated by
rb rb0 exp cb P

36

where
cb cw =1 2:6abs

rb0 rw0 1 0:81abs

where cw is the isothermal compressibility of pure water,


and rw0 is the density of pure water at P = 0 and T. The fit
for brine density (equation (36)) is in good agreement with
density measurements for temperature (0 to 60C), pressure
(up to about 100 MPa), and salinity (up to 100,000 ppm)
values of interest for marine gas hydrates.
[33] Neglecting the heat of dissolution, brine enthalpy hb is
given by
hb abm hg P; T abs hp P; T abw hw P; T

37

[34] The maximum amount of salt that can be dissolved


in water is a function of temperature. If the total amount of
salt is greater than that can be dissolved, then the excess will
form a solid precipitate (Sp > 0). Mathematically, this
constraint can be expressed as
abs  abs max T

38

where absmax(T) denotes the maximum salinity at temperature T. If the salt mass fraction as is insufficient to saturate
the brine, the solid precipitate will not form (Sp = 0).
[35] Above the hydrate equilibrium temperature Teq, the
maximum amount of methane that may be dissolved in
brine is a function of pressure P, temperature T, and salinity
abs. In this work, we use a correlation for methane solubility
in brine, abmmax(P,T, abs), developed by Garg et al. [1977]

6 of 32

GARG ET AL.: FORMATION AND DISSOCIATION OF HYDRATES

B01201

B01201

Table 1. Solubility of Methane in Pure Water Obtained From a Fit by Garg et al. [1977] and the Theoretical
Models of Zatsepina and Buffett [1997] and Duan and Mao [2006]a
Methane Concentration, mM
P = 10 MPa
Temperature, K
286
287
288
289
290
291
292
293
294
295
296
297
298
299
300

ZB
108
103

96

90

P = 20 MPa

DM

GP

124
122
120
118
116
114
112
111
109
107
106
104
103
101

97
96
95
94
94
93
92
91
90
90
89
88
87
87
86

P = 30 MPa

ZB

DM

GP

ZB

156

171
169
166
164
162
160
158
156
155

159
158
157
155
154
153
152
150
149

186

152
151

143

179

DM

GP

199
197
195
192
190
189

197
195
194
193
191
190

ZB, Zatsepina and Buffett [1997] and Davie et al. [2004]; DM, Duan and Mao [2006], tabular values computed using a
program available at http://www.geochem-model.org/programs.htm; GP, Garg et al. [1977] and Pritchett et al. [1979].

and Pritchett et al. [1979]. The correlation utilizes the


solubility data of Culberson and McKetta [1951] for methane in pure water and a correction factor for salinity
presented by Brill and Beggs [1975]. The methane solubility constraint can be written as
abm  abm max P; T ; abs

39

If the methane mass fraction is insufficient to saturate the


brine, then a free gas phase will not form (Sg = 0).
[36] Zatsepina and Buffett [1997] employed a thermodynamic model to compute the solubility of methane in water
over a range of pressures and temperatures. A table of
solubility values based on the work of Zatsepina and Buffett
[1997] was subsequently published by Davie et al. [2004].
Duan et al. [1992] and Duan and Mao [2006] present a
theoretical model for the solubility of methane in water; the
range of applicability extends from 273 to 523 K and from 1
to 200 MPa. According to Duan and Mao [2006], their
model reproduces the experimental data with an overall
accuracy of about 5%. Because of the relative paucity of
solubility data at low temperatures and high pressures, the
accuracy of the theoretical fit of Duan and Mao [2006] may
be significantly worse than 5% in the pressure and temperature region of interest for gas hydrates. A comparison of
the solubility values predicted by the three fits [Garg et al.,
1977; Zatsepina and Buffett, 1997; Duan and Mao, 2006] is
shown in Table 1. Taking into consideration the underlying
experimental data, the agreement between the three fits is
quite good.
[37] The equilibrium temperature Teq is a function of
pressure (P) and salinity (abs), and increases with increasing
pressure and decreases with increasing salinity. If the actual
temperature T exceeds Teq for the particular pressure and
salinity present, solid hydrate cannot form and any excess
methane present beyond the solubility limit in the liquid
phase will form a separate gas phase. Similarly, if T < Teq,
the gaseous phase cannot form and the only possible

combinations are those involving a liquid phase and solid


methane hydrate.
[38] For methane-saturated states with T > Teq, the mass
fraction of dissolved methane in the liquid phase will be
equal to the solubility of the gas in water, which decreases
with increasing temperature, decreases with increasing salinity, and increases approximately in proportion to absolute
pressure. The methane solubility reaches a maximum at
T = Teq. For T < Teq, i.e., within the hydrate stability zone,
methane solubility increases with temperature and decreases
with salinity [Zatsepina and Buffett, 1997; Servio and
Englezos, 2002]; however, the solubility is relatively insensitive to pressure, and may be approximated as a function of
temperature and salinity only [Davie et al., 2004]. Thus
methane solubility in the presence of solid hydrate, aH, is
given by


aH P; T; abs abm max Peq ; T ; abs for T < Teq

40

where Peq is the equilibrium pressure at temperature T and


salinity abs (Peq  P since T  Teq). This simple model
implies that in the presence of hydrate, methane solubility
undergoes a drastic reduction with decreasing temperature
[see also Davie et al., 2004].
[39] For pure water-methane systems, Bakker [1998]
gives the following empirical fit for the hydrate melting
pressure as a function of temperature:
2
3
log10 P a0 b0 Teq c0 Teq
d0 Teq

41

where P is in MPa, and Teq is in C, and


a0
b0
c0
d0

=
=
=
=

0.43115
0.034
0.0010333
1.7994  105

The above relation is valid for temperatures between


0.3C and 40C, and is in close agreement with the fit

7 of 32

GARG ET AL.: FORMATION AND DISSOCIATION OF HYDRATES

B01201

Table 2. Comparison of Hydrate Equilibrium Temperatures


Obtained by Integrating Equation (47) With Bakkers Empirical
Relationship Equation (41)

phase line. The volume change (per kilogram of hydrate)


associated with hydrate formation can be written as:

Teq

DV

Pressure, bars

Equation (47)

Equation (41)

Difference

25
50
75
100
150
200
250
300
400

0.68C
6.42C
10.38C
13.01C
16.42C
18.64C
20.29C
21.62C
23.77C

0.97C
6.68C
10.38C
12.84C
16.19C
18.52C
20.32C
21.80C
24.16C

0.29C
0.26C
0.00C
0.17C
0.23C
0.12C
0.04C
0.18C
0.39C

derived by Moridis [2002]. The stability relation (41) does


not incorporate the effects of salinity. For a given pressure,
an increase in fluid salinity results in a decrease in the
equilibrium temperature. To first order, the effect of salinity
can be represented by the following relation:
DTeq 30abs

42

Equation (42) implies that for normal seawater salinity,


hydrate equilibrium temperature is depressed by about 1C,
in line with the empirical fit of Brown and Bangs [1995].
[40] Given pressure P and salinity abs, equations (41) and
(42) can be solved for Teq. It is most convenient to split the
constitutive description for the pore contents into three parts
as follows:
[41] In part 1, T > Teq. The solid hydrate phase is absent
(Sh = 0), and at most three phases (liquid brine, gas, and
solid precipitate) may be present. In this case, equation (30)
constitutes an implicit relation for temperature T.
[42] In part 2, T = Teq. All four phases (i.e., hydrate, liquid
brine, gas, and solid precipitate) may be present, and
equation (30) constitutes an implicit relation for hydrate
saturation Sh.
[43] In part 3, T < Teq. This case may be subdivided into
two subcases, i.e., a water-rich and a gas-rich system. In a
water-rich system, no free gas will be present (Sg = 0).
Conversely, in a gas-rich system, the liquid brine saturation
is zero (Sb = 0). In either case, at most three phases (hydrate,
solid precipitate, and liquid brine or gas) may coexist.

B01201

1 ahw ahm


rw
rg
rh

44

[46] If the volume of hydrate approximates that of water


incorporated into hydrate, then the volume change associated with hydrate formation or dissociation can be approximated by
DV 

Nh ZRTK
P

45

where (Nh = ahmNg) is the number of moles of hydrate per


kilogram. Substituting from equation (45) into equation
(43), there follows the Clausius-Clapeyron equation:
DH Nh ZRTK2

d ln P
dT

46

[47] Moridis [2002] presents a graph for DH obtained


from equation (46); apparentlyDH increases by a factor of 3
to 4 over the temperature range 0C to 40C.
[48] Equations (45) and (46) are only valid if the condensed phase volume change (i.e., from liquid water to gas
hydrate) is negligible. It is unlikely that this approximation
will hold at pressures of 10 MPa or more. We therefore
decided to investigate the effect of accounting for the
density differences between water and hydrate. Using
equations (43) and (44), it was found that the variation in
DH over the temperature range 0C to 35C is less than
25%; this variation is, however, nonmonotonic. There is
little reason to believe that DH is a nonmonotonic function
of temperature. Most likely, the computed result is due to
inaccuracies in the hydrate stability curve (equation (41))
and the assumed parameter values in the hydrate density
relationship (equation (33)).
[49] An alternate way of looking at the Clapeyron equation is that it serves to specify Teq in differential form
dP=dT eq DH=TK DV

47

43

if the behavior of the heat of formation DH can be


established independently. Assuming that DH = constant =
450 kJ kg1, equation (47) was integrated starting from the
point P = 75 bars and T = 10.37563C; results are shown in
Table 2. Clearly, the difference between the values
computed from equation (47) and those obtained from
Bakkers empirical fit (equation (41)) is small, and most
likely within the experimental error in the determination of
Teq. Pending the availability of more precise experimental
determination of DH, it is therefore reasonable to use a
constant value of 450 kJ kg1 (equivalent to 53.8 kJ mol1)
for DH.

[45] Here DV represents the volume change that accompanies the phase transition, and dP/dT is the slope of the

2.6. Sulfate Reduction Zone


[50] Seawater contains approximately 29 mM of sulfate
ion (about 7.7% of the total dissolved solids). Sulfate

2.5. Heat of Hydrate Dissociation DH


[44] To complete the constitutive description, it is essential to specify a relation for the heat of hydrate dissociation
(or formation) DH. For simple hydrates (e.g., methane
hydrate), DH can be computed using the Clapeyron equation [Sloan, 1998]:
DH TK DV

dP
dT

8 of 32

GARG ET AL.: FORMATION AND DISSOCIATION OF HYDRATES

B01201

concentration decreases from seawater concentration at the


seafloor to zero at depths ranging from a few centimeters to
tens of meters. The chemical and biological processes
involved in sulfate reduction are complex [see, e.g., Claypool
and Kaplan, 1974; Claypool and Kvenvolden, 1983]. For
present purposes, it will suffice to model sulfate reduction by
the following equilibrium reaction [see, e.g., Davie and
Buffett, 2003b]:


CH4 SO2
4 ! HCO3 HS H2 O

48

The reaction is taken to proceed to the right instantaneously


until either (1) the available sulfate ion is exhausted, or (2)
the mass fraction of methane falls to the background
level found in the ocean.
[51] A mass balance relation similar to equation (14) is
used for tracking sulfate ion in the pore brine. The salt mass
fraction abs in equation (14) is replaced by fsulf abs, where
fsulf denotes the sulfate mass fraction of the salt content of
brine. Also, it should be noted that sulfate diffusivity in
brine Dsulf will, in general, be not the same as salt diffusivity
Ds. In the current version of our numerical model, for sake
of simplicity the sulfate diffusivity is taken to be the same as
that for salt. The product species HS and HCO
3 are
considered to rejoin the nonsulfate salt mixture, and the
H2O contributes to the aqueous phase (brine).
2.7. Effect of Pore Size on Hydrate Stability
[52] Numerous observations on hydrate cores [see, e.g.,
Kraemer et al., 2000; Ginsburg et al., 2000; Riedel et al.,
2006] suggest that gas hydrate occurs preferentially in
coarse-grained sands, and that lithology controls the distribution of gas hydrates. Differences in hydrate concentration
between coarse and fine-grained sediments may simply be
due to an increase in permeability (and hence higher fluid
flux) and a decrease in capillary pressure with grain size
[see, e.g., Bear, 1972]. Other anecdotal evidence suggests
that the hydrate may be largely absent from fine grained
sediments and that the maximum hydrate saturation in
porous sediments does not exceed 70 80%. In analogy to
ice in the pores of a freezing soil, Clennell et al. [1999,
2000] have argued that for large hydrate saturations (>70
80%) the effect of small pore size is to depress the melting
temperature below the bulk melting temperature given by
equation (41). Stated somewhat differently, in a sediment
with a distribution of pore sizes, large hydrate saturations
(>70 80%) can be only attained by a significant lowering
of the temperature below the bulk equilibrium temperature.
Lacking experimental data to quantify the effect of all the
relevant parameters on the depression of hydrate melting
temperature, we have implemented a simple formulation to
suppress the formation of additional hydrate above a specified hydrate saturation (critical hydrate saturation).

3. Numerical Model
[53] To solve the system of equations described above, a
Lagrangian finite difference approach is used, with a moving one-dimensional spatial grid that is fixed in the compacting solid sediments, and which advances the solution in
finite time steps of size Dt. A full-backward-implicit
approach is taken to time discretization so that large time

B01201

steps are permissible. The convection terms in the energy


equation are treated using the unconditionally stable and
numerical dispersion free second-order-upstream spatial
technique. The Lagrangian spatial coordinate (m) is subdivided into finite increments. These are all of uniform size
Dm, except for the uppermost interval adjacent to the
sediment-ocean interface (i.e., seafloor). This upper interval
varies in size with time, gradually increasing from Dm
initially to larger and larger values because of continuing
sedimentation. When the uppermost interval reaches 2 Dm
in size, it is subdivided into two intervals, each of size Dm.
Thereafter, the uppermost of these two intervals again
gradually grows. Thus the total number of spatial grid
blocks considered will in general increase with time, so
long as sedimentation continues.
[54] The finite difference scheme has been incorporated
into a prototype computer program called THROBS (an
acronym for Thermal and Hydraulic Response of Ocean
Bottom Sediments), written in the FORTRAN 77 programming language. The computer program provides for the
deposition rate of the sediments, the ocean temperature, the
overlying sea level, the upward seepage rate and heat
content of deep fluids into the bottom of the sediment
layer from the basement, and the upward conductive heat
flux at the sediment-basement interface to be specified
independently as arbitrary functions of time. Provision is
also made for keeping track of mass fractions of biogenic
and thermogenic methane present in the computational
volume.

4. Model Applications
[55] Accumulation of methane hydrates in marine environments is principally determined by the availability of
methane. Although marine hydrates contain methane, which
is mostly biogenic, hydrate accumulations in certain locations (e.g., the Gulf of Mexico) contain thermogenic or
mixed origin gas. Two mechanisms have been proposed for
the supply of methane. In passive margin settings (e.g.,
Blake Ridge), methane may be a mixture of (1) gas
generated in situ in the hydrate stability zone from organic
carbon contained in the sediments, and (2) gas recharge
from depth. By contrast, in active margin settings (e.g.,
Nankai Trough, Hydrate Ridge), methane hydrate is formed
by methane that migrates upward into the hydrate stability
zone. We have used the THROBS numerical model to
simulate natural examples of both types of settings. Model
predictions for Blake Ridge (site 997) and Hydrate Ridge
(site 1249) are described in this section.
4.1. Blake Ridge (Site 997)
[56] The Blake Ridge, located off the eastern coast of the
United States, has been the subject of numerous investigations. Ocean Drilling Program (ODP) Leg 164 was dedicated to understanding the characteristics of gas hydrates in
the area [Paull et al., 1996, 1998]. A transect of holes were
drilled extending from an area where no BSRs are present
(site 994) to two sites where BSRs are well developed (sites
995 and 997). Borehole logging and pore fluid geochemistry
were used to estimate the presence of hydrate in the holes.
At site 997, the hydrate occurs at depths between 185 m
below seafloor (mbsf) and 450 mbsf [Collett and Ladd,

9 of 32

B01201

GARG ET AL.: FORMATION AND DISSOCIATION OF HYDRATES

2000]. The hydrate concentration is relatively low with


estimates ranging from less than 4% to 7% of sediment
porosity [Holbrook et al., 1996; Paull et al., 1996, Egeberg
and Dickens, 1999; Collett and Ladd, 2000]. On the basis of
Br/I ratios, Egeberg and Dickens [1999] estimated an
external upward fluid flow of 0.2 mm a1; this corresponds to a net upward fluid velocity of 0.08 mm a1 at the
seafloor [Davie and Buffett, 2003a]. The Blake Ridge sediments contain high concentrations (1.5 to 2%) of total
organic carbon (TOC). In situ generation of biogenic
methane is thus a possible mechanism for the generation
of gas hydrates at the Blake Ridge [see, e.g., Matsumoto et
al., 2000]. A bottom simulating reflector (BSR) is usually
ascribed to the presence of free gas beneath the hydrate
stability zone [see, e.g., Bangs et al., 1993]. The BSR
observed at site 997 confirms the existence of free gas
below the hydrate stability zone.
[57] The hydrate accumulation at site 997 has previously
been modeled by Xu and Ruppel [1999], Davie and Buffett
[2001, 2003a, 2003b], Nimblett and Ruppel [2003], and
Baba and Katoh [2004]. Xu and Ruppel [1999] used a
steady state analytical model to estimate the upward methane flux required for the base of the hydrate occurrence
zone to coincide with the bottom of the hydrate stability
zone and the top of the free gas zone. Neglecting sedimentation and assuming an upward fluid flux (interstitial velocity) of 0.3 mm a1, the critical methane flux rate is 3 
1011 kg s1 m2 [Xu and Ruppel, 1999]. Later on,
Nimblett and Ruppel [2003] used a two-phase (gas hydrate
and fluid) transient model to show that pore/fracture clogging due to hydrate formation takes place on a timescale
ranging from 0.4 Ma to more than 440 Ma. The latter
authors also concluded that generating hydrates to a concentration of 10 15% of porosity will require 105 years in
conditions typical of Blake Ridge. Davie and Buffett [2001]
investigated the formation of hydrates from the in situ
generation of biogenic methane. To model the inferred
hydrate volume, Davie and Buffett [2001] had to assume
a rather high value (1.1%) for the available organic carbon
at the seafloor. Baba and Katoh [2004], using a onedimensional static model, found that the hydrate saturation
at site 997 could be reproduced using an available organic
carbon content of 0.78%. Hydrate accumulation due to
migration of methane-bearing fluids from a deeper source
was investigated by Davie and Buffett [2003a, 2003b]. They
found that the hydrate distribution at site 997 could be
explained with a deep methane-saturated fluid influx with
a velocity of 0.26 mm a1, or alternately with a combination of in situ biogenic methane generation (available
organic carbon: 0.75%) and an upward fluid velocity of
0.08 mm a1.
[58] An important aspect of hydrate accumulation at site
997 is that the bottom of the hydrate stability zone (depth 
450 mbsf) is about 50 m shallower than that estimated from
the hydrate stability curve using measured pressures and
temperatures. Ruppel [1997] concludes that capillary forces
arising in the fine-grained sediments may be responsible
for the shoaling of the BSR. Davie and Buffett [2003a]
used a somewhat larger than measured temperature gradient (43 K km1 versus 37 K km1) to simulate the
observed depth of the BSR at site 997.

B01201

[59] Hornbach et al. [2004] suggest that provided the free


gas below the bottom of the BSR forms an interconnected
network, faults intersecting the BSR on the Blake Ridge are
critically stressed and prone to failure. Presumably, fluid
upflow can take place along these faults. Flemings et al.
[2003] used core test data from site 994 to predict (1) water
pressures that are 70% of the lithostatic stress, and (2) gas
pressures equal to the lithostatic stress beneath the hydrate
layer at site 997. The lithostatic gas pressure dilates fractures which in turn provide pathways for gas to migrate
through the hydrate zone to the seafloor [Flemings et al.,
2003]. In the following, these conclusions will be examined
in relation to our simulations for site 997.
[60] We have applied the numerical model described in
previous sections to investigate the effect of available
carbon content and fluid influx at the sediment/basement
boundary on hydrate accumulation at site 997. Three cases
are considered. In case 1, methane is generated in situ, and
no upward flux of fluid is allowed at the sediment-basement
boundary. Available carbon content is taken to be zero in
case 2, and a brine (slightly supersaturated in methane)
influx is imposed at the bottom boundary. Finally case 3 is
designed to simulate a mixed origin (in situ generation and
influx from below) for methane. Porosity parameters in
Table 3 approximately reproduce the measured porosity
distribution with depth at site 997 [see Paull et al., 1996,
Table 9 18]. The temperature gradient in Table 3 (41.3 K
km1), selected so as to simulate the bottom of the hydrate
stability zone (BHSZ) at about 450 mbsf, is somewhat
lower than that (43 K km1) assumed by Davie and Buffett
[2003a]. The available organic carbon and the upward fluid
influx at the bottom boundary were regarded as free
parameters and were varied over a range of values in order
to match the measured chloride values [see Paull et al.,
1996, Table 9 17]; only the final values are shown in Table
4. Other model parameters (Tables 3 and 4) were selected so
as to mimic the thermodynamic and geologic conditions
assumed by Davie and Buffett [2003a].
[61] The theoretical model described in the preceding
sections requires an explicit prescription of formation permeability. The formation permeability (Table 3) was selected large enough to minimize pressure gradients due to
single-phase liquid flow through the compacting sediments.
Because of sediment compaction, sediment porosity and
hence absolute permeability decline with depth below the
seafloor (equations (24) and (25)). The simultaneous presence of multiple phases (i.e., gas hydrate, brine, gas) in the
pore space results in flow interference between different
phases; this effect is modeled through the introduction of
relative permeabilities. Permeability reduction due to hydrate formation is modeled with p = 2 in equation (26). As
noted earlier, p = 2 corresponds to the pore wall case
considered by Kleinberg et al. [2003]. Lacking data on
brine and gas relative permeabilities in hydrate-bearing
media, the following linear relationships are adopted for
the brine and gas relative permeability functions in Table 3:

10 of 32

Sb
 Sbr =1  Sbr
Sb Sg


 Sg
frg
 Sgr = 1  Sgr
Sb Sg
frb

49

GARG ET AL.: FORMATION AND DISSOCIATION OF HYDRATES

B01201

B01201

Table 3. Model Parameters for the Blake Ridge Simulations


Parameter

Units

Value

Nominal water depth


Seafloor temperature
Sediment thermal conductivity
Conductive heat flux
Sediment grain density
Porosity at effective stress = 0
Porosity at effective stress = 1
Porosity parameter B, equation (24)
Basement elastic constant Vbm, equation (22)
Permeability at porosity = 0.69
Parameter a, equation (25)
Parameter b, equation (25)
Parameter c, equation (25)
Irreducible brine saturation, Sbr
Critical gas saturation, Sgr
Parameter p, equation (26)
Brine relative permeability function frb, equation (26)
Gas relative permeability frg, equation (26)
Capillary pressure, Pc, equation (32)
Critical hydrate saturation for depression of Teq
Diffusivity of methane in brine, Dm
Diffusivity of salt in brine, Ds
Diffusivity of sulfate in brine, Dsulf
Sedimentation rate
Seawater salinity

m
C
W m1 C1
W m2
kg m3

2781
3
2
0.0826
2650
0.69
0.18
9.444
8  104
5  1017
2
0
0
0.30
0.10
2
see text
see text
0
0.80
109
109
109
5.731  109
0.035

Seawater methane mass fraction


Seawater sulfate mass fraction
Lagrangian spatial interval Dm
Initial number of grid blocks
Height of initial grid
Initial depth of basement/sediment interface
Time step
Calculation time interval

MPa
m3 kg1
m2

MPa
m2 s1
m2 s1
m2 s1
kg m2 s1

3.5  105
0.0027
12,500
2
30
2811
100
10

kg m2
m
m
years
Ma

Our numerical model allows for tabular input of the brine


and gas relative permeabilities, and is not limited to linear
functions (49).
[62] The critical gas saturation Sgr in equation (49) is the
minimum gas saturation (volume fraction of pore space
occupied by gas) at which the gas phase will flow independently of the liquid phase. Stated somewhat differently, the
critical gas saturation determines the maximum free gas
saturation below the bottom of the hydrate stability zone;
any gas in excess of the critical gas saturation will tend to
move upward into the hydrate stability zone. On the basis of
laboratory tests on 17 cores, Schowalter [1979] reported
values for Sgr ranging from 0.045 to 0.17 with an average of
0.10. Other authors [see, e.g., Bennion and Bachu, 2005]

Notes
see text
corresponds to a temperature gradient of 41.3C km1
porosity parameters selected to fit data [Paull et al., 1996]

brine is immobile for (Sb /(Sb + Sg))< Sbr


gas is immobile for (Sg /(Sb + Sg ))< Sgr
(Sb/(Sb + Sg )) Sbr
(Sg/(Sb + Sg )) Sgr
assumed to be zero except as indicated in the text

corresponds to a deposition rate of 0.22 mm a1


these values also equal the initial pore fluid salinity
and methane content
(29 mM); initial pore fluid sulfate assumed to be 0

equals nominal water depth plus initial grid height


equals 100,000 time steps

have reported much lower values for Sgr. For most of the
calculations reported in this paper, Sgr was assumed to equal
0.10. This value for Sgr ensures that for all three cases
(Tables 3 and 4), any methane in the gas phase, whether
separated from fluid influx at the bottom boundary or from
dissociation of the hydrate phase, cannot flow upward into
the hydrate stability zone. As noted elsewhere, a key
assumption invoked by Davie and Buffett [2001, 2003a,
2003b] is that any free gas resulting from hydrate dissociation gets buried along with the sediments and does not
reenter the hydrate stability zone.
[63] The capillary pressure Pc is assumed to be zero for
all three cases (Tables 3 and 4). At site 997, a relatively
sharp interface (BGHS) separates the two-phase hydrate/

Table 4. Available Organic Carbon, Rate of Methanogenesis, and Upward Fluid Flux at Sediment/Basement
Boundary
Parameter

Units

Case 1

Case 2

Case 3

Available organic carbon ac,


equation (6)
Rate of methanogenesis l,
equation (20)
Total upward mass flux
Enthalpy
Methane mass fraction
Salt mass fraction
Sulfate mass fraction

kg kg1

0.0135

0.0

0.00675

s1

3  1013

not applicable

3  1013

kg s1 m2
kJ kg1

9.0  109
124
0.00268
0.032
0

3.6  109
124
0.00268
0.032
0

11 of 32

GARG ET AL.: FORMATION AND DISSOCIATION OF HYDRATES

B01201

B01201

Figure 1. The van Genuchten [1980] capillary pressures as a function of normalized gas saturation
(= Sg/(Sb + Sg)). Curve labeled Pc-low (Pc-high) was evaluated with van Genuchten parameters d = 0.4
and P* = 0.1 MPa (1 MPa). Entry pressure (i.e., capillary pressure corresponding to a critical gas
saturation of 0.10 is 0.064 MPa for Pc-low and 0.64 MPa for Pc-high.
liquid interval from the underlying free gas/liquid zone.
Since the gas saturation Sg is zero above the BGHS, the
capillary pressure Pc is also zero. Below the BGHS, the
capillary pressure is finite. To illustrate the possible impact
of a nonzero capillary pressure, a variation of case 2 that
included a capillary pressure function was simulated. For
the latter calculation (and also for the Hydrate Ridge cases),
the following capillary pressure function [van Genuchten,
1980] was adopted:

1d
Pc P * S* 1=d 1

50

where



S  Sb = Sb Sg  Sbr =1  Sbr

Two van Genuchten capillary pressure curves are plotted in


Figure 1. These curves were computed using the following
parameters:
Sbr
d
P*
P*

=
=
=
=

0.3
0.4
0.1 MPa (Pc-low)
1.0 MPa (Pc-high)

A typical value of P* for low-permeability sediments (e.g.,


shales) is 0.1 MPa [see, e.g., Pruess and Garcia, 2002]; the
value for high-permeability sands is much lower. The Pc-

low (Pc-high) curve in Figure 1 yields an entry pressure of


0.064 MPa (0.64 MPa) at a normalized critical gas
saturation (= Sg/(Sb + Sg )) of 0.10. Although the Pc-high
curve represents an unrealistically high capillary pressure,
this curve will be used in one of the examples to illustrate
the sensitivity of the results to rather large capillary
pressures.
[64] The basement elastic constant Vbm was chosen so as
to minimize the variations in the nominal water depth. More
specifically, the nominal water depth was required not to
vary by more than 100 m from its nominal value (i.e.,
2781 m). The rate constant for methanogenesis (Table 4)
is the same as that used by Davie and Buffett [2003a]. With
this rate constant and the assumed sedimentation rate
(Table 3), essentially all of the biogenic methane generation takes place in sediments above the base of the hydrate
stability zone at 450 mbsf.
[65] Location of the BGHS as a function of time for case 1
is displayed in Figure 2. The sediment column does not attain
a thickness greater than 450 m until a time of 2.3 Ma;
thereafter, the changes in BGHS location are minimal.
Small oscillations in the BGHS location for t > 2.3 Ma are
due to spatial discretization. Temperature, porosity, gas
pressure (equals liquid pressure plus capillary pressure)
referenced to seafloor, and effective stress distributions with
depth at t = 1, 3, 5, and 10 Ma for case 1 are plotted in Figure 3.
Solid dots in the porosity plot denote the measurements and
are taken from Table 9 18 of Paull et al. [1996].
Temporal variations in temperature, porosity, gas pressure

12 of 32

B01201

GARG ET AL.: FORMATION AND DISSOCIATION OF HYDRATES

B01201

Figure 2. Bottom of hydrate stability zone (case 1). The thickness of the sediment column is less than
450 m for t < 2.3 Ma. Oscillations in the location of the bottom of hydrate stability zone are due to spatial
discretization.
and effective stress at a given depth are negligible. Plots for
cases 2 and 3 are very similar to Figures 2 and 3 and are
therefore not displayed.
[66] Free gas saturation (fraction of pore space), upward
fluid mass flux with respect to the compacting sediments,
hydrate saturation (fraction of pore space), and relative
salinity (relative to salinity at seafloor) at t = 3, 5 and
10 Ma for case 1 are shown in Figure 4. The solid dots in
the relative salinity plot denote the measured chlorinity
points normalized relative to the seawater chlorinity. The
salinity curves in Figure 4d were computed assuming
complete hydrate dissociation. The bottom of fluid flux
and relative salinity plots corresponds to the sedimentbasement boundary at the indicated time. The gas saturation
(Figure 4a) immediately below the BHSZ remains below
the critical gas saturation at all times; at greater depths, the
gas is mobile at t = 10 Ma. The increase in free gas
saturation below the BHSZ (Figure 4a) with time is due
to (1) burial and dissociation of hydrate below the base of
the HSZ and (2) sediment compaction and consequent
reduction in pore volume. The hydrate saturation distribution (Figure 4c) is more or less stable after t = 3 Ma. The top
of the hydrate zone is located at about 20 mbsf, and the
maximum hydrate saturation (>13%) occurs at the BHSZ.
Fluid flux (Figure 4b) increases toward the seafloor due to
fluid expulsion resulting from the compaction of sediments.
The discontinuous change in fluid flux at the BHSZ (t = 3 Ma

and 10 Ma) is most probably caused by hydrate dissociation. Although the fluid flow with respect to the compacting
sediments is upward, the fluid flow at the seafloor is in a
downward direction (see Table 5). The computed salinity
values at t = 5 Ma (Figure 4d) are in good agreement with
the measurements above the base of the hydrate stability
zone; the agreement is rather poor below the BHSZ. The
computed salinity values beneath the HSZ at t = 10 Ma are
lower than the measurements. Burial and dissociation of gas
hydrate freshens the pore fluid below the HSZ; some of this
low-salinity fluid is advected upward above the HSZ. The
decrease in salinity with time results in a slight deepening of
the BHSZ at late times (Figure 2). A comparison of
computed salinities at t = 5 Ma and 10 Ma suggests that a
fair agreement with measurements is obtained at an intermediate time. Our estimate for the available organic carbon
(1.35%) required to match the observed salinity profile is
somewhat larger than that derived by Davie and Buffett
[2001]. The required available organic carbon represents
about 67 to 90% of the total organic carbon (1.5 to 2%)
believed to be present in Blake Ridge sediments. According
to Waseda [1998], only about 15% of the total organic carbon
is available for conversion to methane, and it is therefore
unlikely that the available organic carbon is as high as 1.35%.
Finally, based on comparisons between generally accepted
values (see the foregoing discussion) for hydrate saturation (4
to 7% of pore space) and fluid flux (0.08 mm a1) at the

13 of 32

B01201

GARG ET AL.: FORMATION AND DISSOCIATION OF HYDRATES

B01201

Figure 3. Temperature, porosity, gas pressure (equals liquid pressure plus capillary pressure) referenced
to seafloor, and effective stress (equals vertical stress plus gas pressure; negative in compression)
distribution with depth at t = 1, 3, 5, and 10 Ma (case 1). Solid dots in the porosity plot denote the
measurements [Paull et al., 1996]. Temporal variations in temperature, porosity, gas pressure, and
effective stress at a given depth are negligible.
seafloor and calculations for case 1, it is concluded that in situ
methane generation by itself cannot represent the prevailing
conditions at site 997.
[67] Hydrate generation by methane migration into the
HSZ from a deeper source is simulated in case 2. The
prescribed fluid mass upflow at the sediment-basement
boundary (9  109 kg s1 m2) corresponds to an upward
superficial fluid velocity of 0.28 mm a1. The methane
fraction (0.00268 kg kg1) of the inflowing fluid is near
the equilibrium solubility, and yields a methane influx of
2.41  1011 kg s1 m2. The upward fluid velocity
(0.28 mm a1) is not too different from that inferred by
Davie and Buffett [2003a]. The methane influx is also more
or less in agreement with Figure 5a of Xu and Ruppel [1999].
The salt content of the incoming fluid was prescribed so as to
match the measured salinity beneath the HSZ. Computed
salinity curves (see Figure 5) for t = 3, 5, and 10 Ma are
essentially identical, and are in excellent agreement with the
measurements. The top of the hydrate zone (t = 5 and 10 Ma)
is located at about 150 mbsf, and is in fair agreement with
that (185 mbsf) reported by Collett and Ladd [2000]. Note
that the computed hydrate saturations above 185 mbsf are
very small (less than 0.2%) and are probably hard to detect
by well logging. The maximum hydrate saturation occurs

toward the bottom of the HSZ and is less than 3.9% of pore
space at t = 10 Ma; this value is at the lower end of the
estimated hydrate saturation (4 to 7%) at site 997. Apart
from a relatively thin zone close to the sediment-basement
boundary, free gas occupies a very small fraction (1.6 to
1.9%) of the pore space and is immobile. The computed
interstitial fluid velocity at the seafloor (0.31 0.34 mm a1,
Table 5) is significantly higher than that (0.08 mm a1)
estimated by Egeberg and Dickens [1999]. Although the
depth to the top of the hydrate zone and the measured
salinity profile are well reproduced by the theoretical predictions for case 2, agreement between the generally accepted and computed values for hydrate saturation and
interstitial fluid velocity at the seafloor is less satisfactory.
[68] A mixed methane source, i.e., methane generation in
situ and deep methane influx, is assumed for case 3. The
available organic carbon input at the seafloor is taken to be
one half (0.00675 kg kg1 of sediment) of that for case 1.
Deep fluid (and consequently methane) influx is 40% of that
for case 2. The computed fluid interstitial velocity at the
seafloor for the mixed methane source (0.07 0.11 mm a1,
Table 5) is in good agreement with that (0.08 mm a1)
deduced by Egeberg and Dickens [1999]. Beneath the HSZ,
free gas saturation increases from 2.2% at the BSR to

14 of 32

B01201

GARG ET AL.: FORMATION AND DISSOCIATION OF HYDRATES

B01201

Figure 4. Free gas saturation (fraction of pore space), upward fluid mass flux with respect to the
compacting sediments, hydrate saturation (fraction of pore space), and relative salinity (relative to salinity
at seafloor) at t = 3, 5 and 10 Ma (case 1). Relative salinity curves are computed assuming complete
hydrate dissociation. The solid dots in the relative salinity plot denote the measurements [Paull et al.,
1996]. The bottom of fluid flux and relative salinity plots corresponds to the sediment-basement
boundary at the indicated time. At t = 10 Ma, the gas saturation remains below the critical gas saturation
(0.10) immediately beneath the hydrate stability zone; at greater depths, the gas is mobile.
3.5% a short distance above the sediment-basement
boundary (Figure 6). The top of the hydrate zone is located
at 40 mbsf. The maximum hydrate saturation occurs near
the bottom of the HSZ, and (at t = 10 Ma) is 5.1% of the
pore space. The computed salinity profile at t = 10 Ma is in
excellent agreement with the measurements (Figure 6).
Apart from the depth to the top of the HSZ, theoretical
results for the mixed source conform to what is known (or is
generally accepted) for site 997.
[69] The spatial discretization used for cases 1 3 (Dm =
12,500 kg m2) is much too coarse to resolve the sulfate
reduction zone (SRZ). Measurements [Paull et al., 1999,
Table 9 17] indicate that the thickness of the SRZ is about

20 m. To investigate the sulfate distribution in the pore


fluids beneath the seafloor, calculations for case 2 were
repeated using a much finer spatial zoning (Dm = 2500 kg
m12). Because of excessive run time, the latter calculation
was terminated at t = 5 Ma. The computed sulfate distributions at t = 1, 3 and 5 Ma are compared with measurements
and results obtained using original coarse discretization in
Figure 7. The theoretical curve for t = 5 Ma is in good
agreement with the measurements. Also, it is important to
note that the subsurface sulfate values computed with coarse
discretization are in fair agreement with both the measurements and the results of the calculation with fine spatial
zoning. Although our numerical model can be used to

Table 5. Gas Pressure Gradient, Lithostatic Stress Gradient, and Interstitial Fluid Velocity for Blake Ridge (Site 997) Cases 1 3
Hydrostatic pressure gradient above BHSZ at 10 Ma
Gas pressure gradient above BHSZ at 10 Ma
Lithostatic stress gradient above BHSZ at 10 Ma
Interstitial fluid velocity at seafloor at 5 Ma
Interstitial fluid velocity at seafloor at 10 Ma

Units

Case 1

Case 2

Case 3

kPa m1
kPa m1
kPa m1
mm a1
mm a1

10.19
10.30
15.99
0.08
0.05

10.19
10.54
16.00
0.31
0.34

10.19
10.39
16.01
0.07
0.11

15 of 32

B01201

GARG ET AL.: FORMATION AND DISSOCIATION OF HYDRATES

B01201

Figure 5. Free gas saturation (fraction of pore space), upward fluid mass flux with respect to the
compacting sediments, hydrate saturation (fraction of pore space), and relative salinity (relative to salinity
at seafloor) at t = 3, 5, and 10 Ma (case 2). Relative salinity curves are computed assuming complete
hydrate dissociation. The solid dots in the relative salinity plot denote the measurements [Paull et al.,
1996]. The bottom of the fluid flux plot corresponds to the sediment-basement boundary at the indicated
time. Except for a small zone at the bottom of the computational volume, the free gas occupies less than
0.02 of the pore space and is immobile.
resolve the detailed structure of the SRZ, it may well prove
to be impractical in most situations because of the need for
fine spatial zoning, and consequent rather long run
times. On the other hand, it is apparent from Figure 7 that
the effect of the SRZ on the methane (dissolved in
brine, and as solid hydrate) distribution beneath the SRZ
can be adequately modeled using relatively coarse spatial
discretization.
[70] As mentioned above, capillary pressure was assumed
to be zero in the above described calculations for cases 1 3.
In section 4.2, we present a calculation for case 2 that
includes a nonzero capillary pressure. For the present, it will
suffice to note that the inclusion of a nonzero capillary
pressure does not affect the hydrate distribution. The gas
pressure gradient (see Table 5) within the HSZ for all three
cases exceeds the corresponding hydrostatic gradient by a
small amount (1 to 3%), and is about 65 (1)% of the
lithostatic gradient. Thus at the depth of the hydrate stability
zone (450 mbsf), the difference between the gas pressure
and lithostatic stress (i.e., effective stress) is 2.5 MPa.

4.2. Parametric Calculations


[71] In this section, we will employ variations of cases 1
and 2 (section 4.1) to illustrate the effect of certain model
parameters on the computed distributions of hydrate, methane concentration, and effective stress, etc. The critical gas
saturation is the minimum average gas saturation in the pore
space for the gas phase to become mobile. In the Blake
Ridge (site 997) calculations described in section 4.1, the
critical gas saturation was deliberately chosen to be high
enough (= 0.10) to preclude gas phase mobility. For cases 2
and 3, the free gas saturation below the BHSZ is rather
small and lowering the critical gas saturation to 0.05 has no
impact on the computed hydrate distribution. With the
exception of the critical gas saturation, all the model
parameters for case 1a are identical with those for case 1;
the critical gas saturation was, however, lowered to 0.05 so
as to allow free gas from beneath the HSZ to migrate
upward. Case 1a is intended to highlight the importance
of the critical gas saturation in the recycling of gas released
by hydrate dissociation. Free gas saturation (fraction of pore
space), upward fluid mass flux with respect to the compacting sediments, hydrate saturation (fraction of pore space),

16 of 32

B01201

GARG ET AL.: FORMATION AND DISSOCIATION OF HYDRATES

B01201

Figure 6. Free gas saturation (fraction of pore space), upward fluid mass flux with respect to the
compacting sediments, hydrate saturation (fraction of pore space), and relative salinity (relative to salinity
at seafloor) at t = 3, 5, and 10 Ma (case 3). Relative salinity curves are computed assuming complete
hydrate dissociation. The solid dots in the relative salinity plot denote the measurements [Paull et al.,
1996]. The bottom of the fluid flux plot corresponds to the sediment-basement boundary at the indicated
time. Except for a small zone at the bottom of the computational volume, the free gas occupies less than
0.04 of the pore space and is immobile.
and relative salinity (relative to salinity at seafloor) for case
1a at t = 5 and 10 Ma are shown in Figure 8. A zone of
elevated hydrate saturation (>80%) is present just above the
bottom of the hydrate stability zone. Below the BHSZ, free
gas (gas saturation  5%) is present. The free gas results
from hydrate dissociation that accompanies the burial of
sediments containing hydrate; this gas moves upward into
the HSZ and is responsible for the high hydrate concentration above the BHSZ. At t = 10 Ma, a three-phase (hydrate,
free gas and brine) zone extends from 387 mbsf to the
bottom of the hydrate stability zone at 471 mbsf; at
shallow depths, the hydrate concentration is the same as
that for case 1. The bottom of the HSZ for case 1a at t =
10 Ma is about 20 m deeper than for case 1. Unlike case 1,
salinity profiles for case 1a show rather large excursions.
Not surprisingly, the salinity profiles for case 1a display
extremely low salinity values in the depth interval with
elevated hydrate concentration; the low salinity is a consequence of a rather low liquid saturation (and its dilution by
hydrate on dissociation) in this zone. High salinities at
shallower depths (see t = 10 Ma profile) are due to the

upward advection of the high-salinity fluid from near the


BHSZ.
[72] To investigate the effect of nonzero capillary pressure on the hydrate distribution, two cases (cases 2a and 2b)
were run. With the exception of the capillary pressure, all
the model parameters for cases 2a and 2b are identical with
those for case 2. For cases 2a and 2b, capillary pressure was
computed using the curves (Pc-low and Pc-high) displayed
in Figure 1. A detailed examination of the results for cases 2,
2a and 2b shows that the hydrate distribution, gas pressure
and effective stress gradients above the BHSZ are practically indistinguishable for the three cases. This is not
surprising since there is no free gas above the BHSZ.
Inclusion of nonzero capillary pressure does, however,
result in differences in the effective stress distribution
beneath the hydrate zone (Figure 9). Apparent discontinuities in curve labeled Pc-high in Figure 1 are associated with
abrupt changes in gas saturation at these depths (from 0 to
1.5% at the BHSZ, and from 1.9% to 10% at
1335 mbsf). Even with a rather unrealistically high
capillary pressure (Pc-high), the change in the effective
stress is relatively minor and unlikely to result in sediment

17 of 32

B01201

GARG ET AL.: FORMATION AND DISSOCIATION OF HYDRATES

B01201

Figure 7. Comparison of pore fluid sulfate content computed using a refined spatial discretization
(case 2) with measurements [Paull et al., 1996]. Solid symbols denote the sulfate values calculated with
the original discretization (od) for case 2.
failure. For cases 2, 2a and 2b, the computed gas saturations
beneath the BHSZ are small ( less than 2% of pore space)
and are consistent with gas saturations derived from seismic
measurements [see, e.g., Holbrook et al., 1996]. Estimates
of free gas saturation obtained from logging measurements
[Collett and Ladd, 2000; Flemings et al., 2003] are considerably higher. In particular, Flemings et al. [2003] indicate
that a continuous column of mobile gas (gas saturation
equal to or greater than the critical gas saturation) is present
below the BHSZ at site 997. To simulate the latter situation,
we ran a variant of case 2b by increasing the gas fraction of
the inflowing fluid to 0.006 (i.e., over twice the equilibrium
solubility value). The free gas saturation beneath the HSZ in
this calculation is slightly in excess of the critical gas
saturation. The effective stress at the base of the HSZ is
compressive (1.9 MPa). Therefore unlike Hornbach et al.
[2004] and Flemings et al. [2003], we consider it improbable that the sediments beneath the HSZ at site 997 are
critically stressed and close to failure.
[73] The diffusive flux (i.e., diffusion of salt and methane
dissolved in water) helps (1) smooth out gradients in salinity
and (2) transfer methane to seawater. The diffusivity values
given in Table 3 are characteristic of the molecular diffusion
of methane and salt in water. Because of hydrodynamic
dispersion associated with flow in inhomogeneous porous
media, actual diffusivities may be different. To demonstrate
the effect of diffusivity on hydrate distribution, case 2 was

rerun by doubling the diffusivity values (Dm = Ds = Dsulf =


2.0  109 m2 s1). Figures 10a and 10b illustrate the effect
of diffusivity on dissolved methane concentration and
hydrate saturation with depth, respectively. A larger value
for diffusivity (case 2c) results in a smaller dissolved
methane concentration below the seafloor to a depth of
200 mbsf; this is accompanied by a deepening of the top
of the hydrate zone by about 50 m. Also, an increase in
diffusivity results in a significant reduction in the hydrate
saturation. Clearly, methane, salt, and sulfate diffusivities
are important model parameters and exercise a major
influence on dissolved methane concentration and hydrate
saturation with depth.
[74] The sulfate reduction zone (SRZ) serves as a sink of
methane, and inhibits the formation of gas hydrates at
shallow depths below the seafloor. To illustrate the effect
of the SRZ on hydrate distribution, a variant of case 2
excluding the SRZ (case 2d) was run. As expected, exclusion of the SRZ results in a higher dissolved methane
concentration (Figure 11a) below the seafloor; the higher
methane concentration extends to a depth of 150 mbsf.
Exclusion of the SRZ also leads to a higher hydrate
saturation (Figure 11b), and a shoaling of the top of the
hydrate zone. Figures 11a and 11b demonstrate the need to
include the SRZ in order to obtain a realistic estimate of
hydrate saturation in the marine environment.

18 of 32

B01201

GARG ET AL.: FORMATION AND DISSOCIATION OF HYDRATES

B01201

Figure 8. Free gas saturation (fraction of pore space), upward fluid mass flux with respect to the
compacting sediments, hydrate saturation (fraction of pore space), and relative salinity (relative to salinity
at seafloor) at t = 5 and 10 Ma (case 1a). With the exception of the critical gas saturation, all the model
parameters for case 1a are identical with those for case 1. The critical gas saturation for case 1a is 0.05.
Relative salinity curves are computed assuming complete hydrate dissociation. The bottom of the fluid
flux and salinity plots corresponds to the sediment-basement boundary at the indicated time.
4.3. Hydrate Ridge (Site 1249)
[75] Hydrate Ridge is a 25-km-long and 15-km-wide
peanut-shaped accretionary ridge on the Cascadia margin,
located approximately 80 km off the Oregon coast [Milkov
et al., 2004; Trehu et al., 2004a]. As part of ODP Leg 204,
nine sites were drilled on and around the southern part of
the Hydrate Ridge. These sites were selected to complement
ODP site 892 drilled during ODP Leg 146 near the summit
of the northern Hydrate Ridge [Trehu et al., 2004a]. Three
of the ODP Leg 204 sites (sites 1248 1250) are located on
the southern summit, an area characterized by active gas
venting. Site 1249 was cored to 90 mbsf, and pore fluids
with extremely high salinity were observed to depths of 20
40 mbsf [Milkov et al., 2004; Trehu et al., 2004a]. Below
this depth, salinity decreases to near normal seawater
concentrations. Milkov et al. [2004] suggested that active
gas venting may indicate the coexistence of free gas along
with gas hydrate and brine at shallow depths where high
salinity of the pore brine inhibits additional hydrate formation (see also Ruppel et al. [2005] on salt inhibition of
hydrate formation). Although no precise estimates of hydrate saturation at site 1249 are available, the observed
salinity is consistent with a rather high saturation. On the

basis of isotopic and molecular properties of gases, Milkov


et al. [2005] concluded that shallow hydrates at sites 1248
1250 contain mixed microbial and thermogenic methane,
and that the thermogenic gas component was generated at a
depth of 2 2.5 km. Although the microbial gas component
is around 80 85%, very little of it is being generated within
the hydrate stability zone. The molecular and isotopic
properties of gas hydrates at sites 1248 1250 are similar
to gas within horizon A, a permeable zone with high gas
saturation located beneath the southern Hydrate Ridge.
Apparently deep gas moves along horizon A to the base
of the hydrate stability zone [Milkov et al., 2005]. Trehu et
al. [2004b] used log and core data to infer gas saturations as
high as 90% in a coarse-grained turbidite sequence beneath
the HSZ. Gas buoyancy probably results in high gas
pressures. If the gas pressure exceeds the lithostatic stress,
then sediment failure (or slippage along preexisting faults)
will provide a path for gas to flow into the HSZ and to the
seafloor [Trehu et al., 2004b].
[76] Torres et al. [2004] simulated the observed chloride
concentration at site 1249 with an unsteady one-dimensional
transport reaction model. To reproduce the high chloride
concentrations at shallow depths, their model requires the

19 of 32

B01201

GARG ET AL.: FORMATION AND DISSOCIATION OF HYDRATES

B01201

Figure 9. Effect of capillary pressure on effective stress (equals vertical stress plus gas pressure;
negative in compression) distribution with depth at t = 10 Ma. The solid curve (Pc = 0) is for case 2. With
the exception of the capillary pressure, all the model parameters for cases 2a (Pc-low) and 2b (Pc-high)
are identical with those for case 2.
transport of methane in the gas phase from beneath the HSZ
to the seafloor. In addition, Torres et al. [2004] used depthdependent kinetic constants. At shallow depths, the force of
crystallization can overcome effective overburden stress and
hydrate growth takes place by particle displacement [Torres
et al., 2004]. Apparently, hydrate formation at site 1249 is
very rapid; seafloor observations indicate hydrate formation
rates of 100 mol m2 a1 [Torres et al., 2004]. The model
calculations of Torres et al. [2004] indicate that the hydrates
at site 1249 are probably younger than 1500 years. The age
estimate is rather uncertain.
[77] An alternate model for hydrate formation at site 1249
is presented by Liu and Flemings [2006, 2007]. They
proposed that hydrate formation is a self-equilibrating
process in marine settings where a large volume of free
gas is transported into the hydrate stability zone. Upflowing
free gas forms hydrate and elevates pore fluid salinity until
the fluid becomes too saline for additional hydrate formation. Three phases (brine, free gas and gas hydrate) coexist
along the entire path from the bottom of the HSZ to the
seafloor. Pore waters collected from interstitial water samples are usually diluted by hydrate dissociation during
retrieval. Liu and Flemings [2006] used a combination of
logged electrical resistivity, core-derived porosity and pore

fluid salinity to calculate the in situ salinity. A tabular listing


of corrected in situ salinities, plotted in Figure 7b of Liu and
Flemings [2006], has been provided to the authors by X. Liu
(personal communication, 2007). Liu and Flemings [2007]
developed a multiphase numerical model of hydrate formation in the marine environment. They applied the model to
investigate hydrate formation in conditions similar to those
at southern Hydrate Ridge [Liu and Flemings, 2007]; no
attempt was, however, made to reproduce the observed in
situ salinity profile for site 1249. We note in passing that the
numerical model of Liu and Flemings [2007] neglects the
latent heat of hydrate formation, and heat is transferred only
by conduction and advection.
[78] We have applied our numerical model to study
hydrate formation at site 1249. Model parameters for the
Hydrate Ridge simulations are given in Table 6. The
available organic carbon content is assumed to be negligible, and hydrate is formed by methane influx at the
sediment-basement boundary. The inflowing fluid is essentially all methane (99.99% by mass); the methane influx is
about 2.05 kg a1 m2 (128 mol a1 m2) at a temperature of 12.3C. Assuming a gas density of 78 kg m3 at
the sediment-basement boundary, this inflow rate yields an
upflow velocity of 26 mm a1. The methane mass influx

20 of 32

B01201

GARG ET AL.: FORMATION AND DISSOCIATION OF HYDRATES

Figure 10a. Effect of diffusivity on dissolved methane concentration at t = 10 Ma for cases 2 (D =


109 m2 s1) and 2c (D = 2  109 m2 s1).

Figure 10b. Effect of diffusivity on hydrate saturation at t = 10 Ma for cases 2 (D = 109 m2 s1) and
2c (D = 2  10 9 m2 s1).
21 of 32

B01201

B01201

GARG ET AL.: FORMATION AND DISSOCIATION OF HYDRATES

Figure 11a. Effect of the inclusion (case 2, solid line) or exclusion (case 2d, dashed line) of the sulfate
reduction zone on dissolved methane concentration at t = 10 Ma.

Figure 11b. Effect of the inclusion (case 2, solid line) or exclusion (case 2d, dashed line) of the sulfate
reduction zone on hydrate saturation at t = 10 Ma.
22 of 32

B01201

GARG ET AL.: FORMATION AND DISSOCIATION OF HYDRATES

B01201

B01201

Table 6. Model Parameters for the Hydrate Ridge Simulations


Parameter
Nominal water depth
Seafloor temperature
Sediment thermal conductivity
Conductive heat flux
Sediment grain density
Porosity at effective stress of 0
Porosity at effective stress of 1
Porosity parameter B, equation (24)
Permeability at porosity of 0.682
Parameter a, equation (25)
Parameter b, equation (25)
Parameter c, equation (25)
Irreducible brine saturation, Sbr
Critical gas saturation, Sgr
Parameter p, equation (26)
Brine relative permeability function frb, equation (26)
Gas relative permeability frg, equation (26)
Capillary pressure, Pc, equation (32)
Critical hydrate saturation for depression of Teq
Diffusivity of methane in brine, Dm

Units

Value

m
C
W m1 C1
W m2
kg m3

788 m
4.3
1
0.03
2650
0.682

see text
porosity parameters selected to fit the measured porosity
values [Trehu et al., 2003]

0.20
1.524
see note

MPa
m2

calculations were run for three values of


permeability(5  1016, 5  1015, and 5  1014)

2
0
0
0.30
0.10
2
see section 4.1

m2 s1

Pc-low
0.90
see note

Diffusivity of salt in brine, Ds


Diffusivity of sulfate in brine, Dsulf
Sedimentation rate

m2 s1
m2 s1
kg m2 s1

4  109

Available organic carbon ac, equation (6)


Seawater salinity

kg kg1

Seawater methane mass fraction


Seawater sulfate mass fraction
Total upward mass flux
Salt mass fraction
Methane mass fraction
Enthalpy
Basement elastic constant Vbm, equation (22)
Lagrangian spatial interval Dm
Initial number of grid blocks
Height of initial grid
Initial depth of basement/sediment interface
Time step
Calculation time interval

MPa

5

3.5  10
0.0027
6.5  108
106
0.9999
81.7
8  104
3000
50
133
926
1
5000

kJ kg1
m3 kg1
kg m2
m
m
years
years

was chosen so as to produce an outflow at the seafloor of


100 mol a1 m2 once quasi-steady state conditions are
attained. The interstitial methane velocity at the seafloor for
this outflow rate is over 45 mm a1. Both the upflow
velocity at the sediment-basement boundary and the interstitial velocity at the seafloor are about three orders of
magnitude greater than those for the Blake Ridge.
[79] The porosity parameters in Table 6 were selected so
as to fit the measured porosity values [see Trehu et al.,
2003, Table 13, chapter 8]. The initial number of grid blocks
is 50, and the height of computational grid at t = 0 is 133 m.
The sedimentation rate of 4  109 kg m2 s1 corresponds
to a deposition rate of 0.15 mm a1 [Torres et al., 2004]. The
height of the sediment column deposited during the computational time interval of 5000 years is less than 0.75 m, and
consequently. the number of grid blocks remains unchanged
during the calculation.
[80] Our numerical model accounts for the latent heat of
formation. Hydrate formation is an exothermic process, and
rapid hydrate formation can release a large amount of heat

brine is immobile for (Sb /(Sb + Sg )) < Sbr


gas is immobile for (Sg /(Sb + Sg )) < Sgr
(Sb /(Sb + Sg )) Sbr
(Sg /(Sb + Sg )) Sgr
see Figure 1
it is assumed that Dm = Ds = Dsulf = D;
calculations were performed for two
values of D (109 and 0.5  109)

0
0.035

kg s1 m2

Notes
from Table 1 of Torres et al. [2004]

corresponds to a deposition rate of


0.15 mm a1 [Torres et al., 2004]
these values also equal the initial pore fluid
salinity and methane content
(29 mM); initial pore fluid sulfate assumed to be 0
corresponds to a methane influx of 2.05 kg a1 m2
(128 mol a1 m2) at a temperature of 12.3C

see text
equals 5000 time steps

in a relatively short time. Although most of the released heat


gets transported away by conduction and advection, it can at
least locally elevate the sediment temperature. The heat of
hydrate formation is 450 kJ kg1 (section 2.5). Let us
assume that 10% of the pore brine (5% of bulk volume) is
converted to hydrate; with a hydrate density of 920 kg m3,
the evolved heat per cubic meter of bulk sediment volume
is given by
Evolved heat 0:05  920  450  103 J m3 21 MJ m3

The heat capacity of liquid-saturated sediment with 50%


porosity is 3.5 MJ m3 C1. If none of the evolved heat
gets transported away, the local temperature rise would be
6C. This simple calculation illustrates the importance of
accounting for latent heat especially when hydrate is being
formed rapidly. The temperature gradient at site 1249 is
55C km1 [Torres et al., 2004]. With a sediment thermal
conductivity of 1 W m1C1, the overall heat flux is
0.055 W m2. We assume that only a part (0.03 W m2) of

23 of 32

GARG ET AL.: FORMATION AND DISSOCIATION OF HYDRATES

B01201

B01201

Figure 12. Temperature, porosity, gas pressure (equals liquid pressure plus capillary pressure)
referenced to seafloor, and effective stress (equals vertical stress plus gas pressure; negative in
compression) distribution with depth at t = 0, 1500, 3000, and 5000 years (case 4). Solid dots in the
porosity plot denote the measurements reported by Trehu et al. [2003].

the overall heat flux is supplied by conductive flux at the


sediment-basement boundary; the remainder is derived from
methane flux at this boundary and from latent heat of
hydrate formation.
[81] A series of four Hydrate Ridge calculations was
performed with the following input values for absolute
permeability at seafloor and molecular diffusivity (assumed
to be the same for methane, salt and sulfate):
Case
Case
Case
Case

4
5
6
7

k
k
k
k

=
=
=
=

5
5
5
5






1015
1014
1016
1016

m2 ,
m2 ,
m2 ,
m2 ,

D
D
D
D

=
=
=
=

0.5
0.5
0.5
1.0






109
109
109
109

m2
m2
m2
m2

s1
s1
s1
s1

The results for case 4 are typical of all four cases and are
discussed in some detail.
[82] Temperature, porosity, gas pressure referenced to
seafloor, and effective stress (equals vertical stress plus
gas pressure, negative in compression) at t = 0, 1500,
3000, and 5000 years are displayed in Figure 12. The
temperature profile at t = 0 corresponds to a conductive
heat flux of 0.030 W m2. Later temperature profiles reflect
the combined effect of conductive and advective fluxes in
addition to heat evolution accompanying hydrate formation.

It is significant that unlike the Blake Ridge examples, the


temperature profile is not linear. Solid dots in the porosity plot
denote the measurements [see Trehu et al., 2003, Table 13,
chapter 8]; the porosity curve for t = 0 represents a fit to
porosity data. It is apparent from Figure 12 that porosity
increases and effective stress becomes less compressive
between t = 0 and 1500 years. The porosity increase causes
a corresponding increase in the thickness of the sediment
column from 132.7 m at t = 0 to a maximum value of 142.8 m
at t = 2000 years (Figure 13 and Table 7). (The increase in
sediment porosity and hence thickness of the sediment
column is a consequence of the assumed relationship
(equation (24)) between porosity and effective stress. It is
likely that the sediments are much less compliant during
unloading (effective stress becoming less compressive) than
during loading (effective stress becoming more compressive); in this case, the increase in porosity (and sediment
column thickness) would be correspondingly smaller. A
smaller increase in porosity will be accompanied by a larger
fluid pressure gradient.)
[83] Between t = 0 and 1500 years, the gas pressure
gradient increases by about 5%, and the lithostatic stress
gradient declines by 6%. Decline in the lithostatic stress
gradient is a direct result of the expansion of the sediment

24 of 32

B01201

GARG ET AL.: FORMATION AND DISSOCIATION OF HYDRATES

B01201

Figure 13. Sediment column height (case 4). The column height increases from 132.7 m at t = 0 to a
maximum of 142.8 m at t = 2000 years and then slowly declines.
column. As mentioned earlier, the upward fluid velocity is
quite high at site 1249. The sediment column accommodates
the high fluid flux by an increase in the fluid pressure
gradient, and a consequent decrease in the compressive
effective stress which in turn leads to an increase in
sediment porosity. The compressive effective stress at the
base of the sediment column at t = 5000 years is only
0.73 MPa. As noted in Table 6, Pc-low capillary pressure
curve (Figure 1) was used for all the Hydrate Ridge cases.
Use of the Pc-high curve for capillary pressure will yield a
much smaller value for the effective stress at the BHSZ. It is
therefore conceivable that gas buildup in the sediments
below the HSZ could result in sediment failure and/or fault
activation.
[84] Free gas saturation (as a fraction of the pore space),
upward fluid (gas plus liquid) mass flux, gas hydrate
saturation in the pore space, and in situ brine chlorinity at
t = 1500, 3000, and 5000 years are displayed in Figure 14.

Although free gas saturation as a fraction of the pore space


in the HSZ at t = 3000 and 5000 years is less than 10%,
the free gas is mobile since the normalized gas saturation
(=Sg/(Sb + Sg)) exceeds the critical gas saturation. As
mentioned earlier, the fluid influx at the sediment-basement
boundary is essentially all methane (99.99% by mass). The
fluid flux at the seafloor contains both methane and liquid
brine. Prior to about 1750 years, the methane flux at the
seafloor is essentially zero; it then increases to 92 mol a1
at 3000 years and 101 mol a1 at 5000 years. After
3000 years, changes in fluid (liquid brine plus gas) flux at
seafloor are quite small. After an initial transient, the
bottom of the hydrate stability zone starts shoaling and
reaches a minimum depth of 84 mbsf at about 3000 years,
and subsequently begins to deepen (Figure 15). Changes in
the depth of the hydrate stability zone are a result of (1) an
increase in temperature due to the release of latent heat that
accompanies hydrate formation and (2) the evolution of the

Table 7. Gas Pressure Gradient, Lithostatic Stress Gradient, Sediment Column Thickness, and Methane
Outflow at Seafloor at t = 0, 1500, 3000, and 5000 years (Case 4)
Units
Gas pressure gradient
Lithostatic stress gradient
Sediment column thickness
Methane outflow at seafloor

1

kPa m
kPa m1
m
mol a1

t=0

t = 1500

t = 3000

t = 5000

10.12
16.89
132.7
0

10.66
15.93
141.3
0

10.62
15.86
141.9
92

10.62
15.88
141.5
101

25 of 32

B01201

GARG ET AL.: FORMATION AND DISSOCIATION OF HYDRATES

B01201

Figure 14. Free gas saturation, upward fluid flux, hydrate saturation (fraction of pore space), and in situ
pore fluid salinity (relative to seawater salinity) at t = 1500, 3000, and 5000 years (case 4). Solid dots in
the salinity plot denote corrected chlorinity values from Figure 7b of Liu and Flemings [2006].

in situ pore fluid salinity. Hydrate formation results in an


increase in pore fluid salinity, which in turns tends to
inhibit further hydrate formation by shifting the equilibrium
temperature to lower values. Most of the rapid hydrate
formation takes place prior to about 3000 years, and
subsequent increases in hydrate saturation are relatively
small. Thus the latent heat released by hydrate formation
slows down considerably after 3000 years; this results in a
temperature decline (Figure 12) and a deepening of the
bottom of the HSZ. The solid dots in the salinity plot
(Figure 14) correspond to the corrected chlorinity values
plotted in Figure 7b of Liu and Flemings [2006]. Taking
into account the considerable scatter in the corrected
salinity values, it appears that the computed salinity profile
at t = 3000 years provides a good fit. The computed hydrate
saturation at 3000 years exceeds 90% at shallow depths
(less than 30 m), and then declines rapidly with depth.
Corrected chlorinity values range up to five times the
seawater chlorinity; these high chlorinity values suggest
hydrate saturations of at least 80%. Since advection and
molecular diffusion will tend to smooth out gradients in
salinity, the hydrate saturation will need to be greater than
80% in order to generate a relative salinity equal to 5.
[85] Computed hydrate and in situ pore fluid salinity
distributions for all four Hydrate Ridge simulations (cases

4, 5, 6 and 7) at t = 3000 and 5000 years are displayed in


Figures 16a, 16b, 17a, and 17b. Except for the high-permeability case 5 (k = 5  1014 m2), all the other three cases
yield a similar hydrate distribution at t = 3000 years (Figure
16a). Below 40 mbsf, the computed salinity distribution at
3000 years (Figure 16b) for all four cases is in good
agreement with the corrected chlorinity values [Liu and
Flemings, 2006]; at shallower depths, case 4 gives the
best fit. The two low-permeability cases 6 and 7 (k = 5 
1016 m2) yield almost the same hydrate distribution at t =
5000 years (Figure 17a). An increase in permeability (compare cases 6 7 with case 4 and particularly case 5) results in
increasing the thickness of the high-saturation hydrate zone
(Figure 17a). None of the computed salinity profiles at t =
5000 years (Figure 17b) are in agreement with the corrected
chlorinity values below a depth of 50 mbsf. Above the
latter depth, both cases 4 (k = 5  1015 m2, D = 0.5 
109 m2 s1) and 7 (k = 5  1016 m2, D = 109 m2 s1)
provide an adequate fit to the corrected chlorinity values.
On the basis of results presented in Figures 16a and 17a, we
conclude that the computed hydrate and salinity distributions at t = 3000 years for case 4 best represent the
conditions at site 1249. The bottom of the hydrate stability
zone is relatively shallow (84 mbsf) at 3000 years; the
bottom is deeper at earlier and later times (Figure 15). The

26 of 32

B01201

GARG ET AL.: FORMATION AND DISSOCIATION OF HYDRATES

Figure 15. Location of the bottom of the hydrate stability zone (case 4). After an initial transient, the
bottom shoals to a minimum depth of 84 mbsf at t  3000 years and subsequently deepens. The stair
step appearance is due to spatial discretization.

Figure 16a. Hydrate distribution at t = 3000 years for cases 4, 5, 6, and 7. Except for the highpermeability case 5, the computed hydrate distribution is similar for all the other cases.
27 of 32

B01201

B01201

GARG ET AL.: FORMATION AND DISSOCIATION OF HYDRATES

B01201

Figure 16b. In situ pore fluid salinity distribution at t = 3000 years for cases 4, 5, 6, and 7. Solid dots
denote the corrected chlorinity values from Figure 7b of Liu and Flemings [2006]. Below 40mbsf, all
the cases produce an adequate fit to the corrected chlorinity values. Above 40 mbsf, case 4 provides the
best fit.
good agreement between the computed (t = 3000 years) and
corrected chlorinity values implies that the bottom of the
hydrate zone at site 1249 is likely located at about 84 mbsf.
Our results suggest that the hydrates at site 1249 were
created over the past 3000 years. Taking into consideration
the uncertainties inherent in any age determination, this age
estimate is in fair agreement with that (1500 years)
predicted by Torres et al. [2004].

5. Concluding Remarks
[86] We have developed a one-dimensional numerical
model to simulate hydrate formation, burial and decomposition, reformation and distribution in the marine environment. The model includes a description of the burial history
of deep marine sediments and associated phenomenon such
as sediment compaction and consequent reduction in porosity and permeability, fluid and heat flux, the sulfate
reduction zone, and time evolution of pressure and temperature in the subsurface. In situ generation of biogenic
methane from buried organic carbon is treated using the
exponential source function of Davie and Buffett [2001].
Multiphase (i.e., liquid brine with dissolved brine, free gas,
and gas hydrate) flow through a deformable porous matrix
is an essential part of the mathematical formulation. Because of sedimentation and burial, the spatial locations of
both the upper (seafloor) and lower (basement-sediment
interface) boundaries vary with time. To treat these moving
boundaries, a Lagrangian reference frame (with the coordi-

nate system attached to the deforming sediment) is adopted.


The numerical model automatically adds additional grid
blocks in response to continued sedimentation.
[87] The model has been applied to study the hydrate
distributions at Blake Ridge (site 997) and Hydrate Ridge
(site 1249). For the Blake Ridge setting, our results suggest
that the observed hydrate saturation (4 to 7% of the pore
space) can be explained either by gas upflow from the
sediment-basement boundary or by a combination of in situ
methane generation from organic carbon and gas upflow.
Parametric calculations based on the Blake Ridge examples
indicate that the computed hydrate saturation is influenced
by critical gas saturation, diffusivities of dissolved chemical
species (methane, salt, and sulfate) in the pore liquids, and
the presence of the sulfate reduction zone in addition to
pressure, temperature, and methane supply. The computed
effective stress at the BHSZ is relatively large and compressive; this would appear to preclude sediment failure at
site 997 (Blake Ridge). Hydrate Ridge is an accretionary
ridge, with site 1249 located on the southern summit. The
site is characterized by active gas venting and very high
hydrate formation rates. Our simulations indicate that the
bottom of the hydrate zone at site 1249 is rather shallow
(84 mbsf), and that the entire hydrate zone lies along the
three-phase (hydrate, liquid, free gas) equilibrium curve.
The Hydrate Ridge simulations highlight the role played by
the latent heat of hydrate formation and pore fluid salinity
distribution in creating the three-phase hydrate stability
zone, and in moving the bottom of the hydrate stability

28 of 32

B01201

GARG ET AL.: FORMATION AND DISSOCIATION OF HYDRATES

Figure 17a. Hydrate distribution at t = 5000 years for cases 4, 5, 6, and 7. For the two low-permeability
cases 6 and 7, the computed hydrate distribution is similar. An increase in permeability (compare cases 4
and 5 to case 6) causes the high-saturation hydrate zone to extend deeper below the seafloor.

Figure 17b. In situ pore fluid salinity distribution at t = 5000 years for cases 4, 5, 6, and 7. Solid dots
denote the corrected chlorinity values from Figure 7b of Liu and Flemings [2006]. Below 50mbsf, none
of the cases yield an adequate fit to the corrected chlorinity values. Above 50 mbsf, cases 4 and 7
provide the best fit.
29 of 32

B01201

GARG ET AL.: FORMATION AND DISSOCIATION OF HYDRATES

B01201

zone up and down. Unlike the Blake Ridge example, the


computed effective stress at the BHSZ for site 1249 is rather
small, and gas buildup can result in sediment failure.
[88] In summary, the numerical model reported herein can
be used to investigate the distribution of methane hydrates
in marine settings. Work is currently underway to apply the
numerical model to investigate the hydrate distribution
observed in various Nankai Trough boreholes, and will be
reported in a separate publication.

Notation
a,b,c
a0, b0, c0, d0
B
ch,cp
cb, cw
cr, cvp, cph
C
C0
Cc
Dm, Ds, Dsulf
g
Eb, Eg, Eh, Ep, Er,
frb, frg
fsulf
h
hb,hg,hh,hp,hr ,hw
DH
k
k0
krb, krg
m
mmax(t)
M_
Mg, Mb
M0
n
Ng
Nh
p
P
Pb, Pg, Ph, Pp, Pr
Pc
P*

material constants, equation (25).


constants, equation (41).
a material constant, equation (24).
isothermal compressibility of gas hydrate and salt precipitate, equation (33).
isothermal compressibility of brine and
pure water, equation (36).
specific heat for sediments, salt precipitate, and gas hydrate.
= 4 Cc/3.
value of C at seafloor, equation (20).
mass of available organic carbon per
unit volume, equation (6).
diffusivity of methane, salt, and sulfate
ion in brine.
acceleration due to gravity.
brine, gas, hydrate, precipitate, and
sediment internal energy.
relativity permeability functions, equation (26).
sulfate mass fraction of the salt content
of brine.
enthalpy of pore contents, equation (30).
brine, gas, hydrate, solid salt, sediment, and pure water enthalpy.
heat of hydrate dissociation.
absolute permeability.
absolute permeability at 8 = 80.
relative permeability for brine and gas.
Lagrangian coordinate.
seafloor (Lagrangian coordinate).
rate of conversion of organic carbon
plus hydrogen to methane.
gas and brine mass flux.
total (gas plus brine) mass influx at the
sediment-basement boundary.
hydration number, equation (8).
number of moles of gas per kilogram.
number of moles of hydrate per
kilogram.
a constant, equation (26).
thermodynamic pressure (=Pb = Ph =
Pp).
brine, gas, hydrate, salt precipitate, and
sediment pressure.
capillary pressure (=Pg  P b),
equation (32).
a constant, equation (50).

30 of 32

B01201

Q0 heat flux at the sediment basement


boundary.
R universal gas constant.
Sb, Sg, Sh, Sp pore volume fractions occupied by
brine, gas, hydrate and precipitate.
S* a reduced saturation, equation (50).
t time.
T temperature.
Teq hydrate equilibrium temperature.
TK absolute temperature (=T + 273.15).
ur, ub, ug sediment, brine, and gas velocity.
Vbm a material constant, equation (26).
z Eulerian (spatial) coordinate, positive
upward.
z0 (t) location of sediment-basement
boundary.
zmax (t) seafloor location (Eulerian
coordinate).
Z gas compressibility factor.
a = 4 ac/3.
ac mass ratio of available organic carbon
to that of sediment, equation (6).
am, as, aw mass fraction of methane, salt, and
water of pore contents, equations (28)
and (29).
aH methane solubility in the presence of
hydrate, equation (40).
aij mass fraction of chem. species j ( j =
m, w, s) in phase i (i = b, g, h, p),
equation (8).
abm max maximum mass fraction of methane in
brine, equation (39).
abs max maximum salt mass fraction in brine,
equation (38).
bh, b p volumetric thermal expansion coefficient of hydrate and salt precipitate,
equation (33).
d a constant, equation (50).
q(t) sediment mass deposition rate at seafloor, equation (23).
k thermal conductivity.
l rate constant for methanogenesis,
equation (20).
n b, n g kinematic viscosity for brine and gas.
rb, rg, rh, rp density of brine, gas, hydrate, and
solid precipitate.
rr sediment grain density.
r0h ,r0p density of hydrate and salt precipitate
at reference conditions, equation (33).
rb0, rw0 density of brine and pure water at
P = 0 and T, equation (36).
se effective stress (=sz + Pg), positive in
tension.
sr intrinsic sediment stress (=P r ),
positive in tension, equation (19).
sz vertical stress, positive in tension.
8 porosity.
80 porosity at zero effective stress (i.e., at
seafloor), equation (24).
81 porosity in the limit as the effective
stress se approaches 1, equation (24).

GARG ET AL.: FORMATION AND DISSOCIATION OF HYDRATES

B01201

Subscripts
b
g
h
m
p
r
s
w

brine.
gas.
hydrate
methane.
precipitate.
sediment.
salt.
water.

[89] Acknowledgments. This work was performed as a part of the


Japanese MH21 project. Carolyn Ruppel, an anonymous reviewer, and
Associate Editor Ingo Pecher provided constructive and helpful comments on earlier versions of the manuscript. George Hirasaki and Xiaoli
Liu kindly made available prepublication copies of their papers. In
addition, X. Liu provided tabular values of corrected pore fluid salinity
for site 1249.

References
Baba, K., and A. Katoh (2004), Application of a simulation model for the
formation of Methane hydrate to the Nankai Trough and the Blake Ridge:
Natural examples of two end-member cases, Resour. Geol., 54(1), 125
135.
Bakker, R. (1998), Improvements in clathrate modeling II: The H2O-CO2
CH4 N2 C2H6 fluid system, in Gas Hydrates: Relevance to World Margin Stability and Climate Change, edited by J.-P. Henriet, and J. Mienert,
Geol. Soc. Spec. Publ., 137, 75 105.
Bangs, N. L. B., D. S. Sawyer, and X. Golovchenko (1993), Free gas at the
base of the gas hydrate zone in the vicinity of the Chile triple junction,
Geology, 21, 905 908.
Bear, J. (1972), Dynamics of Fluids in Porous Media, Elsevier, New York.
Bennion, B., and S. Bachu (2005), Relative permeability characteristics for
supercritical CO2 displacing water in a variety of potential sequestration
zones in the western Canada sedimentary basin, paper SPE 95547 presented at 2005 SPE Annual Technical Conference, Soc. of Pet. Eng.,
Dallas, Tex.
Bhatnagar, G., W. G. Chapman, G. R. Dickens, B. Dugan, and G. J. Hirasaki
(2007), Generalization of gas hydrate distribution and saturation in marine
sediments by scaling of thermodynamic and transport processes, Am. J.
Sci., 307, 861 900, doi:10.2475/06.2007.01.
Brill, J. P., and H. D. Beggs (1975), Two phase flow in pipes, lecture notes,
Univ. of Tulsa, Tulsa, Okla.
Brown, K. M., and N. L. Bangs (1995), Thermal regime of the Chile Triple
Junction: Constraints provided by downhole temperature measurements
and distribution of gas hydrate, Proc. Ocean Drill. Program Sci. Results,
141, 259 275.
Buffett, B., and D. Archer (2004), Global inventory of methane clathrate:
Sensitivity to changes in the deep ocean, Earth Planet. Sci. Lett., 227,
185 199.
Claypool, G. E., and I. R. Kaplan (1974), The origin and distribution of
methane in marine sediments, in Natural Gases in Marine Sediments,
edited by I. R. Kaplan, pp. 99 139, Plenum, New York.
Claypool, G. E., and K. A. Kvenvolden (1983), Methane and other hydrocarbon gases in marine sediment, Annu. Rev. Earth Planet. Sci., 11, 299
327.
Clennell, M. B., M. Hovland, J. S. Booth, P. Henry, and W. J. Winters
(1999), Formation of natural gas hydrates in marine sediments: 1. Conceptual model of gas hydrate growth conditioned by host sediment properties, J. Geophys. Res., 104(B10), 22,985 23,003.
Clennell, M. B., P. Henry, M. Hovland, J. S. Booth, W. J. Winters, and
M. Thomas (2000), Formation of natural gas hydrates in marine sediments: Gas hydrate growth and stability conditioned by host sediment
properties, in Gas Hydrates: Challenges for the Future, edited by G. D.
Holder, and P. R. Bishnoi, Ann. N. Y. Acad. Sci., 912, 887 896.
Collett, T. S., and J. Ladd (2000), Detection of gas hydrate with downhole
logs and assessment of gas hydrate concentrations (saturations) and gas
volumes on the Blake Ridge with electrical resistivity logs, Proc. Ocean
Drill. Program Sci. Results, 164, 179 191.
Culberson, O. L., and J. J. McKetta Jr. (1951), Phase equilibria in hydrocarbon-water systems III: The solubility of methane in water at pressures
to 10,000 psia, Pet. Trans. AIME, 192, 223 226.
Davie, M. K., and B. A. Buffett (2001), A numerical model for the formation of gas hydrate below the seafloor, J. Geophys. Res., 106(B1), 497
514.

B01201

Davie, M. K., and B. A. Buffett (2003a), Sources of methane for marine gas
hydrate: Inferences from a comparison of observations and numerical
models, Earth Planet. Sci. Lett., 206, 51 63.
Davie, M. K., and B. A. Buffett (2003b), A steady state model for marine
hydrate formation: Constraints on methane supply from pore water sulfate profiles, J. Geophys. Res., 108(B10), 2495, doi:10.1029/
2002JB002300.
Davie, M. K., O. Y. Zatsepina, and B. A. Buffett (2004), Methane solubility
in marine hydrate environments, Mar. Geol., 203, 177 184.
Dins, F. (1956), Thermodynamic Functions of Gases, Butterworth Sci.,
London.
Duan, Z., and S. Mao (2006), A thermodynamic model for calculating
methane solubility, density and gas phase composition of methane-bearing aqueous fluids from 273 to 523 K and from 0 to 2000 bar, Geochim.
Cosmochim. Acta, 70, 3369 3386.
Duan, Z., N. Moller, J. Greenberg, and J. H. Weare (1992), The prediction
of methane solubility in natural waters to high ionic strength from 0 to
250C and from 0 to 1600 bar, Geochim. Cosmochim. Acta, 56, 1451
1460.
Egeberg, P. K., and G. R. Dickens (1999), Thermodynamic and pore water
halogen constraints on hydrate distribution at ODP site 997 (Blake
Ridge), Chem. Geol., 153, 53 79.
Flemings, P. B., X. Liu, and W. J. Winters (2003), Critical pressure and
multiphase flow in Blake Ridge gas hydrates, Geology, 31(12), 1057
1060.
Garg, S. K. (1987), On balance laws for fluid-saturated porous media,
Mech. Mater., 6, 219 232.
Garg, S. K., J. W. Pritchett, M. H. Rice, and T. D. Riney (1977), U. S. gulf
coast geopressured geothermal reservoir simulation, Rep. SSS-R-77-3147,
Sci. Appl. Int. Corp., San Diego, Calif.
Ginsburg, G., V. Soloviev, T. Matveeva, and I. Andreeva (2000), Sediment
grain-size control on gas hydrate presence, sites 994, 995, and 997, Proc.
Ocean Drill. Program Sci. Results, 164, 237 245.
Hesse, R., and W. E. Harrison (1981), Gas hydrates (clathrates) causing
pore-water freshening and oxygen isotope fractionation in deep-water
sedimentary sections of terrigenous continental margins, Earth Planet.
Sci. Lett., 55, 453 462.
Holbrook, W. S., H. Hoskins, W. T. Wood, R. A. Stephen, D. Lizarralde,
and Leg 164 Science Party (1996), Methane hydrate and free gas on the
Blake Ridge from vertical seismic profiling, Science, 273, 1840 1843.
Hornbach, M. J., D. M. Saffer, and W. S. Holbrook (2004), Critically
pressured free-gas reservoirs below gas-hydrate provinces, Nature, 427,
142 144.
Hyndman, R. D., and E. E. Davis (1992), A mechanism for the formation of
methane hydrate and seafloor bottom-simulating reflectors by vertical
fluid expulsion, J. Geophys. Res., 97(B5), 7025 7041.
Kraemer, L. M., R. M. Owen, and G. R. Dickens (2000), Lithology of the
upper gas hydrate zone, Blake Outer Ridge: A link between diatoms,
porosity and gas hydrate, Proc. Ocean Drill. Program Sci. Results,
164, 229 236.
Kleinberg, R. L., C. Flaum, D. D. Griffin, P. G. Brewer, G. E. Malby, E. T.
Pelzer, and J. P. Yesinowski (2003), Deep sea NMR: Methane hydrate
growth habit in porous media and its relationship to hydraulic permeability, deposit accumulation, and submarine slope stability, J. Geophys.
Res., 108(B10), 2508, doi:10.1029/2003JB002389.
Kvenvolden, K. A. (1993), Gas hydrates: Geological perspective and global
change, Rev. Geophys., 31(2), 173 187.
Liu, X., and P. B. Flemings (2006), Passing gas through the hydrate stability zone at southern Hydrate Ridge, offshore Oregon, Earth Planet. Sci.
Lett., 241, 211 226.
Liu, X., and P. B. Flemings (2007), Dynamic multiphase flow model of
hydrate formation in marine sediments, J. Geophys. Res., 112(B3),
B03101, doi:10.1029/2005JB004227.
Matsumoto, R., T. Uchida, A. Waseda, T. Uchida, S. Takeya, T. Hirano,
K. Yamada, Y. Maeda, and T. Okui (2000), Occurrence, structure, and
composition of natural gas hydrate recovered from the Blake Ridge, northwest Atlantic, Proc. Ocean Drill. Program Sci. Results, 164, 3 12.
Milkov, A. V., G. R. Dickens, G. E. Claypool, Y.-J. Lee, W. S. Borowski,
M. E. Torres, W. Xu, H. Tomaru, A. M. Trehu, and P. Schultheiss (2004),
Co-existence of gas hydrate, free gas, and brine within the regional gas
hydrate stability zone at Hydrate Ridge (Oregon margin): Evidence from
prolonged degassing of a pressurized core, Earth Planet. Sci. Lett., 222,
829 843.
Milkov, A. V., G. E. Claypool, Y.-J. Lee, and R. Sassen (2005), Gas hydrate
systems at Hydrate Ridge offshore Oregon inferred from molecular and
isotopic properties of hydrate-bound and void gases, Geochim. Cosmochim. Acta, 69(4), 1007 1026, doi:10.1016/j.gca.2004.08.021.
Moridis, G. J. (2002), Numerical studies of gas production from methane
hydrates, paper SPE 75691 presented at SPE Gas Technology Symposium, Soc. of Pet. Eng., Calgary, Alberta, Canada.

31 of 32

B01201

GARG ET AL.: FORMATION AND DISSOCIATION OF HYDRATES

Morland, L. W., R. Foulser, and S. K. Garg (2004), Mixture theory for a


fluid-saturated isotropic elastic matrix, Int. J. Geomech., 4(3), 207 215,
doi:10.1061/ (ASCE)1532-3641 (2004)4:3 (207).
Nimblett, J., and C. Ruppel (2003), Permeability evolution during the formation of gas hydrates in marine sediments, J. Geophys. Res., 108(B9),
2420, doi:10.1029/2001JB001650.
Paull, C. K., et al. (1996), Proceedings Ocean Drilling Program Initial
Reports, vol. 164, Ocean Drill. Program, College Station, Tex.
Paull, C. K., W. S. Borowski, N. M. Rodriguez, and ODP Leg 164 Shipboard Scientific Party (1998), Marine gas hydrate inventory: Preliminary
results of ODP Leg 164 and implications for gas venting and slumping
associated with the Blake Ridge gas hydrate field, in Gas Hydrates:
Relevance to World Margin Stability and Climate Change, edited by
J.-P. Henriet and J. Mienert and Mienert, Geol. Soc. Spec. Publ., 137,
153 160.
Pritchett, J. W., S. K. Garg, M. H. Rice, and T. D. Riney (1979), Geopressured reservoir simulation, Rep. SSS-R-79-4022, Sci. Appl. Int. Corp.,
San Diego, Calif.
Pruess, K., and J. Garcia (2002), Solutions of test problems for disposal of
CO2 in saline aquifers, Rep. LBNL-51812, Lawrence Berkeley Natl. Lab.,
Berkeley, Calif.
Rempel, A., and B. A. Buffett (1997), Formation and accumulation of gas
hydrate in porous media, J. Geophys. Res., 102(B5), 10,151 10,164.
Rempel, A. W., and B. A. Buffett (1998), Mathematical models of gas
hydrate accumulation, in Gas Hydrates: Relevance to World Margin
Stability and Climate Change, edited by J.-P. Henriet and J. Mienert
and Mienert, Geol. Soc. Spec. Publ., 137, 63 74.
Riedel, M., et al. (2006), Gas hydrate transect across Northern Cascadia
Margin, Eos Trans. AGU, 87(33) , 325, 330, and 332.
Ruppel, C. (1997), Anomalously cold temperatures observed at the base of
the gas stability zone on the U.S. Atlantic passive margin, Geology,
25(8), 699 702.
Ruppel, C., G. R. Dickens, D. G. Castellini, W. Gilhooly, and D. Lizarralde
(2005), Heat and salt inhibition of gas hydrate formation in the northern
Gulf of Mexico, Geophys. Res. Lett., 32, L04605, doi:10.1029/
2004GL021909.
Schowalter, T. T. (1979), Mechanics of secondary hydrocarbon migration
and entrapment, Am. Assoc. Pet. Geol. Bull., 63(5), 723 760.
Servio, P., and P. Englezos (2002), Measurement of dissolved methane in
water in equilibrium with its hydrate, J. Chem. Eng. Data, 47, 87 90.

B01201

Sloan, E. D. (1998), Clathrate Hydrates of Natural Gases, Marcel Dekker,


New York.
Torres, M. E., K. Wallmann, A. M. Trehu, G. Bohrmann, W. S. Borowski,
and H. Tomaru (2004), Gas hydrate growth, methane transport, and
chloride enrichment at the southern summit of Hydrate Ridge, Cascadia
margin off Oregon, Earth Planet. Sci. Lett., 226, 225 241.
Trehu, A. M., et al. (2003), Proceedings Ocean Drilling Program Initial
Reports, vol. 204, Ocean Drill. Program, College Station, Tex.,
doi:10.2973/odp.proc.ir.204.
Trehu, A. M., et al. (2004a), Three-dimensional distribution of gas hydrate
beneath southern Hydrate Ridge: Constraints from ODP Leg 204, Earth
Planet. Sci. Lett., 222, 845 862.
Trehu, A. M., P. B. Flemings, N. L. Bangs, J. Chevallier, E. Garcia, J. E.
Johnson, C.-S. Liu, X. Liu, M. Riedel, and M. E. Torres (2004b), Feeding
methane vents and gas hydrate deposits at south Hydrate Ridge, Geophys.
Res. Lett., 31, L23310, doi:10.1029/2004GL021286.
van Genuchten, M. T. (1980), A closed-form equation for predicting the
hydraulic conductivity of unsaturated soils, Soil Sci. Soc. Am. J., 44,
892 898.
Waseda, A. (1998), Original carbon content, bacterial methanogenesis, and
accumulation processes of hydrates in marine sediments, Geochem. J.,
32, 143 157.
Xu, W. (2004), Modeling dynamic marine gas hydrate systems, Am.
Mineral., 89, 1271 1279.
Xu, W., and C. Ruppel (1999), Predicting the occurrence, distribution, and
evolution of methane gas hydrate in porous marine sediments, J. Geophys. Res., 104(B3), 5081 5095.
Zatsepina, O. Y., and B. A. Buffett (1997), Phase equilibrium of gas hydrate: Implications for the formation of hydrate in the deep sea floor,
Geophys. Res. Lett., 24(13), 1567 1570.


K. Baba, Research Center, Japan Petroleum Exploration Co. Ltd., 1-2-1,


Hamada, Mihama-ku, Chiba, 261-0025, Japan.
T. Fujii, Japan Oil, Gas and Metals National Corporation, 1-2-2 Hamada,
Mihama-ku, Chiba, 261-0025, Japan.
S. K. Garg and J. W. Pritchett, Science Applications International
Corporation, 10260 Campus Point Drive, Mail Stop A-3, San Diego, CA
92121, USA. (gargs@saic.com)
A. Katoh, JGI Inc, Meikei Building, 1-5-21, Otsuka, Bunkyo-ku, Tokyo,
112-0012, Japan.

32 of 32

S-ar putea să vă placă și