You are on page 1of 13

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/271190777

Simulations of the Impact of Co-injected Gases on CO2 Storage, the SIGARRR


Project: First Results on Water-gas Interactions Modeling

Article  in  Energy Procedia · December 2014


DOI: 10.1016/j.egypro.2014.11.341

CITATIONS READS

7 2,254

18 authors, including:

Jerome Corvisier Elise el ahmar


MINES ParisTech, France MINES ParisTech
61 PUBLICATIONS   1,094 CITATIONS    43 PUBLICATIONS   252 CITATIONS   

SEE PROFILE SEE PROFILE

Christophe Coquelet Jérôme Sterpenich


MINES ParisTech University of Lorraine
261 PUBLICATIONS   3,086 CITATIONS    87 PUBLICATIONS   887 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Partition coefficient of mercaptans and hydrocarbon solubilities in alkanolamine solutions View project

Kinetic study of nitric oxide oxidation View project

All content following this page was uploaded by Jérôme Sterpenich on 21 January 2015.

The user has requested enhancement of the downloaded file.


Available online at www.sciencedirect.com

ScienceDirect
Energy Procedia 63 (2014) 3160 – 3171

GHGT-12

Simulations of the impact of co-injected gases on CO2 storage, the


SIGARRR project: first results on water-gas interactions modeling
J. Corvisiera*, E. El Ahmarb, C. Coqueletb, J. Sterpenichc, R. Privatd, J.-N. Jaubertd,
K. Ballerat-Busserollese, J.-Y. Coxame, P. Cézacf, F. Contaminef, J.-P. Serinf,
V. Lachetg, B. Cretong, M. Parmentierh, P. Blanch, L. Andréh, L. de Laryh, E.C. Gaucheri
a
MINES ParisTech, Centre de Géosciences, 35 rue St Honoré,77305 Fontainebleau, France
b
MINES ParisTech, Centre Thermodynamic of Processes, 35 rue St Honoré, 77305 Fontainebleau, France
c
Université de Lorraine, GeoRessources UMR/CNRS 7359, BP70239 Faculté des Sciences et Technologies, 54506 Vandoeuvre-lès-Nancy, France
d
Université de Lorraine, LRGP-ENSIC UMR/CNRS 7274, BP20451 1 rue Granville, 54001 Nancy, France
e
Université Blaise Pascal Clermont-Ferrand 2, ICCF UMR/CNRS 6296, 24 avenue des Landais, 63177 Aubière, France
f
Université de Pau et des Pays de l'Adour, ENSGTI LaTEP EA 1932, BP7511 rue Jules Ferry, 64075 Pau, France
g
IFPEN, 1 & 4 avenue de Bois-Préau, 92852 Rueil-Malmaison, France
h
BRGM, 3 avenue Claude Guillemin BP 36009, 45060 Orléans, France
i
TOTAL, CSTJF, avenue Larribau, 64018 Pau, France

Abstract

When capturing CO2, the collected gas mixture can vary considerably both qualitatively and quantitatively, based on the CO2
origin, the capture process and the industrial sector. Co-injected with CO2, these impurities might be an issue in case of leakage
but may also impact the subsurface storage. Operators of the whole CCS chain are therefore waiting for recommendations in
terms of admissible concentrations while regulators are waiting for tools allowing them to formulate these recommendations. The
SIGARRR project aims at proposing accurate reactive-transport simulations to model the long-term behavior of CO2 and its co-
injected gases within storage sites focusing on the reactivity with reservoirs and the possible inferences on the environment. The
paper presents first numerical results on water-gas equilibriums in agreement with experiments.
©
© 2013 TheAuthors.
2014 The Authors.Published
Publishedbyby Elsevier
Elsevier Ltd.Ltd.
This is an open access article under the CC BY-NC-ND license
Selection and peer-review under responsibility of GHGT.
(http://creativecommons.org/licenses/by-nc-nd/3.0/).
Peer-review under responsibility of the Organizing Committee of GHGT-12
Keywords: water/gas interactions; geochemical modeling; CO2; O2; SO2; N2; CH4.

* Corresponding author. Tel.: +3-316-469-4750 ; fax: +3-316-469-4713 .


E-mail address: jerome.corvisier@mines-paristech.fr

1876-6102 © 2014 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/3.0/).
Peer-review under responsibility of the Organizing Committee of GHGT-12
doi:10.1016/j.egypro.2014.11.341
J. Corvisier et al. / Energy Procedia 63 (2014) 3160 – 3171 3161

1. Introduction and the SIGARRR project

An important part of the total induced CCS cost occurs during the capture, within the separation step, where CO2
is dissociated from other gas components. The obtained gas mixture composition can vary considerably both
qualitatively and quantitatively, based on the CO2 origin, the capture process and the industrial sector. Indeed, in
addition to CO2, several components, including O2, N2, SOx, H2S, NyOx, H2, CO, and Ar, can be present at various
concentration levels [1]. Moreover, these impurities can have an impact within the subsurface storage (water,
reservoir rocks, cap-rocks, and wells). Some of the gas (i.e. SOx, H2S, NyOx, CO) are toxic for human health and the
environment, even at low concentrations and can be an issue in case of leakage. Operators of the whole CCS chain
are therefore waiting for clear recommendations in terms of admissible concentration levels for the various co-
injected impurities while regulators are waiting for tools allowing them to formulate these recommendations. An
increase of the impurities concentrations can reduce the capture costs and this will help to deploy more rapidly the
CCS. Testing scenarios with accurate reactive-transport codes should enable to propose these awaited precise
recommendations, but these models have to be validated and calibrated regarding laboratory experiments.

The role of these gas impurities on the geochemical behavior of the storage sites is presently not very well
documented. Pure CO2 solubility experiments and thermodynamic models are numerous but there are only few
results on CO2 containing mixtures (e.g. mainly for CH4 and H2S). “Gaz Annexes”, a project funded by the French
National Agency for Research (i.e. ANR) between 2006 and 2011 focused on this issue. The database for gas
solubilities is also strongly incomplete at high pressure / high temperature / high salinity. Data for complex salts are
clearly missing. A part of this project devoted to experimental characterization of the impact of gas mixture on the
reactivity of minerals within geological repositories and led to important experimental/modeling results on
individual gases and even to first investigations on the water-gas mixture-rock system [2].

Based on a reinforced consortium, the gained technical experience and the laboratory equipment acquired from these
previous research efforts, a new project, “SIGARRR”, still funded by the ANR, started in late 2013. The main thrust
of purpose of this new project is to be abled to conduct precise geochemical simulations to model the long-term
behavior of co-injected gases within CO2 storage sites focusing on the:
• Impact of CO2 and co-injected gases (SO2, NO, O2) on the minerals and reservoir (silicates + clay
minerals) geochemistry,
• Possible inferences on the environment in case of leak.
Recent improvements in geochemical codes allow them to deal with non-ideal gas mixtures and to reproduce rather
accurately existing solubility measurements for CO2 and even mixtures [3-5].

This paper aims first at presenting these new models and the necessary associated parameters that can be collected
from the literature so far and then at showing the ability of these codes to represent accurately water-gas
equilibriums for CO2, O2, SO2, N2, CH4, CO2+N2 and CO2+CH4 over large temperature and pressure ranges of
interest.

2. Geochemical modeling for water-gas interactions

2.1. Water-gas equilibriums, dissymmetrical approach and mass action laws

In most geochemical codes, chemical reactions and their equilibriums are related to classical mass action laws
and water-gas equilibriums are generally treated using a dissymmetrical approach γ-ϕ (e.g. activity-fugacity). For
example, CO2 dissolution into water:

CO 2(g) ↔ CO 2(aq) (1)

implies the following mass action law:


3162 J. Corvisier et al. / Energy Procedia 63 (2014) 3160 – 3171

mCO2 ( aq )γ CO2 ( aq ) K CO2 ( g ) (T, p) = yCO2 ( g )ϕ CO2 ( g ) p (2)

with m the molality of the dissolved CO2, γ its activity coefficient, K the CO2 Henry’s constant (depending on
both temperature T and pressure p), y the gaseous CO2 molar fraction and ϕ its fugacity coefficient.
A more general form for each gaseous component g and its associated dissolved species aq could be written:

maqγ aq K g (T, p) = ygϕ g p (3)

2.2. Henry’s constants

As mentioned in these last equations, Henry’s constants vary with temperature and pressure. The temperature
dependence is taken into account using tabulated values of these equilibrium constants at water saturation pressure,
while the pressure correction is evaluated as follows [6]:

⎛ p − p sat ⎞
K g (T, p) = K g (T, p sat ) exp ⎜Vaq ⎟ (4)
⎝ RT ⎠

Vaq being the average partial molar volume of the dissolved gas over the pressure interval psat and p and also
over the temperature range of interest. Some authors prefer using the molar volume at infinite dilution.

2.3. Equations of state

Water-gas geochemical calculations require the resolution of an equation of state (EOS) for the gas phase
allowing the computation of the volume when the pressure is fixed (or the other way around). Using the perfect gas
equation is very convenient since the relation between volume and pressure is direct, nevertheless this simple EOS is
satisfactory only at low temperatures and pressures and therefore not in the CO2 storage sites conditions.
Considering non-ideal gases and in the particular case of cubic EOS, the pressure is also directly calculated while
the volume can be determined analytically using the Cardan’s method for example [3-7]. The Peng-Robinson
equation [8], one of the most spread cubic EOS, is now implemented in CHESS/HYTEC [3,4,9] and PHREEQC
[5,10] and can be written as:

p=
RT

a T ( ) (5)
V −b V V +b +b V −b
( ) ( )
R being the perfect gas constant, V the molar volume of the phase and, a and b, some parameters specific to
the Peng-Robinson equation and detailed as follows:

⎧ 2
R 2Tc2 ⎛⎜ ⎛ ⎞⎞
⎜1− T ⎟⎟

( )
⎪⎪ a T = 0.45724
pc ⎜⎝
(
1+ 0.37464 +1.54226ω − 0.26992ω 2 ) ⎜
⎝ Tc ⎟⎠⎟⎠
⎨ (6)
⎪ RT
⎪ b = 0.07780 c
⎪⎩ pc

where Tc is the critical temperature of the considered gas, pc its critical pressure and ω its acentric factor.

In the particular case of gas mixture, a mixing rule shall also be used to calculate the Peng-Robinson parameters for
J. Corvisier et al. / Energy Procedia 63 (2014) 3160 – 3171 3163

the mixture from the parameters of each compound and their molar fractions. The most simple is the classical
mixing rule:

⎧ mix
( )
⎪ a T = ∑∑ yi y j 1− kij

( ) a (T )
i ( )
aj T
i j
⎨ (7)
mix
⎪ b = ∑ yi bi
⎪⎩ i

with kij the binary interaction parameters accounting for the interactions between gaseous compounds i and j.

Following this approach, the gas fugacity of the compound i is given by:

bi ⎛ pV ⎞ ⎛ pV − pbmix ⎞ a mix ⎛ V + 2.414bmix ⎞


ln ϕ i = ⎜ −1⎟ − ln ⎜
⎜ ⎟
⎟ − ln ⎜⎜ ⎟
mix ⎟
bmix ⎝ RT ⎠ ⎝ RT ⎠ 2 2RTbmix ⎝ V + 0.414b ⎠
⎛ 2 y 1− k ⎞
⎜ ∑ j ( ij )
a j ai
bi ⎟
(8)
j
×⎜ − mix ⎟
⎜ a mix b ⎟
⎜ ⎟
⎝ ⎠

2.4. Aqueous activity models

Concerning the electrolytic part, very important for gas-water-salt systems, various models are available to
compute the aqueous activity coefficients. The simplest ones (Debye-Hückel, Davies, b-dot) based on long-distance
electrostatic interaction are valid for diluted solutions only. However, b-dot model [11] shows satisfactory for
simple electrolytes up to 3M. Above this limit, for highly saline solutions, non-electrostatic short distance
interactions could not be neglected. Pitzer model [12,13] and SIT model [14,15], respectively used in PHREEQC
and CHESS, allow to obtain good results for these solutions.

Pitzer model is based on a virial expansion of the excess Gibbs free energy and allows the expression of the aqueous
activity coefficients, notably for neutral species such as dissolved gases:

ln γ i = 2∑ mN λi−N + 2∑ mC λi−C + 2∑ mAλi−A + ∑∑ mC mAζ i−C−A (9)


N C A C A

with mN, mC and mA the molalities of neutral, cationic and anionic aqueous species respectively, λ the 2nd
order interaction parameters and ζ the 3rd order interaction parameters.

The Specific Ion-Interaction Theory (SIT), first mentioned by Bronsted [14], takes into account the short-distance
forces by adding terms to the classical Debye-Hückel law. Theoretical details of the approach are given in Grenthe’s
work [15]. For neutral species like dissolved gas, only the “SIT” terms are of interest, and the SIT model limits to:

log γ i = ∑ m jεi− j (10)


j

with εij interaction parameter between aqueous species i and j.

Comparing the two approaches, and their respective equations (9) and (10), SIT model can easily be seen as a sort of
Pitzer model where the 3rd order interactions are neglected. Applying this assumption and rearranging eq.(9) a little,
3164 J. Corvisier et al. / Energy Procedia 63 (2014) 3160 – 3171

a simple relation between Pitzer and SIT parameters can be written:

2λi− j
εi− j = (11)
ln10

This last relation is proven to be very useful to fill some blanks in databases converting Pitzer parameters into SIT
parameters.

3. Parameters from the literature

From the previous section, necessary models and subsequent parameters for a good representation of water-gas
interactions in geochemical codes were highlighted and can be summed up at this point:
1. Specific parameters for pure gases
• Critical temperature and pressure
• Acentric factor
2. Specific parameters for gas solubilities in water and saline solutions
• Henry’s constants (tabulated for different temperatures and at the water saturated pressure)
• Average partial molar volume for dissolved gas (in pressure and in temperature)
• Interaction parameters between the dissolved gases and various ions in solution (Pitzer or SIT)
3. Specific parameters for binary gas mixtures
• Binary interaction parameters
In this part, the corresponding parameters collected from the literature will be presented

3.1. Specific parameters for pure gases

Critical properties and acentric factors are the first parameters needed for water-gas interactions modeling. They
can be retrieved from many sources and all are in rather good agreement. The retained values are from Yaws’
Handbook [16] and are presented in tab.1

Table 1. Critical properties and acentric factors from [16].


Tc pc ω
Compound
[K] [bar] [-]
CO2 304.19 73.82 0.228
CH4 190.58 46.04 0.011
N2 126.10 33.94 0.040
O2 154.58 50.43 0.022
SO2 430.75 78.84 0.245
H2 O 647.13 220.55 0.345

3.2. Henry’s constant

Most important parameters in such models are probably Henry’s constants since they fix solubilities for gases.
Quite a lot of references can be found depending on the considered gases: experimental values and models of
various forms. Different models were then tested and only the retained ones are cited here. The evolutions of the
selected models for Henry’s constants for CO2 [17], O2 [17], SO2 [18], N2 [19] and CH4 [20] versus the temperature
and at saturation pressure are represented on a same figure (fig.1).
J. Corvisier et al. / Energy Procedia 63 (2014) 3160 – 3171 3165

Fig. 1. Models of Henry’s constants at water saturation pressure vs temperature for CO2, O2, SO2, N2 and CH4 (see text for references).

3.3. Molar volumes of dissolved gases

These molar volumes can be obtained from the revised HKF general equation (Helgeson-Kirkham-Flowers, not
detailed here) [21]. The parameters for this equation can be obtained from various sources of the literature [5,22,23].
These molar volumes can then be determined for chosen temperatures and pressures and the appropriate averages
computed. The evolutions of molar volumes for CO2 [23], O2 [22], SO2 [23], N2 [22] and CH4 [23] versus the
temperature and at saturation pressure are also represented on a same figure (fig.2).

Fig. 2. Molar volume at water saturation pressure (HKF model) vs temperature for CO2, O2, SO2, N2 and CH4 (see text for references).

3.4. Interaction parameters between dissolved gases and ions

These interactions coefficients can be found in various studies such as parameters fitted on experimental data at
25°C or as pressure-temperature laws. Values collected or computed at 25°C are summed up in tab.2 as well as
3166 J. Corvisier et al. / Energy Procedia 63 (2014) 3160 – 3171

associated references. These coefficients are not needed for this paper’s purpose, but for later perspectives and
furthermore this table will have to be upgraded.

Table 2. Pitzer interaction parameters for dissolved gas-ion at 25°C and associated references.
CO2 O2 SO2 N2 CH4
Compound
[24] [25] [26] [19] [20]
+
H - - - - -
Na+ 0.151 0.200 0.013 0.141 0.139
Cl− 0.000 0.000 0.000 0.000 0.000
CO2 0.035
SO2 0.017
Na+-Cl− −0.016 −0.013 −0.002 −0.006 −0.003

3.5. Binary interaction parameters for gases

Such parameters are obtained for experimental data on binary equilibriums and they depend on both the chosen
EOS and mixing rule. They are generally independent regarding temperature. Here again, some existing modeling
studies allow to retrieve most of the needed parameters and collected parameters are presented in tab.3 with the
associated references. Some blanks still remains in this table, but for the simulations presented later, the missing
coefficients are not needed.

Table 3. Binary interaction parameters between gases and associated references (Interactions are considered symmetrical, i.e.
kij = kji).
Compound CO2 O2 SO2 N2 CH4 H2 O
CO2 - 0.114 [27] 0.046 [27] −0.007 [27] 0.100 [27] 0.190 [28]
O2 - -
SO2 - - - −1.103 [29]
N2 - - - - 0.478 [28]
CH4 - - - - - 0.485 [28]

4. First simulations on existing data and validation of the approach

Now, that the whole modeling philosophy has been exposed, some simulations can be run in order to test codes
and even to validate our methodology, comparing the numerical results with existing experimental data. In this
purpose, two kinds of simulation will be run. Firstly, the solubility of binary pure gas + water systems will be
modeled. Secondly, two ternary gas mixture + water systems which have also been studied experimentally will be
modeled. In this section, all the simulations are performed with the geochemical code CHESS [3,4,9] developed by
MINES ParisTech, using the LLNL database (V8.R6) but with the and previously selected and presented Henry’s
constants.

4.1. Verification on pure gas+water systems

Some experimental solubility data exist for all the selected gases at various temperatures and pressures
depending on the compound itself. Calculations are run for 5 gases CO2, O2, SO2, N2 and CH4, for 6 temperatures
25, 50, 75, 100, 125 and 150°C, for pressures up to 500 bar. The aqueous molar faction of dissolved gas obtained for
all these simulations are presented on fig.3 as well as the collected experimental data.
J. Corvisier et al. / Energy Procedia 63 (2014) 3160 – 3171 3167

Fig. 3. Aqueous molar fraction of dissolved gas CO2, O2, SO2, N2 and CH4 (from top to bottom and from left to right) in pure water vs pressure at
different temperatures up to 150°C (symbols correspond to measurements from various sources and solid lines to CHESS simulations).

The obtained results are in rather good agreement for all these binary systems over the whole temperature and
3168 J. Corvisier et al. / Energy Procedia 63 (2014) 3160 – 3171

pressure ranges. For O2 and SO2, pressure has been limited regarding the available experimental data. There is
obviously a correlation between the quantity of available data and the accuracy of the modeling, linked to the simple
fact that the Henry’s constants comes from these data. Some supplementary data may have to be acquired on the
SO2+H2O system so that the Henry’s constant (and its dependency on temperature and pressure) could be more
accurate.

4.2. Validation on gas mixture+water systems

Now that these first simulations were satisfactory and helped selecting the right sets of parameters, some more
complex systems can be handled. Unfortunately, concerning gas mixture+water systems of interest, experimental
data are rather rare. Two interesting datasets were selected.

The first one concerns CO2+N2+H2O at 35 and 45°C, up 160 bar and 7 compositions (0.0, 43.5, 58.7, 74.3, 81.7,
88.3 and 100% mol. CO2) were tested [30]. Simulations are then run trying to reproduce these experimental data and
the calculated aqueous molar faction of both dissolved gas are presented on fig.4 and fig.5 along with the data from
[30].

Fig. 4. Aqueous molar fraction (A) CO2 and (B) N2 vs. pressure for CO2-N2-H2O system at 35°C (symbols correspond to measurements [30],
solid lines to CHESS simulations).

Fig. 5. Aqueous molar fraction (A) CO2 and (B) N2 vs. pressure for CO2-N2-H2O system at 45°C (symbols correspond to measurements [30],
solid lines to CHESS simulations).
J. Corvisier et al. / Energy Procedia 63 (2014) 3160 – 3171 3169

The obtained results are still in really good agreement for this ternary system for both dissolved CO2 and N2 at 35
and 45°C and over the whole pressure range up to 200 bar.

The second dataset is about CO2+CH4+H2O system at 51 and 102°C, up 500 bar and 5 compositions (0.0, 32.8, 59.1,
75.3 and 100% mol. CO2 at 51°C and 0.0, 41.4, 52.9, 71.5 and 100% mol. CO2 at 102°C) were studied [31]. The
calculated aqueous molar faction of both dissolved gas are presented on fig.6 and fig.7 along with the experimental
data from [31].

Fig. 6. Aqueous molar fraction (A) CO2 and (B) CH4 vs. pressure for CO2-CH4-H2O system at 51°C (symbols correspond to measurements [31],
solid lines to CHESS simulations).

Fig. 7. Aqueous molar fraction (A) CO2 and (B) CH4 vs. pressure for CO2-CH4-H2O system at 102°C (symbols correspond to measurements [31],
solid lines to CHESS simulations).

The numerical results are again in really good agreement for this other ternary system for both dissolved CO2 and
CH4 at 51 and 102°C and over the whole pressure range.
These two systems also demonstrate the direct impact of an impurity on CO2, decreasing significantly its solubility
and then its potential storage, even if the quantities involved here are far larger than those met when talking about
CO2 capture.
3170 J. Corvisier et al. / Energy Procedia 63 (2014) 3160 – 3171

5. Conclusion and perspectives

As conclusion, our approach is very satisfactory regarding the simulated results compared to the literature data.
For all the considered systems, numerical simulations are in rather good agreement with the experimental data.
Nevertheless, the quality of such simulations depends a lot on thermodynamic data (e.g. Henry’s constants).
Solubility data for SO2 and O2 at higher pressure and temperature or solubility data are missing for other impurities
such as NO for example. More complex systems involving gas mixtures and/or brines shall also be considered and
thus the necessary data shall be collected in the literature or experimentally acquired in order to propose associated
simulations. As a consequence, the “SIGARRR” project proposes to combine numerical and experimental efforts in
addition to these first geochemical simulations:
• Experimental (using various technical approaches) or pseudo-experimental (i.e. molecular simulation)
acquisition of thermodynamic data on systems with increasing complexity: gas mixtures, gas
mixtures+brine and gas mixtures+brine+rock.
• Collection or determination of missing or inaccurate parameters. For example, parameterization of
adapted equations of state (especially for brines) using existing and collected data, and their
implementation in geochemical and reactive-transport codes.
• Numerical codes should finally enabled us to handle the final phases of our projects: site-scaled
simulations of leakage scenarios involving complex gas mixtures and the associated risk analysis,
which could also lead to the formulation of first recommendations in terms of geochemically
admissible CO2 flux composition.

The project also aims at improving/extending classical thermodynamic databases, like THERMODDEM
[thermoddem.brgm.fr; 32], by the acquisition of lacking experimental data. Moreover, geochemical indicators such
as pH will be measured in these new experiments and will give supplementary hints to ensure that our models are
correct. In addition, a series of experiments achieved in realistic pressure and temperature conditions and
investigating reservoir-rock and gas mixtures systems will also enable us to validate the behavior of geochemical
codes concerning gas-water-rock interactions. At least, the knowledge acquired about the fate of impurities in
reservoir will lead to some conclusions about the possible composition of unexpected leakages and their impacts on
environment and health, but some first recommendations about the composition of injected CO2 stream are expected
in order to minimize environmental risks.

Acknowledgements

The financial support from the French Agence Nationale de la Recherche, ANR, for the SIGARRR project
(ANR-13-SEED-006) is gratefully acknowledged.

References

[1] Effects of impurities on geological storage of CO2, Technical report IEAGHG, 2011; 2011/04.
[2] Sterpenich J, Dubessy J, Pironon J, Renard S, Caumont M-C, Randi A, Jaubert J-N, Favre E, Roizard D, Parmentier M, Azaroual M, Lachet
V, Creton B, Parra T, El Ahmar E, Coquelet C, Lagneau V, Corvisier J, Chiquet, P. Role of impurities on CO2 injection: experimental and
numerical simulations of thermodynamic properties of water-salt-gas mixtures (CO2+co-injected gases) under geological storage conditions,
Energy Procedia 2013; 37; 3638-45.
[3] Corvisier J. Modeling water-gas-rock interactions using CHESS/HYTEC, Goldschmidt Conference,Florence – Italy 2013.
[4] Corvisier J, Bonvalot A-F, Lagneau V, Chiquet P, Renard S, Sterpenich J, Pironon J. Impact of co-injected gases on CO2 storage sites:
geochemical modeling of experimental results, Energy Procedia 2013; 37; 3699-710.
[5] Appelo CAJ, Parkhurst DL, Post VEA. Equations for calculating hydrogeochemical reactions of minerals and gases such as CO2 at high
pressures and temperatures, Geochimica et Cosmochimica Acta 2014; 125; 49-67.
[6] Spycher N, Pruess K, Ennis-King J. CO2-H2O mixtures in the geological sequestration of CO2. I. Assessment and calculation of mutual
solubilities from 12 to 100°C and up to 600 bar, Geochimica et Cosmochimica Acta 2003; 67; 3015-31.
[7] Nickalls RWD. A new approach to solving the cubic: Cardan’s method revealed, the Mathematical Gazette 1993; 77; 354-59.
[8] Peng D-Y, Robinson DB. A new two-constant equation of state, Industrial & Engineering Chemistry Fundamentals 1976; 15; 59-64.
[9] van der Lee J. Thermodynamic and mathematical concepts of CHESS, Technical report, MINES ParisTech 2009; RT-20093103-JVDL.
[10] Parkhurst DL, Appelo CAJ. Description of input and examples for PHREEQC version 3 - a computer program for speciation, batch-reaction,
one-dimensional transport, and inverse geochemical calculations. Technical report, U.S. Geological Survey 2013; Techniques and Methods
J. Corvisier et al. / Energy Procedia 63 (2014) 3160 – 3171 3171

Report book 6.
[11] Helgeson HC. Thermodynamics of hydrothermal systems at elevated temperature and pressures, American Journal of Science 1969; 267;
724–804.
[12] Pitzer KS. Thermodynamics of electrolytes. I. Theoretical basis and general equations, The Journal of Physical Chemistry 1973; 77; 268–77.
[13] Pitzer KS. Activity Coefficients in Electrolyte Solutions. CRC Press, Boca Raton, USA, 1991.
[14] Bronsted JN. Studies on solubility. IV. The principle of the specific interaction of ions, Journal of the American Chemical Society 1922; 44;
877–98.
[15] Grenthe I, Puigdomenech I. Modelling in Aquatic Chemistry. Nuclear Energy Agency, OECD, Paris, France, 1997.
[16] Yaws CL. Chemical properties handbook : physical, thermodynamic, environmental, transport, safety, and health related properties for
organic and inorganic chemicals. McGraw-Hill, New-York, USA, 1999.
[17] Fernandez-Prini R, Alvarez JL, Harvey AH. Henry’s constants of vapor-liquid distribution constants for gaseous solutes in H2O and D2O at
high temperatures, Journal of Physical and Chemical Reference Data 2003; 32; 903–15.
[18] Chapoy A, Haghighi H, Tohidi B. Development of a Henry’s constant correlation and solubility measurements of n-pentane, i-pentane,
cyclopentane, n-hexane, and toluene in water, Journal of Chemical Thermodynamics 2008; 40; 1030-7.
[19] Mao S, Duan Z. A thermodynamic model for calculating nitrogen solubility, gas phase composition and density of the N2–H2O–NaCl
system, Fluid Phase Equilibria 2006; 248; 103-14.
[20] Duan Z, Mao S. A thermodynamic model for calculating methane solubility, density and gas phase composition of methane-bearing aqueous
fluids from 273 to 523 K and from 1 to 2000 bar, Geochimica et Cosmochimica Acta 2006; 70; 3369-86.
[21] Tanger JCIV, Helgeson HC. Calculation of the thermodynamic and transport properties of aqueous species at high pressures and
temperatures : revised equations of state for the standard partial molal properties of ions and electrolytes, American Journal of Science 1988; 288;
19–98.
[22] Shock, EL, Helgeson HC, Sverjensky DA. Calculation of the thermodynamic and transport properties of aqueous species at high pressures
and temperatures : Standard partial molal properties of inorganic neutral species, Geochimica et Cosmochimica Acta 1989; 53; 2157–83.
[23] Plyasunov AV, Shock EL. Correlation strategy for determining the parameters of the revised Helgeson-Kirkham-Flowers model for aqueous
nonelectrolytes, Geochimica et Cosmochimica Acta 2001; 65; 3879–900.
[24] Li D, Duan Z. The speciation equilibrium coupling with phase equilibrium in the H2O-CO2-NaCl system from 0 to 250 C, from 0 to 1000 bar
and from 0 to 5 molality of NaCl, Chemical Geology 2007; 244; 730–51.
[25] Geng M, Duan Z. Prediction of oxygen solubility in pure water and brines up to high temperatures and pressures, Geochimica et
Cosmochimica Acta 2010; 74; 5631–40.
[26] Xia J, Rumpf B, Maurer G. The solubility of sulfur dioxide in aqueous solutions of sodium chloride and ammonium chloride in the
temperature range from 313 K to 393 K at pressures up to 3.7 MPa : experimental results and comparison with correlations, Fluid Phase
Equilibria 1999; 165; 99–119.
[27] Li H, Yan J. Evaluating cubic equations of state for calculation of vapor–liquid equilibrium of CO2 and CO2-mixtures for CO2 capture and
storage processes, Applied Energy 2009; 86; 826–36.
[28] Soreide I, Whitson CH. Peng-Robinson predictions for hydrocarbons, CO2, N2, and H2S with pure water and NaCl brine, Fluid Phase
Equilibria 1992; 77; 217–40.
[29] Ziabakhsh-Ganji Z, Kooi H. An equation of state for thermodynamic equilibrium of gas mixtures and brines to allow simulation of the
effects of impurities on subsurface CO2 storage, International Journal of Greenhouse Gas Control 2012; 11S; S21–S34.
[30] Liu Y, Hou M, Ning H, Yang D, Yang G, Han B. Phase equilibria of CO2+N2+H2O and N2+CO2+H2O+NaCl+KCl+CaCl2 systems at
different temperatures and pressures, Journal of Chemical & Engineering Data 2012; 57; 1928-32.
[31] Qin J, Rosenbauer RJ, Duan Z. Experimental measurements of vapor-liquid equilibria of the H2O+CO2+CH4 ternary system, Journal of
Chemical and Engineering Data 2008; 53; 1246–9.
[32] Blanc P, Lassin A, Piantone P, Azaroual M, Jacquemet N, Fabbri A, Gaucher EC. Thermoddem: A geochemical database focused on low
temperature water/rock interactions and waste materials, Applied Geochemistry 2012; 27; 2107-16.

View publication stats

You might also like