Você está na página 1de 16

SPE 113496

Upscaling of Capillary Trapping Under Gravity Override:


Application to CO
2
Sequestration in Aquifers
R. Juanes, SPE, and C.W. MacMinn, Massachusetts Institute of Technology
Abstract
We present a sharp-interface mathematical model of CO
2
migration in saline aquifers, which accounts for gravity override and
capillary trapping. The major differences with respect to previous investigations is that we model the shape of the plume during
the injection period, we account for regional groundwater ow during the post-injection period, and we introduce rigorously
the reduction in water mobility due to trapping of the CO
2
.
The model leads to a nonlinear advectiondiffusion equation, in which the diffusive term is due to buoyancy forces, not
physical diffusion. The three key dimensionless groups are the mobility ratio between the injected and initial uids (M), the
gravity number (N
g
), and a newly dened trapping coefcient (). For the case of interest in geological CO
2
storage, in which
the mobility ratio is much smaller than 1, the solution is largely insensitive to the value of the gravity number. Under these
conditions, the mathematical model can be simplied to a hyperbolic model.
We present a complete analytical solution to the hyperbolic model that includes the injection, early post-injection, and late
post-injection periods. Despite the fact that the solution involves the interaction of a sharp imbibition front with a drainage
rarefaction front, it admits a closed-form expression. The main outcome of the analytical developments presented here is
a formula that predicts the ultimate footprint on the CO
2
plume, and the time scale required for complete trapping. Both
quantities depend strongly on the shape of the plume at the end of the injection period, which mustthereforebe modeled.
A second application of the analytical solution is a formulation for upscaling the capillary trapping coefcient from the lab-
oratory scale to the basin scale. Explicit expressions are given for the megascopic and gridblock-effective trapping coefcients,
as functions of the local trapping coefcient, the mobility ratio, and the grid resolution. Although the expressions derived are
based on a one-dimensional sharp-interface model, we anticipate that they will have broader applicability to injection scenarios
with unfavorable mobility ratio and dominated by gravity override.
Introduction and Motivation
Deep saline aquifers are attractive geological formations for the injection and long-term storage of CO
2
ipcc-co2. Even if
injected as a supercritical uid, or dense gas (typically, at a depth of 800 m or more depending on the surface temperature and
geothermal gradient), the CO
2
is buoyant with respect to the formation brine. A number of trapping mechanisms act to prevent
the migration of the buoyant CO
2
back to the surface, and these include: (1) hydrodynamic trapping: the buoyant CO
2
remains
as a mobile uid but is prevented from owing back to the surface by an impermeable cap rock (Bachu et al., 1994); (4) capillary
trapping: disconnection of the CO
2
phase into an immobile (trapped) fraction (Flett et al., 2004; Kumar et al., 2005; Juanes
et al., 2006); (3) dissolution trapping: dissolution of the CO
2
in the brine (Pruess and Garca, 2002), possibly enhanced by
gravity instabilities due to the larger density of the brineCO
2
liquid mixture (Ennis-King and Paterson, 2005; Ennis-King
et al., 2005; Riaz et al., 2006); and (4) mineral trapping: geochemical binding to the rock due to mineral precipitation (Gunter
et al., 1997; Pruess et al., 2003; Xu et al., 2003). Because the time scales associated with these mechanisms are believed to be
quite different (t
hydrodyn
t
capil
t
dissol
t
miner
), it is justied to neglect dissolution and mineral trapping in the study of CO
2
migration during the injection and post-injection periodsprecisely when the risk for leakage is higher.
The basis for capillary trapping in aquifer CO
2
storage can be summarized as follows (Juanes et al., 2006). In saline aquifers,
predominantly water-wet, snap-off is the dominant trapping mechanism at the pore scale (Lenormand et al., 1983; Valvatne and
Blunt, 2004). Capillary trapping of the non-wetting CO
2
phase occurs when water invades the pore-space previously occupied
by the CO
2
. During the injection of CO
2
in the geologic formation, the gas saturation increases in a drainage-like process.
Once the injection stops, the CO
2
continues to migrate in response to buoyancy and regional groundwater ow. At the leading
edge of the CO
2
plume, gas continues to displace water in a drainage process (increasing gas saturation), whilst at the trailing
Copyright 2008, Society of Petroleum Engineers
This paper was prepared for presentation at the 2008 SPE/DOE Improved Oil Recovery Symposium held in Tulsa, Oklahoma, U.S.A., 1923April2008.
This paper was selected for presentation by an SPE program committee following review of information contained in an abstract submitted by the author(s). Contents of the paper have not been
reviewed by the Society of Petroleum Engineers and are subject to correction by the author(s). The material does not necessarily reflect any position of the Society of Petroleum Engineers, its
officers, or members. Electronic reproduction, distribution, or storage of any part of this paper without the written consent of the Society of Petroleum Engineers is prohibited. Permission to
reproduce in print is restricted to an abstract of not more than 300 words; illustrations may not be copied. The abstract must contain conspicuous acknowledgment of SPE copyright.

2 SPE 113496
mobile gas
residual gas
(immobile)
CO
2
injection
low
-p
erm
layer
Fig. 1 Schematic of the trail of residual CO
2
(light blue) that is left behind due to snap-off as the plume of mobile CO
2
(dark blue)
migrates upwards during the post-injection period. The inset is a conceptual diagram of the CO
2
trapped as small bubbles (white) in
the pore space of the rock (gray) (from Juanes et al. (2006)).
edge water displaces gas in an imbibition process (increasing water saturations). The presence of an imbibition saturation path
leads to snap-off and, subsequently, trapping of the gas phase. A trail of residual, immobile CO
2
is left behind the plume as it
migrates along the top of the formation (Fig. 1).
Most of the CO2 gas would be immobilized (light blue), trapped as small bubbles (white) in the pore space of the rock (gray).
Only a small portion of the CO2 (dark blue) will continue to ow up towards the impermeable layer of caprock (yellow).
Juanes et al. (2006) discuss ways in which the process of capillary trapping is enhanced:
1. Injecting deeper in thick formations. This provides an opportunity for the CO
2
to be trapped during its vertical ascent.
2. Injection at high injection rates. A shorter injection period leaves less time for the buoyant CO
2
to reach the top of the
formation, from which it is difcult to immobilize.
3. Injection of water slugs alternating CO
2
injection (in the spirit of classical WAG for enhanced oil recovery (Spiteri and
Juanes, 2006)). The injected water forces break-up of large connected CO
2
plumes, enhancing trapping and immobiliza-
tion of the CO
2
. On the other hand, a WAG strategy leads to higher bottom hole pressures at injection wells, which may
be limited by seal integrity, regulatory or economical constraints.
4. Injection in sloping aquifers with regional groundwater ow. In this case, the background ow acts as a natural chase
uid, enhancing the imbibition process.
The effect of capillary trapping can be modeled in conventional simulators by means of relative permeability and capillary
pressure hysteresis (Mo and Akervoll, 2005). However, quantitative predictions of the amount of residual CO
2
require that the
migration paths of the injected CO
2
are captured accurately. This is particularly challenging, since the displacement of brine
by CO
2
is viscously and buoyantly unstable. Coarse simulation models overestimate aquifer sweep and subsequent capillary
trapping of CO
2
(Juanes et al., 2006). The impact of some of the geologic parameters on the fate of the CO
2
(including the
effect on residual trapping) has been investigated by numerical simulation (Mo et al., 2005).
In this paper, we take a parsimonious view toward solving this problem. While high-resolution computational models that
capture the details of the CO
2
brine displacement are possible at the lab scale and (maybe) the pilot scale, this approach is
clearly not feasible for basin-scale models. Modeling (upscaling) of the injection process and of key parameters governing the
CO
2
migration is therefore necessary. In this paper, we develop a model for the upscaling of capillary trapping to the basin
scale. The model is completely analytical and, as such, must involve a number of simplifying assumptions.
The methodology for the upscaling of capillary trapping relies on a sharp-interface model of CO
2
injection and migration
subject to background groundwater ow. The model is one-dimensional, but captures the gravity override due to the density
and mobility contrast between CO
2
and brine. The full mathematical model takes the form of a nonlinear advectiondiffusion
equation with discontinuous coefcients. The diffusion term does not come from physical diffusion but, instead, from the
density difference between CO
2
and brine. It scales with a gravity number, dened below. When the mobility contrast is
sufciently high (as it is in the case of interest), we nd, by solving the full problem numerically, that the footprint and the
CO
2
plume is insensitive to the gravity number. Therefore, we simplify the full model to a hyperbolic model, for which we nd
complete analytical solutions. This solution gives an analytical expression for the footprint of the plume, and the associated time
scale for complete immobilization of CO
2
by residual trapping. The footprint depends on two parameters: the mobility ratio M
between the uids, and a trapping coefcient . By matching the footprint of the model with that of a stable displacement (one
in which gravity override is not accounted for), we immediately obtain a megascopic (aquifer-scale) trapping coefcient .
Using the same logic, we obtain an effective (grid-dependent) trapping coefcient
eff
when the gravity override is only partially
resolved.
SPE 113496 3
100 km
1 km
Cross section
Plan view
Fig. 2 Schematic of the basin-scale model of CO
2
injection. The CO
2
is injected in a deep formation (blue) that has a natural
groundwater ow (East to West in the diagram). The injection wells (red) are placed forming a linear pattern in the deepest section
of the aquifer. Under these conditions, the NorthSouth component of the ow is negligible, and is not accounted for in the one-
dimensional ow model developed here.
Description of the Physical Model
A schematic of the geologic setting for which the ow model is developed is shown in Fig. 2. We are interested in modeling the
injection at the basin scale: injection of gigatons of CO
2
, and a domain of hundreds of kilometers in the horizontal direction.
The CO
2
is injected in a deep formation (blue) that has a natural groundwater ow (East to West in the diagram). The
injection wells (red) are placed forming a line-drive pattern in the deepest section of the aquifer. Under these conditions, the
ow does not have large variations in the NorthSouth direction. This simplication justies the one-dimensional ow model
developed here.
Our physical model is shown in Fig. 3. It consists of a homogeneous and horizontal aquifer, into which CO
2
is injected.
We divide the study of the migration of CO
2
in three stages, also shown in Fig. 3 from top to bottom:
1. Injection period. CO
2
(white) is injected at a high ow rate, displacing the brine (light blue) to its irreducible saturation.
Due to buoyancy, the injected CO
2
will form a gravity tongue, or gravity current (Huppert and Woods, 1995). The
high contrast in mobility between the injected CO
2
and the initial brine exacerbates this gravity tongue, consequently
decreasing the overall sweep of the aquifer.
2. Early post-injection period. Once injection stops, the CO
2
plume continues to migrate due its buoyancy and a background
regional groundwater ow (assumed uniform). The leading edge of the plume is still a drainage front (CO
2
displaces
water). At the trailing edge of the plume, in contrast, water displaces CO
2
, trapping it in residual form (dark gray).
During the initial time after injection, the imbibition and drainage fronts are separated.
3. Late post-injection period. After sufcient time, the drainage and imbibition fronts merge, and the mobile plume detaches
from the bottom of the aquifer. It continues to migrate laterally, progressively decreasing its thickness until all the CO
2
is trapped.
Sharp-interface models of gravity currents in porous media have been studied for a long time (see, e.g., Barenblatt (1996)
and Huppert and Woods (1995)). Analytical solutions for the evolution of an axisymmetric gravity current have been presented
by Kochina et al. (1983), Lyle et al. (2005), and Nordbotten et al. (2005) (this last work in the context of leakage through
abandoned wells). An extension to inclined porous media was developed by (Vella and Huppert, 2006). For a one-dimensional
model in horizontal aquifers, Hesse et al. (2007) found early-time and late-time similarity solutions, which show there is a
change in the scaling of the footprint of the plume from t
1/2
(rapid decay) to t
1/3
(slow decay). Of particular relevance is
the recent work by Hesse et al. (2006): they developed a one-dimensional model in sloping aquifers that includes capillary
trapping. They solved their model numerically, and showed that the presence of slope leads to much more effective trapping
scenarios than if the aquifer is horizontal.
In the next section we develop a sharp-interface mathematical model for the conceptual model of Fig. 3. There are several
important differences with respect to the model of Hesse et al. (2006).
4 SPE 113496
brine
mobile gas
gas
injection
brine
mobile gas
trapped gas
water
flow
brine
mobile gas
trapped gas
water
flow
Fig. 3 Conceptual representation of the three different periods of CO
2
migration in a horizontal aquifer. From top to bottom: injection
period; early post-injection period; late post-injection period (see text for a detailed explanation).
1. The most important difference is that we model the injection period, while Hesse et al. (2006) use a unit-square initial
condition. We show that the shape of the plume at the end of injection leads to exacerbated gravity override, which affects
the subsequent migration of the plume in a fundamental way. Accurate estimates of the footprint of the plume and the
associated time scale for plume immobilization require that the injection period be modeled.
2. Our analysis includes the effect of regional groundwater ow. The impact of background ow is neglected during the
injection period because the high ow rate of the injected CO
2
dominates, but it is essential in the evolution of the plume
after injection stops. Mathematically, the effect of a background ow is similar to that of nonzero aquifer slope.
3. Our model also accounts for the reduced mobility of water due to the trapping of CO
2
. Resistance to ow of water
is different in (a) the initial water-lled medium, and (b) the pore space that has been invaded by CO
2
and contains a
residual fraction of CO
2
. This effect, which has been neglected in all sharp-interface models to date, turns out to be
essential in arriving at a consistent mathematical model during the post-injection period.
Mathematical Model
The essential feature of the mathematical model is a sharp-interface approximation (Huppert and Woods, 1995), by which
the medium is assumed to either be lled with water (water saturation S
w
= 1), or lled with CO
2
to irreducible connate
water saturation (gas saturation S
g
= 1 S
wc
). The other major assumption is that the dimension of the aquifer is much
larger horizontally than vertically, so that the vertical ow equilibrium approximation (Yortsos, 1995), also know as Dupuit
approximation (Bear, 1972), is applicable. This assumption dictates that owis essentially horizontal, and the pressure variation
in the vertical direction is hydrostatic.
For clarity, we develop the mathematical model for the injection period rst. We then extend it to the post-injection period,
during which trapping of CO
2
occurs.
Model for the Injection Period. Consider the encroachment of the injected CO
2
plume into the aquifer, as shown in Fig. 4.
The density of the CO
2
, , is lower than that of the brine, + . Let h
g
be the thickness of the (mobile) CO
2
plume, and
H the total thickness of the aquifer. In view of the sharp-interface and Dupuit approximations, the pressure along a vertical line
is given by:
p =

p
I
g(h
g
z) if 0 < z < h
g
,
p
I
+ ( + )g(z h
g
) if h
g
< z < H,
(1)
where p
I
is the interface pressure, common to both uids (capillary pressure is neglected).
There are two differentiated regions: Region 1, of mobile gas and connate (immobile) water; Region 2, of fully-saturated
mobile water and no gas. The horizontal volumetric ux of each uid is calculated by the multiphase ow extension of
Darcys law, which involves the relative permeability to each phase (Muskat, 1949). In Fig. 5(left) we show typical relative
permeability curves for water and gas during drainage, as functions of water saturation. Since the ow regions are separated by
a sharp interface, only the endpoints of the curves are needed. In Region 1, k
rw
= 1 and k
rg
= 0. In Region 2, k
rw
= 0 and
k
rg
= k

rg
< 1.
SPE 113496 5
h
g
CO
2
injection
H Q

+
Fig. 4 Schematic of the solution during the injection period.
*
rg
k
*
rw
k
w
S
r
k
r
k
w
S
Fig. 5 Typical relative permeability curves for brine (blue) and CO
2
(red) as function of brine saturation. Due to hysteresis, the curves
are different during drainage (left) and imbibition (right).
Therefore, we have:
u
g
=

k
k

rg

g
(
x
p
I
g
x
h
g
) if 0 < z < h
g
,
0 if h
g
< z < H,
u
w
=

0 if 0 < z < h
g
,
k
1

w
(
x
p
I
( + )g
x
h
g
) if h
g
< z < H,
(2)
where k is the permeability of the medium. We assume that the uids are incompressible to establish the following volumetric
ux constraint:
Q
g
+Q
w
=

H
0
u
g
(z) dz +

H
0
u
w
(z) dz = u
g

z<hg
h
g
+u
w

z>hg
(H h
g
) Q = 0, (3)
which allows to eliminate the interface pressure p
I
from the equations, and express the vertically-integrated ux of CO
2
in
fractional ow form (Aziz and Settari, 1979; Chavent and Jaffr e, 1986):
Q
g
= f
g
Qkg
k

rg

g
h
g
(1 f
g
)
x
h
g
. (4)
In the equation above, f
g
is the fractional ow of gas:
f
g
=

g
+

w
, (5)
where

is the vertically-averaged mobility of the -uid. During the injection period:

g
=
k

rg

g
h
g
H
,

w
=
1

w
H h
g
H
. (6)
Mass balance for the CO
2
phase then results in the governing equation for the plume thickness during injection:
(1 S
wc
)
t
h
g
+
x

f
g
QkgH

g
(1 f
g
)
x
h
g

= 0, (7)
where is the aquifer porosity, and S
wc
is the connate water saturation.
Model for the Post-Injection Period. The initial condition for this period is the shape of the plume at the end of injection.
After injection stops, the plume continues to migrate leaving a trail of trapped CO
2
behind, with gas saturation S
gr
(see Fig. 6).
Now, at any point along the horizontal, three regions coexist along the vertical line. Sandwiched between Regions 1 (mobile
gas and connate water) and 2 (initial mobile water and no gas) is Region 3 of mobile water and trapped gas. Because the mobile
6 SPE 113496
brine
mobile gas
trapped gas
water
flow
h
g
UH
h
g,max
(S
g
= S
gr
)
(S
g
= 1 S
wc
)
Fig. 6 Schematic of the solution during the post-injection period.
uid in Regions 2 and 3 is water, the hydrostatic pressure distribution is the same as during the injection case (Equation (1)).
The uxes, however, are affected by the presence of this region of trapped gas, since the resistance to ow of water is higher
than in the region fully-saturated with the initial brine. This can be seen from the plot of the relative permeabilities during
imbibition, shown in Fig. 5(right). The relative permeability to water in the region of residual CO
2
is k

rw
, which may be
signicantly lower than 1 (Oak, 1990; Bennion and Bachu, 2006).
An important variable in the formulation is h
g,max
, the maximum historic value of the plume thickness. Application of
Darcys law to the sharp-interface model leads to the following expression for the gas and water Darcy velocities:
u
g
=

k
k

rg

g
(
x
p
I
g
x
h
g
) if 0 < z < h
g
,
0 if h
g
< z < h
g,max
,
0 if h
g,max
< z < H,
u
w
=

0 if 0 < z < h
g
,
k
k

rw

w
(
x
p
I
( + )g
x
h
g
) if h
g
< z < h
g,max
,
k
1
w
(
x
p
I
( + )g
x
h
g
) if h
g,max
< z < H.
(8)
The incompressibility constraint during the post-injection period reads:
Q
g
+Q
w
=

H
0
u
g
(z) dz+

H
0
u
w
(z) dz = u
g

z<h
g
h
g
+u
w

h
g
<z<h
g,max
(h
g,max
h
g
)+u
w

z>h
g,max
(Hh
g,max
) UH = 0,
(9)
where U is the average Darcy velocity due to regional groundwater ow. This constraint allows, once again, to eliminate the
interface pressure p
I
and express the integrated gas ux in fractional ow form:
Q
g
= f
g
UH kgH

g
(1 f
g
)
x
h
g
, (10)
where f
g
is still dened by Equation (5), with the notable difference that the average mobility of water depends not only on the
plume thickness, but also its maximum historic thickness:

w
=
k

rw

w
h
g,max
h
g
H
+
1

w
H h
g,max
H
. (11)
During the post-injection period, CO
2
is present in the mobile plume (with saturation S
g
= 1 S
wc
) and as a trapped phase
(with saturation S
g
= S
gr
). Mass balance for the CO
2
phase reads:

t
((1 S
wc
)h
g
) +
t
(S
gr
(h
g,max
h
g
)) +
x
Q
g
= 0 (12)
It is important to evaluate the accumulation term carefully. For the drainage front of the plume, the mobile plume thickness
increases (
t
h
g
> 0) and the maximum historic thickness is precisely the current thickness (h
g,max
h
g
). For the imbibition
front, the mobile plume thickness decreases (
t
h
g
< 0) and, since the residual CO
2
is immobile, the maximum historic
thickness does not change with time (
t
h
g,max
0). We dene an accumulation coefcient R as follows:
R =

1 S
wc
if
t
h
g
> 0 (drainage),
1 S
wc
S
gr
if
t
h
g
< 0 (imbibition),
(13)
The governing equation for the plume thickness during the post-injection period is, therefore:
R
t
h
g
+
x

f
g
UH kgH

g
(1 f
g
)
x
h
g

= 0. (14)
This equation is almost identical to Equation (7) governing the plume thickness during injection. There are, however, three
important differences: (1) the coefcient in the accumulation term is discontinuous; (2) the fractional ow is a function of both
h
g
and h
g,max
; and (3) the advection term scales with the integrated groundwater ux, UH, and not the injected CO
2
ow rate.
Dimensionless Form of the Equations. It is insightful (and useful in practice) to express the equations in dimensionless form.
We do it for the injection period rst. We dene the following characteristic quantities:
H = thickness of the aquifer,
T = injection time,
L =
QT
H
= characteristic footprint of a stable displacement at the end of injection,
SPE 113496 7
and the dimensionless variables:
h =
h
g
H
, =
t
T
, =
x
L
. (15)
The plume evolution equation during injection takes the following dimensionless form:
(1 S
wc
)

h +

(f N
g
h(1 f)

h) = 0, (16)
where the fractional ow is now a function of the dimensionless thickness:
f(h) =
h
h +M(1 h)
. (17)
The behavior of the system is entirely governed by the following two dimensionless parameters:
M =
1/
w
k

rg
/
g
= mobility ratio, (18)
N
g
=
kk

rg

g
(Q/H)
H
(QT)/(H)
= gravity number. (19)
Equation (16) is a nonlinear advectiondiffusion equation, where the second-order term comes from buoyancy forces, not
physical diffusion. The gravity diffusion coefcient is a nonlinear function of the plume thickness:
D(h) = h(1 f(h)), (20)
which is degenerate (zero diffusion) at both endpoints (h = 0 and 1).
During the post-injection period, it proves convenient to re-scale time differently given that the integrated groundwater
ow UH may be very different from the injection ow rate Q. Thus, for t > T, we dene
= 1 +
UH
Q
t T
T
. (21)
The scaling in space remains unchanged, something that will prove very convenient later. The dimensionless partial differential
equation during the post-injection period is:
R

h +

f N
g
Q
UH
h(1 f)

= 0. (22)
Let us summarize the main differences with respect to the injection period:
1. As noted above, the time scaling is different.
2. Buoyancy effects reect the difference in time scaling (Q/UH), and are typically much larger during the post-injection
period.
3. The coefcient in the accumulation term is discontinuous, and given by Equation (13).
4. The fractional ow function is now dependent on the (dimensionless) maximum historic thickness h
max
:
f(h; h
max
) =
h
h +k

rw
M(h
max
h) +M(1 h
max
)
. (23)
5. The gravity diffusion coefcient is, therefore, also dependent on h
max
.
In Fig. 7 we plot the fractional ow function (left) and gravity diffusion function (right) for the relative permeability curves in
Fig. 5 and a viscosity ratio
g
/
w
= 0.1. The blue solid line corresponds to the case h
max
= h (drainage front), and the dashed
red line corresponds to the case h
max
= 1 (imbibition front during the early post-injection period).
8 SPE 113496
0 0.2 0.4 0.6 0.8 1
0
0.2
0.4
0.6
0.8
1
h
f
Imbibition
Drainage
0 0.2 0.4 0.6 0.8 1
0
0.01
0.02
0.03
0.04
0.05
0.06
0.07
0.08
h
D
Drainage
Imbibition
Fig. 7 Fractional ow function (left) and gravity diffusion function (right) for the relative permeability curves in Fig. 5 and a viscosity
ratio
g
/
w
= 0.1. The blue solid line (termed drainage in the legend) corresponds to the case h
max
= hdrainage front. The dashed
red line (termed imbibition in the legend) corresponds to the case h
max
= 1imbibition front during the early post-injection period.
0 5 10 15 20 25 30 35 40 45 50
0
1
Drainage, = 1
0 5 10 15 20 25 30 35 40 45 50
0
1
Imbibition, = 2.5
0 5 10 15 20 25 30 35 40 45 50
0
1
Drainage, = 1
0 5 10 15 20 25 30 35 40 45 50
0
1
Imbibition, = 2.5
0 5 10 15 20 25 30 35 40 45 50
0
1
Drainage, = 1
0 5 10 15 20 25 30 35 40 45 50
0
1
Imbibition, = 2.5
Fig. 8 Snapshots of the numerical solution at different times. Left: end of injection period ( = 1). Right: near complete immobiliza-
tion of the plume ( = 2.5). Shown are the mobile CO
2
plume (white) and the trail of residual CO
2
(yellow). Results are for a system
with the relative permeabilities of Fig. 5, M = 0.1, UH/Q = 0.001, and three different gravity numbers. Top row: N
g
= 0. Middle row:
N
g
= 0.1. Bottom row: N
g
= 1.
SPE 113496 9
Numerical Solution of the Full Model
Equations (16) (injection period) and (22) (post-injection period) must be solved with appropriate initial and boundary condi-
tions. The problem during injection is initialized with h( = 0) = 0, and a boundary ux f( = 0, > 0) = 1. The solution
at the end of the injection period, h(, = 1), is then used as initial condition for the post-injection period. The boundary
condition is switched to a ux of water only f( = 0, > 1) = 0.
The problem described above does not, in general, have an analytical solution. It can be solved, however, using standard
numerical discretization methods. Here, we compute numerical simulations using a single-point upstream discretization for
the advection term, a centered second-order scheme for the gravitydiffusion term, and a forward Euler time discretization. In
Fig. 8 we show results for a system with the relative permeabilities of Fig. 5, M = 0.1, UH/Q = 0.001, and three different
gravity numbers, N
g
= 0 (top row), N
g
= 0.1 (middle row), and N
g
= 1 (bottom row).
While certain differences among the solutions exist, they are quantitatively very similar, especially considering the wide
range of gravity numbers considered.
It is useful to see whether what the mobility ratio and gravity number are for a realistic CO
2
sequestration scenario. Consider
an aquifer with k = 100 md = 10
13
m
2
, = 0.2, and H = 100 m. Injection conditions are about 100 bar and 40

C. Under
these conditions, 400 kgm
3
, 600 kgm
3
,
g
0.05 10
3
kgm
1
s
1
, and
w
0.8 10
3
kgm
1
s
1
.
For k

rg
= 0.7, we get M 0.09, very close to the value M = 0.1 used in the simulations. Consider a major sequestration
project, in which 1 gigatonne of CO
2
is injected every year, for a period of T = 30 years. If injection takes place at 100 wells,
with interwell spacing of 1 km. Then, Q = 2.5 10
4
m
2
yr
1
and Q/H = 250 myr
1
. For the ratio UH/Q = 0.001,
this corresponds to a background groundwater ow velocity of a few decimeters per year, which is reasonable. The resulting
gravity number is N
g
0.2510
2
. Clearly, for this range of gravity numbers, the solution is essentially identical to that with
N
g
= 0.
Analytical Solution to the Hyperbolic Model
In the light of the discussion above, it is well justied to drop the second-order diffusive term from the formulation when the
mobility ratio M is sufciently small (as is the case in relevant CO
2
sequestration scenarios). It is interesting (but not surprising)
that for M 1, the solution becomes independent of the density difference between the uids, even though it is buoyancy that
sets the gravity tongue.
In this section, we obtain a closed-form analytical solution to the hyperbolic limit of the problem. We do so for each stage
depicted in Fig. 3: injection, early post-injection, and late post-injection.
Injection Period. During the injection period, 0 < < 1, the dimensionless equation reduces to the scalar hyperbolic
conservation law:
(1 S
wc
)

h +

f = 0. (24)
The problem to be solved is a Riemann problem (Lax, 1957; Smoller, 1994), that is, the evolution of an initial discontinuity at
= 0:
h(, = 0) =

1 0,
0 > 0.
(25)
Since both the partial differential equation and the initial condition are invariant to a stretching of coordinates (, ) (c, c)
with c > 0, the solution can be expressed as a function of the characteristic velocity :
h(, ) = U(), = /. (26)
Since the ux function f is concave (see Fig. 7), the solution to the Riemann problem during injection is a simple rarefaction
wave, that is, a continuous solution that satises:
U

() =
dU
d
=
1
1 S
wc
f

(U). (27)
In Fig. 9(left) we plot the prole of the solution at the end of injection, h(, = 1). In Fig. 9(right) we depict the solution
on the (, )-plane, illustrating the evolution of the discontinuity into a rarefaction fan. The solution shown corresponds to
M = 0.1, typical of CO
2
sequestration scenarios. It is clear that the plume, at the end of the injection period, is far from being
a unit square of homogeneous saturations, as assumed by Hesse et al. (2006).
Early Post-Injection Period. After injection of CO
2
stops, the governing equation is the hyperbolic conservation law:
R

h +

f = 0, (28)
where
R =

1 S
wc
if

h > 0 (drainage),
1 S
wc
S
gr
if

h < 0 (imbibition),
, and f =

h
h+M(1h)
if

h > 0 (drainage),
h
h+k

rw
M(h
max
h)+M(1h
max
)
if

h < 0 (imbibition).
(29)
10 SPE 113496
0 5 10 15
0
1

h
0 5 10 15
0
1

Fig. 9 Analytical solution to the hyperbolic model during the injection period. Top: prole of the mobile CO
2
plume (white) at the
end of injection ( = 1). Bottom: solution on (, )-space, displaying the solution rarefaction fan.
The boundary condition at = 0 switches from h = 1 to h = 0 at = 1. This results in a Riemann problem that describes the
imbibition front during the early time after injection stops. The coefcients in Equation (29) must be evaluated for an imbibition
process, with h
max
= 1. Due to the concave shape of the ux function (Fig. 7), the solution to this Riemann problem is a shock,
that is, a traveling discontinuity that propagates at a speed given by the RankineHugoniot condition (Smoller, 1994):
=
1
1 S
wc
S
gr
. (30)
While the imbibition shock advances, the continuous drainage front continues to propagate exactly as during the injection
period. The characteristic velocity is the same as during the injection period because time was re-scaled. The early post-
injection period ends when the imbibition front catches up with the drainage front at the bottom of the aquifer. The solution at
this time, including the mobile plume (white) and the residual CO
2
(yellow), is shown in Fig. 10.
The end of the early post-injection can be interpreted as the time at which the imbibition shock wave (thick red line) collides
with the slowest ray of the drainage rarefaction wave (thin blue line)see bottom plot in Fig. 10. It can be computed easily by
imposing that
1
1 S
wc
S
gr
(
c
1) =
1
1 S
wc
f

(h)

h=1

c
, (31)
By dening the trapping coefcient:
=
1 S
wc
1 S
wc
S
gr
(always > 1), (32)
the collision time is

c
=

M
. (33)
SPE 113496 11
0 5 10 15
0
1

h
0 5 10 15
0
1

Fig. 10 Analytical solution to the hyperbolic model during the early post-injection period. Top: prole of the mobile CO
2
plume
(white) and trapped CO
2
(yellow) when the drainage front detaches from the bottom of the aquifer ( = c
). Bottom: solution on
(, )-space, showing the collision of the imbibition shock with the slowest ray of the drainage rarefaction.
water
flow
UH
imbibition
drainage
h
m
Fig. 11 Sketch of the continuous interaction of the imbibition front with the leading drainage rarefaction wave during the late post-
injection period. The solution to the problem consists in characterizing the evolution of the state hm as a function of dimensionless
time .
Late Post-Injection Period. Once the drainage front detaches from the bottom of the aquifer, the imbibition front is a shock
from h = 0 to h = h
m
< 1. The solution is no longer a Riemann problem but, instead, the continuous interaction of a
progressively faster shock with a rarefaction wave (see Fig. 11). The problem is solved if one determines the evolution of the
state h
m
as a function of dimensionless time . As it turns out, the problem can be solved in closed form.
The differential equation governing the evolution of the state h
m
can be obtained by nding the intersection (on the (, )-
space) of the imbibition shock wave corresponding to a state h
m
with the rarefaction ray for a state h
m
+ dh
m
, and taking the
limit dh
m
0. The resulting differential equation is:

1
1 S
wc
S
gr
f(h
m
; h
m
)
h
m

1
1 S
wc
f

(h
m
; h
m
)

d =
1
1 S
wc
f

(h
m
; h
m
)dh
m
. (34)
The initial condition is h
m
= 1 at =
c
= /( M), that is, precisely when the drainage front detaches from the bottom
of the aquifer. After separation of variables, Equation (34) can be written as follows:


c
d

hm
1
f

(h; h)

f(h;h)
h
f

(h; h)
dh. (35)
12 SPE 113496
0 5 10 15 20 25 30 35 40 45 50
0
1

h
0 5 10 15 20 25 30 35 40 45 50
0
1
(
max
,
max
)

Fig. 12 Analytical solution to the hyperbolic model during the late post-injection period. Top: prole of the mobile CO
2
plume (white)
and trapped CO
2
(yellow) at some intermediate time (c
< <
max
). Bottom: complete solution on (, )-space until the entire CO
2
plume has been immobilized in residual form.
The integral in Equation (35) can be evaluated analytically, and the solution to the late post-injection period admits a closed-
form representation:
(h
m
) = ( M)

M + (1 M)h
m
M( 1) + (1 M)h
m

2
. (36)
The complete analytical solution to the plume evolution is shown in Fig. 12. On the top frame, we plot the prole of the
mobile CO
2
plume (white) and trapped CO
2
(yellow) at some intermediate time (
c
< <
max
). Particularly interesting is
the bottom frame, which shows the a representation of the solution in characteristic space (the (, )-plane). The thick blue
straight line corresponds to the fastest ray of the drainage front. The thick red line corresponds to the imbibition front. This
wave increases its propagation speed during the late injection period, and interacts continuously with progressively faster rays
of the drainage front. When the imbibition front collides with the fastest ray, the entire CO
2
plume is in residual, immobile
form. This occurs at a dimensionless time
max
(the time required for complete trapping) and a dimensionless distance
max
(the
ultimate footprint of the plume).
Footprint of the CO
2
Plume
One of the most important practical results from the analytical solution derived above is the closed-form expressions for the
length and time scale for complete trapping:

max
=
( M)
( 1)
2
, (37)

max
=
( M)
( 1)
2
M
1 S
wc
. (38)
The newly dened trapping coefcient (Equation (32) emerges as the key parameter in the assessment of CO
2
storage in saline
aquifers. It is always greater than one, and it increases with increasing residual gas saturation. Of course, larger values of
SPE 113496 13
1 1.5 2 2.5 3 3.5 4
1
1.5
2
2.5
3
3.5
4

lab

M=1
M=0.5
M=0.2
M=0.1
M=0
Fig. 13 Plot of the megascopic trapping coefcient vs. the laboratory value
lab
, for different values of the mobility ratio M.
When the mobility ratio is less than 0.1, the curves collapse onto the limiting curve corresponding to M = 0.
result in more effective trapping of the CO
2
plume. It is not surprising that the ultimate footprint of the plume is proportional
to the mobility ratio M.
We emphasize that this analysis takes into account the evolution of the plume during the injection period. The question
arises as to whether the impact of the injection period on the ultimate footprint is signicant or not. This, in turn, depends
directly on the value of
max
, which is the ratio of the plume footprint at complete trapping versus the footprint when injection
stops. Values of
max
1 would indicate that the ultimate footprint is not very sensitive to the injection period. This, however,
is far from being the case. For typical parameters of saline aquifers, > 2, and M 0.1. Therefore,
max
< 4, suggesting that
it is essential to model the injection period for proper assessment of the ultimate footprint of the plume.
Upscaling of Capillary Trapping
Megascopic Trapping Coefcient . We can use the analytical results developed in the previous section to formulate a
model for the upscaling of capillary trapping. The motivation is that, given that gravity override drastically reduces aquifer
sweep, one should modify (upscale) the trapping coefcient if this reduction in sweep is ignored. The premise for the calculation
of a megascopic, aquifer-scale trapping coefcient is to match the footprint of the plume (or, equivalently, its time for
complete trapping) for two scenarios:
1. When gravity override is modeled, and the laboratory-scale value of the trapping coefcient is used,

max
=
max
(
lab
, M) =

lab
(
lab
M)
(
lab
1)
2
. (39)
2. A stable displacement (M = 1, N
g
= 0) with an upscaled value of the trapping coefcient,

max
=
max
() =

1
. (40)
Matching both expressions, we obtain an equation for the megascopic trapping coefcient:
=

lab
(
lab
M)

lab
(
lab
M) (
lab
1)
2
. (41)
Equation (41) is plotted in Fig. 13 as vs.
lab
for different values of M. When the mobility ratio is less than 0.1, the curves
collapse onto the limiting curve corresponding to M = 0.
Block-Effective Trapping Coefcient
eff
. Similar considerations apply for modeling the effective trapping coefcient in a
numerical simulation model. In this case, the model partially (but not completely) resolves gravity override, and the gridblock-
effective trapping coefcient
eff
should be lower than the value measured experimentally in a homogeneous core-scale dis-
placement (Juanes et al., 2006).
The fact that the gravity tongue is partially resolved can be incorporated by modifying the value of the mobility ratio. This
can be done, for instance, by matching:
1. The disparity in the front speeds within the top gridblock in a vertical discretization with resolution z = 1/N
z
and
actual mobility ratio M:
f

(h = 0) f

(h = z) =
1
M

M
(M + (1 M)z)
2
. (42)
14 SPE 113496
N
z
=

1
5
2
0
1
0
0
2
0
0
0
1
0
0
0
0
5
0
0
Fig. 14 Plot of the effective mobility ratio M
eff
as a function of the actual mobility ratio M and the number of gridblocks N
z
in the
vertical direction.
10
0
10
1
10
2
10
3
10
4
1.5
1.6
1.7
1.8
1.9
2
2.1
2.2
2.3
2.4
N
z

e
f
f
M=1
M=0.5
M=0.2
M=0.1
M=0.05
Fig. 15 Plot of the gridblock-effective trapping coefcient
eff
as a function of the actual mobility ratio M and the number of grid-
blocks N
z
in the vertical direction. The curves shown correspond to a laboratory value of
lab
= 2.33.
2. The disparity in front speeds between the top and bottom of the aquifer, for an equivalent mobility ratio M
eff
:
f

eff
(h = 0) f

eff
(h = 1) =
1
M
eff
M
eff
. (43)
Equating both expressions leads to a relation of the effective mobility ratio as a function of the actual mobility ratio and the
number of gridblocks in the vertical direction, N
z
. Such relation is shown graphically in Fig. 14.
Inserting the expression for the equivalent mobility ratio M
eff
(Fig. 14) into the equation for the upscaled trapping coefcient
(Fig. 13), we obtain a model for the gridblock-effective trapping coefcient as a function of
lab
, M and N
z
. In Fig. 15 we
show an example of such relation, for a value of
lab
= 2.33 (corresponding to S
wc
= 0.3 and S
gr
= 0.4).
For coarse discretizations, the block-effective trapping coefcient is close to the megascopic value. If the vertical discretiza-
tion is very ne, the upscaled trapping coefcient converges to the laboratory value. The curves in Fig. 15 also give information
about how much discretization is required to capture the gravity tongue. Naturally, the grid resolution must be higher as the
mobility ratio decreases.
Conclusions
We have presented a sharp-interface mathematical model of CO
2
migration in saline aquifers, that accounts for gravity override
and capillary trapping. The major differences with respect to previous investigations is that we model the shape of the plume
during the injection period, we account for regional groundwater ow during the post-injection period, and we introduce
rigorously the reduced mobility to water through the porous medium with trapped CO
2
.
The model leads to a nonlinear advectiondiffusion equation, in which the diffusive term is due to buoyancy forces, not
physical diffusion. The three key dimensionless groups are the mobility ratio between the injected and initial uids (M), the
gravity number (N
g
), and a newly dened trapping coefcient (). For the case of interest in geological CO
2
storage, in which
the mobility ratio is much smaller than 1, the solution is largely insensitive to the value of the gravity number. Under these
conditions, the mathematical model can be simplied to a hyperbolic model.
We have developed a complete analytical solution to the hyperbolic model that includes the injection, early post-injection,
and late post-injection periods. Despite the fact that the solution involves the continuous interaction of an imbibition shock with
a drainage rarefaction wave, it admits a closed-form expression. The main outcome of the analytical developments presented
here is a formula that predicts the ultimate footprint on the CO
2
plume, and the time scale required for complete trapping. Both
quantities depend strongly on the shape of the plume at the end of the injection period, which mustthereforebe modeled.
SPE 113496 15
A second application of the analytical solution is a model for upscaling the capillary trapping coefcient from the laboratory
scale to the basin scale. Explicit expressions are given for the megascopic and gridblock-effective trapping coefcients, as
functions of the local trapping coefcient, the mobility ratio, and the grid resolution. Although the expressions derived are
based on a one-dimensional sharp-interface model, we anticipate that they will have broader applicability to injection scenarios
with unfavorable mobility ratio and dominated by gravity override.
Acknowledgments
Partial funding was provided by the Charles E. Reed Faculty Initiative Fund. This support is gratefully acknowledged.
References
Aziz, K. and Settari, A. (1979). Petroleum Reservoir Simulation. Elsevier, London.
Bachu, S., Gunther, W. D., and Perkins, E. H. (1994). Aquifer disposal of CO
2
: Hydrodynamic and mineral trapping. Energy Conv. Manag.,
35(4):269279.
Barenblatt, G. I. (1996). Scaling, Self-Similarity, and Intermediate Asymptotics. Cambridge Texts in Applied Mathematics. Cambridge
University Press.
Bear, J. (1972). Dynamics of Fluids in Porous Media. Environmental Science Series. Elsevier, New York. Reprinted with corrections, Dover,
New York, 1988.
Bennion, D. B. and Bachu, S. (2006). Supercritical CO
2
and H
2
Sbrine drainage and imbibition relative permeability relationships for inter-
granular sandstone and carbonate formations. In SPE Europec/EAGE Annual Conference and Exhibition, Vienna, Austria. (SPE 99326).
Chavent, G. and Jaffr e, J. (1986). Mathematical Models and Finite Elements for Reservoir Simulation, volume 17 of Studies in Mathematics
and its Applications. Elsevier, North-Holland.
Ennis-King, J. and Paterson, L. (2005). Role of convective mixing in the long-term storage of carbon dioxide in deep saline formations. Soc.
Pet. Eng. J., 10(3):349356.
Ennis-King, J., Preston, I., and Paterson, L. (2005). Onset of convection in anisotropic porous media subject to a rapid change in boundary
conditions. Phys. Fluids, 17(8):084107. doi:10.1063/1.2033911.
Flett, M., Gurton, R., and Taggart, I. (2004). The function of gaswater relative permeability hysteresis in the sequestration of carbon dioxide
in saline formations. In SPE Asia Pacic Oil and Gas Conference and Exhibition, Perth, Australia. (SPE 88485).
Gunter, W. D., Wiwchar, B., and Perkins, E. H. (1997). Aquifer disposal of CO
2
-rich greenhouse gases: Extension of the time scale of
experiment for CO
2
-sequestering reactions by geochemical modeling. Miner. Pet., 59(12):121140.
Hesse, M. A., Tchelepi, H. A., Cantwel, B. J., and Orr, Jr., F. M. (2007). Gravity currents in horizontal porous layers: transition from early to
late self-similarity. J. Fluid Mech., 577:363383.
Hesse, M. A., Tchelepi, H. A., and Orr, Jr., F. M. (2006). Scaling analysis of the migration of co
2
in saline aquifers. In SPE Annual Technical
Conference and Exhibition, San Antonio, TX. (SPE 102796).
Huppert, H. E. and Woods, A. W. (1995). Gravity-driven ows in porous media. J. Fluid Mech., 292:5569.
Juanes, R., Spiteri, E. J., Orr, Jr., F. M., and Blunt, M. J. (2006). Impact of relative permeability hysteresis on geological CO
2
storage. Water
Resour. Res., 42:W12418, doi:10.1029/2005WR004806.
Kochina, I. N., Mikhailov, N. N., and Filinov, M. V. (1983). Groundwater mound damping. Int. J. Eng. Sci., 21:413421.
Kumar, A., Ozah, R., Noh, M., Pope, G. A., Bryant, S., Sepehrnoori, K., and Lake, L. W. (2005). Reservoir simulation of CO
2
storage in
deep saline aquifers. Soc. Pet. Eng. J., 10(3):336348.
Lax, P. D. (1957). Hyperbolic systems of conservation laws, II. Comm. Pure Appl. Math., 10:537566.
Lenormand, R., Zarcone, C., and Sarr, A. (1983). Mechanisms of the displacement of one uid by another in a network of capillary ducts. J.
Fluid Mech., 135:123132.
Lyle, S., Huppert, H. E., Hallworth, M., Bickle, M., and Chadwick, A. (2005). Axisymmetric gravity currents in a porous medium. J. Fluid
Mech., 543:293302.
Mo, S. and Akervoll, I. (2005). Modeling long-term CO
2
storage in aquifer with a black-oil reservoir simulator. In SPE/EPA/DOE Exploration
and Production Environmental Conference, Galveston, TX. (SPE 93951).
Mo, S., Zweigel, P., Lindeberg, E., and Akervoll, I. (2005). Effect of geologic parameters on CO
2
storage in deep saline aquifers. In 14th
Europec Biennial Conference, Madrid, Spain. (SPE 93952).
Muskat, M. (1949). Physical Principles of Oil Production. McGraw-Hill, New York.
Nordbotten, J. M., Celia, M. A., and Bachu, S. (2005). Analytical solution for CO
2
plume evolution during injection. Transp. Porous Media,
58(3):339360.
Oak, M. J. (1990). Three-phase relative permeability of water-wet Berea. In SPE/DOE Seventh Symposium on Enhanced Oil Recovery, Tulsa,
OK. (SPE/DOE 20183).
Pruess, K. and Garca, J. (2002). Multiphase ow dynamics during CO
2
disposal into saline aquifers. Environ. Geol., 42(23):282295.
Pruess, K., Xu, T., Apps, J., and Garcia, J. (2003). Numerical modeling of aquifer disposal of CO
2
. Soc. Pet. Eng. J., 8(1):4960.
Riaz, A., Hesse, M., Tchelepi, H. A., and Orr, Jr., F. M. (2006). Onset of convection in a gravitationally unstable, diffusive boundary layer in
porous media. J. Fluid Mech., 548:87111.
Smoller, J. (1994). Shock Waves and Reaction-Diffusion Equations, volume 258 of A Series of Comprehensive Studies in Mathematics.
Springer-Verlag, New York, second edition.
Spiteri, E. J. and Juanes, R. (2006). Impact of relative permeability hysteresis on the numerical simulation of WAG injection. J. Pet. Sci.
Eng., 50(2):115139.
Valvatne, P. H. and Blunt, M. J. (2004). Predictive pore-scale modeling of two-phase ow in mixed wet media. Water Resour. Res.,
40:W07406, doi:10.1029/2003WR002627.
Vella, D. and Huppert, H. E. (2006). Gravity currents in a porous medium at an inclined plane. J. Fluid Mech., 555:353362.
16 SPE 113496
Xu, T. F., Apps, J. A., and Pruess, K. (2003). Reactive geochemical transport simulation to study mineral trapping for CO2 disposal in deep
arenaceous formations. J. Geophys. Res., 108(B2):2071. doi:10.1029/2002JB001979.
Yortsos, Y. C. (1995). A theoretical analysis of vertical ow equilibrium. Transp. Porous Media, 18:107129.

Você também pode gostar