Вы находитесь на странице: 1из 8

Marine Pollution Bulletin 76 (2013) 307314

Contents lists available at ScienceDirect

Marine Pollution Bulletin


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

Simulation of CO2waterrock interactions on geologic CO2


sequestration under geological conditions of China
Tianye Wang a, Huaiyuan Wang b, Fengjun Zhang a,, Tianfu Xu a
a
b

Key Laboratory of Groundwater Resources and Environment, Ministry of Education, Jilin University, Changchun 130021, China
Jilin Institute of Geological Environment Monitoring, Changchun 130021, China

a r t i c l e

i n f o

Keywords:
Geologic CO2 sequestration
CO2waterrock interactions
Mineral corrosion and dissolution
New mineral formation

a b s t r a c t
The main purpose of this study focused on the feasibility of geologic CO2 sequestration within the actual
geological conditions of the rst Carbon Capture and Storage (CCS) project in China. This study investigated CO2waterrock interactions under simulated hydrothermal conditions via physicochemical analyses and scanning electron microscopy (SEM). Mass loss measurement and SEM showed that corrosion of
feldspars, silica, and clay minerals increased with increasing temperature. Corrosion of sandstone samples in the CO2-containing uid showed a positive correlation with temperature. During reaction at
70 C, 85 C, and 100 C, gibbsite (an intermediate mineral product) formed on the sample surface. This
demonstrated mineral capture of CO2 and supported the feasibility of geologic CO2 sequestration.
Chemical analyses suggested a dissolutionreprecipitation mechanism underlying the CO2waterrock
interactions. The results of this study suggested that mineral dissolution, new mineral precipitation,
and carbonic acid formation-dissociation are closely interrelated in CO2waterrock interactions.
2013 Elsevier Ltd. All rights reserved.

1. Introduction
With the industrialization and socioeconomic development of
human society, the consumption of fossil fuels has greatly increased. Air pollution and the greenhouse effect associated with
carbon dioxide (CO2) emission have become a major threat to the
environment on a global scale. Despite this, CO2 emissions have
continued to increase over recent decades (Shukla et al., 2010).
The annual CO2 emission rate in China has more than doubled that
of the US since 2001 (Gregg et al., 2008). Consequently, techniques
are urgently needed for the effective management of CO2 emissions in China. Geologic sequestration is one of the most promising
options for long-term CO2 storage and climate change mitigation
(Bachu et al., 2007). In this technique, CO2 is injected into conned
geological formations and sequestered via the process of mineral
capture (Xu et al., 2006). However, the feasibility of geologic CO2
sequestration with regard to the actual geological conditions of
China is an issue requiring urgent attention.
Mineral capture of CO2 occurs when CO2 dissolved in deep saline
aquifers interacts with minerals to form carbonate minerals (Carroll
et al., 2012). CO2waterrock interaction experiments are considered to be one of the most effective ways to understand and explore
the mechanisms and processes of geological CO2 sequestration,
Corresponding author. Tel.: +86 (0)431 88498718.
E-mail address: noforname@hotmail.com (F. Zhang).
0025-326X/$ - see front matter 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.marpolbul.2013.08.014

which, in straightforward terms, reects the formation of carbonate


minerals (Ketzer et al., 2009). Both temperature and pressure inuence the efcacy of geologic CO2 sequestration. Under simulated
conditions of low temperature and pressure, previous experiments
did not produce obvious mineralization reactions (Angeli et al.,
2009). However, mineral corrosion and new mineral formations
were enhanced under conditions of higher temperature and pressure. These results were supported by SEM and XRD analyses, which
indicated intense dissolution of albite, K-feldspar and dolomite
(Wigand et al., 2008). In addition to dissolution reactions, the formation of montmorillonite, illite and dawsonite were observed (Liu
et al., 2011, 2012). The changes in pH and cation concentrations in
the solution were enhanced with increasing temperature owing to
the dissolution of soluble minerals. The formation of new minerals
was also more intense (Alemu et al., 2011; Wigand et al., 2008).
The kinetics equation indicated that the calcite dissolution rate increased as temperature increased (Qingjie et al., 2008). Results of
modeling experiments have predicted that the minerals able to trap
CO2 are dawsonite and dolomite, while siderite or calcite will degrade in high heat conditions. The mineral-trapping capacity for
the sandstone aquifers was relatively low: 1.21.9 kg CO2/m3, and
the solubility trapping capacity did not exceed 4.07 kg CO2/m3
(Labus and Bujok, 2011).
This study focused on the feasibilities of geologic CO2 sequestration and the mechanism of CO2waterrock interaction within the
actual geological conditions of China. Rock samples were cored

308

T. Wang et al. / Marine Pollution Bulletin 76 (2013) 307314

from the reservoir rock in Erdos, which is the location of the rst
CO2 Capture and Storage (CCS) project in China. The present study
investigated the CO2waterrock interactions in the high-pressure
reactor under different temperature conditions (55 C, 70 C, 85 C,
and 100 C) and the actual pressure of the sandstone formation
(18 MPa). Mineral corrosion and dissolution, new mineral
formation, and reactant ion concentrations were studied in detail
to determine the mechanisms of the CO2waterrock interactions.
The results of this study will help in rening the reaction
parameters for modeling predictions, and have signicance for
the implementation of CCS projects in China.

Table 2
Composition of the reaction solution (1 L).
Chemical

Concentration (mmol/L)

Mw.

Mass (g)

CaCl2
MgCl26H2O
KNO3
NaHCO3
FeSO47H2O
NaCl

161.1300
8.2800
0.6300
1.0300
0.1698
166.7600

110.9834
203.3024
101.1032
84.0069
278.0160
58.4425

17.8828
1.6833
0.0637
0.0865
0.0472
9.7459

2. Materials and methods


2.1. Materials
The samples were cored from the reservoir rock (at 1860
1866 m depth) in Erdos, the location of the rst CCS project in
China. The formation pressure was 18.30 MPa, and the stratum
temperature was between 55 and 75 C. A small fraction of the
samples were processed into slices (10  10  1 mm) for the
detection of SEM. The petrological characteristics of the samples
were analyzed by X-ray diffraction (XD-3, Purkinje General Instrument, Beijing, China), as shown in Table 1.

2.2. Methods
A reaction solution (Table 2) with a similar composition to the
underground water at the sampling site was prepared. We loaded
750 ml of this solution and an appropriate amount of sandstone
samples (both of slice and particle forms) into an FYX high-pressure reactor (Fig. 1; Tong-Chan High-pressure Reactor Manufacturer, Dalian, Liaoning, China) (Zhang et al., 2012). The sandstone
samples were weighed before every experiment. The reactor was
set to operate at 18 MPa and a series of temperatures (55 C,
70 C, 85 C, and 100 C). To simulate the actual conditions of geologic CO2 sequestration, no stirring was performed. Meanwhile,
CO2 was continuously purged into the reaction solution. After reaction for various time periods (6 h, 3 d, 6 d, 12 d, 18 d and 24 d), the
solution was sampled for pH and reactant concentrations analysis.
Potassium (K+), sodium (Na+), calcium (Ca2+), and magnesium
(Mg2+) ions concentrations were determined by the inductively
coupled plasma mass spectrometry (ICPMS 7500A, Agilent, Santa
Clara, CA, USA).
The remaining sandstone sample was retrieved, rinsed repeatedly with distilled water, dried at 105 C for 24 h, and weighed.
The surface morphology of the sample was analyzed with a JSM6700 scanning electron microscope (SEM, JEOL, Tokyo, Japan).
The elemental composition of the sample was determined with
an INCAX-SIGHT energy-dispersive X-ray spectrometer (EDS,
Oxford, UK) attached to the SEM.

Table 1
Petrological characteristics of the sandstone sample.
Mineral

V%

Silica
Alkali feldspar
Plagioclase
Calcite
Illite
Illite/smectite interstratied minerals
Total

65
9
16
3
2
5
100

Fig. 1. FYX hydrothermal reactor; this reactor allows reaction under constant
pressure (up to 30 MPa) and temperature (up to 350 C) at adjustable stirring rates.
Air inlet and outlet in the reactor allows air purging; a sampling port under the
reactor allows liquid sampling at regular intervals.

3. Results and discussion


3.1. Sample corrosion and dissolution
3.1.1. Mass change
Table 3 summarizes the changes in the sample mass after the
reaction with CO2 at various temperatures. The mass loss increased
with increasing temperature, indicating enhanced mineral corrosion at higher reaction temperatures. An exception was the mass
loss at 85 C, which was smaller than the loss at other temperatures, probably because of the formation of new minerals on the
sample surface.
3.1.2. Feldspar dissolution
The dissolution of rock minerals in water is classied into congruent and incongruent dissolution (Cai et al., 2002). Feldspar dissolution in water is an incongruent reaction, meaning some
constituent ions enter the solution whereas other constituents
are transformed into new minerals (Zhu et al., 2011). The feldspar
in the Erdos rock samples is primarily composed of K-feldspars, albite, and Anorthite (Oelkers et al., 1994; Wollast and Chou, 1992).
In the present study, these components were eroded by the acidic
CO2-containing uid as follows (Luo et al., 2001):

Table 3
Sample mass changes after reaction at various temperatures.
Temperature (C)

55
70
85
100

Mass (g)
Before reaction

After reaction

Mass change

43.7082
43.8353
43.8125
44.3616

42.638
42.544
43.1147
43.0733

1.0702
1.2913
0.6978
1.2883

T. Wang et al. / Marine Pollution Bulletin 76 (2013) 307314

K-feldspars : KAlSi3 O8 4H2 O 4H


3

! K Al

3H4 SiO4 aq

Albite : NaAlSi3 O8 4H2 O 4H


3

! Na Al

3H4 SiO4 aq

2
3

Anorthite : CaAlSi3 O8 8H ! Ca2 Al

2H4 SiO4 aq

SEM imaging revealed that at 55 C (Fig. 2(a)), the K-feldspars


(Eq. (1)) on the sample surface underwent only negligible dissolution and retained their crystalline form (Fig. 2(a)). After reactions
at 70 C, 85 C and at 100 C (Fig. 2(b)(d)), the surface K-feldspars
underwent weak dissolution. The albite (Eq. (2)) showed no noticeable dissolution at 55 C (Fig. 3(a)), underwent weak dissolution
after the reaction at 70 C (Fig. 3(b)), and was substantially
dissolved to form long columns after the reaction at 85 C and
100 C (Fig. 3(c) and (d)). These ndings showed a progressive
increase in the dissolution of feldspar components with increasing
reaction temperature. Moreover, albite underwent a more severe
dissolution reaction than the K-feldspars at the same temperature
(Dong, 2011; Meng et al., 2006). Salt hydrolysis reactions are
endothermic reactions, while the reactions whereby CO2 dissolves
in water and forms carbonates with the cations are endothermic
reactions. The dissolution reaction rate was consistent with the
change of ion concentrations in the solution, which is claried in
Section 3.3. Consequently, feldspar dissolution increased with
increasing temperature.
3.1.3. Silica corrosion
After the reaction at 55 C, no silica corrosion was observed at
the sample surface and only minor corrosion was observed after
the reactions at 70 C, 85 C and at 100 C (Fig. 4(a)(d)). Earlier
studies have shown that the corrosion of silica is primarily dependent on temperature and pH. For example, Blatt et al. (Blatt et al.,

309

1980) found that the solubility of silica was constant in the pH


range of 2.08.5 and increased when the pH exceeded 8.5. Other
studies have reported that silica underwentlittle corrosion at
100 C, minor corrosion at 200 C, and intense corrosion at
250 C(Alemu et al., 2011; Kaszuba et al., 2003; Zhang et al.,
2012). These results showed that silica is generally resistant to corrosion at <100 C and a higher temperature is required to induce
noticeable corrosion of this mineral. The experimental conditions
(pH and temperature) in the present study were insufcient for
inducing strong corrosion of silica.
3.1.4. Clay mineral dissolution
The dissolution of the clay minerals after the reaction at 55 C
(Fig. 5(a)) was limited, compared with before the reaction
(Fig. 5(e)). After reaction at 70 C (Fig. 5(b)), the clay minerals dissolved to form stratied structures. After reaction at 85 C
(Fig. 5(c)) and 100 C (Fig. 5(d)), the clay minerals dissolved to form
foliate features. The results showed that the dissolution of the clay
minerals was slightly enhanced with increasing temperature, similar to the corrosion pattern of the feldspars and silica.
3.2. New mineral formation
SEM observations suggest that no new mineral phases formed
at the sample surface after the reaction at 55 C. In comparison,
after the reactions at 70 C, 85 C, and 100 C (Fig. 6(a)(c)), new
minerals appeared on the sample surfaces, showing acicular, oral,
or spherical morphologies. The EDS analysis (Table 4) revealed
that these new minerals consisted mainly of oxygen (O), aluminum
(Al), and silicon (Si). Taking into account SiO2 interference,
these results indicated that the new minerals were gibbsite
(Al2O3nH2O).
The formation of gibbsite precipitate on the sample surface
suggested that this mineral is an intermediate product of the reaction. If the system is allowed to react at a higher temperature for an

Fig. 2. Scanning electron micrograph showing K-feldspar dissolution after reaction at (a) 55 C, (b) 70 C, (c) 85 C and (d) 100 C.

310

T. Wang et al. / Marine Pollution Bulletin 76 (2013) 307314

Fig. 3. Scanning electron micrograph showing albite dissolution after reaction at (a) 55 C, (b) 70 C, (c) 85 C and (d) 100 C.

Fig. 4. Scanning electron micrograph showing silica corrosion at (a) 55 C, (b) 70 C, (c) 85 C and (d) 100 C.

extended period, this intermediate mineral may be further


transformed into more insoluble carbonates such as magnesite,
siderite, or dawsonite (Berg and Banwart, 2000; Kaszuba et al.,
2003; Xu et al., 2005). Although such insoluble carbonates were

not detected by SEM in our study, the formation of gibbsite showed


that CO2 can be captured by rocks via mineral sequestration, thus
supporting the feasibility of geologic CO2 sequestration by chemical entrapment (Berg and Banwart, 2000; Xiyu et al., 2008).

T. Wang et al. / Marine Pollution Bulletin 76 (2013) 307314

311

Fig. 5. Scanning electron micrograph showing clay mineral dissolution after reaction at (a) 55 C, (b) 70 C, (c) 85 C, (d) 100 C and (e) before the experiment.

3.3. Solution condition changes


3.3.1. pH changes
During the reaction, the solution pH (Fig. 7) initially decreased
sharply, then rose again, and nally stabilized. During day 1, the
pH decreased and the solution remained slightly acidic. At this initial stage, the CO2 injection reduced the pH via two reactions:

CO2 H2 O
H2 CO3

H2 CO3
H HCO3

Between days 3 and 24, the pH increased and nally stabilized.


The continuous dissolution of feldspar resulted in the release of
cations into the solution. The now acidic uid reacted with the
mineral components (Eqs. (1)(3)), thereby consuming H+ and
increasing the pH. Although the formation of new minerals tended
to reduce the pH during the reaction, the rate of mineral dissolution exceeded that of new mineral formation (Qu, 2007). Thus,
the pH did not decrease markedly. At the end of the 24-day reaction, the nal pH values were higher the in the experiments with
higher reaction temperature: 5.68 at 55 C, 5.85 at 70 C, 6.33 at

85 C, and 6.65 at 100 C. This trend indicates a positive correlation


between mineral corrosion and reaction temperature. A higher
reaction temperature resulted in accelerated mineral corrosion,
which led to increased H+ consumption and a greater pH increase.
3.3.2. Metal cation concentration changes
Most of K+ and Na+ were derived from the dissolutions of feldspar. During the reaction, the K+ and Na+ concentrations (Fig. 8(a)
and (b)) increased continuously, indicating continuous corrosion
of K-feldspars and albite. Feldspar corrosion also accelerated with
increasing temperature. After 6 h, intermediate complexes were
generated by the reactions of CO2
and Ca2+ and Mg2+ in the
3
solution, which absorbed some K+ and Na+. Where the corrosions
of K-feldspars and albite occurred only weakly, the concentrations
of K+ and Na+ decreased. The dissolution of feldspar increased gradually over time, which increased the concentrations of K+ and Na+.
The concentrations of Ca2+ and Mg2+ (Fig. 8(c) and (d)) changed
substantially with time but their nal concentrations (i.e., at the
end of the reaction) varied only moderately with temperature.
During the reaction at 55 C, the Ca2+ and Mg2+ concentrations
decreased initially, reached a nadir at 6 h, and subsequently

312

T. Wang et al. / Marine Pollution Bulletin 76 (2013) 307314

Fig. 6. The formation of new minerals; (a) acicular minerals formed after reaction at 70 C; (b) spherical minerals formed after reaction at 85 C; and (c) acicular and oral
minerals formed after reaction at 100 C.

Table 4
Energy dispersive X-ray data for new minerals formed on samples.
Temperature (C)

pH

70
85
100

8
7
6
5
4
3
2
1
0

xB (%)

xB (%)

AL

Si

AL

Si

56.19
52.91
29.45

3.58
2.48
6.93

40.23
44.61
63.62

69.17
66.31
42.19

2.61
1.85
5.89

28.21
31.85
51.93

2
via the reaction of Ca2+, Mg2+ and HCO
3 and CO3 (forming minerals such as dolomite) (Holdren and Speyer, 1985). Consequently,
the Ca2+ and Mg2+ concentrations initially decreased. Over time,
the corrosion of the sandstone was gradually enhanced by the
CO2 uid, which caused soluble minerals to dissolve constantly
in the water. Six days later, the rate of dissolution exceeded that
of new mineral formation and, as a result, the Ca2+ and Mg2+ concentrations started to increase. The reaction processes were as
follows:

CO2 H2 O
H2 CO3
H HCO3
H CO2
3

Fe=Mg5 Al2 Si3 O10 OH8 5CaCO3


2CO2
5CaFe=MgCO3 2 Al2 Si2 O5 OH4 SiO2
55

0.25

70

85

12

2H2 O

100

18

24

Time (d)

CaCO3 s H2 O CO2
HCO3
CaHCO3 ;

CaHCO3
HCO3 Ca2

Fig. 7. Change of solution pH during the reaction.

MgCO3 s H2 O CO2
HCO3
increased between 0.25 and 24 d. During the reactions at 70 C,
85 C, and 100 C, the Ca2+ and Mg2+ concentrations decreased in
the rst 6 d. They subsequently increased, and tended to stabilize
within 1224 d. This V-shaped pattern of concentration changes
may be explained as follows. The increasing concentrations of
Ca2+ and Mg2+ in the solution were mainly derived from the corrosion of silicate mineral and the dissolution of clay minerals. At the
early stage of the reaction, the rocks had not undergone any dissolution. The CO2 injected into the solution dissolved to form HCO
3
and CO2
3 . New Ca- and Mg-containing carbonate intermediate
complexes or carbonate precipitations may have been generated

MgHCO3 ;

MgHCO3
HCO3 Mg2

The different reaction temperatures produced different degrees


and rates of mineral corrosion and dissolution, which gave different ion concentration proles. However, at the end of the 24-d
reaction, all the ion concentrations had stabilized to similar regardless of the reaction temperature.
SEM observations and physicochemical analyses both revealed
that the corrosion of feldspars, silica, and clay minerals was enhanced with increasing temperature. The corrosion rate of albite
exceeded that of K-feldspars at the same temperature.

313

T. Wang et al. / Marine Pollution Bulletin 76 (2013) 307314

80

8000

70

7000

60

6000

Na+ (mg/L)

K+ (mg/L)

50
40
30

5000
4000
3000

20

2000

10

1000

0
0

0.25

12

18

24

0.25

70

12

18

24

Time (d)

Time (d)
55

85

55

100

c 12000

70

85

100

700
600

Ca2+ (mg/L)

10000
500
8000

400
300

6000

200
4000
100
2000

0.25

12

18

24

0.25

Time (d)
55

70

85

12

18

24

Time (d)
100

55

70

85

100

Fig. 8. Change of ion concentration during the reaction (a) K+, (b) Na+, (c) Ca2+, (d) Mg2+.

4. Conclusions
CO2waterrock interactions were studied in a simulated
hydrothermal environment at four temperatures (55 C, 70 C,
85 C, and 100 C). The corrosion and dissolution of the samples,
new mineral formation, and the conditions of the reaction solution
were investigated in detail. The results provide insights into the
mechanisms of CO2waterrock interactions.
The corrosion of feldspars, silica, and clay minerals by the
CO2-containing uid was positively correlated with the reaction
temperature. During the reactions at 70 C, 85 C, and 100 C,
gibbsite formed on the sample surface. The changes in the solution
pH and mineral ion concentrations suggest that mineral dissolution, new mineral precipitation, and carbonic acid formationdissociation are closely interrelated in CO2waterrock interactions. The corrosion of feldspars, silica, and clay minerals increases
with increasing temperature. At the same temperature, the
corrosion rate of albite feldspar is higher than that of K-feldspars.
Although SEM analysis did not nd any insoluble carbonates
(e.g. magnesite, siderite, or dawsonite), the formation of gibbsite
suggested that CO2 can be captured by rocks via mineral entrapment, thus supporting the feasibility of geologic sequestration
(underground storage) of CO2 by chemical entrapment in the rst
CCS project of China.
References
Alemu, B.L., Aagaard, P., et al., 2011. Caprock interaction with CO2: a laboratory
study of reactivity of shale with supercritical CO2 and brine. Applied
Geochemistry 26 (12), 19751989.

Angeli, M., Soldal, M., et al., 2009. Experimental percolation of supercritical CO2
through a caprock. Energy Procedia 1 (1), 33513358.
Bachu, S., Bonijoly, D., et al., 2007. CO2 storage capacity estimation: methodology
and gaps. International Journal of Greenhouse Gas Control 1 (4), 430443.
Berg, A., Banwart, S.A., 2000. Carbon dioxide mediated dissolution of Ca-feldspar:
implications for silicate weathering. Chemical Geology 163 (14), 2542.
Blatt, H., Middleton, G.V., et al., 1980. Origin of Sedimentary Rocks. Prentice2Hall,
Englewood Cliffs, New Jersey.
Cai, J., Xie, Z., et al., 2002. Diagenesis and pore evolution of deep sandstones in
Jiyang depression. Oil & Gas Geology (01), 8488.
Carroll, S.A., McNab, W.W., et al., 2012. Reactivity of Mount Simon sandstone and
the Eau Claire shale under CO2 storage conditions. Environmental Science and
Technology 47 (1), 252261.
Dong, L., 2011. The Characteristics and Mechanism of CO2Pyroclastic Rock
Interaction Doc., Jilin University.
Gregg, J.S., Andres, R.J., et al., 2008. China: emissions pattern of the world leader in
CO2 emissions from fossil fuel consumption and cement production.
Geophysical Research Letters 35 (8), L08806.
Holdren, J.G.R., Speyer, P.M., 1985. PH dependent changes in rates and
stoichiometry of dissolution of an alksli feldspar at room temperature.
American Journal of Science 285, 9541026.
Kaszuba, J.P., Janecky, D.R., et al., 2003. Carbon dioxide reaction processes in a
model brine aquifer at 200 C and 200 bars: implications for geologic
sequestration of carbon. Applied Geochemistry 18 (7), 10651080.
Ketzer, J.M., Iglesias, R., et al., 2009. WaterrockCO2 interactions in saline aquifers
aimed for carbon dioxide storage: experimental and numerical modeling
studies of the Rio Bonito Formation (Permian), southern Brazil. Applied
Geochemistry 24 (5), 760767.
Labus, K., Bujok, P., 2011. CO2 mineral sequestration mechanisms and capacity of
saline aquifers of the Upper Silesian Coal Basin (Central Europe) modeling and
experimental verication. Energy 36 (8), 49744982.
Liu, N., Liu, L., et al., 2011. Genesis of authigene carbonate minerals in the Upper
Cretaceous reservoir, Honggang Anticline, Songliao Basin: a natural analog for
mineral trapping of natural CO2 storage. Sedimentary Geology 237 (34), 166
178.
Liu, F., Lu, P., et al., 2012. CO2brinecaprock interaction: reactivity experiments on
Eau Claire shale and a review of relevant literature. International Journal of
Greenhouse Gas Control 7, 153167.

314

T. Wang et al. / Marine Pollution Bulletin 76 (2013) 307314

Luo, X., Yang, W., et al., 2001. Effects of pH on the solubility of the feldspar and the
development of secondary porosity. Bulletin of Mineralogy Petrology and
Geochemistry 02, 103107.
Meng, Z., Sijing, H., et al., 2006. Thermodynamics model for the characteristic of the
dissolution of the dissolution of primary minerals related to clastic diagensis.
Xinjiang Geology 02, 187191.
Oelkers, E.H., Schott, J., et al., 1994. The effect of aluminum, pH, and chemical
afnity on the rates of aluminosilicate dissolution reactions. Geochimica et
Cosmochimica Acta 58 (9), 20112024.
Qingjie, G., Jun, D., et al., 2008. Calcite dissolution in deionized water from 50 C to
250 C at 10 MPa: rate equation and reaction order. Acta Geologica Sinica
English Edition 82 (5), 9941001.
Qu, X., 2007. The Experiment Research of CO2-Sandstone Interaction, and
Application CO2 Gas Reservoir. doc., Jilin University.
Shukla, R., Ranjith, P., et al., 2010. A review of studies on CO2 sequestration and
caprock integrity. Fuel 89 (10), 26512664.
Wigand, M., Carey, J.W., et al., 2008. Geochemical effects of CO2 sequestration in
sandstones under simulated in situ conditions of deep saline aquifers. Applied
Geochemistry 23 (9), 27352745.

Wollast, R., Chou, L., 1992. Surface reactions during the early stages of weathering of
albite. Geochimica et Cosmochimica Acta 56 (8), 31133121.
Xiyu, Q., Li, L., et al., 2008. Experiment on Debris-Arkosic sandstone reformation by
CO2 uid. Journal of Jilin University(Earth Science Edition) (06), 959964.
Xu, T., Apps, J.A., et al., 2005. Mineral sequestration of carbon dioxide in a
sandstoneshale system. Chemical Geology 217 (34), 295318.
Xu, T., Sonnenthal, E., et al., 2006. TOUGHREACTa simulation program for nonisothermal multiphase reactive geochemical transport in variably saturated
geologic media: applications to geothermal injectivity and CO2 geological
sequestration. Computers and Geosciences 32 (2), 145165.
Zhang, F., Wang, H., et al., 2012. Experiment on mechanism of CO2 uid interacting
with sandstone layer. Journal of Jilin University (Earth Science Edition) 42 (03),
821826.
Zhu, H., Qu, X., et al., 2011. Study on interaction between the feldspar and CO2 uid.
Journal of Jilin University (Earth Science Edition) 41 (03), 697706.

Вам также может понравиться