Sie sind auf Seite 1von 14

G Model

ARTICLE IN PRESS

IJGGC-688; No. of Pages 14

International Journal of Greenhouse Gas Control xxx (2012) xxxxxx

Contents lists available at SciVerse ScienceDirect

International Journal of Greenhouse Gas Control


journal homepage: www.elsevier.com/locate/ijggc

An Equation of State for thermodynamic equilibrium of gas mixtures and brines


to allow simulation of the effects of impurities in subsurface CO2 storage
Zaman Ziabakhsh-Ganji , Henk Kooi
Faculty of Earth and Life Sciences (FALW), VU University Amsterdam De Boelelaan 1085, 1081 HV Amsterdam, The Netherlands

a r t i c l e

i n f o

Article history:
Received 9 March 2012
Received in revised form 29 June 2012
Accepted 26 July 2012
Available online xxx
Keywords:
Carbon Capture and Storage (CCS)
Equation of State (EOS)
Brine
Solubility
Activity coefcient
Thermodynamic model

a b s t r a c t
Comprehensive understanding and prediction of chemical and reactive processes during and following
injection of CO2 in depleted gas reservoirs and saline aquifers is important for the assessment of the
performance and impacts of planned and existing Carbon Capture and Storage (CCS) projects. Over the
last decade signicant improvements have been made in numerical modelling of the complex, coupled
processes involved. Among the many remaining issues where progress is still called for, is the consistent
simulation of impacts of gas mixtures. In particular the presence of impurities or co-contaminants in
the injected CO2 stream that are retained from the original ue-gases, such as H2 S and SO2 , have the
potential, upon dissolution in the pore water, to alter aqueous and watermineral reactions. Moreover,
presence of these and other injected or in situ gases (e.g. CH4 ) affect CO2 solubility and thermodynamic
properties of the uid and gas phases, which, in turn, impact transport processes.
As an important step towards evaluation of the impact of gas mixtures on these processes, a new
Equation of State (EOS) has been developed which allows accurate and efcient modelling of thermodynamic equilibrium of gas mixtures and brines over a large range of pressure, temperature and salinity
conditions. Presently the new EOS includes CO2 , SO2 , H2 S, CH4 and N2 . This model is based on equating
the chemical potentials in the system, using the PengRobinson EOS to calculate the fugacity of the gas
phase. The model performs favourably with respect to existing EOSs and experimental data for single
gas systems and accurately reproduces available data sets for gas mixtures. Preliminary analysis shows,
amongst others, that CO2 solubility is most sensitive to CH4 admixture and least sensitive to the presence
of SO2 in the injected gas.
2012 Elsevier Ltd. All rights reserved.

1. Introduction
There is broad consensus that anthropogenic emission of
carbon dioxide (CO2 ) into the atmosphere is an important driver of
global warming and regional climate change and that the current
trend of increasing CO2 emissions should be reversed (IPCC,
2007). Carbon dioxide capture and storage (CCS) is a signicant
component of the portfolio of mitigation technologies required to
achieve such reversal (Smith et al., 2009). CCS generally involves
capture of CO2 at major stationary sources such as power plants
and injection in underground geological structures such as saline
aquifers, depleted hydrocarbon reservoirs, or producing hydrocarbon reservoirs to enhance oil or gas recovery. Depths greater
than 800 m are generally needed for aquifers to allow CO2 to be
sequestered in supercritical and hence, fairly dense state (Bachu
and Adams, 2003). While demonstration projects for storage in
depleted gas/oil reservoirs have started only recently, the number

Corresponding author. Tel.: +31 20 5987264; fax: +31 20 5989940.


E-mail address: z.ziabakhshganji@vu.nl (Z. Ziabakhsh-Ganji).

of aquifer storage pilots conducted on various continents is progressively growing since rst injection in a saline aquifer overlying
the Sleipner eld in the North Sea began in 1996 (AAAS, 2009;
http://www.globalccsinstitute.com).
Coeval with these developments, efforts are ongoing to study
the processes and factors that control the injectivity, containment,
and long-term safety of geological storage. Numerical models play
an important role in these efforts and are essential, together with
experimental studies, to help develop a comprehensive understanding of the complex, coupled physico-chemical processes at
play in subsurface storage. Over the last decades, capabilities of
these numerical simulators have been progressively improving and
they have been used to clarify key system behaviours such as the
role of heterogeneity and density inuences on spreading of CO2
(Bachu, 2008), the pressure response of aquifers (Birkholzer et al.,
2009), the thermal response of depleted gas reservoirs (Oldenburg
et al., 2009), and CO2 brine-mineral chemical reactions (White
et al., 2005; Andr et al., 2007; Gaus, 2010).
While simulation of injection of pure, dry CO2 and its interaction
with aquifer/reservoir constituents is well developed, capabilities to investigate presence and impacts of other (gaseous)

1750-5836/$ see front matter 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.ijggc.2012.07.025

Please cite this article in press as: 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 in subsurface CO2 storage. Int. J. Greenhouse Gas Control (2012),
http://dx.doi.org/10.1016/j.ijggc.2012.07.025

G Model
IJGGC-688; No. of Pages 14

ARTICLE IN PRESS
Z. Ziabakhsh-Ganji, H. Kooi / International Journal of Greenhouse Gas Control xxx (2012) xxxxxx

Nomenclature
a
A
b
B
kij
P
T
R
x
y
Nw
K
Z
m
kH

energy parameter for PR EOS


dimensionless energy parameter
volume parameter of PR EOS
dimensionless volume parameter
binary interaction coefcient
pressure (bar)
temperature (K)
universal gas constant 83.1447 cm3 bar K1 mol1
component mole fraction in AqP
component mole fraction in NaqP
number of mole of water in one kilogram of water
(55.508)
equilibrium constant
compressibility factor
molality (mol kg1 )
Henrys constant

Greek symbols

acentric factor

constant parameter
adjustable parameter (cm3 g1 )


adjustable parameter (cm3 K0.5 g1 )


adjustable parameter (bar)


second-order interaction parameter
third-order interaction parameter


temperature ( C)

fugacity coefcient

activity coefcient
chemical potential

components than CO2 are still limited. These additional components can either be so-called impurities or co-contaminants (SO2 ,
H2 S, N2 , NOx ) retained from the original ue-gases from which
CO2 was separated, or pre-injection in situ gases in the reservoir
(Ghaderi et al., 2011) or aquifer, notably methane (CH4 ). Evaluation
of the impacts of the presence of such components is important
because they modify dynamical and thermal properties (viscosity, density, JouleThomson coefcient) of the gas/liquid streams,
chemical partitioning among the CO2 -rich phase and brine, and
brine mineral reactions (Jacquemet et al., 2009; Gaus, 2010). Moreover, knowledge of the consequences of impurities in the CO2
stream is of particular interest as high-level purication of CO2 is
costly and injection of co-contaminants with the CO2 may therefore
reduce the front-end processing costs of CCS (Knauss et al., 2005).
Simplied approaches to evaluate the impacts of additional
gases have been used in a couple of studies. Gunter et al.
(2000) performed batch-type modelling (without transport) of
geochemical interaction of carbonate minerals with brine containing dissolved H2 S and H2 SO4 (latter can be considered to be
derived from SO2 brine system). However, CO2 was not included
and concentrations of the acid gases were not obtained from
solubility calculations. Knauss et al. (2005) used a more comprehensive approach involving 1D transport modelling, simulating
the response of an aquifer to injection of a brine pre-equilibrated
with imposed fugacities of CO2 , H2 S and SO2 . Here, the separate
non-aqueous (gas-mixture) phase was not considered in transport
and simplications were involved regarding activity calculations.
Xu et al. (2007) further improved on this by simulating injection
of pure CO2 together with an H2 S and SO2 equilibrated brine. In
their simulations the ECO2N module (Equation of State) of the
TOUGH2 code (Pruess et al., 2004) handles equilibrium calculations
of the CO2 H2 ONaCl system during transport. However, presence

of non-aqueous H2 S and SO2 in the CO2 -stream could not be modelled explicitly.
To further extend simulation-capabilities for analysing the
impacts of associated gases in subsurface CO2 storage, there is a
need for efcient and accurate equations of state (EOS) that model
the thermodynamic equilibrium of a large suite of gas mixtures and
brines with a wide range of composition (Jacquemet et al., 2009).
Some authors such as Zhang et al. (2011) have used a modied
version of the TMVOC simulator of the TOUGH2 family to model
the fate and transport of co-injected H2 S with CO2 in deep saline
formations. Battistelli and Marcolini (2009) recently took an important step by presenting an EOS module (TMGAS) for the TOUGH2
reservoir simulator (Pruess et al., 1999) that handles mixtures of
several inorganic gases and hydrocarbons. They showed TMGAS
reproduces density, viscosity and specic enthalpy data for several
binary gas mixtures and does well in predicting solubility of several
pure gases (CH4 , CO2 , H2 S) in brine and in predicting water content
of pure CO2 . Moreover, they demonstrated injection of a H2 SCO2
mixture in a hydrocarbon reservoir containing methane does alter
behaviour compared to pure CO2 system modelling. Battistelli et al.
(2011) have used TMGAS-TOUGHREACT to model the injection of
an acid gas mixture (CO2 + H2 S + CH4 ) into a high-pressure undersaturated sour oil reservoir. Geloni et al. (2011) have used the same
software to analyse the effect of CO2 + CH4 in rock and cement and
caprock around the wellbore in an exploited hydrocarbon reservoir. Unfortunately, TMGAS still is proprietary software. Moreover,
not all details of the EOS appear to have been disclosed in publication(s), which makes it not easily available or reproducible for
broader academic investigations in CCS.
In this paper, we present a new or alternative Equation of State
(EOS) which, similar to TMGAS, allows accurate and efcient modelling of thermodynamic equilibrium of gas mixtures and brines
over a large range of pressure (up to 600 bar), temperature (up to
110 C) and salinity (up to 6 m) conditions. Presently the model
includes CO2 , SO2 , H2 S, CH4 and N2 , but the suite of gases can be
readily extended. Non-NaCl brines can be handled and activity of
aqueous species is based on the Pitzer formalism for high ionic
strength. In the following we rst present the EOS in detail. Then we
show that the model performs favourably with respect to existing
EOSs and experimental data for single gas systems and accurately
reproduces available data sets for gas mixtures. Focus is on solubility of impurities and special attention is paid to SO2 which is likely
to have prominent geochemical impacts (Knauss et al., 2005; Xu
et al., 2007). Finally, we illustrate that CO2 solubility is most sensitive to CH4 admixture and least sensitive to the presence of SO2 in
the injected gas mixture.

2. Thermodynamic model
The model describes thermodynamic equilibrium between a
non-aqueous phase (NaqP), basically a multi-component mixture
that can be in gas, supercritical or condensed conditions, and an
aqueous phase (AqP), that may include dissolved hydrocarbons
(here methane) and gases in addition to water and dissolved solids.
The EOS does not include solid/minerals as a separate phase.

2.1. Equilibrium between AqP and NaqP


Thermodynamic equilibrium implies that the chemical potential of each component in the AqP and the NaqP are equal. For the
NaqP phase we use
NaqP (T, P) = 0NaqP (T, P) + RT ln(f ),

(1)

Please cite this article in press as: 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 in subsurface CO2 storage. Int. J. Greenhouse Gas Control (2012),
http://dx.doi.org/10.1016/j.ijggc.2012.07.025

G Model

ARTICLE IN PRESS

IJGGC-688; No. of Pages 14

Z. Ziabakhsh-Ganji, H. Kooi / International Journal of Greenhouse Gas Control xxx (2012) xxxxxx

where is chemical potential and 0 is chemical potential at reference temperature, R is the gas constant, T is temperature, P is
pressure, and f is fugacity which is equal to
f = P y,

(2)

assumed to be 1 bar. The equilibrium constant at reference pressure


KH0 O (T, P0 ) is obtained from
2

log(KH0

2O

) = 2.209 + 3.097 102  1.098 104  2


+ 2.048 107  3 ,

(12)

where is the fugacity coefcient, y is the mole fraction of the


component in the NaqP and P is total pressure.
The chemical potential of the AqP is written in terms of activity
a rather than fugacity.

where  is temperature in C (Spycher et al., 2003). By combining


(10), (11) and (2) we obtain

AqP (T, P) = 0AqP (T, P) + RT ln(a),

yH2 O =

(3)

KH0

2O

aH2 O

H2 O P


exp

(P P0 )VH2 O

RT

(13)

Equating the two chemical potentials gives


0AqP (T, P) 0NaP (T, P)
RT

= ln

 a 

(4)

P y

In terms of equilibrium constant K0 this reads


0AqP (T, P) 0NaP (T, P)
RT

= ln(K 0 ).

a
Nw
=
kH
P y

(6)

where Nw is the number of moles per kilogram of water (55.508)


and kH is Henrys constant. We further use a = Nw
x, which is reasonable because the solubility of the gas species is small (Spycher
and Pruess, 2005). The effect of salt is accounted for in the activity
coefcient of the gas species,
(Eq. (26)). Therefore the nal equation for dissolved gas yields (Akinev and Diamond, 2003; Zirrahi
et al., 2012)
P y = kH
x

(7)

Hence, for each gas we have


(P i yi )NaqP = (kHi
i xi )AqP

(8)

where subscript i denotes individual gases like CO2 , SO2 , N2 , H2 S,


CH4 except water. Similar to Battistelli and Marcolini (2009), we
ignore binary interaction between different dissolved gases in the
aqueous phase. Therefore, the activity coefcient in Eq. (8) for individual gas species does not depend on presence of other gases. This
assumption is an important one in the present EOS since it allows
use of a rather simple, non-iterative solving method.
Since the EOS should be able to quantify thermodynamic equilibrium between gas mixtures and brine, water is an important
system component. For equilibrium between H2 O in the AqP and
the NaqP we follow the approach of Spycher et al. (2003):
H2 O(l) H2 O(g)

(9)

fH2 O(g)

(10)

aH2 O(l)

where K is Spychers true equilibrium constant, f is the fugacity


of NaqP water, and a is the activity of AqP water. The equilibrium
constant of water is a function of temperature and pressure as given
by Eq. (11)

KH2 O (T, P) =

KH0 O (T, P0 ) exp


2

(5)

By using kH = Nw/K 0 (Prausnitz et al., 1986) Eqs. (4) and (5)


yield

KH2 O =

At the range of consideration for pressure and temperature


(5110 C), the solubility of gases in water is low, and the activity
of the water component can be approximated by its mole fraction
in the liquid phase. Therefore, in Eq. (13) the effect of dissolved salt
is accounted for in the activity of water. Using these considerations
yields

(P P0 )VH2 O
RT

KH0 O
2

exp

(P P0 )VH2 O

RT

xH2 O = H2 O PyH2 O

(14)

Eq. (14) is used in the model for equilibrium of H2 O in the system.


For the full set of equilibrium equations, Eq. (14) is combined with
Eq. (8).
2.2. Non aqueous phase
In Eqs. (8) and (14) the fugacity coefcient must be derived from
PVT or PVT-X properties of brines and different gas mixtures by utilizing an Equation of State. There are many EOSs in the literature.
Some authors such as Duan and Sun (2003) and Duan and Mao
(2006) used a virial like Equation of State. However, the complexity of the equation makes it not very practical for our aims. Cubic
equations in volume were developed and improved over the last
century. Redlich and Kwong (1949) (RK) and Peng and Robinson
(1976) (PR) are well-known examples. Other examples are the
equations of Schmidt and Wenzel (1980) and Soave RedlichKwong
(1972) which basically are modications of the Van der Waals EOS.
Still other studies used a modied form of RK to represent the
properties of gas mixtures such as CO2 H2 O (Spycher et al., 2003;
King et al., 1992; Zirrahi et al., 2010; Hassanzadeh et al., 2008).
Similar to Battistelli and Marcolini (2009), we use the classical
PengRobinson (1976) EOS. Although the PR EOS is more elaborate than, for instance, RK, it has the advantage that is has greater
accuracy around the liquidvapour boundary.
In our model, calculation of the fugacity coefcient in the non
aqueous phase is as follows. The compressibility factor obeys
Z 3 (1 B)Z 2 + (A 2B 3B2 )Z (AB B2 B3 ) = 0

(15)

Parameters A and B are a function of pressure and temperature


and are dened as follows:
A=

B=

a(T )P

(16)

(RT )2
bP
RT

(17)

where
(11)

where T is temperature in K; VH2 O is the average partial molar


volume of the water in the AqP over the pressure interval from
P to P0 which is equal to 18.1; P0 is a reference pressure, which is

a(T ) = 0.45724
b = 0.07780

R2 Tc2
(T )
Pc

RTc
Pc

(18)
(19)

Please cite this article in press as: 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 in subsurface CO2 storage. Int. J. Greenhouse Gas Control (2012),
http://dx.doi.org/10.1016/j.ijggc.2012.07.025

G Model

ARTICLE IN PRESS

IJGGC-688; No. of Pages 14

Z. Ziabakhsh-Ganji, H. Kooi / International Journal of Greenhouse Gas Control xxx (2012) xxxxxx

Table 1
Experimental data used in calibration of the model.
Solution
(m = mol kg1 )

T (K)

P (bar)

CO2
Wiebe and Gaddy (1939)
Wiebe and Gaddy (1940)
Todheide and Franck (1963)
Zawisza and Malesinska (1981)
Muller et al. (1988)
King et al. (1992)
Drummond (1981)
Nighswander et al. (1989)
Rumpf et al. (1994)
Prutton and Savage (1945)

Water
Water
Water
Water
Water
Water
06.5 m NaCl
00.17 m NaCl
46 m NaCl
03.9 m CaCl2

323373
285313
323623
323473
373473
288298
290673
353473
313433
348394

25710
25510
2003500
154
380
60250
35400
20100
1100
10700

H2 S
Selleck et al. (1952)
Lee and Mather (1977)
Gillespie and Wilson (1982)
Drummond (1981)
Suleimenov and Krupp (1994)
Xia et al. (2000)

Water
Water
Water
06 m NaCl
02.5 m NaCl
46 m NaCl

310.15443.15
283.15453.15
311.15477.15
373.15653.15
393.15593.15
313.15398.15

6210
170
10210
6200
0140
10100

N2
Goodman and Krase (1931)
Wiebe et al. (1932)
Wiebe et al. (1933)
Saddington and Krase (1934)
Smith et al. (1962)
OSullivan and Smith (1970)
Alvarez and Prini (1991)
Chapoy et al. (2004)

Water
Water
Water
Water
06.2 m NaCl
04.6 m NaCl
Water
Water

273.15442.15
298.15
298.15373.15
338.15513.15
303.15
324.65398.15
336.3636.5
274.19363.02

101.3303.9
25.331013.25
25.331013.25
101.3304
1172.6
101.3616.1
5.34256
9.7170.43

Wang et al. (2003)


Chapoy et al. (2005)

Water
1.016.59 m NaCl
2.912.93 m CaCl2
0.00.6 m brine
Water
Water
Water
0.56.1 m NaCl
07.35 m CaCl2
Water
1.014.41 m NaCl
Water
1.0 m CaCl2
01.54 m NaCl
05.9 m NaCl
Water
Water
04.0 m NaCl
04.0 m KCl
02.16 m MgCl2
02.0 m CaCl2
Water
0.814.7 m NaCl
Water
Water
Water
Water
Water
Water
0.993.99 m KCl
Water
Water

298.15423.15
298.15423.15
298.15
311.15394.15
298.15
298.2444.3
298.15303.15
303.15
298.15303.15
324.65398.15
324.65398.15
310.93344.26
298.2398.2
323623
372.15513.15
297.5518.3
298.15
298.15
298.15
298.15
298.15
277.2573.2
277.2573.2
298.2338.2
298.15323.15
298.15323.15
273.2290.2
344
313473
313.51373.19
283.2303.2
275.11313.11

40.6469.1
41.8456.0
56.2209.9
35345
36.2667.4
22.3689.1
3.1751.71
214.8957.5
3.274.8
101.3616.1
101.3616.1
41.4344.7
101608
295
751570
13.2764.51
24.151.7
24.151.7
24.151.7
24.151.7
24.151.7
11132
19124
25125
3080
5.6790.82
34.5
2001000
3.493
4.297.9
20400.3
9.7180

SO2
Rumpf and Maurer (1992)
Xia et al. (1999)

Water
2.9425.928 m NaCl

293.15413.15
313393

0.125
0.137

CO2 + CH4 + H2 S
Huang et al. (1985)

Water

37.8176.7

48.2173.1

CH4
Michels et al. (1936)

Dodson and Standing (1944)


Culberson et al. (1950)
Culberson and Mcketta (1951)
Duffy et al. (1961)

OSullivan and Smith (1970)


Amirijafari and Campbell (1972)
Blanco and Smith (1978)
Namiot et al. (1979)
Blount and Price (1982)
Crovetto et al. (1982)
Stoessell and Byrne (1982b)

Cramer (1984)
White et al. (1985)
Yokoyama et al. (1988)
Lekvam and Bishnoi (1997)
Song et al. (1997)
Dhima et al. (1998)
Kiepe et al. (2003)

where is the acentric factor. To calculate the fugacity coefcient


for each species, i , in gas mixtures, we use standard simple mixing
rules and binary interaction coefcients (Prausnitz et al., 1986).

and


(T ) =


1+(0.37646+1.4522 0.269922 )

T
Tc


2
(20)

a=

yi yj aij ,

aij =

ai aj (1 kij ),

b=

bi yi

(21)

Please cite this article in press as: 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 in subsurface CO2 storage. Int. J. Greenhouse Gas Control (2012),
http://dx.doi.org/10.1016/j.ijggc.2012.07.025

G Model

ARTICLE IN PRESS

IJGGC-688; No. of Pages 14

Z. Ziabakhsh-Ganji, H. Kooi / International Journal of Greenhouse Gas Control xxx (2012) xxxxxx
Table 2
Calibrated parameter values for kH for various species (Eq. (24)).
Gas

CO2
H2 S
SO2
N2
CH4

0.114535
0.77357854
0.198907
0.008194
0.092248

5.279063
0.27049433
1.552047
5.175337
5.779280

6.187967
0.27543436
2.242564
6.906469
7.262730

B
A
ln i = i (Z 1) ln(Z B) +
B
2.828B
ln

Table 5
Calibrated binary interaction coefcients.


2
Bi

ya
j j ij

0
0
0.009847
0
0

 Z + 2.414B 

(22)

Z 0.414B

The binary interaction coefcients for CO2 SO2 , CO2 H2 S,


CO2 CH4 and CO2 N2 were obtained from Li and Yan (2009). As
observed by Spycher et al. (2003) the current approach neglects the
mole fraction of water in the mixing rule (same as assuming innite dilution of NaqP H2 O). This is practical since it reduces iteration
demands as will be explained in the section on model calibration.
To obtain the proper value of Z in Eq. (22), we follow the
approach described by Danesh (1998). When Eq. (15) has three
roots, the intermediate one is ignored and the root yielding the
lowest Gibbs free energy between the remaining two is selected.
G
G
Let Zh and Zl be the two real roots with RTh and RTl being the Gibbs
free energy, and where the subscripts denote the high and low Z
value, respectively. The difference in Gibbs free energy is given by:
(Gh Gl )
= (Zh Zl ) + ln
RT



ln

Z B
l

Zl + 1 B
Zl + 2 B

Zh B



H2 O

CO2

H2 S

SO2

N2

CH4

0.19014

0.105

1.10329

0.32547

0.47893

2.3. Aqueous phase


In the next step of model development, Henrys constant, kHi ,
and the activity coefcient,
i , on the right-hand side of Eq. (8), need
to be quantied. For the temperature and pressure dependency of
Henrys constant, we use a correlation established by Akinev and
Diamond (2003). The correlation is a virial-like equation for the
thermodynamic properties of the aqueous phase species at innite
dilution and requires but a few empirical parameters (constrained
by experimental data), and these parameters are independent of
temperature and pressure
ln(kH ) = (1 ) ln fH0

2O



(23)

where 1 and 2 for PengRobinson EOS are 1 + 2 and 1 2,


respectively. If (Gh Gl )/RT in Eq. (23) is positive Zl is selected;
otherwise Zh is the correct root.

+  ln

 RT

Mw

0
H

2O

0
+ 2 H

2O

(24)

where  is a constant for each dissolved gas in water, T is tempera0


ture in K; and fH0 O and H
are fugacity and density of pure water,
2
2O
respectively. For calculation of properties of pure water we use the
correlation of Fine and Millero (1973); see Appendix A. It should
be noted that we quantify fH0 O using Eq. (A6) because Eq. (22) is
2

not sufciently accurate for pure water. In Eq. (24) B (cm3 g1 )


stands for the difference in interaction between dissimilar solvent
molecules (Akinev and Diamond, 2003) and is calculated as follows:

B =  + P +

B(2 1 )

Zh + 2 B
Zh + 1 B

103
T

(25)

where  (cm3 g1 ), (cm3 K0.5 g1 ) and  (bar1 ) denote adjustable


parameters. In the original version of Eq. (25) (Akinev and
Diamond, 2003, their Eq. (15)) the term  P does not occur. It is
added here specically for SO2 to allow for the fact that the solubility of this species is two to three orders of magnitude higher than
other considered gas species. For the latter we, therefore, set  = 0.
It should further be noted that we only used the parameter values

Table 3
Calibrated second-order interaction parameters () for various gases (Eq. (27)).
Constant

CO2 Na

SO2 Na

N2 Na

H2 SNa

CH4 Na

c1
c2
c3
c4
c5
c6
c7
c8
c9
c10

0.0652869
1.6790636E04
40.838951
0
0
3.9266518E02
0
2.1157167E02
6.5486487E06
0

5.0961510E02
2.8865149E04
0
0
1.1145002E02
0
2.4878170E05
0
0
0

2.0939363
3.1445269E03
3.9139160E+02
2.9973977E07
0
1.5918098E05
0
0
0
0

1.03658689
1.1784797E03
1.7754826E+02
4.5313285E04
0
0
0
0
0
0.47.751650E+02

5.7066455E01
7.2997588E04
1.5176903E+02
3.1927112E05
0
1.6426510E05
0
0
0
0

Table 4
Calibrated third-order interaction parameters () for various gases (Eq. (27)).
Constant

CO2 Na

SO2 Na

N2 Na

H2 SNa

CH4 Na

c1
c2
c3
c4
c5
c6
c7
c8
c9
c10

1.144624E02
2.8274958E05
0
0
0
1.3980876E02
0
1.4349005E02
0
0

7.1462699E03
0
0
0
0
0
0
0
0
0

6.3981858E03
0
0
0
0
0
0
0
0
0

0.010274152
0
0
0
0
0
0
0
0
0

2.9990084E03
0
0
0
0
0
0
0
0
0

Please cite this article in press as: 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 in subsurface CO2 storage. Int. J. Greenhouse Gas Control (2012),
http://dx.doi.org/10.1016/j.ijggc.2012.07.025

G Model

ARTICLE IN PRESS

IJGGC-688; No. of Pages 14

Z. Ziabakhsh-Ganji, H. Kooi / International Journal of Greenhouse Gas Control xxx (2012) xxxxxx

Fig. 1. Comparison between results for the new EOS (solid lines) and experimental data (symbols) for CO2 solubility (left panels) and H2 O content of the CO2 -rich phase
(right panels) for various temperatures. The data shown are identical to those used by Spycher et al. (2003; their Appendix A).

Table 6
Absolute deviation (%) from experimental data for the new model and for the models of Spycher et al. (2003) and Duan and Sun (2003).
T (K)

P bar

xCO2 100

New model

Spycher et al. (2003)

Duan and Sun (2003)

323.15
323.15
323.15
323.15
323.15
323.15
323.15
323.15
323.15
323.15
323.15
323.15
323.15
323.15
323.15
323.15
AAD

25.3
40.5
50.6
68.2
75.3
101.33
111
121
141.1
152
200
304
405.3
500
608
709.3

0.774
1.09
1.37
1.651
1.75
1.98
2.10
2.14
2.17
2.174
2.30
2.457
2.606
2.80
2.868
2.989

2.550519
6.983281
0.250958
0.787999
0.410151
0.425269
2.974057
2.906066
1.233456
0.007072
0.528422
0.945205
1.094991
1.321428
1.041145
0.931
1.219551

1.720475
6.886282
0.607855
1.83888
1.682105
1.674057
2.164378
2.367484
0.969925
0.154437
0.530476
0.980066
1.218304
1.187714
1.074843
0.741758
1.289952

2.456373
7.069009
0.465618
1.273293
1.01269
1.368011
2.227906
2.407302
1.091927
0.004453
0.780196
0.862124
1.505869
0.402989
2.563775
2.914181
1.420286

AD% = 100|(xexp xcal )/xexp |, ADD% =

100|(xexp xcal )/xexp | /N.

Please cite this article in press as: 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 in subsurface CO2 storage. Int. J. Greenhouse Gas Control (2012),
http://dx.doi.org/10.1016/j.ijggc.2012.07.025

G Model

ARTICLE IN PRESS

IJGGC-688; No. of Pages 14

Z. Ziabakhsh-Ganji, H. Kooi / International Journal of Greenhouse Gas Control xxx (2012) xxxxxx

Fig. 2. (a) Comparison between results for the new EOS (solid lines) and predictions of the EOSs of Spycher and Pruess (2005) and Duan and Sun (2003) (symbols) for CO2
solubility as function of pressures at 90 C and 2 m salinity of NaCl. (b) Same for the EOS of Spycher and Pruess (2005) and 76 C and various salinity of CaCl2 .

for Eqs. (24) and (25) tabulated by Akinev and Diamond (2003) as
initial guesses, and obtained new estimates of these parameters for
the pressure and temperature ranges of interest during calibration
(new parameter values are listed in Table 2).
For calculation of the activity coefcient of the various gas
species, we developed an approach similar to Duan and Sun (2003).
Reduction of the activity coefcient due to interaction with solute
present in the brine (Pitzer, 1973), is based on a virial expansion of
Gibbs excess energy
ln
i =

2mC iC +

2mA iA +

mA mC iAC

(26)

where subscript of mC and mA denote anions and cations molality,


respectively, iC and  iAC are second and third-order interaction parameters, respectively. Following Duan and Sun (2003) we
also assume iA = 0. Temperature and pressure dependency of the
interaction parameters is modelled using

potassium and magnesium salts, Eq. (26) effectively gives

ln
i = 2iNa (mNa + 2mCa + mK + 2mMg ) + iNaCl mCl
(mNa + mK + mMg + mCa )

(28)

It may be worthy to note that, in contrast to Duan and Sun


(2003) who use different equations for the pressure and temperature dependency for the various gas species, we use the single form
shown in Eq. (27). Furthermore, it may be pointed out that in the
absence of dissolved salts (gas dissolution in pure water), Eq. (26)
yields activity coefcients equal to 1.

3. Model calibration
c3
c5
c8 P
P
T
Par(T, P) = c1 + c2 T +
+ c4 P +
+ c6 + c7 2 +
T
P
T
630 T
P
+ c9 T ln P + c10

P
T2

(27)

where Par(T, P) is either  or , P is pressure in bar and T is temperature in K and the ci , i = 1, . . ., 10 are constants which are calculated by
using a tting procedure and are given in Table 3. Following Duan
and Sun (2003) we reduce the number of interaction parameters
by using iC = iNa , and by interpreting mC for the second-order
interaction term in terms of equivalents per mass of brine (that
is, molality multiplied by valency). For the third-order interaction
parameters we assume  iAC =  iNaCl while maintaining molalities in Eq. (26). For instance, for a brine containing sodium, calcium,

The Weighted Nonlinear Least Square (WNLS) method was used


in model calibration. Calibration for each gas component (CO2 , CH4 ,
N2 , H2 S, SO2 ) was conducted in two steps. First, the parameters in
the Henry constant (Eqs. (24) and (25); , , , and ) and the binary
interaction coefcient between H2 O and the gas species (Eq. (21))
were obtained using experimental data for equilibrium with pure
water. Second, the parameters in the relationship quantifying the
activity coefcient (Eq. (27); ci , i = 1, . . ., 10) were determined from
data for brinegas systems. The calibration targets (experimental
data sets) that were used were obtained from the literature listed in
Table 1. These include both solubility data for the gas species in the
aqueous phase and data on H2 O vapour content in the non aqueous
phase. The optimized parameter values are listed in Tables 25.

Please cite this article in press as: 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 in subsurface CO2 storage. Int. J. Greenhouse Gas Control (2012),
http://dx.doi.org/10.1016/j.ijggc.2012.07.025

G Model

ARTICLE IN PRESS

IJGGC-688; No. of Pages 14

Z. Ziabakhsh-Ganji, H. Kooi / International Journal of Greenhouse Gas Control xxx (2012) xxxxxx

Fig. 3. (a) Comparison between results for the new EOS (solid lines) and predictions
of the EOS of Duan et al. (2007) (symbols) for H2 S solubility in NaCl brines with
different salinities and at a temperature of 333.15 K. (b) Comparison between results
for the new EOS (solid line) and experimental data (Soreide and Whitson, 1992) for
H2 O content in the H2 S-rich phase.

4. Solving the model equations for systems consisting of


single or mixed gases
For single gasbrine systems, partitioning of the gas species and
H2 O among the aqueous and non-aqueous phase can be directly
obtained from the following equations. The determination of the
phase composition for single gasbrine systems is non iterative
because of the assumptions made that the fugacity of the gas
in NAqP does not depend on composition, but can be computed
accounting only for know P and T (illustrated here for CO2 ).

1
yH2 O =



K0

1/

H2 O

H2 O P

exp

P CO

2
kH

H2 O CO2

 (PP

0 )VH2 O

P CO2
kHCO
CO2

1




K0

H O
1/ 2 P
H2 O

exp

P CO

(29a)

2
kH

CO2 CO2

P CO
kH

i=CO2 ,H2 O

xi = 1,

i=CO2 ,H2 O

yi = 1

CO2

CO

(PP0 )VH O
2
RT



 (29b)

P CO
kH

CO2

zi (Ki 1)
=0
1 + (Ki 1)nV

yi
,
xi

(30)

i = H2 O, N2 , CO2 , CH4 , H2 S, SO2

(32)

According to Eq. (14) the K-value for water is

CO

(31)

is solved for nV , the mole fraction of NaqP in the system (value


between 0 and 1). In Eq. (31) Ki -values are dened as follows:
Ki =

Eq. (29) is obtained by combination of Eqs. (8) and (14) and using
the fact that

For gas mixtures the method is a little different. Because the


number of equations and the number of unknowns (mole fraction
of each species in gas and liquid phase) are not equal, vapourliquid
ash calculations are used (Danesh, 1998). First the mole fraction
of each component in the total system, zi , (feed) is dened. The
problem is to solve for xi and yi for given P, T, and salt molality. To
do this, the RachfordRice equation

i=1

RT

xCO2 =

Fig. 4. (a) Comparison between results for the new EOS (solid lines) and Duan
et al. (2007) data (symbols) for H2 S solubility in NaCl brines at 363.15 K for various salinities. (b) Comparison between results for the new EOS (solid lines), and
AQUAlibrium (2010) software-predicted (symbols) water content in the H2 S-rich
phase. The dashed lines illustrate that if ideal mixing would be assumed for calculation of water content in the H2 S-rich phase, the model would yield too low values
for pressures exceeding about 40 bar.

KH2 O =

KH0

2O

exp[((P P0 )VH2 O )/RT ]


P H2 O

(33)

And according to Eq. (8) for other components, the K-value is


Ki =

kHi
i
P i

i = N2 , CO2 , CH4 , H2 S, SO2

(34)

Please cite this article in press as: 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 in subsurface CO2 storage. Int. J. Greenhouse Gas Control (2012),
http://dx.doi.org/10.1016/j.ijggc.2012.07.025

G Model
IJGGC-688; No. of Pages 14

ARTICLE IN PRESS
Z. Ziabakhsh-Ganji, H. Kooi / International Journal of Greenhouse Gas Control xxx (2012) xxxxxx

Fig. 5. Comparison between results for the new EOS (solid lines) and experimental data (symbols) for SO2. (a) Solubility in pure water; data from Rumpf and Maurer (1992).
(b) Solubility in NaCl brines for different salinity and temperature; data from Xia et al. (2000).

Fig. 6. (a) Comparison between results for the new EOS (solid lines) and predictions of the EOS of Mao and Duan (2006) (symbols) for N2 solubility in NaCl brines with
different salinities and at a temperature of 333.15 K. (b) Comparison between results for the new EOS (solid line) and experimental data (Namiot and Bondareva, 1959)
(symbols) for water content in the N2 -rich phase at 366.45 K.

Please cite this article in press as: 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 in subsurface CO2 storage. Int. J. Greenhouse Gas Control (2012),
http://dx.doi.org/10.1016/j.ijggc.2012.07.025

G Model

ARTICLE IN PRESS

IJGGC-688; No. of Pages 14

Z. Ziabakhsh-Ganji, H. Kooi / International Journal of Greenhouse Gas Control xxx (2012) xxxxxx

10

Fig. 7. (a) Comparison between results for the new EOS (solid lines) and predictions of the EOS of Duan and Mao (2006) (symbols) for CH4 solubility in NaCl brines with
different salinities and at a temperature of 333.15 K. (b) Comparison between results for the new EOS (solid line) and experimental data (Olds et al., 1942) (symbols) for
water content in the CH4 -rich phase at 377.59 K.

With the known value of nV , Ki s and zi s, the mole fraction of


each phase are subsequently obtained from
xi =

zi
,
1 + (Ki 1)nV

yi =

Ki zi
.
1 + (Ki 1)nV

(35)

In the case of gas mixtures the binary interaction between


dissolved gases in the aqueous phase is ignored (Battistelli and
Marcolini, 2009). Binary interaction coefcients for CO2 SO2 ,
CO2 H2 S, CO2 CH4 , CH4 H2 S and CO2 N2 for the non-aqueous
phase are adopted from Li and Yan (2009), while the gasH2 O
interaction coefcients, listed in Table 5 are known from model
calibration.
It may be worthwhile repeating that in the mixing rule (Eq. (21))
the mole fraction of water has been neglected. This is convenient
since it eliminates the need for iteration procedures.
5. Result and discussion
In this section the model performance is assessed and illustrated
by showing: (1) results of calibration, (2) comparison with existing
models for H2 ONaClCO2 and H2 OCaCl2 CO2 systems (Spycher
et al., 2003; Spycher and Pruess, 2005; Duan and Sun, 2003; Duan
and Mao, 2006; Mao and Duan, 2006), and (3) how model predictions for gas mixtures compare with experimental data.

up to pressures of 600 bars. The Spycher model forms the basis


of the ECO2N-EOS module in TOUGH2 (Pruess, 2005). To allow a
further quantitative comparison with these models, we calculated
the absolute deviation (AD) of these models and our model with
respect to experimental data shared by all studies. Table 6 shows
that our model performs favourably with respect to the existing
ones. Table 6 only gives results for a small selection of data. For the
full data set of 405 measurements the (average absolute deviation)
AAD value of our model is about 8.3%.

5.2. H2 Sbrine
For H2 S our model shows good agreement with the solubility
model of Duan et al. (2007) (Fig. 3a) and with experimental data for
the mole fraction of water in the gas phase (Soreide and Whitson,
1992) (Fig. 3b). Fig. 4a highlights the importance of accounting for
non-ideality of H2 O in the non-aqueous phase as we do in our
model. The gure indicates that the assumption of ideal mixing
for water in the gas-rich phase strongly underestimates the actual
water content at relatively high pressures. Fig. 4b illustrates model
behaviour with respect to H2 S solubility in brine for various salinities.

5.1. CO2 brine

5.3. SO2 brine

In Fig. 1 a comparison is shown of the model with respect to


experimental data used in the calibration for pure water. Both CO2
solubility and water content in the CO2 -rich phase are shown for
different temperatures. Fig. 2 demonstrates that our model predictions are virtually identical to those of Spycher et al. (2003), Spycher
and Pruess (2005) and Duan and Sun (2003) for different brines

Experimental data on SO2 water or SO2 brine systems are very


scarce. Fig. 5a and b shows how the model compares with respect
to the data reported in the two studies listed in Table 1. It should
be noted that although the model t is good, predictions at pressures beyond about 10 bar are not well constrained and, hence, are
associated with relatively large uncertainty.

Please cite this article in press as: 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 in subsurface CO2 storage. Int. J. Greenhouse Gas Control (2012),
http://dx.doi.org/10.1016/j.ijggc.2012.07.025

G Model
IJGGC-688; No. of Pages 14

ARTICLE IN PRESS
Z. Ziabakhsh-Ganji, H. Kooi / International Journal of Greenhouse Gas Control xxx (2012) xxxxxx

Fig. 8. Comparison between results for the new EOS (solid lines) and experimental data (symbols) for a three-component gas mixture consisting of CO2 , CH4 and
H2 S with mole fractions of 0.6, 0.3 and 0.1, respectively. (a) Solubility of CO2 in
pure water; experimental data from (Huang et al., 1985). (b) H2 O content of the gas
mixture (NaqP); experimental data from (Huang et al., 1985).

11

Fig. 9. According to our EOS CH4 solubility increases in the presence of CO2 (a) and
CO2 solubility increases in the presence of CH4 . This behaviour is shown here for
T = 375 K and is consistent with behaviour inferred in an experimental study by Qin
et al. (2008). For the horizontal axis, n denotes the amount of gas species in the
system in moles.

5.4. N2 brine
Fig. 6a demonstrates model-predicted solubilities of N2 show
good correspondence with solubilities by Mao and Duan (2006)
for a large range of salinities. Model performance with regard to
experimental data for the mole fraction of water in rich N2 phase
(Namiot and Bondareva, 1959) is illustrated in Fig. 6b.
5.5. CH4 brine
Fig. 7 depicts model performance for methane. Both the correspondence with the existing model of Duan and Mao (2006) (Fig. 7a)
and experimental data (Olds et al., 1942) are very satisfactory.
5.6. Gas mixturebrine
We have been able to nd two studies with relevant experimental data regarding equilibrium conditions for brine and gas
mixtures (Huang et al., 1985; Qin et al., 2008). The system of Huang
et al. (1985) is for a 3-component gas mixture consisting of CO2 ,
CH4 and H2 S with mole fractions of 0.6, 0.3 and 0.1, respectively.
This gas mixture is equilibrated with various amounts of water.
Data and model performance is shown in Fig. 8. It should be noted

that the experimental data (Huang et al., 1985) were not involved
in the model calibration, which was done for single gasbrine systems only. The favourable correspondence between data and model
therefore provides some condence that the employed method for
handling gas mixtures is adequate.
The system of Qin et al. (2008) is for a 2-component gas mixture consisting of CO2 and CH4 . Their experimental data showed
that CO2 solubility (dened as the reciprocal of apparent Henrys
law constant, (P yCO2 /xCO2 )1 ) increases in the presence of CH4 .

Similarly, CH4 solubility, (P yCH4 /xCH4 )1 , was shown to increase


in the presence of CO2 . Fig. 9 shows that our model reproduces
this behaviour. Unfortunately, we could not test our model against
their measurements in detail because the amount of water in the
experimental system was not reported.
Finally, we used our model to investigate the sensitivity of CO2
solubility to the presence of various impurities. Fig. 10 shows the
impact of 5% (by weight) admixture of each of the contaminant
gases (as listed in Table 7) on the mole fraction of carbon dioxide in the aqueous phase. Results show that CO2 mole fraction in
the aqueous phase is most sensitive to CH4 admixture and least
sensitive to the presence of SO2 in the injected gas.

Please cite this article in press as: 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 in subsurface CO2 storage. Int. J. Greenhouse Gas Control (2012),
http://dx.doi.org/10.1016/j.ijggc.2012.07.025

G Model

ARTICLE IN PRESS

IJGGC-688; No. of Pages 14

Z. Ziabakhsh-Ganji, H. Kooi / International Journal of Greenhouse Gas Control xxx (2012) xxxxxx

12

allows use of a simple, non-iterative solving method. More


comprehensive experimental data for gas mixtures are needed to
ascertain the validity of this assumption. It should be further noted
that uncertainty of model predictions will be strongly correlated
with the P and T ranges of experimental data of the various gas
species used in model calibration.
The EOS may be suitable for incorporation in reactive transport
simulators in order to address, for instance, the chemical impact
of co-components. Presently the model includes fugacity calculations using the PengRobinson EOS. However, it can be easily
modied to include another EOS. In that case the model calibration parameters would have to be re-calculated. We did so for the
modied RK EOS (Spycher et al., 2003), but for H2 S the prediction of
the water content proved to be considerably less accurate than for
the PR EOS.
A key advantage of the present model is that it is fairly simple
and non-iterative and fully explicated, which makes that it can be
readily adopted for use in other studies.

Fig. 10. Results for mole fraction of CO2 in the aqueous phase for the gas mixtures
listed in Table 7.

Acknowledgments
Table 7
Weight fraction in the composition of injected gas (before equilibration with water).
Gas

Case 1

Case 2

Case 3

Case 4

CO2
SO2
N2
H2 S
CH4

95
5
0
0
0

95
0
5
0
0

95
0
0
5
0

95
0
0
0
5

This work was made possible by support of the Dutch National


Research Programme CATO2 on CO2 Capture, Transport and Storage: Work Package 3.2: Reservoir behaviour. We would further
like to thank Boris van Breukelen for help with issues concerning calibration methods and general hydrochemistry, Ali
Akbar Eftekhari for contributions regarding thermodynamic concepts, and Mariene Gutierrez-Neri for general assistance and
discussions.

6. Concluding remarks

Appendix A.

In CCS or acid gas disposal, and natural gas sweetening and


gas transportation, accurate estimation of water content in the
non-aqueous phase as well as solubility of gases in water/brine is

For calculation of properties of pure water we use the correlation


of Fine and Millero (1973)
V=

V0 =

V 0 V 0P
,
B + A1 P + A2 P 2

(A1)

1 + 18.159725 103 
0.9998396 + 18.224944 10

 7.922210 106  2 55.44846 109  3 + 149.7562 1012  4 393.2952 1015  5

B = 19654.320 + 147.037 2.21554 2 + 1.0478 102  3 2.2789 105  4 ,


3

A1 = 3.2891 2.3910 10
A2 = 6.245 10

4 2

 + 2.8446 10
6

3.913 10

 2.8200 10
8 2

 3.499 10

(A2)

(A3)

6 3

9 4

 + 8.477 10

 + 7.942 10

10 3

 ,

(A4)

12 4

 3.299 10

 ,

(A5)
C

important. In this paper, a new EOS has been presented which quanties the thermodynamic equilibrium between gas mixtures and
brines. Presently the model includes CO2 , SO2 , H2 S, CH4 and N2 ,
but the suite of gases can be readily extended. Non-NaCl brines can
be handled and activity of aqueous species is based on the Pitzer
formalism for high ionic strength. This model predicts the water
content in non-aqueous phase and composition of the various gas
components in both the aqueous and non-aqueous phase at moderate temperatures (5110 C), a wide pressure range (1600 bar)
and various salinities (06 m).
The model predictions are consistent with existing EOSs and
experimental data for single gas systems and accurately reproduces
available, albeit still very limited data sets for gas mixtures. Analysis
shows, amongst others, that the amount of dissolved CO2 is most
sensitive to CH4 admixture and least sensitive to the presence of
SO2 .
An important assumption of the EOS is that it neglects binary
interaction between dissolved gases in the aqueous phase. This

where  is temperature in
and V = 1/ is the reciprocal of the
density of pure water in cubic centimetre per gram.
For the fugacity of pure water we use the relation of King et al.
(1992)
fH0

2O

= Ps exp

 (P P )v 
s
RT

(A6)

where v is molar volume, calculated by multiplying molecular


weight of water (18.0152) by V in Eq. (A1). For calculation of Ps
we utilize Shibues (2003) correlation
ln

P 
s

Pc

Tc
[a1  + a2  1.5 + a3  3 + a4  3.5 + a5  4 + a6  7.5 ]
T

(A7)

where  = 1 T/Tc and Tc and Pc are critical temperature and pressure of water.
a1 = 7.85951783, a2 = 1.84408259, a3 = 11.7866497,
a4 = 22.6807411, a5 = 15.9618719, a6 = 1.80122502.

Please cite this article in press as: 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 in subsurface CO2 storage. Int. J. Greenhouse Gas Control (2012),
http://dx.doi.org/10.1016/j.ijggc.2012.07.025

G Model
IJGGC-688; No. of Pages 14

ARTICLE IN PRESS
Z. Ziabakhsh-Ganji, H. Kooi / International Journal of Greenhouse Gas Control xxx (2012) xxxxxx

References
AAAS (The American Association for the Advancement of Science), 2009. Carbon
sequestration. Science 325, 16441645.
Akinev, N.N., Diamond, L.W., 2003. Thermodynamic description of aqueous
nonelectrolytes at innite dilution over a wide range of state parameters.
Geochimica et Cosmochimica Acta 67 (4), 613627.
Amirijafari, B., Campbell, J.M., 1972. Solubility of gaseous hydrocarbon mixtures in
water. Society of Petroleum Engineers Journal 1 (1), 2127.
Alvarez, J., Prini, R.F., 1991. A semiempirical procedure to describe the thermodynamics of dissolution of non-polar gases in water. Fluid Phase Equilibria 66,
309326.
Andr, L., Audigane, P., Azaroual, M., Menjoz, A., 2007. Numerical modeling of
uidrock chemical interactions at the supercritical CO2 liquid interface during
supercritical CO2 injection into a carbonate reservoir, the Dogger aquifer (Paris
Basin, France). Energy Conversion and Management 48, 17821797.
AQUAlibrium, FlowPhase, 2010. Calgary, http://www.telusplanet.net/public/
jcarroll/AQUA.HTM.
Bachu, S., Adams, J.J., 2003. Estimating CO2 sequestration capacity in solution in deep
saline aquifers. Energy Conversion and Management 44 (20), 31513175.
Bachu, S., 2008. CO2 storage in geological media: role, means, status and barriers to
deployment. Progress in Energy and Combustion Science 34, 254273.
Battistelli, A., Ceragioli, P., Marcolini, M., 2011. Injection of acid gas mixtures in sour
oil reservoirs: analysis of near-wellbore processes with coupled modeling of
well and reservoir ow. Transport in Porous Media 90, 233251.
Battistelli, A., Marcolini, M., 2009. TMGAS: a new TOUGH2 EOS module for the
numerical simulation of gas mixtures injection in geological structures. International Journal of Greenhouse Gas Control 3, 481493.
Birkholzer, J.T., Zhou, Q., Tsang, C.-F., 2009. Large-scale impact of CO2 storage in deep
saline aquifers: a sensitivity study on pressure response in stratied systems.
International Journal of Greenhouse Gas Control 3, 181194.
Blanco, L.H., Smith, N.O., 1978. High-pressure solubility of methane in aqueous calcium-chloride and aqueous tetraethylammonium bromidepartial molar
properties of dissolved methane and nitrogen in relation to water-structure.
Journal of Physical Chemistry 8 (2), 186191.
Blount, C.W., Price, L.C., 1982. Solubility of methane in water under natural conditions: a laboratory study. DOE Contract No. DE-A508-78ET12145, Final Report.
Chapoy, A., Mohammadi, A.H., Tohidi, B., Richon, D., 2004. Gas solubility measurement and modeling for the nitrogen plus water system from 274.18 K to
363.02 K. Journal of Chemical and Engineering Data 49, 11101115.
Chapoy, A., Mohammadi, A.H., Tohidi, B., Richon, D., 2005. Estimation of water
content for methane + water and methane + ethane + n-butane + water systems
using a new sampling device. Journal of Chemical and Engineering Data 50,
11571161.
Cramer, S.D., 1984. Solubility of methane in brines from 0-degrees-C to 300-degreesC. Industrial & Engineering Chemistry Process Design and Development 2 (3),
533538.
Crovetto, R., Fernandez-Prini, R., Japas, M.L., 1982. Contribution of the cavityformation or hard-sphere term to the solubility of simple gases in water. Journal
of Physical Chemistry 8 (21), 40944095.
Culberson, O.L., Mcketta, J.J., 1951. Phase equilibria in hydrocarbonwater systems
3. The solubility of methane in water at pressures to 10,000 Psia. Transactions of the American Institute of Mining and Metallurgical Engineers 192,
223226.
Culberson, O.L., Horn, A.B., Mcketta, J.J., 1950. Phase equilibria in hydrocarbonwater
systems: the solubility of ethane in water at pressures to 1200 pounds per
square inch. Transactions of the American Institute of Mining and Metallurgical
Engineers 189, 16.
Danesh, A., 1998. PVT and Phase Behavior of Petroleum Reservoir Fluids. Elsevier.
Dhima, A., Hemptinne, J.C.D., Moracchini, G., 1998. Solubility of light hydrocarbons
and their mixtures in pure water under high pressure. Fluid Phase Equilibria
145, 129150.
Dodson, C.R., Standing, M.B., 1944. Pressurevolumetemperature and solubility
relations for natural gaswater mixtures. Drilling and Production Practice, API,
173178.
Drummond, S.E., 1981. Boiling and mixing of hydrothermal uids: chemical effects
on mineral precipitation. Ph.D. Dissertation. Pennsylvania State University, State
College, PA.
Duan, Z., Mao, S., 2006. A thermodynamic model for calculating methane solubility,
density and gas phase composition of methane-bearing aqueous uids from
273 to 523 K and from 1 to 2000 bar. Geochimica et Cosmochimica Acta 70,
33693386.
Duan, Z., Sun, R., 2003. An improved model calculating CO2 solubility in pure water
and aqueous NaCl solutions from 273 to 533 K and from 0 to 2000 bar. Chemical
Geology 193, 257271.
Duan, Z., Sun, R., Liu, R., Zhu, C., 2007. Accurate thermodynamic model for the calculation of H2 S solubility in pure water and brines. Energy & Fuels 21, 20562065.
Duffy, J.R., Smith, N.O., Nagy, B., 1961. Solubility of natural gases in aqueous salt
solutions: I. Liquidus surfaces in the system CH4 H2ONaClCaCl2 at room temperatures and at pressures below 1000 psia. Geochimica et Cosmochimica Acta
2 (12), 2331.
Fine, R.A., Millero, F.J., 1973. Compressibility of water as a function of temperature
and pressure. Journal of Chemical Physics 59 (10), 55295536.
Gaus, I., 2010. Role and impact of CO2 rock interactions during CO2 storage in sedimentary rocks. International Journal of Greenhouse Gas Control 4, 7389.

13

Geloni, C., Giorgis, T., Battistelli, A., 2011. Modeling of rocks and cement alteration
due to CO2 injection in an exploited gas reservoir. Transport in Porous Media 90,
183200.
Ghaderi, S.M., Keith, D.W., Lavoie, R., Leonenko, Y., 2011. Evolution of hydrogen
sulde in sour saline aquifers during carbon dioxide sequestration. International
Journal of Greenhouse Gas Control 5 (2), 347355.
Gillespie, P.C., Wilson, G.M., 1982. Vaporliquid and liquidliquid equilibria:
watermethane, watercarbon dioxide, waterhydrogen sulde, waternpentane, watermethanen-pentane. Research report RR-48, Gas Processors
Association, Tulsa, OK.
Goodman, J.B., Krase, N.W., 1931. Solubility of nitrogen in water at high pressures
and temperatures. Industrial and Engineering Chemistry 23 (4), 401404.
Gunter, W.D., Perkins, E.H., Hutcheon, I., 2000. Aquifer disposal of acid gases: modeling of waterrock reactions for trapping of acid wastes. Applied Geochemistry
15, 10851095.
Hassanzadeh, H., Pooladi-Darvish, M., Elsharkawy, A.M., Keith, D.W., Leonenko,
Y., 2008. Predicting PVT data for CO2 brine mixtures for black-oil simulation
of CO2 geological storage. International Journal of Greenhouse Gas Control 2,
6577.
Huang, S.S., Leu, A.D., Ng, H.I., Robinson, D.B., 1985. The phase behaviour of two
mixtures of methane, carbon dioxide, Hydrogen sulde, and water. Fluid Phase
Equilibria 19, 2132.
IPCC, 2007. In: Solomon, S., Qin, D., Manning, M., Chen, Z., Marquis, M., Averyt, K.B.,
Tignor, M., Miller, H.L. (Eds.), Contribution of Working Group I to the Fourth
Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge, United Kingdom/New York.
Jacquemet, N., Gallo, Y.L., Estublier, A., Lachet, V., von Dalwigk, I., Yan, J., Azaroual,
M., Audigane, P.w, 2009. CO2 streams containing associated components a
review of the thermodynamic and geochemical properties and assessment of
some reactive transport codes. Energy Procedia 1, 37393746.
Kiepe, J., Horstmann, S., Fischer, K., Gmehling, J., 2003. Experimental determination
and prediction of gas solubility data for methane + water solutions containing different monovalent electrolytes. Industrial and Engineering Chemistry
Research 42, 53925398.
King, M.B., Mubarak, A., Kim, J.D., Bott, T.R., 1992. The mutual, solubilities of water
with supercritical and liquid carbon dioxide. Journal of Supercritical Fluids 5,
296302.
Knauss, K.G., Johnson, J.W., Steefel, C.I., 2005. Evaluation of the impact of CO2 , cocontaminant gas, aqueous uid and reservoir rock interactions on the geologic
sequestration of CO2 . Chemical Geology 217, 339350.
Lee, J.I., Mather, A.E., 1977. Solubility of hydrogen sulde in water. Berichte der
Bunsen-Gesellschaft Physical Chemistry Chemical Physics 81, 10201023.
Lekvam, K., Bishnoi, P.R., 1997. Dissolution of methane in water at low temperatures
and intermediate pressures. Fluid Phase Equilibria 131, 297309.
Li, H., Yan, J., 2009. Evaluating cubic equations of state for calculation of vaporliquid
equilibrium of CO2 and CO2 -mixtures for CO2 capture and storage processes.
Applied Energy 86, 826836.
Mao, S., Duan, Z., 2006. A thermodynamic model for calculating nitrogen solubility,
gas phase composition and density of the N2 H2 ONaCl system. Fluid Phase
Equilibria 248, 103114.
Michels, A., Gerver, J., Bijl, A., 1936. The inuence of pressure on the solubility of
gases. Physica 3, 797808.
Muller, G., Bender, E., Maurer, G., 1988. Das Dampf-Flussigkeits gleichgewicht des
ternaren Systems Ammoniak-Kohlendioxid- Wasser bei holen Wassergehalten
im Bereich zwischen 373 und 473 K. Ber. Bunsenges. Phys. Chem. 92, 148160.
Namiot, Yu.A., Bondareva, M.M., 1959. Rastvorimost gazov v vode. Gostekhizdat
(Quoted in Mao, S., Duan, Z., 2006), Moscow (In Russian).
Namiot, A.Y., Skripka, V.G., Ashmyan, K.D., 1979. Inuence of water dissolved
salt upon methane solubility under the temperatures 50 to 350-degrees-C.
Geokhimiya 1, 147149.
Nighswander, J.A., Kalogerakis, N., Mehrotra, A.K., 1989. Solubilities of carbon dioxide in water and 1 wt.% NaCl solution at pressures up to 10 MPa and temperatures
from 80 to 200 C. J. Chem. Eng. Data 34, 355360.
Oldenburg, C.M., Bryant, S.L., Nicot, J., 2009. Certication framework based on
effective trapping for geological carbon sequestration. International Journal of
Greenhouse Gas Control 3, 444457.
Olds, R.H., Sage, B.H., Lacey, W.H., 1942. Phase equilibria in hydrocarbon systems:
composition of the dew-point gas of the methanewater system. Industrial and
Engineering Chemistry Research 34 (10), 12231227.
OSullivan, T.D., Smith, N.O., 1970. Solubility and partial molar volume of nitrogen
and methane in water and in aqueous sodium chloride from 50 to 125 degrees
and 100 to 600 atm. Journal of Physical Chemistry 74 (7), 14601466.
Peng, D.Y., Robinson, D.B., 1976. A new two-constant equation of state. Industrial
and Engineering Chemistry Fundamentals 15, 5964.
Pitzer, K.S., 1973. Thermodynamics of electrolytes: 1. Theoretical basis and general
equations. J. Phys. Chem. 77, 268277.
Prausnitz, J.M., Lichtenthaler, R.N., Gomes, D.A.E., 1986. Molecular Thermodynamics
of Fluid-Phase Equilibria, 2nd ed. Prentice Hall, New York.
Pruess, K., Garcia, J., Kovscek, T., Oldenburg, C., Rutqvist, J., Steefel, C., Xu, T., 2004.
Code intercomparison builds condence in numerical simulation models for
geologic disposal of CO2 . Energy 29 (910), 14311444.
Pruess, K., 2005. ECO2N: A TOUGH2 uid property module for mixtures of water,
NaCl and CO2. Lawrence Berkeley National Laboratory Report.
Pruess, K., Oldenburg, C., Moridis, G., 1999. TOUGH2 Users Guide V. 2. 0. LBNL Report
43134. Lawrence Berkeley National Laboratory, Berkeley, CA.

Please cite this article in press as: 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 in subsurface CO2 storage. Int. J. Greenhouse Gas Control (2012),
http://dx.doi.org/10.1016/j.ijggc.2012.07.025

G Model
IJGGC-688; No. of Pages 14
14

ARTICLE IN PRESS
Z. Ziabakhsh-Ganji, H. Kooi / International Journal of Greenhouse Gas Control xxx (2012) xxxxxx

Prutton, C.F., Savage, R.L., 1945. The solubility of carbon dioxide in calcium chloride
water solutions at 75, 100, 120 C and high pressures. J. Am. Chem. Soc. 67,
15501554.
Qin, J., Rosenbauer, R.J., Duan, Z., 2008. Experimental measurements of vaporliquid
equilibrium of the H2 O + CO2 + CH4 ternary system. Journal of Chemical and
Engineering Data 53, 12461249.
Redlich, O., Kwong, J.N.S., 1949. On the thermodynamics of solutions: V. An equation
of state. Fugacities of gaseous solutions. Chemical Reviews 44, 233244.
Rumpf, B., Maurer, G., 1992. Solubilities of hydrogen cyanide and sulfur dioxide in
water at temperatures from 293.15 to 413.15 K and pressures up to 2.5 MPa.
Fluid Phase Equilibria 81, 241260.
Rumpf, B., Nicolaisen, H., Ocal, C., Maurer, G., 1994. Solubility of carbon dioxide in
aqueous solutions of sodium chloride: experimental results and correlation. J.
Solution Chem. 23 (3), 431448.
Saddington, A.W., Krase, N.W., 1934. Vaporliquid equilibria in the system
nitrogenwater. Journal of the American Chemical Society 56, 353361.
Schmidt, G., Wenzel, H., 1980. A modide Van der Waals type equation of state.
Chem. Eng. Sci. 35, 15031512.
Selleck, F.T., Carmichael, L.T., Sage, B.H., 1952. Phase behavior in the hydrogen
suldewater system. Industrial and Engineering Chemistry 44, 22192226.
Shibue, Y., 2003. Vapor pressure of aqueous NaCl and CaCl2 solutions at elevated
temperatures. Fluid Phase Equilibria 213, 3951.
Song, K.Y., Feneyrou, G., Fleyfel, F., Martin, R., Lievois, J., Kobayashi, R., 1997. Solubility
measurements of methane and ethane in water at and near hydrate conditions.
Fluid Phase Equilibria 128, 249260.
Soreide, I., Whitson, C.H., 1992. PengRobinson predictions for hydrocarbons,
CO2 N2 , and H2 S with pure water and NaCl brine. Fluid Phase Equilibria 77,
217240.
Smith, N.O., Kelemen, S., Nagy, B., 1962. Solubility of natural gases in aqueous salt
solutionsII: Nitrogen in aqueous NaCl, CaCl2 Na2 SO4 and MgSO4 at room temperatures and at pressures below 1000 Psia. Geochimica et Cosmochimica Acta
26, 9219260.
Smith, H.J., Fahrenkamp-Uppenbrink, J., Coontz, R., 2009. Clearing the air. Science
325, 1641.
Soave, G., 1972. Equilibrium constants from a modied RedlichKwong equation of
state. Chemical Engineering Science 27, 11971203.
Spycher, N., Pruess, K., Ennis-King, J., 2003. CO2 H2 O 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 67,
30153031.
Spycher, N., Pruess, K., 2005. CO2 H2 O mixtures in the geological sequestration of
CO2 : II. Partitioning in chloride brines at 12100 C and up to 600 bar. Geochimica et Cosmochimica Acta 69, 33093320.
Stoessell, R.K., Byrne, P.A., 1982b. Salting-out of methane in single-salt solutions
at 25-degrees-C and below 800 psia. Geochimica et Cosmochimica Acta 46 (8),
13271332.
Suleimenov, M., Krupp, R.E., 1994. Solubility of hydrogen sulde in pure water and
in NaCl solutions, from 20 to 320 C and at saturation pressures. Geochimica et
Cosmochimica Acta 58 (11), 24332444.

Todheide, K., Franck, E.U., 1963. Das Zweiphasengebiet und die kritische Kurve im
System Kohlendioxid-Wasser bis zu Drucken von 3500 bar. Z. Phys. Chem. 37,
387401.
White, S.P., Allis, R.G., Moore, J., Chidsey, T., Morgan, C., Gwynn, W., Adams, M., 2005.
Simulation of reactive transport of injected CO2 on the Colorado Plateau, Utah,
USA. Chemical Geology 217 (34), 387405.
White, C.M., Smith, D.H., Jones, K.L., Goodman, A.L., Jikich, S.A., La Count, R.B., DuBose,
S.B., Yarym-Agaev, N.L., Sinyavskaya, R.P., Koliushko, I.L., Levinton, L.Y., 1985.
Phase-equilibria in the water methane and methanol methane binary-systems
under high-pressures. Journal of Applied Chemistry of the USSR 58 (1), 154157.
Wang, L.K., Chen, G.J., Han, G.H., Guo, X.Q., Guo, T.M., 2003. Experimental study on the
solubility of natural gas components in water with or without hydrate inhibitor.
Fluid Phase Equilibria 207, 143154.
Wiebe, R., Gaddy, V.L., 1939. The Solubility in water in carbon dioxide at 50, 75,
and 100 C, at pressures to 700 atmospheres. Journal of the American Chemical
Society 61, 315318.
Wiebe, R., Gaddy, V.L., 1940. The solubility of carbon dioxide in water at various
temperatures from 12 to 40 and at pressures to 500 atmospheres Critical phenomena. Journal of the American Chemical Society 62, 815817.
Wiebe, R., Gaddy, V.L., Heins, C., 1932. Solubility of nitrogen in water in 250c from
25 to 1000 atmospheres. Industrial and Engineering Chemistry 24 (8), 927927.
Wiebe, R., Gaddy, V.L., Heins, C., 1933. The solubility of nitrogen in water at 50, 75 and
100 from 25 to 1000 atmospheres. Journal of the American Chemical Society
55 (3), 947953.
Xia, J., Kamps, A.P.-S., Rumpf, B., Maurer, G., 2000. Solubility of hydrogen sulde in
aqueous solutions of the single salts sodium sulfate, ammonium sulfate sodium
chloride, and ammonium chloride at temperatures from 313 to 393 K and total
pressures up to 10 MPa. Industrial and Engineering Chemistry Research 39,
10641073.
Xia, J., Rumpf, B., Maurer, G., 1999. 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 165, 99119.
Xu, T., Apps, J.A., Pruess, K., Yamamoto, H., 2007. Numerical modeling of injection and
mineral trapping of CO2 with H2 S and SO2 in a sandstone formation. Chemical
Geology 242, 319346.
Yokoyama, C., Wakana, S., Kaminishi, G.-I., Takahashi, S., 1988. Vaporliquid equilibria in the methanediethylene glycolwater system at 298.15 and 323.15 K.
Journal of Chemical and Engineering Data 33, 274276.
Zawisza, A., Malesinska, B., 1981. Solubility of carbon dioxide in liquid water and of
water in gaseous carbon dioxide in the range 0.25 MPa and at temperatures up
to 473 K. J. Chem. Eng. Data 26 (4), 388391.
Zhang, W., Xu, T., Li, Y., 2011. Modeling of fate and transport of coinjection of H2 S
with CO2 in deep saline formations. Journal of Geophysical Research 116, 113.
Zirrahi, M., Azin, R., Hassanzadeh, H., Moshfeghian, M., 2010. Prediction of water
content of sour and acid gases. Fluid Phase Equilibria 299, 171179.
Zirrahi, M., Azin, R., Hassanzadeh, H., Moshfeghian, M., 2012. Mutual solubility of
CH4 , CO2 H2 S, and their mixtures in brine under subsurface disposal conditions.
Fluid Phase Equilibria 324, 8093.

Please cite this article in press as: 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 in subsurface CO2 storage. Int. J. Greenhouse Gas Control (2012),
http://dx.doi.org/10.1016/j.ijggc.2012.07.025

Das könnte Ihnen auch gefallen