Assessing The Potential of Solubility Trapping in

Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

www.nature.

com/scientificreports

OPEN Assessing the potential of solubility


trapping in unconfined aquifers
for subsurface carbon storage
Mouadh Addassi 1*, Abdirizak Omar 1, Hussein Hoteit 1, Abdulkader M. Afifi 1,
Serguey Arkadakskiy 2, Zeyad T. Ahmed 2, Noushad Kunnummal 2, Sigurdur R. Gislason 3 &
Eric H. Oelkers 4
Carbon capture and storage projects need to be greatly accelerated to attenuate the rate and degree
of global warming. Due to the large volume of carbon that will need to be stored, it is likely that the
bulk of this storage will be in the subsurface via geologic storage. To be effective, subsurface carbon
storage needs to limit the potential for ­CO2 leakage from the reservoir to a minimum. Water-dissolved
­CO2 injection can aid in this goal. Water-dissolved ­CO2 tends to be denser than ­CO2-free water, and
its injection leads immediate solubility storage in the subsurface. To assess the feasibility and limits
of water-dissolved ­CO2 injection coupled to subsurface solubility storage, a suite of geochemical
modeling calculations based on the TOUGHREACT computer code were performed. The modelled
system used in the calculations assumed the injection of 100,000 metric tons of water-dissolved
­CO2 annually for 100 years into a hydrostatically pressured unreactive porous rock, located at 800
to 2000 m below the surface without the presence of a caprock. This system is representative of an
unconfined sedimentary aquifer. Most calculated scenarios suggest that the injection of ­CO2 charged
water leads to the secure storage of injected ­CO2 so long as the water to ­CO2 ratio is no less than ~ 24
to 1. The identified exception is when the salinity of the original formation water substantially exceeds
the salinity of the ­CO2-charged injection water. The results of this study indicate that unconfined
aquifers, a generally overlooked potential carbon storage host, could provide for the subsurface
storage of substantial quantities of ­CO2.

Carbon capture and storage (CCS) is one of the key strategies available to combat the rising levels of greenhouse
gases in the atmosphere. Some estimates suggest that global CCS efforts must successfully store in excess of sev-
eral gigatons annually to limit global warming to no more than 1.5 or 2 °C above pre-industrial l­ evels1. Current
CCS efforts, however, store less than ~ 50 million tons of ­CO2 annually. It is, therefore, critical to accelerate the
adoption of carbon storage s­ olutions2–5.
Numerous subsurface C ­ O2 storage methods have been advocated and studied over the past few decades to
enable large-scale storage. Some of the targeted geologic media, such as some deep confined saline aquifers and
depleted oil and gas reservoirs, have robust caprocks capable of maintaining buoyant supercritical ­CO2 in the
subsurface for extended time-frames6–9. Other potential subsurface carbon storage formations, such as fractured
reactive basalts or unconfined saline aquifers, may not contain caprocks and have fluid pathways that would
enable buoyant single-phase C ­ O2 to rapidly return to the ­surface8. Injection of carbonated water has been pro-
posed as a solution to this ­issue10–14. This method introduces an already stable and non-buoyant phase into the
reservoir, which allows the storage of C ­ O2 directly by solubility t­ rapping14–18. Moreover, water-dissolved C
­ O2 can
also promote the dissolution of divalent metal-bearing silicate minerals leading to the formation of carbonate
minerals fixing the injected dissolved gas in the solid-state17,19–22.
The injection of ­CO2-charged water into the subsurface overcomes the slow dissolution of ­CO2 into reser-
voir fluids if injected as a single phase. Although carbon dioxide dissolution into water is fast, and equilibrium
is generally assumed for this reaction at the ­CO2-water ­interface23–25, in the subsurface the overall dissolution
process is slow. Carbon dioxide dissolution in subsurface rock formations is limited by ­CO2 diffusion into the
water at the ­CO2-water interface and the subsequent ­CO2-charged water convection driven by density contrasts

1
King Abdullah University of Science and Technology (KAUST), Thuwal 23955‑6900, Saudi Arabia. 2Environmental
Protection, Saudi Arabian Oil Company, Dhahran, Saudi Arabia. 3Institute of Earth Sciences, University of Iceland,
Sturlugötur 7, 102 Reykjavík, Iceland. 4Géosciences Environnement Toulouse (GET), CNRS UMR 5563, Toulouse,
France. *email: mouadh.addassi@kaust.edu.sa

Scientific Reports | (2022) 12:20452 | https://doi.org/10.1038/s41598-022-24623-6 1


Content courtesy of Springer Nature, terms of use apply. Rights reserved
Vol.:(0123456789)
www.nature.com/scientificreports/

between the C­ O2-saturated and unsaturated b ­ rines26. The slow reservoir scale dissolution of C
­ O2 is a bottleneck
in exploiting solubility trapping as a primary storage method in geologic formations.
The goal of this study is to assess the potential of C­ O2 storage in unconfined aquifers though the injection
of carbonated water through a series of comprehensive chemical transport calculations. Particular attention is
focused on the effect of pressure, temperature, and salinity gradients on the stability of carbonated water injected
into the subsurface. The purpose of this manuscript is to report the results of this computational study and to use
these results to assess the potential for subsurface carbon storage in unreactive unconfined aquifers.

Methods
The currently study considers only reactions among the fluid phases in the system; mineral dissolution and pre-
cipitation reactions have been excluded for several reasons. First, with the exception of relatively rapidly reacting
minerals such as evaporites, carbonates and some silicates such as basaltic glass and olivine, the dissolution rates
of many minerals, including those of clays and quartz, which dominate many sedimentary rocks are too slow
to affect greatly model results. Secondly, the inclusion of mineral-fluid reactions would require the choice of
host-rock mineralogy making computed results less general.
TOUGHREACT version 3.3227 was used to calculate the multiphase flow of C ­ O2, including C ­ O2 solubility
and the transport in aqueous fluids. The results of solubility calculations from this code, which are critical for
this study, were validated by comparison with corresponding solubilities calculated using CMG-GEM version
2018.1028, and with available experimental data. A close correspondence was found between all calculated and
experimental values.

Transport model. The governing equations for multiphase multicomponent transport in porous media are
given by the material/mass balance equations, Darcy’s law, and the thermodynamic equilibrium between the
­phases29,30. The general form of these equations is described in this section, highlighting relations relevant to
this work. The mass of each component in the phases present in the system can be described by material balance
given by:
∂Mκ
= −∇Fκ + qκ κ = 1, . . . , nc , (1)
∂t
where M represents the mass accumulation, F refers to the mass flux, and q denotes a source or sink term. The
components, e.g., water and ­CO2 are represented by the index κ for each of the nc components in the system.
The mass accumulation term Mκ is given by:

Mκ = φ Sβ ρβ Xκ,β ,
(2)
β

where φ denotes the porosity,Sβ refers to the saturation,β represents the phase i.e., gas or aqueous fluid, ρβ refers
to the mass density, and Xκ,β represents the mass fraction of component κ in phase β.
The flux of the κth component, Fκ , is written as:

Fκ = Xκ,β ρβ uβ ,
(3)
β

where uβ denotes the phase velocity, which is described by Darcy’s l­ aw31 as:
kkr,β
uβ = − (∇pβ − ρβ g), (4)
µβ
where k signifies the absolute permeability of the porous medium, and kr,β , µβ , and ρβ represent the relative
permeability, viscosity, and mass density, respectively, of the phase β . The symbol p denotes the pressure and g
the gravitational acceleration.
The diffusive flux, J , is derived from Fick’s ­law32:
Jβ = −cβ Dβ ∇xβ , (5)
where Dβ corresponds to the diffusion coefficient in phase β.
Local thermodynamic equilibrium is expressed as the equality of the fugacities of each component in the
phases β1 and β2:
fi,β1 (T, p, xj,β1 ) = fi,β2 (T, p, xj,β2 ), i = 1, ..., nc , j = 1, ..., nc − 1, (6)
where fi,β stands for the fugacity of a component i in phase β. The fugacity can be determined using an equation
of state. The equations presented above are implemented in TOUGHREACT​33,34 and control multicomponent
transport in the physical domains represented in the model calculations presented below.

Carbon‑dioxide solubility. The interaction between ­CO2 and water is an important phenomenon during carbon
storage. Solubility trapping, a key sequestration mechanism, is driven by this ­interaction15,23,35,36. The dissolution
of gaseous ­CO2 in water is a relatively rapid reaction at the local scale, and therefore local thermodynamic equi-
librium is assumed between the gaseous ­CO2 and aqueous phases. This local equilibrium is expressed in terms
of the fugacities in each phase, such that:

Scientific Reports | (2022) 12:20452 | https://doi.org/10.1038/s41598-022-24623-6 2


Content courtesy of Springer Nature, terms of use apply. Rights reserved
Vol:.(1234567890)
www.nature.com/scientificreports/

fi,g = fi,aq , i = 1, ..., nc , (7)


where, fi,g represents the fugacity of the ith component in the gas phase, and fi,aq represents the fugacity of the
ith component in the aqueous phase. The fugacity of ­CO2 in the gas phase is related to its partial pressure by:
fCO2 ,g = ϕPCO2 , (8)
37
where ϕ denotes the fugacity coefficient estimated using the Peng-Robinson ­EOS , and PCO2 refers to the partial
pressure of ­CO2. In the aqueous phase, the fugacity of ­CO2 is calculated.
using the extended Henry’s l­ aw38:
fCO2 ,aq = yCO2 HCO2 , (9)
where, yCO2 stands for the mole fraction of ­CO2 in the aqueous phase and HCO2 represents Henry’s constant. In
pure water, Henry’s constant, which is a function of pressure, temperature, the universal gas constant, and the
partial molar volume with respect to a reference pressure, is given by:
v i (p − p∗ )

ln HCO2 = ln HCO + . (10)
2 RT
In saline water, a correction is made to Henry’s constant to account for salinity, such that:
 
Hbrine,CO2
ln = ks,CO2 ms , (11)
HCO2
where Hbrine,CO2 denotes Henry’s constant for ­CO2 in the brine, HCO2 corresponds to Henry’s constant for ­CO2 in
pure water, ks,CO2 designates the Setchenov salting-out coefficient for a C
­ O2-water aqueous phase, and ms refers
to the molality of the dissolved salt.
Carbon-dioxide solubility in TOUGHREACT and CMG-GEM is calculated following Henry’s law. In
TOUGHREACT, the effect of salinity, temperature, and pressure on the fugacity coefficients of ­CO2 is calculated
within the ECO2N module using the ­H2O-CO2 mutual solubility model of Spycher and ­Pruess39–41. In CMG-
GEM, the effects of salinity, temperature, and pressure on Henry’s constants are accounted for by either Harvey’s
­correlation42 or the Li and Nghiem ­correlation43.

Numerical examples
This study aims to assess the limits of solubility trapping in unconfined aquifers by the injection of ­CO2-charged
water. Towards this goal, several numerical examples have been conducted. The results of each example are com-
pared to a reference case to illustrate the effects of varied parameters. As a reference case, a simplified reservoir
with a hydrostatic pressure gradient represented by a radial model is considered. In this reference case, carbon-
ated water is continuously injected for 100 years. The distribution of ­CO2 during injection and post injection is
computed over and beyond this time period.
A single system geometry was selected to illustrate the effect of varying parameters on water-charged C ­ O2
injection efforts. The choice of a single system geometry allows for direct comparison of the model results. A
schematic illustration of the modelled subsurface system is shown in Fig. 1. The domain consists of a radial model
with a 1 km radius and 2 km total depth. The injection interval is 800 m long, starting from a depth of 800 m and
extending down to a depth of 1600 m. The thickness of the injection zone was set to minimize pressure buildup.
It might be uncommon to find continuous reservoir rock without heterogeneities in a vertical interval. However,
this could be thought of as a well penetrating multiple formations with a net pay thickness of 800 m, which may
include non-reservoir streaks. It should be possible to have a ‘large’ injection zone in unconfined aquifers, as we
are not targeting a specific zone below a seal. This contrasts with conventional reservoirs where the injection zone
is limited to the targeted traps. An example of that is when using commingle wells to access stacked reservoirs,
including geothermal reservoirs where the perforated intervals are commonly within 1000-2000m44,45. A no-flow
boundary is located at a depth of 2 km. This represents a non-permeable formation. A fixed-pressure open flow
boundary is located at the top and the edges of the radial model.
The total pore volume in the system is approximately 6 × 108 ­m3. A rough estimate of the maximum solubil-
ity storage capacity of this system is calculated to be 12 million metric tons of C ­ O2, assuming the average C ­ O2
solubility in the reservoir aqueous phase is 4% by mass, and that half of the pore space is available for storage.
This mass of ­CO2 storage volume would allow an injection rate of 100,000 metric tons/year for 120 years to fill the
domain completely with C ­ O2 saturated water. Such an injection would require a total injection of 2.5 to 3.5 mil-
lion metric tons of fluid into the subsurface over 120 years. This injection rate is within the range of geothermal
wells. Average water re-injection rates from geothermal sites around the world range from 0.2 to 5 Mt/year46–48,
see Fig. S1. The injected water is assumed to enter the reservoir with a temperature of 40 °C. The specific heat
and the wet heat conductivity of the rock are assumed to be 1000 J/(kg·K) and 2.0 W/(m·K), respectively. These
values are within the range of reported values for different rock types, including sedimentary and volcanic r­ ock49.
A linear function is assumed for the relative permeability with a critical gas saturation of 0.1.

Model parameters. Numerous parameters have been varied to illuminate their effects on the fate of
injected water dissolved C
­ O2. Table 1 summarizes the limits of the parameters considered in this study. The
main parameters varied include the temperature and its gradient, the salinity and its gradient, the vertical to
horizontal permeability ratio and the perforation thickness of the dissolved gas injection.

Scientific Reports | (2022) 12:20452 | https://doi.org/10.1038/s41598-022-24623-6 3


Content courtesy of Springer Nature, terms of use apply. Rights reserved
Vol.:(0123456789)
www.nature.com/scientificreports/

Figure 1.  A schematic of the radial simulation domain considered in this study, including boundary conditions,
injection zone, and the temperature gradient for the reference system. The plot background shows the numerical
domain with 20 m blocks in the z direction and varying grid size in the radial direction. The finest grid blocks in
the radial direction are 5 m, near the well, increasing to 50 m further away from the well.

Injection water Injection zone Permeability


Temperature gradient Salinity Salinity gradient Temperature Salinity Top Bottom kv Kh W\CO2 ratio*
°C/km ppm ppm/m °C ppm m m md md kg/kg
Reference case 30 35,000 0 40 35,000 800 1600 100 100 25
18 35,000 0 40 35,000 800 1600 100 100 25
Temperature gradient
50 35,000 0 40 35,000 800 1600 100 100 26
Ignoring thermal effects 30 35,000 0 Same as reservoir 35,000 800 1600 100 100 28
30 10,000 0 40 35,000 800 1600 100 100 25
Constant reservoir salinity
30 100,000 0 40 35,000 800 1600 100 100 35*
30 10,000 0 40 10,000 800 1600 100 100 22
Injecting reservoir fluid
30 100,000 0 40 100,000 800 1600 100 100 35
30 – 10 40 35,000 800 1600 100 100 25
Salinity gradient
30 – 100 40 35,000 800 1600 100 100 25
30 35,000 0 40 35,000 800 1200 100 100 26
Perforation zone thickness
30 35,000 0 40 35,000 800 2000 100 100 25
30 35,000 0 40 35,000 1000 1800 100 100 23
Injection depth
30 35,000 0 40 35,000 1700 2000 100 100 19
30 35,000 0 40 35,000 800 1600 10 100 20
Permeability ratio
30 35,000 0 40 35,000 800 1600 200 100 27

Table 1.  Model parameters and their impact on solubility storage potential. Significant values are in bold.
*The water to ­CO2 ratio was fixed to be the minimum value without allowing C ­ O2 to escape to the surface
within 100 years of carbonated water while injecting 100,000 metric tons C
­ O2/year.

Results
The solubility of C
­ O2 and the density of ­CO2‑charged water as a function of depth. The solu-
bility of C
­ O2 in water, and the density of C
­ O2-charged water as a function of depth depend on many factors
including the gas pressure, temperature, and salinity. Figure 2 shows examples of how these properties vary
versus depth, assuming a hydrostatic pressure gradient. At a constant temperature of 50 °C, the solubility and

Scientific Reports | (2022) 12:20452 | https://doi.org/10.1038/s41598-022-24623-6 4


Content courtesy of Springer Nature, terms of use apply. Rights reserved
Vol:.(1234567890)
www.nature.com/scientificreports/

Figure 2.  The left pair of plots show ­CO2 solubility and density of ­CO2-saturated water as a function of depth.
The temperature in the system is constant at 50 °C and two salinity cases are shown (0 ppm—pure water; and
35,000 ppm—seawater). The experimental data shown for C ­ O2 solubility was compiled by (circle filled with blue
colour)Appelo et al.50 and (square filled with green colour) Koschel et al.51. The right pair of plots show ­CO2
solubility and ­CO2-saturated water density as a function of depth, assuming a temperature gradient of 30 °C/km.
Three salinity cases are shown; pure water, seawater, and a vertical salinity gradient of 50 ppm/m. The model
results were calculated using ECO2N EOS in TOUGHREACT. A normal hydrostatic pressure gradient was
assumed for all cases, as shown on the secondary y-axis of the left-most plots.

density of carbonated water increase with increasing depth. The C ­ O2 solubility and density of pure water and
saline seawater differ somewhat. Increasing salinity lowers ­CO2 solubility but increases the density of the water.
The calculated C
­ O2 solubility values from the model calculations are in close agreement with experimental
data compiled by Appelo et al.50 and Koschel et al.51. The thermodynamic properties of ­CO2/brine mixtures
calculated using TOUGHREACT, and GEM were compared to experimental data ­in52. Both numerical codes
captured the solubility trends well, and good matches were obtained against experimental data for different
ranges of temperature, pressure, and salinity investigated.
The presence of a temperature or salinity gradient, however, changes significantly the variation of ­CO2 solu-
­ O2 saturated water with depth. A temperature gradient of 30 ◦C/km results in lower
bility and the density of C
­CO2 solubility with depth compared to the constant temperature case as shown in Fig. 2. The solubility of ­CO2
in water increases with increasing pressure and decreases with increasing temperature and salinity. Down to a
depth of approximately 600–800 m, the pressure effects dominate, resulting in an increase of ­CO2 solubility with
depth (for a temperature gradient of 30 ◦C/km). Below approximately 600–800 m depth, the solubility curve
flattens for pure water and seawater, and it reverses in the presence of a 50 ppm/m salinity gradient. The carbon-
ated water density curve, thereby, shows how the temperature gradient results in reduced water density with
depth after a few hundred meters of depth. The effect of fluid density decreasing due to increasing temperature
dominates the density increase due to increasing pressure with depth for both pure water and seawater. The
presence of a 50 ppm/m salinity gradient in the reservoir, combined with the effect of pressure and solubility,
overrides the effect of the temperature gradient on water density, and results in a net increase of water density
with increasing depth.

The fate of injected ­CO2 into a homogeneous hydrostatically pressured reservoir. The first
model calculation focuses on a system having a constant porosity of 0.1 and constant permeability of 100 mD in
the vertical and horizontal directions. Although it is unrealistic to encounter a homogenous reservoir with con-
stant porosity and permeability, the selected values are within the lower range for ­sandstone53, as well as other
rock types, including basaltic ­rocks54. The impact of the ratio between vertical and horizontal permeability are
investigated later in this study. The surface temperature is taken as 30 ◦C and temperature gradient as 30 °C/km.
The salinity of the reservoir and the injected water is assumed constant and equal to a 35,000 ppm NaCl solution,
which is approximately equal to that of seawater.
Figure 3 illustrates 2D cross-section time snapshots for the pressure, temperature, C ­ O2 mass fraction in the
liquid, gas saturation, and water density. Note that gas saturation in this study represents the fraction of exsolved
­CO2 mass in the pore fluids. The fluid pressure profile remains fairly constant as the injection rate is kept constant.

Scientific Reports | (2022) 12:20452 | https://doi.org/10.1038/s41598-022-24623-6 5


Content courtesy of Springer Nature, terms of use apply. Rights reserved
Vol.:(0123456789)
www.nature.com/scientificreports/

Figure 3.  Simulation results for the reference case considered in this study. 2D cross-sections of the (from top
to bottom) pressure, temperature, ­CO2 mass fraction in the liquid, gas saturation, which is the mass fraction
of exsolved ­CO2 in the pore fluids, and water density for the reference system after 1, 10, 50 and 100 years of
carbonated water injection. Color bars indicating the range for each of the properties are displayed on right
side of the plots. The dotted line marks the gas-charged water injection zone. This homogeneous hydrostatically
pressured reference system assumes a 30 °C/km gradient, constant salinity of 35000 ppm (seawater), and water
injection temperature of 40 °C. The water to C­ O2 injection mass ratio is 25:1.

Scientific Reports | (2022) 12:20452 | https://doi.org/10.1038/s41598-022-24623-6 6


Content courtesy of Springer Nature, terms of use apply. Rights reserved
Vol:.(1234567890)
www.nature.com/scientificreports/

Water den. Gas saturation


After 100 years of
carbonated water inj. 25:1
18 ◦C/km

Water den. Gas saturation


1

25:1
30 ◦C/km

Water den. Gas saturation

26:1
50 ◦C/km

Figure 4.  Left plot: A comparison of ­CO2 concentration profiles at the well location (r = 0) as a function of
depth after 100 years of injection for three temperature gradients of 18, 30 (reference case) and 50 °C/km. The
water to ­CO2 injection mass ratio is 25:1 for the 18, and 30 °C/km gradient, and 26:1 for the 50 °C/km gradient.
Center and right plots: 2D cross-sections for the water density and the gas saturation, which is the mass fraction
of exsolved ­CO2 in the pore fluids, for the three temperature gradients after 50 and 100 years of carbonated
water injection.

The pressure buildup in these systems was insignificant due to the large injection zone, open-flow boundary
conditions, and the assumed porosity and permeability conditions. The temperature near the well decreases over
time due to cooling by the injected carbonated water, which improves the storage capacity during injection. The
gas saturation profiles in Fig. 3 show that part of the C
­ O2 exsolves at the top of the region containing saturated
carbonated water, then redissolves as it moves into the non-CO2 saturated zones higher up in the system.
Furthermore, we modeled the fate of the injected C ­ O2 post-injection. We monitored the fluid flow for
100 years after the 100-year injection period, as presented in Fig. S2. After the end of the 100-year injection, the
carbonated water sinks towards the bottom of the system then is transported away due to the open boundary
­ f55.
conditions. This result is consistent with those o
The water to ­CO2 mass injection ratio for this homogeneous hydrostatically pressured reference system is
25:1. This ratio was the maximum injection ratio of this system, where the exsolved ­CO2 zone is stable and does
not reach the surface. Lowering the water to C ­ O2 ratio to 24:1 results in some of the exsolved C
­ O2 reaching the
surface after approximately 40 years of injection (see Fig. S3).

Impact of temperature gradient on the fate of injected gas‑charged ­CO2. The impact of the tem-
perature gradient of ­CO2 solubility storage was explored by considering gradients of 18, 30 (reference system),
and 50 °C/km. Figs. S5,S6 illustrate 2D cross-section time snapshots for the temperature, C ­ O2 mass fraction in
the liquid, gas saturation, and water density for the 30 and 50 °C/km gradient. Figure 4 compares the calculated
dissolved ­CO2 concentration versus depth profiles at the well location (r = 0) for the three temperature gradients.
Both the 18 and 30 °C/km gradient cases have the same injected water to ­CO2 ratio of 25:1. A less steep tempera-
ture gradient (18 °C/km) should improve the storage capacity in the reservoir as C ­ O2 is more soluble at lower
temperatures. However, Fig. 4 shows the top of the C ­ O2 concentration profiles versus depth is slightly deeper
for 30 °C/km compared to the 18 ◦C/km case, leading to marginally better storage security for the 30 °C/km
case. This counter-intuitive result stems from differing fluid densities; the carbonated water tends to sink more
rapidly in the lower temperature gradient system. The contrast between the density of the injected carbonated
water and the initial water in place increases with increasing temperature gradient, which accelerates the flow of
carbonated water downwards. This behavior results in a more concentrated zone of carbonated water near the
injection well.
For the lowest temperature gradient of 18 °C/km, the exsolved gas is only seen at the top of the injection zone,
as shown in Fig. 4. The exsolved gas zone expands deeper around the injection zone with an increasing tempera-
ture gradient. The increasing density gradient between the injected carbonated water and the formation fluids
due to increasing temperature gradients, pushes the injected water to travel more downwards than upwards or
vertically. Further, the ­CO2 solubility decreases with increasing temperature contributing to a solubility difference

Scientific Reports | (2022) 12:20452 | https://doi.org/10.1038/s41598-022-24623-6 7


Content courtesy of Springer Nature, terms of use apply. Rights reserved
Vol.:(0123456789)
www.nature.com/scientificreports/

Gas saturation

25:1
10,000-ppm

Gas saturation

50 years 100 years 25:1


of inj. of inj. 35,000-ppm

Gas saturation

35:1
100,000-ppm

Figure 5.  Left and center plots: Comparison of ­CO2 concentration profiles at the well versus depth after 50
and 100 years of injection for the injection of ­CO2 charged water having a seawater salinity with into reservoirs
having an initial salinity of 10,000, 35,000 (the reference case), and 100,000 ppm. The injection water to C ­ O2
mass ratio is 25:1 for the 10,000 and 35,000, ppm initial salinities, and 35:1 for the 100,000 ppm initial salinity.
Right plots: 2D cross-sections for the gas saturation, which is the mass fraction of exsolved C­ O2 in the pore
fluids, for the three salinities after 50 and 100 years of carbonated water injection.

between the zone cooled by the injected water and the surrounding reservoir water. This leads to increased C ­ O2
exsolution. In fact, the 50 ◦C/km gradient case required a slightly higher water to ­CO2 ratio of 26:1 to avoid the
leakage of exsolved ­CO2 at the surface.

Impact of subsurface temperature changes due to fluid injection. In general, if the injected water
is colder than the reservoir water, it results in a positive impact on ­CO2 solubility in the subsurface system due
to the retrograde solubility of ­CO2. A number of past studies have assumed that the temperature of the injected
water has no significant effect on the subsurface fluid temperatures, and thus the thermal effects were i­ gnored12,56.
The model simulations performed in this study indicate that the injection of fluid into a porous reservoir with
a temperature gradient of 30 ◦C/km or above will significantly alter the temperature near the injection well, as
shown in Fig. 3. A comparative calculation for a system neglecting the heating or cooling effect of injected fluids
was run to assess the impact of this assumption on computed results. The results of this calculation are shown in
Fig. S4. In the absence of formation fluid heating or cooling due to fluid injection, the minimum water to C ­ O2
mass ratio that can be injected while avoiding exsolved C ­ O2 escaping to the surface increases to 28:1. Thus, the
secure injection of ­CO2 into this system requires 12% more water per unit mass of C ­ O2 injected compared to the
reference system, which has a thermal gradient of 30 ◦C/km. In the case of injecting C ­ O2-charged water into a
system having an 18 ◦C/km temperature gradient, the average temperature in the injection zone is ~ 40 °C. Thus,
the isothermal system behaves similarly to that of the 18 ◦C/km temperature gradient ­system56. For a tempera-
ture gradient of 30 °C/km or above, however, ignoring thermal effects would lead to an overestimation of the
water needed to avoid ­CO2 leakage to the surface.

Impact of formation and injection water salinity. The effect of salinity on water-dissolved ­CO2 injec-
tion was studied by varying the initial salinity of the formation water from 35,000 ppm for the reference system,
to either 10,000 or 100,000 ppm. Two cases are considered in each of these systems. In the first, it is assumed that
the injected water has the same salinity as the reference system, which is 35,000 ppm. In the second, it is assumed
that the injected water has the same salinity as the reservoir fluids.
Figures S7 and S8 illustrate 2D cross-section time snapshots for the C ­ O2 mass fraction in the liquid, gas
saturation, water density, and salinity for the 10,000 and 100,000 ppm initial salinity systems when injecting ­CO2
charged water with a seawater salinity of 35,000 ppm. Figure 5 compares the model results for the three initial
salinities after 50 and 100 years of carbonated water injection. Lowering the salinity results in an improved storage
capacity as indicated by the ­CO2 concentration profiles in Fig. 5. However, the injected water to C­ O2 ratio was the
same for the 10,000- and the 35,000 ppm initial salinity calculations. Decreasing the ratio to 24:1 results in C ­ O2

Scientific Reports | (2022) 12:20452 | https://doi.org/10.1038/s41598-022-24623-6 8


Content courtesy of Springer Nature, terms of use apply. Rights reserved
Vol:.(1234567890)
www.nature.com/scientificreports/

gas escaping to the surface even for 10,000 ppm initial salinity. For a 100,000 ppm salinity, it was not possible to
inject 100,000 metric tons/year of ­CO2 for 100 years without ­CO2 gas escaping to the surface, as demonstrated in
the 2D cross-sections for the gas saturation after 100-years in Fig. 5. The density of 100,000 ppm saline water is
approximately 1050 kg/m3 near the injection zone, which is denser than that of the injected carbonated seawater
with a density of approximately 1020 kg/m3. This means that solubility trapping by injecting saline carbonated
water having a significantly lower salinity than the formation water is not an effective trapping mechanism, as
the injected carbonated water prefers to flow upwards towards the surface, as shown in Fig. 5. Reducing the
­CO2 concentration in the injected water does not help, as lowering the C ­ O2 concentration of the injected water
further reduces the density of the injected water and increases the convection upwards towards the surface. A
demonstration of carbonated water movement upwards during and after injection stops after 50 years of injec-
tion into 100,000 ppm saline reservoir can be seen in Fig. S9.
When the injected water has the same salinity as the formation water, the storage potential for the 10,000 ppm
system improves, and solubility storage in the 100,000 ppm system is possible by the injection of C ­ O2 charged
water. The minimum water to ­CO2 ratio waters having salinities of 10,000, and 100,000 ppm, respectively, with-
out ­CO2 escaping to the surface are 22:1 and 35:1, respectively. This reflects the decrease in C ­ O2 solubility due
to increasing s­ alinity57. The overall flow behavior for both systems is similar to the reference system. In each of
these cases, the carbonated water is denser and tends to flow downwards.

Impact of salinity gradient on carbon storage and security. The effect of the presence of a salin-
ity gradient in the reservoir on the fate of injected water-dissolved C ­ O2 injection is considered assuming the
presence of an initial 10 and 100 ppm/m salinity gradient. The results of these model calculations are shown
in Figs. S10,S11. A salinity gradient of 10 ppm/m results in a lower salinity throughout the reservoir compared
to the reference system. The storage capacity of this system is identical to that of the reference system as the
injection water has a salinity of 35,000 ppm, the same as the reference system. A salinity gradient of 100 ppm/m
results in significantly higher salinity with depth compared to the reference system, reaching 200,000 ppm at a
depth of 2000 m. As in the case of the lower salinity gradient system, the storage capacity in this system is dic-
tated by the salinity of the injected water (35,000 ppm). The high salinity gradient introduces a fluid density gra-
dient that increases with increased depth, which helps push the injected carbonated water downwards towards
the bottom of the system, as shown in Figs. S11,S12.
Figure 6 illustrates the fluid flow paths during and after carbonated water injection into a system with initial
constant 100,000 ppm salinity and for a system having a 100 ppm/m salinity gradient. Both during fluid injection
and after the injection, the majority of ­CO2 near the injection zone travels upwards in the constant 100,000-ppm
initial salinity system. In contrast, the majority of C
­ O2 travels downwards during injection for the 100 ppm/m
salinity gradient, and the ­CO2 solely travels downwards after the end of its injection.

Impact injection zone thickness and injection depth on carbon storage and security. The
impact of injection zone thickness was explored by considering thicknesses of 400- and 1200 m and comparing
model results to the reference system of 800 m. The results of these model calculations are shown in Figs. S13,S14.
Increasing the thickness of the injection zone to 1200 m does not improve the storage capacity compared to the
reference system. Decreasing the thickness of the injection zone to 400 m slightly reduces the storage capacity,
as it requires a 26:1 water to ­CO2 ratio, compared to 25:1 for the reference system to avoid the escape of C
­ O2 to
the surface. This is due to the increased pressure in the injection zone in the 400 m case, which drives more of
the carbonated water upwards towards the surface as shown in Fig. S13. Note that the risk of fracturing due to
pressure buildup may increase with decreased injection zone thickness.
Changing the injection depth has a clearer impact on storage potential compared to the thickness of the
injection zone. A deeper injection results in a better storage capacity due to the increased ­CO2 solubility with
increasing depth and the presence of an open flow boundary at the margins of the modelled system. For example,
injecting at a depth of 1000 to 1800 m (with an 800 m thick injection zone) improves the minimum water to ­CO2
ratio to 22:1, and injecting in at a depth of 1700 to 2000 m (a 300 m thick injection zone) improves the minimum
water to C­ O2 ratio to 19:1. The results of these model calculations are shown in Figs. S15,S16. In this case, the
­CO2 storage capacity is improved due to injection depth despite the smaller perforation thickness.

Impact of horizontal to vertical permeability ratio on carbon storage and security. The effect
of heterogeneous permeability on the fate of water dissolved C ­ O2 injection was assessed by altering the ratio
between the vertical and horizontal permeability ­(kv/kh) from the homogeneous reference system. Three cases
were considered: systems having a ­kv/kh of 0.1, 1 (reference system) and 2. To this end, the horizontal perme-
ability was held constant, and the vertical permeability was changed to 10 and 200 mD in the modelled systems.
Figure 7 shows a comparison of the fate of ­CO2 injected into the systems having the three permeability ratios.
The ­kv/kh = 0.1 system restricts fluid flow from moving in the vertical direction, creating more preferential flow
paths in the horizontal direction, as seen by the 2D cross-section for aqueous ­CO2% mass in Fig. 7. The preferred
horizontal movement results in a better storage security allowing for a reduced water to ­CO2 injection mass ratio
of 22:1. In addition, the enhanced horizontal movement might improve the effective storage capacity as larger
portion of the midsection of the reservoir is accessed compared with the reference system. In contrast, the ­kv/
kh = 2 system induces preferential flow in the vertical direction, resulting in enhanced C
­ O2 exsolution, as seen in
Fig. 7. The ­kv/kh = 2 system requires the same water to ­CO2 ratio as the reference system with ­kv/kh = 1 to avoid
­CO2 leakage to the surface. In this case, the denser C
­ O2-charged water enhances the flow path downwards, com-
pensating for the limited flow in the horizontal direction.

Scientific Reports | (2022) 12:20452 | https://doi.org/10.1038/s41598-022-24623-6 9


Content courtesy of Springer Nature, terms of use apply. Rights reserved
Vol.:(0123456789)
www.nature.com/scientificreports/

Figure 6.  Comparison of the dissolved ­CO2 flow paths during the final stages of injection and after the
injection has stopped. The left plots represent the flow paths in a system having a constant initial salinity of
100,000 ppm. The right plots represent the flow paths in a system with an initial salinity gradient of 100 ppm/m.
Carbonated water was injected for 50 years for the constant initial salinity of 100,000 ppm, and for 100 years
for the initial salinity gradient of 100 ppm/m. The non-blue colors highlight the higher flow rate zones, and the
arrows indicate the flow direction.

Discussion
The results described above suggest that the injection of water dissolved ­CO2 can provide significant and safe
long-term subsurface C ­ O2 storage even in the absence of traditional caprocks. Such results potentially open the
possibility of using unconfined subsurface systems for the solubility storage of ­CO2 injected as a water dissolved
phase around the world. Notably, the impact of most physical parameters changes little this result.
The impact of the various physical parameters considered in this study are Table 1. Most of the investigated
scenarios showed similar maximum injection water to ­CO2 mass ratios ranging from 20 to 30. The calculated
results suggest that the temperature and salinity of the injected ­CO2 charged water dictated the storage potential
more than these parameters in the original reservoir fluids. This behavior is because fluid injection cools and
dilutes the reservoir fluids over time. The storage security of the system, however, is highly dependent on the
initial reservoir temperature and salinity gradients. These factors impact the fluid flow direction of the injected
fluid. In the most extreme situation, the injection of a mildly saline ­CO2-charged water into a highly saline
reservoir could introduce upwards convective flow. The depth of the injection zone also plays a significant role

Scientific Reports | (2022) 12:20452 | https://doi.org/10.1038/s41598-022-24623-6 10


Content courtesy of Springer Nature, terms of use apply. Rights reserved
Vol:.(1234567890)
www.nature.com/scientificreports/

Gas saturation

22:1

kv/kh =0.1

Gas saturation

25:1
kv/kh=1

100 years Gas saturation


of inj.
25:1
kv/kh =2

Figure 7.  Left plot: Comparison of ­CO2 concentration profiles at the well versus depth after 100 years of
injection for systems having ­kv/kh ratios of 1:10, 1:1 and 2:1. The water to ­CO2 injection mass ratio is 22:1 for
the ­kv/kh ratio of 1:10, and 25:1 for the k­ v/kh ratio of 1:1 (reference system), and ­kv/kh = 2:1 as indicated in the
plots.Center and right plots: 2D cross-sections for the ­CO2 mass fraction and gas saturation, which is the mass
fraction of exsolved ­CO2 in the pore fluids, for the three modelled systems after 50 and 100 years of carbonated
water injection.

on the solubility storage potential, assuming hydrostatic p ­ ressure58. Injecting in a deeper interval improves the
­CO2 solubility, thus reducing the required amount of injection water required. Injecting deeper, however, might
introduce higher wellbore pressure because of potentially lower formation permeability and fluid friction effect
within the well, which was not accounted for in this study. Furthermore, the injectivity cost might increase with
increased injection d ­ epth59. The impact of reservoir heterogeneity in terms of permeability anisotropy showed
improved storage potential with higher horizontal to vertical ratio.
The calculations in this study illustrate the potential for subsurface solubility carbon storage in unreactive
unconfined aquifers by the injection of carbonated water. Note that the current study considered only a simplified
systems that had no provision for capillary or mineral trapping. These trapping mechanisms, when accounted
for, will provide additional storage potential. This choice was made in this study to focus directly on solubility
trapping as a primary storage mechanism to define the temperature, pressure, salinity, and permeability limits
on the fate of water dissolved C ­ O2 in the subsurface. Other factors could also come to play in real subsurface
systems. For instance, porosity often decreases with depth in sandstone ­reservoirs53,60. The rock permeability
tends to decrease as porosity ­decreases61. The effect of subsurface heterogeneities such as layering, salt bodies,
and dipping strata can have a profound effect on the subsurface flow dynamicsand storage e­ fficiency62. Note
that all calculations performed in this study assumed the original formation waters are ­CO2 free. Most natural
subsurface aquifers, however, contain at least some dissolved C ­ O263. Other factors not considered directly in the
calculations presented in this study include the effects of regional flow patterns, wellbore pressure, and subsurface
mineral-fluid interaction. Each factor may have an effect on subsurface carbon storage mechanisms. The impact
of these processes should be considered and examined in detail for each individual subsurface system scenario.
The capital and operational costs and dissolution efficiency are key factors when assessing different engineer-
ing solutions, including the injection of carbonated water into the ­subsurface13. The dissolution of ­CO2 into water
at the surface, followed by its injection into the subsurface has been considered by Burton and B ­ ryant11. Their
analysis concluded that compared to supercritical C ­ O2 injection, the capital costs increase by 60%, and additional
energy consumption is anticipated during operation for surface C ­ O2 dissolution and the subsequent injection
of this fluid. This process was modified and implemented as part of the CarbFix2 project, where above surface
water dissolution was used to capture ­CO2 from an impure exhaust ­stream64. In this case, the use of surface dis-
solution yielded substantial cost savings compared to alternative ­CO2 capture approaches. Alternatively, wellbore
dissolution of C ­ O2 into water has been proposed and implemented as part of the original CarbFix1 p ­ roject14,65.
This approach combines the engineering benefits of surface dissolution while avoiding most extra costs for cases
where a pure C ­ O2 stream is a­ vailable13. Wellbore dissolution, compared to supercritical C ­ O2 injection, has a lower
operating cost per well with up to one-third less energy consumed during ­compression66. In some cases, ­CO2
can just be added to existing waste water streams. This could lead to subsurface solubility storage of ­CO2 with
limited additional cost. There are more than 180,000 wells for oil and gas applications in the United States and
more than 500,000 wells for other a­ pplications67. More than 9 million m ­ 3/day of waste water from oil and gas

Scientific Reports | (2022) 12:20452 | https://doi.org/10.1038/s41598-022-24623-6 11


Content courtesy of Springer Nature, terms of use apply. Rights reserved
Vol.:(0123456789)
www.nature.com/scientificreports/

operation are injected in the United States. This corresponds to more than 3 Gigatons of fluid per y­ ear68. Adding
­CO2 to the existing waste water streams in the United States alone has the potential to store tens to hundreds of
megatons of C ­ O2 annually at little extra cost.
Alternatively, dedicated wells could be drilled for the injection of C
­ O2 charged water for subsurface solubility
storage. The injection of C­ O2 charged water requires injection of a substantially larger mass than the injection of
a pure ­CO2 stream. This will likely limit the overall mass of ­CO2 that can be injected into each well, so therefore
more wells would be required for this carbon storage a­ pproach17. Bodnar et al.69, however, concluded that the
formation volume needed to store dissolved ­CO2 is less than the formation volume needed to store equivalent
amount as supercritical ­CO2. Moreover, Burton and ­Bryant70, suggested that less net formation volume is required
to extract and re-inject formation water with dissolved ­CO2 compared to net formation volume needed for
separate ­scCO2 phase. Burton and B ­ ryant11 demonstrated a theoretical case study on how to store ­CO2 emissions
from a 500-MW power plant near Mt. Simon Formation in the US. They suggested the use of 50 injection and 50
extraction wells to store 3.65 Mt C ­ O2 annually, equivalent to 90% of the power plants emissions. At least some
of the costs of such required additional wells may be overcome because (1) the depth of ­CO2 injection can be
less for the injection when this gas is in the aqueous phase as opposed to a supercritical phase, and (2) the injec-
tion into hydrostatically pressured systems require less compression or wells that can support high p ­ ressure71.
The model calculations in the present study are based on the injection of 100,000 metric tons of ­CO2 into the
subsurface annually. The maximum injection rate will be site dependent, which is a function of depth, wellbore
size, reservoir connectivity, permeability, and porosity of the subsurface formation.. Nevertheless, an annual
injection of 100,000 metric tons C ­ O2 is approximately equivalent to the total emissions of a reverse-osmosis
desalination plant, producing desalinated water at a rate of 30 million cubic meters per y­ ear72,73. Furthermore,
according to the 2021 Global Status of CCS ­Report74, a significant number of operating a planned CCS facilities
that have an annual injection capacity of around 100,000 metric tons C ­ O2. For example, an annual operational
­CO2 storage capacity of 59,000 metric tons is reported for a C ­ O2 EOR facility in Hungary, and an annual opera-
tional ­CO2 storage capacity of 100,000 metric tons, for Karamay Dunhua Oil Technology CCUS EOR project in
China. The list also includes five planned Biorefinery Carbon Capture and Storage project in the US, all with a
projected annual storage capacity around 100,000 metric tons. Multiple wells would be required to store larger
quantities of C ­ O2 from larger emission sources. For example, the Gorgon Carbon Dioxide Injection project in
Australia includes 17 ­wells75. However, large emission sources often have numerous existing water disposals wells
nearby that can potentially be incorporated in the carbon storage strategy to reduce the initial ­cost68.

Conclusions
The presented results encourage further consideration of solubility trapping in unconfined subsurface aquifers,
potentially unlocking significant ­CO2 storage capacity. This would make ­CO2 storage more accessible at different
locations around the world in systems without caprocks. In this study, the only scenario in which the injected
­CO2 was not securely stored was the case in which the injected carbonated sea water was of much lower density
than the in-situ water. All other scenarios investigated suggest that there is substantial potential for the safe
long-term storage of ­CO2 by solubility trapping.
The injection of C­ O2-charged water has the advantage of requiring less compression and shallower, cheaper
wells, thus resulting in a comparable storage cost to the injection of supercritical C
­ O2. The need for large volumes
of water can be overcome by using seawater or withdrawing and reinjecting in-situ water from the same forma-
tion. Furthermore, the withdrawal wells can be used to monitor the fate of the injected ­CO2.
The process of water-dissolved ­CO2 injection into hydrostatically pressured subsurface systems eliminates
the need for a structural or stratigraphic seal, assuming fixed pressure open boundaries. Unconfined reservoirs
are more abundant than confined reservoirs and can potentially be found in closer proximity to major carbon
emitting sources. This significantly increases the number of potential subsurface carbon storage sites around the
world. Furthermore, solubility trapping promotes carbon mineralization resulting in higher storage potential.
Although the structures and hydraulics of many hydrostatically pressured saline aquifers have not yet been
studied in detail, the fact that water-charged C
­ O2 tends to sink overcomes the need for extended knowledge on
the subsurface geology prior to injection into a potential storage site. Overall, the results of this work further
compel the consideration of carbonated water injection into unconfined aquifers, a generally overlooked potential
carbon storage host, for the long-term storage of carbon-dioxide.

Data availability
The datasets used and/or analyzed during the current study available from the corresponding author on reason-
able request.

Received: 6 September 2022; Accepted: 17 November 2022

References
1. Friedlingstein, P. et al. Global carbon budget 2020. Earth System Science Data 12, 3269–3340 (2020).
2. Oelkers, E. H. & Cole, D. R. Carbon dioxide sequestration a solution to a global problem. Elements 4, 305–310 (2008).
3. IPCC et al. Climate Change 2021: The Physical Science Basis. Contribution of Working Group I to the Sixth Assessment Report of the
Intergovernmental Panel on Climate Change. Cambridge University Press (2021).
4. Cuéllar-Franca, R. M. & Azapagic, A. Carbon capture, storage and utilisation technologies: A critical analysis and comparison of
their life cycle environmental impacts. J. CO2 Util. 9, 82–102 (2015).
5. Gibbins, J. & Chalmers, H. Carbon capture and storage. Energy Policy 36, 4317–4322 (2008).
6. Orr, F. M. Onshore geologic storage of C ­ O2. Science 325, 1656–1658 (2009).

Scientific Reports | (2022) 12:20452 | https://doi.org/10.1038/s41598-022-24623-6 12


Content courtesy of Springer Nature, terms of use apply. Rights reserved
Vol:.(1234567890)
www.nature.com/scientificreports/

7. Bachu, S. et al. ­CO2 storage capacity estimation: Methodology and gaps. Int. J. Greenh. Gas Control 1, 430–443 (2007).
8. Gislason, S. R. & Oelkers, E. H. Carbon storage in basalt. Science 344, 373–374 (2014).
9. Bachu, S. Review of C ­ O2 storage efficiency in deep saline aquifers. Int. J. Greenh. Gas Control 40, 188–202 (2015).
10. Eke, P. E., Naylor, M., Haszeldine, S. & Curtis, A. ­CO2/brine surface dissolution and injection: ­CO2 storage enhancement. SPE
Projects Facil. Constr. 6, 41–53 (2011).
11. Burton, M. & Bryant, S. L. Eliminating buoyant migration of sequestered C ­ O2 through surface dissolution: Implementation costs
and technical challenges. SPE Reservoir Eval. Eng. 12, 399–407 (2009).
12. Shariatipour, S. M., Mackay, E. J. & Pickup, G. E. An engineering solution for ­CO2 injection in saline aquifers. Int. J. Greenh. Gas
Control 53, 98–105 (2016).
13. Emami-Meybodi, H., Hassanzadeh, H., Green, C. P. & Ennis-King, J. Convective dissolution of C ­ O2 in saline aquifers: Progress
in modeling and experiments. Int. J. Greenh. Gas Control 40, 238–266 (2015).
14. Sigfusson, B. et al. Solving the carbon-dioxide buoyancy challenge: The design and field testing of a dissolved C ­ O2 injection system.
Int. J. Greenh. Gas Control 37, 213–219 (2015).
15. Gilfillan, S. M. V. et al. Solubility trapping in formation water as dominant ­CO2 sink in natural gas fields. Nature 458, 614–618
(2009).
16. Riaz, A. & Cinar, Y. Carbon dioxide sequestration in saline formations: Part I—Review of the modeling of solubility trapping. J.
Petrol. Sci. Eng. 124, 367–380 (2014).
17. Snæbjörnsdóttir, S. Ó. et al. Carbon dioxide storage through mineral carbonation. Nature Rev. Earth Environ. 1, 90–102 (2020).
18. Pool, M., Carrera, J., Vilarrasa, V., Silva, O. & Ayora, C. Dynamics and design of systems for geological storage of dissolved C ­ O2 .
Adv. Water Resour. 62, 533–542 (2013).
19. McGrail, B. P. et al. Potential for carbon dioxide sequestration in flood basalts. J. Geophys. Res. https://d​ oi.o
​ rg/1​ 0.1​ 029/2​ 005JB
​ 0041​
69 (2006).
20. Oelkers, E. H. Geochemical aspects of C ­ O2 sequestration. Chem. Geol. 217, 183–186 (2005).
21. Liu, Q., Benitez, M. D., Xia, Z. & Santamarina, J. C. Pore-scale phenomena in carbon geological storage (Saline aquifers—Miner-
alization—Depleted oil reservoirs). Front. Energy Res. https://​doi.​org/​10.​3389/​fenrg.​2022.​979573 (2022).
22. Clavijo, S. P., Addassi, M., Finkbeiner, T. & Hoteit, H. A coupled phase-field and reactive-transport framework for fracture propa-
gation in poroelastic media. Sci. Rep. 12, 17819 (2022).
23. Suekane, T., Nobuso, T., Hirai, S. & Kiyota, M. Geological storage of carbon dioxide by residual gas and solubility trapping. Int. J.
Greenh. Gas Control 2, 58–64 (2008).
24. Duan, Z. & Sun, R. An improved model calculating ­CO2 solubility in pure water and aqueous NaCl solutions from 273 to 533 K
and from 0 to 2000 bar. Chem. Geol. 193, 257–271 (2003).
25. Zhaoa, H., Fedkin, M. V., Dilmore, R. M. & Lvov, S. N. Carbon dioxide solubility in aqueous solutions of sodium chloride at geo-
logical conditions: Experimental results at 323.15, 373.15, and 423.15 K and 150 bar and modeling up to 573.15 K and 2000 bar.
Geochim. Cosmochim. Acta 149, 165–189 (2015).
26. Ennis-King, J. & Paterson, L. Rate of dissolution due to convective mixing in the underground storage of carbon dioxide. In
Greenhouse Gas Control Technologies-6th International Conference 1653–1656 (Elsevier, 2003).
27. Xu, T. et al. Toughreact version 2.0: A simulator for subsurface reactive transport under non-isothermal multiphase flow condi-
tions. Comput. Geosci. 37, 763–774 (2011).
28. Nghiem, L., Sammon, P., Grabenstetter, J. & Ohkuma, H. Modeling CO2 storage in aquifers with a fully-coupled geochemical
EOS compositional simulator. In Proceedings—SPE Symposium on Improved Oil Recovery (Society of Petroleum Engineers (SPE),
2004).
29. Hoteit, H. & Firoozabadi, A. Numerical modeling of diffusion in fractured media for gas injection and recycling schemes. In SPE
Annual Technical Conference and Exhibition (Society of Petroleum Engineers, 2006).
30. Xu, T. & Pruess, K. Modeling multiphase non-isothermal fluid flow and reactive geochemical transport in variably saturated
fractured rocks: 1 Methodology. Am. J. Sci. 301(1), 16–33 (2001).
31. Darcy, H. P. G. Les Fontaines publiques de la ville de Dijon. Exposition et application des principes à suivre et des formules à employer
dans les questions de distribution d’eau, etc. (V. Dalamont, 1856).
32. Fick, A. Ueber diffusion. Ann. Phys. 170, 59–86 (1855).
33. Pruess, K., Oldenburg, C. & Moridic, G. TOUGH2 User’s Guide Version 2.0. Report LBNL-43134, Lawrence Berkeley National
Laboratory, California Lawrence Berkeley National Laboratory 210 (1999).
34. Spycher, N., Pruess, K. & Ennis-King, J. C ­ O2-H2O mixtures in the geological sequestration of C ­ O2. I. Assessment and calculation
of mutual solubilities from 12 to 100°C and up to 600 bar. Geochim. Cosmochim. Acta 67, 3015–3031 (2003).
35. Omar, A., Addassi, M., Vahrenkamp, V. & Hoteit, H. Co-optimization of C ­ O2 storage and enhanced gas recovery using carbonated
water and supercritical ­CO2. Energies 14, 7495 (2021).
36. Omar, A. A., Addassi, M., Hoteit, H. & Vahrenkamp, V. A new enhanced gas recovery scheme using carbonated water and super-
critical ­CO2. SSRN Electron. J. https://​doi.​org/​10.​2139/​ssrn.​38221​40 (2021).
37. Peng, D.-Y. & Robinson, D. B. A new two-constant equation of state. Ind. Eng. Chem. Fundam. 15, 59–64 (1976).
38. Henry, W. I. I. I. Experiments on the quantity of gases absorbed by water, at different temperatures, and under different pressures.
Philos. Trans. R. Soc. Lond. 93, 29–274 (1803).
39. Spycher, N. & Pruess, K. ­CO2-H2O mixtures in the geological sequestration of ­CO2. II. Partitioning in chloride brines at 12–100°C
and up to 600 bar. Geochim. Cosmochim. Acta 69, 3309–3320 (2005).
40. Pruess, K. & Spycher, N. ECO2N–A fluid property module for the T ­ OUGH2 code for studies of C ­ O2 storage in saline aquifers.
Energy Convers. Manage. 48, 1761–1767 (2007).
41. Spycher, N., Pruess, K., Spycher, N. & Pruess, K. A phase-partitioning model for ­CO2–brine mixtures at elevated temperatures and
pressures: Application to C ­ O2-enhanced geothermal systems. Springer 82, 173–196 (2010).
42. Harvey, A. H. Semiempirical correlation for Henry’s constants over large temperature ranges. AIChE J. 42, 1491–1494 (1996).
43. Li, Y. & Nghiem, L. X. Phase equilibria of oil, gas and water/brine mixtures from a cubic equation of state and Henry’s law. Can.
J. Chem. Eng. 64, 486–496 (1986).
44. Grant, M. A. & Bixley, P. F. Chapter 5—Downhole measurement. In Geothermal Reservoir Engineering 2nd edn (eds Grant, M. A.
& Bixley, P. F.) 75–92 (Academic Press, 2011). https://​doi.​org/​10.​1016/​B978-0-​12-​383880-​3.​10005-8.
45. Gul, S. & Aslanoglu, V. Drilling and Well Completion Cost Analysis of Geothermal Wells in Turkey. In 43rd workshop on geothermal
reservoir engineering (2018).
46. Hoteit, H., He, X., Yan, B. & Vahrenkam, V. Optimization and uncertainty quantification model for time-continuous geothermal
energy extraction undergoing re-injection. arXiv Preprint https://​doi.​org/​10.​48550/​arxiv.​2112.​05544 (2021).
47. Kaya, E., Zarrouk, S. J. & O’Sullivan, M. J. Reinjection in geothermal fields: A review of worldwide experience. Renew. Sustain.
Energy Rev. 15, 47–68 (2011).
48. Bertani, R. Geothermal power generation in the world 2005–2010 update report. Geothermics 41, 1–29 (2012).
49. Čermák, V. & Rybach, L. Thermal conductivity and specific heat of minerals and rocks. In Landolt-Börnstein: Numerical Data and
Functional Relationships in Science and Technology, New Series, Group V (Geophysics and Space Research), Volume Ia, (Physical
Properties of Rocks) (ed. Angenheister, G.) 305–343 (Springer, 1982).

Scientific Reports | (2022) 12:20452 | https://doi.org/10.1038/s41598-022-24623-6 13


Content courtesy of Springer Nature, terms of use apply. Rights reserved
Vol.:(0123456789)
www.nature.com/scientificreports/

50. Appelo, C. A. J., Parkhurst, D. L. & Post, V. E. A. Equations for calculating hydrogeochemical reactions of minerals and gases such
as ­CO2 at high pressures and temperatures. Geochim. Cosmochim. Acta 125, 49–67 (2014).
51. Koschel, D., Coxam, J.-Y., Rodier, L. & Majer, V. Enthalpy and solubility data of CO2 in water and NaCl(aq) at conditions of interest
for geological sequestration. Fluid Phase Equilib. 247, 107–120 (2006).
52. Addassi, M., Omar, A., Ghorayeb, K. & Hoteit, H. Comparison of various reactive transport simulators for geological carbon
sequestration. Int. J. Greenh. Gas Control 110, 103419 (2021).
53. Ehrenberg, S. N. & Nadeau, P. H. Sandstone vs. carbonate petroleum reservoirs: A global perspective on porosity-depth and
porosity-permeability relationships. AAPG Bull. 89, 435–445 (2005).
54. Saar, M. O. & Manga, M. Permeability-porosity relationship in vesicular basalts. Geophys. Res. Lett. 26, 111–114 (1999).
55. Lyle, S., Huppert, H. E., Hallworth, M., Bicle, M. & Chadwick, A. Axisymmetric gravity currents in a porous medium. J. Fluid
Mech. 543, 293–302 (2005).
56. Javaheri, M. et al. Linear stability analysis of double-diffusive convection in porous media, with application to geological storage
of CO2. Transp. Porous Media 84(2), 441–456 (2009).
57. Gunter, W. D., Perkins, E. H. & McCann, T. J. Aquifer disposal of CO2-rich gases: Reaction design for added capacity. Energy
Convers. Manage. 34, 941–948 (1993).
58. Kampman, N., Bickle, M., Wigley, M. & Dubacq, B. Fluid flow and CO2–fluid–mineral interactions during CO2-storage in sedi-
mentary basins. Chem. Geol. 369, 22–50 (2014).
59. Shevenell, L. The estimated costs as a function of depth of geothermal development wells drilled in Nevada. GRC Trans. 36, 121–128
(2012).
60. Medina, C. R., Rupp, J. A. & Barnes, D. A. Effects of reduction in porosity and permeability with depth on storage capacity and
injectivity in deep saline aquifers: A case study from the Mount Simon Sandstone aquifer. Int. J. Greenh. Gas Control 5, 146–156
(2011).
61. Nelson, P. H. Permeability-porosity relationships in sedimentary rocks. The log analyst 35(3), 38–61 (1994).
62. Kitanidis, P. K. Persistent questions of heterogeneity, uncertainty and scale in subsurface flow and transport. Water Resour. Res.
51, 5888–5904 (2015).
63. De Silva, G. P. D., Ranjith, P. G. & Perera, M. S. A. Geochemical aspects of CO2 sequestration in deep saline aquifers: A review.
Fuel 155, 128–143 (2015).
64. Gunnarsson, I. et al. The rapid and cost-effective capture and subsurface mineral storage of carbon and sulfur at the CarbFix2 site.
Int. J. Greenh. Gas Control 79, 117–126 (2018).
65. Paterson, L. et al. Overview of the CO2CRC Otway residual saturation and dissolution test. Energy Procedia 37, 6140–6148 (2013).
66. Shafaei, M. J., Abedi, J., Hassanzadeh, H. & Chen, Z. Reverse gas-lift technology for CO2 storage into deep saline aquifers. Energy
45, 840–849 (2012).
67. EPA. Protecting underground sources of drinking water and public health. (2020).
68. US EPA. Class I Industrial and Municipal Waste Disposal Wells. https://​www.​epa.​gov/​uic/​class-i-​indus​trial-​and-​munic​ipal-​waste-​
dispo​sal-​wells. Accessed 10 June 2022.
69. Bodnar, R. J. et al. PVTX properties of H2O-CO2-“salt” at PTX conditions applicable to carbon sequestration in saline formations.
Rev. Mineral. Geochem. 77, 123–152 (2013).
70. Burton, M. & Bryant, S. L. Surface dissolution: Minimizing groundwater impact and leakage risk simultaneously. Energy Procedia
1, 3707–3714 (2009).
71. Bai, M., Sun, J., Song, K., Li, L. & Qiao, Z. Well completion and integrity evaluation for CO2 injection wells. Renew. Sustain. Energy
Rev. 45, 556–564 (2015).
72. Cornejo, P. K., Santana, M. V. E., Hokanson, D. R., Mihelcic, J. R. & Zhang, Q. Carbon footprint of water reuse and desalination:
A review of greenhouse gas emissions and estimation tools. J. Water Reuse Desal. 4, 238–252 (2014).
73. Hamieh, A. et al. Quantification and analysis of CO2 footprint from industrial facilities in Saudi Arabia. Energy Convers. Manag.
16, 100299 (2022).
74. Global Status Report 2021. Global CCS Institute https://​www.​globa​lccsi​nstit​ute.​com/​resou​rces/​global-​status-​report/ (2021).
Accessed 10 June 2022.
75. Trupp, M., Ryan, S., Barranco Mendoza, I., Leon, D. & Scoby-Smith, L. Developing the world’s largest CO2 injection system—A
history of the gorgon carbon dioxide injection system. Proc. 15th Greenh. Gas Control Technol. Conf. https://​doi.​org/​10.​2139/​ssrn.​
38154​92 (2021).

Acknowledgements
The authors thank King Abdullah University of Science and Technology (KAUST) for supporting this work.
Thanks to LBNL and CMG Ltd. For providing the simulators TOUGHREACT and GEM, respectively.

Author contributions
M.A. and E.O. wrote the main manuscript. All authors reviewed the manuscript.

Competing interests
The authors declare no competing interests.

Additional information
Supplementary Information The online version contains supplementary material available at https://​doi.​org/​
10.​1038/​s41598-​022-​24623-6.
Correspondence and requests for materials should be addressed to M.A.
Reprints and permissions information is available at www.nature.com/reprints.
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and
institutional affiliations.

Scientific Reports | (2022) 12:20452 | https://doi.org/10.1038/s41598-022-24623-6 14


Content courtesy of Springer Nature, terms of use apply. Rights reserved
Vol:.(1234567890)
www.nature.com/scientificreports/

Open Access This article is licensed under a Creative Commons Attribution 4.0 International
License, which permits use, sharing, adaptation, distribution and reproduction in any medium or
format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the
Creative Commons licence, and indicate if changes were made. The images or other third party material in this
article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the
material. If material is not included in the article’s Creative Commons licence and your intended use is not
permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from
the copyright holder. To view a copy of this licence, visit http://​creat​iveco​mmons.​org/​licen​ses/​by/4.​0/.

© The Author(s) 2022

Scientific Reports | (2022) 12:20452 | https://doi.org/10.1038/s41598-022-24623-6 15


Content courtesy of Springer Nature, terms of use apply. Rights reserved
Vol.:(0123456789)
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for small-
scale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:

1. use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
2. use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
3. falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
4. use bots or other automated methods to access the content or redirect messages
5. override any security feature or exclusionary protocol; or
6. share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at

onlineservice@springernature.com

You might also like