Вы находитесь на странице: 1из 22

Introduction

CO2 has been injected into the Utsira formation since September 1996. In the
SACS project (Saline Aquifer CO2Storage) the basic objective is to monitor the
migration of CO2 in order to be able to determine the fate of CO2 both on short
time scale (< 50 years) and on a long time scale (1000’s of years).Simulation of
CO2 injection in aquifers can be simulated by must ordinary black-oil simulators
as these can handle oil, gas and water phase and by simply leaving the oil out.
Since typical black-oil simulators have a more sophisticated handling of the oil
phase than the water phase, more flexibility can be achieved by rather leaving out
the water phase and to give the oil phase the physical properties of water including
CO2 /water phase properties (Lindeberg 1996). The input parameters for these
simulators are mostly tabulated data. The formation volume factor and viscosity
for respectively brine and CO2 are the most important fluid parameters because
they are the source for the buoyancy of CO2 in the reservoir and the viscous drag
of the fluid through the pores. Other important parameters are the pressure
dependent solubility of gas in brine and capillary pressure. For simulation
scenarios of less than 25 years the least important parameter is the molecular
diffusion constant of CO2in brine, but this parameter can be important on long
time scales when the final fate of CO2 shall be predicted. There are of course
several other parameters, which are important in simulation, but these parameters
are related to the porous medium and are treated in a separate chapter.
Permeability, porosity, capillary pressure and relative permeability including end-
point saturations are important parameters on a microscopic scale and can be
obtained from laboratory core experiments. For a reservoir scale simulation the
most important input data for the reservoir grid is the distribution of
heterogeneities and these must be obtained from seismic, well-logs of other
geological input

Thermodynamic data for the CO/CH4/H2O system


Experimental data
There exists a of experimental a lot of experimental data from phase studies of
CO2. Much of this interest is due its many industrial applications, e.g. as a solvent
and extractant in process and food industry, as a coolant in the nuclear industry, as
a refrigerant in the cooling machines and injectant in the petroleum industry. Many
studies has been performed near the critical point not only because its practical
interest, but also because it is easy to create critical conditions in the laboratory
(31°C, 73 bar). CO2 is therefor particularly suited for studies of near-critical
phenomena.
The CO2 injected in Utsira is not pure CO2, but contaminated with between 1 and
2.5% methane. This can have a pronounced effect on the density of the fluid
especially near the saturation point. Experimental data for the CO2 /methane
system is therefor also required.
The injected CO2 will also be saturated with water vapour. The partial pressure of
waterat these temperatures is so low that its contribution to total composition can
beneglected. It will typically be below 0.07% of the total composition. The
solubility of CO2 in water can, however, not be neglected especially when
modelling the long-term fate of CO2 in the aquifer. A lot of experimental data is
available and most of the studies include experimental conditions that are relevant
for the problems related to CO2 disposal in aquifers.

Thermodynamic conditions in offshore aquifers


There exist a lot of thermodynamic models for CO2 but many of these are valid
oroptimal in only limited temperature and pressure ranges. Some models, that can
model near critical behaviour very well, may have large error at higher pressure
and temperature and vice versa. Before a comprehensive test of various models it
is therefor necessary to know what condition we shall simulate.

Temperature gradient
Heat is transported form the core of the earth to the surface due to primordial
heataccumulated in the earth’s history and heat generated by decay of radiogenic
elements.The resulting temperature gradient in rock is controlled by this heat flow,
the thickness of the crust, the conductivity of the rock, sediments and pore fluid
sand the surface temperature, in this case sea bottom temperature. Below60m the
average sea bottom temperature is relatively constant over the region and is here
set to 4.8°C. The conductivity of the rock varies laterally, and on a map reported
by Eggen (1984) thegradient varies between 20 and 50 mK/m at various locations
in the north Sea. The gradient can also vary vertically because the different
sediments may have different conductance of heat, but in this study it is assumed
that the gradient is linear between the sea floor and the formation. There exists
only scattered temperature measurements at the depths and can at the present so the
possibility to make a more advanced model is limited.

Pressure gradient
In geological formation down to approximately 1500 m the pressure in the North
Sea is typically controlled by the hydrostatic head. From the density profile of the
brine the pressure profile can be calculated. The change in pressure with depth:
dp = ρ (p, t , c,) g dz
where g is gravity constant, ρ (p,t,c) is the density of brine as function of
pressure, p, temperature, t, and concentration., c. The brine at these depths have a
salinity and composition that can be compared to the sea water (3.5% salt).
Because both the reservoir brine and the sea water brine has the same
concentration it is assumed that the brine in column between the sea and the Utsira
formation also has the same concentration. Experimental density of 3.5% brine has
been taken from Engineering Science Data (1968) and the data set has been used to
fit the density to a simple second order polynomial in p and t

Models for thermodynamic properties.


Pure CO2
In reservoir simulation simple cubic equation of states (EOS) are often used to
provide the phase data for hydrocarbon mixtures. Some of these equations were
therefor tested on a set of experimental data at varying pressure at five different
isotherms 25, 29, 32, 35 and 40°C. These were selected because they are relevant
for the temperatures at various depth in a 700 to 1300 m deep aquifer. The cubic
EOS that were investigated were the Soave-Redlich- Kwong (SRK) EOS and the
Peng-Robinson (PR) EOS with and without the Peneloux translation volume
correction. These four equations were compared with the experimental data of
Michels and Krillin (Table 2.1). A fifth EOS that is included in this test it is a Lee-
Kesler type EOS modified by Duan et al. (1992). This EOS has specifically been
designed for the ternarysystem carbon dioxide/methane/ water and could therefor
be a useful tool for the disposal problems since contamination of methane can be
expected when the CO2 source is not flue gas, but process waste from gas
processing (e.g. Sleipner, Snøvit, Natuna). Also the other four EOS can handle
mixtures. This equation was also used to model gas properties in a previous EU
study (“Underground Carbon Dioxide Disposal”). Examples of the results are
illustrated as deviation between measured and modelled densities for the isotherms
29.9°C, 32°C and 40°C in Figure 2.2 to Figure 2.4. The vertical grey line indicates
the reservoir pressure for each isotherm in the Utsira formation.
The results show that in the gas phase the density prediction is relatively good for
all EOS. The two Peng-Robinson EOS are the least accurate but with deviations
less than 3%. In the dense phase, however, the four cubic EOS have significant
deviation in the whole pressure range, but the error is particularly large near the
Utsira conditions, up to 18%. The driving force for migration in the reservoir is not
the density, but the\ buoyancy of CO2 which is proportional to the density
difference between brine and CO2. The error in buoyancy is approximately 3 times
larger than the error in density. The Lee-Kesler type EOS has also a significant
deviation at low pressures, it performs relatively well. This trend is the same at all
the five isotherms that were tested.

Application on Utsira temperature and pressure conditions


For a given temperature and pressure profile as function of depth it is now possible
to model the density profile of CO2. Using the simulated pressure and temperature
profiles obtained in previous chapter as input, the CO2 densities with varying
amount of CO2 is simulated of between sea floor at 80 m to the bottom of the
formation at 1250 m depth was computed. The most accurate profile is the curve
for pure CO2 , but the densities of CO2/methane mixtures are also sufficient
accurate as input in reservoir simulation. The jump in the densities illustrate the
depth where the state of CO2 changes form dense (liquid) phase to gas. Above this
depth the underground storage space is used much more inefficient. Previously
(Holloway et al. 1996) estimated the minimum for storage depth has 800 meter
based on a criterion of high-density requirement and simulation of Van der Sluijs
(1991). The present simulation shows that this range could be extended to 530 m
forpure CO2 while CO2 with 2.5% methane could be effectively stored in
reservoirs withminimum depth of 680 meter (bold curve) if the CO2 is injected
offshore with a minimum water depth of 80 meter. At greater water depths CO2
can be stored in dense phase in even in shallower formations. The difference in this
results and previous estimates are most likely due to different surface temperature.
Holloway et al. used average land temperature as lower temperature but here the
offshore seafloor temperature is used This typically 5 to 10 °C lower than the
average surface land

Solubility of CO2 in brine


Solubility of CO2 in brine is an important parameter for modelling the reservoir
behaviour of CO2 especially on long time scales. If the CO2 during vertical
migration has to pass through labyrinths of horizontalshales or other
heterogeneities that canincrease the amount of brine that is contacted by CO2,
solubility can play a significant role also on shorter time scales. On short time
scale, however, great care must be exercised not to overestimate the solubility due
to numerical dispersion. Application of fine girding or up-scaling of the solubility
are measures to reduce the effect of numerical dispersion. One unit volume of
brine can theoretically store approximately 10% of the one unit volume of pure
CO2 at reservoir conditions. Enick and Klara (1990) have reviewed a number of
different models for CO2 solubility in water and Krichevsky-Ilinskaya equations
appears to be the most reliable. Solubility of CO2 decreases with increased

Derivation of thermodynamic properties


Above only the density was tested for the different models. However, from the
EOS all other thermodynamic properties like fugacity, entropy, enthalpy, internal
energy, heat capacity and speed of sound can be calculating using standard
thermodynamic method by mathematical manipulation of the EOS.

Transport properties: Diffusion and viscosity of in CO2 and brine


Vukalovich (1968) has developed a model for the viscosity of CO2. This model
requires the density of CO2 and the temperature as input. The model was tested on
an experimental data set that has been obtained by Golubev (1959). The standard
deviation between modelled and experimental data for these 18 data is 3.9%.
Experimental viscosities for CO2 /methane mixtures have not been found and it
remains to show how accurate the model of Vukalovich and Altunin is for mixtures
with low methane concentration. The method to determine viscosity of brine as
function of temperature, pressure and salinity has been developed by Hewlett
Packard and is reported in their “Petroleum Fluid Pack” which is a manual for a
programme package. This method is based on a work by Numbere et al.( 1977)
and has been re-coded for use in this project.
Renner (1988) has correlated experimental diffusion coefficients for CO2 in water
to both his own high-pressure measurement and previously measured low-pressure
data. It was shown that the diffusion coefficient, D0, correlated well to the gas and
water viscosity, μ.
This correlation is based on measurements up to 59 bar at 37.8°C and extrapolation
to into the dense-phase CO2 region gives non-consistent results. The available
reservoir simulators do, however, not use the diffusion constant data as function of
pressure and temperature, so a single average value is sufficient. It is therefor
recommended that the experimental value for the highest pressure measured by
Renner is used for the entire depth. Considering the relative importance of
diffusion compared to the other parameters in the simulation this should be
sufficient.

Studies of cores in the Utsira formation


Core preparation
The cores where delivered as frozen samples in 4” aluminium tubes 1 meter long.
Because the cores where totally unconsolidated, special care had to be observed to
avoid unnecessary disturbance of the sand grains. Through some experimenting a
recommended method was obtained. The core sample contained so little formation
water and was too loose for coring or cutting. The core was therefor carefully
melted in its barrel, evacuated and then saturated with 3.5% brine. The core was
then frozen again to –40°C in its barrel. In frozen condition the barrel was opened
by milling two seems on each side and lifting of the two resulting half-pipes. In
this condition it was possible to take 1.5” cores for the remaining measurements.

Porosity and permeability measurements


The porosity was measured by both a helium porosimeter and when the evacuated
core in i core holder 100 bar overburden pressure as sleeve pressure. The
measurements were performed on different cores, but until more measurements
exists, it is recommended to use the average value. 3.3

Capillary pressure and relative permeability


Capillary pressures were measured at ambient conditions in an ultra centrifuge.
The core was originally saturated with brine and the production of brine was
recorded as function of increased centrifugal acceleration. From the production
data the corresponding air/brine capillary curve could be computed. This curve was
reprocessed by help of air/brine and CO2 /brine interfacial tension data form
Hough et al. (1959) and Eggers and Jäger (1994). The consistency of the two data
sets is questionable, but little relevant published data are available. Their data were
obtained with pure water and not with brine. The final CO2 /brine capillary
pressure curve is illustrated in Figure 3.2. No further work on relative permeability
in Utsira cores has been completed so far in this project. Until better analysis
exists, it is recommended to use the gas-brine relative permeability data for high
permeable sandstone of Holt et al. (1995).

Reservoir simulation
Previous simulation of CO2 injection into aquifers (Korbøll and Kaddour 1995,
Lindeberg 1997) has shown that CO2 readily will migrate to the sealing cap due
togravitational forces. A specific simulation on how fast this accumulation will
occur is the objective for this part of the study.

Selection of reservoir segment in Utsira and basic data and


assumptions.
From the supplied map of the Utsira top, Figure 4.1, a shallow anticline trap can be
identified above the injection point. A simplified reservoir model has been built on
the basis of a cylinder with 1600 m diameter below the circle shown on the figure.

true vertical depth relative to the rotary table on the Sleipner A platform). Initial
hydrostatic pressure is applied and this pressure is maintained at the bottom of the
periphery of the model corresponding to a situation with infinite extension of the
whole formation. The injection well is horizontal, and the perforation is 40 m of
which only 20 m will reach into the 60° sector, which is studied. The simulation is
carried out with Eclipse 100 according to the method, fluid and relative
permeability data used by Lindeberg (1996). The permeability used was 3.3 Darcy
(~3.3·10-12m2) in both vertical and horizontal direction. The numerical grid
consists of up to 56 000 radial grid blocks. Molecular diffusion was not included
during this simulation, while capillary pressure and solubility of CO2 in brine
wereaccounted for.

Simulations on a homogeneous reservoir model


In this case it assumed that the reservoir is both a homogenous and isotropic body
with a permeability of approximately 3.3 Darcy. Injection is simulated from start
September 15th 1996. The CO2 will migrate from the injection point to the cap seal
in approximately three weeks. The CO2 bubble will then gradually increase in
radius untilit reaches the spill point at December 1st. 1998 at 800 m radius. The
CO2 will then start to migrate to one of the three traps north, west or south of the
injection trap. The maximum thickness of the CO2 gas cap at start August 15th
1999 is 20 m.12.5 m of this is due to the topography of the seal, while 7.5 m is a
down-dip cone in the proximity to where CO2 ascending from the injection well
reaches the gas cap.
An alternative approach on a regional scale, the CO2 distribution under a perfectly
flat cap rock was simulated with a streamline simulator. The solubility of CO2 is
neglected, but this is not a big error when only short time intervals are considered.
The advantage with this method is that numerical dispersion due block size is
avoided. A result of these runs it was found that in a worse case scenario the CO2
bubble would extent over a distance of 3000 meter from the injection point with an
average thickness of 25 meter after an injection period of 20 years. For all these
runs the injection rate was set at 1 million tonneCO2 per year. Only small
reservoir pressure increments were observed during these runs, this mainly due to
the large extent of the Utsira formation (infinite aquifer) and the relative high
permeability levels.

Heterogeneous reservoir model


In the heterogeneous reservoir model five impermeable layers of 400 m length
have been distributed around the injection point corresponding to thin shale layers
in a grid consisting of 10 000 blocks (250 x 40 blocks with size 1 m x 20 m). This
is illustrated in transmissibility plot in Figure 4.5. The impermeable layers are
visible as dark lines in the figure. The impermeable layers are supposed to
resemble possible heterogeneities seen on well logs from the Utsira formation in
the Sleipner area. The transport properties of these shales are not known, but they
are assumed to perfectly impermeable in this model in order to introduce an
extreme perturbation in flow the pattern of ascending CO2.

Capacity Estimation of CO2 Storage in Saline Aquifers


The storage capacity of CO2 is an estimate of the amount of CO2 that can be
stored in subsurface geologic formations. Because of uncertainties inherent to
subsurface evaluation, exact quantification of geological properties is not possible
and therefore storage capacity is always at best an approximation of the amount of
CO2 that can be stored. Storage capacity estimates rely on the integrity, skill and
judgment of the evaluator and are affected by the geological complexity, stage of
exploration or development, amount of existing storage and of available data.
Global storage estimates are imperfectly known due to inconsistency in assessment
methodologies and gaps in global, regional and local estimates, particularly data
from Africa, South America and large parts of Asia (IPCC, 2005). As a result, two
major works providing methodologies for the estimation of storage capacity of
CO2 in geological formations have been published since the IPCC Special Report
in 2005. These are the ‘Methodology for Development of Carbon Sequestration
Capacity Estimates’ prepared for the Carbon Sequestration Program of the
National Energy Technology Laboratory, U.S. Department of Energy (DOE) by
the Capacity and Fairways Subgroup of the Geologic Working Group of the DOE
Regional Carbon Sequestration Program in December 2006, and the ‘Estimation of
CO2 Storage Capacity in Geological Media – Phase II’ prepared by the Task Force
on CO2 Storage Capacity Estimation for the Technical Group of the Carbon
Sequestration Leadership Forum (CSLF) in April, 2007 (CSLF, 2007). The results
and summary of the CSLF study were published subsequently in the Journal of
Greenhouse Gas Technologies (Bachu et al., 2007).
Methodologies for Capacity Estimations in Saline Aquifers
Both the DOE (2007) and the CSLF (2007) publications provide methodologies for
the estimation of storage capacity of CO2 in saline aquifers,
hydrocarbon reservoirs and coal seams. Only the former will be reviewed in this
section because the others are beyond the scope of this study. Detailed
comparisons between the methodologies implemented in the two studies were
performed by the (CSLF, 2008) and the CO2CRC (2008). Former methodologies,
proposed for example by (Hendriks et al., 2004) or (Li et al., 2005), represent
simplifications of or derivations from the DOE and CSLF methods and are only
reviewed briefly the end of the this section.

DOE-Proposed Methodology

The DOE (2007) study provides a volumetric equation for capacity calculations in
saline aquifers. Displacement of saline water in the pore volume by immiscible
CO2 is the fundamental mechanism implicit in the calculations. No distinction is
made between CO2 that is stored as an immiscible phase within structural or
stratigraphic geologic traps; CO2 that is stored as an immiscible phase outside of
traps (for example, trapped in pores by capillary processes); CO 2 that is stored as
dissolved phase in saline water; and CO2 that is precipitated as minerals. The
expected uncertainty in efficiency of the storage capacity from a combination of
trapping mechanisms is estimated using a Monte Carlo approach. The equation for
capacity calculations in saline aquifers as proposed by the DOE (2006) is as
follows:

Gco2 = A × hg Φtot × ρ × E (1)

Variable and units are defined in Table 14.

Table 1. Volumetric equation parameters for capacity calculation in


saline formations (DOE, 2007).
ParameterUnits*Description
GCO2 M Mass estimate of saline-formation CO2 storage capacity
Geographic area that defines the basin or region being assessed
A L2
for CO2 storage-capacity calculation
Gross thickness of saline formations for which CO2 storage is
hg L
assessed within the basin or region defined by A
Average total porosity of entire saline formation over thickness
Φtot L3/L3
hg
Density of CO2 evaluated at pressure and temperature that
Ρ M/L3 represents storage conditions anticipated for a specific geologic
unit averaged over the depth range associated with hg
CO2 storage efficiency factor that reflects a fraction of the total
E L3/L3
pore volume that is filled, or contacted, by CO2 (see Table 15)
*
L is length; M is mass

The storage efficiency factor (E) adjusts total gross thickness to net gross
thickness, total area to net area and total porosity to effective (interconnected)
porosity actually containing CO2. Without the CO2 storage efficiency factor (E),
Equation 1 presents the Total Pore Volume or maximum upper limit to storage
capacity in free phase. Inclusion of the storage efficiency factor provides a means
of estimating storage volume for a basin or region with the level of knowledge
(uncertainty) in specific parameters determining the type of CO 2 storage capacity
estimated. Monte Carlo simulations for parameter ranges typical for US
sedimentary basins estimated a range of E between 1 and 4 percent of the bulk
volume of saline formations for a 15 to 85% confidence range.

A reasonable maximum Effective Storage Capacity may be estimated by assuming


CO2 injection wells are placed regularly throughout the basin, or region, and
multiplying the storage efficiency terms in Table 15 together to determine the
storage efficiency factor for the basin or region. In Table 14, the first three terms
(An/At, hn/hg and Φe/Φtot) are used to define the regional effective pore volume, and
the next three terms (EA, EI and Eg) are used to define the fraction of that volume
accessed by CO2 from injection wells. The remaining term (Ed) in Table
15 accounts for the proportion of the effective pore volume filled with CO2.
Ideally, the last four terms would be determined for each potential well site and
combined to estimate the total storage volume; however, practically it may be
easier to determine the average volume of pore space accessed by injecting
CO2 into a single well and then multiplying by the expected number of wells.

Although the capacity estimates given in the Atlas are based on CO2 stored as a
separated fluid phase, some Regional Partnerships initially used the assumption of
CO2 dissolved in formation water in their calculations. In those cases, the
efficiency factor E for dissolved CO2 was converted to an E for free-phase CO2,
using a single conversion factor representative of the average pressure, temperature
and salinity conditions for each basin or region.

Table 2. Terms included in storage efficiency factor for saline


formations (from DOE, 2007).

Term SymbolDescription
Terms used to define the entire basin/region pore volume
Net to total Fraction of total basin/region area that has a suitable
An/At
area formation present
Net to gross Fraction of total geologic unit that meets minimum
hn/hg
thickness porosity and permeabilityrequirements for injection
Effective to
Fraction of total porosity that is effective, i.e.
total porosityΦe/Φtot
interconnected
ratio
Terms used to define the fraction of pore volume accessed by CO2 from
injection wells
Fraction of immediate area surrounding an injection well
Areal
that can be contacted by CO2; most likely influenced by
displacement EA
areal geologic heterogeneity such as faults or permeability
efficiency
anisotropy
Fraction of vertical cross section (thickness), with the
Vertical volume defined by the area (A) that can be contacted by
displacement EI the CO2 plume from a single well; most likely influenced
efficiency by variations in porosity and permeability between
sublayers in the same geologic unit. If one zone has higher
Term SymbolDescription
permeability compared with others, the CO2will fill this
one quickly and leave the other zones with less CO2 or no
CO2 in them
Fraction of net thickness that is contacted by CO2 as a
consequence of the density difference between CO2 and in
Gravity Eg situ water. In other words, (1-Eg) is that portion of the net
thickness not contacted by CO2 because the CO2 rises
within the geologic unit
Portion of the CO2-contacted, water-filled pore volume that
Microscopic
can be replaced by CO2. Ed is directly related to irreducible
displacement Ed
water saturation in the presence of CO2 or residual
efficiency
CO2 saturation, dependent on position within the plume

Due to the variability in data quality and data distribution, the complexity of the
subsurface and various levels of subsurface characterisation by different assessors
or scientist, the DOE suggests the use of a 1 (low) to 9 (high) relative index that
indicates the level of confidence in the capacity estimations:
CSLF-Proposed Methodology

CSLF (2007) provides individual equations for structural and stratigraphic


(volumetric) trapping, residual-gas saturation trapping, dissolution trapping, as
well as a detailed discussion on the processes and time frames involved for mineral
precipitation and hydrodynamic trapping. Separated equations are given for the
estimation of theoretical and effective storage capacities. In the following only the
equations for effective storage capacities are shown, which can be converted easily
to theoretical capacity by omitting the capacity coefficient C, which describes
trapping efficiency.
For calculating the storage capacity in structural and stratigraphic traps the
CSLF Taskforce proposes the use of two equations that result in either estimates of
storage volume or corresponding mass of CO2. The effective storage volume is
given by:

Vco2 = A × h × Φ × (1 − Swirr) × Cc (2)

where A = trap area, h = average thickness, Φ = average porosity, S wirr =


irreducible water saturation and Cc = capacity coefficient that incorporates trap
heterogeneity, CO2 buoyancy and sweep efficiency.

The corresponding mass of CO2, MCO2, that can be effectively stored depends on
the average subsurface temperature, T, and the subsurface pressure, which can
range between the initial formation pressure, Piand the maximum approved
bottomhole injection pressure Pmax:

minMco2 = ρco2 (Pi, T) × Vco2 ≤ Mco2 ≤ maxMco2 = ρco2 (Pmax, T) × Vco2 (3)

Relations (2) and (3) can be applied to both theoretical and effective storage
capacity estimates for basin-and regional-scale assessments by applying them
individually to all the structural and stratigraphic traps identified as potential
candidates for CO2 storage and summing the resulting individual capacities. They
can be applied also to the case of a plume of CO2 that is not necessarily contained
in a stratigraphic or structural trap (CSLF, 2008).

Unlike in the case of structural and stratigraphic traps, residual-gas trapping is


time-dependant and the corresponding CO2 storage capacity has to be evaluated at
a specific point in time. The storage volume can be estimated by:

Vco2t = ΔVtrap × h × Φ × Sco2t (4)

whereΔVtrap = rock volume previously saturated with CO2 that is invaded by water
and SCO2t = trapped CO2 saturation after flow reversal.

The mass of stored CO2 by residual gas trapping is obtained by multiplying the
storage volume by the density of CO2 at in-situ conditions, but this density is both
time- and position-dependent as pressure and temperature vary along the flow path
and as, for the same location, pressure builds up or decays, depending on the stage
of the storage operation.

The dissolution of CO2 in formation water is a continuous, time-dependent process,


which depends on pressure temperature and water salinity, as well as the amount of
CO2 coming into contact with formation water under-saturated with CO2.
Therefore, the CO2 storage capacity through solubility trapping has to be
evaluated for a specific period of time. For average aquifer thickness and porosity,
the following equation gives the effective storage capacity of CO2 in solution:

Mco2t = A × h Φ × (ρsXsco2 − ρ0X0co2) × C (5)

Where ρ = formation water density, X = CO2 content (mass fraction), subscripts 0


and S denote initial CO2content and CO2 content at saturation, respectively, and C
= coefficient that includes the effect of all factors that affect the spread and
dissolution of CO2 in the whole aquifer volume under consideration.

The estimation of CO2 storage capacity in mineral and hydrodynamic traps is


discussed only with respect to the processes and timeframes involved. Like
residual-gas and solubility trapping, mineral trapping is a time-dependent process,
operating on the scale of millennia and storage capacity needs to be estimated for a
particular point in time. Because hydrodynamic trapping is based on several
CO2 trapping mechanisms acting at times simultaneously and sometimes being
mutually exclusive, the CO2 storage capacity has to be evaluated at a specific point
in time as the sum of the individual storage capacities. Given the complexity of the
processes involved, only local- and site-scale numerical modelling backed up by
laboratory experiments and field data can provide capacity estimates and
timeframes for mineral and hydrodynamic trapping of CO2.

Other Methodologies for Calculating CO2 Storage Capacity

The methodology employed by Hendriks et al. (2004) to estimate global storage


capacity values used general values for aquifer thickness and porosity, and
represents a somewhat simplified version of the DOE-proposed methodology
(Equation 1). Low, best, and high storage estimates for supercritical CO2 in
gigatonnes are calculated according to:
Gco2i = A × hi × Φi × ρco2 × 0.01 × 0.02 × 10−12 (6)

where the subscript i denotes minimum, best and maximum values for aquifer
thickness (50 m, 100 m 300 m) and porosity (0.05, 0.2, 0.3), A = surface area of
sedimentary basin (m2), and ρCO2 = average density of CO2 (750 kg/m3).
Multiplication by 0.01 and 0.02 assumes that 1 % of the aquifer is part of a
structural trap and 2 % sweep efficiency, respectively.

(Brennan and Burruss, 2006) define a Specific Storage Volume (m3/t) for saline
aquifer storage, which incorporates free-phase CO2 storage and solubility trapping:

(7)

whereSwr = residual water saturation after CO2 injection, ρCO2 = CO2 density (t/m3),
ρaq = density of formation water at given pressure and temperature conditions and
CO2 saturation (t/m3) and XCO2 = concentration of CO2 dissolved in formation
water (t of CO2 per t of aqueous solution).

The SSVSA increases with time as free-phase CO2 dissolves into formation water
and the residual water saturation increases until the saturation point is reached. In
other words, a given mass of injected CO2 in solution will occupy a larger volume
of the aquifer than free-phase supercritical CO2.

The Research Institute of Innovative Technology for the Earth (RITE) uses a
hybrid of the DOE and CSLF-proposed technologies to estimate storage capacity
and select preferred storage sites in Japan. Following the classification by (Tanaka
et al., 1995), Li et al. (2005) assess the capacity of selected storage sites in four
separate categories: 1) oil and gas reservoirs and neighbouring aquifers, 2) aquifers
in anticlinal structures, 3) aquifers in monoclinal structure – onshore, and 4)
aquifers in monoclinal structure – offshore. For the sites in Category 1, a
volumetric storage capacity is estimated assuming trapping of supercritical CO 2,
whereas capacity estimates (in grams CO2) for categories 3 and 4 are based on
solubility trapping. A site is divided into 500 m thick blocks (Figure 12) and the
capacity of each block is then calculated based on its average temperature and
pressure.

Figure 3. Dividing a site into blocks. Only the grey blocks are considered to
contribute to storage, (Li et al., 2006). A site is bounded on the sides by
vertical surfaces through coastline, the 1000-m sedimentary isopach, and/or
the 500-m sea-depth isogram. At its top is the 500-m isopachous surface, and
at the bottom are the 3000-m sedimentary isopachous surface and/or
basement..

Conclusion
For pure CO2there exists a wealth of experimental data at conditions which are
interesting for underground CO2 disposal. There exists also several equation of
state that can model physical properties in for this purpose and the recommended
model is the IUPAC EOS. For CO2 /methane mixtures there are only a few data
sets that are relevant for Utsiratemperature and pressure conditions. Between 15
and 38°C there exists no data and there data for methane concentrations below
20% are scarce. It is recommended to obtain measurements for these conditions for
further improving the LK-Duan EOS model, which is the most promising of the
EOS tested for CO2 /methane mixtures.

References

Bateson L., Vellico M., Beaubien S.E., Pearce J.M., Annunziatellis A., Ciotoli G., Coren
F., Lombardi S., Marsh S., (2008), The application of remote sensing techniques to
monitor CO2 storage sites for surface leakage: method development and testing at Latera
(Italy) where naturally-produced CO2 is leaking to the atmosphere, International Journal
of Greenhouse Gas Control, 2, 388-400.
Bielinski A., Kopp A., Schütt H., Class H., (2008), Monitoring of CO2 plumes during
storage in geological formations using temperature signals: Numerical investigation.
International Journal of Greenhouse Gas Control, 2, 319-328.
Celia M.A., Nordbotten J.M., Court B., Dobossy M., Bachu S., (2011), Field-scale
application of a semianalytical model for estimation of CO2 and brine leakage along old
wells, International Journal of Greenhouse Gas Control, 5, 257-269.
Chadwick R.A., Arts R., Bernstone C., May F., Thibeau S., Zweigel P., (2007), Best
practice for the storage of CO2 in saline aquifers. Observations and guidelines from the
SACS and CO2 STORE projects, British Geological Survey.
Doughty, C., Pruess, K., Benson, S. M., Hovorka, S. D., Knox, P. R., and Green, C. T.:
2001 Capacity investigation of brine-bearing sands of the Frio Formation for geologic
sequestration of CO2, in: Proceedings of the First National Conference on CO2 Capture
and Sequestration, (CD-ROM), Washington, DC, May 14–17.
Gasda, S. E.: 2004, CO2 Sequestration into a Mature Sedimentary Basin: Determining
the Capacity and Leakage Potential of a Saline Aquifer Formation, Master’s Thesis,
Princeton University, 121 pp.
Altunin, V.V., Gadetskii, 1971: (title translated to English:) “Equation of State and
Thermophysical properties of Liquid and. Gaseous Carbon Dioxide”,
Teploenergetica, 18, (3), 81-84
Angus, S., Armstrong, B., de Reuck, K.M. (editors.) “International
Thermodynamic Tables of the Fluid State, 1973: Carbon Dioxide”, International
Union of Pure and Applied Chemistry,. Project Centre. London, UK
Arai, Y., Kamanishi, G.I.,Saiyo, S., 1971: “ The experimental determination of the
P-V- T-X relations for the Carbon dioxide-nitrogen and the carbon dioxide-
methane systems”, J-Chem. Eng. Japan 4, p113 – 122
Duan, Z., Møller, N., Weare, J.H., 1992: “An equation of the for the CH4-CO2-
H2O system: II. Mixtures from 50 to 1000°C and 0 to 1000 bar, Geochimica et
CosmochimicaActa, Vol 56, p2605-2617.
Eggen, 1984: “Modelling of subsidence, hydrocarbon generation and heat transport
in the Norwegian North Sea”, IFP
Eggers, R., Jäger, Ph.T. 1994: “Die EinflussvomGrenzflächenerscheinungen auf
denStoffübergang und die Prozessführung in Gegenstromkolonnenmit
überkritschemCO2”, Wärme und Stoffübertragung, 29, p 373- 377.
Encyclopedie Des Gaz, L’AIR LIQUIDE,.p 333 – 368.
Engineering Science Data, 1968: “Density of Water Substance”, Institute of
Mechanical Engineers , London.
Enick, R.M. and Klara, S.M., 1990: “CO2 Solubility in Water and Brine Under
Reservoir Conditions”, Chem. Eng. Comm., Vol. 90, pp. 23-33.
Golubev, I.F., 1969: “Viscosity if gases and gasmixtures, Fizmatigiz, 1950,
Reproduced by Vukalovich, M.P., Altunin, V.V., “Thermopysical Properties of
Carbon Dioxide”, Collet’s Publishers LTD. London &Wellingborough p 387
Holloway, S., Bateman, K., Barbier, J., Doherthy, P., Fabriol, H., Harrison,. R.,
Heederik, J.P., van der Meer, L.G.H., Czernichowski-Lauriol, I., Lindeberg,
E.G.B, Pearce, J.M., Summerfield, I.R., Rochelle, C., Sanjuan, B., Schartzkopf, T.,
Kaarstad, O., Berger,B. 1996: "The Underground Disposal of Carbon Dioxide",
Final Report of JOULE II Project no. CT92-0031, Non Nuclear Energy R&D
Programme

Вам также может понравиться