You are on page 1of 9

Applied Geochemistry 39 (2013) 69–77

Contents lists available at ScienceDirect

Applied Geochemistry
journal homepage: www.elsevier.com/locate/apgeochem

Do organic ligands affect forsterite dissolution rates?


Julien Declercq, Olivier Bosc, Eric H. Oelkers ⇑
GET/Université Paul Sabatier, Observatoire Midi-Pyrénées, UMR 5563 (CNRS/UPS/IRD/CNES), 14 Avenue Edouard Belin, 31400 Toulouse, France

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

Article history: Far-from equilibrium, steady state forsterite dissolution rates were measured at pH 3 and 25 °C in aque-
Received 8 November 2012 ous solutions containing 0.1 m/kg NaCl and up to 0.1 mol/kg of 13 distinct dissolved organic ligands in
Accepted 27 September 2013 mixed-flow reactors. The organic ligands considered in this study include those common in Earth surface
Available online 4 October 2013
environments and those considered as potential catalysts for use in CO2 sequestration efforts: acetate,
Editorial handling by M. Hodson
oxalate, citrate, EDTA4, glutamate, gluconate, malonate, aspartate, tartrate, malate, alginate, salycilate
and humate. The presence of up to 0.1 mol/kg of each organic ligand altered forsterite dissolution rates
less than 0.2 log units, which is the estimated uncertainty of the measured rates. Results obtained in this
study, therefore, suggest that the presence of aqueous organic anions negligibly affects forsterite far-from
equilibrium dissolution rates in most natural environments, and indicate that forsterite carbonation may
not be appreciably accelerated by organic ligand catalysis.
Ó 2013 Elsevier Ltd. All rights reserved.

1. Introduction on an industrial scale (e.g. O’Connor et al., 2000a,b; Wolf et al.,


2004; Giammar et al., 2005; Maroto-Valer et al., 2005; Chen et al.,
A significant number of studies have been aimed at characteriz- 2006; Gerdemann et al., 2007; Oelkers et al., 2008; Broecker,
ing forsterite dissolution rates at various solution compositions 2012). Reaction (1) involves the coupling of two processes: forste-
and temperatures (Luce et al., 1972; Sanemasa et al., 1972; Grand- rite dissolution and magnesite precipitation. The overall rate of car-
staff, 1978, 1986; Murphy and Helgeson, 1987, 1989; Blum and bonation reaction (1) is, therefore, a function of these two processes
Lasaga, 1988; Van Herk et al., 1989; Wogelius and Walther, 1991, (e.g. Saldi et al., 2013). Catalysis which could accelerate the rates of
1992; Casey and Westrich, 1992; Jonckbloedt, 1998; Awad et al., forsterite dissolution thus may also accelerate its carbonation in ac-
2000; Chen and Brantley, 2000; Rosso and Rimstidt, 2000; Pokrov- cord with reaction (1).
sky and Schott, 1999, 2000a,b; Oelkers, 2001; Giammar et al., A vast number of past studies have suggested that organic li-
2005; Golubev et al., 2006; Hänchen et al., 2006; Olsen and Rims- gands can increase the dissolution rates of silicate minerals sub-
tidt, 2008; Rimstidt et al., 2012; Saldi et al., 2013; Wang and Gaim- stantially (e.g. Huang and Kiang, 1972; Manley and Evans, 1986;
mar, 2013). There has been increased recent interest in these rates Mast and Drever, 1987; Bennett et al., 1988; Welch and Ullman,
owing to the potential application of forsterite in carbon capture 1993, 1996; Poulson et al., 1997; Cama and Ganor, 2006; Golubev
and storage efforts (e.g. Giammar et al., 2005; Oelkers and Schott, and Pokrovsky, 2006; Golubev et al., 2006; Ganor et al., 2009;
2005; Marini, 2007; Kelemen and Matter, 2008; Oelkers et al., Pokrovsky et al., 2009; Schott et al., 2009). The effect of the pres-
2008; Dufaud et al., 2009; Prigiobbe et al., 2009; Garcia et al., ence of organic ligands on forsterite dissolution has been consid-
2010; King et al., 2010a,b; Daval et al., 2011; Guyot et al., 2011; ered in detail by Grandstaff (1986), Wogelius and Walther
Olsson et al., 2012). Forsterite is commonly thought of as the best (1991), Hänchen et al. (2006), Olsen and Rimstidt (2008) and
source of the divalent metal cations required for mineral carbon- Prigiobbe and Mazzotti (2011). Grandstaff (1986) reported an in-
ation due to its fast dissolution rates and global abundance. The crease of forsterite dissolution rates when the reactive aqueous
carbonation of forsterite in CO2-rich fluids can occur via fluid contained citrate, EDTA, oxalate, tannic acid, succinate and
phthalate at 25 °C and pH 3.5 and 4.5. Wogelius and Walther
Mg2 SiO4ðsÞ þ2CO2 þ 2H2 O ¼ 2MgCO3ðsÞ þH4 SiO4ðaqÞ ð1Þ
Forsterite
(1991) reported that the addition of either 0.05 mol/kg potassium
Magnesite
phthalate or 103 mol/kg ascorbic acid increases forsterite disso-
where the resulting aqueous silica could eventually precipitate as lution rates at 25 °C by 0.75 log units at pH 4, but this effect de-
amorphous silica or other silicate minerals (cf. Weres et al., 1981; creased with decreasing pH. Hänchen et al. (2006) reported that
Teir et al., 2007; Orlando et al., 2001). Numerous studies have fo- 103 mol/kg citric acid increased forsterite dissolution rates at
cused on the application of reaction (1) to carbon sequestration 90 °C by 0.25 log units at pH 3.4 and 0.5 log units at pH 4.5. Ol-
sen and Rimstidt (2008) reported that 104 mol/kg oxalic acid in-
⇑ Corresponding author. Tel.: +33 561332575. creases forsterite dissolution rates by a factor of 6 at 25 °C and
E-mail address: oelkers@get.obs-mip.fr (E.H. Oelkers). 2.5 < pH < 6.5.

0883-2927/$ - see front matter Ó 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.apgeochem.2013.09.020
70 J. Declercq et al. / Applied Geochemistry 39 (2013) 69–77

There have been a number of distinct interpretations of the ob-


served effect of organic ligands on silicate dissolution rates. Stumm
and collaborators (Furrer and Stumm, 1986; Stumm and Wollast,
1990) proposed that a parallel organic ligand promoted dissolution
pathway catalyzes the formation of organic anion surface species
leading to enhanced rates. Oelkers and Schott (1998) proposed that
organic ligands enhance silicate dissolution rates by forming aque-
ous complexes with those aqueous metal cations that would other-
wise slow rates. Yet others have suggested the formation of surface
chelates by organic ligands as a rate promoting mechanism (e.g.
Wogelius and Walther, 1991).
This study aims to further understanding of organic ligands on
mineral reactivity by measuring the effect of the presence of organ-
ic ligands on forsterite dissolution kinetics at 25 °C and pH 3. The
ligands investigated included acetate, oxalate, citrate, EDTA4, glu-
tamate, gluconate, malonate, aspartate, tartrate, malate, alginate,
salycilate and humate. The purpose of this paper is to report the re-
sults of this experimental study and to use these results to further
illuminate the effect of the presence of organic ligands on forsterite
dissolution rates in natural and carbon storage processes.

2. Materials and methods

2.1. Materials

Xenoliths containing forsteritic (Fo89) olivine were collected


from the San Carlos volcanic field east of Globe, AZ, USA. The xeno-
liths were crushed and forsterite was recovered by handpicking.
This forsterite was then ground and sieved to obtain the 100– Fig. 1. SEM images of the forsterite powder. Image A shows the forsterite before
200 lm size fraction. X-ray diffraction analysis of this material re- dissolution experiments while image B shows the forsterite powder after it had
vealed it to be pure forsterite, free of clays and secondary phases. been dissolved during experimental series 6. Note the initial forsterite is free of fine
particles, and appears to have sharp edges while the reacted mineral powder
Following the protocol established by Olsen and Rimstidt (2008), exhibits rounded edges.
the ground forsterite was first ultrasonically cleaned in 0.01 M
HCl for 10 min, left to settle for 5 min, and washed with alcohol
to remove the fine particles. The resulting forsterite was then dried
and used in the experiments without further treatment. The BET kg NaCl, sufficient HCl or NaOH to adjust the pH to 3, and from
surface area was measured by N2 adsorption using an Absorb-1 0.0001 to 0.1 M of acidic acid, apartic acid, citric acid, EDTA diso-
Quantachrome™ and calculated to be 980 cm2/g (±10%). Micro- dium salt dehydrate, sodium gluconate, L-glutamic acid, malic acid,
scopic analysis of fresh and altered forsterite surfaces was per- malonic acid, oxalic acid, salicylic acid, tartaric acid, or humic acid.
formed using a Jeol JSM840a scanning electron microscope. A All chemicals used in the preparation of inlet fluids were analytical
photomicrograph of the forsterite prior to the experiments is grade and purchased from either Fluka, Sigma, or Sigma–Aldrich.
shown in Fig. 1a. The surfaces are free of both fine particles and The organic acids or salts were used as purchased except for the
secondary phases. humic acid, which was pre-dissolved using techniques reported
Forsterite was dissolved in mixed flow reactors in fluids pre- by Pokrovsky et al. (2005). A list of the acids or salts used in this
pared from 18 MX ultrapure water (MilliQ Plus system), 0.1 mol/ study and their chemical formulas is provided in Table 1.

Table 1
List of the dissociation constants and chemical formula for the organic ligands and the identity of the acid or salt used to make the inlet fluids used in this study. With the
exception of humate, dissociation constants were compiled from Perrin et al. (1981) and Ullman and Welch (2002).

Ligand Formula pKaa Salt or acid used Salt formula


Acetate C2 H3 O
2 4.76 Acidic acid CH3COOH
Alginate (C6H8O6)n 3.5 Alginic acid (C6H8O6)n
Aspartate C2 H3 ðNH2 ÞðCOOÞ2
2
2.0–3.9 Aspartic acid C2H3(NH2)(COOH)2
Citrate C6 H5 O3 3.14–4.75–6.40 Citric acid C6H8O7
7
EDTA4 C10 H12 N2 O4 2.0–2.7–6.2–10.3 EDTA disodium salt dihydrate C10H14N2Na2O8
8
Gluconate C6 H11 O
7 3.86 Sodium gluconate C6H11NaO7
Glutamate C3 H5 ðNH2 ÞðCOOÞ2 2.1–4.1 L-Glutamic acid C3H5(NH2)(COOH)2
2
Malate C4 H4 O2 3.4–5.2 Malic acid C4H6O5
5
Malonate CH2 ðCOOÞ2 2.85–5.70 Malonic acid CH2(COOH)2
2
Oxalate C2 O2 1.2–4.2 Oxalic acid C2H2O4
4
Salicylate C7 H5 O
3 2.97 Salicylic acid C7H6O3
Tartrate C4 H4 O2 3.0–4.3 Tartaric acid C4H6O6
6

Data for humic acid is not available.


a
Where more than one value is given the values refer to the equilibrium constants sequentially for the first, second, third, and fourth deprotonation reaction where
relevant.
J. Declercq et al. / Applied Geochemistry 39 (2013) 69–77 71

Table 2
Results of all experiments performed in the absence of aqueous organic ligands.

Exp cSi (mol/kg)  105 cMg (mol/kg)  105 cMg/cSi pH Flow (mL/min) m (g) log r+,Si (mol/cm2/s) log r+,Mg (mol/cm2/s)
1.s 12.9 26.21 2.03 3.33 0.14 0.49 8.65 8.60
1.f 15.1 31.01 2.05 3.37 0.14 0.49 8.58 8.53
2.s 11.1 20.51 1.85 3.42 0.10 0.31 8.66 8.65
2.f 7.57 15.35 2.03 3.05 0.10 0.31 8.82 8.78
3.s 17.2 31.44 1.83 3.34 0.15 0.31 8.29 8.28
3.f 4.46 8.35 1.87 2.97 0.15 0.31 8.87 8.86
4.s 14.5 27.01 1.86 3.04 0.12 0.30 8.45 8.44
4.f 3.60 7.21 2.00 3.00 0.14 0.30 8.98 8.94
5.s 20.7 33.72 1.63 3.38 0.15 0.30 8.20 8.24
5.f 4.75 8.61 1.81 3.01 0.16 0.30 8.80 8.80
6.s 3.28 6.75 2.06 3.12 0.40 0.31 8.58 8.53
6.f 1.87 4.29 2.29 3.11 0.40 0.31 8.83 8.73
7.s 14.7 26.02 1.77 3.06 0.22 0.50 8.41 8.40
7.f 4.82 9.42 1.95 3.08 0.22 0.50 8.89 8.93
8.s 14.8 26.09 1.76 3.49 0.14 0.31 8.59 8.60
8.f 4.68 8.06 1.72 3.04 0.10 0.31 9.03 9.05
9.s 9.27 10.04 1.08 3.24 0.15 0.33 8.56 8.78
9.f 1.90 3.56 1.87 3.09 0.14 0.33 9.32 9.31
10.s 4.29 8.67 2.02 3.04 0.15 0.31 8.88 8.83
10.f 7.72 13.39 1.73 3.00 0.15 0.31 8.62 8.64
11.s 4.66 9.09 1.95 3.03 0.20 0.30 8.72 8.69
11.f 5.02 8.86 1.76 2.90 0.20 0.30 8.69 8.70
12.s 1.93 4.14 2.15 2.90 0.30 0.35 8.93 8.86
12.f 1.61 2.86 1.78 2.90 0.30 0.35 9.01 8.96
13.s 1.64 3.21 1.95 2.90 0.17 0.35 9.25 9.21
13.f 2.25 4.10 1.82 2.97 0.17 0.35 9.11 9.11

X.s and X.f designates the first and last experiments of the series.

2.2. Experimental methods series dissolution rates measured in the presence of organic ligands
can be compared directly to rates measured in organic ligand-free
All experiments were performed in 30 mL polypropylene fluids on the identical forsterite samples. After the steady-state
mixed-flow reactors (cf. Chaïrat et al., 2007; Flaathen et al., was obtained for the final organic ligand-free fluid, the reactor
2010); schematic illustrations of the reactor design have been pre- was dismantled, the remaining forsterite powder was recovered
sented by Kohler et al. (2005). The reactors were continuously stir- and the reactor cleaned. In the case of EDTA and L-glutamate, or-
red with floating Teflon stirring bars. These reactors were ganic ligand concentrations of 0.1 M could not be achieved due
immersed in a water bath held at a 25 ± 2 °C. The fluid was injected to solubility constraints. Approximately 10 mL of outlet fluid was
using a Gilson peristaltic pump, which allows fluid flow rates from collected twice daily during each experiment. Steady-state was
0.01 to 10 g/min. The fluid left the reactor through a 0.45 lm ace- confirmed by the outlet fluids maintaining a constant Mg and Si
tate filter. No additional filtering was performed on outlet fluid concentration within analytical uncertainty for a minimum of 3
samples prior to chemical analysis. Inlet and outlet fluid pH was residence times. The residence time is defined as the volume of
measured using a combination glass electrode calibrated with NIST the reactor divided by the reactive fluid flow rate, which was fixed
buffers (pH = 4.002, 6.865, and 9.180 at 25 °C). The precision of pH at 0.15–0.40 g/min in the experiments. Further details of each
measurements was ±0.02 units. Magnesium concentrations were experimental series are provided in Tables 2 and 3.
determined by flame atomic absorption spectrophotometry using
a Perkin Elmer 5100 PC spectrometer equipped with an AS-90
3. Results
autosampler, with an uncertainty of ±2%. Silica concentrations
were determined using the molybdate blue colorimetric method
An example of the reactive fluid Mg and Si concentrations mea-
(Govett, 1961) using a Technicon analyzer with an uncertainty of
sured during one experimental series is depicted in Fig. 2. This fig-
±2% for most fluids, though some interference was evident in some
ure shows the reactive fluid evolution during experiments
organic ligand-rich fluids, see below.
designed to assess the effect of aqueous malonate on forsterite dis-
Seventy-seven mixed-flow reactor experiments were run at a
solution rates. It can be seen in this figure that the addition of the
pH of 3 ± 0.4 throughout the study. To allow direct comparison be-
organic ligand leads to the attainment of a new steady state condi-
tween rates in the presence and absence of organic ligands, exper-
tion as indicated by the dashed lines in the figure.
iments were organized into thirteen series. Each series is
Forsterite steady-state dissolution rates were calculated from
designated by the prefix of the experiment. For example, series 2
the aqueous Mg and Si concentrations using the relationship:
consists of experiments 2-b, 2-1, 2-2, 2-3, 2-4 and 2-f. Each exper-
imental series was initiated by passing an organic-free, pH 3 fluid qðc i;out  c i;in Þ
through the reactor at a constant flow rate for approximately rþ;i ¼ ð2Þ
S  m  mi
6 days. Once an initial steady state was obtained and confirmed,
the original inlet fluid was replaced by a pH 3 fluid containing where q refers to the fluid flow rate, ci,in and ci,out represent the
104 mol/kg of a selected organic ligand until a second steady state aqueous concentration of the ith element in the inlet and outlet, S
was obtained and confirmed. This sequence was generally repeated denotes the specific surface area, m designates the amount of for-
using pH 3 fluids containing 103, 102, 101 and 0 mol/kg of this sterite powder introduced in the reactor at the beginning of the
organic ligand. The total run time for each complete experimental experiment series and mi stands for the stoichiometric number of
series ranged from 3 to 6 weeks. By running the experiments in moles of the ith element in one mole of dissolving forsterite.
72 J. Declercq et al. / Applied Geochemistry 39 (2013) 69–77

Table 3
Results of all experiments performed in the presence of aqueous organic ligands.

EXP Ligand Ligand conc. cSi (mol/kg)  105 cMg (mol/kg)  105 cMg/cSi pH Flow m (g) log r+,Si log r+,Mg
(mol/kg) (mL/min) (mol/cm2/s) (mol/cm2/s)
1-1 Malonate 0.0001 15.8 28.9 1.83 3.38 0.14 0.49 8.55 8.54
1-2 0.0001 6.50 12.2 1.87 3.39 0.16 0.30 8.67 8.66
1-3 0.001 5.18 9.13 1.76 3.14 0.16 0.30 8.77 8.78
1-4 0.01 5.39 9.13 1.67 3.01 0.16 0.30 8.76 8.78
1-5 0.1 6.14 12.2 1.98 3.01 0.16 0.30 8.70 8.66
2-1 Acetate 0.0001 14.9 26.9 1.81 3.50 0.10 0.31 8.51 8.51
2-2 0.001 15.3 28.5 1.86 3.22 0.10 0.31 8.50 8.49
2-3 0.01 13.5 25.1 1.85 3.17 0.10 0.31 8.55 8.54
2-4 0.1 8.36 24.5 2.93 3.05 0.10 0.31 8.76 8.55
3-1 Oxalate 0.0001 11.2 21.8 1.94 3.34 0.30 0.31 8.17 8.14
3-2 0.001 8.37 16.4 1.96 3.11 0.30 0.31 8.30 8.26
3-4 0.01 8.45 16.7 1.98 2.95 0.30 0.31 8.29 8.26
3-5 0.1 1.42 5.80 4.09 3.03 0.30 0.31 9.07 8.71
4-1 Citrate 0.0001 6.86 12.1 1.77 3.17 0.30 0.30 8.38 8.39
4-2 0.001 5.75 10.9 1.90 3.13 0.30 0.30 8.46 8.44
4-3 0.01 6.76 13.2 1.96 3.04 0.30 0.30 8.39 8.35
4-4 0.1 4.12 n.d. n.d. 3.07 0.30 0.30 8.60 n.d.
5-1 EDTA 0.0001 11.0 22.9 2.09 2.95 0.30 0.30 8.17 8.11
5-2 0.001 5.71 11.2 1.96 3.10 0.30 0.30 8.45 8.42
5-3 0.01 3.82 9.01 2.36 3.20 0.30 0.30 8.63 8.51
6-1 Tartrate 0.001 3.26 6.72 2.06 3.07 0.39 0.31 8.60 8.54
6-3 0.01 3.00 5.94 1.98 3.20 0.40 0.31 8.62 8.58
7-1 Malate 0.0001 4.88 9.23 1.89 3.12 0.22 0.30 8.67 8.65
7-2 0.0001 5.11 9.79 1.92 3.13 0.30 0.30 8.51 8.65
7-3 0.001 4.63 8.52 1.84 2.99 0.30 0.30 8.55 8.55
7-4 0.01 4.93 8.85 1.79 2.99 0.30 0.30 8.53 8.53
7-5 0.1 12.4 24.1 1.94 3.08 0.14 0.50 8.67 8.65
8-1 Salicylate 0.0001 15.0 27.4 1.83 3.01 0.14 0.50 8.59 8.58
8-2 0.001 4.11 7.45 1.81 3.03 0.10 0.31 9.08 9.08
8-3 0.01 4.93 8.43 1.71 3.01 0.10 0.31 9.01 9.03
8-4 0.1 5.18 44.3 8.72 3.11 0.10 0.31 8.98 8.30
9-1 Alginate 0.0001 4.69 9.53 2.03 3.15 0.15 0.30 8.85 8.80
9-2 0.001 1.27 2.61 2.05 3.29 0.31 0.31 9.11 9.05
9-3 0.01 2.08 3.84 1.85 3.30 0.31 0.31 8.90 8.89
9-4 0.1 1.64 3.41 2.08 3.04 0.31 0.31 9.00 8.94
10-1 Humatea 0.0002 3.43 6.28 1.83 3.07 0.33 0.14 8.30 8.29
10-2 0.0006 2.92 5.43 1.86 3.04 0.30 0.31 8.76 8.75
10-3 0.0006 3.24 5.08 1.57 3.05 0.15 0.33 9.05 9.11
10-4 0.001 3.70 6.83 1.60 3.04 0.30 0.31 8.65 8.65
10-5 0.001 14.5 n.d. n.d. 3.22 0.10 0.33 8.57 n.d.
11-1 Gluconate 0.0001 4.36 9.83 2.26 3.15 0.15 0.30 9.70 8.79
11-2 0.001 1.91 4.52 2.36 3.51 0.20 0.30 9.11 8.99
11-3 0.01 2.00 4.10 2.05 3.43 0.20 0.30 9.09 9.04
11-4 0.1 3.46 6.79 1.96 3.15 0.20 0.30 8.85 8.82
13-1 Glutamate 0.0001 1.65 3.29 1.99 2.77 0.21 0.35 9.21 9.17
13-2 0.001 1.86 3.81 2.05 2.88 0.21 0.35 9.16 9.11
13-3 0.01 1.93 3.76 1.95 2.81 0.21 0.35 9.15 9.11
12-1 Aspartate 0.0001 2.20 5.07 2.30 2.92 0.20 0.35 9.11 9.01
12-2 0.001 2.08 5.07 2.43 2.95 0.20 0.35 9.14 9.01
12-3 0.01 2.00 5.07 2.53 2.92 0.20 0.35 9.15 9.01
12-4 0.1 1.80 5.07 2.82 2.92 0.20 0.35 9.20 9.01

n.d. Not determined.


a
The ligand concentration for the humates is given in units of volume fraction.

Steady-state Si and Mg outlet concentrations and the dissolu- method; these observations will be expanded upon in detail in
tion rates calculated from all experiments are provided in Tables a future communication. As such, rates based on Mg release are
2 and 3. With few exceptions the molar Mg to Si ratio of all stea- used for the interpretation of these experiments in the discussion
dy-state outlet flow is 1.78 ± 0.25, consistent with stoichiometric below.
dissolution of the forsterite. The logarithm of the dissolution rates All experimentally determined forsterite dissolution rates are
calculated based on Si release is plotted as a function of those plotted as a function of organic ligand concentration in Figs. 4
based on Mg release in Fig. 3. Only 3 experiments are significantly and 5. Plots in these figures are grouped by experimental series
non-stoichiometric: 11-1, 3-4 and 8-4, containing respectively allowing the direct assessment of the effect on forsterite dissolu-
104 mol/kg of gluconate, and 0.1 mol/kg of oxalate and salicylate. tion rates of each organic ligand. Measured rates are plotted in
Tests of the molybdate blue method used to analyze dissolved Si each figure in the sequence they were measured; first rates were
in this study show that the presence of 0.1 mol/kg gluconate and measured in an organic ligand-free fluid, then rates were measured
oxalate significantly affected Si concentrations measured by this in a sequence of fluids containing increasing organic ligand
J. Declercq et al. / Applied Geochemistry 39 (2013) 69–77 73

Fig. 2. The temporal evolution of reactive fluid aqueous Si and Mg concentrations during experimental series 1. The solid symbols correspond to measured concentrations,
the vertical solid lines indicate times when the inlet fluid composition was changed, and the dashed lines indicate the steady state concentration values for each experiment.

release, is as much as 0.6 log units different from that measured


in the first organic ligand-free fluid. This difference is not system-
atic; in some experiments the final steady state forsterite dissolu-
tion rate is higher and in some the final rate is lower than the
initial rate. The origin of these variations remains unclear, but a
number of previous studies have suggested that significant tempo-
ral variations in rates during experiments are due to changes in
reactive surface area (Lüttge et al., 1999; Gautier et al., 2001; Hod-
son, 2006; Lüttge, 2006); mineral dissolution rates are commonly
assumed to be proportional to the mineral–fluid reactive surface
area, yet the variation of these surface areas as minerals dissolve
is still poorly defined (e.g. Noiriel et al., 2009). A second relevant
observation is that rates measured in the presence of organic
ligands are equal to those measured in organic ligand-free fluids
within analytical uncertainty.

4. Discussion

4.1. Comparison with past studies

Fig. 3. Measured far-from-equilibrium steady-state forsterite dissolution rates at Wogelius and Walther (1991) reported that the degree to
25 °C and pH 3 based on Si release as a function of corresponding rates based on which ascorbic acid and K-phthalate affect forsterite dissolution
Mg release. Rates measured in the presence of aqueous organic ligands are depicted rates depends on pH. They observed no effect of these organic
as solid squares were whereas those measured in organic ligand-free fluids are
compounds at pH 2, but reported that these ligands increased
depicted as solid diamonds. Unfilled symbols correspond to experiments showing
an apparent non-stochiometric metal release. Note that 74 of the 77 plotted rates forsterite dissolution rates by up to 0.75 log units at pH 6.
are within 0.25 log unit of the 1:1 stoichiometric line. Grandstaff (1986) noted a similar pH dependence of the effect
of organic ligands on forsterite dissolution rates. Hänchen et al.
(2006) reported that the effect of citric acid on forsterite disso-
lution rate was small under acid conditions but increases with
concentrations and, finally, a second rate measurement was per- increasing pH at 90 °C. In contrast, Olsen and Rimstidt (2008)
formed in an organic ligand-free fluid. First it is evident that the fi- suggested that the presence of oxalate on forsterite dissolution
nal rate measured in organic ligand-free fluids, based on Mg rates at 25 °C was pH independent from 2.5 < pH < 6.5. One
74 J. Declercq et al. / Applied Geochemistry 39 (2013) 69–77

Fig. 4. Forsterite far from equilibrium dissolution rate conditions at 25 °C and pH 3 measured in the present study as a function of the concentration of the indicated organic
ligand concentration. Rates obtained in the presence of aqueous organic ligands are shown as solid diamonds. Rates obtained at the beginning and end of the corresponding
experimental series in organic ligand-free fluids are plotted as filled triangles at a log organic ligand concentration of 5 and 0, respectively, to illuminate the effect of the
ligands on rates.

possible reason why this latter study is inconsistent with the of forsterite dissolution experiments can be inconsistent with the
previous studies is that their conclusions were based on forste- longer-term steady state rates.
rite dissolution experiments that lasted only 2 h. As can be seen Rates in the present study demonstrate that at pH 3 forsterite
in Fig. 2, the Mg and Si release rates during the first few hours dissolution rates are negligibly impacted by the presence of a wide
J. Declercq et al. / Applied Geochemistry 39 (2013) 69–77 75

Hood, 1965; Suess, 1970; Otsuki and Wetzel, 1972; Reddy, 1977;
Berner et al., 1978; Reynolds, 1978; Reddy and Wang, 1980; Inns-
keep and Bloom, 1986; Giannimaras and Koutsoukos, 1988; Dove
and Hochella, 1993; Katz et al., 1993; Paquette et al., 1996).

4.3. Implications for silicate dissolution models

There are at least two distinct models for the enhancement of


silicate mineral dissolution rates in the presence of organic ligands.
A number of studies have argued that organic ligands could adsorb
to the mineral surface, providing a new parallel reaction mecha-
nism for the detachment of material from the mineral surface
(e.g. Furrer and Stumm, 1986; Amrhein and Suarez, 1988). In con-
trast, Oelkers and Schott (1998) proposed that the observed accel-
eration of the dissolution rates of some silicate minerals with
organic ligand concentration stems from their ability to make
aqueous complexes with aqueous metals that would otherwise in-
Fig. 5. Forsterite far from equilibrium dissolution rates at 25 °C and pH 3 hibit rates. Although the degree to which organic ligands adsorb
measured in the present study as a function of aqueous humate concentration. onto forsterite surfaces has not been measured directly in this
Rates obtained in the presence of aqueous humate are shown as solid diamonds. study, it seems likely that significant adsorption would occur be-
Rates obtained at the beginning and end of the experimental series in organic
cause at pH 3 forsterite surfaces are positively charged (Pokrovsky
ligand-free fluids are plotted as filled triangles at an aqueous humate concentration
of 0 and 0.0011, respectively, to illuminate the effect of the aqueous humate on
and Schott, 2000a; Oelkers et al., 2009) whereas most of the organ-
rates. ic species considered in this study are at least partly present as
aqueous anions at this pH due to their low dissociation constants
(see Table 3). The lack of forsterite dissolution rate acceleration
in the presence of most organic ligands seems to be inconsistent
variety of organic ligands at 60.1 mol/kg. These observations are in with the hypothesis that rates are accelerated by ligand adsorption.
general agreement with the previous reports that aqueous organic Although aqueous Mg-organic ligand complexation does occur,
ligands have at most a small effect on forsterite dissolution rates in forsterite dissolution rates have been shown to be independent
strongly acidic conditions but may have an effect at higher pH. A of aqueous Mg2+ activity (Oelkers, 2001). As such this aqueous
reason for the contrasting effects of aqueous organic ligands on Mg-organic ligand complexing would not lead to accelerated rates
steady-state forsterite dissolution rates with increasing pH may according to the Oelkers and Schott (1998) hypothesis, consistent
stem from their aqueous speciation. As can be seen in Table 1, with the observations reported above.
many of the organic ligands considered in this study have a first
dissociation constant near pH 3. As such these organic species tend
to be present as neutral aqueous species at acidic pH, but as nega- 5. Conclusions
tively charged aqueous species in mildly acidic and neutral condi-
tions. Another possible reason for the contrasting effect of organic The results of this study suggest that the presence of organic li-
ligands on forsterite dissolution rates with increasing pH is the gands at concentrations of <0.1 mol/kg have at most a small effect
composition of the forsterite surface. Both the charge and Mg/Si ra- on forsterite dissolution rates at pH 3 and 25 °C. As this pH is char-
tio of forsterite surfaces are strong functions of pH, factors which acteristic of the CO2 charged waters that might be found in the
could affect its reactivity (Pokrovsky and Schott, 2000a,b; Oelkers subsurface during carbon storage efforts it seems unlikely that
et al., 2009). the presence of organic ligands would enhance the availability of
those divalent metals required for mineral carbonation from ultra-
4.2. Implications for carbon capture and storage mafic rocks prior to the neutralization of this fluid. In contrast, the
presence of organic ligands have been shown to enhance the disso-
One of the major goals of this study is to assess the degree to lution rates of the intermediate feldspars and basaltic glass at
which the addition of organic ligands might accelerate Mg release acidic conditions (e.g. Blum and Stillings, 1995; Thorseth et al.,
from forsterite thus aiding mineral carbonation efforts. The pH of 1995; Ullman et al., 1996; Oelkers and Schott, 1998; Alt and Mata,
CO2-charged water depends on the partial pressure of CO2 in the 2000; Oelkers and Gíslason, 2001); as such the use of organic li-
system; the pH of CO2 saturated water decreases from 3.9 to 3.0 gands might prove effective in accelerating mineral carbonation
with increasing CO2 partial pressure from 1 to 64 bars (Gíslason in basaltic and/or rocks of intermediate composition. Nevertheless,
et al., 2010). The results summarized above suggest that the addi- organic ligands may hinder mineral carbonation due to their effect
tion of most organic ligands will have little effect on Mg release on the solubility and/or precipitation kinetics of carbonate miner-
from forsterite in these CO2-charged waters. This conclusion agrees als. As such the potential of aqueous organic ligands to enhance
with those of Wolff-Boenisch et al. (2006, 2011) who concluded carbon mineralization remains questionable.
that the addition of various organic ligands negligibly affected
the rates of peridotite dissolution at 25 °C and pH 3.6. As forsterite Acknowledgements
dissolves in CO2 charged waters, however, the fluid will neutralize
(e.g. Siever and Woodfort, 1979). Under such conditions, previous We would like to thank Alain Castillo for technical assistance
studies suggest that some acceleration of Mg release could be at- throughout the duration of the experimental work, Carole Causser-
tained from the presence of a number of organic ligands. The pres- and for her generous help during the analytical part of the work,
ence of these organic ligands, however might inhibit the ultimate and Philippe de Parseval for aid creating SEM images. We thank
precipitation of carbonates either through increasing the solubility Per Aagaard, Stacey Callahan, Oleg Pokrovsky, Jacques Schott, Pas-
of these minerals (Innskeep and Bloom, 1986; Lebrón and Suárez, cale Bénézeth and Morgan T. Jones for helpful discussions during
1998) or by inhibiting their precipitation rates (e.g. Kitano and the course of this study. Support from Centre National de la
76 J. Declercq et al. / Applied Geochemistry 39 (2013) 69–77

Recherche Scientifique, and the European Community through Grandstaff, D.E., 1986. The dissolution rate of forsteritic olivine from Hawaiian
beach sand. In: Colman, S.M., Dethier, D.P. (Eds.), Rates of Chemical Weathering
CARB-FIX (Collaborative Project-FP7-ENERGY-2011-1-283148) is
of Rocks and Minerals. Academic Press, pp. 41–57.
gratefully acknowledged. Guyot, F., Daval, D., Dupraz, S., Martinez, I., Menez, B., Sissmann, O., 2011. CO2
geological storage: the environmental mineralogy perspective. C. R. Geosci. 343,
246–259.
Hänchen, M., Prigiobbe, V., Storti, G., Seward, T.M., Mazotti, M., 2006. Dissolution
References kinetics of forsteritic olivine at 90–150 °C including effects of the presence of
CO2. Geochim. Cosmochim. Acta 70, 4403–4416.
Alt, J.C., Mata, P., 2000. On the role of microbes in the alteration of submarine Hodson, M.E., 2006. Does reactive surface area depend on grain size? Results from
basaltic glass: a TEM study. Earth Planet. Sci. Lett. 181, 301–313. pH 3, 25 °C far-from-equilibrium flow-through dissolution experiments on
Amrhein, C., Suarez, D.L., 1988. The use of a surface complexation model to describe anorthite and biotite. Geochim. Cosmochim. Acta 70, 1655–1667.
the kinetics of ligand-promoted dissolution of anorthite. Geochim. Cosmichim. Huang, W.H., Kiang, W.C., 1972. Laboratory dissolution of plagioclase feldspar
Acta 52, 2785–2793. in water and organic acids at room temperature. Am. Mineral. 57, 1848–
Awad, A., van Groos, A.F.K., Guggenheim, S., 2000. Forsteritic olivine: effect of 1859.
crystallographic direction on dissolution kinetics. Geochim. Cosmochim. Acta Innskeep, W.P., Bloom, P.R., 1986. Kinetics of calcite precipitation in the presence of
64, 1765–1772. water-soluble organic ligands. Soil Sci. Soc. Am. J. 50, 1167–1172.
Bennett, P.C., Melcer, M.E., Siegel, D.I., Hasset, J.P., 1988. The dissolution of quartz in Jonckbloedt, R.C.L., 1998. Olivine dissolution in sulfuric acid at elevated
dilute aqueous solutions of organic acids at 25 °C. Geochim. Cosmochim. Acta temperatures, implications for the olivine process, and alternative waste acid
52, 1521–1530. neutralizing process. J. Geochem. Explor. 62, 337–346.
Berner, R.A., Westrich, J.T., Smith, J., Martens, C.S., 1978. Inhibition of aragonite Katz, J.L., Reick, M.R., Herzog, E.R., Parsiegla, K.L., 1993. Calcite growth inhibition by
precipitation from supersaturated seawater: a laboratory and field study. Am. J. iron. Langmuir 9, 1423–1430.
Sci. 278, 816–837. Kelemen, P.B., Matter, J., 2008. In situ carbonation of peridotite for CO2 storage. Proc.
Blum, A.E., Lasaga, A.C., 1988. Role of surface speciation in the low-temperature Natl. Acad. Sci. USA 105, 17295–17300.
dissolution of minerals. Nature 331, 431–433. King, H.E., Plumper, O., Putnis, A., 2010a. Effect of secondary phase formation of the
Blum, A.E., Stillings, L.L., 1995. Chemical weathering of feldspars. Rev. Mineral 31, carbonation of olivine. Environ. Sci. Technol. 44, 6503–6509.
173–233. King, H.E., Stimpfl, M., Deymier, P., Drake, M.J., Catlow, C.R.A., Putnis, A., de Leeuw,
Broecker, W.S., 2012. The carbon cycle and climate change: memoirs of my 60 years N.H., 2010b. Computer simulation of water interaction with low-coordinated
in science. Geochem. Perspect. 1, 221–351. forsterite surface sites: implications for the origin of water in the inner solar
Cama, J., Ganor, J., 2006. The effect of organic acids on the dissolution of silicate system. Earth Planet. Sci. Lett. 300, 11–18.
minerals: a case study of oxalate catalysis of kaolinite dissolution. Geochim. Kitano, Y., Hood, D.W., 1965. The influence of organic material on the polymorphic
Cosmochim. Acta 70, 2191–2209. crystallization of calcium carbonate. Geochim. Cosmochim. Acta 29, 29–41.
Casey, W.H., Westrich, H., 1992. Control of dissolution rates of orthosilicate Kohler, S.J., Harouiya, N., Chairat, C., Oelkers, E.H., 2005. Experimental studies of REE
minerals by divalent metal–oxygen bonds. Nature 355, 157–159. fractionation during water–mineral interactions: REE release rates during
Chaïrat, C., Schott, J., Oelkers, E.H., Lartigue, J.-E., Harouiya, N., 2007. Kinetics and apatite dissolution from pH 2.8 to 9.2. Chem. Geol. 222, 168–182.
mechanism of natural fluorapatite dissolution at 25 °C and pH from 3 to 12. Lebrón, I., Suárez, D.L., 1998. Kinetics and mechanisms of precipitation of calcite as
Geochim. Cosmochim. Acta 71, 5901–5912. affected by pCO2 and organic ligands at 25 °C. Geochim. Cosmochim. Acta 62,
Chen, Y., Brantley, S.L., 2000. Forsterite dissolution at 65 °C and 2 < pH < 5. Chem. 405–416.
Geol. 165, 267–281. Luce, R.W., Bartlett, R.W., Parks, G.A., 1972. Dissolution kinetics of magnesium
Chen, Y., Lundqvist, P., Johansson, A., Platell, P., 2006. A comparative study of the silicates. Geochim. Cosmochim. Acta 36, 35–50.
carbon dioxide transcritical power cycle compared with an organic rankine Lüttge, A., 2006. Crystal dissolution kinetics and Gibbs free energy. J. Electron
cycle with R123 as working fluid in waste heat recovery. Appl. Therm. Eng. 26, Spectrom. Relat. Phenom. 150, 248–259.
2142–2147. Lüttge, A., Bolton, E.W., Lasaga, A.C., 1999. An interferometry study of the
Daval, D., Sissmann, O., Menguy, N., Saldi, G.D., Guyot, F., Martinez, I., Corvisier, J., dissolution kinetics of anorthite: the role of reactive surface area. Am. J. Sci.
Garcia, B., Machouk, I., Knauss, K.G., Hellmann, R., 2011. Influence of amorphous 299, 652–678.
silica layer formation on the dissolution rate of olivine at 90 °C and elevated Manley, E.P., Evans, L.J., 1986. Dissolution of feldspar by low-molecular-weight
pCO2. Chem. Geol. 284, 193–209. aliphatic and aromatic acids. Soil Sci. 141, 106–112.
Dove, P.M., Hochella Jr., M.F., 1993. Calcite precipitation mechanisms and inhibition Marini, L., 2007. Geological sequestration of carbon dioxide: thermodynamics,
by orthophosphate. In situ observations by scanning force microscopy. kinetics and reaction path modeling. Developments in Geochemistry, vol. 11.
Geochim. Cosmochim. Acta 57, 705–714. Elsevier, New York.
Dufaud, F., Martinez, I., Shilobreeva, S., 2009. Experimental study of Mg-rich Maroto-Valer, M.M., Fauth, D.J., Kuchta, M.E., Zhang, Y., Andresen, J.M., 2005.
silicates carbonation at 400 and 500 °C and 1 kbar. Chem. Geol. 265, 79–87. Activation of magnesium rich minerals as carbonation feedstock materials for
Flaathen, T.K., Gislason, S.R., Oelkers, E.H., 2010. The effect of aqueous sulphate on CO2 sequestration. Fuel Process. Technol. 86, 1627–1645.
basaltic glass dissolution rates. Chem. Geol. 277, 345–354. Mast, A.M., Drever, J.I., 1987. The effect of oxalate on the dissolution rate of
Furrer, G., Stumm, W., 1986. The coordination chemistry of weathering: I. oligoclase and tremolite. Geochim. Cosmochim. Acta 51, 2559–2568.
Dissolution of d-Al2O3 and BeO. Geochim. Cosmochim. Acta 50, 1847–1860. Murphy, W.M., Helgeson, H.C., 1987. Thermodynamic and kinetic constraints on
Ganor, J., Reznik, I.J., Rosenberg, Y.O., 2009. Organics in water–rock interactions. reaction rates among minerals and aqueous solutions. III. Activated complexes
Rev. Mineral. Geochem. 70, 259–369. and the pH-dependence of the rates of feldspar, pyroxene, wollastonite, and
Garcia, B., Baumont, V., Perfetti, E., Rouchon, V., Blanchet, D., Oger, P., Dromart, G., olivine hydrolysis. Geochim. Cosmochim. Acta 51, 3137–3153.
Huc, A.-Y., Haeseler, F., 2010. Experiments and geochemical modeling of CO2 Murphy, W.M., Helgeson, H.C., 1989. Thermodynamic and kinetic constraints on
sequestration by olivine: potential, quantification. Appl. Geochem. 25, 1383– reaction rates among minerals and aqueous solutions. IV. Retrevial of rate
1396. constants and activation parameters for the hydrolysis of pyroxene,
Gautier, J.-M., Oelkers, E.H., Schott, J., 2001. Are quartz dissolution wollastonite, olivine, analusite and quartz. Am. J. Sci. 288, 17–101.
rates proportional to BET surface areas? Geochim. Cosmochim. Acta 65, Noiriel, C., Luquot, L., Madé, B., Raimbault, L., Gouze, P., van der Lee, J., 2009.
1059–1070. Changes in reactive surface area during limestone dissolution: an experimental
Gerdemann, S.J., O’Connor, W.K., Dahlin, D.C., Penner, L.R., Rush, H., 2007. Ex situ and modeling study. Chem. Geol. 265, 160–170.
aqueous mineral carbonation. Environ. Sci. Technol. 41, 2587–2593. O’Connor, W.K., Dahlin, D.C., Nilsen, D.N., Walters, R.P., Turner, P.C., 2000a. Carbon
Giammar, D.E., Bruant, R.G., Peters, C.A., 2005. Forsterite dissolution and magnesite dioxide sequestration by direct carbonation with carbonic acid. In: Proc. 25th
precipitation at conditions relevant for deep saline aquifer storage and Int. Tech. Conf. Coal Utilization & Fuel Systems, Coal Technology Assoc.,
sequestration of carbon dioxide. Chem. Geol. 217, 257–276. Clearwater, FL.
Giannimaras, E.K., Koutsoukos, P.G., 1988. Precipitation of calcium carbonate in O’Connor, W.K., Dahlin, D.C., Turner, P.C., Walters, R., 2000b. Carbon dioxide
aqueous solutions in the presence of oxalate anions. Langmuir 4, 855–861. sequestration by ex-situ mineral carbonation. In: Proc. 2nd Ann. Dixy Lee Ray
Gíslason, S.R., Wolff-Boenisch, D., Stefansson, A., Oelkers, E.H., Gunnlaugsson, E., Memorial Symp., Am. Soc. Mech. Eng., Washington, DC.
Sigurdardottir, H., Sigfusson, B., Broecker, W.S., Matter, J.M., Stute, M., Axelsson, Oelkers, E.H., 2001. An experimental study of forsterite dissolution rate as a
G., Fridriksson, T., 2010. Mineral sequestration of carbon dioxide in basalt: a function of temperature and aqueous Mg and Si concentration. Chem. Geol. 148,
pre-injection overview of the CarbFix project. Int. J. Greenhouse Gas Control 4, 485–494.
537–545. Oelkers, E.H., Gíslason, S.R., 2001. The mechanism, rates and consequences of
Golubev, S.V., Pokrovsky, O.S., 2006. Effect of pH and organic ligands on diopside basaltic glass dissolution: I. An experimental study of the dissolution rates of
dissolution kinetics. Chem. Geol. 235, 377–389. basaltic glass as a function of aqueous Al, Si and oxalic acid concentration at
Golubev, S.V., Bauer, A., Pokrovsky, O.S., 2006. Effect of pH and organic ligands on 25 °C and pH = 3 and 11. Geochim. Cosmochim. Acta 65, 3671–3681.
the kinetics of smectite dissolution at 25 °C. Geochim. Cosmochim. Acta 70, Oelkers, E.H., Schott, J., 1998. Does organic acid adsorption affect alkali-feldspar
4436–4451. dissolution rates? Chem. Geol. 155, 235–245.
Govett, G.J.S., 1961. Critical factors in the colorimetric determination of silica. Oelkers, E.H., Schott, J., 2005. Geochemical aspects of CO2 sequestration. Chem.
Geochim. Cosmochim. Acta 25, 69–80. Geol. 217, 183–186.
Grandstaff, D.E., 1978. Changes in surface area and morphology and the mechanism Oelkers, E.H., Gislason, E.H., Matter, J., 2008. Mineral carbonation of CO2. Elements
of forsterite dissolution. Geochim. Cosmochim. Acta 42, 1899–1901. 4, 333–337.
J. Declercq et al. / Applied Geochemistry 39 (2013) 69–77 77

Oelkers, E.H., Golubev, S.V., Chairat, C., Pokrovsky, O.S., Schott, J., 2009. The surface limiting reactions at 90 and 150 °C in open and closed systems. Geochim.
chemistry of multi-oxide silicates. Geochim. Cosmochim. Acta 73, 4617–4634. Cosmochim. Acta 118, 157–183.
Olsen, A.A., Rimstidt, D., 2008. Oxalate promoted forsterite dissolution at low pH. Sanemasa, I., Yoshida, L., Ozawa, T., 1972. The dissolution of olivine in aqueous
Geochim. Cosmochim. Acta 72, 1758–1766. solutions of inorganic acids. Bull. Chem. Soc. Jpn. 45, 1741–1746.
Olsson, J., Bovet, N., Makovicky, E., Bechgaard, K., Balogh, Z., Stipp, S.L.S., 2012. Schott, J., Pokrovsky, O.S., Oelkers, E.H., 2009. The link between mineral dissolution/
Olivine reactivity with CO2 and H2O on a microscale: implications for carbon precipitation kinetics and solution chemistry. Rev. Mineral. Geochem. 70, 207–
sequestration. Geochim. Cosmochim. Acta 77, 86–97. 258.
Orlando, A., Borrini, D., Marini, L., 2001. Dissolution and carbonation of a Siever, R., Woodfort, N., 1979. Dissolution kinetics and the weathering of mafic
serpentine: inferences from acid attack and high P–T experiments performed minerals. Geochim. Cosmochim. Acta 43, 717–724.
in aqueous solutions at variable salinity. Appl. Geochem. 26, 1569–1583. Stumm, W., Wollast, R., 1990. Coordination of chemistry of weathering: kinetics of
Otsuki, A., Wetzel, R.G., 1972. Coprecipitation of phosphate with carbonates in a the surface-controlled dissolution of oxide minerals. Rev. Geophys. 28, 53–69.
marl lake. Limnol. Oceanogr. 17, 735–737. Suess, E., 1970. Interaction of organic compounds with calcium carbonate – I.
Paquette, J., Vali, H., Mucci, A., 1996. TEM study of Pt–C replicas of calcite Association phenomena and geochemical implications. Geochim. Cosmochim.
overgrowths precipitated from electrolyte solutions. Geochim. Cosmochim. Acta 34, 157–168.
Acta 60, 4689–4699. Teir, S., Revitzer, H., Eloneva, S., Fogelholm, C., Zevenhoven, R., 2007. Dissolution of
Perrin, D.D., Dempsey, B., Serjeant, E.P., 1981. pKa Predictions for Organic Acids and natural serpentine in mineral and organic acids. Int. J. Miner. Process. 83, 36–46.
Bases. Chapman & Hall, London. Thorseth, I.H., Furnes, H., Tumyr, O., 1995. Textural and chemical effects of bacterial
Pokrovsky, O.S., Schott, J., 1999. Olivine surface speciation and reactivity in aquatic activity on basaltic glass: an experimental approach. Chem. Geol. 119, 139–160.
systems. In: Geochemistry of the Earth’s Surface. Balkema, Rotterdam, pp. 461– Ullman, W.J., Welch, S.A., 2002. Organic ligands and feldspar dissolution. In:
463. Hellmann, R., Wood, S.A. (Eds.), Water–Rock Interactions, Ore Deposits, and
Pokrovsky, O.S., Schott, J., 2000a. Forsterite surface composition in aqueous Environmental Geochemistry: A Tribute to David A. Crerar. Geochem. Soc.
solutions: a combined potentiometric, electrokinetic, and spectroscopic Special Pub. 7, pp. 3–35.
approach. Geochim. Cosmochim. Acta 64, 3299–3312. Ullman, W.J., Kirchman, D.L., Welch, S.A., Vandevivere, P., 1996. Laboratory evidence
Pokrovsky, O.S., Schott, J., 2000b. Kinetics and mechanism of forsterite dissolution at for microbially mediated silicate mineral dissolution in nature. Chem. Geol. 132,
25 °C and pH from 1 to 12. Geochim. Cosmochim. Acta 64, 3313–3325. 11–17.
Pokrovsky, O.S., Dupré, B., Schott, J., 2005. Fe–Al-organic colloids control of trace Van Herk, J., Pietersen, H.S., Schuiling, R.D., 1989. Neutralization of industrial-waste
elements in peat soil solutions: results of ultrafiltration and dialysis. Aquat. acids with olivine – the dissolution of forsteric olivine at 40–70 °C. Chem. Geol.
Geochem. 11, 241–278. 76, 341–352.
Pokrovsky, O.S., Shirokova, L.S., Bénézeth, P., Schott, J., Golubev, S.V., 2009. Effect of Wang, F., Gaimmar, D.E., 2013. Forsterite dissolution in saline water at elevated
organic ligands and heterotrophic bacteria on wollastonite dissolution kinetics. temperature and high CO2 pressure. Environ. Sci. Technol. 47, 168–173.
Am. J. Sci. 309, 731–772. Welch, S.A., Ullman, W.J., 1993. The effect of organic acids on plagioclase dissolution
Poulson, S.R., Drever, J.I., Stillings, L.L., 1997. Aqueous Si-Oxalate complexing, rates and stoichiometry. Geochim. Cosmochim. Acta 57, 2725–2736.
oxalate adsorption onto quartz and the effect of oxalate upon quartz dissolution Welch, S.A., Ullman, W.J., 1996. Feldspar in acidic and organic solutions:
rates. Chem. Geol. 140, 1–7. compositional and pH dependence of dissolution rate. Geochim. Cosmochim.
Prigiobbe, V., Mazzotti, M., 2011. Dissolution of olivine in the presence of oxalate, Acta 60, 2939–2948.
citrate and CO2 at 90 and 120 °C. Chem. Eng. Sci. 66, 6544–6554. Weres, O., Yee, A., Tsao, L., 1981. Kinetics of silica polymerization. J. Colloid Interface
Prigiobbe, V., Costa, G., Baciocchi, R., Hänchen, M., Mazzotti, M., 2009. The effect of Sci. 84, 379–402.
CO2 and salinity on olivine dissolution kinetics at 120 °C. Chem. Eng. Sci. 64, Wogelius, R.A., Walther, J.V., 1991. Olivine dissolution at 25 °C: effects of pH, CO2
3510–3515. and organic acids. Geochim. Cosmochim. Acta 55, 943–954.
Reddy, M.M., 1977. Crystallization of calcium carbonate in the presence of trace Wogelius, R.A., Walther, J.V., 1992. Olivine dissolution kinetics at the near-surface
concentrations of phosphorous containing anions. I. Inhibition by phosphate conditions. Chem. Geol. 97, 101–112.
and glycerophosphate ions at pH 8.8 and 25 °C. J. Cryst. Growth 41, 287–295. Wolf, G.H., Chizmeshya, A.V.G., Diefenbacher, J., McKelvy, M.J., 2004. In-situ
Reddy, M.M., Wang, K.K., 1980. Crystallization of calcium carbonate in the presence observation of CO2 sequestration reactions using a novel microreaction
of metal ions. I. Inhibition by magnesium ion at pH 8.8 and 25 °C. J. Cryst. system. Environ. Sci. Technol. 38, 932–936.
Growth 50, 470–480. Wolff-Boenisch, D., Gislason, S.R., Oelkers, E.H., 2006. The effect of crystallinity on
Reynolds Jr., R.C., 1978. Polyphenol inhibition of calcite precipitation in Lake Powell. dissolution rates and CO2 consumption capacity of silicates. Geochim.
Limnol. Oceanogr. 23, 585–597. Cosmochim. Acta 70, 858–870.
Rimstidt, D., Brantley, S.L., Olsen, A.A., 2012. Systematic review of forsterite Wolff-Boenisch, D., Wenau, S., Gislason, S.R., Oelkers, E.H., 2011. Dissolution of
dissolution rate data. Geochim. Cosmochim. Acta 99, 159–178. basalts and peridotite in seawater, in the presence of ligands, and CO2:
Rosso, J.J., Rimstidt, J.D., 2000. A high resolution study of forsterite dissolution rates. implication for CO2 sequestration of carbon dioxide. Geochim. Cosmochim. Acta
Geochim. Cosmochim. Acta 64, 797–811. 75, 5510–5525.
Saldi, G.D., Daval, D., Morvan, G., Knauss, K.G., 2013. The role of Fe and redox
conditions in olivine carbonation rates: an experimental study of the rate

You might also like