Sie sind auf Seite 1von 9

PROCEEDINGS, TOUGH Symposium 2012

Lawrence Berkeley National Laboratory, Berkeley, California, September 17-19, 2012

DEVELOPMENT AND APPLICATION OF A CHEMICAL OSMOSIS SIMULATOR


BASED ON TOUGH2
M. Takeda1*, T. Hiratsuka1, K. Ito1, S. Finsterle2
1

National Institute of Advanced Industrial Science and Technology (AIST)


Higashi 1-1-1, Central 7
Tsukuba, Ibaraki 305-8567, Japan
e-mail: mikio-takeda@aist.go.jp
2

Lawrence Berkeley National Laboratory, Earth Sciences Division


1 Cyclotron Rd., MS 74-0120
Berkeley, CA 94720, United States
e-mail: safinsterle@lbl.gov
thermal, hydrological, mechanical, and chemical
(T-H-M-C) processes, have been one of the
main research subjects in geological disposal
projects. However, a different type of H-C
coupled phenomenon caused by the membrane
property of clayey rock has recently received
attention in the context of geological disposal of
radioactive waste (e.g., Soler, 2001; Gonalvs
et al., 2004; Garavito et al., 2007; Horseman et
al., 2007; Cruchaudet et al., 2008; RousseauGueutin et al., 2009).

ABSTRACT
One of the key issues for radioactive waste
disposal in clayey formations is uncertainty with
respect to fluid pressure anomalies, which could
be caused by chemical osmosis. For the examination of chemical osmosis in groundwater
systems, a 3D, site-scale simulator with chemical osmosis simulation capabilities is needed. In
this study, the TOUGH2 code with the EWASG
module (hereafter referred to as the T2+EWASG
code) was modified to simulate chemical osmosis in clayey geologic media. In the modified
code, chemical osmosis is treated as a diffusion
of water rather than the advective flux of a
solution (consisting of water and salt), which has
been the conventional approach in chemical
osmosis modeling. The modified code was
tested against a semi-analytical solution and a
laboratory experiment of chemical osmosis. The
simulation results agree well with both the
experimental data and the semi-analytical solution, and attest to the practical effectiveness of
the chemical osmosis simulation capabilities in
the modified T2+EWASG code.

The membrane property of clay works to partially restrict the passage of ions while allowing
water to migrate into the clayey materials (Fritz,
1986). This causes a coupled flow phenomenon,
termed chemical osmosis, in which the gradient
of solute concentration causes water migration.
Strictly speaking, chemical osmosis is water
migration through a semi-permeable membrane
driven by the difference in the chemical potential of water across the membrane (Mitchell and
Soga, 2005). The net migration of water continues until the difference in potential energy is
compensated by the other potential differences,
such as the pressure difference. Accordingly, if
geologic media can act as semi-permeable
membranes, and if the salt concentration is not
uniform in the formation, localized, erratic fluid
pressures may be generated by chemical osmosis
(Neuzil, 2000). Ignoring osmotically induced
pressures may lead to erroneous predictions of
groundwater flow direction (Marine and Fritz,
1981). Therefore, a numerical code capable of
simulating chemical osmosis must be developed.
Simulation codes for chemical osmosis have so
far been developed for the interpretation of

INTRODUCTION
Clayey rocks are potential host rocks for the
geological disposal of radioactive wastes, due to
their low hydraulic conductivities and efficient
retention properties. For the long-term stable
isolation of radioactive wastes, the processes
relevant to the migration of radionuclides
through groundwater need to be identified and
incorporated into the radionuclide transport
model for safety assessment. Interactions between different processes, such as coupled
-1-

chemical osmosis experiments in the laboratory


and the field; however, these codes are mostly
based on 1D linear or radial models (Gonalvs
et al., 2004; Bader and Kooi, 2005; Garavito et
al., 2006, 2007; Gueutin et al., 2007; Horseman
et al., 2007; Gonalvs, 2008; Cruchaudet et al.,
2008; Rousseau-Gueutin et al., 2008, 2009). In
order to develop a chemical osmosis simulator
capable of using 3D, site-scale models based on
the T2+EWASG code, we revisit the function of
clay as a selective permeable membrane in mass
transport through clayey geologic media and the
conventional mathematical formulation of the
coupled flow phenomenon caused by the membrane effect. We then describe the mathematical
formulation of the coupled flow phenomenon for
the modification of the T2+EWASG code. The
modified simulation code is then tested against a
laboratory experiment of chemical osmosis
through a mudstone rock sample, and verified
against a semi-analytical solution derived from a
conventional mathematical formulation.

However, the clayey media also contain large


pores and minerals without surface charges, in
which the EDLs do not overlap or do not exist,
resulting in the preferential migration of charged
ions through large pores with less electrical
repulsion force. Accordingly, the clayey media
act as imperfect membranes.

MATHEMATICAL FORMULATION OF
CHEMICAL OSMOSIS

The membranes ability to retard the transport of


solutes relative to water is represented by a
reflection coefficient or osmotic efficiency (e.g.,
Neuzil and Provost, 2009). The relationship
between the reflection coefficient and the solute
selective passage through a semi-permeable
membrane is expressed as (Katchalsky and
Curran, 1965):

Figure 1. Membrane effects of clayey minerals on


mass transport in pore spaces.

Membrane Effect on Mass Transport


The membrane property of clay causes two
major processes in mass transport through pore
spaces. One process is chemical osmosis, which
is the diffusion of water through pores driven by
the gradient of the chemical potential of water.
The other process is the filtration of solutes at
the pore entrance, by which the advective and
diffusive solute migrations are partially restricted (Ingebritsen, 2006).

vs
vw

= 1

(1)

=0

where vs and vw are the velocities of solute and


water, respectively, represents the difference
in water osmotic pressures between solutions
sandwiching the membrane, and is the reflection coefficient, which ranges in value between
0, indicating no retardation, and 1, indicating a
perfect semi-permeable membrane. For an ideal
membrane of =1, vs becomes zero, meaning
the solute is completely repelled by the membrane. As mentioned above, the extent of EDL
overlap determines the filtration effect in the
clayey media, and hence, the reflection coefficient depends on both the geological medium
and solution properties (e.g., Bresler, 1973).

These clay membrane effects arise from the


existence of electrical diffuse layers (EDLs)
formed between pore spaces with negative
charges on the surfaces, as shown in Figure 1
(e.g., Appelo and Postma, 2005). The negative
electrical potentials formed between pore spaces
generate an electrical repulsive force on anions
at the pore entrance. The migration of cations is
also hindered at the pore entrance since the
solution must remain electrically neutral at the
entrance. Consequently, neutral components,
such as water molecules and uncharged species,
can migrate through pore spaces without electrical restrictions. Because of this process, clays
behave like selectively permeable membranes.
-2-

The filtration effect on the advective mass flux


of salt, F|adv, is expressed as:

Conventional Mathematical Formulation


Chemical osmosis is essentially the diffusion of
water molecules through a semi-permeable
membrane. Chemical osmosis has been conventionally described as an advective flux of a
solution consisting of water and salt (e.g., Bader
and Kooi, 2005):

Similarly, the diffusive mass flux of salt, F|dif,


under the influence of the filtration effect, can be
expressed as:

k
q = ( P + g z + )

F dif = (1 ) D X

F adv = (1 ) X q

(2)

1
w
Vw

( X ) + F = 0
t

F = F adv + F dif

(9)

In the modeling of chemical osmosis, the medium and solution parameters, such as k, , , and
D, are usually treated as constants. The reflection coefficient is also treated as a constant
parameter (Soler, 2001; Gonalvs et al., 2004;
Bader and Kooi, 2005; Horseman et al., 2007;
Gonalvs, 2008), or as a function of properties
of the medium and solution (Garavito et al.,
2006, 2007; Cruchaudet et al., 2008; Gueutin et
al., 2007; Rousseau-Gueutin et al., 2008, 2009).
In the latter case, the reflection coefficient is
often represented by Breslers (1973) model, in
which is obtained based on a graphical approximation of Breslers curve of , as described by Neuzil and Provost (2008).

(3)

X (4)
Ms
where is the number of constituent ions of the
dissociating salt, R is the universal gas constant,
T is the temperature, C is the salt concentration,
Ms is the molar mass of salt, and X is the mass
fraction of salt. Using Equation (2), the massbalance equation for solution is generally expressed as (e.g., Bader and Kooi, 2005):

( ) + ( q) = 0
t

(8)

where F is the mass flux of salt and is given by:

where Vw is the molar volume of water and w is


the chemical potential of water. Conventionally,
the gradient of osmotic pressure given by Equation (3) is approximated using the concentration
or the mass fraction of salt (e.g., Fritz, 1986;
Gonalvs et al., 2004; Bader and Kooi, 2005):

R T C or R T

(7)

where D is the diffusion coefficient of salt. The


salt mass-balance equation is expressed as (e.g.,
Bader and Kooi, 2005):

where q is the volumetric flux of the solution, k


is the intrinsic permeability, is the dynamic
viscosity of solution, P and are the pressure
and density of the solution, g is the gravitational
acceleration, and z is the elevation. In Equation
(2), chemical osmosis is expressed as if it were
driven by the gradient of osmotic pressure;
however, the gradient of osmotic pressure is a
measure of the gradient of the chemical potential
in water (e.g., Spiegler and Kedem, 1966):

(6)

Governing Equations in TOUGH2


The balance equations solved by TOUGH2 can
be written in the following general form (Pruess,
1991):
d
M dVn = F ndn + Q dVn
dt V

(5)

where t is the time and is the porosity. The


third term on the right-hand side of Equation (2)
corresponds to the volumetric solution flux due
to chemical osmosis, and the second term on the
left-hand side of Equation (5) represents the
solution mass transport, which implicitly describes the salt as also transported along with
water by chemical osmosis.

(10)

The quantity M on the left-hand side represents


mass or energy per volume, with =1, , NK
corresponding to the mass components, and
=NK+1 denoting heat. F denotes mass or heat
flux, and Q denotes sinks and sources. n is a
normal vector on surface element dn, pointing
inward into Vn. The total mass, mass flux, and
sink/source of component are obtained by
-3-

summing over fluid phases =1, , NPH. In the


EWASG module formulation, NK=3, NPH=3,
and =1, 2, 3, 4 indicates water, NaCl, a noncondensible gas (NCG) and heat, respectively,
and =1, 2, 3 indicates the gas, liquid, and solid
salt phases, respectively (Battistelli et al., 1997).
The flux of component is expressed as

F = F

adv

+ F

dis

APPLICATION TO A CHEMICAL
OSMOSIS EXPERIMENT
Coupled Processes in the Chemical Osmosis
Experiment
Figure 2 shows a schematic of coupled processes
in the chemical osmosis experiment. In the
experiment, solutions of different NaCl concentrations are separated by a rock sample. The
concentration difference between solutions
induces chemical osmosis through the rock
sample, from the low-concentration solution to
the high-concentration solution. The migration
of water increases the pressure of the highconcentration solution. The pressure difference
evolving between the two solutions causes an
advective flow of the solutions that counter
chemical osmosis. The solute also diffuses
through the rock sample toward the lowconcentration solution. Although the solute
migration due to advection and diffusion is
partially hindered by the filtration effect of the
rock sample, the solute concentration eventually
becomes uniform throughout the system. The
pressure difference evolving between solutions
dissipates as the concentration becomes uniform.
In the actual experiment, the pressure of the lowconcentration solution was kept constant, and
the pressure of the high-concentration solution
was allowed to change.

(11)

where F|adv is the advective mass flux and F|dis


is the dispersive and/or diffusive mass flux.
In the governing equations of the T2+EWASG
code, water and salt are treated as components.
This enables explicit description of chemical
osmosis as a diffusion process of water in the
mass flux equation. The general forms of
advective and diffusive mass fluxes of component can now be given as:
F
F

adv

= 1 X F

dis

= 1 D X

(12)
(13)

where is the reflection coefficient of component , and is the density of phase . The
advective mass flux of phase is given by:
F = k

k r

(P g )

(14)

Reservoir
- High concentration
solution
- Variable pressure
(isolated)

where kr is the relative permeability to phase ,


is viscosity, and P is the fluid pressure in
phase . Equations (12) and (13) are the generalized forms of mass fluxes, and hence, of
water needs to be assigned as zero in the numerical computation.

Pressure
change

Rock sample

Reservoir
- Low concentration
solution
- Constant pressure

Chemical osmosis
by concentration gradient

Advection
by pressure gradient

Concentration
change

Salt diffusion
by concentration gradient

In the T2+EWASG code, the fluid density and


viscosity, and , are calculated based on the
equation of state, and the reflection coefficient
is assigned as a constant, or is represented by a
graphical approximation of Breslers (1973)
model.

Figure 2. Diagram of the chemical osmosis experiment using a closed system.

Experiment
A chemical osmosis experiment was performed
on a siliceous mudstone, taken at a depth of 982
m from the Wakkanai formation in the
Horonobe research area of Japan, where the
Japan Atomic Energy Agency (JAEA) has been
operating an Underground Research Laboratory
-4-

(URL) for research and development related to


the geological disposal of high-level radioactive
waste (Matsui et al., 2007). A disc-shaped
sample with a diameter of 5 cm and length of
1 cm was taken from the drill core, then immersed in 0.1 mol/L (M) NaCl solution and
vacuumed to remove the air in the pore space.

pump embedded in the bottom reservoir. After


the solution replacement, the cross-over valve of
the bottom reservoir was closed, and the increasing solution pressure in the bottom reservoir was
monitored. The NaCl concentrations of reservoir
solutions were also measured. Both the permeability and chemical osmosis experiments were
conducted at temperature of 313K.

Figure 3 shows the schematic of the experimental system developed for performing permeability and chemical osmosis experiments. The
system mainly consisted of a stainless-steel
pressure vessel, a syringe pump for imposing
confining pressures, two solution reservoirs,
pressure-control equipment for reservoir solutions, a data acquisition system, and a temperature controlled incubator accommodating the
reservoirs and pressure vessel. The solution
reservoirs were symmetrical, and each reservoir
contained a magnetic gear pump for the circulation of reservoir solution, a pressure transducer,
an electric conductivity sensor for concentration
measurement, a porous stone, valves, and a
pressure buffer tank. The stainless-steel pressure
vessel supported hydrostatic pressures of up to
40 MPa, and accommodates a disc-shaped
specimen with a diameter of 5 cm and lengths
ranging from 1 to 2.5 cm. The maximum allowable back pressure for the reservoirs is 0.7 MPa.
The electric conductivity was calibrated against
NaCl solutions of 0.05 to 0.8 M.
The prepared rock sample was first installed in
the pressure vessel, and a confining pressure of
15 MPa was applied to the sample. Each reservoir was filled with 100 mL of 0.1 M NaCl
solution. To remove the entrapped air in the
reservoir tubing, a vacuum was applied to each
reservoir while the magnetic gear pumps circulated the reservoir solutions. After the release of
air entrapped in the reservoir tubing, a back
pressure of 100 kPa was applied to each reservoir. In order to obtain the intrinsic permeability
and pore compressibility of the rock sample, we
performed a permeability experiment before the
chemical osmosis experiment. After the permeability experiment, a back pressure of 200 kPa
was applied to the rock sample and reservoirs.

Figure 3. Schematic of experimental setup for the


chemical osmosis experiment.

Modeling and Results


Figure 4 shows the 1D geometrical model and
the zones representing the reservoirs and rock
sample in the chemical osmosis experiment. The
elements RSVR1 and RSVR2 represent the
bottom and top reservoirs, respectively. The
rock sample is discretized into 190 elements.
The left side of the rock sample is discretized by
100 elements with an element length of 10-5 m,
and the right side of the rock sample is represented by 90 elements with an element length of

The chemical osmosis experiment was initiated


by replacing the bottom reservoir solution with a
0.6 M NaCl solution using the magnetic gear
-5-

10-4 m. The nodal points of the reservoir elements are located at the interfaces between the
reservoir and rock sample. The nodal points of
rock sample elements are located at the center of
each element.

at later time. This may be because the diffusion


of water is accounted for in the T2+EWASG
code, resulting in a dilution effect in the bottom
reservoir caused by water diffusion.
The pressure and NaCl concentration profiles in
the rock sample were also simulated, as shown
in Figure 6. For t 104 s, negative pressures (i.e.,
pressures less than the initial pressure of
3.01105 Pa) evolved in the rock sample. This
can be interpreted as indicating that the gradients
of NaCl concentration in the rock sample are so
large at early times that they induce a water flux
due to chemical osmosis that is larger than the
advection-fueled water flux from the top reservoir. After the NaCl concentration profile
evolved across the rock sample, i.e., at t 105 s,
the solution pressures become positive. At
t = 107 s (about 116 days), the osmotically
induced pressures dissipate, as the NaCl concentration gradient becomes almost uniform in the
rock sample.

The physical properties and simulation parameters assigned to the reservoir and rock sample
elements are summarized in Table 1. The intrinsic permeability and pore compressibility of the
rock sample were determined from the permeability experiment performed before the chemical
osmosis experiment, using iTOUGH2 (Finsterle,
2007). The compressible storage of the bottom
reservoir was also determined by the permeability experiment and is represented by the pore
compressibility of RSVR1. In the top reservoir,
the fluid pressure was kept constant in the
chemical osmosis experiment. In order to keep
the fluid pressure constant in the numerical
computation, the pore compressibility of RSVR2
is assigned a large value. The physical properties
of water and NaCl are summarized in Table 2.
The diffusion coefficient of water is the selfdiffusion coefficient at 313 K. The diffusion
coefficient and the reflection coefficient of NaCl
were selected so that the numerically calculated
pressures and concentrations fit the experimentally obtained pressures and concentrations in
the reservoirs.

Nodal distance
RSVR1

RSVR2

Elements of rock sample


Element of reservoir

Figure 4. Element array and connections in the


TOUGH2 EWASG model.

Figure 5 shows the pressures and NaCl concentrations obtained from the experiment and the
numerical computation using the modified
TOUGH2 EWASG code. The reservoir pressures and concentrations were also simulated by
the semi-analytical solution derived from the
conventional mathematical formulation defined
by Equations (2)(9) (Takeda et al., 2009), using
the same parameters shown in Tables 1 and 2.
The pressures and concentrations simulated by
the T2+EWASG code generally agreed well
with both the experimental data and the simulated curves calculated by the semi-analytical
solution. The simulated bottom reservoir concentrations are slightly higher than the experimental data because the initial concentration of
RSVR1 was selected to match the simulated
curves with the late part of the experimental data.
The bottom reservoir concentrations simulated
by the T2+EWASG code is slightly lower than
those calculated by the semi-analytical solution

Table 1. Physical properties and simulation parameters for 1D chemical osmosis experiment.
Parameter
Intrinsic permeability (m2)
Pore compressibility (Pa-1)
Porosity (-)
Tortuosity (-)
Interface area (m2)
Volume (m3)
Initial pressure (Pa)
Initial NaCl concentration
(mol/L)

Value
RSVR1

Rock

RSVR2

1.0010-13
3.0010-8
1.00
1.00
1.9610-3
2.110-5
3.01105

1.5510-19
1.8810-8
0.28
1.00
1.9610-3
1.9610-5
3.01105

1.0010-13
1.001050
1.00
1.00
1.9610-3
1.0010-4
3.01105

0.55

0.10

0.10

Table 2. Physical properties of water and NaCl.


Parameter
Diffusion coefficient (m2/s)
Reflection coefficient (-)

-6-

Value
water

NaCl

3.2110-9
0.00

1.0110-10
4.2310-2

Reservoir pressure, P (Pa)

105
4.0

CONCLUSION
Bottom reservoir

In order to implement the computational capability for simulating chemical osmosis, this
study outlined the membrane effect of clay and
described the mathematical formulation necessary for the modification of the T2+EWASG
code. Chemical osmosis is incorporated as the
diffusion of water in the water mass flux term,
rather than the advective flux of solution consisting of water and salt, which has been conventionally adopted in chemical osmosis modeling.

Experimental data
3.5

TOUGH2 EWASG
Semi-analytical solution
Top reservoir

3.0

Reservoir NaCl concentration, C (mol/L)

0x100 1x105 2x105 3x105 4x105 5x105 6x105 7x105 8x105


0.6

0.4
0.3

Experimental data
TOUGH2 EWASG
Semi-analytical solution

0.2

Top reservoir

0.1
0
0x100 105 2x105 3x105 4x105 5x105 6x105 7x105 8x105
Time, t (s)

Figure 5.
105
4.0
Pressure, P (Pa)

The modified T2+EWASG code was tested


against a laboratory chemical osmosis experiment and was compared with a semi-analytical
solution derived from the conventional mathematical formulation. The simulation results
agreed well with both the results calculated by
the semi-analytical solution and the experimental data. This agreement verifies the correct
implementation of the equations governing
chemical osmosis and provides confidence that
the relevant processes are captured. It confirms
the chemical osmosis simulation capabilities of
the modified T2+EWASG code.

Bottom reservoir

0.5

Simulated pressures and NaCl concentrations in reservoirs vs experimental data.


t =10 2 (s)
t =10 3 (s)
t =10 4 (s)

In simulating the chemical osmosis experiment,


we selected the physical parameters of both the
rock sample and the solution to match the experimental data. The interpretation of the experiment may be further improved by the use of
iTOUGH2, which can also provide parameter
uncertainties necessary for the field-scale examination of the chemical osmosis effect. The
implementation of the chemical osmosis simulation capabilities into iTOUGH2 is currently
being conducted.

t =10 5 (s)
t =10 6 (s)
t =10 7 (s)

3.5
3.0
2.5
2.0
0x100

NaCl concentration, C (mol/L)

0.6
0.5
0.4

5x10-3
t =10 2 (s)
t =10 3 (s)
t =10 4 (s)

1x10-2
t =10 5 (s)
t =10 6 (s)
t =10 7 (s)

In the modified code, the reflection coefficient


can be assigned as a constant value or can be
treated as a function graphically approximated
from Breslers curve of , as described by
Neuzil and Provost (2008). However, the reflection coefficient can be directly calculated from
Breslers model or other models, which enables
the chemical osmosis simulation under
nonisothermal conditions. The physicochemical
model accounting for the dependency of the
membrane effect on geological and solution
properties will be incorporated into the modified
T2+EWASG code for further study.

0.3
0.2
0.1
0
0x100

Figure 6.

5x10-3
Distance from bottom reservoir, z (m)

10-2

Simulated pressure and NaCl concentration profiles in the rock sample.

-7-

laboratory, Physics and Chemistry of the


Earth, Parts A/B/C, 32, 421433, 2007.
Goncalvs, J., S. Violette, and J. Wendling,
Analytical and numerical solutions for alternative overpressuring processes: Application
to the Callovo-Oxfordian sedimentary in the
Paris basin, France, Journal of Geophysical
Research,
109:
B02110,
doi:
10.1029/2002JB002278, 2004.
Gonalvs, J., A slug test to assess the osmotic
and hydraulic properties of argillaceous
formations, Water Resources Research, 44,
W07501,
doi:10.1029/2007WR006547,
2008.
Gueutin, P., S. Altmann, J. Gonalvs, P. Cosenza, and S. Violette, Osmotic interpretation of overpressures from monovalent
based triple layer model, in the CallovoOxfordian at the Bure site, Physics and
Chemistry of the Earth, Parts A/B/C, 32,
434440, 2007.
Horseman, S.T., J.F. Harrington, and D.J. Noy,
Swelling and osmotic flow in a potential
host rock, Physics and Chemistry of the
Earth, Parts A/B/C, 32, 408420, 2007.
Ingebritsen, S.E., W.E. Sanford, and C.E. Neuzil,
Groundwater in Geologic Processes, 2nd
edn., Cambridge University Press, Cambridge, 2006.
Katchalsky,
A.,
and
P.F.
Curran,
Nonequilibrium Thermodynamics in Biophysics, Harvard University Press, Cambridge, 1965.
Matsui, H., H. Kurikami, T. Kunimaru, H.
Morioka, and K. Hatanaka, Horonobe URL
projectpresent status and future plans, In:
Eberhardt, E., Stead, D., Morrison, T. (Eds.),
Rock Mechanics: Meeting Society's Challenges and Demands. Taylor & Francis
Group, London, 2007.
Marine, I.W., and S.J. Fritz, Osmotic model to
explain anomalous hydraulic heads, Water
Resources Research, 17, 7382, 1981.
Mitchell, J.K., and K. Soga, Fundamentals of
Soil Behavior, 3rd edn., John Wiley & Sons,
Hoboken, 2005.
Neuzil, C.E., Osmotic generation of anomalous
fluid pressures in geological environments,
Nature, 403, 182-184, 2000.

ACKNOWLEDGMENT
The authors wish to thank Masashi Nakayama
and JAEA for providing the mudstone samples.
This study is regulatory support research funded
by the Nuclear and Industrial Safety Agency,
Ministry of Economy, Trade and Industry, Japan.
The last co-author was supported, in part, by the
U.S. Department of Energy under Contract No.
DE-AC02-05CH11231.

REFERENCES
Appelo, C.A.J., and D. Postma, Geochemistry,
groundwater and pollution. 2nd edn., A.A.
Balkema Publishers, Leiden, 2006.
Bader, S., and H. Kooi, Modelling of solute and
water transport in semi-permeable clay
membranes: Comparison with experiments,
Advances in Water Resources, 28, 203214,
2005.
BattistelliA., C. Calore, and K. Pruess, The
simulator TOUGH2/EWASG for modelling
geothermal reservoirs with brines and
noncondensible gas, Geothermics, 26, 437
464, 1997.
Bresler, E., Anion exclusion and coupling
effects in nonsteady transport through unsaturated soils: I. Theory, Proceedings - Soil
Science Society of America, 37, 663669,
1973.
Cruchaudet, M., J. Crois, and J. M. Lavanchy,
In situ osmotic experiment in the CallovoOxfordian argillaceous formation at the
Meuse/Haute-Marne URL (France): Data
and analysis, Physics and Chemistry of the
Earth, Parts A/B/C, 33, S114S124, 2008.
Finsterle, S., iTOUGH2 User's Guide, Report
LBNL-40040, Lawrence Berkeley National
Laboratory, Berkeley, Calif., 2007.
Fritz, S.J., Ideality of clay membranes in osmotic processes; a review, Clays and Clay Minerals, 34, 214232, 1986.
Garavito, A.M., H. Kooi, and C.E. Neuzil,
Numerical modeling of a long-term in situ
chemical osmosis experiment in the Pierre
Shale, South Dakota, Advances in Water Resources, 29, 481492, 2006.
Garavito, A.M., P.De Cannire, and H. Kooi, In
situ chemical osmosis experiment in the
Boom Clay at the Mol underground research
-8-

Spiegler, K., and O. Kedem, Thermodynamics


of hyperfiltration (reverse osmosis): criteria
for efficient membranes, Desalination, 1,
311326, 1966.
Soler, J.M., The effect of coupled transport
phenomena in the Opalinus Clay and implications for radionuclide transport, Journal of
Contaminant Hydrology, 53, 6384, 2007.
Takeda, M., T. Hiratsuka, and K. Ito, Laboratory
characterization of chemico-osmotic, hydraulic and diffusion properties of rocks:
Apparatus development, Proceedings of
Waste Management 2009, Phoenix, Arizona,
2009.

Neuzil, C.E., and A.M. Provost, Recent experimental data may point to a greater role for
osmotic pressures in the subsurface, Water
Resources
Research,
45,
W03410,
doi:10.1029/2007WR006450, 2009.
Pruess, K., C. Oldenburg, and G. Moridis,
TOUGH2 Users Guide, Version 2.0, Report
LBNL-43134, Lawrence Berkeley National
Laboratory, Berkeley, Calif., 1999.
Rousseau-Gueutin, P., J. Gonalvs, and S.
Violette, Osmotic efficiency in CallovoOxfordian argillites: Experimental vs. theoretical models, Physics and Chemistry of the
Earth, Parts A/B/C, 33, S106S113, 2008.
Rousseau-Gueutin, P., V. de Greef, J. Gonalvs,
S. Violette, and S. Chanchole, Experimental
device for chemical osmosis measurement
on natural clay-rock samples maintained at
in situ conditions: Implications for formation pressure interpretations, Journal of
Colloid and Interface Science, 337, 106116,
2009.

-9-

Das könnte Ihnen auch gefallen