Hesse 08
Hesse 08
A DISSERTATION
SUBMITTED TO THE DEPARTMENT OF ENERGY
RESOURCES ENGINEERING
AND THE COMMITTEE ON GRADUATE STUDIES
OF STANFORD UNIVERSITY
IN PARTIAL FULFILLMENT OF THE REQUIREMENTS
FOR THE DEGREE OF
DOCTOR OF PHILOSOPHY
ii
Abstract
Carbon capture and storage (CCS) is a key technology to reduce CO2 emissions
from industrial processes, in particular from fossil-fuel based electricity generation.
One important aspect of CCS is the safe long-term storage of the captured CO2 in
geological formations, especially in deep regional saline aquifers. Predicting the long-
term evolution of the injected CO2 requires an understanding of the basic physical
mechanisms and the ability to capture them in field-scale numerical simulations.
Simple mathematical models of trapping processes are developed to allow the
identification of the dominant physical processes during CO2 storage and their as-
sociated length and time scales. First-order estimates of the duration of the active
storage period and the migration distance are obtained as a function of the average
properties of the aquifer. These estimates support the selection of storage sites, in
particular at the early stages when limited data is available. They also show that
the length scales associated with the physical processes in regional aquifers can span
several orders of magnitude.
Multiscale simulation techniques are necessary to resolve physical processes and
geological heterogeneity. In particular, robust multiscale methods of elliptic flow
problems, must be developed. The multiscale finite volume method is analyzed in
the context of multipoint flux approximations and shown to lose monotonicity for
anisotropic problems. Strong anisotropy arises in the simulation of CO2 storage,
because of the large aspect ratios of regional aquifers. A new compact coarse operator
and new local fine-scale problems are introduced to obtain monotone coarse pressure
solutions for anisotropic domains. This development presents a major step towards
multiscale simulation of CO2 storage in large regional saline aquifers.
iv
Acknowledgements
I would like to thank my PhD advisors Hamdi Tchelepi and Franklin Orr Jr. who
have provided the best possible intellectual environment for me. The foundation of
this dissertation is their enthusiasm for open enquiry, their appreciation of a good
argument, their willingness to question common assumptions, their patience when
results were not forthcoming, and their trust in my abilities that has allowed me
to explore and develop my own ideas. They have set an example that I hope to
emulate in the future. I am also grateful that they have always encouraged me to
seek collaborations with other scientists, which has lead to a rich and intellectually
diverse graduate experience. Amongst my collaborators I would like to acknowledge
Amir Riaz and Brad Mallison, who have shared their knowledge and experience with
me and who have made significant intellectual contributions to my dissertation. I
would also like to thank Prof. Brian Cantwell for his encouragement and interest in
my work, and Prof. Patrick Jenny who hosted me at ETH-Zürich in the summer of
2008 and who has allowed me to participate in the development of iterative multiscale
methods.
I want to thank Roland Horne, who carefully assigns students to each office, for
my office mates Rami Younis and Felix Kwok. Our office has been a stimulating place
to work that has been central to my graduate experience. I thank Felix and Rami
for their friendship, and I will miss their company. I hope our paths will cross in the
future. I would also like to thank my parents and my family for their support with all
my decisions and their love. Finally, I would like to thank Valentina for all her love
and the happiness she has brought to my life in the last two years of this dissertation.
v
Contents
Abstract iv
Acknowledgements v
1 Introduction 1
1.1 Carbon capture and storage . . . . . . . . . . . . . . . . . . . . . . . 1
1.1.1 Clean coal technology . . . . . . . . . . . . . . . . . . . . . . . 2
1.1.2 Geological CO2 storage . . . . . . . . . . . . . . . . . . . . . . 3
1.2 Saline aquifer storage . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.2.1 Storage security . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.2.2 Leakage mechanisms . . . . . . . . . . . . . . . . . . . . . . . 8
1.2.3 Trapping mechanisms . . . . . . . . . . . . . . . . . . . . . . . 9
1.3 Thesis overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.3.1 Trapping models . . . . . . . . . . . . . . . . . . . . . . . . . 13
1.3.2 Multiscale simulation . . . . . . . . . . . . . . . . . . . . . . . 14
vi
2.2.6 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
2.3 Convective dissolution rates . . . . . . . . . . . . . . . . . . . . . . . 48
2.3.1 Convection in open systems . . . . . . . . . . . . . . . . . . . 49
2.3.2 Convection in closed systems . . . . . . . . . . . . . . . . . . . 51
2.3.3 Convective time scales . . . . . . . . . . . . . . . . . . . . . . 53
2.3.4 Further work . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
vii
4.1.2 Incompressible two-phase flow in porous media . . . . . . . . . 122
4.1.3 Multiscale methods . . . . . . . . . . . . . . . . . . . . . . . . 124
4.1.4 Fine-scale problem . . . . . . . . . . . . . . . . . . . . . . . . 125
4.1.5 Multipoint flux approximations . . . . . . . . . . . . . . . . . 127
4.2 Multiscale finite volume method . . . . . . . . . . . . . . . . . . . . . 129
4.2.1 Coarse operator . . . . . . . . . . . . . . . . . . . . . . . . . . 130
4.2.2 Local problems . . . . . . . . . . . . . . . . . . . . . . . . . . 132
4.2.3 Fine-scale reconstruction . . . . . . . . . . . . . . . . . . . . . 134
4.2.4 MSFV properties . . . . . . . . . . . . . . . . . . . . . . . . . 137
4.3 Anisotropy and monotonicity . . . . . . . . . . . . . . . . . . . . . . 138
4.3.1 Effective tensor . . . . . . . . . . . . . . . . . . . . . . . . . . 138
4.4 Homogeneous coarse operator . . . . . . . . . . . . . . . . . . . . . . 141
4.4.1 Original operator . . . . . . . . . . . . . . . . . . . . . . . . . 141
4.4.2 Compact operator . . . . . . . . . . . . . . . . . . . . . . . . . 144
4.5 Heterogeneous compact operator . . . . . . . . . . . . . . . . . . . . 146
4.6 Localization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148
4.6.1 Transverse equilibrium . . . . . . . . . . . . . . . . . . . . . . 149
4.6.2 Bilinear local problems . . . . . . . . . . . . . . . . . . . . . . 150
4.6.3 Linear local problems . . . . . . . . . . . . . . . . . . . . . . . 152
4.7 Numerical results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154
4.8 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162
4.9 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166
5 Conclusions 167
5.1 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
5.1.1 Dissolution trapping & convection . . . . . . . . . . . . . . . . 168
5.1.2 Residual trapping & gravity currents . . . . . . . . . . . . . . 169
5.1.3 Multiscale methods for porous media . . . . . . . . . . . . . . 170
5.2 Future work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 170
viii
List of Figures
1.1 (a) Predicted emission scenarios. The line labeled 500 ppm is the
emissions scenario necessary to stabilize the CO2 concentration at ap-
proximately twice the preindustrial average. (b) Schematic version of
figure a, showing the 7 stabilization wedges. Both graphs are modified
from Pacala & Socolow (2004). . . . . . . . . . . . . . . . . . . . . . 2
1.2 The different states of CO2 in the underground together with the phys-
ical and chemical processes responsible for the transfer between these
states. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.3 A schematic illustration showing the evolution of the fraction of CO2
in each state as a function of time, modified from Metz et al. (2006). 9
1.4 Sketches of the limiting cases investigated with simple process models:
(a) Dissolution trapping of CO2 ponded in structural trap. (b) Resid-
ual trapping of CO2 migrating in a horizontal aquifer. (c) Residual
trapping of CO2 migrating in a sloping aquifer. In all cases CO2 plume
is dark gray, residual CO2 light gray, and dissolved CO2 shaded. . . . 13
ix
2.2 Disturbance concentration profile for Ra = 500 and k = 0 at different
times, obtained from, (a) the IVP in the ξ-t coordinate system with
(2.21-2.23) and (b) the IVP in the z-t coordinate system with (2.14-
2.18). Random initial conditions are used for both cases. Plot (a)
shows that the concentration profile resolves into the correct solution
within a very short time, starting from any given set of white noise
initial conditions. For the z-t coordinate system on the other hand,
plot (b), convergence is much slower and is strongly dependent on
initial conditions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
2.3 (a) Growth rate vs. wavenumber curves for Ra = 500 computed by the
dominant mode method and the IVP. The flow is stable for small times.
The growth rate increases with time to become positive at the critical
time tc and the critical wavenumber kc . Flow instability increases with
time beyond tc with both shortwave and longwave cutoffs. Compar-
ison with the IVP shows exact agreement for small times and small
wavenumbers. (b) Comparison of the dominant mode method and the
QSSA with the initial value problem for two perturbation wavenum-
bers at Ra = 500. The dominant mode solution gives exact results for
small times but becomes inaccurate for later times particularly for large
wavenumbers. The QSSA on the other hand is reasonably accurate for
all times. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.4 (a) Critical time vs. Ra. (b) Critical wavenumber vs. Ra, given by the
dominant mode solution. The critical time varies as the 1/Ra, while
the critical wavenumber scales linearly with Ra. . . . . . . . . . . . . 31
2.5 (a) Maximum growth rate σmax as a function of time, and (b) The
most dangerous wavenumber kmax as a function of time, for different
Rayleigh numbers. All the results are obtained using the QSSA. . . . 32
2.6 (a) The longwave cutoff wavenumber kL as a function of time for dif-
ferent Rayleigh numbers, computed by the dominant mode method.
kL behaves similarly for all Ra and varies approximately linearly with
time. (b) The shortwave cutoff kS computed by the QSSA. . . . . . . 33
x
2.7 Comparison of the growth rates obtained from nonlinear direct numer-
ical simulation and the linear initial value problem, for two different
initial perturbations with wavenumber, k. . . . . . . . . . . . . . . . 35
2.8 Concentration contours at different times for Ra = 4000. (a) t = 1,
(b) t = 1.8, (c) t = 2.3 and (d ) t = 3.8 A large number of fingers
consistent with the linear stability analysis develop initially. Nonlinear
interactions rapidly reduce the number of fingers and give rise to large
scale structures at later times. The fingers at later times are connected
to the top boundary at discrete locations. These connections act as
feeding sites of the high density fluid to the convecting fingers below. 37
2.9 Vorticity contours in the background with overlapping streamlines for
Ra = 4000, t = 2.3. Corresponding concentration contours are shown
in figure 2.8(c). The vorticity field has a dipole structure that drives the
high density fluid through the fingers. The stream lines show how the
fluid travels laterally and then descends through the isolated feeding
sites for the fingers. The direction of fluid circulation is shown by the
arrows. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
2.10 The dominant mode of the nonlinear simulations as a function of time
for different values of the Rayleigh number. Also shown are the most
dangerous modes given by the linear stability analysis. Good agree-
ment is observed between the two for early times. At later times the
onset of nonlinear behavior leads to a significant deviation from linear
results. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
2.11 Concentration contours for (a) Ra = 1000, t = 2.2 and (b) Ra =
8000, t = 1.6. The larger Rayleigh number case is marked by intense
nonlinear interaction between competing fingers. The small Ra case
shows larger fingers with a strong diffusive spreading. . . . . . . . . . 41
xi
2.12 Concentration contours at time tb , when the concentration front reaches
the bottom. (a) Ra = 1000, tb = 8.9. (b) Ra = 2000, tb = 7. (c)
Ra = 4000, tb = 8.0. (d ) Ra = 8000, tb = 8.5. The late time dy-
namics is governed by large scale fingers. Large Rayleigh number cases
continue to display vigorous fingering interactions. . . . . . . . . . . . 42
2.13 The position of the most advanced portion of the front as a function
of time for different values of the Rayleigh number. The tip moves
faster for larger Rayleigh numbers for early times. Fingering interac-
tions significantly influence the rate of front propagation at later times,
particularly for large Ra cases. . . . . . . . . . . . . . . . . . . . . . . 43
2.14 (a) Variation with permeability of the critical time, tc , the critical
wavelength, λc and the penetration depth at the critical time, δc , for
∆ρ = 5 kg/m3 , φ = 0.3, µ = 0.5 cP and D = 10−9 m2 /s. For this range
of permeability variation tc varies between 2000 yrs and 10 days while
λc goes from about 100 m to less than a meter. The penetration depth
δc gives information regarding the applicability of our analysis with
respect to the layer thickness H, such that δc H. (b) The advance
of the fastest finger tip is shown until the bottom of the simulation
domain is reached. A diffusive, t0.5 , behavior is observed at early times,
while the fingers advance proportional to t at later times. . . . . . . . 45
2.15 Evolution of the concentration in an open aquifer for Ra = 4000. . . . 49
2.16 (a) Evolution of the dissolution rate at Ra = 4000. Average dissolution
rate over time is shown as dashed line. (b) Time averaged of the
dissolution rate for different Ra. . . . . . . . . . . . . . . . . . . . . . 50
2.17 Evolution of the concentration in a closed aquifer, for Ra = 1000. (a)
Plumes of dissolved have not reached the base of the aquifer. Convec-
tion similar to an open aquifer. (b-d ) Convection in a closed aquifer. . 52
2.18 (a) Evolution of the dissolution rate, dC/dτ , for various Ra in a closed
aquifer. (b) Variation of the time scales discussed in the text as a
function of Ra. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
xii
3.1 (a) Microtomogram of non-wetting CO2 bubbles (black) and water
(white) in the pore space between grains (grey), from Benson et al.
(2006). (b) Comparison between non-wetting phase saturation profile
after imbibition, inferred from a wire-line log (Koperna & Kuuskraa,
2006), and the simplified step function saturation profile corresponding
to the sharp-interface approximation. (c) The geometry of the porous
layer and the variables used in the derivation are shown. . . . . . . . 60
3.2 The geometry of the initial condition and the three associated length
scales H, Ld , and Lf . A particular initial condition h0 [x] and the
corresponding idealized step function initial condition h∗0 [x] are shown.
The arrows indicate that the fluid has been injected over the entire
thickness of the reservoir. . . . . . . . . . . . . . . . . . . . . . . . . . 63
3.3 The evolution of buoyant CO2 -vapor (gray) released into a horizontal
porous layer saturated by brine (white). The outward propagating
tip of the interface is marked by xo . All figures are exaggerated in
the vertical direction, to make the late solution (d ) visible. In many
situations of interest the width of the invaded region in figure (a) is
several times larger than the aquifer thickness. . . . . . . . . . . . . . 68
3.4 Similarity solution is shown for different values of M. The variable θ
is the non-dimensional thickness of the CO2 , it is measured from the
top of the aquifer. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
3.5 The numerical values for the outer (a) and inner (b) tip positions are
shown as a function of M (solid lines). The scaling laws (3.17) and
(3.18) are shown as dashed lines. . . . . . . . . . . . . . . . . . . . . 71
3.6 (a) The numerical solution to (3.2) with initial condition (3.22) and
M = 10 is plotted at various times, t, and compared to the self-similar
solution obtained from (3.15) (solid grey line). (b) Numerical solutions
to (3.2) with M = 20 and initial condition (3.25) is shown for several
values of the parameter a in the initial condition and compared to the
solution of (3.15) with a = 0 (solid grey line). . . . . . . . . . . . . . 73
xiii
3.7 The transition of the numerical solution (light lines) from the early to
the late similarity solution (both heavy lines) is shown in scaled coordi-
nates. (a) The numerical solution is shown at ϑ = ϑb +{1, 4, 10, 30, 100},
where ϑb = 0.61; (b) The numerical solution is shown at ϑ = ϑb +
{101 , 102 , 103 , 104 , 105 }, where ϑb = 7.1. . . . . . . . . . . . . . . . . . 79
3.8 The numerical results for the non-dimensional tip position ξo are shown
as a function of non-dimensional time ϑ, for different mobility ratios
M. In all figures the numerical solution is given by dots (· · · ), the tip
scaling from the early similarity solution by a dashed line (- - -), and
the tip scaling from the late similarity solution as a full line (—). (a)
ϑt = 2.0, ϑb = 0.1; (b) ϑt = 2.5, ϑb = 0.24; (c) ϑt = 29.3, ϑb = 1; (d )
ϑt = 811.3, ϑb = 1; (d ) ϑt = 811.3, ϑb = 7.1. . . . . . . . . . . . . . . 81
3.9 Regime diagram for a finite release of fluid into a horizontal porous
slab, showing the non-dimensional time scales obtained in this study,
and the shapes of the gravity current as a function of the mobility
ratio M. The shaded region indicates the transition period between
the similarity solutions. The characteristic time to dimensionalize all
results is td = L2d κ−1 H −1 . . . . . . . . . . . . . . . . . . . . . . . . . 84
3.10 Current evolution in an aquifer with 30 mD permeability, for details of
the physical properties see § 3.5.2. CO2 currents (dark grey) and the
zone containing residual saturation in their wake. Figures a to d show
the evolution in a horizontal aquifer. These figures are equivalent to
figure 3.3, but they include the effect of residual trapping. Figures e
to h show the evolution in an equivalent aquifer sloping with 5◦ . . . . 86
3.11 (a) Interface shapes for M = 1 and increasing from 0 to 1 in 0.2
increments. (b) Interface shapes for M = 10 and increasing from 0
to 1 in 0.2 increments. (c) Contours of the outward propagating tip
position, ζo . (d ) Contours of the inward propagating tip position, -ζi . 88
xiv
3.12 Change of the solution from the early to the late scaling law (M = 1).
(a) The scaling laws for ξo ∝ ϑβ at early and late times as a function
of the residual, . The symbols at of 0, 1/4, 1/2, and 3/4 show the
exponents obtained by fitting the numerical data shown in figure b. For
the early scaling law the data was fit for ϑ between 0.1 and 10 and for
the late scaling law the data was fit for ϑ > 1000. (b) The shifted tip
position, ξˆ = ξo − ξ,
˜ is shown as function of ϑ, for increasing residuals,
where ξ˜ = 1. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
3.13 (a) The flux function f is shown for M = 10 and increasing trapping .
(b) The corresponding solution profiles given by (3.56) in self-similar
coordinates. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
3.14 Shock construction for M = 10. (a) In the absence of trapping, = 0,
the construction for advancing and receding shocks is the same. (b) In
the presence of trapping, = 1/2, the shock construction for advancing
shocks is modified. (c) Shock speed as function of ηL , for ηL > ηc the
shock speed depends on the amount of trapping. . . . . . . . . . . . . 97
3.15 Characteristic portraits of the evolution of initial data (3.71), in the
ξτ -plane. The current is bounded by the shock, A to D, in the back
and the leading characteristic of the rarefaction, F to D, in the front.
(a) Analytic solution for M = 1 and = 0.4. (b) Semi-analytic solution
for M = 5 and = 0.4. (c) Evolution of the current interface (black)
and the residual surface (grey). . . . . . . . . . . . . . . . . . . . . . 100
3.16 (a) Portrait of the shock path and the leading characteristic of the rar-
efaction, for increasing residual. The locus of the termination points,
τf = τf [ξ↑ ], is shown as a dotted line. (b) The corresponding tem-
poral evolution of the shock strength, ηs . (c) The corresponding final
residual surfaces, ηr = ηr [τf ]. . . . . . . . . . . . . . . . . . . . . . . 102
3.17 Contour plots of the up-dip migration distance, log(ξ↑ ), on the left and
of the total migration time, log(τf ), on the right. The data for several
potential storage aquifers in Alberta, Canada, is shown as circles to
indicate typical parameter values (Bachu & Bennion, 2007). . . . . . 103
xv
3.18 The effect of the governing parameters on the evolution of the current
volume. The passing and final migration times, are shown as circles
and triangles. (a) Increasing the residual, , from 0.1 to 0.9 at constant
M = 1. (b) Increasing the mobility ratio, M, from 1 to 40 at constant
= 0.4. Circles denote . . . . . . . . . . . . . . . . . . . . . . . . . . 105
3.19 This figure shows several comparisons between the numerical solution
(∆ξ = 0.01) to (3.7a) and the hyperbolic solution for Pe = 2 at τ = 50.
(a) Increasing M at constant = 0. (b) Increasing at constant M = 10.106
3.20 Evolution of the current volume, V, as a function of diffusive dimen-
sionless time, ϑ, for M = 5 and increasing Pe. The time axis at the
top gives dimensional values corresponding to the parameter choice
discussed in § 3.5.2, assuming cos θ ≈ 1. . . . . . . . . . . . . . . . . . 107
3.21 Time and length scales obtained from the numerical solution of (3.7)
are shown as symbols (∆ξ = 0.01). All results are for a residual of
= 0.4. The limiting hyperbolic and self-similar solutions are indicated
by full and dashed lines. (a) Down-dip migration distance, ξ↓ . (b)
Transition time, τt . (c) Up-dip migration time, ξ↑ . (d ) Final migration
time, τf . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
3.22 Contour plots of the critical angle, θc , assuming d0 = 2 and the sweep,
S, as function of M and . The data for several potential storage
aquifers in Alberta, Canada, is shown as circles to indicate typical
parameter values (Bachu & Bennion, 2007). . . . . . . . . . . . . . . 115
xvi
4.2 (a) Nine cells of the finite volume grid are shown by dashed lines. The
four dual interaction regions, Ω̃(1) to Ω̃(4) , which connect the shaded
finite volume cell in the center to its eight nearest neighbors are shown
by solid lines. (b) A dual interaction region is shown with the four
coarse cell boundary segments labeled s1 to s4 and the corresponding
fluxes f1 to f4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
4.3 The top row shows contour plots of the four standard basis functions of
the MSFV method, calulated numerically for a heterogeneous, isotropic
fine permeability distribution. The second row shows the underlying
dual interaction region with the fine grid and the fine fluxes t∗m,k along
the coarse boundaries (small arrows) and the total coarse fluxes tm,k
integrated over each face (big arrows). . . . . . . . . . . . . . . . . . 133
4.4 The fine-scale solution as well as the reconstructed conservative and
the interpolated multiscale solutions are shown for the same problem.
The log of the permeability is shown at the base of each figure. A
unit pressure gradient is applied in the x-direction, and homogeneous
Neumann boundary conditions are applied in the y-directions. (a)
The fine-scale solution, ph . (b) The conservative solution from the
flux reconstruction on the fine grid, p̄h . (c) The coarse solution, pH ,
interpolated onto the fine grid using the MSFV basis functions, p̂h . In
(b) and (c) the coarse solution, pH , is shown as large black dots. . . . 136
4.5 Three solution profiles in the x-direction located at y = 1/2 are shown
for increasing values of the ratio a/b with c = 0. (a) anisotropy ratio,
a/b = 1 > 1/3, corresponding to A = ∆x/∆y = 1 . (b) anisotropy
ratio, a/b = 1/100 < 1/3, corresponding to A = 10. (c) anisotropy
ratio, a/b = 1/2500 < 1/3, corresponding to A = 50. . . . . . . . . . . 139
4.6 Numerical results for the stencil components m1,0 and m1,1 of the
MSFV method are compared to the analytic expressions of the MPFA
O(1) method. The components are shown as a function of the orien-
tation of the tensor, θ, for several ratios of the principal components,
λ1 /λ2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
xvii
4.7 The monotonicity regions of the original and compact coarse operator
are shown in the parameter space a/b vs. |c|/b. A gray grid shows the
corresponding angle, θ, and the ratio of principal components, λ1 /λ2
in the effective permeability tensor given by (4.35). All methods are
monotone above the labeled line. (a) The monotonicity regions of both
the MPFA O(1) method and the coarse operator of the MSFV method
are compared. (b) The monotonicity region of the CMSFV method
equals that of the 7-point stencil. . . . . . . . . . . . . . . . . . . . . 143
4.8 The triangular domains of dependence of the fluxes in a dual interaction
region are shown for both signs of c. The corresponding orientation of
the permeability tensor is indicated by the ellipses. . . . . . . . . . . 145
4.9 The approach of the global fine-scale solution to transverse equilibrium,
as the effective anisotropy a/b increases. The boundary conditions for
this problem are p(x = 0, y) = 1, p(x = 1, y) = 0, and py (x, y = 0, 1) =
0 and the effective permeability on the fine-scale is diagonal, c = 0.
The log of the permeability field is shown at the base. . . . . . . . . . 148
4.10 Comparison between the fine-scale solution ph and the interpolated
pressure field p̂h = φh1 + φh4 in a dual crossing the high permeabil-
ity channel at a/b = 1/1000. (a) Fine-scale solution, ph , in the dual,
rescaled to a unit pressure gradient, is in transverse equilibrium. (b)
Interpolated fine pressure p̂h with reduced boundary conditions. (c)
Interpolated fine pressure p̂h with linear boundary conditions. (d) In-
terpolated fine pressure p̂h with hybrid boundary conditions. . . . . . 151
4.11 (a) The unit square domain Ω, with a 20 × 20 coarse grid. The back
squares indicate coarse blocks with sources fH = ±1. In these blocks
the fine-scale permeability is constant at the mean to give a uniform
distribution of the fine-scale flux. (b) A realization of the permeability
field with small correlation length. Figures c and d show realizations of
the layered permeability fields at angles of 15◦ and −30◦ respectively.
The logarithm of the permeability is shown in all cases. . . . . . . . . 154
xviii
4.12 Comparison of time of flight maps between multiscale methods and the
fine-scale solution for three isotropic but heterogeneous fine-scale per-
meability fields. All solutions are shown at the same injected volume,
corresponding to the breakthrough of the fine-scale solution. The fine-
scale solution, fs, is shown in the first row, a to c, the original MSFV -
red method in the second row, d to f, the CMSFV -lin method in the
third row, g to i, and the CMSFV -hyb method in the fourth row, j
to l. The first column shows solutions for a realization of the perme-
ability field shown in figure 4.11b, columns 2 to 3 show solutions for
realizations of layered permeability fields at 15◦ and -30◦ (figure 4.11c
and d). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
4.13 Comparison of time of flight maps between multiscale methods and the
fine-scale solution for three heterogeneous fine-scale permeability fields
with grid aligned fine-scale anisotropy, a/b = 0.01. All solutions are
shown at the same injected volume, corresponding to the breakthrough
of the fine-scale solution. The figure layout is the same as in figure 4.12.156
4.14 Figures a and b show the mean L2 -errors of the interpolated solution p̂h
and the reconstructed flux field ūh = −k∇p̄h as function of the orien-
tation, θ, of the fine-scale layering for isotropic fine-scale permeability,
a/b = 1. Figures c and d show the same for grid aligned fine-scale
anisotropy a/b = 0.04 (∆x/∆y = 5). All values shown are means over
20 realizations and the standard deviations are shown as vertical bars,
if they are larger than the symbols. . . . . . . . . . . . . . . . . . . . 158
4.15 Figures a and b show the mean L2 -errors of the interpolated solution
p̂h and the reconstructed flux field as function of increasing fine-scale
anisotropy a/b. All values shown are means over 20 realizations of the
permeability field shown in figure 4.11a. . . . . . . . . . . . . . . . . 160
xix
4.16 Results for the quarter 5-spot shown in figure 4.11a with homogeneous
Dirichlet boundary conditions. The first column shows contour maps
of p̂h , the second column shows a horizontal view of p̂h along the y-axis,
and the third column shows streamlines obtained from the conservative
reconstruction p̄h . Figures a to c show the results for MSFV-lin, figures
d to f show results for MSFV-hyb, and figures g to i show results for
CMSFV-hyb. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
4.17 The L∞ and L2 -norms of M (LH ) are shown as a function of the angle
of the layering. Figures a and b show results for a isotropic fine-scale
permeability field (a/b = 1), while figures c and d are for a fine-scale
anisotropy a/b = 1/100. . . . . . . . . . . . . . . . . . . . . . . . . . 163
xx
Chapter 1
Introduction
1
CHAPTER 1. INTRODUCTION 2
Figure 1.1: (a) Predicted emission scenarios. The line labeled 500 ppm is the emis-
sions scenario necessary to stabilize the CO2 concentration at approximately twice the
preindustrial average. (b) Schematic version of figure a, showing the 7 stabilization
wedges. Both graphs are modified from Pacala & Socolow (2004).
approximately 14 GtC/yr in 2060 (figure 1.1a). Pacala & Socolow (2004) have intro-
duced the concept of a stabilization wedge (figure 1.1b), as an activity that reduces
emissions to the atmosphere that starts at zero today and increases linearly until it
accounts for 1 GtC/yr of reduced carbon emissions in 50 years. A few examples of
such activities are: efficient vehicles and buildings, substitution of gas baseload power
for coal baseload power, substitution of nuclear, wind or photovoltaic power for coal
power, biomass fuel for fossil fuel, and CO2 capture at baseload power plants, H2
plants or coal-to-synfuel plants. Seven stabilization wedges are necessary to avoid a
doubling of preindustrial atmospheric CO2 concentrations. Currently, no single avail-
able technology can provide even half of the necessary emission reductions. However,
Pacala & Socolow (2004) argue that the necessary reductions in carbon emissions are
already possible with a portfolio of options available today. They have identified 15
activities, including those listed above, which can contribute one stabilization wedge
each.
Assuming that CCS prevents 90% of the fossil carbon from reaching the at-
mosphere, a stabilization wedge could be achieved by the installation of 800 GW
of baseload coal power plants with carbon capture. The associated geological storage
would require that current CO2 injection, for enhanced oil recovery, be scaled up by
a factor of 100 over the next 50 years, or the equivalent of 3500 storage projects
comparable to the current CO2 storage operations at the Sleipner-Vest natural gas
field, operated by Statoil (Torp & Gale, 2004).
basins. As of today, there are no economic benefits associated with storage in saline
aquifers that can help offset the storage cost. Therefore, other incentives, such as the
carbon-emissions tax in Norway, must be in place for the adoption of aquifer storage.
Since CO2 emission regulations, taxes, or both, are expected in the near future, saline
aquifer storage is currently the most promising large-scale geological storage option.
Therefore, this thesis focuses on accurate mathematical modeling of the complex flow
behaviors associated with CO2 storage in saline aquifers.
An impermeable seal, such as a shale or clay layer, overlying the storage formation is
therefore always necessary to prevent upward migration of the injected CO2 towards
the surface. To prevent upward migration of the CO2 along an inclined seal towards
the surface a structural trap, (figure 1.4a), is thought to be necessary for long-term
storage security. Based on these and other criteria, Bachu (2003) have proposed a
screening and ranking methodology to identify suitable storage aquifers.
CCS projects targeting saline aquifers are under discussion in many countries,
including Norway, the US, Japan, Canada, Australia, the UK, France, Germany,
Italy, and Poland (Metz et al., 2006). The Sleipner Project, a gas field operated by
Statoil on the Norwegian continental shelf, is a good example of CO2 storage in a
saline aquifer. Approximately 1 MtCO2 /yr is removed from the natural gas stream
produced at the Sleipner-Vest platform and injected into the Utsira formation. The
operation started in October 1996, and over the lifetime of the project a total of 20
MtCO2 is expected to be stored. The evolution of the Sleipner CO2 plume has been
successfully monitored by seismic time-lapse surveys. The surveys also show that the
caprock prevents migration out of the storage formation (Torp & Gale, 2004).
atmospheric CO2
leakage
geological seal
on re
uti g tra sidual
sol ppin ppi
tra ng
primary
trapping
dissolved CO2 residual CO2
trapping
mineral
solution
trapping
secondary
trapping
mineralized CO2
Figure 1.2: The different states of CO2 in the underground together with the physical
and chemical processes responsible for the transfer between these states.
CHAPTER 1. INTRODUCTION 8
Gasda et al. (2004) have identified abandoned wells as one of the most probable leak-
age pathways for CO2 storage projects, due to their high density in many sedimentary
basins, i.e., 350000 wells in the Alberta basin. Nordbotten et al. (2004) have used
semi-analytic solutions to show that leakage through multiple passive wells is not a
simple sum of single well leakage rates, due to leakage induced draw-down around
passive wells. They have also identified that multiple aquifers and aquitards mitigate
leakage into shallow zones, because of loss of CO2 into the intervening aquifers. Nu-
merical comparisons under less restrictive assumptions have shown that the inclusion
of additional processes leads to reductions in leakage rates (Ebigbo et al., 2006).
Pruess (2005) showed that non-isothermal effects are important during leakage
along faults and lead to a self-limiting behavior of the leakage rate. In case CO2
does leak into the shallow subsurface, CO2 will displace the lighter resident soil gas.
Therefore, CO2 concentrations may approach 100% in some parts of the vadose zone,
CHAPTER 1. INTRODUCTION 9
CO2 Reservoir %
natural
trapping
engineered
trapping
mineralized CO2
0
1 10 102 103 104 105 106
Figure 1.3: A schematic illustration showing the evolution of the fraction of CO2 in
each state as a function of time, modified from Metz et al. (2006).
even for small leakage fluxes (Oldenburg & Unger, 2005). Leakage of CO2 plumes is
similar to migration of plumes of volatile organic compounds, and similar remediation
approaches may be applicable (Zhang et al., 2005).
1. Residual trapping, which refers to the formation of residual CO2 through capil-
lary snap-off in the wake of a migrating immiscible CO2 plume. Residual CO2 is
still buoyant, but it is immobilized because it is disconnected from the flowing
saturation. Holtz (2002) was among the first to point out the potentially large
CHAPTER 1. INTRODUCTION 10
residual saturations in CO2 –brine systems. Kumar et al. (2005) showed that
the formation of residual saturation may be the dominant trapping mechanism
in saline aquifer storage, and they highlighted the importance of the magnitude
and variation of the residual saturation in the formation as a key petro-physical
property.
2. Dissolution trapping, which refers to the dissolution of CO2 into the brine. Dis-
solved CO2 is still mobile, through the migration of the CO2 -rich brine. How-
ever, the brine density increases with increasing CO2 saturation (Yang & Gu,
2006), and the negative buoyancy prevents leakage of dissolved CO2 (Lindeberg
& Wessel-Berg, 1997). Convective motion induced by dense plumes of CO2 -rich
brine may have the potential to increase the overall dissolution rate of the CO2
significantly (Ennis-King et al., 2005).
Figure 1.2 illustrates the relationship between the different CO2 trapping processes
and the states of CO2 in the subsurface. Mineralization offers the hope for perma-
nent storage on geological timescales, but CO2 is effectively trapped by dissolution
and residual trapping much earlier. Therefore, we refer to direct dissolution of the
mobile CO2 plume and residual trapping as primary trapping processes, and to the
dissolution of residual CO2 and the mineralization of dissolved CO2 as secondary
trapping processes. Figure 1.3 is a conceptual sketch of the evolution of the different
states of CO2 over time.
During the injection period, the CO2 plume contacts fresh brine as it is advancing
into the aquifer, leading to some CO2 dissolution. Residual trapping is expected
to be very effective in the period immediately following the end of injection, as the
CHAPTER 1. INTRODUCTION 11
CO2 migrates toward the top of the aquifer (Kumar et al., 2005). It is possible to
enhance trapping during the injection period and for some period after the end of
injection, leading to engineered trapping. Examples of engineered trapping include
water-alternating-gas injection (Kovscek & Cakici, 2005; Juanes et al., 2006; Ide et al.,
2007), and the circulation of brine to increase dissolution (Leonenko & Keith, 2008).
However, engineered trapping at large scales will increase the storage cost and is
therefore unlikely in aquifers, where economic benefits do not offset the increased
cost.
In most situations we have to rely on the trapping of CO2 by natural processes,
in the post-injection period. This reliance on natural trapping processes makes the
proper selection of the storage site very important. This thesis focusses on natural
trapping processes in the post-injection period. In situations where the immiscible
CO2 plume has ponded in a structural trap (figure 1.4a) and has come to rest, no
further residual saturation forms, and dissolution of CO2 into the underlying brine
is the main process reducing the volume of the stationary CO2 plume. As long as
the immiscible CO2 plume continues to migrate, CO2 dissolves into fresh brine and
residual CO2 forms in the wake of the moving CO2 plume (figures 1.4b-c).
We may also distinguish an active storage period, during which mobile buoyant
CO2 is present and leakage is possible, from a passive storage period, where CO2 is
only redistributed between trapped phases (figure 1.3). During the active period,
leakage and trapping processes compete to reduce the mobile buoyant CO2 volume.
The active period ends when all injected CO2 has either leaked or been trapped. The
time when a CO2 storage site switches from the active to the passive period is an
important time scale that provides an indication of the storage security of a particular
site.
1. Chapters 2 and 3 present simple mathematical models that aim to capture the
fundamental dynamics of dissolution and residual trapping.
Figure 1.4: Sketches of the limiting cases investigated with simple process models:
(a) Dissolution trapping of CO2 ponded in structural trap. (b) Residual trapping of
CO2 migrating in a horizontal aquifer. (c) Residual trapping of CO2 migrating in a
sloping aquifer. In all cases CO2 plume is dark gray, residual CO2 light gray, and
dissolved CO2 shaded.
porous media are not well developed. This is partly because of upscaling and down-
scaling problems when the mesh size is changed. Recently, multiscale methods that
work on a single coarse grid and incorporate finescale effects through numerical ba-
sis functions have been developed. Multiscale methods offer adaptivity through the
selective updating of basis functions. In § 4 we discuss the multiscale finite volume
(MSFV) method introduced by Jenny et al. (2003). The MSFV method is currently
the only multiscale method for porous media transport that has been extended to in-
clude wells (Lee et al., 2008), compressibility (Lunati & Jenny, 2006; Zhou & Tchelepi,
2008), and gravity (Lunati & Jenny, 2008). However, the original MSFV method is
only robust for uniform grids, which prevents the application to saline aquifers that
typically have very large aspect ratios. In chapter 4 we analyze the monotonicity of
the original MSFV method, and we develop a new Compact Multiscale Finite Volume
(CMSFV) discretization that is robust for large aspect ratios. Once these different
results have been combined into a single multiscale simulator, the MSFV method
should be able to simulate CO2 storage in saline aquifers at dramatically reduced
computational cost.
Chapter 2
2.1 Introduction
Dissolution of the supercritical CO2 into the brine is an important primary trapping
process that reduces the volume of the CO2 plume, and it therefore determines the
duration of the active storage period (figure 1.3). Lindeberg & Wessel-Berg (1997)
pointed out that the density of the brine increases with increasing CO2 saturation,
and Yang & Gu (2006) verified this for a CO2 –brine system at reservoir conditions.
Dissolved CO2 is considered trapped, because the CO2 -rich brine migrates downward
and leakage is not likely. The associated reduction in the total volume of the system
also allows the pressure in the aquifer to relax.
At an interface where both the supercritical CO2 and the brine are at rest the
dissolution of CO2 will be slow. In this case, the dissolution rate of the CO2 across
the interface is controlled by molecular diffusion of dissolved CO2 in the brine, and
we expect the dissolution rate of CO2 to decay as t−1/2 . Much more rapid dissolution
of CO2 occurs across an interface where the supercritical CO2 continues to contact
fresh brine, due to advection in either phase. Both the buoyancy driven migration
of the supercritical CO2 as well as convective motion in the brine can lead to rapid
dissolution.
16
CHAPTER 2. DISSOLUTION TRAPPING & CONVECTION IN THE BRINE 17
Once the CO2 plume has ponded in an anticline, only convective motion in the
underlying brine can increase the dissolution rate of the CO2 (figure 1.4a). We
assume that the supercritical CO2 is separated from the brine below by a horizontal
interface. Across this horizontal interface, CO2 dissolves into the brine to form a
diffusive boundary layer that grows with time (Lindeberg & Wessel-Berg, 1997). The
CO2 -rich brine in this boundary layer is heavier than the underlying fresh brine. Once
the boundary layer has grown to a sufficient thickness, it will become unstable, so
that fingers of dense CO2 -rich brine propagate downward. In this case the dissolution
rate is determined by convective mass transport in the brine.
We use both a linear stability analysis and direct numerical simulation to study
the onset of this instability. Linear stability analysis is a standard method used to
determine the stability of a particular base-state. In 2.2.3 we study the stability of
the concentration profile of dissolved CO2 in the brine below the interface. If the
concentration profile is unstable to small perturbations, convective motion in the
brine will develop. The advantage of linear analysis is that it can detect very small
unstable wavelengths, because it is not limited by numerical resolution. However,
the linear analysis is valid only at early times when the perturbations are small. In
§ 2.2.4 we validate the direct numerical simulations against the linear theory, and
we use simulations to investigate the nonlinear evolution of the instability at larger
times. These direct numerical simulations are used in § 2.3 to study the long-term
dissolution rate of CO2 in saline aquifers.
z=0 g
CO2 dissolved in brine
z
Brine
Figure 2.1: Supercritical CO2 accumulates along the impermeable top boundary. It
slowly dissolves into the underlying brine, forming a heavier boundary layer. The
resulting gravitational instability leads to the convective transport of CO2 saturated
brine plumes.
and the brine can be assumed across the interface. We assume, therefore, that the
supercritical CO2 layer acts as a horizontal upper boundary, with a constant CO2
concentration. These authors propose that the gravitational instability can be stud-
ied in the context of a finite domain, bounded at the top by a constant concentration
boundary, using the Boussinesq approximation. This problem is analogous to ther-
mal convection in a porous medium with insulated boundaries, that is rapidly heated
from below, or cooled from the top, and initially at a fixed temperature. Elder (1967)
showed that the diffusive layer at the boundary becomes unstable only after an ini-
tial period of decaying perturbations. After this critical time, plumes or fingers of
hot fluid rise into the porous medium. This onset time for convection is an impor-
tant timescale that determines the efficiency of dissolution trapping in saline aquifer
storage.
CHAPTER 2. DISSOLUTION TRAPPING & CONVECTION IN THE BRINE 19
For unbounded domains Pritchard (2004) and Ben et al. (2002) argue that a fun-
damental limitation on the accuracy of the solution at short times is the assumption
of decoupled normal modes in the streamwise direction. Since the disturbances are
confined within a narrow diffusive zone, the global Fourier eigenfunctions of the diffu-
sion operator do not provide an optimal basis for streamwise perturbations (Pego &
Weinstein, 1994). Ben et al. (2002) note that the Hermite polynomial based discrete
eigenspectrum of the self-similar diffusion operator is the natural basis for streamwise
perturbations in an unbounded domain. This eigenspectrum is the solution of the
stability problem for the zero wavenumber. The first mode is neutral, it has zero
growth rate, while the rest of the modes decay with time. Therefore, the first mode
becomes dominant after a relatively short time. For small wavenumbers, the per-
turbation dynamics can then be projected onto this dominant mode to increase the
growth rate from zero to positive values.
A similar approach is adopted here for the stability analysis in the semi-infinite
domain. We use the self-similar transform of the diffusion operator in the inhomoge-
neous direction with localized eigenfunctions. The zero wavenumber solution shows
that the first mode of the self-similar diffusion operator decays with time, while the
rest of the spectrum decays more rapidly. We use this dominant first mode to capture
the perturbation dynamics for larger wavenumbers, which shift the growth rate from
negative to positive values at later times. We show that in contrast to the longwave
instability in an unbounded domain, a critical time, critical wavenumber, and a long
wavelength cutoff characterize the stability problem in the semi-infinite domain.
It is important to note that the dominant mode solution becomes inaccurate for
large times and large wavenumbers. For these cases, we use the QSSA in self-similar
coordinates, and we show by comparion with solutions to the IVP that the QSSA
gives good results. Compared to the poor accuracy in the original coordinates, the
success of both QSSA and IVP in self-similar coordinates is due to the localized basis
functions in the streamwise direction.
In order to analyze the long-term evolution of the unstable modes predicted by
the linear stability theory, we carry out high-accuracy direct numerical simulations
(DNS) using a vorticity based formulation (Tan & Homsy, 1988). Voss & Sousa
CHAPTER 2. DISSOLUTION TRAPPING & CONVECTION IN THE BRINE 21
(1987) have also carried out numerical simulations of density-driven flows. We use
the methodology developed by Ruith & Meiburg (2000) using Fourier-Galerkin de-
composition in the lateral direction and compact finite differences in the streamwise
direction to solve for the velocity field. A 4th order Runga-Kutta method is used
for time integration. The nonlinear simulations are validated by comparison to the
linear stability analysis. Excellent agreement is observed at early times for growth
rates associated with individual wavenumbers. The long time behavior, however, is
characterized by complex nonlinear interactions that require DNS.
K
u = − (∇P − ρgẑ) , (2.1)
µ
∂C
φ = −u · ∇C + φD∇2 C , (2.2)
∂t
∇·u = 0 , (2.3)
ρ = ρ0 + ∆ρC . (2.4)
The normalized concentration C of the heavier fluid, u = (u, w) is the Darcy velocity.
K is the permeability, D is the diffusion coefficient, φ is the porosity, and µ is the
viscosity. The unit vector in the direction of the gravitational acceleration is ẑ.
Density ρ is specified as a linear function of concentration, and ρ0 is the density of
the lighter fluid. The initial conditions are u(x, z, t = 0) = 0 and C(x, z, t = 0) = 0,
and the boundary conditions are given by
CHAPTER 2. DISSOLUTION TRAPPING & CONVECTION IN THE BRINE 22
w(x, z = 0, t) = 0 , w(x, z = H, t) = 0 ,
∂C
C(x, z = 0, t) = 1 , =0.
∂z x,z=H,t
K∆ρg
U = , (2.5)
µ
ρ∗ = ρ1 − ρ0 = ∆ρ , (2.6)
µU H
P∗ = = ∆ρgH , (2.7)
K
φH φµH
t∗ = = . (2.8)
U K∆ρg
where H is the domain thickness and U is the buoyancy velocity. The corresponding
dimensionless equations are
∇·u = 0 , (2.9)
u = − (∇P − Cẑ) , (2.10)
∂C 1 2
= −u · ∇C + ∇C. (2.11)
∂t Ra
w(x, z = 0, t) = 0 w(x, z = 1, t) = 0 ,
∂C
C(x, z = 0, t) = 1 =0.
∂z x,z=1,t
CHAPTER 2. DISSOLUTION TRAPPING & CONVECTION IN THE BRINE 23
We eliminate pressure by taking the curl of (2.10) and substitute the transverse
velocity from the continuity equation. The streamwise velocity, w, and concentration
are decomposed into the base-state and perturbation components
The base velocity, wo , is zero, and the base concentration, Co , is assumed constant
in the x-direction and given by the solution of (2.11), with u = 0. The perturbation
variables are decomposed into eigenfunctions, which depend on time and the stream-
wise coordinate, and normal modes in the transverse x-direction with wavenumber k.
The linearized perturbation equations can then be expressed as
∂2
2
− k ŵ = −k 2 ĉ ,
2
(2.14)
∂z
2
∂ĉ 1 ∂ 2 ∂Co
− 2
− k ĉ = − ŵ , (2.15)
∂t Ra ∂z ∂z
4 1
∞
2 2
Co (t, z) = 1 − sin ((n − 1/2)πz) e−(n−1/2) π t/Ra .
π n=1 2n − 1
The length δ(t) over which C0 is significantly different from zero is the so-called
“penetration depth” of the diffusive boundary layer. For δ ∝ 4t/Ra 1, the
domain can be considered semi-infinite in the positive z-direction, and the base-state
is given by
Ra
Co (z, t) = 1 − erf z on z ∈ (0, ∞) (2.18)
4t
CHAPTER 2. DISSOLUTION TRAPPING & CONVECTION IN THE BRINE 24
Co (z = 0, t) = 1 , (2.19)
Co (z → ∞, t) → 0 . (2.20)
The problem has now been redefined as the stability of a diffusive boundary layer
in a semi-infinite domain. The parameter range over which the results are valid,
for the original layer geometry in a finite domain, is given by δ ∝ 4t/Ra 1.
Note that the semi-infinite domain does not impose any external length scale on the
problem. An internal length scale in the problem is the time dependent penetration
depth δ(t). A Rayleigh number based on δ(t) is itself time dependent, such that a
critical Rayleigh number, Rac = Ra(δ(tc )), is merely a function of the critical time,
tc , at which the boundary layer becomes unstable. On the other hand, the Rayleigh
number can also be scaled out of the equations by specifying the length scale as a
ratio of diffusion to buoyancy velocity, H = D/U . In this case, the critical time, tc ,
is the only criterion for the onset of instability. For our analysis, we choose to work
with the latter. However, we retain the Rayleigh number, with an arbitrary length
scale H, for convenience in comparing the linear stability analysis with the nonlinear
results discussed in § 2.2.4. In the discussion of the dimensional results (§ 2.2.5), we
show that the imposed length scale cancels out, so that the choice of length scale is
arbitrary.
The perturbation equations (2.14-2.17) can be solved in a straight forward manner
using the quasi-steady-state approximation (QSSA). Earlier investigations have shown
that the QSSA gives accurate results for relatively long times. A fundamental problem
with such an approach is that the concentration eigenfunctions are localized in the
boundary layer, while the eigenfunctions of the operator ∂ 2 /∂z 2 , z ∈ (0, ∞) have
global support. Hence, they do not provide an appropriate basis for streamwise
perturbations (Chang et al., 1998). Therefore, a finite time is required, in the initial-
value-problem (IVP), before the global eigenfunctions can accurately represent the
localized structure of the streamwise perturbations. Since the critical time of this
problem can relatively short, the QSSA in the original coordinate system may not
CHAPTER 2. DISSOLUTION TRAPPING & CONVECTION IN THE BRINE 25
Note that the self-similarity applies only to the base concentration. The amplitude
and the spatial structure of perturbations are time dependent. The streamwise oper-
ator of the concentration perturbation in the transformed coordinate, ξ, is
1 ∂2 ξ ∂
L= 2
+ , ξ ∈ (0, ∞) . (2.26)
4 ∂ξ 2 ∂ξ
∞
ĉ(ξ, t) = An (t)φn (ξ) , (2.27)
n=1
CHAPTER 2. DISSOLUTION TRAPPING & CONVECTION IN THE BRINE 26
(a) (b)
1
t=10-5 t=0.1
t=10-4 t=0.5
1 t=10-3 t=1.0
0.5
c 0.5 c0
〈
〈
0 -0.5
-1
0 1 2 3 4 5 0 0.2 0.4 0.6 0.8 1
ξ z
with
2
Lφn = λn φn (ξ) = λn e−ξ Hn (ξ) ; n = 1, 2, 3, ..... (2.28)
dAn
= λn An . (2.29)
dt
All modes decay as t−n for k = 0, so that the flow is stable in the longwave limit.
2
The perturbation eigenfunction related to the largest eigenvalue, φ1 = ξ e−ξ , decays
CHAPTER 2. DISSOLUTION TRAPPING & CONVECTION IN THE BRINE 27
as 1/t.
We solve the initial-value problem given by (2.21-2.23) to show that the preferred
mode of the streamwise perturbations for k = 0 is φ1 . Figure 2.2(a) shows that the
2
concentration eigenfunction resolves into the dominant mode of L, given by ξ e−ξ , in
a very short time (t = O(1/Ra)) for any given set of white-noise initial conditions.
We show in § 2.2.3 that the time required for convergence to the first eigenfunction is
several orders of magnitude smaller than the critical time for the onset of instability.
Therefore, the IVP in ξ-t coordinates can yield accurate results for short times.
In order to highlight this phenomenon, figure 2.2(b) plots the perturbation eigen-
function in the z-t coordinates, obtained by solving the IVP, given by (2.14-2.18).
2 Ra/4t
Although the white noise initial conditions converge to ĉ ∼ z e−z , the time re-
quired for convergence is several orders of magnitude larger than in the self-similar
coordinates, as shown in figure 2.2(a). Hence, the small-time dynamics in the z-t
coordinate system are obscured during the period it takes for the random initial
perturbations to resolve into the dominant mode. This explains why earlier investi-
gations using the IVP in the z-t coordinate system did not produce accurate results.
For the ξ-t coordinate system, on the other hand, the localized eigenfunctions of L
converge rapidly to the exact solution, thereby giving an accurate growth rate of the
disturbance at small times.
Robinson (1976) used a one-term approximation to the solutions in his QSSA analysis
in z-t coordinates, and he found that the error is small when compared to expansions
using many terms. In the ξ-t coordinates such a one-term approximation is expected
to be even better, because we use an eigenfunctions expansion. We use the leading
order approximation for ĉ from (2.27), substitute it into the IVP, given by (2.23), and
integrate across the domain to obtain
dA1 A1 A1 k 2 Ra −ξ2
=− − +
e ŵ , (2.30)
dt t Ra πt
CHAPTER 2. DISSOLUTION TRAPPING & CONVECTION IN THE BRINE 28
where ∞
2
e−ξ ŵ dξ
2
e−ξ ŵ = 0 ∞ . (2.31)
2
ξ e−ξ dξ
0
Again using only the leading order estimate of ĉ to solve for ŵ, we obtain
∂2 4tk 2 4tk 2 2
− ŵ = − A1 ξ e−ξ . (2.32)
∂ξ 2 Ra Ra
where
ξ √ 1 √ k πt k2 t
t/Ra −x2 2
ξ e−2kx dx = − e−2kξ t/Ra − ξ − e erf ξ + k , (2.34)
0 2 2 Ra Ra
and
ξ √ 1 √ k πt k2 t
t/Ra −x2 2
ξ e2kx dx = − e2kξ t/Ra − ξ + e erf ξ − k . (2.35)
0 2 2 Ra Ra
dA1
= σ(t; k) A1 , (2.36)
dt
with growth rate σ. Equation 2.36 shows that the perturbations grow exponentially.
CHAPTER 2. DISSOLUTION TRAPPING & CONVECTION IN THE BRINE 29
2.5
2 t=1
2
t=tc=0.34 k=8π
0
1.5
growth rate
-2
σ t=0.05 1 k=4π
-4
0.5
-6 IVP
IVP 0 QSSA
dominant mode dominant mode
-8
-0.5
0 20 40 60 80 0 1 2 3 4
k t
Figure 2.3: (a) Growth rate vs. wavenumber curves for Ra = 500 computed by
the dominant mode method and the IVP. The flow is stable for small times. The
growth rate increases with time to become positive at the critical time tc and the
critical wavenumber kc . Flow instability increases with time beyond tc with both
shortwave and longwave cutoffs. Comparison with the IVP shows exact agreement for
small times and small wavenumbers. (b) Comparison of the dominant mode method
and the QSSA with the initial value problem for two perturbation wavenumbers at
Ra = 500. The dominant mode solution gives exact results for small times but
becomes inaccurate for later times particularly for large wavenumbers. The QSSA on
the other hand is reasonably accurate for all times.
The growth rate σ(t; k) is obtained without the QSSA and is expected to yield accu-
rate results.
It is interesting to compare the above development for the semi-infinite case with
that of the infinite case analyzed by Ben et al. (2002). The eigenfunctions of the
self-similar operator in the latter case are the full range of Hermite polynomials, The
associated eigenvalues are λn = −n/2. for n = 0, 1, 2... resulting in a neutral mode for
the zeroth eigenvalue. In a semi-infinite domain, on the other hand, the self-similar
operator has eigenfunctions based upon only those Hermite polynomials which satisfy
the boundary condition ĉ(ξ = 0, t) = 0. The associated eigenvalues are −1, −2, ... ;
therefore, a neutral mode is not present.
The presence of a zero eigenvalue in the unbounded case implies a longwave insta-
bility (σ = 0) for k = 0, such that the flow is always unstable for small wavenumbers.
For the semi-infinite domain, on the other hand, the dominant mode decays as t−1 ,
CHAPTER 2. DISSOLUTION TRAPPING & CONVECTION IN THE BRINE 30
hence σ = −1/t for k = 0, so that the flow is always stable for small wavenumbers. A
longwave cutoff therefore exists along with a critical time at which the flow becomes
unstable.
The existence of a critical time as a function of the Rayleigh number has been
noted previously (Ennis-King & Paterson, 2003; Ennis-King et al., 2005). However,
its exact value as well as the fundamental mechanism have not been given explicitly.
Consider the growth rate, given by
1 k2 k
σ(t; k) = − − + √ F (t; k) , (2.37)
t Ra π
where F (t; k) is computed numerically from (2.31). Since F (t; k) > 0 the stabilizing
effects come from the first two negative terms on the left hand side. The first term,
−1/t, which is due to the nonzero eigenvalue of the dominant mode of L, insures that
σ < 0 for very small times. In physical terms, the flow can become unstable only
when the perturbations grow at a rate faster than the decay rate of the first mode of
L.
It is important to note that selecting only the first mode to capture the pertur-
bation dynamics cannot be accurate for the entire range of length and time scales
of interest, in view of the absence of a neutral mode of L. We show in § 2.2.3 that
the dominant mode solution gives exact results in comparison with the IVP only for
small values of k t/Ra. Therefore, we use the QSSA for (2.21-2.25), to solve for
the growth rates for larger values of k t/Ra. We also show that when the QSSA is
used with the governing equations in the self-similar coordinates, accurate results are
obtained.
Results
The growth rate vs. wavenumber curves given by both the dominant mode method
and the numerical solution of the IVP are shown in figure 2.3(a) for different times.
High accuracy numerical simulations of the IVP were carried out for each wavenumber
k on a fine computational grid, by using standard methods for 1-D problems. The
CHAPTER 2. DISSOLUTION TRAPPING & CONVECTION IN THE BRINE 31
(a) (b)
101 103
critical wavenumber, kc
102
critical time, tc
100
101
-1
10
100
-2
10
101 102 103 104 101 102 103 104
Ra Ra
Figure 2.4: (a) Critical time vs. Ra. (b) Critical wavenumber vs. Ra, given by
the dominant mode solution. The critical time varies as the 1/Ra, while the critical
wavenumber scales linearly with Ra.
evolution of the maximum value of either the concentration or the velocity eigen-
function forms the basis of the growth rate plotted at a particular time for each
wavenumber. Completely stable behavior is indicated by σ < 0 for all wavenumbers
at early time. A critical time, tc , is also shown when the growth rate just becomes
positive at a critical wavenumber kc . At larger times, the stability curve displays a
maximum growth rate at a corresponding most dangerous wavenumber, and a long-
wave and a shortwave cutoff. Comparison of the dominant mode solution with the
IVP results shows exact agreement for all wavenumbers at small times when k t/Ra
is small. For longer times, the growth rate begins to deviate from the IVP result at
larger wavenumbers. However, the critical time and the longwave cutoff are computed
exactly by the dominant mode solution.
Since the dominant mode solution does not give accurate results for large values of
k t/Ra, we use the QSSA in self-similar coordinates to compute the growth rates.
Figure 2.3(b) compares the growth rate as a function of time obtained from the
IVP with the results computed from the dominant mode method and the QSSA, for
two wavenumbers. The dominant mode method again gives exact results for small
times, but deviates from the IVP solution for large times, particularly for the larger
wavenumber. The QSSA, on the other hand, gives reasonably accurate results for
CHAPTER 2. DISSOLUTION TRAPPING & CONVECTION IN THE BRINE 32
(a)
(b)
(b) (a)
(b)
1 1
10
250
150 200
10
250
200
125 200
150
100
150 150
σkmaxcutoff, kS
cutoff, kS
100
75
100 100
max
max
σkmax
50 0 500
10 10
shortwave
shortwave
50 50
Ra=500 Ra=500
25 Ra=1000 Ra=500 Ra=1000 Ra=500
Ra=500 Ra=1000 Ra=500 Ra=1000
Ra=2000
Ra=1000 Ra=2000
Ra=1000
Ra=2000 Ra=2000
Ra=2000 Ra=2000
1020
-1
-1
10-1-1
0
20
-1
1020
10
10 10010
10 0
1011
10 10-1-1-1
10
10 10000
10
10 10111
10
10
ttt ttt
Figure 2.5: (a) Maximum growth rate σmax as a function of time, and (b) The most
dangerous wavenumber kmax as a function of time, for different Rayleigh numbers.
All the results are obtained using the QSSA.
all times. This clearly shows that the QSSA in the self-similar coordinates can be
employed to obtain reliable results even for short times.
Figures 2.4(a) and (b) respectively show the critical time and the critical wavenum-
ber as a function of the Rayleigh number computed by the dominant mode method.
The critical time varies as Ra−1 , while the critical wavenumber scales linearly with
Ra. Similar scalings have been obtained by Caltagirone (1980), but our analysis elim-
inates the effect of the initial condition. We obtain tc ≈ 146/Ra and kc ≈ 0.07 Ra for
the critical time and wavenumber. These relationships apply only when 4t/Ra 1.
The maximum growth rate as a function of time and the corresponding most dan-
gerous mode for various Ra, computed by the QSSA, are shown in figures 2.5(a) and
(b) respectively. The maximum growth rate, σmax , increases rapidly at early time
beyond tc , reaches a maximum value and decays approximately as t1/4 at late times.
The most dangerous wavenumber kmax also displays approximately t1/4 scaling.
The evolution of the longwave, kl , and the shortwave, ks , cutoff is plotted in
figure 2.6. The former decays approximately as t1/5 , while the latter decays much
faster as approximately t4/5 at long times.
CHAPTER 2. DISSOLUTION TRAPPING & CONVECTION IN THE BRINE 33
(a) (b)
(a)
(b)
103 250
10
150
8
200
125
Ra=500 6
100
150
kmaxcutoff, kS
longwave cutoff, kL
Ra=1000
Ra=2000
102 4
75
100
max
50
shortwave
2
σ
50
101
1 Ra=500
25 Ra=1000 Ra=500
Ra=500 Ra=1000
Ra=2000
Ra=1000 Ra=2000
Ra=2000
100 -2 0.5
20
10 10-1 100 101 10-1-1
10
10
0 0
100
1010 10
1
t ttt
Figure 2.6: (a) The longwave cutoff wavenumber kL as a function of time for different
Rayleigh numbers, computed by the dominant mode method. kL behaves similarly
for all Ra and varies approximately linearly with time. (b) The shortwave cutoff kS
computed by the QSSA.
Nonlinear fingering dynamics govern the long term flow behavior. We solve the non-
linear problem with a high accuracy, vorticity based method proposed by Ruith &
Meiburg (2000). This method has been employed successfully to obtain highly accu-
rate results for various miscible flow problems in porous media. (Camhi et al., 2000;
Riaz & Meiburg, 2003b, 2004b). High resolution finite-difference simulations have
been reported recently by Otero et al. (2004), for high Rayleigh number convection
in a porous layer of finite depth. They compare the numerical results with analytical
heat flow estimates. We compare our numerical simulations with the results from the
linear stability analysis.
The governing equations used for direct numerical simulations are (2.10) and
(2.11). We use a vorticity formulation for (2.10) to eliminate pressure. By taking
the curl of (2.10) we obtain,
∂C
ω=− = −∇2 ψ , (2.38)
∂x
CHAPTER 2. DISSOLUTION TRAPPING & CONVECTION IN THE BRINE 34
∂ψ
w = , (2.39)
∂x
∂ψ
u = − . (2.40)
∂z
∂C
C(z = 0, x, t) = 1 , (z = 1, x, t) = 0 , (2.41)
∂z
∂C ∂C
(z, x = 0, t) = 0 , (z, x = A, t) = 0 , (2.42)
∂x ∂x
w(z = 0, x, t) = 0 , w(z = 1, x, t) = 0 , (2.43)
u(z, x = 0, t) = 0 , u(z, x = A, t) = 0 . (2.44)
The aspect ratio A = L/H, where L is the lateral extent of the computational domain
and H is the layer thickness (figure 2.1). The streamfunction ψ = 0 on all boundaries
while ω = 0 at x = 0 and x = A. Boundary conditions for vorticity at z = 0, 1 are
obtained from (2.38).
We solve the Poisson equation (2.38) by expanding ω and ψ in Fourier modes in
the x-direction. We then solve the resulting ODE for the decoupled z-direction eigen-
functions with 6th order compact finite differences (for details see Riaz & Meiburg,
2003b). The velocities are then computed from (2.39) and (2.40), where the deriva-
tives of ψ are evaluated with 6th order compact finite differences. Time integration
of (2.11) is carried out using a standard 4th order Runga-Kutta method, where all
the spatial derivatives are again evaluated with 6th order compact finite differences.
The resulting numerical scheme resolves all relevant length and time scales accurately.
The initial condition for the concentration is given by (2.18) with a starting time of
t = 0.2. The initial condition for velocity is w = u = 0.
CHAPTER 2. DISSOLUTION TRAPPING & CONVECTION IN THE BRINE 35
2.5
k=8π
2
1.5
growth rate
1 k=4π
0.5
-0.5
0 1 2 3 4
t
Figure 2.7: Comparison of the growth rates obtained from nonlinear direct numerical
simulation and the linear initial value problem, for two different initial perturbations
with wavenumber, k.
The explicit nature of time integration imposes a strict limit on the time steps.
Although stable time steps are given by the CFL condition, we use even smaller time
steps, on the order of 10−5 , to insure accuracy. Spatial resolution of the computational
grid ranges from 512 × 512 grid points for small Rayleigh number cases to 2048 × 2048
grid points for larger Rayleigh number cases. These fine spatial and temporal reso-
lutions produce converged results. Appropriate grid spacing for different parameter
combinations is obtained by consideration of the cutoff mode provided by the linear
stability analysis. The grid spacing is chosen such that it is smaller than the cutoff
wavelength. Additionally, the divergence of the velocity field is checked throughout
to ensure exact mass conservation for a given grid spacing.
Validation of the numerical simulations was performed by comparing the growth
rates with those obtained from the linear stability analysis. The numerical simulations
are perturbed with pure sinusoidal modes in the transverse direction, superimposed
on the initial concentration profile. The growth rate is measured using the norm of
vorticity, defined as
CHAPTER 2. DISSOLUTION TRAPPING & CONVECTION IN THE BRINE 36
1 A
ω(t) = ω(z, x, t) dx dz . (2.45)
0 0
Figure 2.8: Concentration contours at different times for Ra = 4000. (a) t = 1, (b)
t = 1.8, (c) t = 2.3 and (d ) t = 3.8 A large number of fingers consistent with the linear
stability analysis develop initially. Nonlinear interactions rapidly reduce the number
of fingers and give rise to large scale structures at later times. The fingers at later
times are connected to the top boundary at discrete locations. These connections act
as feeding sites of the high density fluid to the convecting fingers below.
CHAPTER 2. DISSOLUTION TRAPPING & CONVECTION IN THE BRINE 38
Figure 2.9: Vorticity contours in the background with overlapping streamlines for
Ra = 4000, t = 2.3. Corresponding concentration contours are shown in figure 2.8(c).
The vorticity field has a dipole structure that drives the high density fluid through
the fingers. The stream lines show how the fluid travels laterally and then descends
through the isolated feeding sites for the fingers. The direction of fluid circulation is
shown by the arrows.
Tchelepi & Orr , 1994; Manickam & Homsy, 1995). The driving force for instability,
i.e. the density gradient, is weakened progressively as the fingers move away from
the top boundary. In addition to the diffusive spreading, nonlinear finger interactions
tend to further smooth out the concentration gradients of a large number of competing
fingers. At a later time of t = 2.3 shown in figure 2.8(c), some of the smaller fingers
disappear due to diffusive smearing, while others merge to form large-scale structures
that develop relatively independently. This trend continues for later times, as shown
in figure 2.8(d ) for t = 3.8. It is interesting to note that even at late times, large-
scale fingers are connected to the diffusive boundary layer at discrete narrow locations,
which serve as feeding sites of high density fluid.
Fingering dynamics are revealed by analysis of the vorticity profile that generates
the rotational flow responsible for driving the fingers. Figure 2.9 plots the vorticity
field contours for Ra = 4000 at t = 2.3. The corresponding concentration field
is shown in Fig, 2.8(c). Superimposed on the vorticity field in figure 2.9 are the
streamlines showing the negative (dashed lines) and positive (solid lines) circulation
paths. The vorticity field displays a dipole structure of negative (light) and positive
(dark) vorticity pairs. Very high vorticity pairs are concentrated at the root of the
fingers, which act as feeding sites for the finger portions away from the top boundary.
The corresponding streamlines show how the fluid is drawn from the sides to flow
CHAPTER 2. DISSOLUTION TRAPPING & CONVECTION IN THE BRINE 39
laterally along the top boundary layer and down through the high-vorticity regions
at the finger roots. Although the highest vorticity magnitude occurs at the finger
roots, the tips of the fingers also display a moderate accumulation of vorticity.
As noted above in § 2.2.4, the early time wavelength developed by the nonlinear
simulations is in close agreement with that predicted by the linear stability analysis.
In order to compare the time evolution of the preferred mode of the nonlinear flow,
we define a dominant mode as
K
kE(k) dk
0
n̂ = K
, (2.47)
E(k)dk
0
where n̂ is the dominant wavenumber, k is the Fourier mode and E(k) is the energy
spectrum associated with the Fourier transform of the vorticity field. We compute
E(k) as
x 1
E(k; t) = ω(z, x, t)dz e−ikx dx . (2.48)
0 0
40
〈
30 DNS (n)
linear stability (kmax)
20
wavenumber
10
Ra=500
1000
2 2000
4000
0.4 2 4 6 8 10
t
Figure 2.10: The dominant mode of the nonlinear simulations as a function of time for
different values of the Rayleigh number. Also shown are the most dangerous modes
given by the linear stability analysis. Good agreement is observed between the two
for early times. At later times the onset of nonlinear behavior leads to a significant
deviation from linear results.
Figure 2.10 shows that the nonlinear behavior is strongly dependent on the Rayleigh
number. In order to estimate the influence of Ra on the flow dynamics, we com-
pare the relatively early-time behavior at two Rayleigh numbers. Figure 2.11 plots
the concentration contours for Ra = 1000 at t = 2.2 and Ra = 8000 at t = 1.6.
These simulations employ 512 × 512 and 2048 × 2048 grid points, respectively. The
concentration front moves much faster for the latter case. The Ra = 1000 case de-
velops a few large fingers and also shows a substantial amount of diffusive smearing
in other regions of the concentration front. The large-scale fingers appear to move
independently, without interacting with neighboring fingers. The Ra = 8000 case,
on the other hand, develops a vigorous instability that results in complex fingering
structures. Compared to the Ra = 4000 case shown in figure 2.8(b) at t = 0.18, the
Ra = 8000 case already shows well developed discrete feeding sites. Note that new
feeding sites develop as the old ones are abandoned. In other regions, multiple fingers
CHAPTER 2. DISSOLUTION TRAPPING & CONVECTION IN THE BRINE 41
Figure 2.11: Concentration contours for (a) Ra = 1000, t = 2.2 and (b) Ra = 8000,
t = 1.6. The larger Rayleigh number case is marked by intense nonlinear interaction
between competing fingers. The small Ra case shows larger fingers with a strong
diffusive spreading.
can attach to a single feeding site, nonlinear interactions then select finger one over
the others as the preferential flow path.
The complexity of the fingering structures suggest that the long term fate of the
nonlinear competition cannot be predicted from the concentration profiles at earlier
times, especially for large Rayleigh numbers. In order to analyze the fingering dynam-
ics at long times, we plot the concentration profiles in figure 2.12 for various Rayleigh
numbers at times when the concentration front reaches the bottom boundary. We
designate this time as tb . The small Rayleigh case, Ra = 1000 at time tb = 8.9,
shows that two large fingers survive to reach the bottom boundary. Comparison with
figure 2.11(a) shows that the fingering configuration is completely different from that
at an earlier time of t = 2.2. The Ra = 2000 case at tb = 7 displays slightly narrower
fingers, with one isolated finger in the middle making it to the bottom boundary,
while two other fingers, attached to a single feeding site, are still in competition. For
larger Rayleigh number cases, Ra = 4000, tb = 8 and Ra = 8000, tb = 8.5, more
fingers reach the bottom boundary. Many of these fingers can be observed to undergo
strong interactions while others are in the process of fading out. Figures 2.12(c) and
CHAPTER 2. DISSOLUTION TRAPPING & CONVECTION IN THE BRINE 42
Figure 2.12: Concentration contours at time tb , when the concentration front reaches
the bottom. (a) Ra = 1000, tb = 8.9. (b) Ra = 2000, tb = 7. (c) Ra = 4000, tb = 8.0.
(d ) Ra = 8000, tb = 8.5. The late time dynamics is governed by large scale fingers.
Large Rayleigh number cases continue to display vigorous fingering interactions.
CHAPTER 2. DISSOLUTION TRAPPING & CONVECTION IN THE BRINE 43
0
10
tip position
10-1
Ra=500
Ra=2000
Ra=8000
10-2 -1
10 100 101
t
Figure 2.13: The position of the most advanced portion of the front as a function
of time for different values of the Rayleigh number. The tip moves faster for larger
Rayleigh numbers for early times. Fingering interactions significantly influence the
rate of front propagation at later times, particularly for large Ra cases.
2.12(d) clearly show the fading of the fingers is due to shifts in the feeding sites as well
as due to lateral pinch-off, where one diagonally moving finger cuts the fluid supply
of a neighboring finger. An interesting consequence of these fingering interactions is
an increase in tb for larger values of the Rayleigh number, as compared to the small
Ra cases. Note that tb = 8.9 for Ra = 1000, it decreases to 7 for Ra = 2000 but then
again increases to 8 for Ra = 4000 and then to 8.5 for Ra = 8000.
The position of the most advanced section of the concentration front, the tip
position, as a function of time is shown in figure 2.13 for various Rayleigh numbers.
For the small Ra = 500 case, the front initially propagates as t1/2 and then switches to
a linear growth for larger times. Higher Rayleigh number cases display the diffusive
t1/2 behavior for relatively shorter times. The Ra = 2000 case displays a faster
than linear growth of the tip position at later times. The tip travels quickly for the
Ra = 8000 case, but for a very small initial time. It is clear that by t = 5 the tip is
moving faster for the Ra = 2000 case as compared the Ra = 8000 case. Consequently
the time for a finger to reach the bottom boundary is shorter for the Ra = 2000 case.
CHAPTER 2. DISSOLUTION TRAPPING & CONVECTION IN THE BRINE 44
As shown by the concentration contours in figure 2.12 this behavior is due to more
intense finger interaction at higher Rayleigh number cases, such that the interacting
fingers do not allow any one finger to clearly breakaway ahead of the front. For the
smaller Ra cases, on the other hand, isolated fingers travel faster due to the absence
of interference from neighboring fingers.
2.2.5 Discussion
We analyze the stability of a diffusive boundary layer in a semi-infinite domain. This
analysis is applicable when the penetration depth of the diffusive boundary layer, δ,
is small relative to the domain thickness, H. The penetration depth at the onset
of instability is given in dimensional form as δc ≈ 24µD/(K∆ρg). We must have
δc H for the assumption of the semi-infinite domain to be valid. The dimensional
critical time, which is then independent of the length scale, is given by
φµ2 D
tc = 146 . (2.49)
(K∆ρg)2
2πµD
λc = . (2.50)
0.07K∆ρg
To validate both the linear analysis and the numerical simulations, the predictions
of tc and λc should be compared with experimental observations. Unfortunately
Elder (1968) only states the Rayleigh number of his experiments and does not give
all the data necessary to calculate tc and λc . Green & Foster (1975) have reported
experiments of a salt solution diffusing into the top of a Hele Shaw cell. The Rayleigh
number in their experiment is Ra = KρgH/(µD) = 90500. They do not report the
onset time, but they give the wavelength of the first observed fingers as λ = 1.8 mm
and note that this may be an over estimate. For the same parameters, our linear
stability analysis predicts a critical time of tc = 4 s and a critical wavelength of
λc = 0.6 mm. The critical wavelength predicted by linear theory is generally smaller
than the wavelength first observed, because the fingers are initially not visible in
CHAPTER 2. DISSOLUTION TRAPPING & CONVECTION IN THE BRINE 45
103 tc
0
λc
1.
tc (yrs), λc (m), δc (m)
2
10 102
≈
pe
slo
depth (m)
101
δc
0 .5
10 ≈0
101 pe
slo
10-1
10-2
1 2 3 4 100 1 2 3 4
10 10 10 10 10 10 10 10
k (mD) time (yrs)
Figure 2.14: (a) Variation with permeability of the critical time, tc , the critical wave-
length, λc and the penetration depth at the critical time, δc , for ∆ρ = 5 kg/m3 ,
φ = 0.3, µ = 0.5 cP and D = 10−9 m2 /s. For this range of permeability variation
tc varies between 2000 yrs and 10 days while λc goes from about 100 m to less than
a meter. The penetration depth δc gives information regarding the applicability of
our analysis with respect to the layer thickness H, such that δc H. (b) The ad-
vance of the fastest finger tip is shown until the bottom of the simulation domain
is reached. A diffusive, t0.5 , behavior is observed at early times, while the fingers
advance proportional to t at later times.
the experiments. By the time they become observable, they have already coarsened.
Given this experimental limitation on the detection of the critical parameters, high
resolution numerical studies as reported above are valuable because the growth of
perturbations can be detected before they become visible.
The critical time and the critical wavelength can vary by orders of magnitude
depending on the properties of the geological formation. While viscosity, diffusion and
density difference have more or less similar values for typical aquifers, permeability
can vary over a large range of values, and therefore introduces the largest variations
in tc and λc . Figure 2.14(a) shows that δc decreases from 55 m to about 0.07 m when
the permeability increases from 1 mD to 3 D. Hence our results for the semi-infinite
domain will apply to high permeability layers with a thickness of few tens of meters,
but for low permeability formations the thickness has to be several hundred meters
for our results to be applicable. The critical wavelength decreases from λc ≈ 200 m
to λc ≈ 0.3 m and the critical time for the onset of the instability decreases from
CHAPTER 2. DISSOLUTION TRAPPING & CONVECTION IN THE BRINE 46
2.2.6 Conclusions
Accurate numerical simulation of density fingering over long times at the field scale
is one of the main challenges in predicting the movement of CO2 underground. Our
theoretical and numerical results are valuable for understanding density driven con-
vection during CO2 storage in saline aquifers. Our analysis is directly applicable to
the thermal instability problem and resolves the issue of the critical time and the
critical wavenumber. We also highlight the physical mechanisms of instability that
give rise to both the long wavenumber cutoff and the critical time for the onset, in a
semi-infinite domain.
The disturbances only become experimentally observable a finite period after the
critical time, and it is therefore difficult to measure the critical time with reasonable
accuracy. For typical aquifers with moderate permeability, the onset time can be as
large as hundreds of years. The prediction of the critical wavelength is crucial in
choosing the grid resolution in the numerical simulation. Although the large scale
CHAPTER 2. DISSOLUTION TRAPPING & CONVECTION IN THE BRINE 48
fingers at later times can be resolved with fewer grid blocks, the errors introduced by
ignoring the small-scale dynamics at early times, may have a significant influence on
the late time behavior.
The simplifications necessary to treat the problem theoretically may introduce
significant errors when such results are applied to real aquifers. The important as-
sumptions are the homogeneity and isotropy of the porous medium, as well as the
assumption of single-phase flow and the absence of velocity induced dispersion. The
assumption of single-phase flow will break down if capillary forces are significant.
Additionally, physical mechanisms related to dissolution, precipitation and geochem-
ical reactions, which are not accounted for in our analysis, can be expected to play
a role. Some of these processes may be incorporated in the linear stability analysis,
while others can only be investigated with high-resolution nonlinear simulation. The
results for the basic gravitational fingering instability presented here, form the basis
to investigate the effect and importance of these additional processes.
0.9
0.8
1 1 1 1
0.7
0.6
z 2 2 2 2 0.5
0.4
0.3
3 3 3 3
0.2
0.1
4 4 4 4 0.0
0 1 0 1 0 1 0 1
x x x x
scales (2.5) to (2.8), and the concentration of CO2 in the brine, C = c/ceq , is normal-
ized by the equilibrium concentration at the CO2 -brine interface, ceq . The concentra-
tion is given by c = xc ρc , where xc is the mass-fraction of CO2 in the brine and ρc is
the density of the dissolved CO2 . As before, the governing parameter is the Rayleigh
number, Ra = k∆ρgl/φµD, where l is a suitable length scale specified below.
10-2
3 2.5
dC dC 1.5
dτ 2 dτ
1
0.5
1 0
1 10 20 30 40 50 0 1000 2000 3000 4000
τ Ra
Figure 2.16: (a) Evolution of the dissolution rate at Ra = 4000. Average dissolution
rate over time is shown as dashed line. (b) Time averaged of the dissolution rate for
different Ra.
of dissolved CO2 have not reached the bottom of the domain. Figure 2.15 shows the
evolution of gravity driven fingers, or plumes, of dense CO2 -rich brine in a typical
simulation. The nonlinear interactions of the fingers have already been described in
§ 2.2.4-2.2.4, and here we focus on the evolution of the dissolution rate, dC/dτ . We
define the dissolution rate as the rate of change of the total dissolved CO2 in the
domain, given by
L H
dC d
≡ Cdx dz. (2.51)
dτ dτ 0 0
Figure 2.16(a) shows the evolution of dC/dτ , for a typical simulation at Ra = 4000.
This evolution can be divided into three stages. For τ < 1 the diffusive boundary
layer is stable and the dissolution rate decays as, dC/dτ ∝ τ −1/2 . After the onset of
convection the dissolution rate fluctuates, and the peaks correspond to the growth of
large fingers. After an initial transient, 1 < τ < 10, the dissolution rate fluctuates
around a constant value of dC/dτ ≈ 0.017. We refer to this average value as the
open-system dissolution rate,
dC/dτ . Figure 2.16(b) shows that
dC/dτ is nearly
independent of Ra, because Ra can be scaled out of the problem in a domain of
infinite depth, see also the discussion in § 2.2.3.
This open-system dissolution rate allows simple estimates of the time necessary
CHAPTER 2. DISSOLUTION TRAPPING & CONVECTION IN THE BRINE 51
to dissolve a given mass of CO2 injected into a saline aquifer based on its average
petrophysical properties. The mass dissolution rate of CO2 per unit width in the
open convection regime is given by
dmc dC k∆ρgceq
≈ , (2.52)
dt dτ φµ
where mc is the total mass of dissolved CO2 . The onset time, τo , of the convective
mass transport is given by the first minimum in the evolution of the dissolution rate
(figure 2.16a, see also figure 2.18a). For the range of Ra investigated (see figure 2.18b),
it is given by
φµ11/5 D6/5
to = 6215 , (2.53)
(k∆ρg)11/5 H 1/5
as computed from DNS. This onset time, to , decreases with Ra and it is larger than
the theoretically predicted critical time, given by (2.49), obtained from linear theory
in § 2.2.3.
(a) ¿ = 7.4 C
0 1.0
0.8
z 0.6
0.4
0.2
1 0.0
0 1 2 3 4
(b) ¿ = 22.1 C 1.0
0
0.8
z 0.6
0.4
0.2
1 0.0
0 1 2 3 4
(c) ¿ = 36.7 C
0 1.0
0.8
0.6
z
0.4
0.2
1 0.0
0 1 2 3 4
(d) ¿ = 124.7 C
0 1.0
0.8
z 0.6
0.4
0.2
1 0.0
0 1 2 3 4
x
Figure 2.17: Evolution of the concentration in a closed aquifer, for Ra = 1000. (a)
Plumes of dissolved have not reached the base of the aquifer. Convection similar to
an open aquifer. (b-d ) Convection in a closed aquifer.
CHAPTER 2. DISSOLUTION TRAPPING & CONVECTION IN THE BRINE 53
10-1 102
τi (a) (b)
101
10-2
dC τ0 τ
dτ 100
DNS:
10-3 τc
10 -1
τ0
τi
τc (theory)
10-4 10-2
10-1 100 101 102 102 103 104
τ Ra
Figure 2.18: (a) Evolution of the dissolution rate, dC/dτ , for various Ra in a closed
aquifer. (b) Variation of the time scales discussed in the text as a function of Ra.
1. The critical time, tc ∝ k −2 , is the time after which small perturbations grow, it
gives an analytic lower bound for the onset of convection.
2. The onset time to ∝ k −2.2 , marks the transition from diffusive to convective
mass transfer, and corresponds to the appearance of the first fingers in the
concentration profile. For large Ra, we expect to to approach tc from above.
The first two time scales are independent of the thickness of the aquifer, because
they are determined by the growth of the boundary layer, which we assume to be
much smaller than H. The interaction time, ti , is of course directly proportional to
H. These time scales identify three dynamic regimes of mass transport in the brine,
shown in figure 2.18(b). These three regimes may be characterized as:
1. In the diffusive regime (t < to ) the concentration of dissolved CO2 in the aquifer
increases proportional to t1/2 , and the dissolution rate of CO2 decreases rapidly
as t1/2 . In the absence of the gravitational instability this would be the only
regime.
2. In the open-system regime (to < t < ti ) the dynamics are independent of the
domain size. In this regime the total concentration increases linearly with t,
and the open-system dissolution rate is given by
dC/dτ ≈ 0.017. This is the
most effective regime of mass transfer (gray triangle in figure 2.18b), and is
characterized by rapidly growing fingers with strong nonlinear interactions.
3. The closed-system regime (t > ti ), where the finite size of the aquifer influences
the dynamics of convection. This regime is characterized by large, slow convec-
tive cells. The dissolution rate decays rapidly and dissolved CO2 approaches its
equilibrium concentration throughout the aquifer.
of the aquifer. In general, the dissolution of CO2 into the brine will be greatly
enhanced by density driven convection in large high-permeability aquifers, because
the onset time is short, the dissolution rate in the open-system regime is high, and
the transition to finite-acting convection is late.
3.1 Introduction
The formation of residual saturations is the second primary trapping process that
reduces the volume of mobile CO2 , and therefore determines the duration of the active
storage period (figure 1.3). When the saturation of the non-wetting CO2 decreases,
a process referred to as imbibition, an interplay of viscous and capillary forces at
the pore-scale leads to the formation of disconnected essentially-immobile bubbles
of residual CO2 (3.1a). Imbibition takes place in the wake of the migrating CO2
plume, and residual trapping continues as long as the CO2 migrates. To understand
residual trapping it is therefore essential to understand the buoyancy driven migration
of the CO2 plume. In the two aquifers shown in figure 1.4(b) and (c) the CO2
plume migrates as a gravity current and residual trapping is likely to be an important
trapping process. In § 3.1.1 we give a brief introduction to the pore-scale phenomenon,
and we introduce a nomenclature suitable for the simple model we adopt. A discussion
of the current research on the formation of residual saturations is beyond the scope
of this thesis. Instead we adopt the simplest trapping model, and we focus on the
interplay between sweep and residual trapping, which has so far not received much
attention in CO2 storage.
56
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 57
After the end of injection the advective forces are small compared to gravitational
and capillary forces, and we study the migration of the CO2 using a simplified model
assuming gravity-capillary equilibrium in the vertical direction and a sharp interface
between the CO2 plume and the ambient brine. Vertical-equilibrium sharp-interface
models are commonly used in hydrology, petroleum engineering, and applied math-
ematics, because they allow analytic and semi-analytic solutions that give insight
into the underlying dynamics. CO2 storage in saline aquifers raises new questions
that have not been addressed by the literature on vertical-equilibrium sharp-interface
models, reviewed in § 3.1.2. We focus on the effects of confinement, residual trap-
ping and slope on the migration of CO2 plumes in saline aquifers. In § 3.2 we derive
the vertical-equilibrium sharp-interface equations for a confined sloping aquifer with
constant residual saturations.
In § 3.3 we study the effect of confinement on the spreading of the CO2 plume,
in absence of residual trapping. We use the scaling laws, associated with similarity
solutions for confined and unconfined aquifers, to detect the influence of confinement
in more general numerical solutions. Using this novel approach, we are able to show
that confinement has a strong effect on the migration of the CO2 plume, due to the
less mobile ambient brine. In other words the mobility ratio between the CO2 and
the brine is an important parameter governing the migration of the CO2 .
In § 3.4 we turn our attention to the relation between the sweep of the gravity
current and residual trapping. We extend the work on horizontal confined aquifers,
presented in § 3.3, to include residual trapping, and show that the solutions remain
self-similar. This implies that the volume of the current decreases as a power-law
in horizontal aquifers. Using the horizontal aquifer as a base case we show that
residual trapping is much more effective in sloping aquifers. The hyperbolic limit of
the equations allows us to study analytically how far the CO2 migrates up-dip and
how long it remains mobile.
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 58
Figure 3.1: (a) Microtomogram of non-wetting CO2 bubbles (black) and water (white)
in the pore space between grains (grey), from Benson et al. (2006). (b) Comparison
between non-wetting phase saturation profile after imbibition, inferred from a wire-
line log (Koperna & Kuuskraa, 2006), and the simplified step function saturation
profile corresponding to the sharp-interface approximation. (c) The geometry of the
porous layer and the variables used in the derivation are shown.
(Bear, 1972; Lake, 1989). Coats et al. (1971) used a vertical-equilibrium model, which
accounts for presence of a capillary transition zone, to formulate efficient quasi-three-
dimensional reservoir simulators. Parker & Lenhard (1989) introduced a similar model
for the migration of non-aqueous phases along the groundwater table.
Nordbotten et al. (2005), Lyle et al. (2005), Vella & Huppert (2006), and Bickle
et al. (2007) used sharp-interface models to study the development of CO2 plumes
during the injection period in confined and unconfined aquifers. During injection
the CO2 saturations increase, and no residual CO2 forms. After several decades the
injection of CO2 ends, and the CO2 plume migrates as a gravity current for several
hundreds or thousands of years, so that the long term evolution of the CO2 plume
may be modeled as an instantaneous release of finite volume. In the post-injection
period the CO2 saturation decreases in the wake of the gravity current, and residual
trapping reduces the current volume (figure 3.10). Kochina et al. (1983) extended
the sharp-interface model to allow for a constant residual saturation, and studied the
decay of the plume volume after a finite release. We extend their work to confined
sloping aquifers and consider the limiting cases.
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 61
3.2.1 Derivation
We consider the flow of supercritical CO2 with density, ρc = ρ, and viscosity, µc , and
of brine with density, ρb = ρ + ∆ρ, and viscosity, µb , in a sloping porous layer of
constant thickness, H, with dip angle, θ ≥ 0, and infinite lateral extent (figure 3.1c).
Both fluids are considered incompressible, and we assume that the porous medium
is homogeneous and isotropic with permeability, k, and porosity, φ, and that the
top and bottom boundaries are impermeable. We also assume that gravity–capillary
equilibrium is maintained in any vertical cross section of the current (figure 3.1b), and
we replace the transition zone by sharp-interface and assume that the saturations
above and below the interface are constant. We denote the thickness of the CO2
plume by hc = h so that the depth of the brine is given by hb = H − h. In this case
the pressure in the aquifer is given by
pI − gρ(z − hb ) + Pc , for z > hb ,
p= (3.1)
pI − g(ρ + ∆ρ)(z − hb ), for z ≤ hb ,
where pI is the unknown pressure at the interface, Pc is the constant capillary pressure,
and g is the gravitational acceleration. The volume flux per unit width, qp , of phase
p ∈ {c, b} is given by the multiphase extension of Darcy’s law qp = −kλp ∂φp /∂x,
where φp = p − gρp (x sin θ + z cos θ) is the potential of phase p, and λp = krp /µp is the
mobility of phase p. The flow rate per unit width of phase p is given by Qp = hp qp .
In the absence of a source term, and with the assumption of incompressibility the
global conservation of volume is given by Qc + Qb = 0. Using this constraint we can
eliminate ∂pI /∂x from the expressions for the flow rates, and we obtain
hλc (H − h)λb ∂h
Qc = −Qb = kg∆ρ sin θ − cos θ .
hλc + (H − h)λb ∂x
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 62
To obtain an equation for the evolution of the interface, we consider the conser-
vation of the CO2 volume Vc over region ∆x and time ∆t as shown in figure 3.1(c).
∆Vc = φ (1 − Sbr ) ∆h∆x = Qc |x − Qc |x+∆x ∆t + Rc .
The source term, Rc , accounts for the volume of CO2 that is lost as residual saturation
Scr in the wake of the plume. Following Kochina et al. (1983) we assume a constant
residual saturation, Scr , so that
−∆h∆xScr , for ∂h/∂t < 0,
Rc =
0, for ∂h/∂t ≥ 0.
Taking limits for small ∆x and ∆t the equation for the evolution of the interface is
given by
∂h ∂ h(H − h) ∂h
=κ − sin θ + cos θ , (3.2)
∂t ∂x h(M − 1) + H ∂x
where we have introduced two parameters. The conductivity, κ, of the CO2 is given
by
kλc ∆ρg
κ1 = φ(1−Sbr −Scr )
, for ∂h/∂t < 0,
κ= kλc ∆ρg
(3.3)
κ0 = φ(1−Sbr )
, for ∂h/∂t > 0,
where κ1 ≥ κ0 , and the mobility ratio that is given by M = λc /λb . The mobility ratio
M measures the change in the mobilities across the advancing part of the interface,
because we neglect the effect of the residual CO2 on the mobility of the brine. For
geological CO2 storage the ambient brine is less mobile, M > 1 (Adams & Bachu,
2002), and we restrict the discussion to this range. For Scr = 0 the coefficient 3.3
becomes continuous and equation 3.2 reduces to the form given by Bear (1972).
Note that the parameters are not independent; λc occurs in both. This choice of
parameters allows a simple reduction of (3.2) in the limits max(h) H or M 1,
as discussed in § 3.3.2. In these limits the equation loses its dependence on M, but
retains the parameter κ. To allow this reduction κ must be defined in terms of the
mobility of the CO2 , and must be independent of the mobility of the ambient brine.
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 63
Ld xo,0 x
0
plane of symmetry
CO2 brine
h0 h0
?
H xi;0
h Lf
Figure 3.2: The geometry of the initial condition and the three associated length
scales H, Ld , and Lf . A particular initial condition h0 [x] and the corresponding
idealized step function initial condition h∗0 [x] are shown. The arrows indicate that
the fluid has been injected over the entire thickness of the reservoir.
This transformation can be used to obtain a similar expression for the evolution of
the thickness of the ambient brine hb from (3.2).
The thickness of the zone saturated with residual saturation is given by hr [x, t1 ] =
hmax [x, t1 ] − h[x, t1 ], where hmax is the largest thickness the CO2 plume has achieved
at a particular location up to the time of consideration
We also refer to hr as the residual surface that identifies the fraction of the aquifer
swept by the CO2 plume. In the absence of residual trapping this surface has no
significance.
We consider the evolution of the CO2 plume after injection has stopped (t ≥ t0 ).
Figure 3.2 illustrates the plume shape at the end of the CO2 injection, which serves
as the initial condition for the post-injection period under consideration here. We
assume that CO2 has been injected along the whole depth of the aquifer, and that
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 64
it has formed a gravity tongue along the upper boundary of the porous layer (Riaz
& Tchelepi, 2006). Near the injection site the CO2 has displaced the water over an
average distance of Ld . The initial saturations in the CO2 plume and in the brine are
Sc = 1 − Sbr and Sc = 0, respectively. This corresponds to a porous medium that is
invaded by the released fluid for the first time, which is appropriate for CO2 storage
in saline aquifers. The lateral extent of the fluid invasion is determined by the viscous
to gravity ratio Rvg = (u/Ld )/(ug /H) = (uµg H)/(k∆ρgLd ), where u is the average
horizontal flow velocity, and ug = k∆ρg/µ is a gravitational velocity (for detailed
discussion see Tchelepi, 1994). When Rvg is small, gravitational forces dominate the
flow, and a thin gravity tongue forms at the top of the aquifer. When Rvg is large
the interface advances over the entire thickness of the aquifer. Ld increases with
time during the injection period, and Rvg generally decreases over time. During CO2
storage large quantities of fluid are injected, and the horizontal velocity, u, is high. As
a result Rvg is initially large, and the interface advances over across the full thickness
of the aquifer. Over time Rvg decreases, and a gravity tongue will form, leading to
the initial condition shown in figure 3.2.
The CO2 -brine interface transitions from h0 [xi,0 ] = H to h0 [xo,0 ] = 0 over a
frontal region of width Lf = xo,0 − xi,0 . The volume of CO2 is given by the integral
over the initial distribution
xo,0
V = 2φ(1 − Sbr ) H − h0 dx. (3.5)
0
The length scale Ld is chosen so that an idealized step function initial profile located
at x = Ld has the same CO2 volume as the particular initial condition (V = 2φLd H).
The idealized initial condition is
H, for |x| ≤ Ld ,
h∗0 =
0, for |x| > Ld .
This initial configuration imposes three length scales: the layer height H, the average
displacement distance Ld , and the width of the front at the end of injection Lf
(figure 3.2). The two boundary conditions for (3.2) require that h[x, t] → 0 for
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 65
|x| → ∞.
so that ϑ = t/td and τ = t/ta . Substituting these definitions into equation 3.2 we
obtain two dimensionless equations
depending on the choice of the characteristic time scale. The flux function is defined
as
η(1 − η)
f= , (3.8)
η(M − 1) + 1
and the dimensionless discontinuous coefficient is given by
1, ητ < 0, or ηϑ < 0,
σ= (3.9)
1 − , ητ > 0, or ηϑ > 0,
depending on the choice of the characteristic time. Equations 3.7a and b have the
same dimensionless governing parameters
krc µb Scr
M= , Pe = A0 tan θ, = .
µc krb 1 − Sbr
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 66
3.2.3 Discretisation
High accuracy numerical solutions of equation 3.7 were obtained as follows. The spa-
tial domain was divided into N grid cells of width ∆ξ centered at ξi = i − 12 ∆ξ,
where i ∈ [1, N ]. The temporal domain has been divided into T constant time steps
of size, ∆τ , so that the solution is obtained at times τn = n∆τ , for n ∈ [1 T ]. The nu-
ξi+1/2
merical approximation of the cell average in the i-th cell is given by ξi−1/2 η(ξ, τn )dx =
ηin +O(∆ξ 2 ). The right hand side of (3.7) was discretised in divergence form to ensure
discrete conservation (Leveque, 2002), and central differences were used for all spa-
tial derivatives. The time derivative was discretised using the explicit forward Euler
method with a constant time step ∆τ . The update formula is given by
∆τ n
ηin+1 = ηin − Fi+1/2 − Fi−1/2
n
. (3.10)
∆ξ
n
The numerical flux function Fi+1/2 = F ηin , ηi+1
n
is given by
n
ηi+1/2 1− n
ηi+1/2
n
ηi+1 − ηin
F ηin , ηi+1
n
=− , (3.11)
n
ηi+1/2 (M − 1) + 1 ∆ξ
n
n
where ηi+1/2 = ηi+1 + ηin /2. The numerical results were validated against the early
similarity solution derived in § 3.3.1 (see figure 3.6a).
ratio. The current is initially given by two tilting interfaces and later by a spreading
mound. The tilting interfaces correspond to self-similar solutions of (3.2) for a con-
fined aquifer that predict that the tips of the interface propagate as x ∝ t1/2 . At late
time the governing equation simplifies to the porous medium equation for an uncon-
fined aquifer. The spreading mound is a similarity solution to the porous-medium
equation and predicts tip propagation as x ∝ t1/3 . We use these scaling laws, which
are predicted by the similarity solutions, to study when this transition occurs and
how the transition from confined to unconfined flow depends on the mobility ratio.
The early evolution of the interface is independent of the front separation 2Ld , but
the duration of this early period depends on Ld .
Dimensional analysis
We follow the general procedure for dimensional analysis given by Barenblatt (1996).
The problem defined above has three dimensions: length L, height H ∗ , and time T .
The dimensions of the variables and parameters appearing in (3.2) and the initial
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 68
plane of
symmetry
(a) xo xo∝ t1/2
(b) xo xo
(c) xo xo
(d) xo xo ∝ t1/3
Figure 3.3: The evolution of buoyant CO2 -vapor (gray) released into a horizontal
porous layer saturated by brine (white). The outward propagating tip of the interface
is marked by xo . All figures are exaggerated in the vertical direction, to make the late
solution (d ) visible. In many situations of interest the width of the invaded region in
figure (a) is several times larger than the aquifer thickness.
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 69
1.0
0.8
0.6
M=1
0.4
M = 10
M = 102
M = 103
0.2
M = 104
0.0
-1.0 -0.5 0.0 0.5 1.0
Figure 3.4: Similarity solution is shown for different values of M. The variable θ is
the non-dimensional thickness of the CO2 , it is measured from the top of the aquifer.
condition are
−1
[ĥ] = [H] = H ∗ , [x̂] = [Lf ] = L, [t] = T, [κ] = L2 H ∗ T −1 , [M] = 1.
The dimensions of the parameters κ, H, and t are independent, and the parameters
1
give the length scale l = (κHt) 2 . We obtain the dimensionless parameters
ĥ x̂ Lf
Π= , Π1 = ζ = 1 , Π2 = ζf = 1 , Π3 = M. (3.13)
H (κHt) 2 (κHt) 2
1
ĥ = Hθ [ζ, M] , x̂ = ζ [M] (κHt) 2 . (3.14)
Dimensional analysis shows that the tip propagation is proportional to t1/2 when
this scaling analysis is valid. The inner tip position is given by x̂i = ζi [M] (κHt)1/2 ,
and the outer tip position by x̂o = ζo [M] (κHt)1/2 , where ζi and ζo are dimensionless
quantities that depend only on the mobility ratio M. Substituting relationships (3.14)
into (3.2), we obtain a nonlinear ordinary differential equation for θ:
ζ dθ d θ(1 − θ) dθ
− = . (3.15)
2 dζ dζ θ(M − 1) + 1 dζ
The mobility ratio M is the only parameter determining the shape of the similarity
solution at early times. The inner and outer boundaries of integration ζi and ζo are
unknown, and must be determined as part of the solution. The boundary conditions
are:
dθ ζi M dθ ζo
θ (ζi ) = 1, = , θ (ζo ) = 0, =− .
dζ ζi 2 dζ ζo 2
The boundary conditions on θ are the non-dimensional form of (3.12), and the condi-
tions on dθ/dζ come from inserting the conditions on θ into (3.15). Equation (3.15)
and the boundary conditions are invariant under reflection in ζ, so that if θ1 (ζ) is
a particular solution θ1 (−ζ) is also a solution. The physical interpretation of this
reflection is exchanging the position of the fluids on either side of the initially ver-
tical interface. The evolution of the interface at early times has been reduced to a
nonlinear eigenvalue problem for a second order ordinary differential equation, with
two unknown eigenvalues and four boundary conditions. The two additional bound-
ary conditions allow the unique determination of the eigenvalues as a function of the
mobility ratio M. For unit mobility ratio (3.15) reduces to a simpler equation
ζ dθ d dθ
− = θ(1 − θ) (3.16)
2 dζ dζ dζ
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 71
0 0
10 10
−1 −1
i
−ζ
ζo
10 10
(a) (b)
−2 −2
10 10
0 1 2 3 4 0 1 2 3 4
10 10 10 10 10 10 10 10 10 10
M M
Figure 3.5: The numerical values for the outer (a) and inner (b) tip positions are
shown as a function of M (solid lines). The scaling laws (3.17) and (3.18) are shown
as dashed lines.
that has been obtained by Huppert & Woods (1995). The solution is symmetric with
respect to the origin, and the eigenvalues become ζi = −ζo = 1. Huppert & Woods
1
(1995) have obtained the solution θ = 2
(1 + ζ). We note that ζi = −ζo = −1 and
1
θ = 2
(1 − ζ) is also a solution. In the numerical solutions for the case M = 1 we
have chosen ζi < 0 and ζo > 0 (see figure 3.4). This choice places the CO2 on the
left side and the brine on the right side of the tilting interface and is consistent with
geometry shown in figure 3.2.
The nonlinear eigenvalue problem is solved numerically for the shape of the interface
and the tip positions as a function of M. We only need to obtain numerical solutions
for M > 1, the corresponding solutions for M < 1 can be obtained from the transfor-
mation (3.4). A shooting method is used to integrate inward from both boundaries of
the domain. The mismatch of θ and dθ/dζ at the origin was minimized to determine
the eigenvalues for a given value of M. The analytical solution for M = 1 was used as
an initial guess for ζi and ζo , and M was increased incrementally to obtain solutions
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 72
for M > 1. The resulting interface shapes are shown in figure 3.4.
As the mobility ratio increases a gravity tongue develops along one of the horizon-
tal boundaries. In the limit of M → ∞ the interface appears to approach a vertical
line. The CO2 viscosity, µc , is kept constant, because the similarity variable ζ de-
pends on µc through κ. Assuming krc and krb are constant, increasing the mobility,
M = µb /µc · krc /krb , we increase µb , and in the limit µb → ∞ the brine becomes
immobile and the interface remains vertical.
Figure 3.5(a, b) shows the position of the inner and outer tips as a function of
increasing mobility ratio. For large values of M the positions of the tips follow scaling
laws given by
1.24
ζi = −e0.2210 M−0.4997 ≈ − √ , for M > 10, (3.17)
M
2.37
ζo = e0.8645 M−0.4163 ≈ 0.42 , for M > 200, (3.18)
M
which are shown as dashed lines in figure 3.5. The outward propagating non-dimensional
tip position for M < 1 is obtained from the following argument
12
1 κp Ht
x̂o [Mp < 1] = −x̂i [Mq > 1] = −ζi [Mq ] (κq Ht) = 2 −ζi [M−1
p ] .
Mp
The inward propagating tip xi [Mp < 1] can be obtained by an analogous argument.
√
The self-similar tip positions for M < 1 are given by ζo [M < 1] = −ζi [M−1 ]/ M
√
and ζi [M < 1] = −ζo [M−1 ]/ M. The position of the outward propagating tip of the
interface at early times is given by
1
Ld + ζo [M] (κHt) 2 , M ≥ 1,
xeo = 1 (3.19)
Ld − ζi [M−1 ] κHt
M
2
, M < 1,
where the superscript e is used to indicate the scaling for the early similarity solution.
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 73
1.0 1.0
t = 10-2
a = 0.1
0.8 t = 10-1 0.8 a = 0.2
t = 100 a = 0.3
t = 101 a = 0.4
0.6 0.6
θ(ζ) a = 0.5
θ(ζ)
0.4 0.4
0.2 0.2
(a) (b)
0.0 0.0
-5 -2.5 0 2.5 5 -0.5 -0.25 0 0.25 0.5 0.75
ζ ζ
Figure 3.6: (a) The numerical solution to (3.2) with initial condition (3.22) and
M = 10 is plotted at various times, t, and compared to the self-similar solution
obtained from (3.15) (solid grey line). (b) Numerical solutions to (3.2) with M = 20
and initial condition (3.25) is shown for several values of the parameter a in the initial
condition and compared to the solution of (3.15) with a = 0 (solid grey line).
The numerical values of ζo and ζi can be obtained from figure 3.5 or from (3.17) and
(3.18) in the appropriate limits.
The similarity solutions described above were obtained under the assumption of com-
plete similarity in Π2 , which corresponds to a step function initial profile (Lf = 0).
Barenblatt & Zeldovich (1972) have shown that similarity solutions are intermedi-
ate asymptotic solutions for a much larger class of initial conditions. Therefore,
the analysis presented above also applies to initial profiles with a finite front width
(Lf = 0), for which Π2 = 0. For this larger class of initial conditions the similarity
solution will be valid after the details of the initial conditions have dissipated, because
Π2 = Lf / (κHt)1/2 approaches zero for Lf (κHt)1/2 . Hence every particular initial
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 74
L2f
t
te = . (3.21)
κH
This is illustrated for a particular initial condition in figure 3.6(a). The initial condi-
tion is a ramp defined by
⎧
⎪
⎪ x̂ < −0.5,
⎨ 0,
ĥ0 = x̂ + 0.5, −0.5 ≤ x̂ ≤ 0.5, (3.22)
⎪
⎪
⎩ 1, x̂ > 0.5.
Hence the early self-similar solution is valid for te t tb . We can also define
a new length scale Lb = 2xo (tb ), the width of the current at the back-propagation
time . For small M, Lb provides a suitable initial length scale for the late similarity
solution in § 3.3.2. Lb is given by
⎧
−1 √
⎪
⎨ 2Ld 1 + ζi [M −1] M , M < 1,
ζo [M ]
Lb = (3.24)
⎪
⎩ 2Ld 1 + ζo [M] ,
ζi [M]
M ≥ 1.
In some situations the CO2 may not fill the entire depth of the domain, so that the
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 75
where a < H is the thickness of the brine in the injection zone. Figure 3.6(b) compares
numerical solutions to (3.2) with M = 20 for various values of a, with the solution to
(3.15), which corresponds to a = 0. Solutions for a = 0 are also self-similar in the
early similarity variable, so that the outward propagating tip propagates as t1/2 for
a = 0. The position of the tip ζo in self-similar coordinates decreases with increasing
a, but for a/H < 0.2 the difference is less than 10% compared to the values of ζo
given in figure 3.5(a). The evolution of currents that do not occupy the full thickness
of the layer is similar to those investigated here and follows the same early scaling
law. These results indicate clearly that the viscosity of the brine cannot be neglected,
even if the CO2 plume does not occupy the full thickness of the aquifer. A full
investigation of this larger family of similarity solutions for a = 0 is beyond the scope
of this investigation.
At the back-propagation time tb the interface detaches from one of the horizontal
boundaries, and the thickness, h, of the CO2 decreases monotonically as a function
of time (figure 3.3c, d ). At late times h H, and we expect the solution for a finite
layer to be similar to the solution in a half-space. The equation for the half-space
can be obtained from (3.2) by taking the limit for H → ∞, for finite h and M, or
equivalently taking the limit h → 0 for finite H and M. Consider the limit of the
nonlinear diffusion coefficient in (3.2) for small h, keeping M and H constant,
h (H − h)
lim = h. (3.26)
h→0 h(M − 1) + H
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 76
which has been studied intensively, and the similarity solution for a finite release of
fluid into a two-dimensional porous half-space was found by Barenblatt (1952).
We expect that the limit (3.26) becomes a good approximation even if h H
is finite, and (3.27) becomes a good approximation for (3.2) after some time. The
parameter M can vary over several orders of magnitude, and we need to consider its
effect on the validity of approximation (3.26). Consider the approximation in the
denominator of (3.26) for finite but small values of h
h(M − 1) + H ≈ H.
Large values of M require even smaller values of h to allow this approximation. Since
h is a monotonically decreasing function of time, the half-space approximation will
become valid for all M eventually. In § 3.3.3 we develop an expression for the onset of
half-space behaviour as a function of M. For small M the half-space approximation
becomes valid very quickly. In the limit of small mobility ratios we obtain
h (H − h)
lim = h, (3.28)
M →0 h(M − 1) + H
and (3.2) reduces to (3.27) at all times and for all values of h and H. As mentioned
in § 3.1.2, the simplification in this limit is responsible for the success of (3.27) in
problems of unconfined flow, where the ambient fluid is a gas (M 1).
Equation (3.27) depends only on the CO2 mobility, λc , not on the mobility ratio
M. From the global conservation of mass Qc + Qb = 0, we can obtain an expression
for the Darcy velocity qb in the ambient fluid
hc qc
qb = − . (3.29)
(H − hc )
For finite h and qc the flux in the brine qb becomes negligible as H → ∞. In contrast
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 77
to the early evolution, where M is the governing parameter, the problem becomes
independent of the mobility ratio at late times, because the brine is stationary.
Barenblatt’s solution
In either limit the initial condition for the porous medium equation is a particular gas
distribution h̃[x], with finite width Lb , in a half-space otherwise saturated by water.
The volume of current is given by
V = h̃[x]dx = 2Ld H. (3.30)
From the definition of the self-similar coordinate, ς, the tip propagation at late times
is proportional to t1/3 , and the front position at late time is given by
1
xlo = (9κLd Ht) 3 . (3.33)
The superscript l identifies the tip scaling for the late similarity solution. The late
similarity solution depends on the CO2 volume V = 2φLd H, but it is independent of
the local length scale Lf of the initial front, and the mobility ratio M. In contrast
the early tip scaling (3.19) is independent of the global length scale Ld , but depends
on Lf and M.
The similarity solution obtained for the idealized initial condition is an intermedi-
ate asymptotic solution for a larger range of initial conditions with Lb = 0, for times
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 78
larger than
L3b
t
tl = t̃ + , (3.34)
2κg Ld H
where t̃ is the time at which (3.26) becomes valid. This lower bound becomes impor-
tant when M 1, and the porous medium equation becomes valid very quickly. In
this case t̃ = tb and (3.24) is a suitable initial length scale Lb , so that tl becomes
⎛ √ 3 ⎞
L2d −1
⎝8 1 + Mζi [M ] M ⎠.
tl = − (3.35)
κH −1
ζo [M ] ζo [M−1 ]2
The problem is symmetric with respect to the origin, so that the boundary condition
at the origin is ∂η(0, ϑ)/∂ξ = 0, and the outer boundary condition is η(a, ϑ) = 0.
Transition time
The two examples in figures 3.7(a) and 3.7(b) show the numerical transition from
the early to the late similarity solution. The initial condition is the early similarity
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 79
1 1
la M = 1/2 lat M = 10
t e
0.8 0.8 s
e
sim
im
ea
il a
rl
ila
rit
ea
´ ´
rit
sm
y
0.6 0.6
y
r
y
ly
i
so
max(´) max(´)
so
ila
sim
lut
lut
rit
ion
ila
ion
y
0.4 0.4
rit
so
lu
y
ol
tio
s
u tio
n
0.2 0.2 n
(a) (b)
0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
»/»o »/»o
Figure 3.7: The transition of the numerical solution (light lines) from the early to
the late similarity solution (both heavy lines) is shown in scaled coordinates. (a) The
numerical solution is shown at ϑ = ϑb + {1, 4, 10, 30, 100}, where ϑb = 0.61; (b) The
numerical solution is shown at ϑ = ϑb + {101 , 102 , 103 , 104 , 105 }, where ϑb = 7.1.
In figure 3.7(a) the evolution is shown for M = 1/2. In this case the curvature of
the early and the late similarity solution is of the same sign. The main difference
between them is the slope at the origin, where the early similarity solution has a
finite slope, but the late similarity solution has zero slope. In this case the transition
period is relatively short, and the late similarity solution is a good approximation to
the solution for ϑ > 100. Figure 3.7(b) shows the transition for the case M = 10. In
this case the curvature of the early and the late similarity solution is of opposite sign.
The numerical solution adjusts very slowly and the transition period is very long.
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 80
Figure 3.8(a-d ) shows the numerical results for the non-dimensional position of the
outward propagating tip ξo of the CO2 as a function of non-dimensional time ϑ. The
figure shows the effect of increasing the mobility ratio M on the tip propagation and
the timing of the transition. The scaling laws for the tip position obtained from the
early and late similarity solutions are also shown. In non-dimensional coordinates
these scaling laws (3.19, 3.33) simplify to
1
1 + ζo [M] ϑ 2 , M ≥ 1,
ξoe = −1 1
− 12
(3.38)
1 − ζi [M ] ϑ M 2 , M < 1,
1
ξol = (9ϑ) 3 , (3.39)
1. The numerical tip position initially follows the early scaling law ξo ∝ ϑ1/2 , and
then the scaling law for late times ξo ∝ ϑ1/3 .
3. The transition from early to late scaling is short for M ≈ 10−1 (figure 3.8b),
and increases rapidly for M > 10−1 (figure 3.8d ).
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 81
2
10
1
(a) M =
100
1
10 ϑb
ξo − 1
0
10
ϑt
−1
10
−2 −1 0 1 2 3 4 5
10 10 10 10 10 10 10 10
2
10
1
(b) M =
10
1
10 ϑb
ξo − 1
0
10
ϑt
−1
10
−2 −1 0 1 2 3 4 5
10 10 10 10 10 10 10 10
2
10
(c) M = 1
1
10 ϑb
ξo − 1
0
10
ϑt
−1
10
−2 −1 0 1 2 3 4 5
10 10 10 10 10 10 10 10
2
10
(d) M = 10
1
10 ϑb
ξo − 1
0
10
ϑt
−1
10
−2 −1 0 1 2 3 4 5
10 10 10 10 10 10 10 10
ϑ
Figure 3.8: The numerical results for the non-dimensional tip position ξo are shown
as a function of non-dimensional time ϑ, for different mobility ratios M. In all figures
the numerical solution is given by dots (· · · ), the tip scaling from the early similarity
solution by a dashed line (- - -), and the tip scaling from the late similarity solution
as a full line (—). (a) ϑt = 2.0, ϑb = 0.1; (b) ϑt = 2.5, ϑb = 0.24; (c) ϑt = 29.3,
ϑb = 1; (d ) ϑt = 811.3, ϑb = 1; (d ) ϑt = 811.3, ϑb = 7.1.
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 82
4. The tip position follows the early scaling law ξo ∝ ϑ1/2 even after the early
similarity solution has become invalid at ϑb (3.37). In figure 3.8(a-c) the early
scaling law continues to be valid almost up to ϑt . This shows that the finite
depth of the layer continues to have a strong effect on the solution, even after
the interface has detached from one of the boundaries.
Although the early and late scaling behaviours are separated by a transition period,
it is useful to define a dimensionless transition time ϑt that falls within this transition
period. This transition time defines a lower bound for the validity of the late similarity
solution. The difference between the early and the late scaling laws is given by
f [ϑ; M] = ξol − 1 − ξoe , (3.40)
where ξol is given by (3.39), and ξoe by (3.38). We use the substitution ϑ = y 6 to
eliminate ϑ1/2 and ϑ1/3 and obtain a cubic in y. Let Mt denote the value of M, where
the early scaling law is tangent to the late scaling law such that f (ϑ; Mt ) = 0. For
M ≥ Mt the non-dimensional transition time ϑt can be defined as the intersection
of the early and late time scaling laws (figure 3.8c, d ), and is therefore given by
the largest real root of f (ϑ; M ≥ 1) = 0. For M < Mt the two scaling laws do
not intersect, but the transition time can be defined as the point of minimal vertical
distance between the two scaling laws (figures 3.8c and 3.8d ) given by the local
minimum of (3.40). Solving for the appropriate root and the minimum we obtain the
following expression for the non-dimensional transition time
⎧ π θ 6
1
⎪
⎪ 1 + 2 cos −3 , M ≥ 1,
⎨ 9ζo [M] 6
3
M3 6
ϑt = 9ζi [M−1 ]6
1 + 2 cos π3 − 3θ , Mt ≤ M ≤ 1, (3.41)
⎪
⎪
⎩ 64M3
, M ≤ Mt ,
9ζi [M−1 ]6
Due to the change in the definition of the transition time at Mt the graph is not smooth
at this point (figure 3.9). For M < Mt the transition time increases very slowly with
M, while it increases strongly for M ≥ Mt (figure 3.9). Mt can be obtained by finding
the value of M for which the local minimum of (3.40) is zero.
3 2
Mt − ζi M−1 t = 0. (3.42)
4
This equation must be solved numerically, because ζi M−1
t is not known analytically,
and we obtain Mt = 0.1839. For large values of M, ξo (ϑt )
1, and (3.40) simplifies
to f ≈ fˆ = ξol − ξoe and gives a scaling law for the transition time ϑt = 0.45M5/2 .
For small values of M, (3.20) can be used to simplify (3.41) to obtain a constant
ϑt = 1.96. Equation 3.41 is complicated to evaluate, and we therefore introduce
a simple expression based on the two limits discussed above and a simple fit for
intermediate values. ⎧
⎪
⎪
5
2
⎨ 0.45M 2 , 10 < M,
3
ϑt ≈ 36.6M 2 , Mt ≥ M ≤ 102 , (3.43)
⎪
⎪
⎩ 2, M < Mt .
1010
? ϑt ∝ M 5/2
?
105
e
im
iont
it
ns
tra ϑb ∝ M
# ?
ime
on t
8 ←ϑl lower bound late soln. p agati
ϑt -pro
1.96 ← back
100
ϑb ∝ M1/6
10-5
10-4 10-2 100 102 104
M
Figure 3.9: Regime diagram for a finite release of fluid into a horizontal porous slab,
showing the non-dimensional time scales obtained in this study, and the shapes of the
gravity current as a function of the mobility ratio M. The shaded region indicates
the transition period between the similarity solutions. The characteristic time to
dimensionalize all results is td = L2d κ−1 H −1
After the details of the initial condition are lost, the interface shape and dynamics
are asymptotic to an early similarity solution that corresponds to a tilting interface.
The early similarity variable is ζ = x(κHt)−1/2 , so that the non-dimensional tip posi-
tion is given by ξo ∝ ϑ1/2 . During this period the left and the right interfaces evolve
independently, and the length scale of their separation, 2Ld , does not appear in the
similarity variable. In this phase, both fluids move with non-zero velocities, and there-
fore the mobility ratio M determines the shape of the interface. We have not plotted
the lower bound for the onset of the early similarity solution (3.21), because this time
scale depends on the initial width of the front Lf , which is given by ϑe = (Lf /Ld )2 .
The validity of the early similarity solution ends at the back-propagation time, ϑb ,
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 85
because the inward propagating tips of the two initially separated fronts begin to
interact at the origin (figure 3.3b). Figure 3.9 shows that ϑb increases monotonically
with time, and follows simple scaling laws for small and large M. Physically we can
explain the increase of ϑb with increasing M by the increasing viscosity of the brine,
which slows down the inward propagating tip of the tilting interface. The period
during which the early similarity solution is valid increases with increasing M, be-
cause the ϑe = ϑe [M]. As long as Lf = 0 there are always values M ≤ Mc such that
ϑe ≥ ϑb , and hence the early similarity solution is not realized. Mc is determined by
√
the equation Lf /Ld = Mc /ζo [M−1 c ] for Mc < 1, a similar equation can be found for
Mc > 1. Even in the case Lf = 0 the early similarity solution will not be realized
in the limit M → 0, because ϑb → 0. Figure 3.8 shows that the scaling law for the
non-dimensional tip position ξo ∝ ϑ1/2 is valid for a significant time even after the
early similarity solution itself has become invalid at ϑb .
The initial similarity solution is followed by a period where the solution is not
self-similar and must be obtained numerically (figure 3.7). For M 1 we can define
the transition period as ϑb < ϑ < ϑl , where ϑl is given by (3.35). The duration of
the transition period increases as M → 0, because the upper boundary is constant
ϑl = 8, while the lower boundary is proportional to ϑb ∝ M1/6 . For M > 1 we have
no estimate of the upper boundary of the transition period. The numerical results in
figure 3.8(b-d ) show that the transition period increases with increasing M. From the
transition of the scaling laws for the tip position we have defined a transition time ϑt ,
that provides a lower bound on the onset of the late similarity solution. Equation 3.41
or (3.43) shows a rapid increase of ϑt with increasing M for M > Mt .
After the transition period the late similarity solution becomes valid, because
the CO2 occupies only a small fraction of the thickness of the aquifer, and (3.2)
reduces to (3.27). Equation 3.27 admits a similarity transformation in the variable
ς = x/(κV t)−1/3 , and the analytical solution was obtained by Barenblatt (1952). In
contrast to the early similarity solution, this late similarity solution is independent
of M and depends on the volume of the gravity current, given by V = 2Ld H. In
this limit the velocity of the brine is negligible, which explains why the problem
is now independent of the mobility of the brine, and hence the mobility ratio M.
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 86
(a) (e)
-15 10 -5 0 5 10 15 km -5 0 10 20 30 40 50 km
trapped CO2: 4% # = 0.3 trapped CO2: 6% # = 0.3
t = 112 yrs t = 112 yrs
(b) (f)
20% # = 10 19% # = 1.5
t = 3730 yrs t = 560 yrs
(c) (g)
32% # = 100 44% #=4
t = 37300 yrs t = 1500 yrs
(d) (h)
41% # = 1_000 69% #=8
t = 373000 yrs t = 3000 yrs
When the late similarity solution is valid, the non-dimensional tip position is given
by ξo ∝ ϑ1/3 . Again the scaling for the tip position becomes valid before the solution
is fully self-similar.
that maximize residual trapping during or shortly after the injection period (Mo et al.,
2005; Juanes et al., 2006; Ide et al., 2007). We focus on the effect of the vertical sweep
efficiency, S, which is defined as the fraction of the aquifer contacted by the CO2 plume
during its buoyancy-driven migration in the post-injection period. A volume balance
argument shows that the up-dip migration distance, x↑ , is approximately given by
x↑ ≈ Vc / (φHSScr ) , (3.44)
where Vc is the volume of CO2 , H the thickness of the aquifer, and φ the porosity.
Ennis-King & Paterson (2002) used this relationship to highlight the sensitivity of the
migration distance to the magnitude of Scr . Equation (3.44) shows that the effect of S
on the migration distance of the CO2 plume is comparable to that of Scr . The sweep
is expected to be less than unity, because gravity segregation and viscous instabilities
lead to the formation of a gravity tongue along the top of the aquifer.
Initially the two interfaces of the current evolve independently and symmetrically
(figure 3.10a). Until they interact they evolve as tilting interfaces described by a self
similar solution given by a step function initial condition
˜
a, ξ < ξ,
η= (3.45)
˜
b, ξ ≥ ξ,
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 88
1 1
ζi ζi
ζo ζo
η η
=
=
0.5 0 0.5
=
0
=
1
1
(a) (b)
0 0
-1.5 -1 -0.5 0 0.5 1 -0.5 -0.25 0 0.25 0.5 0.75
ζ ζ
20 20
(c) ζo 0.1 (d) ζi
0.2 -0.3
15 0.3 15
0.4
M M
10 0.5 10 -0.4
0.6
-0.5
5 0.7 5
-0.6
0.8 -0.7 -0.8
1 1
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Figure 3.11: (a) Interface shapes for M = 1 and increasing from 0 to 1 in 0.2
increments. (b) Interface shapes for M = 10 and increasing from 0 to 1 in 0.2
increments. (c) Contours of the outward propagating tip position, ζo . (d ) Contours
of the inward propagating tip position, -ζi .
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 89
where ξ˜ is the initial position of the interface. Here we consider only the cases a = 1
and b = 0, but self-similar solutions are not limited to this choice. The outward and
inward propagating tips of the interface are denoted by ξo and ξi respectively. For
ξ˜ < ξ ≤ ξo the interface is advancing, and the thickness of the current is increasing,
ηϑ > 0, so that σ = 1 − . For ξi ≤ ξ < ξ˜ the interface is receding, and the thickness
of the current is decreasing, ηϑ < 0, so that σ = 1. The location ξ˜ of the discontinuity
in the coefficient is constant at early times.
Equation 3.7b and the initial condition (3.45) are self-similar in the variable ζ =
ξ/ϑ1/2 . Therefore, the propagation of the plume tip at early time is given by the
scaling law ξo = a[, M]ϑ1/2 , where a is a constant that decreases with increasing
. Inserting Pe = 0 and the similarity variable into (3.2b), we obtain the ordinary
differential equation
ζ dη d η(1 − η) dη
− = σ̃ , (3.46)
2 dζ dζ η(M − 1) + 1 dζ
where σ[ηζ ζϑ ] = σ̃[ηζ ]. The mobility ratio, M, and the residual, , determine the
shape of the similarity solution at early times. The inner and outer boundaries of
integration ζi and ζo are unknown, and they are generally functions of both M and .
These eigenvalues must be determined as part of the solution, requiring two additional
constraints. The boundary conditions on the inward and outward propagating tips
are given by
dη ζi M
η [ζi ] = 1, = , (3.47)
dζ ζi 2
dη −ζo
η [ζo ] = 0, = . (3.48)
dζ ζo 2(1 − )
respectively. The conditions on η specify the vertical extent of the interface. The
conditions on ηζ have no direct physical interpretation, but they are required to satisfy
(3.15) at the boundaries ζi and ζ0 , and they provide the two additional constraints
necessary to determine the eigenvalues. Together (3.46) and (3.47) form a nonlinear
eigenvalue problem for the shape of the interface and the two unknown tip positions,
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 90
In § 3.3.2 we show that f → η at late times when the current is much thinner than
the thickness of the aquifer, so that (3.7b) reduces to
ηϑ = σ (ηηξ )ξ , (3.49)
for all M. The only parameter in this equation is the residual, , in the discontinuous
coefficient, σ. Semi-analytical solutions for the interface shape are available for the
radial (Kochina et al., 1983; Barenblatt, 1996), and linear cases (Bear & Ryzhik,
1998).
For a compact initial condition, the solution of a parabolic equation with a moving
discontinuous coefficient is a similarity solution of the second kind (Barenblatt, 1996).
These solutions are characterized by an anomalous exponent in the scaling laws that
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 91
102
1/2 (b)
early
late 101
1/3
β ξ̂o
1/4
100 =0
1/8 = 0.25
= 0.5
(a) = 0.75
0 10-1
0 0.25 0.5 0.75 1 10-2 10-1 100 101 102 103 104 105
ϑ
Figure 3.12: Change of the solution from the early to the late scaling law (M = 1).
(a) The scaling laws for ξo ∝ ϑβ at early and late times as a function of the residual,
. The symbols at of 0, 1/4, 1/2, and 3/4 show the exponents obtained by fitting
the numerical data shown in figure b. For the early scaling law the data was fit for
ϑ between 0.1 and 10 and for the late scaling law the data was fit for ϑ > 1000. (b)
˜ is shown as function of ϑ, for increasing residuals,
The shifted tip position, ξˆ = ξo − ξ,
where ξ˜ = 1.
ξo = c1 [, M, η0 ] ϑβ() , (3.50)
ηmax = c2 [, M, η0 ] ϑ2β()−1 , (3.51)
V = c3 [, M, η0 ] ϑ3β()−1 , (3.52)
for the tip of the plume, the height of the plume, and the volume of the plume.
Figure 3.12(a) shows the decrease of the anomalous exponent β with increasing .
For (3.49) the exponent must be determined numerically from a nonlinear eigenvalue
problem (see Bear & Ryzhik, 1998). As expected the increasing residual trapping
leads to a faster decrease of the plume volume and slower tip propagation. Another
characteristic of this type of solution is that the constants of proportionality (c1 -c3 )
in these scaling laws depend on the initial condition.
Figure 3.12(b) shows transition in the scaling law for the tip position from the
early scaling law, with the exponent 1/2, to the late scaling law with the anomalous
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 92
exponent β ≤ 1/3. Initially only the constant in the scaling law is affected by ,
but at late times both the coefficient and the exponent depend on . The sharp-
interface model presented here predicts that the volume of the current always decays
as a power-law in horizontal aquifers. In this case the current is never exhausted and
continues to spread. Capillary forces, which have been neglected in our model, will
arrest the movement of the current once it becomes too thin.
ητ + σληξ = 0, (3.53)
Riemann problems
First we consider the two Riemann problems to understand the effect of trapping,
i.e. the discontinuous coefficient σ, on rarefactions and shocks. Then we consider a
hat function initial condition, representing a finite release of fluid, and the resulting
wave interaction. For non-zero values of loss, the wave interaction will lead to an
extinction of the current in finite time.
0.06 1
(a) ²=0 (b)
0.25 0.75
0.04
¾f 0.5 ´
0.5
0.02
0.75
0.25
² = 1
0 1 ²=0
0
0 0.25 0.5 0.75 1 -0.5 0 0.5 1
´ »=¿
Figure 3.13: (a) The flux function f is shown for M = 10 and increasing trapping .
(b) The corresponding solution profiles given by (3.56) in self-similar coordinates.
˜
ξ = σ [ητ ] λ[η]τ + ξ, (3.55)
therefore all solutions are self-similar in the coordinate ξ/τ . The discontinuity in
˜ but the interface itself
σ leads to a discontinuity of the slope of the interface at ξ,
˜ The equation for the evolution of η[ξ, τ ],
remains continuous, because λ = 0 at ξ.
given by inverting (3.55) is
⎧
⎪
⎪ 1, ξ ≤ ξi ,
⎪
⎪
⎪
⎪
⎪
⎪
−τ +M̂(ξ0 −ξ)+ Mτ (τ +M̂(ξ−ξ0 ))
˜
⎪
⎪ , ξi < ξ < ξ,
⎨ M̂(τ +M̂(ξ−ξ0 ))
1 ˜
η [ξ, τ ; < 1, M = 1] = √ , ξ = ξ, (3.56a)
⎪
⎪
(1+ M)
⎪
⎪ −τ ˆ+M̂(ξ0 −ξ)+ Mτ ˆ(τ ˆ+M̂(ξ−ξ0 ))
⎪
⎪ , ξ˜ < ξ ≤ ξo ,
⎪
⎪ M̂(τ ˆ+M̂(ξ−ξ0 ))
⎪
⎪
⎩ 0, ξo < ξ,
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 94
˜ τ ˜
ξo = (1 − )τ + ξ, and ξi = − + ξ. (3.57)
M
For tilting interfaces in the parabolic limit we have shown that ξi is a stronger func-
tion of M than . In the hyperbolic limit the entire receding interface has lost its
dependence on . Where the interface is advancing, f is scaled by σ = 1 − (fig-
ure 3.13a), and the residual reduces the propagation speed of the advancing interface
(figure 3.13b). This behaviour is analogous to the self-similar solutions in the par-
abolic limit. The solution for = 1 is still given by (3.56), but the advancing interface
is replaced by a discontinuity at ξ˜ with a jump from 0 to ηmax . A comparison of fig-
ures 3.11b and figure 3.13b shows that M and have a similar effect on the shape of
the interface in both the parabolic and hyperbolic limits.
For M = 0, equation 3.53 reduces to the linear advection equation
ητ + σηξ = 0, (3.58)
where ηL < ηR . The weak solution of (3.53) with (3.60) is a shock of strength,
ηs = ηR − ηL , given by
˜
ηR , ξ ≤ Λτ + ξ,
η [ξ, τ ; M] = (3.61)
ηL , Λτ + ξ˜ < ξ,
where Λ is the shock speed. Similar to (3.58) an ambiguity arises, because ητ is singu-
lar at the shock. Therefore we define the shock speed considering (3.53) regularized
by a diffusion term
ητ + σ [ητ ] fη ηξ = νηξξ , (3.62)
where ν > 0. For initial data (3.60) the entropy-satisfying solution to (3.53) is defined
by the solution of (3.62) in the limit of small ν, the so-called vanishing viscosity solu-
tion. We assume the solution of (3.62) with (3.60) takes the form of a traveling wave
with constant speed Λν , so that η [ξ, τ ] = η̂ [ζ], where the traveling wave coordinate
is, ζ = ξ − Λν τ . For a steady traveling wave the time derivative that determines the
value of σ is given by
ητ = ζτ η̂ζ = −Λν η̂ζ . (3.63)
For initial data (3.60) we expect the solution to be monotonically increasing, η̂ζ > 0,
so that the sign of ητ is determined by the direction of wave propagation, given by the
sign of Λν . The constant wave speed implies that ητ does not change sign, and hence
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 96
σ is continuous, but its magnitude depends on the sign of the wave speed, σ = σ̂ [Λν ].
In the moving coordinate system (3.62) reduces to the ordinary differential equa-
tion with the boundary conditions η̂ → ηR for ζ → +∞, and η̂ → ηL for ζ → −∞.
After integrating once and eliminating the constant of integration with either bound-
ary condition, the following expression is obtained
This confirms our assumption that η̂ζ > 0 for ηL < η̂ < ηR . Equation (3.64) does
not depend on the wave speed, because consistency with both boundary conditions
requires that the wave speed be given by
1 − M̂ηR ηL − ηL − ηR
Λν = σ , (3.65)
M̂ηL + 1 M̂ηR + 1
which has already been substituted into (3.64). For M = 1 the solution to (3.64) is
1 (ζ − Cν)σ (ηR − ηL )
η̂ = ηL + ηR + (ηR − ηL ) tanh , (3.66)
2 2ν
f [ηL ] − f [ηR ]
Λ = σ̂ = Λν , (3.68)
ηL − ηR
so that the shock speed of the limiting solution is identical to the wave speed of the
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 97
10 -2 10 -2 10 -1
6 6 1
(a) (b) (c)
0. 5
4 4
σf σf Λ rec.
0
adv.
2 2
adv. −0. 5
rec.
0 0 −1
0 0.25 0. 5 0.75 1 0 0.25 0. 5 0.75 1 0 0.25 0. 5 0.75 1
η η ηL
Figure 3.14: Shock construction for M = 10. (a) In the absence of trapping, = 0,
the construction for advancing and receding shocks is the same. (b) In the presence
of trapping, = 1/2, the shock construction for advancing shocks is modified. (c)
Shock speed as function of ηL , for ηL > ηc the shock speed depends on the amount
of trapping.
viscous solution.
The geometric interpretation of the jump condition (3.68), is a cord joining f [ηL ]
and f [ηR ], as shown in figure 3.14a and b. For a given ηR and M the direction of
wave propagation as function of ηL is given by
⎧
⎪
⎪
⎨ < 0, ηc < ηL < ηR ,
Λν = 0, ηL = ηc , (3.69)
⎪
⎪
⎩ > 0, 0 < η < η ,
L c
where
1 − ηR
ηc = , (3.70)
(M − 1)ηR + 1
which corresponds to a horizontal tangent and a stationary shock.
The speed of shocks moving to the right is not affected by trapping, while the
speed of waves traveling to the left is reduced due to trapping (figure 3.14c). This
behavior of the shock is analogous to the behavior of the rarefaction. In both cases
the speed of the interface is reduced, by a factor 1 − due to trapping, when it is
advancing, but it is not affected if the interface is receding. However, the parameter
that identifies whether the interface recedes or not is different for the rarefaction and
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 98
the shock. For an interface with a finite slope, i.e. a rarefaction, the change of the
height of the interface with time, ητ , determines which parts of the interface recede.
For a vertical interface, i.e. a shock, the sign of shock speed, Λ, determines whether
the interface recedes across the shock or not.
Wave interaction
The evolution of a finite release of fluid is governed by the interaction of the shock
at the back of the current with the rarefaction at the front of the current. In the
presence of trapping this interaction leads to a rapid reduction of current volume,
and give rise to a finite migration distance and time. Consider the piecewise constant
initial data
1, |ξ| ≤ 1/2,
ηH = (3.71)
0, |ξ| > 1/2.
Initially the solution of the front is a rarefaction given by 3.56 with ξ˜ = 12 , for all
M > 0, and the evolution of the back is a stationary shock (A to B in figure 3.15)
with constant shock strength (figure 3.16b), because Λ = 0 for ηL = 0 and ηR = 1
for all M > 0. When the back end of the rarefaction reaches the shock, the shock
strength decreases, and it begins to move. The solution, η, will then be given by
(3.56), but η [ξ < ξs ] = 0, where the ξs is the shock position.
ξηs = ληs [ηs ] τ + λ [ηs ] τηs and ηs ξηs = f [ηs ] τηs . (3.72)
for the evolution of the shock strength. Together with the initial condition ηs [τt ] = 1,
it determines the evolution of the shock strength (figure 3.16b). Once ηs is known,
the shock position is given by ξs = λ [ηs ] τ + 12 . We solve for ηs instead of ξs , because
(3.73) is always separable.
If trapping occurs, = 0, then the evolution of ηs is divided into an early period,
τt ≤ τ ≤ τp , during which the shock interacts with the receding portion of the
rarefaction (B to C in figure 3.15), and a later period τp < τ ≤ τf , where the shock
interacts with the advancing portion of the rarefaction (C to D). Hence, the problem
has three time scales: the transition time, τt , when interaction between waves begins
(B ), the passing time, τp , when the shock passes the stationary point of the rarefaction
and starts to interact with the advancing section of the rarefaction (C ), and the final
migration time, τf , when the volume of the current goes to zero (D).
In the early period the shock strength and position are independent of and are
given by
1 1 √ √ 2
ηs = √ and ξs = − + M− τ . (3.74)
1 − M + Mτ 2
We have shown in (§3.4.2) that the receding section of the interface is not affected by
trapping. Therefore, the shock strength and position are independent of the amount
of trapping. The early period extends until the shock migrates past the initial extent
of the current, ξs < 12 . The passing time and the shock strength at the passing time
are given by
√ 2 1
τp = 1 + M and ηs [τp ] = √ . (3.75)
1+ M
In the absence of trapping (3.74) remain valid for all τ > τt (dashed line in figure 3.16).
After τp an explicit expression for ηs and ξs is only obtained for M = 1. We give the
full details for the case M = 1 to illustrate the structure of the solution, and for M > 1
we integrate (3.73) numerically, using (3.75) as an initial condition. For τp ≤ τ ≤ τf
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 100
16 50
(a) (b)
D D
40
12
30
τ 8 τ
20
C
4 C
10
B
B
A
0 A 0
-1.5 0F 2 4 6 8 10 -5 0F 10 20 30
ξ ξ
(c)
η 16
12
8
τ
0 4
-1 0 2 4 6 8 10 0
ξ
Figure 3.15: Characteristic portraits of the evolution of initial data (3.71), in the
ξτ -plane. The current is bounded by the shock, A to D, in the back and the leading
characteristic of the rarefaction, F to D, in the front. (a) Analytic solution for M = 1
and = 0.4. (b) Semi-analytic solution for M = 5 and = 0.4. (c) Evolution of the
current interface (black) and the residual surface (grey).
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 101
1
For = 2
the right hand side of (3.73) loses the explicit dependence on ηs , leading to
a simplified expression. The expression for the shock path is given by
⎧
⎨ (−1)
t−2
1
2+ −1
t
1
2−2 + 12 , = 12 ,
2−1
ξs = (3.77)
⎩ 1 (log( t )t + 1), = 12 .
2 4
In the second phase the shock speed and position depend strongly on the amount
of trapping. Figure 3.16a illustrates the effect of increasing trapping on the current
evolution in the ξτ -plane for M = 1, and similar behavior is observed numerically for
other M (figure 3.15b). From (3.77), (3.76), and (3.56) the aspect ratio of the current
is given by ⎧
⎪
ξo − ξs ⎨ 1 + (1 − )τ, τ < 1,
⎪
A= = 2τ − τ 3/2 , 1 < τ ≤ 4, (3.78)
ηs ⎪
⎪
⎩ 2(1 − )τ, 4 < τ < τf ,
for M = 1. Even in the presence of trapping A is always increasing for M ≥ 1. The
final residual surface, ηr , is the interface that divides the part of the aquifer that has
been invaded by the current, and now contains trapped fluid, from the region of the
1
aquifer that has not been invaded. The final residual surface for ξ < 2
is the initial
1
condition. For ξ > 2
the final residual surface is defined as ηr = ηs [ξs ] and can be
calculated by eliminating τ between (3.76) and (3.77) to obtain ξ = ξ [ηr ]. However,
it is not possible to invert this relationship to obtain an explicit expression for ηr
⎧
⎨ 1
+ (−1) 1 1
ϕ(ηr ) − 22+ −1 ϕ(ηr ) 2−2 , = 12 .
2 2−1
ξ= (3.79)
⎩ e1−2ηr (2 − 4ηr ) + 1 , = 12 .
2
2−1
1
−1
− −1
were ϕ [ηr ] = 2 (−2ηr + ηr + ) . The shape of the final residual surface
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 102
16
(b) (a)
14
12
10
τ 8 4
0.
=
0
5
=
0.
6
=
.6
τp
0
0.7
=
4 C
=
=1
2
τt B
A F
0
1 0.5 0
1 (c)
ηs
0.5
ηr
0
-1 0 2 4 6 8 10
ξ
Figure 3.16: (a) Portrait of the shock path and the leading characteristic of the
rarefaction, for increasing residual. The locus of the termination points, τf = τf [ξ↑ ],
is shown as a dotted line. (b) The corresponding temporal evolution of the shock
strength, ηs . (c) The corresponding final residual surfaces, ηr = ηr [τf ].
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 103
20 20
(a) log(ξ↑) (b) log(τf)
15 8 15
8
6 6
M M
10 4 10
4
2
5 5
0 2
1 1
0 0.2 0.4 0.6 0. 8 1 0 0. 2 0.4 0.6 0.8 1
Figure 3.17: Contour plots of the up-dip migration distance, log(ξ↑ ), on the left and of
the total migration time, log(τf ), on the right. The data for several potential storage
aquifers in Alberta, Canada, is shown as circles to indicate typical parameter values
(Bachu & Bennion, 2007).
Final migration time and distance The finite migration distance is an impor-
tant characteristic of the hyperbolic limit, in contrast to the self-similar solutions for
the horizontal case. The maximum migration distance and the corresponding total
migration time are amongst the most important time and length scales for CO2 stor-
age, and the analytic and semi-analytic results (figure 3.17) presented here show how
these scales depend on the governing parameters.
The total migration distance, ξ↑ , is given by
1 1
1
2
− 21+ 2−1 ( − 1) 2−1 −1 , = 12 ,
ξ↑ = 1
(3.80)
2
+ 2e, = 12 .
Increasing mobility increases ξ↑ , because a gravity tongue forms at the top of the
aquifer and only a small fraction of the aquifer is swept and contributes to residual
trapping.
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 104
The dimensionless final migration time increases strongly with increasing M. The
dimensionless time is only based on the properties of the released fluid. As M increases
the released fluid becomes increasingly less mobile, which retards the movement of
the released fluid.
Volume evolution The final migration time, τf , bounds the time available for
leakage from the storage reservoir. For smaller times the volume of mobile CO2 is a
basic quantity necessary to estimate maximum potential leakage over time. Analytic
expressions for the evolution of the current volume are given by integrating (3.56)
over the extent of the plume. The normalized current volume V = V (τ )/V (τ0 ) is
given by ξo √ −2
V= η dξ = 1 − τ 1 + M (3.82a)
ξs
for τ < τp . In this early period both the shock and the receding part of the rarefaction
are independent of , and therefore the area containing residual saturation is also
independent of . Hence, the current volume decreases proportional to . The decay
of the current volume slows down with increasing M, because the receding interface
slows down. This behavior that is also observed in semi-analytic solutions for τ > τp .
At later times the lower limit of the integral is only known for M = 1, and in this
case the current volume is given by
⎧ 2
⎨ 1− 1
2− 1− t 2(1−) − t , = 12 ,
V= t(1−2)2 (3.82b)
⎩ t 2
t
8
log 4 − 1 , = 12 ,
for τ < τf = 4e. Figure 3.18(a) shows how strongly the current volume is affected by
the magnitude of . This evolution is distinctly different from the power-law decay
obtained from the self-similar solutions for horizontal aquifers. It suggests that sloping
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 105
1 1
(a) (b)
0.8 0.8
0.6 0.6
V V
0.4 0.1 0.4
1
0.2 5
0.2 0.3 0.2 10
15 20 40
0.4 25
0 0
0 25 50 0 100 200 300
τ τ
Figure 3.18: The effect of the governing parameters on the evolution of the current
volume. The passing and final migration times, are shown as circles and triangles.
(a) Increasing the residual, , from 0.1 to 0.9 at constant M = 1. (b) Increasing the
mobility ratio, M, from 1 to 40 at constant = 0.4. Circles denote
0.2 hyp.
(a) 1 0 num.
sim.
´ 0.1 M = 20 15 10 5
0
0 10 20 30 40 50 60
»
0.1
(b) hyp.
num.
´ ² = 0
0.1
0.2
0.3
0
10 20 30 40 50 60
»
Figure 3.19: This figure shows several comparisons between the numerical solution
(∆ξ = 0.01) to (3.7a) and the hyperbolic solution for Pe = 2 at τ = 50. (a) Increasing
M at constant = 0. (b) Increasing at constant M = 10.
In CO2 storage Pe is not necessarily large, but the mobility ratio is generally larger
than unity, M ≈ 5. Here we show that the particular shape of f leads to near
hyperbolic behavior of (3.7a) for small Pe and large M. Consider the differentiated
form of (3.7a) given by
σ −1 ητ + ληξ = Pe−1 λ (ηξ )2 + f ηξξ . (3.83)
In the absence of source terms the thickness η of the current decreases monoton-
ically with time and for M
1 the current takes the shape of a triangle composed
of a very elongate, flat, advancing tongue in the front and a short, steep, receding
tongue in the back (figure 3.10g-h). This asymmetry increases with Pe, M, and τ .
The hyperbolic approximation introduced in § 3.4.2 replaces the short, steep, receding
interface of the current by a shock.
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 107
t [yrs]
37 373 3’730 37’300 373’000
100 100
10-1 10-1
V
full equation:
Pe = 0 (θ = 0 ◦ )
◦
Pe = 0 .44 (θ = 12 )
◦
10-2 Pe = 0 .87 (θ = 1 ) 10-2
Pe = 4 .37 (θ = 5 ◦ )
Pe = 8 .82 (θ = 1 0 ◦ )
Pe = 1 3 .4 (θ = 1 5 ◦ )
limiting cases:
similarity soln.
hyperbolic soln.
10-3 10-3
10-1 100 101 102 103
ϑ = t/td = tκ1 H cos θ/L2
In the advancing tongue ηξξ ηξ 1, and hence both the diffusive, f ηξξ , and the
second order term, ληξ2 are small. The diffusive term is reduced further over the entire
√
current, because max(f ) = 1/(1 + M)2 decreases with increasing M. The second
order term ηξ2 is not small in the steep receding back of the current, and figure 3.19
shows that the error in the hyperbolic approximation is localized there.
Figure 3.19(a) illustrates how the error introduced by the hyperbolic approxima-
tion decreases with increasing M. For M > 5 and τ > 2τp the only significant error in
the hyperbolic approximation is due to small off-set in the shock position. In the limit
M = 0 a similarity solution to the full equation is available (Huppert & Woods, 1995),
and it is shown to provide an indication of the accuracy of the numerical method used.
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 108
Once loss is introduced, > 0, the small differences in the sweep of the full solution
and the hyperbolic solution lead to differences in plume volume at any given time.
This volume error accumulates over time and leads to differences between the shock
location in the hyperbolic model and the back of the plume in the numerical solution
(figure 3.19b).
In gently sloping media the current transitions from initial parabolic behavior to
hyperbolic behavior at late times (figure 3.20). Due to the distinctly different behavior
of these limiting cases for > 0 the transition is apparent in the evolution of the
current volume. An initial power-law decay of the current volume is followed by a
much more rapid decay characteristic of the hyperbolic limit.
For very large Pe the plume volume is given 3.82 and its semi-analytic extension
to M = 1. This limiting solution is plotted in figure 3.20 as a function of the diffusive
dimensionless time, ϑ = τ /Pe, so that changes in Pe correspond to a stretching of
the time axis. We observe that the limiting solution describes the rapid decay of the
current volume at late times, even for Pe < 1.
The length and time scales that are important for gravity currents in sloping layers are
the propagation distance against gravity, ξ↓ , and the time of the associated reversal
in the direction of the interface movement, τt . For a buoyant current the down-dip
propagation of the interface is due to the parabolic part of the equation. The reversal
of the interface movement indicates the transition from early near-parabolic to late
near-hyperbolic behavior. The continued volume loss and the resulting extinction
of the current give rise to additional length and time scales, the up-dip migration
distance, ξ↑ , and the final migration time, τf . Numerical results for all four scales as
a function of increasing Pe are shown in figure 3.21. To determine the scales associated
with the end of the current the numerical simulation was terminated at V = 10−3 .
This value was chosen to allow a reasonable detection of the gravity current tips,
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 109
2
20
(a) numerical: (b)
M=0
1.5 M=1 15
M=5
-»↓ M = 10
M = 15 ¿t
1 10
M = 20
0.5 5
similarity soln.
hyperbolic soln. 0
0
0 5 10 15 20 0 5 10 15 20
Pe Pe
100 160
(c) (d)
75 120
»↑ ¿f
50 80
25 40
0 0
0 5 10 15 20 0 5 10 15 20
Pe Pe
Figure 3.21: Time and length scales obtained from the numerical solution of (3.7)
are shown as symbols (∆ξ = 0.01). All results are for a residual of = 0.4. The
limiting hyperbolic and self-similar solutions are indicated by full and dashed lines.
(a) Down-dip migration distance, ξ↓ . (b) Transition time, τt . (c) Up-dip migration
time, ξ↑ . (d ) Final migration time, τf .
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 110
which become strongly affected by numerical diffusion for smaller cut-off volumes.
The rapid decay of the plume volume at late times (figure 3.20) suggests that the
results are not very sensitive to the cut-off, and the reasonable agreement between
numerical and limiting analytic results in figure 3.17 confirms this.
The down-dip migration distance, ξ↓ , is not very sensitive to M or . For small Pe
√
it follows the relation τt = 1/ 6Pe obtained for the similarity solution for M = = 0.
As the τt increases with decreasing Pe the current thickness decreases as well, and
the similarity solution for the unconfined aquifer becomes a better approximation.
As Pe increases ξ↓ approaches the position of the initial condition, which is also the
down-dip migration distance in the hyperbolic limit. The transition time, τt , appears
to approach the limiting hyperbolic solution much slower than the other parameters.
The up-dip migration distance, ξ↑ , and the final migration time, τf , show a more
complicated behavior as Pe increases. As Pe goes to zero both become very large,
and this is consistent with the late similarity solution for the limiting case Pe = 0,
introduced in section § 3.4.2. As Pe increases the hyperbolic limit is approached
quickly, except for M = 0. However, for Pe ≈ 0.2 − 0.5 a minimum in both quantities
is observed for M > 1. Hesse et al. (2006) showed that both ξ↑ and τf decrease with
increasing Pe for M = 0. Figures 3.17(c-d ) show a similar behavior of ξ↑ and τf for
M < 1, but we observe an increase of ξ↑ and τf for M > 1 for Pe > 1.
3.5 Discussion
3.5.1 Assumptions
The most limiting assumptions in the derivation of the governing equations are the
homogeneity of the porous medium, the uniform saturation in the CO2 plume, and the
incompressibility of the CO2 . While simple forms of heterogeneity in the across slope
direction can be included (Huppert & Woods, 1995) general forms of heterogeneity
require a full numerical solution. Ide et al. (2007) investigated gravitational spreading
of a CO2 plume in a horizontal, two-dimensional, heterogeneous aquifers with a reser-
voir simulator. They showed that the effect of capillary forces on residual trapping
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 111
which is likely to exhibit this behavior. This model could still be solved by the method
of characteristics allowing fast and accurate solutions to be obtained by numerical
integration.
In terms of geological CO2 storage we have neglected the dissolution of CO2 into
the brine. Two types of dissolution must be distinguished, direct dissolution of CO2
into the residual brine and the enhanced dissolution of CO2 at the interface between
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 112
CO2 and brine due to convective currents in the underlying brine, see § 2. The direct
dissolution into the residual brine can be modeled in this framework, but the effect is
likely to be small compared with residual trapping. The enhanced dissolution due to
convective transport may become important at later times, because the gravity tongue
provides a large interfacial area. This process could be modeled conceptually through
an effective loss term. Detailed information regarding the convective dissolution rates,
similar to those presented in § 2.3, will allow us to estimate the magnitude of this loss
term. An extension of the model presented here would then allow for a comparison
of the relative contributions of dissolution and residual trapping.
The leading edge of the CO2 plume is both by gravitationally and viscously un-
stable and this leads to the formation of a gravity tongue in our solutions. Riaz &
Tchelepi (2006) have shown that this gravity tongue is also the dominant feature
in high resolution numerical simulations of two-dimensional vertical displacements.
The displacement may also be viscously unstable in the transverse direction, similar
to the Saffman-Taylor instability in a Hele-Shaw cell (Saffman & Taylor, 1958), or
the instability of viscous currents on inclined planes discussed by Lister (1992). The
linear instability of an immiscible displacement in a porous medium has been studied
by Yortsos & Hickernell (1989) and Riaz & Meiburg (2004a). However, these studies
consider advection dominated one dimensional base flows and not the two-dimensional
base flow in gravity-capillary equilibrium that is of interest here. The study of the
stability of the two-dimensional flows presented here is an interesting direction for
future research, but beyond the scope of this contribution.
Let us consider a deep saline aquifer with the following constant properties: H ≈
20 m, k ≈ 30 mD, and φ ≈ 0.1. In addition, we assume the following constant
fluid properties: ∆ρ ≈ 300 kg/m3 , Scr = Sbr = 0.2, µc ≈ 0.06 · 10−3 , µb /µc ≈ 10,
krc ≈ 0.2, and at the advancing interface krb = 1. In this case, the mobility ratio is
M ≈ 5, so that we expect the formation of a gravity tongue of CO2 along the top
of the aquifer. Finally, we assume that the aquifer has been invaded by CO2 over a
distance of Ld ≈ 500 m, so that the aspect ratio at the end of injection is given by
A = 2 · Ld /H ≈ 50.
Horizontal aquifers The evolution of the CO2 plume in a horizontal aquifer with
these properties is shown in figure 3.10(a-d ), and the evolution of the volume of the
mobile CO2 plume is given in figure 3.20. For gravity currents with M ≈ 5 we expect
an extended early period where the tip position is given by x ∝ t1/2 (figure 3.9).
The regime diagram in figure 3.9 does not account for the effect of residual trapping,
but figure 3.12(b) shows that the transition from the early to the late scaling law is
similar for = 0.25 and = 0. We can therefore use the theory developed in § 3.3 to
estimate the duration of the early period, given by tb , and the time, tt , after which
we can expect the late scaling law to hold.
For M = 5, the dimensionless time scales are given by ϑb ≈ 4 and ϑt ≈ 265,
and the characteristic time is td = L2d (κ1 H)−1 ≈ 250 yrs. The interface will detach
from the aquifer after approximately tb ≈ 1000 yrs. At this time the footprint of the
CO2 plume is 4.6 km. In the period right after the detachment of the CO2 plume
residual trapping is the most effective, and the fraction of CO2 trapped as residual
saturation increases to 20% (figure 3.10b). The period that corresponds to rapid
residual trapping in a horizontal aquifer is approximately given by tb to 3tb .
The tip of the CO2 plume will continue to propagate proportional to t1/2 until
tt ≈ 66000 yrs, when the footprint of the plume is 20 km. Therefore, although the
CO2 plume shown in figure 3.10(c) is already very thin compared to the thickness of
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 114
the aquifer, the plume tips are is still propagating according to the early scaling law,
x ∝ t1/2 . This illustrates that the effects of confinement, expressed in the mobility
ratio, must be taken into account to understand the evolution of the CO2 plume. In
particular, confinement has an effect on the propagation of the gravity current long
after the CO2 plume has detached from the base of the aquifer at tb .
After tt the spreading of the CO2 plume slows down significantly, and in fig-
ure 3.10(d ) the tips of the plume propagate according to the late scaling law. In this
late stage, residual trapping decreases the propagation velocity in a manner similar
to figure 3.12. As the plume slows down, residual trapping decreases rapidly. From
figure 3.10(a) to (d ) we can see that it takes 3,000 yrs to trap the first 20%, about
30,000 yrs to trap the next 10%, and approximately 300,000 yrs to trap an additional
10%. This drastic decay of the rate of residual trapping in horizontal aquifers is a
consequence of the self-similar nature of the solution at late times, which requires
that the volume of the mobile CO2 plume decays as a power-law.
Sloping aquifers Figure 3.20 shows that the effectiveness of residual trapping in-
creases dramatically even for small dip angles. The evolution of the interface shape
for θ = 5◦ is shown in figure 3.10(e-h). We see that the migration time of the cur-
rent is reduced from approximately 40,000 yrs for a slope of 1/2◦ to approximately
2,000 yrs for a slope of 15◦ . The up-dip migration distance decreases from 85.6 km
at 1/2◦ to a minimum of 82.8 km at 1◦ and slowly increases again to 84.5 km at 15◦ .
As the slope of the aquifer increases the distance the plume can migrate before
it reaches the surface decreases. For an injection depth of 2 km, only angles of 1/2◦
and 1◦ prevent the CO2 from reaching the surface, and even for θ = 1/2◦ the CO2
rises 750 m vertically. Therefore, the depth of the current at the end of migration
is the important criterion for storage security. Given the dimensionless initial depth
d0 = D0 /L the dimensionless final depth is d = −d0 + ξ↑ sin θ. We can define the
critical angle at which the injected CO2 reaches the surface as
d0
θc = arcsin (3.85)
ξ↑
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 115
20 20
(a) θc (b) S
0.1° 0.01%
15 15
0.5°
1°
2° 0.1%
M 4° M
10 6° 10 1%
10°
20° 10%
5 35° 5
90° 50%
90%
1 1
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Figure 3.22: Contour plots of the critical angle, θc , assuming d0 = 2 and the sweep,
S, as function of M and . The data for several potential storage aquifers in Alberta,
Canada, is shown as circles to indicate typical parameter values (Bachu & Bennion,
2007).
Figure 3.22(a) shows that the critical angle, θc , increases with and decreases with
M. For a given M and the CO2 will not reach the surface for θ < θc . The optimal
storage aquifer therefore has a few degrees of slope, just enough to cause lateral up-
dip migration and the corresponding efficient residual trapping, but without reducing
the distance to the surface too much.
Despite rapid trapping compared to the horizontal case, CO2 storage in saline
aquifers is limited by the poor vertical sweep, S, and the resulting long migration
distances. The vertical sweep can be defined as S =
hr /H =
ηr , where
hr is the
average residual surface. The initial volume of mobile CO2 is given by φ(1 − Sbr )HL,
and the residual volume of CO2 at the end of the migration is given by φScr
hr (x↑ −
x↓ ). An expression for the vertical sweep, S, can be obtained from the conservation
of volume and is given by
S= , (3.86)
ξ↑ − ξ↓
where ξ↓ ≈ 0 in the hyperbolic limit. Figure 3.22(b) shows that the sweep is a strong
function of , increasing monotonically from zero at = 0 to unity for = 1. For the
data from Bachu & Bennion (2007), the sweep, S, can vary from less than 10−3 to
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 116
more than 10−1 , roughly three orders of magnitude, while the residual, , itself only
varies between 0.1 and 0.7. Therefore, the residual CO2 saturation, Scr , influences
the up-dip migration distance given by (3.44) more strongly through its influence on
the sweep than through its direct volumetric effect.
Nicot (2008) considered the storage of roughly a fifth of the CO2 emissions from
coal power stations in the state of Texas over the next 50 years. This would re-
quire the injection of approximately 370·106 m3 /yr of supercritical CO2 , equivalent
to 50 MtCO2 /yr, into the central section of the Carrizo-Wilcox aquifer. They envi-
sioned a line of 50 wells aligned perpendicular to the dip of the aquifer with a 1.6 km
spacing. These wells act as a 80 km long line source and our one-dimensional solution
is a reasonable approximation of the displacement away from the edges.
To obtain first order estimates of the migration distance of the injected CO2 we
assume the following constant aquifer properties estimated from the data given in
Nicot (2008): H ≈ 200 m, k ≈ 500 mD, φ ≈ 0.15, the average depth of injection
2.7 km, the average distance to the outcrop 100 km, and therefore and average dip
angle, θ ≈ 1.5◦ . We assume the following fluid properties: ∆ρ ≈ 300 kg/m3 , Scr =
Sbr = 0.2, µc ≈ 0.06·10−3 , µb /µc ≈ 10, krc ≈ 0.2, krb = 1 (at the advancing interface).
Given this data the appropriate length scale at the end of 50 years of injection is
L ≈ 10 km, so that the initial aspect ratio is A = L/H ≈ 50. The three governing
parameters are therefore M ≈ 5, ≈ 0.25, and Pe = A tan θ ≈ 1.4. Figure 3.21 shows
that the migration distance and time for M = 5 are already close to the hyperbolic
limit for Pe = 1.4, so that we obtain ξ↑ ≈ 80 and τf ≈ 110 from figure 3.17.
The along slope migration distance would be approximately ξ↑ L = 800 km, so
that the migration distance exceeds the up-dip extent of the aquifer significantly.
Under the simplifying assumptions made here, this particular injection scenario does
not ensure that all CO2 would be trapped by residual saturation alone. The initial
length scale L would have to be reduced to 100 km/ξ↑ = 1.25 km to achieve residual
trapping of all injected CO2 . This could be achieved by increasing the width of the
injection zone from 80 to approximately 620 km. Another option would be to reduce
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 117
the volume injected into this aquifer from 50 to roughly 6.5 MtCO2 /yr, assuming the
original injection width of 80 km. In this case the CO2 would remain mobile for 1550
years after the end of injection.
3.6 Conclusion
We derive a vertical-equilibrium sharp-interface model governing the migration of an
immiscible gravity current in a two-dimensional confined aquifer. A simple model of
residual trapping allows the analysis of the relationship between the vertical sweep of
CO2 plume and residual trapping.
In the parabolic limit of the governing equations, corresponding to a horizontal
aquifer, the evolution of the CO2 plume is divided into three regimes. The mobility
ratio M is the parameter that determines the time scales separating these regimes.
We have obtained new similarity solutions in the variable ζ = x(κHt)−1/2 that are
valid at early times when the interface is tilting, due to a horizontal exchange flow.
These similarity solutions are a strong function of the mobility ratio M, and of the
residual, . In this regime the position of the tip of the interface is given by x ∝ t1/2 ,
so that the exponent is independent of . The numerical solution continues to follow
the early scaling law long after the interface has detached from one boundary. This
indicates that the finite thickness of the aquifer is important for gravity currents with
M > 1, and solutions that assume infinite depth are not valid until the current has
become very thin.
In the limits h → 0 and M → 0, the governing equation simplifies to the porous
medium equation. For = 0, this equation admits a similarity solution in the variable
ς = x/(κV t)−1/3 , where the position of the tip of the interface is given by x ∝ t1/3 .
For = 0, the exponent in is a function of , x ∝ tβ() . This dependence is given by a
nonlinear eigenvalue problem. We obtain an expression for the transition time tt from
the t1/2 to the tβ() scaling. The transition time, tt , increases monotonically with M,
but it is a weak function of M for M < 0.18, and it increases rapidly for M > 0.18. The
two self-similar regimes are separated by a transition period that is roughly centered
on tt . Numerical solutions of the governing partial differential equation are used to
CHAPTER 3. RESIDUAL TRAPPING & CO2 GRAVITY CURRENTS 118
describe the evolution during the non-self-similar transition period. Numerical results
show good agreement with the early and the late similarity solutions.
During CO2 storage in horizontal saline aquifers M ≈ 5, and we expect a prolonged
period where the current spreads according to the early scaling law. Residual trapping
will be most effective in the time just after the current detaches from the base of the
aquifer. In the long term limit, the the current spreads very slowly and residual
trapping becomes less important. Despite continued loss of volume, the current is
never reduced to zero and continues to spread. This behavior is an artefact of the
model that neglects capillary pressure.
In the hyperbolic limit of the governing equations, appropriate for dipping aquifers,
the current is described by a travelling wave with hysteresis. The leading edge of the
interface is a rarefaction, and the trailing edge is a shock. The current volume is
reduced continuously by the residual, and the volume is reduced to zero in finite
time. We have obtained expressions for the up-dip migration distance and the final
migration time of the current.
In both the parabolic self-similar solutions and the hyperbolic rarefaction and
shock we observe that residual trapping affects the advancing section of the interface
more strongly than the receding section of the interface. Although trapping occurs
at the receding interface, the advancing interface is affected more strongly because
less fluid is supplied to it from the receding section.
In gently sloping aquifers, an initial near-parabolic regime with power-law decay
of the volume is followed by a near-hyperbolic regime with very rapid volume decay.
Increasing the slope of the aquifer or the initial aspect ratio of the current reduces
the duration of initial parabolic period.
Our results suggest that lateral migration of the injected CO2 along the seal will
trap the CO2 relatively quickly as residual saturation. Residual trapping is optimized
in sloping aquifers with small mobility ratios and high residual CO2 saturations.
However, the long migration distances of CO2 , due to the formation of a gravity
tongue, may limit the volume of CO2 that can be stored in sloping regional aquifers.
Chapter 4
Multiscale simulation of
CO2 storage in saline aquifers
4.1 Introduction
119
CHAPTER 4. MULTISCALE SIMULATION 120
recovery, much larger domains have to be simulated and over much longer time pe-
riods. The CO2 plume may migrate laterally over 10’s to 100’s of kilometers in
regional aquifers, and toward the surface along wells and fractures. Nordbotten et al.
(2004) and Silin et al. (2006) showed that the interaction of the CO2 with shallower
aquifers and aquitards along the leakage pathway can reduce the leakage rate to the
atmosphere significantly. It may therefore be necessary to simulate fluid movement
in the overburden of the storage site, leading to even larger simulation domains. In
addition, large quantities of CO2 have to be injected to lead to a meaningful reduction
of CO2 emissions, and this may lead to displacements of the interface between brine
and potable groundwater on a basin scale (Nicot, 2008).
On the other hand, geological porous media, including saline aquifers, are hetero-
geneous on all length scales (Journel, 1986), and numerical simulation with very high
spatial resolution is necessary to capture the complex flow patterns that develop in
unstable displacements (Tchelepi & Orr , 1994). The simulation of CO2 storage also
requires high numerical resolution to capture the small length-scales associated with
important physical processes, such as diffusion and dispersion. In § 2 we show that
buoyancy driven fingers in the brine determine the dissolution rate of CO2 , and may
require even higher decimeter-scale resolution. In § 3 we show that the adverse mobil-
ity ratio leads to dynamical instabilities such as long thin gravity tongues and viscous
fingers that require sub-meter-scale resolution in numerical simulation. The length
and time scales associated with the physical mechanisms during CO2 storage may
span six orders of magnitude in a sloping high-permeability aquifer, and this range
of scales considers only the continuum description where Darcy’s law is thought to
apply.
Finally, the uncertainty in the description of the geological storage site requires
many simulations to assess the uncertainty in model predictions (Kovscek & Wang,
2005). Similarly, the updating of the geological model through monitoring data re-
quires a large number of simulations (Doughty et al., 2007). Current reservoir simula-
tion technology may not be able to provide adequate numerical resolution to capture
physical processes and at the same time allow for the reasonable exploration of the
CHAPTER 4. MULTISCALE SIMULATION 121
0 km
Carrizo Fm.
1 km
10 km
2 km
Figure 4.1: An example of a typical regional aquifer under consideration for CO2
storage, the Carrizo Formation in east Texas is shown in grey. The figure is vertically
exaggerated 16:1, so that the true aspect ratio is much larger than it appears in the
figure. The sketch is simplified from Nicot (2008).
et al., 2008).
Grids with high aspect ratios are common in reservoir simulation, because oil
reservoirs are kilometers in areal extent and tens of meters thick, leading to simula-
tion domains with high aspect ratios (100:1 in 2D). In addition, geological layering
and fractures can introduce severe anisotropy, which is often miss-aligned with the
computational grid. Grid aspect ratios and permeability anisotropy have similar ef-
fects on the numerical discretization (see § 4.3.1) and can lead to significant challenges
in numerical modeling.
Large regional aquifers extend for several tens or even a few hundred kilometers,
but they are at most a few hundred meters thick (figure 4.1), leading to even larger
aspect ratios than in reservoir simulation of oil and gas fields (1000:1 in 2D). There-
fore, the problems with accurate and robust numerical discretization of anisotropic
domains with high aspect ratios are even more relevant in CO2 storage than in con-
ventional reservoir simulation. The original MSFV method, however, is only robust
for uniform grids and isotropic fine-scale permeability (Kippe et al., 2007; Lunati &
Jenny, 2007). The development of a multiscale finite volume discretization that is
robust for large aspect ratios is therefore an important step towards the simulation
of CO2 storage at the field scale.
∂Sc
φ + ∇ · uc = −fp , (4.1)
∂t
∂Sb
φ + ∇ · ub = −fp , (4.2)
∂t
where Si , ui , and fi denote the saturation, Darcy velocity and source term (well) of
fluid phase i, and φ is the porosity. The subscripts c and b for the phase denote CO2
and brine. The saturation, Si , is the volume fraction of phase i in the pore space, so
that Sc + Sb = 1. For our problem, the viscosities of the CO2 and brine phases can be
different, but the viscosity of each phase is assumed to be constant. We also assume
that capillarity and gravity effects are negligible, so that Darcy’s law in terms of the
phase mobility, λi , is given by
kri (Si )
ui = −λi ∇p = −K(x) ∇p, (4.3)
µi
where K(x) is the absolute permeability tensor field, p is the pressure, and kri and
µi are the relative permeability and viscosity of phase i. Summing the conservation
equations of both phases we obtain
∇ · u = f, (4.4)
where f = −fc − fb , u = uc + ub = −λ∇p is the total velocity, and the total mobility,
λ, is defined as
krc (Sc ) krb (Sb )
λ(x, Si ) = K(x) + . (4.5)
µc µb
This system of equations is supplemented with appropriate initial and boundary con-
ditions. In reservoir simulation, the system is often assumed to be closed, so that
no-flow boundary conditions are imposed at the periphery.
The total velocity, u, defined above allows us to rewrite (4.1) in terms of the
fractional flow, fi = ui /u, of phase i. Incompressible two-phase flow in porous media
CHAPTER 4. MULTISCALE SIMULATION 124
Equations (4.6) and (4.7) are usually referred to as the flow and transport problems,
respectively. The flow problem is an elliptic equation for p, and the transport problem
is a hyperbolic conservation law for Si . They are coupled through the dependence of
λ on the saturations Si , which are a function of space and time. These equations are
representative of the type of system that must be solved accurately and efficiently by
a simulator for CO2 storage.
At every time step, tn , in the solution process (4.6) has to be solved for the pressure
field, and in this context the total mobility is only a function of space. Therefore, we
drop the dependence of λ on µp and krp in (4.5) and set λ(x, tn ) = K(x) to emphasize
its tensorial nature.
Aarnes, 2004), the two-scale conservative subgrid approach (Arbogast, 2002), the fi-
nite difference heterogeneous multiscale method (Abdulle & E, 2003), the variational
multiscale method (Juanes & Dub, 2008), and many others.
We discuss the multiscale finite volume (MSFV) method for elliptic problems,
which was introduced by Jenny et al. (2003, 2005, 2006). The method described in
§4.2 is identical to the original MSFV method, but we describe it from a different
perspective that motivated the new results and developments presented in §4.4 & §4.6.
For simplicity, we restrict the discussion to a two-dimensional (2D) orthogonal grid.
In § 4.1.2 we show that the pressure field in incompressible porous media flows is given
by the scalar elliptic boundary value problem (4.6). Introducing operator notation
the complete definition of the flow problem is
Here x = (x, y), and Ω ⊂ R2 is a rectangular open domain with boundary ∂Ω. L
is the linear elliptic operator on Ω and LΓ represents the linear boundary operators.
The elliptic operator L is defined by
where the tensor K is symmetric positive definite. In problems of flow and transport
in porous media the pressure equation, (4.8), is coupled to one or more transport
equations, given by (4.7).
Here h is a discretization parameter defining the centers of the finite control volumes
or cells Gh := {(x, y) : xi = (i + 12 )h, yj = (j + 12 )h; i ∈ (0, ..., n − 1), j ∈ (0, ..., m − 1)},
where m and n are the number of cells in the x- and y-directions. We only consider
square grids, because simple forms of grid stretching and skewness can be absorbed
into an effective permeability tensor, see § 4.3.1. The control volume centered on
(xi , yj ) is denoted Ω̄hij , and the four interaction regions that connect it to the sur-
(d)
rounding eight neighbors are labeled Ω̃hij , where d ∈ (1, 2, 3, 4). When we discuss
a control volume and its associated interaction regions, we drop the i, j subscript
(figure 4.2). The discrete operator for the ij-th cell is given by
1
Lh ph (xi , yj ) = ma,b ph (xi + ah, yj + bh) = [ma,b ]h ph , (4.11)
a,b=−1
where we have dropped the i, j subscript on the discrete operator, Lh , and the coef-
ficients, ma,b , for clarity. For homogeneous problems it is convenient to express the
discrete operator using the stencil notation (Trottenberg et al., 2001)
⎡ ⎤
m−1,1 m0,1 m1,1
⎢ ⎥
Lh ph = [ma,b ]h ph = ⎢
⎣ m −1,0 m 0,0 m 1,0
⎥ ph .
⎦ (4.12)
m−1,−1 m0,−1 m1,−1
h
1
ma,b = 0. (4.13)
a,b=−1
~ s1 f1
f2(3) f4(4) ~
a) b)
Figure 4.2: (a) Nine cells of the finite volume grid are shown by dashed lines. The
four dual interaction regions, Ω̃(1) to Ω̃(4) , which connect the shaded finite volume
cell in the center to its eight nearest neighbors are shown by solid lines. (b) A dual
interaction region is shown with the four coarse cell boundary segments labeled s1 to
s4 and the corresponding fluxes f1 to f4 .
With (4.13) the stencil [ma,b ]h has eight degrees of freedom. These eight coefficients
are determined by considering the fluxes induced across the cell boundaries by unit
sources in the eight neighboring cells (figure 4.2a). Every cell is linked to its eight
neighbors by four interaction regions, Ω̃(1) to Ω̃(4) . Each interaction region contains
four cell boundary segments s1 to s4 (figure 4.2b). When considering a single inter-
action region the pressures in the corners are labeled p1 to p4 as shown in figure 4.2b.
In this contribution we only consider MPFA schemes based on the interaction
regions shown in figure 4.2a. Other choices of the interaction regions are possible
and have been discussed by Lambers et al. (2008). The MPFA schemes considered
here assume that the fluxes f1 to f4 across cell boundaries in each interaction region
can be expressed as a linear combination of the four coarse pressures p1 to p4 in the
corners of the dual
⎛ ⎞ ⎛ ⎞⎛ ⎞
f1 t1,1 t1,2 t1,3 t1,4 p1
⎜ ⎟ ⎜ ⎟⎜ ⎟
⎜ f2 ⎟ ⎜ t2,1 t2,2 t2,3 t2,4 ⎟ ⎜ p2 ⎟
⎜ ⎟ ⎜ ⎟⎜ ⎟
⎜ f ⎟=⎜ t ⎟⎜ p ⎟. (4.15)
⎝ 3 ⎠ ⎝ 3,1 t3,2 t3,3 t3,4 ⎠⎝ 3 ⎠
f4 t4,1 t4,2 t4,3 t4,4 p4
where [ma,b ]h differs from Lee et al. (1998) by a minus sign due to the definition of
Lh . The form of these equations can be understood from figure 4.2a.
In this framework, any method that determines T will lead to a conservative dis-
cretization, Lh . Originally MPFA methods were developed for a full tensor permeabil-
ity field that is constant in every control volume, Ω̄h , and analytical considerations
have been used to obtain the entries in T (Aavatsmark, 2002). In reservoir engi-
neering the tensor permeability field is usually computed using upscaling techniques
(Durlofsky, 1991).
The MSFV method addresses this problem by a three step procedure (Jenny et al.,
2003): 1) Local numerical fine-scale solutions in each interaction region, are used to
construct a coarse operator that accounts for fine-scale information on a coarser grid.
2) Problem 4.8 is solved on the coarse grid with the coarse operator. 3) A locally
conservative fine-scale flux field is reconstructed solving local fine-scale problems in
each coarse cell. The MSFV flux reconstruction on the fine-scale is consistent with
the fluxes of the coarse pressure field. The MSFV method is distinguished from
other upscaling methods by the construction of the coarse operator and the locally
conservative flux reconstruction that is consistent with the coarse operator.
The geostatistical or fine-scale permeability field is commonly isotropic Journel
(1986), although skewed and stretched grids can introduce an effective fine-scale per-
meability anisotropy (see § 4.3.1). All multiscale methods introduced in § 4.1.3 con-
sider isotropic fine-scale permeability fields on rectangular grids, corresponding to
a diagonal effective fine-scale permeability tensor field. The multiscale method dis-
cussed here also assumes a diagonal fine-scale permeability tensor, except in § 4.4
where fine-scale anisotropy is introduced to allow an analysis of the MSFV method
for the homogeneous case.
The coarse grid has a grid size H = nc h, where the integer nc is the coarsening ratio,
assumed to be the same in both directions. In porous media applications values of nc
between 3 and 15 are a common choice, so that the coarse linear system can be up
two orders of magnitude smaller than the fine system in two dimensional problems
(Jenny et al., 2003). The centers of the coarse finite control volumes are defined by
GH := {(x, y) : xi = (i + 12 )H, yj = (j + 12 )H; i ∈ (0, ..., N − 1), j ∈ (0, ..., M − 1)}.
The coarse control volume centered on node (xi , yj ) is denoted by Ω̄Hij . The coarse
CHAPTER 4. MULTISCALE SIMULATION 131
LH pH = [ma,b ]H pH , (4.25)
The t∗m,k are column vectors and the matrix is of size (4 · ns ) × 4, where ns is the
number of fine cells on the coarse boundary segment. The transmissibilities on the
coarse-scale, T, can be obtained by summing the column vectors in T∗ . Therefore, the
CHAPTER 4. MULTISCALE SIMULATION 132
Here the linear operator L̃h is the restriction of Lh to the open domain Ω̃H ⊂ Ωh , and
L̃Γh is the linear boundary operator. The local support of the L̃h requires that artificial
boundary conditions are imposed on Γ̃. The choice of L̃Γhl is an important design choice
in the MSFV method. The two most common choices for L̃Γhl , are the linear boundary
condition and the oscillatory boundary condition, both were introduced by Hou &
Wu (1997) in the context of the multiscale finite element method.
For any vector pk that contains the solution in the corners of Ω̃H , the boundary
condition g̃hl (pk ) interpolates the solution values onto the sides of Ω̃H . The simplest
choice is linear interpolation, in this case L̃Γhl := I, where I is the identity and the
boundary condition is given by
0 0 0 0 0 1 1 0
1 0 0 1 0 0 0 0
t4,1 t2,2
t3,1 t3,2 t3,3 t3,4
t2,1 t4,2
Figure 4.3: The top row shows contour plots of the four standard basis functions of the
MSFV method, calulated numerically for a heterogeneous, isotropic fine permeability
distribution. The second row shows the underlying dual interaction region with the
fine grid and the fine fluxes t∗m,k along the coarse boundaries (small arrows) and the
total coarse fluxes tm,k integrated over each face (big arrows).
where n ∈ (1, 2, 3, 4) is circular, and xb is the appropriate coordinate along the bound-
ary. The oscillatory boundary condition is obtained by solving a reduced elliptic
problem along the boundary of Ω̃H . The reduced boundary operator is defined by
shown that local oversampling can reduce this error at higher computational cost
(Hou & Wu, 1997; Hou et al., 1999). In all results presented in this paper the domain
has not been oversampled, but all developments presented below can be used with
oversampling.
In the MSFV method these four local solutions are interpreted as basis functions,
and the vectors of corner pressures are chosen as pk = [δ1k , δ2k , δ3k , δ4k ]T , where δik is
the Kronecker delta. In the homogeneous case the solution to (4.27) reduces to the
bilinear basis functions known from finite element methods. We also use the term
bilinear local problem or basis function for heterogeneous problems with the same pk .
Similarly we use the term linear basis function or local problem for cases where the
homogeneous limit admits a linear solution. Figure 4.3 shows these solutions for a
heterogeneous, isotropic permeability field. From each of the four local solutions we
obtain a vector of fine transmissibilities, t∗m,k , across each of the coarse cell boundary
segments. These vectors contain the fine transmissibilities ordered by increasing x or
y-coordinate. At the boundaries the dual interaction regions overlap by a fine cell so
that the corresponding entries in t∗m,k are half of the associated fine-scale flux. The
, s ∗
coarse transmissibilities are given by tm,k = ni=1 (tm,k )i , where ns is the number of
fine cells along the coarse boundary segment. The four local problems give a total
of sixteen constraints that determine T uniquely in a given dual interaction region
Ω̃H . The MSFV basis functions are a particularly convenient way of determining the
entries of T, because the entries can be obtained directly by integrating the fluxes in
the local problems along the coarse cell boundaries (figure 4.3). In general any set of
four linearly independent local pressure solutions can be used to determine T.
(d)
Once the transmissibility matrices are known for all dual interaction regions Ω̃Hij ,
the coarse operator can be assembled. The coarse solution pH is obtained from the
coarse linear system, LH pH = fH .
Here the linear elliptic operator L̄h is the restriction of Lh onto the open domain
Ω̄H ⊂ Ωh , and L̄Γh is the linear boundary operator. To obtain a fine reconstruction
that is locally conservative on Ωh the linear boundary operator L̄Γh must impose the
fluxes on Γ̄, and is hence defined by
∂ p̄h
L̄Γh p̄h := −kn = ḡ(pH ), (4.31)
∂xn
where ns is the length of the boundary segments and (p1 , p2 , p3 , p4 )T are the values
of the coarse solution pH in the corners of the dual interaction region. Once the fine
CHAPTER 4. MULTISCALE SIMULATION 136
local local
maximum maximum
Figure 4.4: The fine-scale solution as well as the reconstructed conservative and the
interpolated multiscale solutions are shown for the same problem. The log of the
permeability is shown at the base of each figure. A unit pressure gradient is applied
in the x-direction, and homogeneous Neumann boundary conditions are applied in
the y-directions. (a) The fine-scale solution, ph . (b) The conservative solution from
the flux reconstruction on the fine grid, p̄h . (c) The coarse solution, pH , interpolated
onto the fine grid using the MSFV basis functions, p̂h . In (b) and (c) the coarse
solution, pH , is shown as large black dots.
fluxes in all dual interaction regions are known, the fluxes on the entire boundary
of the coarse cell ḡ(pH ) can be assembled. Again, care should be taken to properly
treat the fine-scale fluxes at the boundaries of dual interaction regions; these fluxes
receive contributions from two interaction regions. To specify the unknown constant
in (4.30) the fine solution is set to the coarse solution at the center of the coarse cell.
The reconstructed solution is conservative on the fine grid, and the corresponding
fine flux field can be used in transport calculations.
In §4.4.2 the coarse operator, LH , is modified by changing the entries of T∗ ,
and it is important to maintain consistency between the coarse operator and the
reconstruction. The reconstruction of the flux boundary condition given in (4.32) is
always consistent with LH .
CHAPTER 4. MULTISCALE SIMULATION 137
where ∆x and ∆y are the length of the finite control volumes on the rectangular
grid. The relation between the aspect ratio, A = ∆x/∆y, used by Kippe et al.
(2007) and Lunati & Jenny (2007), and the effective permeability tensor used here is
therefore a/b ∝ A−2 . In all heterogeneous numerical examples Kr is isotropic and K
is diagonal, in this case a/b = A−2 .
To understand the loss of monotonicity of the MSFV method we need to under-
stand the effect of coarse-scale anisotropy introduced by heterogeneity on the fine
scale. However, it is difficult to analyze the properties of multiscale methods for
heterogeneous cases with long correlation length. Durlofsky (1991) shows that the
effect of the fine-scale heterogeneity on the coarse scale is equivalent to a homoge-
neous anisotropic fine-scale permeability. This equivalent fine-scale permeability is
generally not aligned with the grid, c = 0. In § 4.4 we use this analogy to study the
properties of the coarse operator of the MSFV method on a homogeneous anisotropic
fine-scale permeability field. Numerical examples in § 4.7 show that the behavior of
MSFV for these problems gives a good indication of its accuracy and robustness for
CHAPTER 4. MULTISCALE SIMULATION 139
−3 −5
x 10 x 10
0.04 pH 1 4
a) b) c)
ph
0.03
pH , ph
2
0.02 0.5
0.01 0
0 0
0 0.5 1 0 0.5 1 0 0.5 1
x x x
Figure 4.5: Three solution profiles in the x-direction located at y = 1/2 are shown
for increasing values of the ratio a/b with c = 0. (a) anisotropy ratio, a/b = 1 > 1/3,
corresponding to A = ∆x/∆y = 1 . (b) anisotropy ratio, a/b = 1/100 < 1/3,
corresponding to A = 10. (c) anisotropy ratio, a/b = 1/2500 < 1/3, corresponding
to A = 50.
where λ1 and λ2 are the principle components of the effective tensor (λ1 ≥ λ2 ) and θ
is the angle between the x-axis and the eigenvector associated with λ1 .
Loss of monotonicity
Incompressible flow in porous media leads to elliptic problems with strongly heteroge-
neous anisotropic coefficients. Many numerical discretizations of the elliptic operators
lead to unphysical oscillatory solutions, and the design of robust schemes has received
a lot of attention. Generally it is not possible to construct a 9-point discretization
that is locally conservative, exact for linear pressure fields, and monotonic for all
CHAPTER 4. MULTISCALE SIMULATION 140
Kippe et al. (2007) and Lunati & Jenny (2007) have reported unphysical oscillatory
solutions that occur when the MSFV method is applied to heterogeneous anisotropic
problems. In heterogeneous domains it is not clear whether the loss of monotonicity is
caused by the assumptions on the boundary condition of the local problems 4.27, or by
the properties of LH itself. The following example will show that pH has oscillations
even in homogeneous domains, indicating loss of monotonicity of the coarse operator
LH .
Consider (4.10), with Ωh = (0, 1) × (0, 1), a square grid of 135 × 135 fine cells,
homogeneous Dirichlet boundary conditions, so that LΓh = I and gh = 0, the source
term
1 1/3 < x, y < 2/3,
fh = (4.40)
0 otherwise,
5 0
a) b)
-1
0
-2 λ /λ = 2.5: O(1)
1 2
MSFV
m1,0
m1,1
λ /λ = 5.0: O(1)
1 2
-3 MSFV
-5
λ /λ = 7.5: O(1)
1 2
-4 MSFV
λ1/λ2 = 10 : O(1)
MSFV
-10 -5
0 45 90 135 180 0 45 90 135 180
θ θ
Figure 4.6: Numerical results for the stencil components m1,0 and m1,1 of the MSFV
method are compared to the analytic expressions of the MPFA O(1) method. The
components are shown as a function of the orientation of the tensor, θ, for several
ratios of the principal components, λ1 /λ2 .
Consider the same problem solved with the MSFV method with a coarsening ratio
nc = 5, giving 27 × 27 coarse cells. Figures 4.5a & b show that the coarse solution pH
is monotonic for a/b close to unity, but figure 4.5c shows that oscillations and local
minima occur as a/b decreases further.
region. For grid aligned anisotropy, c = 0, the basis functions are given by
The analytical solutions for the bases functions are not known for anisotropy not
aligned with the grid (c = 0), but these basis functions can be calculated numerically.
The discussion is therefore, initially restricted to problems with aligned anisotropy,
and in this case the matrix T is given by
⎛ ⎞
3a −3a −a a
⎜ ⎟
1⎜ b 3b −3b −b ⎟
T= ⎜
⎜
⎟. (4.41)
8⎝ a −a −3a 3a ⎟
⎠
3b b −b −3b
From (4.16) to (4.23) we obtain the coarse operator of the MSFV method as
⎡ ⎤
− (a + b) , 2 (a − 3b) , − (a + b)
1⎢ ⎥
⎢
LH = [ma,b ]H = ⎣ 2 (b − 3a) , 12(a + b), 2 (b − 3a) ⎥⎦ , (4.42)
8
− (a + b) , 2 (a − 3b) , − (a + b)
H
for homogeneous problems with aligned anisotropy, c = 0. The coefficients m±1,0 and
m0,±1 in the stencil are only non-positive for an anisotropy ratio, a/b, between 1/3 and
3, outside this narrow interval LH violates (4.37). This is the cause of the oscillations
in homogeneous problems with aligned anisotropy observed in §4.3.1. The violation
of (4.36) to (4.39) does not immediately lead to a coarse solution pH that violates
Hopf’s maximum principle. In figure 4.5b the coarse solution pH is still oscillation
free although LH violates (4.37), but oscillations occur eventually if the anisotropy
increases further.
The coarse operator of the MSFV method is similar to the MPFA O(η) method,
for η = 1. This family of schemes was first introduced by Edwards & Rogers (1998),
however we follow the parametrization of Nordbotten et al. (2007) in terms of the
parameter η ∈ [0, 1). Although the general construction of the O(η) method requires
CHAPTER 4. MULTISCALE SIMULATION 143
θ= θ=
40 º 40 º
0.8 0.8
r
d rato
θ=
r
o
θ=
to
h
et p e i a
30
30
se c l
0.6 0.6
er
-m se o
θ=2
º
θ=2
e p
a n
)
θ = 10
θ = 10
1
O( coar
o
a/b
b
co st
a/b
b
0º
t
0º
=a
r
=a
V in
FV po
º
º
c2
0.4
c2
0.4
MS 7
SF -
it:
it:
lim
lim
M
C
0.2 0.2
ic
pt
ic
i i pt
ell ell
a) b)
0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
|c|/b |c|/b
Figure 4.7: The monotonicity regions of the original and compact coarse operator are
shown in the parameter space a/b vs. |c|/b. A gray grid shows the corresponding
angle, θ, and the ratio of principal components, λ1 /λ2 in the effective permeability
tensor given by (4.35). All methods are monotone above the labeled line. (a) The
monotonicity regions of both the MPFA O(1) method and the coarse operator of the
MSFV method are compared. (b) The monotonicity region of the CMSFV method
equals that of the 7-point stencil.
η ∈ [0, 1), for the homogeneous case the O(1) method is well defined Nordbotten et al.
(2007). Previously the O(0) method has been emphasized by Aavatsmark (2002), the
O(1/2) method by Edwards & Rogers (1998), and the O(0.9) method by Pal et al.
(2007).
For homogeneous problems with aligned anisotropy (c = 0) the MSFV coarse
operator is identical to the MPFA O(1) method. For c = 0 we have used numerical
solutions on the fine grid to compute the MSFV coarse operator. The MPFA O(0)
method was used to compute the basis functions on the fine grid. Figure 4.6 compares
the numerical results for the MSFV method to the analytic expression for the O(1)
method. Both methods have a similar behavior as the tensor is rotated. The difference
between the methods increases as the ratio of the principal components of K increases.
CHAPTER 4. MULTISCALE SIMULATION 144
The region where a stencil satisfies (4.36) to (4.39) will be referred to as the
monotonicity region of the stencil. The monotonicity region of the MSFV method is
shown in figure 4.7a and has the same characteristics as that of the O(1) method.
The monotonicity of both methods is very limited when anisotropy is aligned with
the grid and largest, if the tensor is rotated by 45◦ . This unusual property of the
original MSFV method is an unanticipated consequence of using bilinear dual basis
functions within the finite volume framework of § 4.1.5.
is second order accurate. Two of the diagonal neighbors are always zero depending
on the sign of c, and therefore the stencil has 7 non-zero entries for c = 0. All four
diagonal neighbors are zero for c = 0, and the method reduces to the standard 5-
point stencil. Nordbotten et al. (2007) have shown that the 7-point stencil has an
optimal monotonicity region, in the sense that no other 9-point stencil satisfies con-
ditions 4.37 to 4.39 over a larger range of parameters (figure 4.7b). For homogeneous
problems and inside its monotonicity region this 7-point stencil also leads to a linear
system, where Lh is a M-matrix. The 7-point stencil can be characterized using a
3-point flux approximation (figure 4.8), leading to one zero entry in every row of the
CHAPTER 4. MULTISCALE SIMULATION 145
P4 P3 P4 P3
f3 f3
f4 f2 f4 f2
f1 f1
P1 P2 P1 P2
a) c > 0, 0° < θ < 90° b) c < 0, 0° > θ > −90°
Figure 4.8: The triangular domains of dependence of the fluxes in a dual interaction
region are shown for both signs of c. The corresponding orientation of the permeability
tensor is indicated by the ellipses.
transmissibility matrix
⎧ ⎛ ⎞
⎪
⎪ a, c − a, −c, 0
⎪
⎪ ⎜ ⎟
⎪
⎪ ⎜ c, b − c, −b, ⎟
⎪
⎪ 1 ⎜ 0 ⎟
⎪
⎪ 2 ⎜ c, ⎟ c ≥ 0,
⎪
⎪ ⎝ 0, −a, a − c ⎠
⎪
⎪
⎪
⎪
⎪
⎨ b, 0, −c, c − b
T= (4.44)
⎪
⎪ ⎛ ⎞
⎪
⎪ a − c, −a,
⎪
⎪
0, c
⎪
⎪ ⎜ ⎟
⎪
⎪ ⎜ 0, c − b, −c ⎟
b,
⎪
⎪ 1⎜ ⎟ c < 0,
⎪ 2⎜
⎪
⎪
⎪ ⎝ 0, −c, c − a, a ⎟⎠
⎪
⎩
b − c, c, 0, −b
whose position depends on the sign of the off-diagonal component, c. Figure 4.8 shows
that the domain of dependence of each flux is aligned with the direction of maximum
connectivity. We refer to a stencil with three non-zero entries in each row of T as
a compact operator. The definition is based on T rather than the resulting stencil,
because the latter may lose its zero entries in the heterogeneous case, while the zero
entries in T always remain.
CHAPTER 4. MULTISCALE SIMULATION 146
4
Φ1 = φi , Φ2 = φ1 + φ4 , Φ3 = φ1 + φ2 , and Φ4 = φ4 ,
i=1
can be used to determine the transmissibilities tm,k . This change in the basis does
not affect the coarse operator, but it separates the linear and the bilinear parts of the
solution. Linear solutions can be represented exactly by the first three bases, and we
eliminate Φ4 to obtain the four degrees of freedom necessary to modify T. The first
three basis functions Φ1 to Φ3 provide the following twelve constraints
0 = T̄( 1 1 1 1 )T , (4.45)
f x = T̄( 1 0 0 1 )T , (4.46)
f y = T̄( 1 1 0 0 )T , (4.47)
on T̄. The definition of T̄ is completed by placing a zero entry in each row. As in the
homogeneous case the zeros are placed to align the stencil with the principle direction
of anisotropy, which can be determined from the local flow information in (4.46) and
CHAPTER 4. MULTISCALE SIMULATION 147
(4.47). The compact transmissibility matrix, T̄, has the following rows,
⎧
⎨ f x , f y − f x , −f y , 0 , f1y ≥ 0,
1 1
t̄1,1 , t̄1,2 , t̄1,3 , t̄1,4 = 1 1
(4.48)
⎩ f x − f y , −f x , 0, f y , f1y < 0,
1 1 1 1
⎧
⎨ f x , f y − f x , −f y , 0 , f2x ≥ 0,
2 2
t̄2,1 , t̄2,2 , t̄2,3 , t̄2,4 = 2 2
(4.49)
⎩ 0, f y , f x − f y , −f x , f2x < 0,
2 2 2 2
⎧
⎨ f y , 0, −f x , f x − f y , f3y ≥ 0,
t̄3,1 , t̄3,2 , t̄3,3 , t̄3,4 = 3 3 3 3
(4.50)
⎩ 0, −f y , f y − f x , f x , f3y < 0,
3 3 3 3
⎧
⎨ f y , 0, −f x , f x − f y , f4x ≥ 0,
t̄4,1 , t̄4,2 , t̄4,3 , t̄4,4 = 4 4 4 4
(4.51)
⎩ f y − f x , f x , 0, −f y , f4x < 0,
4 4 4 4
Figure 4.9: The approach of the global fine-scale solution to transverse equilibrium,
as the effective anisotropy a/b increases. The boundary conditions for this problem
are p(x = 0, y) = 1, p(x = 1, y) = 0, and py (x, y = 0, 1) = 0 and the effective
permeability on the fine-scale is diagonal, c = 0. The log of the permeability field is
shown at the base.
The second term is zero due to the homogeneous Neumann boundary conditions. In
a boundary layer, δy ∝ A, near the Neumann boundaries the transverse gradients in
the solution vanish, ∂p/∂y ≈ 0. In the limit of large A the boundary layers from the
opposite sides of the domain merge so that p is in TVE and no longer a function of
y, and (4.53) is approximately equal to
1
1 ∂ ∂p(x) ∂ ∂p(x)
b(x)dy =
ay (x) ≈ 0, (4.54)
A2 ∂x 0 ∂x ∂x ∂y
CHAPTER 4. MULTISCALE SIMULATION 150
where
ay (x) is the average of a(x) in the y-direction.
a) b) c) d)
1 1 1 1
p̂h
p̂h
p̂h
ph
Figure 4.10: Comparison between the fine-scale solution ph and the interpolated
pressure field p̂h = φh1 +φh4 in a dual crossing the high permeability channel at a/b =
1/1000. (a) Fine-scale solution, ph , in the dual, rescaled to a unit pressure gradient,
is in transverse equilibrium. (b) Interpolated fine pressure p̂h with reduced boundary
conditions. (c) Interpolated fine pressure p̂h with linear boundary conditions. (d)
Interpolated fine pressure p̂h with hybrid boundary conditions.
y-direction.
Consider a domain with a high permeability channel in the y-direction, small
amplitude random noise, an effective anisotropy a/b = 1/1000, and the boundary
conditions p(x = 0, y) = 1, p(x = 1, y) = 0, and py (x, y = 0, 1) = 0. Figure 4.10a
shows that the fine-scale solution is in transverse equilibrium, and the reduced one
dimensional problem in the x-direction reflects the high permeability channel.
The advantage of the reduced boundary conditions is that they are also solutions
to one dimensional problems, and therefore recognize the high permeability channel.
Figure 4.10b shows that the channel is reflected in the superposition of the local
solutions. The reduced boundary conditions sample different realizations of the per-
meability and are therefore not identical. This introduces finite transverse gradients,
∂ p̂h /∂y = 0, even if the coarse pressure is in transverse equilibrium. Even small
transverse pressure gradients lead to very large fluxes as the effective anisotropy in-
creases, and this leads to the loss of monotonicity of the coarse operator even for
the compact stencil. Eliminating these gradients is therefore the key to a monotone
coarse operator and a robust MSFV method for heterogeneous anisotropic problems.
Linear boundary conditions are identical by construction and do not introduce
artificial transverse gradients. Therefore, linear boundary conditions give rise to a
CHAPTER 4. MULTISCALE SIMULATION 152
more robust coarse scale operator, but figure 4.10c shows that they give poor accu-
racy. In this case the linear boundary conditions have a large error. For isotropic
problems this error would be localized in a narrow boundary layer, but for high effec-
tive anisotropy the error spreads through the entire dual interaction region. Linear
boundary conditions therefore give a robust operator, but they may lead to poor
accuracy.
for linear gradient, Φh3 , in the y-direction. We can use these two solutions to define
the boundary conditions for a bilinear local problem Φh4
Together with the constraint, 0 = T1, these three problems determine the coarse
operator uniquely. The standard set of bilinear basis functions can be obtained from
Φh2 to Φh4 through linear superposition. These bilinear basis functions have hybrid
boundary conditions and hence the transmissibility matrix T and the coarse operator
LH are modified. This modification is independent from the implementation of the
compact coarse operator.
The local problems defined above attain transverse equilibrium at high anisotropy
by construction (figure 4.10d ). Transverse gradients in the local problems are re-
duced faster than in the global fine-scale solution, because the domain is smaller and
therefore boundary layers merge earlier. Therefore, hybrid boundary conditions have
robustness similar to linear boundary conditions. Section 4.11 shows that the accu-
racy of the hybrid boundary conditions is similar to that of the reduced boundary
conditions at low anisotropy. At high anisotropy the accuracy of the hybrid boundary
conditions is good if
ay in the dual interaction region is a good representation of
Γ fH = -1
1
a) b) c) d)
0 fH = 1 1
Figure 4.11: (a) The unit square domain Ω, with a 20 × 20 coarse grid. The back
squares indicate coarse blocks with sources fH = ±1. In these blocks the fine-scale
permeability is constant at the mean to give a uniform distribution of the fine-scale
flux. (b) A realization of the permeability field with small correlation length. Figures
c and d show realizations of the layered permeability fields at angles of 15◦ and −30◦
respectively. The logarithm of the permeability is shown in all cases.
a) fs b) fs c) fs
Figure 4.12: Comparison of time of flight maps between multiscale methods and
the fine-scale solution for three isotropic but heterogeneous fine-scale permeability
fields. All solutions are shown at the same injected volume, corresponding to the
breakthrough of the fine-scale solution. The fine-scale solution, fs, is shown in the first
row, a to c, the original MSFV -red method in the second row, d to f, the CMSFV -lin
method in the third row, g to i, and the CMSFV -hyb method in the fourth row, j to
l. The first column shows solutions for a realization of the permeability field shown
in figure 4.11b, columns 2 to 3 show solutions for realizations of layered permeability
fields at 15◦ and -30◦ (figure 4.11c and d).
CHAPTER 4. MULTISCALE SIMULATION 156
a) fs b) fs c) fs
Figure 4.13: Comparison of time of flight maps between multiscale methods and
the fine-scale solution for three heterogeneous fine-scale permeability fields with grid
aligned fine-scale anisotropy, a/b = 0.01. All solutions are shown at the same injected
volume, corresponding to the breakthrough of the fine-scale solution. The figure
layout is the same as in figure 4.12.
CHAPTER 4. MULTISCALE SIMULATION 157
0.15
||uh ||2
||ph ||2
0.2
0.1
0.1
0.05
0 0
-45 -30 -15 0 15 30 45 -45 -30 -15 0 15 30 45
θ θ
0.2 3
c)
2.5
d)
0.15
2
||ūh −uh ||2
||p̂h −ph ||2
||uh ||2
||ph ||2
0.1 1.5
1
0.05
0.5
0 0
-45 -30 -15 0 15 30 45 -45 -30 -15 0 15 30 45
θ θ
Figure 4.14: Figures a and b show the mean L2 -errors of the interpolated solution
p̂h and the reconstructed flux field ūh = −k∇p̄h as function of the orientation, θ, of
the fine-scale layering for isotropic fine-scale permeability, a/b = 1. Figures c and d
show the same for grid aligned fine-scale anisotropy a/b = 0.04 (∆x/∆y = 5). All
values shown are means over 20 realizations and the standard deviations are shown
as vertical bars, if they are larger than the symbols.
CHAPTER 4. MULTISCALE SIMULATION 159
detail of the tracer distribution. In the presence of fine-scale anisotropy the CMSFV-
hyb solution is a good approximation to the fine-scale solution, while the MSFV-red
solution has large errors and unphysical behavior.
The linear boundary conditions give comparable results for isotropic problems,
but lose accuracy as the aspect ratio increases. For flow across layers the errors are
largest (figure 4.13i ), as expected from the discussion in §4.6.2.
Error norms for the pressure and the reconstructed flux obtained by both methods
are shown in figure 4.14. The interpolated pressure p̂h is proportional to the error
in the coarse solution, while the reconstructed, fine-scale flux ūh is proportional to
the error in the reconstructed pressure p̄h . Figure 4.14a shows that linear boundary
conditions lead to larger errors than reduced or hybrid boundary conditions for both
coarse operators. For clarity, the results for linear boundary conditions are omitted
from the other graphs. Figures 4.14a and b show that for isotropic fine-scale perme-
ability fields the errors in the MSFV-red and the CMSFV-hyb solutions are of the
same order of magnitude for flow along layers (0◦ < θ < 45◦ ). For flow across layers
(−45◦ < θ < 0◦ ) the error in the MSFV-red solution is significantly smaller, because
the reduced boundary condition is optimal for this special case. However, for any sig-
nificant anisotropy on the fine-scale the CMSFV-hyb solution becomes more accurate
than the MSFV-red for all angles (figure 4.14c and d ).
Figures 4.15a and b show that for both the MSFV-red and the CMSFV-red
solutions the errors in the flux, ūh , grow exponentially with increasing fine-scale
anisotropy, while the errors for both the MSFV-hyb and the CMSFV-hyb remain
bounded. Therefore, in problems where the fine solution approaches transverse equi-
librium, the hybrid local boundary condition rather than the compact operator lead
to a monotone multiscale solution.
Consider the domain shown in figure 4.11a with homogeneous Dirichlet bound-
ary conditions, so that the solution cannot approach transverse equilibrium with
increasing anisotropy. Figure 4.16 shows that both the MSFV-lin and the MSFV-hyb
solution have lost monotonicity and the streamlines show unphysical recirculation
cells. The CMSFV-hyb solution has remained monotone and captures the fine-scale
transport. Similar behavior has been observed in §4.3.1 for homogeneous problems,
CHAPTER 4. MULTISCALE SIMULATION 160
2
0.4 10
a) MSFV-red b)
MSFV-hyb
0.3 CMSFV-red
1
10
||uh ||2
||ph ||2
0.2
0
10
0.1
-1
0 10
-4 -3 -2 -1 0 -4 -3 -2 -1 0
10 10 10 10 10 10 10 10 10 10
a a
b b
Figure 4.15: Figures a and b show the mean L2 -errors of the interpolated solution
p̂h and the reconstructed flux field as function of increasing fine-scale anisotropy a/b.
All values shown are means over 20 realizations of the permeability field shown in
figure 4.11a.
where all local boundary conditions become linear. Therefore, the compact operator
is necessary to obtain a monotone solution for strongly constrained pressure fields.
A similar situation is shown in figure 4.4c where a local maximum occurs in the
coarse solution computed by the MSFV-red method even for an isotropic problem
with Dirichlet constraints on only two boundaries. The compact operator gives a
monotone coarse solution in this case, demonstrating that it improves the monotonic-
ity even for isotropic problems with few constraints. Such situations arise commonly
in reservoir engineering in areas with horizontal wells that constrain the pressure over
many gridblocks.
In cases with or without transverse equilibrium the compact operator leads to a
linear system, LH , much closer to an M -matrix. A simple measure M (LH )i given by
by the ratio of the largest and smallest off-diagonal entries,
max( LHi,j=i
M (LHi,j )i = − , (4.65)
min( LHi,j=i
−4
x 10
1 3 1
(a) (b) (c)
1.5
p̂h
0
y
y
−1.5
0 −3 0
−4
x 10
1 3 1
(d) (e) (f)
1.5
p̂h
0
y
y
−1.5
0 −3 0
−4
x 10
1 3 1
(g) (h) (i)
1.5
p̂h
0
y
−1.5
0 −3 0
0 1 0 1 0 1
x x x
Figure 4.16: Results for the quarter 5-spot shown in figure 4.11a with homogeneous
Dirichlet boundary conditions. The first column shows contour maps of p̂h , the second
column shows a horizontal view of p̂h along the y-axis, and the third column shows
streamlines obtained from the conservative reconstruction p̄h . Figures a to c show
the results for MSFV-lin, figures d to f show results for MSFV-hyb, and figures g to
i show results for CMSFV-hyb.
CHAPTER 4. MULTISCALE SIMULATION 162
in the i-th row of the system matrix LH . If M (LH )i is non-positive, then the i-th row
of LH satisfies the M -matrix condition. As the magnitude of M (LH )i increases the
violation of the M -matrix condition becomes stronger. The norm ||M (LH )|| allows
us to quantify the quality of the coarse linear system resulting from a multiscale
discretization.
For heterogeneous problems LH does generally not satisfy the M-matrix property,
for any choice of operator and local boundary conditions, but the compact operator
always gives a system much closer to an M -matrix. The violations of the M -matrix
property are larger in magnitude (figure 4.17a and c) and much more common (fig-
ure 4.17b and d ) in the coarse linear system arising from the original coarse operator.
For isotropic fine-scale permeability fields the dependence of ||M (LH )|| on the ori-
entation of the layering, shown in figures 4.14a&b, is analogous to the homogeneous
analysis in §4.4. For the original operator the violation of the M -matrix property
is largest for grid aligned layering and smallest for layering oriented at 45◦ , corre-
sponding to the region of monotonicity shown in figure 4.7a, which is most limited
for anisotropy at 0◦ and unlimited for anisotropy at 45◦ to the grid. The compact
operator on the other hand produces linear systems with a maximum in ||M (LH )|| for
layering oriented between 15◦ and 30◦ , corresponding to the region of monotonicity
shown in figure 4.7b, which is unlimited for anisotropy at 0◦ and 45◦ . The improve-
ment in the quality of LH relative to the original operator is largest for layering that
is aligned, or close to aligned, with the grid. Comparing the first and second row of
figure 4.17, we see the that the violations of the M -matrix criterion produced by the
MSFV operator increase rapidly with increasing fine-scale anisotropy, a/b → 0, while
the CMSFV-hyb remains close to a M -matrix.
4.8 Discussion
In the formulation presented in §4.2 the MSFV method is defined by the properties
of the transmissibility matrix, T∗ . The integrated form of T∗ directly defines the
coarse operator through (4.16) to (4.23) and the reconstruction of the conservative
fine-scale velocity is defined by T∗ through (4.32). The properties of T∗ are therefore
CHAPTER 4. MULTISCALE SIMULATION 163
MSFV-red
2 MSFV-lin 0.02
a) MSFV-hyb b)
CMSFV-red
0.015
CMSFV-lin
||M(LH)||∞
||M(LH)||2
CMSFV-hyb
1 0.01
0.005
0 0
-45 -30 -15 0 15 30 45 -45 -30 -15 0 15 30 45
θ θ
15 0.06
0.05
c) d)
||M(LH)||∞
10 0.04
||M(LH)||2
0.03
5 0.02
0.01
0 0
-45 -30 -15 0 15 30 45 -45 -30 -15 0 15 30 45
θ θ
Figure 4.17: The L∞ and L2 -norms of M (LH ) are shown as a function of the angle
of the layering. Figures a and b show results for a isotropic fine-scale permeability
field (a/b = 1), while figures c and d are for a fine-scale anisotropy a/b = 1/100.
CHAPTER 4. MULTISCALE SIMULATION 164
the starting point for the analysis of MSFV formulations and T∗ provides a direct
link to MPFA methods. This link has made it possible to take advantage of the
existing analysis of MPFA methods and to formulate the compact coarse operator
for the MSFV method. The construction of many other coarse operators is possible
in this framework, and as long as the modification is done at the level of T∗ , the
resulting MSFV method will be: 1) conservative on the coarse level; 2) the fine scale
reconstruction of the fluxes will be locally conservative; 3) the fine-scale fluxes will
be consistent with the coarse fluxes.
The CMSFV method described here has been designed to give a coarse operator
with optimal monotonicity, but other design criteria, e.g. rotational invariance of the
discretization, may lead to other preferred coarse operators. The formulation given
here provides a consistent framework for the design of coarse operators by modifying
T∗ . In a parallel Aavatsmark et al. (2008) have developed a compact MPFA method
with improved monotonicity for non-orthogonal quadrilateral grids with piecewise
constant, full-tensor permeability fields.
Our framework emphasizes that local elliptic solutions determine the numerical
entries of T∗ , rather than their role as basis functions used for the interpolation of
the coarse solution. The concept of the basis function has been important in the
development of the MSFV method, but the formation of the coarse operator and the
fine-scale reconstruction of the conservative velocity field do not require explicit basis
functions. In the framework presented here coarse operators, which are not naturally
associated with a set of basis functions, can be considered, e.g. the compact operator.
In these cases the interpolation of the coarse solution with the standard basis functions
still gives a good approximation of the fine-scale solution, but the corresponding fine
fluxes are not consistent with the coarse fluxes. This is generally not a problem since
the fluxes derived from the interpolated solution are not used for transport.
The original bilinear basis functions require Dirichlet boundary conditions, which
are too restrictive for many anisotropic problems. Once the local elliptic problems
are not interpreted as basis functions, a larger range of local problems and local
boundary conditions can be considered. This allows the design of local problems
that take advantage of simplifications in the fine-scale solution, such as the linear
CHAPTER 4. MULTISCALE SIMULATION 165
flow problems with hybrid boundary conditions that are designed to capture TVE.
Lunati & Jenny (2007) modify the boundary condition for the flux reconstruction,
L̃Γh , to eliminate recirculation cells in the velocity field. Here we modify the boundary
condition of the local elliptic problems, L̄Γh , and hence the coarse operator to achieve
a monotone coarse pressure, which eliminates the unphysical recirculation cells.
The CMSFV-hyb method introduced here is only based on local information, be-
cause local methods are computationally efficient. It has been shown that the accu-
racy of purely local methods may be low if the permeability field has structures with
very long correlation lengths. Several authors have suggested the incorporation of
global data to increase accuracy at increased cost (Chen & Durlofsky, 2006; Efendiev
et al., 2006; Durlofsky et al., 2007). The coarse operator introduced here may also
increase the robustness of these applications.
Most recent studies of multiscale methods for flow in porous media have been
restricted either to tracer flow or to immiscible two phase flow (Chen & Hou, 2002;
Jenny et al., 2003, 2005; Aarnes, 2004; Juanes & Dub, 2008). In these problems only
the flux field is necessary for the solution of the transport problem, while the quality
of the solution itself - the pressure field - is not important. In many applications,
such as the three-phase black-oil formulation recently presented by Lee et al. (2008),
mass transfer between phases is an important physical mechanism. In particular
strongly non-linear phenomena such as phase appearance and disappearance are a
strong function of the pressure. For these applications the increased robustness of the
CMSFV-hyb method for the pressure equation will increase the stability of the coupled
problem. In these cases the pressure interpolation, p̂h , may be more appropriate for
the evaluation of the phase properties.
In the general heterogeneous anisotropic case neither coarse operator leads to a
linear system where LH is an M-matrix. However, compared to the original coarse
operator the compact operator leads to a LH with fewer positive off-diagonal entries
and these entries have a smaller magnitude. The performance of algebraic multigrid
solvers will not deteriorate seriously, unless the off-diagonal entries are substantial
(Ruge & Stüben, 1987). Therefore, we expect that the CMSFV method will avoid
problems with the performance of the linear solver.
CHAPTER 4. MULTISCALE SIMULATION 166
4.9 Conclusion
The construction of the MSFV coarse operator is identical to the construction of a
certain type of MPFA methods. The MSFV method is a natural extension of these
MPFA methods to numerical evaluation of the transmissibility matrix. Therefore,
the rich literature on the analysis of MPFA methods is directly relevant to the design
and analysis of MSFV methods.
The original MSFV coarse operator has a limited region of monotonicity for
aligned anisotropy. This leads to loss of monotonicity of the coarse pressure in prob-
lems with Dirichlet boundary conditions with increasing anisotropy.
A compact coarse operator for the MSFV method that reduces to a 7-point stencil
in the homogeneous case has been introduced. This compact operator has the largest
possible monotonicity region. In particular it is always monotone for grid aligned
anisotropy in homogeneous problems. In heterogeneous problems the compact opera-
tor leads to coarse linear systems that are close to a M -matrix. The compact operator
eliminates oscillations in anisotropic problems with Dirichlet boundary conditions.
For anisotropic problems with homogeneous Neumann boundary conditions the
fine-scale solution approaches transverse equilibrium as the anisotropy increases. To
obtain a monotone, accurate, multiscale solution the local boundary conditions must
be able to reach transverse equilibrium as well. A hybrid boundary condition has been
introduced that naturally allows transverse equilibrium and gives accurate solutions
for anisotropic problems.
The compact coarse operator in combination with the hybrid boundary condition
is the most robust multiscale finite volume formulation for large anisotropy and similar
in accuracy to the original MSFV method for isotropic problems.
Chapter 5
Conclusions
Saline aquifer storage of CO2 has the potential to reduce the emissions from fossil
energy significantly. The leakage of CO2 from the storage aquifer back into the
atmosphere is a main concern in developing this technology. Carbon dioxide can leak
from the geological storage formation back into the atmosphere, because it is mobile
and buoyant relative to the resident brine. Therefore, any process that immobilizes
the injected CO2 , makes it negatively buoyant, or both, is considered a trapping
mechanism. The competition between leakage and trapping processes will determine
what fraction of the injected CO2 will remain in the subsurface permanently. While
mineral trapping offers permanent storage, leakage is unlikely after the CO2 has
dissolved into the brine or has become disconnected at the pore-scale, as a residual
saturation. Consequently dissolution and residual trapping determine the length of
the active storage period, during which leakage is possible.
5.1 Results
We studied dissolution and residual trapping using two different approaches. In order
to resolve the small length and time scales associated with buoyancy driven miscible
convection, we employed an innovative linear stability analysis in self-similar coordi-
nates, and we performed high accuracy nonlinear simulations to study the nonlinear
evolution of the instability.
167
CHAPTER 5. CONCLUSIONS 168
To study the migration and residual trapping of immiscible CO2 plumes in the
post-injection period, we employed vertical-equilibrium sharp-interface models, in
which the dimensionality of the problem is reduced in favor of the ability to focus on
the first-order behavior of the gravity current.
The dominant physical and chemical processes during CO2 storage in saline aquifers
change with time, and natural geologic formations are strongly heterogeneous. Pre-
dictions of the flow and transport behaviors for a particular storage site must account
for the presence of heterogeneity and be able to model the wide range of length and
time scales associated with the physical and chemical processes. To make numerical
modeling of such systems possible, we extend the multiscale finite volume (MSFV)
method to highly anisotropic problems. This will allow the robust numerical modeling
of CO2 storage in saline aquifers that commonly have large aspect ratios.
for CO2 storage, because they have relatively short active storage periods and large
storage capacities.
Multiscale simulation of CO2 storage The MSFV method has been extended to
include all the necessary physics: gravity and capillary forces, compressibility, wells,
and three-phase problems. We have extended it to problems with severe anisotropy,
expected in regional aquifers. Adaptivity criteria have been developed for the pressure
solver as well as for the solution of the transport equation promising large speed-
up compared to conventional simulations. We therefore plan to demonstrate the
application of the MSFV method to saline aquifer storage in the near future.
BIBLIOGRAPHY 172
Bibliography
Aarnes, J. E. 2004 On the use of a mixed multiscale finite element method for
greater flexibility and increased speed or improved accuracy in reservoir simulation.
Multiscale Model. Simul. 2 (3), 421–439.
Adams, J. J. & Bachu, S. 2002 Equations of state for basin geofluids: algorithm
review and intercomparison for brines. Geofluids 2, 257271.
Bear, J. 1972 Dynamics of Fluids in Porous Media. American Elsevier Pub. Co.
Ben, Y., Demekhin, E. A. & Chang, H.-C. 2002 A spectral theory for small-
amplitude miscible fingering. Phys. Fluids 14 (3), 999.
Benson, S. M., Tomutsa, L., Silin, D. & Kneafsey, T. 2006 Core scale and
pore scale studies of carbon dioxide migration in saline formations. In Proceedings
GHGT-8 . Trondheim, Norway.
Bickle, M., Chadwick, A., Huppert, H. E., Hallworth, M. & Lyle, S. 2007
Modelling carbon dioxide accumulation at Sleipner: Implications for underground
carbon storage. Earth Planet. Sc. Lett. 255, 164–176.
van der Burgt, M. J., Cantle, J. & Boutkan, V. K. 1992 Carbon dioxide
disposal from coal-based IGCC’s in depleted gas fields. Energy Convers. Manage.
33 (5-8), 603–610.
Chen, Z. & Hou, T. Y. 2002 A mixed multiscale finite element method for elliptic
problems with oscillating coefficients. Math. Comp. 72 (242), 541–576.
Coats, K. H., Dempsey, J. R. & Henderson, J. H. 1971 The use of vertical equi-
librium in two-dimensional simulation of three-dimensional reservoir performance.
Soc. Petrol. Eng. J. 251 (March), 63–71.
BIBLIOGRAPHY 175
Ebigbo, A., Class, H. & Helmig, R. 2006 CO2 leakage through an abandoned
well: problem-oriented benchmarks. Computat. Geosci. 11 (2), 103–115.
Efendiev, Y., Ginting, V., Hou, T. & Ewing, R. 2006 Accurate multiscale
finite element methods for two-phase flow simulations. J. Comput. Phys. 220,
155174.
Elder, J. W. 1968 The unstable thermal interface. J. Fluid Mech. 32, 69–96.
BIBLIOGRAPHY 176
Henry, H. R. 1959 Salt intrusion into fresh-water aquifers. J. Geophys. Res. 64,
1911–1919.
Holloway, S & Savage, D. 1993 The potential for aquifer disposal of carbon-
dioxide in the UK. Energy Convers. Manage. 34 (9-11), 925–932.
Hou, T. Y. & Wu, X.-H. 1997 A multiscale finite element method for elliptic
problems in composite materials and porous media. J. Comput. Phys. 134, 169–
189.
Hou, T. Y., Wu, X.-H. & Cai, Z. 1999 Convergence of a multiscale finite ele-
ment method for elliptic problems with rapidly oscillating coefficients. Math. Comp.
68 (227), 913–93.
Ide, S. T., Jessen, K. & F. M. Orr Jr. 2007 Storage of CO2 in saline aquifers:
effects of gravity, viscous, and capillary forces on amount and timing of trapping.
Int. J. Greenh. Gas Control 1 (4), 481–491.
Jenny, P., Lee, S. H. & Tchelepi, H. A. 2005 Adaptive multiscale finite volume
method for multi-phase flow and transport. SIAM Multiscale Model. Simul. 3 (1),
50–64.
Jenny, P., Lee, S. H. & Tchelepi, H. A. 2006 Adaptive fully implicit multi-scale
finite-volume method for multi-phase flow and transport in heterogeneous porous
media. J. Comput. Phys. 217, 627–641.
Jessen, K., Kovscek, A. R. & F. M. Orr Jr., 2005 Increasing CO2 storage in
oil recovery. Energy Conv. Manag. 46 (2), 293–311.
Jhavery, B. S. & Homsy, G. M. 1982 The onset of convection in fluid layer heated
rapidly in a time-dependent manner. J. Fluid Mech. 114, 251.
Journel, A. G. 1986 Geostatistics - models and tools for the Earth-sciences. Math-
ematical Geology 18 (1), 119–140.
Juanes, R. & Dub, F.-X. 2008 A variational multiscale method with locally con-
servative treatment of multiscale source terms. Comput. Geosci., Special Issue on
Multiscale Methods for Flow and Transport in Heterogeneous Porous Media, doi
10.1007/s10596-007-9070-x (In Press) .
BIBLIOGRAPHY 179
Juanes, R., Spiteri, E. J., F. M. Orr Jr. & Blunt, M. J. 2006 Impact
of relative permeability hysteresis on geological CO2 storage. Water Resour. Res.
42 (W12418), 1–13.
Kaarstad, O. 1992 Emission-free fossil energy from norway. Energy Convers. Man-
age. 33 (5-8), 781786.
Katzer, J. 2007 The future of coal, options for a carbon constrained world. Tech.
Rep.. Massachusetts Institute of Technology.
Kippe, V., Aarnes, J. E. & Lie, K.-A. 2007 A comparison of multiscale meth-
ods for elliptic problems in porous media flow. Comput. Geosci., Special Issue on
Multiscale Methods for Flow and Transport in Heterogeneous Porous Media, DOI
10.1007/s10596-007-9074-6 (In Press) .
Koide, H., Shindo, Y., Tazaki, Y., Iijima, M., Ito, K., Kimura, N. &
Omata, K. 1997 Deep sub-seabed disposal of CO2 - the most protective storage.
Energy Convers. Manage. 38, S253–S258.
Koide, H., Tazaki, Y., Noguchi, Y., Nakayama, S., Iijima, M., Ito, K.
& Shindo, Y. 1992 Subterranean containment and long-term storage of carbon
dioxide in unused aquifers and in depleted natural gas reservoirs. Energy Convers.
Manage. 33 (5-8), 619626.
BIBLIOGRAPHY 180
Kovscek, A. R. & Wang, Y. 2005 Geological storage of carbon dioxide and en-
hanced oil recovery I. Uncertainty quantification employing a streamline based
proxy for reservoir simulation. Energy Conv. Manag. 46 (11-12), 1920–1940.
Kumar, A., Ozah, R., Noh, M., Pope, G. A., Bryant, S., Sepehrnoori, K.
& Lake, L. W. 2005 Reservoir simulation of CO2 storage in deep saline aquifers.
Soc. Petrol. Eng. J. pp. 336–348.
on Multiscale Methods for Flow and Transport in Heterogeneous Porous Media, doi
10.1007/s10596-007-9069-3 (In Press) .
Legg, J. F. 1992 Overview of carbon dioxide removal and disposal in canada. Energy
Convers. Manage. 33 (5-8), 787–794.
Lick, W. 1964 The instability of a fluid layer with time-dependent heating. J. Fluid
Mech. 21, 565–576.
Lindeberg, E. & Bergmo, P. 2003 The long-term fate of CO2 injected into an
aquifer. In Proceedings of the 6th International Conference on Greeenhouse Gas
Technologies (ed. J. Gale & Y. Kaya), , vol. 1, pp. 489 – 494. Elsevier Science Ltd.
Lister, J. R. 1992 Viscous flows down an inclined plane from point and line sources.
J. Fluid Mech. 242, 631–653.
Lunati, I. & Jenny, P. 2007 Treating highly anisotropic subsurface flow with the
multiscale finite-volume method. SIAM Multiscale Model. Simul. 6 (1), 308–318.
BIBLIOGRAPHY 182
Lyle, S., Huppert, H. E., Hallworth, M., Bickle, M. & Chadwick, A. 2005
Axisymmetric gravity currents in a porous medium. J. Fluid Mech. 543, 293–302.
McGrail, B. P., Schaef, H. T., Ho, A. M., Chien, Y.-J., Dooley, J. J. &
Davidson, C. L. 2006 Potential for carbon dioxide sequestration in flood basalts.
J. Geophys. Res. 111 (B12), B12201.
van der Meer, L. G. H. 1992 Investigations regarding the storage of cabon dioxide
in aquifers in the Netherlands. Energy Convers. Manage. 33 (5-8), 611–618.
Metz, B., Davidson, O., de Coninck, H., Loos, M. & Meyer, L., ed.
2006 Special Report on Carbon Dioxide Capture and Storage. Cambridge University
Press.
Mo, S. & Akervoll, I. 2005 Modeling long-term CO2 storage in aquifer with a
black-oil reservoir simulator. In SPE/EPA/DOE Exploration and Production En-
vironmental Conference. Galveston, TX.
Mo, S., Zweigel, P., Lindeberg, E. & Akervoll, I. 2005 Effect of geologic
parameters on CO2 storage in deep saline aquifers. In SPE Eurospec/EAGE Annual
Conference. Madrid, Spain.
BIBLIOGRAPHY 183
Nield, D. A. & Bejan, A. 1999 Convection in porous media, 2nd edn. Springer
Verlag.
Oldenburg, C. M. & Unger, A. J. A> 2005 Coupled vadose zone and at-
mospheric surface-layer transport of carbon dioxide from geologic carbon seques-
tration sites. Vadose Zone J. 3 (3), 848–857.
Pritchard, D. 2004 The instability of thermal and fluid fronts during radial injec-
tion in porous media. J. Fluid Mech. 508, 133.
Pruess, K. 2005 Numerical studies of fluid leakage from a geologic disposal reservoir
for CO2 show self-limiting feedback between fluid flow and heat transfer. Geophys-
ical Research Letters 32 (L14404), 1–4.
Reed, A. C., Mathews, J. L., Bruno, M. S. & Olmstead, S. E. 2002 Safe dis-
posal of one million barrels of NORM in Louisiana through slurry fracture injection.
SPE Drill. Completion 17 (2), 72–81.
Riaz, A. & Meiburg, E. 2003a Radial source flows in porous media: Linear sta-
bility analysis of axial and helical perturbations in miscible displacements. Phys.
Fluids 15 (4), 938.
Ruge, J. W. & Stüben, K. 1987 Multigrid Methods, chap. 4, pp. 73–130. SIAM.
Stevens, S. H., Kuuskra, V. A., Gale, J. & Beecy, D. 2001 CO2 injection
and sequestration in depleted oil and gas fields and deep coal seams: worldwide
potential and costs. Environmental Geosciences 8 (3), 200–209.
Voss, C. I. & Sousa, W. R. 1987 Variable density flow and solute transport
simulation of regional aquifers containing a narrow freshwater-saltwater transition
zone. Water Resour. Res. 28, 1851.
Xu, T., Apps, J. A. & Pruess, K. 2003 Reactive geochemical transport simulation
to study mineral trapping for CO2 disposal in deep arenaceous formations. Journal
of Geophysical Research 108 (doi:10.1029/2002JB001979,).
Yang, C. & Gu, Y. 2006 Accelerated mass transfer of CO2 in reservoir brine due
to density-driven natural convection at high pressures and elevated temperatures.
Ind. Eng. Chem. Res. 45, 2430–2436.
Zhou, H. & Tchelepi, H. A. 2008 Operator based multiscale method for com-
pressible flow. Soc. Petrol. Eng. J. (June), pp. 1–7.
Zhou, Q. L., Bear, J. & Bensabat, J. 2003 Saltwater upconing and decay
beneath a well pumping above an interface zone. Transport Porous Med 61 (3),
337–363.