Vous êtes sur la page 1sur 11

See

discussions, stats, and author profiles for this publication at: http://www.researchgate.net/publication/248360458

Solubility product of siderite (FeCO 3) as a


function of temperature (25250 C)

ARTICLE in CHEMICAL GEOLOGY JULY 2009


Impact Factor: 3.52 DOI: 10.1016/j.chemgeo.2009.03.015

CITATIONS READS

37 287

3 AUTHORS, INCLUDING:

Pascale Bnzeth
French National Centre for Scientific Resea
94 PUBLICATIONS 1,413 CITATIONS

SEE PROFILE

All in-text references underlined in blue are linked to publications on ResearchGate, Available from: Pascale Bnzeth
letting you access and read them immediately. Retrieved on: 31 December 2015
Chemical Geology 265 (2009) 312

Contents lists available at ScienceDirect

Chemical Geology
j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / c h e m g e o

Solubility product of siderite (FeCO3) as a function of temperature (25250 C)


P. Bnzeth , J.L. Dandurand, J.C. Harrichoury
Universit de Toulouse, UPS (OMP), LMTG, 14 Avenue Edouard Belin, 31400 Toulouse, France
CNRS, LMTG, 31400 Toulouse, France
IRD, LMTG, 31400 Toulouse, France

a r t i c l e i n f o a b s t r a c t

Article history: The solubility of natural siderite was investigated from 25 to 250 C in 0.1 mol kg 1 NaCl aqueous solutions using a
Received 25 January 2008 hydrogen-electrode concentration cell, which provided continuous in situ measurement of hydrogen ion molality.
Received in revised form 9 March 2009 Iron(II) was analyzed by a revised Ferrozine-spectrophotometric method. The obtained apparent solubility
Accepted 9 March 2009
products (Q sp-siderite) were extrapolated to innite dilution to generate the solubility products (Ksp-siderite),
allowing calculation of the thermodynamic properties of siderite. Of all the temperature functions tested, the
Keywords:
Siderite
equation giving a reliable t of our data in the investigated temperature range (25250 C) has the following form:
Solubility product log10 Ksp-siderite =a +b (T / K) +c (T /K) 1 +d log10 (T/K) with: a = 175.568, b = 0.0139, c = 6738.483 and
Carbonates d = 67.898. Based on this equation and its rst and second derivatives with respect to T, we were able to derive
CO2 sequestration values for the Gibbs energy of formation: f Go298.15 = (680.71 2) kJ mol 1, enthalpy of formation: f Ho298.15 =
(749.59 2) kJ mol 1, entropy: So298.15 = (109.54 2) J mol 1 K 1 and heat capacity: Cp298.15 o
= (83.26 2) J
1 1 o o
mol K of siderite (uncertainties are 3). The values of f G298.15 and f H298.15 are in very good agreement with
the values reported by Robie et al. [Robie, R.A., Haselton, H.T. Jr., Hemingway, B.S., 1984. Heat capacities and
entropies of rhodochrosite (MnCO3) and siderite (FeCO3) between 5 and 600 K. Am. Mineral. 69, 349357] and
Chai and Navrotsky [Chai, L., Navrotsky, A., 1994. Enthalpy of formation of siderite and its application in phase
equilibrium calculation. Am. Mineral. 79, 921929], respectively. The density model [Anderson, G.M., Castet, S.,
Schott, J., Mesmer, R.E., 1991. The density model for estimation of thermodynamic parameters of reactions at high
temperatures and pressures. Geochim. Cosmochim. Acta 55, 17691779] reproduced correctly our experimental
data and allowed the extrapolation of the siderite solubility product up to 350 C by using our values of the Gibbs
energy and enthalpy of formation of siderite combined with its entropy and the heat capacity equation given by
Robie et al. [Robie, R.A., Haselton, H.T. Jr., Hemingway, B.S., 1984. Heat capacities and entropies of rhodochrosite
(MnCO3) and siderite (FeCO3) between 5 and 600 K. Am. Mineral. 69, 349357].
2009 Elsevier B.V. All rights reserved.

1. Introduction (Romanek et al., 1994; Valley et al., 1997; Treiman and Romanek, 1998)
and interplanetary dust particles (Keller et al., 1994). Siderite
Iron is the fourth most abundant metal on Earth occurring in a formation is known to be facilitated by both mesophilic and
variety of rock and soil minerals in oxidation states (II) and (III). Under thermophilic iron reducing bacteria (e.g., Zhang et al., 2001), and
anoxic conditions, the concentration of ferrous iron (Fe(II) or Fe2+) is has been interpreted to be microbially mediated in many natural
frequently controlled by the ferrous carbonate, siderite (FeCO3(cr)), environments (see Mortimer and Coleman, 1997).
according to a reaction that can be expressed as: Lately, siderite has also been mentioned as potential CO2 mineral
trapping in numerous computer simulations of CO2 geological
2 + 2
FeCO3cr fFe + CO3 1 sequestration (e.g., Johnson et al., 2002; Xu et al., 2003; Palandri
and Kharaka, 2005; Palandri et al., 2005; Zerai et al., 2006). This was
Siderite is commonly found in hydrothermal veins associated with conrmed in supercritical CO2brinerock experiments performed at
uorite, galena and barite and as concretions in shales and sandstones 200 C by Kaszuba et al. (2005) and in the study of Palandri et al.
(Mozley, 1989; Munoz et al., 1999; Munoz et al., 2005). Siderite has (2005). Glauconitic, arkosic and illitic sediments are discussed in
also been identied in extraterrestrial materials, such as meteorites these previous studies for their potential to trap CO2 in siderite,
ankerite (CaFe(CO3)2), calcite (CaCO3), magnesite (MgCO3) or
eventually dawsonite (NaAlCO3(OH)2). However, these sediments,
Corresponding author. Gochimie et Biogochimie Exprimentale, Universit de
Toulouse; UPS (SVT-OMP); CNRS; LMTG, 14 Avenue Edouard Belin, 31400 Toulouse,
in particular Fe(II)-bearing glauconitic beds, are generally of limited
France. Tel.: +33 5 61 33 26 17; fax: +33 5 61 33 25 60. thickness and geographical occurrence as oppose to other iron-rich
E-mail address: benezeth@lmtg.obs-mip.fr (P. Bnzeth). sediments (including redbeds), which in addition present great

0009-2541/$ see front matter 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.chemgeo.2009.03.015
4 P. Bnzeth et al. / Chemical Geology 265 (2009) 312

Table 1 smoothed regression of the heat capacity data corrected for deviations
Literature values of the siderite solubility product Ksp-siderite (Eq. (1)). from exact stoichiometry of the solid, which differed from the value
t (C) log10Ksp-siderite References that can be calculated at the same temperature from Eq. (2) below
30 10.46 Smith (1918) (82.25 J mol 1 K 1), also reported by Robie et al. (1984) and Robie and
30 10.55 Recalculated value from the data of Smith (1918) (see text) Hemingway (1995):
17 10.12 Singer and Stumm (1970)
22.5 10.22 o 1 1
25 10.24
Cp = J mol K = 257:38 0:04620 T = K
0:5 6 2
30 10.25 3081:9 T = K + 1:523 10 T = K
20 10.40 Bardy and Pr (1976)
2
50 11.20 Reiterer et al. (1981), Extrapolated to innite dilution
by Bruno et al. (1992)
25 10.60 Calculated by Robie et al. (1984) from Smith (1918) Chai and Navrotsky (1994) determined the enthalpy of formation
and Langmuir (1969)
of siderite at 25 C by drop calorimetric measurements of enthalpies
30 11.11 Braun (1991)
40 11.27 of decomposition of FeCO3(cr) at 978 and 1075 K. The value of
50 11.88 f Ho298.15 = 750.60 kJ mol 1 obtained by these authors is in good
60 12.43 agreement with Helgeson et al. (1978). However, the literature values
70 12.31 of the standard enthalpy of formation differ by more than 10 kJ mol 1
80 12.57
25 10.80 Bruno et al. (1992)
(see Table 2). Recently, Preis and Gamsjger (2002) performed a
25 10.77 Greenberg and Tomson (1992) critical evaluation of solubility data and enthalpy of formation of
43 10.94 siderite and gave a new optimized value of f Ho298.15 = 752.00 kJ
62 11.03 mol 1. They also provided support for the validity of the properties
83 11.25
of Fe2+ in the survey performed by Parker and Khodakovskii (1995).
94 11.42
25 10.90 Silva et al. (2002) In this study, we measured the solubility products of Eq. (1) from
25 10.93 Ptacek and Reardon (1992) 25 to 250 C in 0.1 mol kg 1 NaCl aqueous solutions using a hydrogen-
25 11.03 Ptacek and Blowes (1994) electrode concentration cell (HECC) and re-evaluated the thermo-
25 11.03 Jensen et al. (2002) dried crystals dynamic properties of siderite from these solubility measurements.
25 10.43 Jensen et al. (2002) wet crystals

2. Materials and methods

2.1. Starting solutions and materials


thickness and high porosity and permeability. As iron must be in its
ferrous oxidation state to precipitate into carbonate minerals, All solutions were prepared from reagent grade chemicals and
sedimentary Fe(III) requires a reductant, which may be organic distilled deionized water (resistivity 0.18 M m). For the test solution,
matter, sulfur dioxide (SO2) or hydrogen sulde (H2S) (see for
example, Palandri et al., 2005; Palandri and Kharaka, 2005). Never-
theless, in order to model CO2 mineral trapping in these geological
reservoirs, it is essential to know precisely the solubility product and
thermodynamic properties of siderite (this paper) and its dissolution/
precipitation rates (see Golubev et al., 2009-this issue).
A number of previous studies have focused on the determination of
the solubility product of FeCO3(cr) at low temperature (b90 C),
various ionic strengths (from 0.1 to 1.0 mol kg 1 NaClO4 or 0.1 to
5.5 mol kg 1 NaCl medium) and CO2 pressure from 0.01 to 0.93 bars
(e.g., Smith, 1918; Singer and Stumm, 1970; Bardy and Pr, 1976;
Reiterer et al., 1981; Braun, 1991; Bruno et al., 1992; Greenberg and
Tomson, 1992; Ptacek and Reardon, 1992; Ptacek and Blowes, 1994;
Jensen et al., 2002; Silva et al., 2002). A summary of literature values
for the solubility product of Eq. (1) is given in Table 1 and the values
are reported as a function of reciprocal temperature in Fig. 1. It can be
seen from this gure that the values at 25 C are dispersed, ranging
from 10 11.03 to 10 10.24. Few values are also available for tempera-
ture greater than 30 C, including the data of Reiterer et al. (1981)
(50 C), Braun (1991) (3080 C) and Greenberg and Tomson (1992)
(2594 C), with large scatters among the values obtained from one
set to the other. The standard state thermodynamic properties of
siderite at 298.15 K, 1 bar are tabulated in a number of thermo-
chemical data compilations (Wagman et al., 1982; Robie and
Hemingway, 1995). Most of them are reported in Table 2, including,
at absolute temperature T, the Gibbs energy of formation f Go298.15, the
enthalpy of formation f Ho298.15, the entropy So298.15 and the heat
capacity Cop298.15. Robie et al. (1984) measured the heat capacity of
siderite by adiabatic calorimetry from 5 to 380 K, and by differential
scanning calorimetry in the temperature range 350550 K. The heat
capacity and entropy reported by these authors at 298.15 K are 82.44
0.10 J mol 1 K 1 and 95.47 0.15 J mol 1 K 1, respectively. Note that Fig. 1. logKsp-siderite values from the literature as a function of the reciprocal of
this value of the heat capacity was deduced by Robie et al. (1984) from a temperatures.
P. Bnzeth et al. / Chemical Geology 265 (2009) 312 5

Table 2
Standard state properties of siderite and aqueous species involved in Eq. (1) at 25 C and 1 bar.

Species f Go(a)
298.15 So(b)
298.15 f Ho(a)
298.15
o(b)
Cp298.15 References
FeCO3(cr) 749.70 Helgeson et al. (1978, from Langmuir, 1969)
741.90 Reiterer et al. (1981)
666.67 92.90 740.57 82.13 Wagman et al. (1982)
680.03 0.6 95.47 0.15 753.22 0.6 82.44 0.1 Robie et al. (1984)
95.50 761.20 0.9 82.30 Holland and Powell (1990)
750.60 1.1 Chai and Navrotsky (1994)
682.80 95.50 755.90 5.5 82.25 Robie and Hemingway (1995)
752.00 1.2 Preis and Gamsjger (2002)
680.71 2 109.54 2 749.59 2 83.26 2 This study
Fe2+ 90.53 1 101.60 3.7 90.00 0.5 Parker and Khodakovskii (1995)
33.05 Shock et al. (1997)
CO2
3 527.98 50.00 675.24 289.53 Shock and Helgeson (1988)

The values are in bold because they are from this study (to differentiate from literature ones).
When not specied, uncertainties correspond to 3.
(a)
in kJ mol 1.
(b)
in kJ mol 1.

deionized water was thoroughly sparged with argon for 1 h (see point Krypton adsorption, BET method) was 7.08 m2 g 1 for the
details below in the experimental procedure section). Concentrated siderite from Montroc, and 1.05 m2 g 1 for the siderite from
stock solutions of NaCl and HCl were used to make up the desired Peyrebrune.
experimental solutions. The composition of the starting reference and
test solutions for each experimental run is given in Table 3. 2.2. Experimental procedure
The siderite starting material was from the Montroc uorite
deposit (southern Massif Central, France) described by Munoz et al. 2.2.1. The hydrogen-electrode concentration cell (HECC)
(1999, 2005) and from Peyrebrune, Quarry, France. Both siderites The design and function of the HECC have been described in
were manually extracted from the rock sample containing respec- numerous publications (e.g., Bnzeth et al., 1997, 1999; Palmer et al.,
tively assemblages of uoritequartzchalcopyrite (for Montroc) and 2001; Bnzeth et al., 2007). The present version shown in Fig. 4 is a
quartzchalcopyritesphalerite (for Peyrebrune). The Montroc's side- newly cell constructed at LMTG based on the HECC described in details
rite (named Sid-A in Table 3) was used in run #1. The other runs were by Palmer et al. (2001). It consists of a 300 ml capacity Stainless Steel
performed with the siderite from Peyrebrune (named Sid-B in 316 pressure vessel containing two concentric Teon cups separated
Table 3). After gently grinding the siderite crystals with an agate by a porous Teon plug, which acts as a liquid junction completing the
mortar and pestle, the obtained powder was submitted to sedimenta- electric circuit. Teon-insulated platinum electrodes (coated with
tion cycles in alcohol to remove ultra ne particles and sieved to platinum black) protrude into each cup where the solutions are stirred
obtain various fractions. Only the fraction between 10 and 100 m was magnetically using Teon-coated magnets. The solution in the inner
retained and stored in alcohol to prevent oxidation. Before each cup serves as the reference of known stoichiometric hydrogen ion
experiment, few grams of the stored powder (~1 to 2.5 g) were dried molality, whereas the outer, or test solution, contains a suspension of
in a dessicator for several hours. Powder X-ray Diffraction (XRD, siderite (initially ca. 1 to 2.5 g of solid in ca. 100 ml of solution). The
Fig. 2a), Scanning Electron Microscopy (SEM, Fig. 2b), and Energy head space of the pressure vessel is thoroughly purged with hydrogen
Dispersive Spectrometry (EDS) examinations indicated the absence of at ambient temperature. Upon completion of the hydrogen purge, the
any phases other than small amount of quartz. Differential Scanning hydrogen pressure is set at ~10 bars at 25 C prior placing the pressure
Calorimetry (DSC) was performed on siderite samples. Fig. 3 shows vessel in the block tube furnace for equilibration at the experimental
the results of the DSC realized on the siderite from Peyrebrune in air temperature. Note that the effect on the dissociation constant of water
and under argon atmosphere from 30 C to 1000 C at 20 C/min. (as an example) for a 10 bar pressure change at 25 C or a 20 bar
Siderite undergoes an exothermic reaction at around 530 C during change at 300 C (where the solvent is much more compressible) is
heating in air, which is assigned to the sum of endothermic trivial, less than the uncertainties assigned to the values. In some runs,
decarbonation and exothermic oxidation of the decarbonated iron prior purging the system, the stirred test solution was bubbled with
oxide. In an argon atmosphere, the reaction is endothermic and H2 through the sampling line and released from the head of the vessel
corresponds to the decarbonation of siderite, without any further to optimize the O2 removal from the solution. A simulation test was
exothermic oxidation reaction. The surface area (measured by multi- performed before a run showing that distilled deionized water purged
with argon for one hour still contained 3% of O2 with regard to the
saturation (0.3 mg l 1) as measured with an O2 probe. After
bubbling with H2 for 30 min (inside the pressure vessel), the water
Table 3
only contained 0.6% of O2 (0.06 mg l 1). Following this purging
Stoichiometric molal compositions (m in mol kg 1 H2O) of starting solutions,
temperature (t) ranges and solid phase used in each run (Sid-A: Montroc, Massif process, the hydrogen line is put back onto its inlet port and both
Central, France; Sid-B: Peyrebrune, Quarry, France). solutions are sparged again with H2 few times before regulating the
pressure to approximately 10 bars at 25 C.
Run # t (C) range Reference cell Test cell Siderite
investigated sample
The initial conguration of the cell in a typical experiment
103 103
containing identical HCl (m1) and NaCl (m2) solutions in the reference
m(HCl) m(NaCl) m(HCl) m(NaCl)
and test cups is as follows:
1 10025 1.9986 0.0981 1.9994 0.0980 Sid-A
2 20075 1.9984 0.0981 1.9984 0.0981 Sid-B
3 25100 2.0009 0.0979 2.0009 0.0979 Sid-B H2 ; Ptj HClm1 ; NaClm2 j jHClm1 ; NaClm2 ; FeCO3cr jPt; H2
4 250 1.9985 0.0981 1.9985 0.0981 Sid-B Reference jj Test
5 150200 2.0072 0.0979 2.0072 0.0979 Sid-B
6 25 1.9993 0.0980 1.9993 0.0980 Sid-B where the ratio m1/m2 is b0.1 in order to minimize both liquid
4 bars of CO2 injected at 25 C. junction contribution to the measured potential and activity
6 P. Bnzeth et al. / Chemical Geology 265 (2009) 312

and the difference in potential between the electrodes is described by


the Nernst equation:

E = RT = FlnaH test = aH ref + Elj 4

where E is the measured cell potential; F, R, and T represent the


Faraday constant, the universal gas constant and the temperature in
Kelvin, respectively. aH+ is the activity of the hydrogen ion and Elj refers
to the liquid-junction potential based on the full Henderson equation
(Baes and Mesmer, 1986), which involves the limiting conductivities
of the individual ions. The limiting equivalent conductances (o) of Na+,
Cl, H+ and OH are taken from Quist and Marshall (1965), whereas the
o of Fe2+ is assumed to be the same as that for Mg2+ (Quist and
Marshall, 1965). Note that Kubota et al. (1988) determined the limiting
conductivity of Fe2+ at 25 C from the measurements of the conductivity
of Fe(II) sulphate aqueous solution. The obtained value is comparable to
those of various divalent metals (Mn2+, Co2+, Ni2+, Cu2+ and Zn2+)
including the value for Mg2+ determined by Quist and Marshall (1965),
which supports our assumption.
During the steady state achievement, the cell potential, the tem-
perature and the pressure are recorded on a computer via an Agilent
34970A Data Acquisition Switch Unit. In most of the experiments,
the solutions were initially allowed to equilibrate with the solid at
a starting pHm ranging from 4 to 6 and at temperature indicated in
Table 3. Once the cell attained thermal equilibrium, solution samples
were withdrawn from the test solution over time (see Table 4). During
each sampling episode, the stirring motor was turned off (ca. 30 min)
to allow the solid to settle, whereupon approximately 1 ml of test
solution was withdrawn (and discarded) via an Hastelloy valve,
through a platinum dip tube provided at its bottom (see Fig. 4) with a
porous Teon frit (35 m) (to prevent particles from entering
the platinum tube that could act as seeds for precipitation during the
sampling process), and then through a 13 mm, 0.2 m polyvinylidene
uoride lter (PVDF Acrodisc LC13). 2 to 4 ml samples were then
collected into sterilized, polypropylene/polyethylene syringes con-
taining a known mass of high purity 0.1 mol l 1 HCl (double distilled
HCl) for subsequent analyses of iron by spectrophotometry using
a modied Ferrozine method (see Section 2.2.2. for details).
Additional samples (~5 ml) were periodically taken for alkalinity
analysis (see Section 2.2.3.). Note that a new porous Teon frit was
used for each run. Equilibrium was attained within ca. 24 h (at
temperature t 100 C); nevertheless samples were taken at time
intervals of several days whereupon temperature was decreased (see

Fig. 2. (a) X-ray Diffractogram (CuK radiation) and (b) Scanning Electronic Microscopy
micrographs (top: 100 m scale and bottom: close-up view) of the 10100 m fraction of
the natural siderite from the Montroc uorite deposit (southern Massif Central, France).

coefcient differences between the two solutions. Note that the


working denition of pHm throughout this paper is pHm = log [H+]
(stoichiometric molal concentration units). Each platinum-hydrogen
electrode responds to the half cell reaction (with H2(g) standing for
gaseous hydrogen):

+ Fig. 3. Differential Scanning Calorimetry (DSC) of the Peyrebrune siderite performed in


H2g = 2H + 2e 3 (a) air and (b) argon atmosphere (20 C/min).
P. Bnzeth et al. / Chemical Geology 265 (2009) 312 7

Fig. 4. Schematic of the hydrogen-electrode concentration cell (HECC).

Table 4), approaching then equilibrium from under-saturation as the acid)-1,2,4-triazine, monosodium salt, monohydrate) has been used to
solubility of siderite increase with decreasing temperature. In some determine Fe(II) concentration spectrophotometrically. The method
runs (runs #3 and 5), the temperature was raised (from 25 to 100 C recently revisited by Viollier et al. (2000), was optimized in this
and from 150 to 200 C, respectively) approaching then equilibrium study to be able to analyze Fe(II) in acidic solutions. The effect of pH,
from over-saturation. Furthermore, in run #5, the experiment was the nature of acids (HNO3, HCl), and NaCl concentration was also
performed with previous addition, at room temperature, of 4 bars of investigated in this work. Absorbance measurements were made with
pure CO2 and 10 bars of H2 (total pressure of 14 bars), after the a UV visible dual-beam spectrophotometer (Varian Scan-50) with a
hydrogen purging procedure. 10 mm width polypropylene cell. A 10 2 mol l 1 Ferrozine (97%,
Aldrich) solution was prepared in 0.1 mol l 1 ammonium acetate
2.2.2. Analysis of Fe(II) (CH3COONH4) solution. The reducing agent for Fe(III) analyses was
A number of sensitive analytical methods are available for the 1.4 mol l 1 hydroxylamine hydrochloride (H2NOH.HCl, 99.9999%,
determination of Fe(II). Some of the most commonly used include Aldrich), prepared in a solution of 2 mol l 1 analytical grade hydro-
spectrophotometry (e.g., Sung and Morgan, 1990; Viollier et al., 2000) chloric acid. The buffer solution is an ammonium acetate (10 mol l 1)
with 1,10-phenanthroline or Ferrozine, capillary electrophoresis solution adjusted to pH 9.5 with a solution of ammonium hydroxide
(Pozdniakova et al., 1997), chemiluminiscence (see King et al., 1995) (2830% NH4OH, JT Baker). Fe(III) standard solutions were pre-
and ow injection analysis (Measures et al., 1995). Since Stookey pared from a 1000 g g 1 Fe(III) stock solution (Fe(NO3)3 in HNO3,
(1970), Ferrozine iron reagent (3-(2-pyridyl)-5,6-bis(4-phenylsulfonic 0.5 mol l 1, Merck Certipur) and Fe(II) standards were prepared
8 P. Bnzeth et al. / Chemical Geology 265 (2009) 312

Table 4 0.05 mol l 1 HCl. The two sets of Fe(III) and Fe(II) standards
Experimental results for siderite solubility experiments in NaCl media. prepared in 0.5 mol l 1 HNO3 (or HCl, not shown) are compared in
Run # t (C) Time of log[H+]a log [Fe2+]ameas [Alkalinity]a [CO2
3 ]calc
a Fig. 5 (open symbols) with Fe(II) standards directly prepared in
sampling measured in situ measured 108 0.05 mol l 1 HCl or 0.05 mol l 1 HNO3 (lled symbols). The rst
(h) at 25 C observation is that no complexation with Ferrozine of Fe(II) standards
1a 99.3 16 5.690 3.289 1.859 (prepared initially in 0.5 mol l 1 HNO3) occurred during step 1 (A1 = 0)
1b 99.4 112 5.778 3.298 2.596
of the procedure described above, unlike standards prepared in less
1c 99.1 136 5.791 3.307 0.00024 2.579
1d 74.6 16 5.922 3.252 5.099 concentrated HNO3 or HCl solutions (see lled circle symbols in Fig. 5).
1e 74.4 58 5.977 3.227 6.559 This observation implies that Fe(II) is fully oxidized after 24 h by the
1f 74.3 153 5.959 3.271 5.554 concentrated nitric acid (0.5 mol l 1) as after the reduction step the
1g 49.4 36 6.075 3.167 8.597 concentration of Fe(II) (open circle symbols in Fig. 5) falls on the linear
1h 49.6 80 6.100 3.163 9.462
1i 49.3 154 6.076 3.176 0.00034 8.430
function of Fe(III)-reduced standards prepared in concentrated HNO3
1j 26.9 19 6.091 3.149 6.014 (open square symbols). The same procedure was performed by using
1k 26.5 67 6.120 3.135 6.806 concentrated HCl (0.5 mol l 1) and no effect has been observed for Fe
1l 26.8 139 6.146 3.115 7.857 (II) analysis (not shown in Fig. 5). The purpose of the second test we
1m 26.7 188 6.365 3.136 15.26
performed was to analyze the effect of NaCl concentration. As an
2a 198.8 20 5.474 3.301 0.129
2b 198.0 68 5.521 3.324 0.00014 0.155 example, Fig. 6 shows the results obtained with Fe(III) standards
2c 148.7 63 5.467 3.283 0.404 solutions prepared in 0.05 mol l 1 HCl/0.1 mol l 1 NaCl and in
2d 148.7 107 5.476 3.257 0.00016 0.447 0.025 mol l 1 HCl/0.05 mol l 1 NaCl. As can be seen in Fig. 6 when
2e 74.0 33 5.633 3.149 2.151 compared to standards without NaCl (circle, crossed symbols), there is
2f 73.9 105 5.682 3.131 2.720
no effect of NaCl on the analysis of iron up to 0.1 mol l 1 NaCl.
2g 74.0 177 5.705 3.119 3.066
2h 74.0 273 5.715 3.115 3.212
3a 26.4 144 6.232 3.005 13.327 2.2.3. Alkalinity
3b 26.7 192 6.300 2.993 17.269 Analysis of alkalinity ([HCO 2 +
3 ] + 2[CO3 ] + [OH ] [H ], where [i]
3c 26.4 336 6.364 2.974 21.961
refers to the ith species molal concentration) was carried out at 25 C
3d 26.5 480 6.399 2.965 25.027
3e 26.5 597 6.426 2.961 27.485 immediately after sampling by acidimetric titration of c.a. 48 ml of
3f 98.2 87 5.866 3.091 0.00014 1.535 ltered solutions taken without acidication. The alkalinity was
4a 251.5 24 6.002 3.844 0.0941 measured using either a Schott titrator instrument (Titroline alpha
4b 251.5 47 6.010 3.861 0.0938 plus, TA10plus), or by hand acidimetric titration using the Gran Function
4c 251.5 115 5.993 3.915 0.0769
plot method. Total dissolved inorganic carbon is calculated at 25 C from
5a 149.1 72 4.776 3.062 0.0005 0.2619
5b 149.3 240 4.787 3.073 0.0004 0.2303 the measured pHm at this temperature, the titration result and the
5c 198.0 144 4.745 3.140 0.00017 0.0261 speciation of carbonates generated by using the apparent dissociation
5d 198.0 240 4.740 3.140 0.00018 0.0345 constants of water in NaCl media measured by Busey and Mesmer
5e 198.0 312 4.728 3.149 0.00017 0.0341
(1978) and the apparent hydrolysis constants (Q) of the following
6a 25.4 256 6.035 2.986 0.00053 8.844
6b 25.3 424 6.081 2.980 8.269
reactions:
a
Molal concentrations in the experimental solutions. The subscript letters associated

to run numbers correspond to the successive time of sampling. CO2aq + H2 O = HCO3 + H 5
2
HCO3 = H + CO3 6

freshly before analyses from solid FeCl24H2O dissolved in acidic


at the ionic strength of interest. The value of Q 5 (= [HCO +
3 ][H ] /
solution (HCl or HNO3). The analytical procedure is as follow:
[CO2](aq)) was from Patterson et al. (1982) and Q 6 (=[CO2 3 ][H+
]/

1) 3 ml of each standard (Fe(II) and/or Fe(III) from 1 to 5 g g 1) are [HCO3 ]) from Patterson et al. (1984). The speciation of carbonates is
added to 300 l of Ferrozine and 50 l of buffer. The addition of then recalculated at the temperature of the experiments using the
this amount of buffer is part of our optimization compared to total dissolved inorganic carbon calculated at 25 C combined with the
Viollier et al. (2000) to ensure the formation of Fe(II)-Ferrozine
complex, which occurs between pH 4 and 9. The absorbance (A1) is
then recorded at 563.0 nm for Fe(II) analysis. In such condition, no
absorbance is recordable for Fe(III) standard solutions.
2) An aliquot of 2.4 ml is then taken from above and 450 l of the
reducing agent solution is added. A reaction time of 10 min allows
the complete reduction of Fe(III).
3) 150 l of buffer solution is then added and the absorbance (A2)
recorded at 563.0 nm, which corresponds to total iron determina-
tion [Fe]T, yielding the concentration of Fe(III) by difference ([Fe]T
(A2)[Fe(II)] (A1)).

The detection limit of this method, using a 10 mm cell, is 0.5 g g 1


and the uncertainty of the analysis is better than 5%.
Following the procedure above, various tests were performed. The
rst one was to check the effect of HNO3 and HCl concentrated
solutions on Fe redox status. Concentrated nitric acid or hydro-
chloric acid (0.5 mol l 1) were used to prepared stock solutions of Fe
(II) and Fe(III) standards (from 10 to 40 g g 1). 24 h later, 10
folds dilutions of these stock solutions were realized with deionized Fig. 5. Absorbance versus Fe concentration for oxidation test with concentrated
water to obtain standards from 1 to 4 g g 1 in 0.05 mol l 1 HNO3 or standard solution of HNO
3 .
P. Bnzeth et al. / Chemical Geology 265 (2009) 312 9

where (NaCl) is the mean molal stoichiometric activity coefcient of


NaCl. These latter values were calculated from Archer (1992) and are
reported in Table 5 together with the ionic strength and the calculated
solubility products (Ksp-siderite).
After some runs, the solid phases were analyzed by XRD. During run
#1 (Sid-A) and run #5 (Sid-B), parts of the solid changed color from gray-
light brown to red, and tiny amount of goethite associated with siderite
was detected by XRD (see Fig. 7 as an example for Sid-A, run #1). For the
solid Sid-B, even though the solution has been initially bubbled with
hydrogen, the color of the solid generally changed, to a certain extent,
from gray-light brown to grey-black. The solid was analyzed by XRD,
which showed traces of magnetite (see Fig. 8, run #2 as an example).
The logarithms of the solubility products taken from Table 5 are
reported in Fig. 9 as a function of the reciprocal of temperature and
compared with the literature values (reported in Fig. 1, using the same
symbols). As can be seen, our value at 25 C is in good agreement with
Fig. 6. Absorbance versus Fe(III) concentration for NaCl test. some of the previous studies (see also Table 1 for clarity) but in
disagreement with the data of Singer and Stumm (1970), the recent
experimental measured pHm and the hydrolysis constants taken from value of Jensen et al. (2002), particularly when using wet crystal, and
the above authors at the temperature of interest. the values of Smith (1918) (30 C) and Bardy and Pr (1976) (20 C).
In the study of Smith (1918), no inert ionic medium was used, which
3. Results and discussion leads to a variation of the ionic strength during the course of the
experiment. Moreover, Smith calculated the solubility product of
The results of the solubility experiments obtained for each run are siderite (as reported in Table 1) by using the rst and second
listed in Table 4, which show the measured temperature, the time of ionization constants of carbonic acid (Eqs. (5), (6)), both at 25 C
sampling (in hours) and pHm (log[H+]), along with the measured (instead of 30 C) and innite dilution. We recalculated the value of
concentration of Fe2+ and alkalinity (when available). The last column Smith (1918) at 30 C using the constants of reactions (5) and (6) from
reports the carbonate (CO2 3 ) species concentrations calculated as Patterson et al. (1982, 1984, respectively) and found a value of 10.55
explained above or in most of the cases by considering that [Fe2+] = (reported in Table 1 and Figs. 1 and 9). In the work of Singer and
[carbonates] when alkalinity is not available. Note that this Stumm (1970), a synthesized siderite precipitated at low temperature
assumption was veried when alkalinity was measured and no CO2 (17 and 30 C) was used in free-drift experiments. Their experimental
injected. Note also that for all the temperatures and pHm investigated procedures were described briey and no information was provided
in this study, Fe2+ ion was the dominant iron species. No Fe3+ was on the evolution of solution chemistry as a function of time. However,
detected by spectrophotometry at any temperature and pH. From this was the rst study where equilibrium was approached from
Table 4, several observations can be made, which allowed us to select super-saturation, which leads to the highest values at low tempera-
the runs and data to be used for the regression procedure discussed ture as can be seen in Figs. 1 and 9. In another study from Greenberg
afterwards. At temperature t 100 C, equilibrium is reached very and Tomson (1992), equilibrium was also approached from super-
quickly (no more than 24 h). At lower temperatures, at least 3 to saturation. It can be seen again that higher values are obtained from
5 days for 50 t 75 C, and 18 days at 25 C are necessary to ensure 25 to 94 C, when compared to our data. In the present study,
that equilibrium has been reached. As mentioned earlier (see equilibrium was approached twice from over-saturation by increasing
experimental procedure section), equilibrium was approached at
some temperature by both under-saturation and super-saturation (for
Table 5
example at 150 C, runs #2 and 5, respectively). For run #5 and
Siderite apparent solubility products, Qsp-siderite, at I 0.1 molal NaCl, ionic strengths,
t N 150 C, difculties were encountered to measure accurately the I, activity coefcients, (NaCl), and siderite solubility products, Ksp-siderite, calculated for
alkalinity due to the injection of CO2(g) (see Table 3). Rapid degassing, each run.
along with a pH drift, occurred at the beginning of the alkalinity
Run # t (C) log10Qsp-siderite I mol kg 1 (NaCl) log10Ksp-sideritea
analysis inducing a large uncertainty in the calculation of total
1b 99.4 10.88 0.0981 0.7469 11.90 0.2
carbonates, subsequently in the calculated solubility product. For this 1c 99.1 10.89 0.0981 0.7471 11.91 0.2
reason runs #5c, 5d and 5e were not used in the tting procedure. 1d 74.6 10.54 0.0981 0.7599 11.50 0.3
The apparent solubility products for Eq. (1), dened as Q sp-siderite = 1f 74.3 10.53 0.0981 0.7600 11.48 0.3
[Fe2+][CO23 ], are reported in Table 5 (only for the selected data as
1g 49.4 10.23 0.0981 0.7703 11.14 0.4
1h 49.6 10.19 0.0981 0.7702 11.10 0.4
explained above) and at specic ionic strength I. The ionic strength and
1i 49.3 10.25 0.0980 0.7704 11.16 0.4
the carbonate speciation were determined iteratively from the mea- 2a 198.8 12.19 0.0995 0.6691 13.59 0.2
sured pHm, total ferrous iron, and carbonate concentrations reported in 2b 198.0 12.13 0.0995 0.6699 13.53 0.2
Table 4. The solubility products (at innite dilution), Ksp-siderite, can then 2c 148.7 11.68 0.0993 0.7143 12.84 0.3
be expressed as: 2d 148.7 11.61 0.0994 0.7142 12.78 0.3
2g 74.0 10.63 0.0996 0.7601 11.59 0.3
   2h 74.0 10.61 0.0996 0.7600 11.56 0.3
2+ 2
Ksp siderite = Qspsiderite Fe CO3 7 3f 98.2 10.91 0.0997 0.7474 11.92 0.2
4a 251.5 12.87 0.0994 0.6043 14.62 0.3
4b 251.5 12.89 0.0995 0.6040 14.64 0.3
where i is the activity coefcients of the ith aqueous species. The 4c 251.5 13.03 0.0995 0.6037 14.78 0.3
mean stoichiometric activity coefcients were then derived from the 5a 149.1 11.64 0.0999 0.7130 12.82 0.3
Meissner equation (Lindsay, 1989) with the implicit assumption that 5b 149.3 11.71 0.0999 0.7130 12.89 0.3
6a 25.4 10.04 0.1000 0.7771 10.92 0.4
for an ion of charge z:
6b 25.3 10.06 0.1000 0.7771 10.94 0.4
a
z2 Uncertainties calculated from the combined experimental uncertainties. The subscript
j z j = FNaCl 8 letters associated to run numbers correspond to the successive time of sampling.
10 P. Bnzeth et al. / Chemical Geology 265 (2009) 312

Fig. 9. Logarithms of siderite solubility product, log10Ksp-siderite, (Eq. (1)) at innite


Fig. 7. X-ray Diffractogram (CuK radiation) of siderite (Sid-A) after run #1 showing
dilution obtained in this study compared with literature data (symbols as in Fig. 1) as a
few peaks of goethite (FeOOH).
function of reciprocal T.

temperature during runs #3 and 5, conrming (especially for run #3


for which the increase of temperature is larger) that steady state has increasing temperature and non-linearly with 1/T. Therefore, of all the
been reached. In the work of Jensen et al. (2002), super-saturated temperature functions tested, the equation giving a reliable t of our
solutions at 25 C did not reach equilibrium with respect to data (Table 5 and Fig. 9) has the form:
precipitated siderite after 474 days. However, their experiments
with re-suspension of the precipitated crystals reached equilibrium 1
log10 Ksp siderite = a + b T = K + c T = K + d log10 T = K
after 10 days, but their solubility product depended on whether the
9
crystals were dried at 105 C (called dried crystal in Table 1 and Fig. 1)
or kept wet before re-suspension (called wet crystal in Table 1 and
Fig. 1). Bruno et al. (1992) obtained equilibrium at 25 C after only The regression then yielded the four coefcients: a = 175.568,
three days of equilibration, which seems to be fast compared to our b=0.0139, c= 6738.483 and d =67.898 corresponding to the solid
study where, following Fe(II) concentration as a function of time at curve shown in Figs. 9 and 11. This equation yields the thermodynamic
25 C, steady state appears to be reached only after about 18 days, as values for reaction (1) at 25 C,1 bar: rGo =(R ln(10){aT+bT 2 +c +
can be seen from Table 4 and Fig. 10 (in the case of run #3). Note that dT logT}) / 1000 = 62.204 kJ mol 1, and log10Kspsiderite = 10.90;
the data of run #3 at 25 C were only used to ensure that equilibrium rHo = (R ln(10){bT2 c + dT/ln(10)}) / 1000 = 15.654 kJ mol 1;
was reached, but not used in the regression. The other experimental rCpo = R ln(10){2b T + d/ln(10)} = 405.845 J mol 1 K 1 and
data available at temperature above 25 C are from Reiterer et al. consequently rSo = 261.136 J mol 1 K 1. Combining these values
(1981) at 50 C (value extrapolated by Bruno et al., 1992 at innite with the thermodynamic data of Fe2+ and CO2 3 from respectively
dilution, see Table 1), in good agreement with our value, and from Parker and Khodakovskii (1995), Shock et al. (1997) and Shock and
Braun (1991), obtained from 30 to 80 C and deviating considerably Helgeson (1988) (reported in Table 2), we obtained for siderite at
from our experimental values at temperature higher than 40 C. 25 C, 1 bar: f Go298.15 = ( 680.71 2) kJ mol 1; f Ho298.15 =
Finally, it can be seen from Fig. 9 that values of log10Ksp-siderite, ( 749.59 2) kJ mol 1; So298.15 = (109.54 2) J mol 1 K 1, and
obtained herein over the temperature range 25250 C, decrease with
o
Cp298.15 = (83.26 2) J mol 1 K 1 (uncertainties are 3). These
values are reported in Table 2 for comparison with literature data. The
values of the Gibbs energy of formation and heat capacity of siderite

Fig. 8. X-ray Diffractogram (CuK radiation) of siderite (Sid-B) after run #2 showing
few peaks of magnetite (Fe3O4). Fig. 10. Fe(II) concentration as a function of time at 25 C (run #3).
P. Bnzeth et al. / Chemical Geology 265 (2009) 312 11

dawsonite (Bnzeth et al., 2007). However, it should be emphasized


that thermodynamic properties of a solid phase particularly when
deduced from solubility measurements are highly dependent on
various factors: 1) the nature of the solid phase plays a major role in
solubility experiments. Most previous solubility studies used a
synthesized siderite precipitated at low temperature, with probably
a poor crystallinity and very ne grains; 2) the duration of experiment
and attainment of equilibrium; 3) the temperature range and the
equation used to t the data; for instance we encountered that
without the results of the highest temperature runs (t N 150 C), a
function of the form: log10Ksp-siderite = a + b T + c T 1 could be
used but gave a poorest t and unreliable thermodynamic data,
particularly the enthalpy of formation and heat capacity of siderite; 4)
the availability of suitable and reliable thermodynamic data for the
aqueous species involved in the equilibrium; 5) nally, part of the
differences in the solubility products can also be related to the
different methods used to extrapolated at innite dilution the results
obtained in many previous works at high ionic strength (1.0 mol kg 1
Fig. 11. Logarithms of siderite solubility product obtained in this study compared with
values generated from the density models (I and II, see text) and from Helgeson (1969) NaClO4). Moreover, solubility measurements of minerals containing
and SUPCRT92 (Johnson et al., 1992). reduced chemical elements could be complicated by oxidation of the
released aqueous species and subsequent precipitation of (hydr)
oxides, which are orders of magnitude less soluble than the carbonate
minerals.
are in good agreement with the values of Robie et al. (1984) and Robie
and Hemingway (1995). We also obtained a good agreement Acknowledgments
for the value of the enthalpy of formation when compared with the
experimental one obtained by Chai and Navrotsky (1994), and the one Research sponsored by the French national programs: INSU
recommended recently by Preis and Gamsjger (2002) (see Table 2). (Institut National des Sciences de l'Univers du CNRS) and ANR
However, our entropy value is higher compared to the one directly (National Agency for Research, within the framework of the
measured by Robie et al. (1984), which should be preferred over the Geocarbone-Carbonatation Project under contract ANR05CO2-009-
one we obtained from the second derivative of the t of the solubility 02). Michel Thibaut, Alain Castillo and Thierry Aigouy (LMTG) are
product versus reciprocal of temperature. thanked for their assistances with XRD, BET and SEM analyses,
The density model (Anderson et al., 1991) has been used to respectively. The paper was improved by constructive comments from
extrapolate our data to higher temperature (up to 350 C), using the two anonymous reviewers and Bndicte Mnez.
heat capacity equation (Eq. (2)) from Robie et al. (1984) combined
with our values of the Gibbs energy and enthalpy of formation of References
siderite and either the entropy of siderite determined in this study or
by Robie et al. (1984) (density models I and II, respectively, see Fig. 11). Anderson, G.M., Castet, S., Schott, J., Mesmer, R.E., 1991. The density model for
estimation of thermodynamic parameters of reactions at high temperatures and
A comparison of the dependence of log10Ksp-siderite on reciprocal pressures. Geochim. Cosmochim. Acta 55, 17691779.
temperature is shown in Fig. 11 including the experimental values and Archer, D.G., 1992. Thermodynamic properties of the NaCl + H2O system. II. Thermo-
the t derived from Eq. (9) (red symbols and solid line), the values dynamic properties of NaCl(aq), NaCl2H2O(cr), and phase equilibria. J. Phys. Chem. Ref.
Data 21, 793829.
generated by the density models I and II (green and blue dashed lines,
Baes Jr., C.F., Mesmer, R.E., 1986. The Hydrolysis of Cations (Reprint with corrections of
respectively), the values proposed by Helgeson (1969) (dashed dot the 1976 edition by John Wiley and Sons, New York). Robert E. Krieger Publishing
dot line) and calculated by using the database from SUPCRT92 Co., Inc., Malabar, Florida. 489 pp.
Bardy, J., Pr, C., 1976. Dtermination exprimentale du coefcient de solubilit du
(Johnson et al., 1992) (black dashed line). The density model I
carbonate ferreux en milieux aqueux. Trib. CEBEDEAU 29 (387), 7581.
reproduces correctly our experimental data, whereas model II begins Bnzeth, P., Palmer, D.A., Wesolowski, D.J., 1997. The aqueous chemistry of aluminum.
to deviate above 75 C, however within the uncertainty limits. A new approach to high-temperature solubility measurements. Geothermics 26,
Moreover, good agreement is also obtained with the data reported 465481.
Bnzeth, P., Palmer, D.A., Wesolowski, D.J., 1999. The solubility of zinc oxide at 0.03 m
by Helgeson (1969) between 25 and 150 C, with signicant NaTr as a function of temperature, with in situ pH measurement. Geochim.
deviations only becoming apparent above 175 C down to 300 C. Cosmochim. Acta 63, 15711586.
Finally the values generated from SUPCRT92 are systematically higher Bnzeth, P., Palmer, D.A., Anovitz, L.M., Horita, J., 2007. Dawsonite synthesis and
reevaluation of its thermodynamic properties from solubility measurements:
than all the other values reported in Fig. 11 from 25150 C, and then implications for mineral trapping of CO2. Geochim. Cosmochim. Acta 71, 44384455.
become lower above 225 C. This difference is mainly due to Braun, R.D., 1991. Solubility of iron (II) carbonate at temperatures between 30 and 80 C.
inconsistent data sets used to compute these values. Talanta 38, 205211.
Bruno, J., Wersin, P., Stumm, W.,1992. On the inuence of carbonate in mineral dissolution:
II. The solubility of FeCO3(s) at 25 C and 1 atm total pressure. Geochim. Cosmochim.
4. Conclusion Acta 56, 11491155.
Busey, R.H., Mesmer, R.E., 1978. Thermodynamic quantities for the ionization of water in
sodium chloride media to 300 C. J. Chem. Eng. Data 23, 175176.
In this study we measured the solubility of siderite from 25 to Chai, L., Navrotsky, A., 1994. Enthalpy of formation of siderite and its application in
250 C, and 0.1 mol kg 1 ionic strength solutions. An empirical phase equilibrium calculation. Am. Mineral. 79, 921929.
equation for the temperature dependence of the siderite solubility Greenberg, J., Tomson, M., 1992. Precipitation and dissolution kinetics and equilibria of
aqueous ferrous carbonate vs temperature. Appl. Geochem. 7, 185190.
constants (Eq. (9)) was derived and the parameters deduced from this
Golubev, S.V., Bnzeth, P., Schott, J., Dandurand, J.L., Castillo, A., 2009. Siderite dissolution
regression were used to generate the thermodynamic properties of kinetics in acidic aqueous solutions from 25 to 100 C and 0 to 50 atm pCO2. Chem.
siderite. The values obtained are in good agreement with previous Geol. 265 (1-2), 1319 (this issue).
data obtained from calorimetric measurements, providing that Helgeson, H.C., 1969. Thermodynamics of hydrothermal systems at elevated tempera-
tures and pressures. Am. J. Sci. 267, 729804.
reliable thermodynamic data can be obtained from solubility Helgeson, H.C., Delany, J.M., Nesbitt, H.W., Bird, D.K., 1978. Summary and critique of the
measurements as, for example, recently demonstrated in the case of thermodynamic properties of rock-forming minerals. Am. J. Sci. 278-A, 1229.
12 P. Bnzeth et al. / Chemical Geology 265 (2009) 312

Holland, T.J.B., Powell, R., 1990. An enlarged and updated internally consistent Ptacek, C.J., Reardon, E.J., 1992. Solubility of siderite (FeCO3) in concentrated NaCl and
thermodynamic dataset with uncertainties and correlations: the system Na2O Na2SO4 solutions at 25 C. WaterRock Interaction, Proc. Int. Symp., 7th, pp. 181183.
K2OCaOMgOMnOFeOFe2O3Al2O3TiO2SiO2CH2O2. J. Metamorph. Geol. Ptacek, C.J., Blowes, D.W., 1994. Inuence of siderite on the pore-water chemistry of
8, 89124. inactive mine-tailings impoundments. In: Alpers, C.N., Blowes, D.W. (Eds.),
Jensen, D.L., Boddum, J.K., Tjell, J.C., Christensen, T.H., 2002. The solubility of Environmental Geochemistry of Sulde Oxidation. American Chemical Society,
rhodochrosite (MnCO3) and siderite (FeCO3) in anaerobic aquatic environments. Washington, DC, pp. 172189.
Appl. Geochem. 17, 503511. Quist, A.S., Marshall, W.L., 1965. Assignment of limiting equivalent conductances for
Johnson, J.W., Oelkers, E.H., Helgeson, H.C., 1992. SUPCRT92: a software package for single ions to 400 C. J. Phys. Chem. 69, 29842987.
calculating the standard molal thermodynamic properties of minerals, gases, Reiterer, F., Johannes, W., Gamsjger, H., 1981. Semimicro determination of solubility
aqueous species, and reactions from 1 to 5000 bar and 0 to 1000 C. Comput. Geosci. constants: copper (II) carbonate and iron (II) carbonate. Mikrochim. Acta 1, 6372.
18, 899947. Robie, R.A., Hemingway, B.S., 1995. Thermodynamic Properties of Minerals and Related
Johnson, J.W., Nitao, J.J., Steefel, C.I., 2002. Fundamental elements of geologic CO2 Substances at 298.15 K and 1 bar (105 Pascals) Pressure and at Higher
sequestration in saline aquifers. ACS Fuel Chem. Div. Symp. Prepr. 47, 4142. Temperatures. US Geological Survey Bulletin, Washington D.C., vol. 1231. 461 pp.
Kaszuba, J.P., Janecky, D.R., Snow, M.G., 2005. Experimental evaluation of mixed uid Robie, R.A., Haselton Jr., H.T., Hemingway, B.S., 1984. Heat capacities and entropies of
reactions between supercritical carbon dioxide and NaCl brine: relevance to the rhodochrosite (MnCO3) and siderite (FeCO3) between 5 and 600 K. Am. Mineral. 69,
integrity of a geologic carbon repository. Chem. Geol. 217, 277293. 349357.
Keller, L.P., Thomas, K.L., McKay, D.S., 1994. The nature of carbon-bearing phases in Romanek, C.S., Grady, M.M., Wright, I.P., Mittlefehldt, D.W., Socki, R.A., Pillinger, C.T.,
hydrated interplanetary dust particles. Meteoritics 29, 480481. Gibson Jr., E.K., 1994. Record of uid-rock interactions on Mars from the meteorite
King, W.D., Lounsbury, H.A., Millero, F.J., 1995. Rates and mechanism of Fe(II) oxidation ALH84001. Nature 372, 655657.
at nanomolar total iron concentrations. Environ. Sci. Technol. 29, 818824. Shock, E.L., Helgeson, H.C., 1988. Calculation of the thermodynamic and transport
Kubota, E., Mochizuki, Y., Yokoi, M., 1988. Conductivity of iron(II) sulfate in aqueous properties of aqueous species at high pressures and temperatures: correlation
solution at various temperatures. Bull. Chem. Soc. Jpn. 61, 37233724. algorithms for ionic species and equation of state predictions to 5 kb and 100 C.
Langmuir, D., 1969. The Gibbs free energies of substances in the system FeO2H2O Geochim. Cosmochim. Acta 52, 20092036.
CO2 at 25 C. U. S. Geol. Surv. Prof. Pap. 650-B. Shock, E.L., Sassani, D.C., Willis, M., Sverjensky, D.A., 1997. Inorganic species in geologic
Lindsay Jr., W.T., 1989. Chemistry of steam cycle solutions: principles. In: Cohen, P. (Ed.), uids: correlations among standard molal thermodynamic properties of aqueous
The ASME Handbook on Water Technology for Thermal Power Plants. The American ions and hydroxide complexes. Geochim. Cosmochim. Acta 61, 907950.
Society of Mechanical Engineers, New York, pp. 341544. Chapter 7. Silva, C.A.R., Liu, X., Millero, F.J., 2002. Solubility of siderite (FeCO3) in NaCl solutions.
Measures, C.I., Yuan, J., Resing, J.A., 1995. Determination of iron in seawater by ow J. Solution Chem. 31, 97108.
injection analysis using in-line preconcentration and spectrophotometric detec- Singer, P.C., Stumm, W., 1970. The solubility of ferrous iron in carbonate-bearing waters.
tion. Mar. Chem. 50, 312. J. Am. Water Works Assoc. 62, 198202.
Mortimer, R.J.G., Coleman, M.L., 1997. Microbial inuence on the oxygen isotopic Smith, H.J., 1918. On equilibrium in the system ferrous carbonate, carbon dioxide, and
composition of diagenetic siderite. Geochim. Cosmochim. Acta 61, 17051711. water. J. Am. Chem. Soc. 40, 879883.
Mozley, P.S., 1989. Relation between depositional environment and the elemental Stookey, L.L., 1970. Ferrozine: a new spectrophotometric reagent for iron. Anal. Chem.
composition of early diagenetic siderite. Geology 17, 704706. 42, 779781.
Munoz, M., Boyce, A.J., Courjault-Rade, P., Fallick, A.E., Tollon, F., 1999. Continental Sung, W., Morgan, J.J., 1990. Kinetics and product of ferrous iron oxygenation in aqueous
basinal origin of ore uids from southwestern Massif Central uorite veins systems. Environ. Sci. Technol. 14, 561568.
(Albigeois, France): evidence from uid inclusion and stable isotope analyses. Appl. Treiman, A.H., Romanek, C.S., 1998. Bulk and stable isotopic composition of carbonate
Geochem. 14, 447458. minerals in Martian meteorite Allan Hills 84001: no proof of high formation
Munoz, M., Premo, W.R., Courjault-Rade, P., 2005. SmNd dating of uorite from World temperature. Meteorit. Planet. Sci. 33, 737742.
class Montroc uorite deposit, southern Massif Central, France. Miner. Depos. 39, Valley, J.W., Eiler, J.M., Graham, C.M., Gibson Jr, E.K., Romanek, C.S., Stolper, E.M., 1997.
970975. Low-temperature carbonate concretions in the Martian meteorites ALH 84001:
Palandri, J.L., Kharaka, Y.K., 2005. Ferric iron-bearing sediments as a mineral trap for CO2 evidence from stable isotopes and mineralogy. Science 275, 16331638.
sequestration: iron reduction using sulfur-bearing waste gas. Chem. Geol. 217, Viollier, E., Inglett, P.W., Hunter, K., Roychoudhury, A.N., Van Cappellen, P., 2000. The
351364. ferrozine method revisited: Fe(II)/Fe(III) determination in natural waters. Appl.
Palandri, J.L., Rosenbauer, R.J., Kharaka, Y.K., 2005. Ferric iron in sediments as a novel Geochem. 15, 785790.
CO2 mineral trap: CO2SO2 reaction with hematite. Appl. Geochem. 20, 20382048. Wagman, D.D., Evans, W.H., Parker, V.B., Schumm, R.H., Halow, I., Bailey, S.M., Churney,
Palmer, D.A., Bnzeth, P., Wesolowski, D.J., 2001. Aqueous high temperature solubility K.L., Nuttal, R.L., 1982. The NBS tables of chemical thermodynamic properties.
studies. I. The solubility of boehmite at 150 C as a function of ionic strength and pH as Selected values for inorganic C1 and C2 organic substances in SI units. J. Phys. Chem.
determined by in situ measurements. Geochim. Cosmochim. Acta 65, 20812095. Ref. Data 11, 1392.
Parker, V.B., Khodakovskii, I.L., 1995. Thermodynamic properties of the aqueous ions (2+ Xu, T., Apps, J.A., Pruess, K., 2003. Reactive geochemical transport simulation to study
and 3+) of iron and the key compounds of iron. J. Phys. Chem. Ref. Data 24, 16991745. mineral trapping for CO2 disposal in deep arenaceous formations. J. Geophys. Res.
Patterson, C.S., Slocum, G.H., Busey, R.H., Mesmer, R.E., 1982. Carbonate equilibria in 108 (B2), 2071.
hydrothermal systems: rst ionization of carbonic acid in NaCl media to 300 C. Zhang, C.L., Horita, J., Cole, D.R., Zhou, J., Lovley, D.R., Phelps, T.J., 2001. Temperature-
Geochim. Cosmochim. Acta 46, 16531663. dependent oxygen and carbon isotope fractionation of biogenic siderite. Geochim.
Patterson, C.S., Busey, R.H., Mesmer, R.E., 1984. Second ionization of carbonic acid in Cosmochim. Acta 65, 22572271.
NaCl media to 250 C. J. Solution Chem. 13, 647661. Zerai, B., Saylor, B.Z., Matisoff, G., 2006. Computer simulation of CO2 trapped through
Pozdniakova, S., Padarauskas, A., Schwedt, G., 1997. Simultaneous determination of iron mineral precipitation in the Rose Run Sandstone, Ohio. Appl. Geochem. 21,
(II) and iron (III) in water by capillary electrophoresis. Anal. Chim. Acta 351, 4148. 223240.
Preis, W., Gamsjger, H., 2002. Critical evaluation of solubility data: enthalpy of
formation of siderite. Phys. Chem. Chem. Phys. 4, 40144019.

Vous aimerez peut-être aussi