Journal Pre-proof CO2-brine-caprock interaction: Reactivity experiments on mudstone caprock of South-west Hub geo-sequestration project D.W. Jayasekara, P.G. Ranjith, W.A.M. Wanniarachchi, T.D. Rathnaweera, D. Van Gent PII:
S0920-4105(20)30106-6
DOI:
https://doi.org/10.1016/j.petrol.2020.107011
Reference:
PETROL 107011
To appear in:
Journal of Petroleum Science and Engineering
Received Date: 20 October 2019 Revised Date:
27 January 2020
Accepted Date: 28 January 2020
Please cite this article as: Jayasekara, D.W., Ranjith, P.G., Wanniarachchi, W.A.M., Rathnaweera, T.D., Van Gent, D., CO2-brine-caprock interaction: Reactivity experiments on mudstone caprock of Southwest Hub geo-sequestration project, Journal of Petroleum Science and Engineering (2020), doi: https:// doi.org/10.1016/j.petrol.2020.107011. This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version will undergo additional copyediting, typesetting and review before it is published in its final form, but we are providing this version to give early visibility of the article. Please note that, during the production process, errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain. © 2020 Published by Elsevier B.V.
CRediT author statement D. W. Jayasekara:: Conceptualization, Methodology, Investigation, Visualization, Writing -
Original Draft P.G. Ranjith: Resources, Supervision, Project administration and Funding acquisition W. A. M. Wanniarachchi T. D. Rathnaweera:: Formal analysis and Writing - Review & Editing, D. Van Gent: Sample analysis, geological informations, sample sourced
1
Cover Page
2
Manuscript Title: CO2-brine-caprock interaction: Reactivity experiments on mudstone
3
caprock of South-west Hub geo-sequestration project
4
Authors’ names:
5
D. W. Jayasekara1, P.G. Ranjith*1, W. A. M. Wanniarachchi1, T. D. Rathnaweera1, D. Van
6
Gent2
7
1
8
3800, Australia.
9
2
Department of Civil Engineering, Monash University, Building 60, Melbourne, Victoria,
Department of Mines and Petroleum, 100 Plain Street, East Perth, WA 6004, Australia.
10 11 12 13 14
*Corresponding author:
15
Prof. Ranjith PG
16
Deep Earth Energy Laboratory, Monash University, Building 60,
17
Melbourne, Victoria, 3800, Australia.
18
Phone/Fax: 61-3-9905 4982
19
E-mail:
[email protected]
20 1
21
Abstract
22
This study aims at focusing on the geo-chemical reactions of caprock upon injection of
23
supercritical CO2 (ScCO2) under deep saline aquifer’s conditions. The caprock samples were
24
obtained at a depth of 979 m in the DMP Harvey-2 well, which is recognized as mudstone in
25
the Lesueur Formation of the South-west Hub geo-sequestration project located at south of
26
Perth in Australia. Geo-chemical reactions were conducted in a reaction chamber to
27
determine the chemical reactivity of brine-saturated caprock under actual reservoir conditions
28
(40 °C, 10 MPa and salinity 4.5%). The reaction was conducted at 10 MPa ScCO2 pressure
29
and allowed for 37 weeks. The reacted fluid samples were subjected to several chemical
30
analyses, including alkalinity tests, pH measurements and inductively-coupled plasma optical
31
emission spectrometry (ICP-OES) tests. The solid minerals were tested using X-ray
32
Diffraction (XRD), Scanning electron microscopy (SEM) and Energy-Dispersive X-ray
33
(EDX) to obtain a better understanding related to of their chemical characterization of the
34
caprock. Finally, a geochemical model (PHREEQC) was developed to forecast mineral
35
dissolution/precipitation equilibrium and redox reactions.
36
Fluid chemistry showed that the concentration of major elements such as Ca, Mg and K
37
increased with time due to the dissolution of minerals such as K-feldspar, anorthite, and
38
chlorite. On the other hand, the release concentration of Si, Fe and Al ions decreased with
39
time due to the precipitation of secondary minerals such as kaolinite, gibbsite, amorphous
40
silica and Fe(OH)3. Accordingly, the dissolution of minerals is very significant compared to
41
precipitation of secondary minerals in the short term which can increase the pore volume of
42
the caprock.
43
Keywords: South-west Hub geo-sequestration project, caprock, mudstone, minerals,
44
chemical reactivity
45 46 47 48 49
2
50
1. Introduction
51
The increase of anthropogenic CO2 in the atmosphere has caused significant climate changes
52
in the world (Nguyen et al., 2017). In order to mitigate global climate change, it is crucial to
53
implement substantial emission reduction strategies to reduce anthropogenic CO2 in the
54
atmosphere. Storing CO2 in deep saline aquifers plays a significant role in CO2 sequestration
55
because of their high storage capacity, easy accessibility and extensive availability (De Silva
56
et al., 2015; Rathnaweera et al., 2014). Generally, deep saline aquifers lie at a depth range of
57
800- 2000 m and such a depth allows to retain the injected CO2 in its supercritical state due to
58
the high pressure and temperature prevailing at the depth, which is higher than the critical
59
point of CO2 (temperature >31.1 °C and pressure > 7.39 MPa) (Kharaka et al., 2009). Since
60
the density and viscosity of CO2 are lower than those of the reservoir pore fluid, the injected
61
CO2 undergoes viscous fingering and gravitational segregation while moving upward to the
62
top of the reservoir rock. During this process, the contamination of upwardly moving CO2
63
with overlying groundwater bodies is prevented by the caprock, a less permeable layer made
64
up of siliciclastic sedimentary rocks such as mudstone, claystone, shale or siltstone (Shukla et
65
al., 2010).
66
Injecting CO2 into deep saline aquifers can cause hydro-dynamic, thermal, chemical and
67
mechanical changes in the caprock. The main chemical variation which appears in the
68
caprock is the loss of geochemical equilibrium between the caprock minerals and the pore
69
fluid due to massive CO2 injection into the reservoir rock (Shao et al., 2010). The dissolution
70
of supercritical CO2 in brine, making an acidic medium capable of reacting with the caprock
71
matrix, leads to the dissolution of primary minerals such as quartz, plagioclase and biotite and
72
the precipitation of secondary minerals such as analcime and carbonate minerals (Kaszuba et
73
al., 2003). Such changes can significantly alter the porosity, morphology and permeability
74
jeopardising the entire CO2 storage process (Liu et al., 2012). It is therefore important to
75
understand the mechanics of the interactions of various caprock minerals with CO2 in order to
76
characterise the caprock behaviour in both short- and long-term injection scenarios.
77
In most cases, the effective minerals in terms of caprock integrity are clay minerals such as
78
kaolinite, montmorillonite, illite and chlorite (Shao et al., 2010). Due to the presence of
79
abundant clay minerals in the caprock, some characteristics such as ultra-low permeability,
80
low diffusion and high capillary entry pressure enhance the sealing capacity of the caprock
81
which reduces the risk of back-migration of injected CO2 into the surface and freshwater 3
82
aquifers (Bennion and Bachu, 2007). Generally, clay minerals in caprocks can be divided into
83
different groups based on their characteristics, as in Table 1, where different clay minerals act
84
differently according to the prevailed conditions in the aquifer.
85
A deep knowledge of the geo-chemical reactions in the caprock upon injection of CO2 is
86
critical for the site selection process. Several studies have been recently conducted to evaluate
87
the reactivity of caprock under different temperatures (Alemu et al., 2011; Credoz et al., 2009;
88
Kaszuba et al., 2005; Liu et al., 2012; Xiao et al., 2017). This is important to investigate as
89
temperature varies with the depth of the aquifer and has a significant influence on the
90
reactivity of caprock. A detailed summary and the related key findings of each study,
91
including geo-chemical models, are shown in Table 2. According to the previous studies, it is
92
clear that the laboratory hydro-thermal tests conducted so far are based on higher
93
temperatures (80-250 ºC) than the reservoir temperatures (30-50 ºC), and such high
94
temperatures accelerate the silicate mineral reaction rate in sedimentary rocks to predict long-
95
term mineral reactivity. For instance, Liu et al., (2012) conducted a study under the
96
temperature of 200 ºC and pressure of 30 MPa and Alemu et al., (2011) performed a test at a
97
temperature of 250 ºC and pressure of 11 MPa to accelerate the mineral reaction rates.
98
However, the results of hydro-thermal reactions are not very reliable because high
99
temperatures can change the mineral reaction mechanism (Liu et al., 2012). For example,
100
when the aquifer temperature is less than 60 ºC, the carbonate caprocks, which contain
101
dolomite and calcite start to dissolve in the presence of CO2, while at high temperatures (> 60
102
ºC) precipitation can be expected as carbonate minerals (Song and Zhang, 2013). Moreover,
103
smectite becomes unstable at the temperature of 80-100 ºC and starts to dissolve, proving the
104
temperature effect on the reaction mechanism (Stephansson et al., 2004). Although Kohler et
105
al. (2009) state that a lot of mineralogical changes due to temperature shift are related to
106
dehydration, there are a few reactions that change its mechanism due to temperature. For
107
instance, Alemu et al., (2011) conducted three different batch reactions at 80, 150 and 200 ºC
108
for clayey shale which consisted of quartz 26 %, albite 8 %, pyrite 1 %, siderite 5 %, illite
109
22 %, and chlorite 38 % and found that different mineral dissolutions and precipitations took
110
place according to the temperature. Results depicted that albite, K-feldspar and illite
111
precipitated at 80 and 150 ºC while dissolved at 200 ºC even though quartz behaviour was
112
opposite of it. Moreover, siderite mineral dissolved at 80 ºC whereas it precipitated at 150
113
and 200 ºC. Furthermore, the SiO2 concentration identified using brine chemistry showed that
114
its concentration reduced with the increase of temperature. In addition, K-feldspar- NaCl 4
115
water mixture is transformed into gibbsite, kaolinite, kaolinite + quartz, aragonite + quartz,
116
and finally into albite as a result of mineral-water interactions (Bjørlykke et al., 1992). But, if
117
the temperature is greater than 200 ºC, boehmite is formed instead of gibbsite formation in
118
this reaction series (Fu et al., 2009). Thus, it is clear that the high temperature has an ability
119
to change the reaction mechanism, which can induce different primary mineral dissolutions
120
and secondary mineral precipitations (Liu et al., 2012). As a result, if we provide high
121
temperatures in batch reactions which are greater than the reservoir conditions, unrealistic
122
chemical reactions can result, which are different from the real aquifer reactions. In addition,
123
the temperature has a great impact on the CO2 dissolution rate, changing the pH value in the
124
solution. According to Alemu et al., (2011), clayey shale saturated in brine showed a low pH
125
value due to high dissolution of CO2 at lower temperatures. As a result, high silica and iron
126
(Fe) concentrations were observed in the solution because low pH creates a highly acidic
127
medium. Therefore, conducting experiments at simulated aquifer temperatures is crucial to
128
identify the realistic chemico-mineralogical changes of caprock in deep saline aquifers (Liu et
129
al., 2012).
130
Although there are several studies done on caprock (hydro-thermal experiments) (Alemu et
131
al., 2011; Anabaraonye et al., 2019; Credoz et al., 2009; Kaszuba et al., 2005; Kohler et al.,
132
2009; Liu et al., 2012), very limited studies have focused on the reactivity of caprock under
133
reservoir conditions (especially the temperature < 50 ºC) (Dávila et al., 2017; Ellis et al.,
134
2011; Ilgen et al., 2018; Pearce et al., 2019; Tarkowski et al., 2015). Among them, the
135
reactivity tests conducted for mudstone are very limited (Pearce et al., 2019; Tarkowski et al.,
136
2015; Wdowin et al., 2014). Thus, the understanding regarding chemico-mineralogical
137
behaviour of mudstone caprock is narrow, which needs to be investigated in detail.
138
Tarkowski et al. (2015) conducted reactivity tests for 20 months at room temperature under
139
CO2 pressure of 6 MPa for mudstone obtained from Zaosie Anticline which is considered as a
140
potential CO2 storage site. Results depicted that the surface area and porosity of mudstone
141
were increased while bulk density was reduced after the batch reaction. Similarly, Wdowin et
142
al. (2014) also conducted reactivity tests for mudstone obtained from Chabowo Anticline at a
143
temperature of 25 ºC and pressure of 6 MPa for 18 months and found that there were
144
dissolution of feldspar and precipitation of kaolinite. However, the use of powdered sample
145
to accelerate the chemical reactions, identification of possible chemical reactions and
146
pressurizing the chamber with supercritical CO2 could improve the quality of these studies. In
147
addition, Pearce et al. (2019) conducted a combined reactivity and µCT study on mudstone 5
148
obtained from Surat Basin, Queensland, Australia to identify the relationship between the
149
permeability changes and reactivity of caprocks after CO2 injection. The pressure and
150
temperature in lined reactors were 12 MPa and 60 ºC respectively and found that there was a
151
slight dissolution of chlorite, feldspar and calcite, which could change the permeability. But
152
still, the findings would be more meaningful if authors could conduct geo-chemical
153
modelling to verify the possible reactions. According to Pearce et al. (2019), the reactivity
154
studies conducted on caprocks are very rare, especially under low salinity conditions. Thus,
155
reactivity tests were conducted in the current study including a geo-chemical modelling on
156
crushed mudstone obtained from South-west Hub geo-sequestration project, Australia to
157
increase the reaction rate providing reservoir temperature and pressure minimizing the
158
drawbacks in previous studies (Kaszuba et al., 2005; Tarkowski et al., 2015; Wdowin et al.,
159
2014). In addition, since the current study is based on low salinity formation fluid (4.5 %),
160
the results are very important as limited studies have been done on low salinity conditions
161
(Pearce et al., 2019). As a summary, this paper focuses on the evaluation of geo-chemical
162
reactions in mudstone caprock upon CO2 injection to better understand the caprock behaviour
163
under CO2 sequestration environment.
164
Table 1: Characteristics of clay minerals present in the caprocks (Kloprogge, 2017; Shainberg
165
and Levy, 2005) Clay group
Clay minerals
Cation exchange
Remarks
capacity (cmolckg-1) Kaolin
Kaolinite, dickite, halloysite,
about 1-10
and nacrite Smectite
Montmorillonite, nontronite
capacity 80-120
and beidellite Illite
Clay micas
low shrinkage
High swelling capacity
20-40
Non-expanding clay minerals
Chlorite
Clinochlore, pennantite, nimite and chamosite
10-40
Non-swelling minerals
166 167 168
2. Experimental approach 2.1 Sample selection and description
6
169
The caprock samples for the current study was obtained from Harvey-2 well of South-west
170
Hub geo-sequestration project. The main idea behind the South-west Hub geo-sequestration
171
project establishment was to mitigate the CO2 emissions in the area using CCS techniques. It
172
is located in the Perth Basin, which extends over 1350 km along the southwestern margin of
173
Australia (Olierook et al., 2014). The stratigraphy of the Perth Basin is shown in Figure 1.
174
The storage site consists of two formations known as Upper Lesueur (Yalgorup or Myalup
175
member) and Lower Lesueur (Wonnerup member) which were developed during the late
176
Triassic period. The Wonnerup member, which consists of coarse sands can be used as a
177
good reservoir to store CO2, and the Yalgorup member can act as a sealing layer due to the
178
high presence of floodplains and paleosols (shaley and clayey sediments which can provide
179
better resistance to CO2 leakage) (Stalker et al., 2013). In order to conduct a detailed analysis
180
of the storage site, four wells known as DMP Harvey 1, 2, 3 and 4 were drilled (Figure 2) to
181
obtain core samples and log data. The temperature measurements of different DMP Harvey
182
wells are displayed in Figure 3.
183
The caprock sample for reactivity tests was obtained from a stratum between 979.10 – 979.40
184
m that belongs to Myalup member in the DMP Harvey-2 well, which was identified as
185
mudstone. The physical and mineralogical properties of the selected mudstone are shown in
186
Table 3. For the reactivity test, the mudstone sample was crushed in order to enhance the
187
reaction rate by increasing the surface area (Kweon and Deo, 2017). The crushed caprock
188
sample was dry-sieved to retain the fraction between 425 and 75 µm.
189
7
190 191
Table 2: Recent studies of the chemico-mineralogical behaviour of caprock due to CO2 injection
Purpose To identify the limestone caprock reactivity after interaction with mixed CO2-brine fluids under different physicochemical conditions (Credoz et al., 2009).
To
evaluate
Testing methods Key findings Reacted brine- Inductively coupled Intact caprock mineralogy plasma atomic emission • 45% calcite, 15% mixed-layer illite/smectite (I/S), 10% kaolinite, 10% spectroscopy (ICP –AES). quartz, 5% gypsum, 5% pyrite and 5% other components. Reacted solid fractions- Caprock-CO2 (aq)-brine tests Simultaneous thermal analysis • Fe-dolomite totally dissolved while calcite destabilized. There, carbonate (STA), X-ray powder diffraction minerals are the most reactive minerals. (XRD), scanning electron • New form of mixed carbonated mineral precipitated (composition was not microscopy (SEM), energy determined). dispersive spectrometry (EDS). • New smectite precipitated due to I/S destabilization and kaolinite dissolution. I/S + Kaolinite → Illitic I/S + Fe-Mg-Smectite • Pyrite was significantly dissolved and a new carbonated mineral precipitated. Here HCO3- ions in the solution substituted for S2- ions (All reactions happened in 365 days at a temperature and pressure of 150 °C and 0.1 MPa respectively). Caprock-CO2(Sc)-brine tests • Massive dissolution of Fe-dolomite (after 30 days at a temperature and pressure of 80 °C and 15 MPa respectively). • Destabilization of pyrite (after 90 days at a temperature and pressure of 80 °C and 15 MPa respectively). • Kaolinite dissolved, but I/S was stable (after 90 days at a temperature and pressure of 80 °C and 15 MPa respectively). • Smectite in the mixed-layer decreased so that the relative illite fraction increased (after 45 days at a temperature and pressure of 80 °C and 15 MPa respectively).
CO2–brine–shale Reacted solid fractions- SEM and Intact caprock mineralogy 8
caprock interactions in Eau Claire XRD tests Formation using hydrothermal experiments (Liu et al., 2012).
To evaluate long-term geochemical Chemical modelling reactions of caprock in the presence PHREEQC software of CO2 (Gaus et al., 2005)
• Major minerals are quartz, orthoclase, illite and minor mineral is chlorite. Caprock-CO2 (aq)-brine tests • Minor dissolution of K-feldspar and anhydrite. • Precipitation of illite and/or smectite, and siderite in the vicinity of pyrite. (All reactions happened after 60 days at a temperature and pressure of 200 °C and 30 MPa respectively). using Intact caprock mineralogy • 24.7% mica/illite, 21.5% quartz, 18% kaolinite, 12.3% plagioclase, 8.8% smectite, 4.1% chlorite, 2.8% pyrite, 2.1% K-feldspar, 1.6% siderite, 1% calcite and 3.1% other components. Short term (over 9 years at a temperature and pressure of 37 °C and 10.13 MPa respectively) • Calcite dissolution. Long term (over 15,000 years at a temperature and pressure of 37 °C and 10.13 MPa respectively) • Albite dissolution forming kaolinite and dawsonite. • Anorthite dissolution forming kaolinite. • Dissolution of illite and smectites, and precipitation of kaolinite, chalcedony, K-feldspar and large amounts of carbonate.
To investigate the effect of Reacted brine - Inductively coupled Intact caprock mineralogy mineralogical compositions on the plasma mass spectrometry (ICP – • 13% quartz, 6% plagioclase, 19% chlorite, 26% illite, 7% ankerite and 29% reactivity of shale (Alemu et al., MS) calcite. 2011) Reacted solid fractions Caprock-CO2(Sc)-brine tests X-ray fluorescence (XRF), XRD • Ankerite dissolution. • Precipitation of calcite. • Chlorite dissolution. • Illite dissolution to form smectite (All reactions happened after 5 weeks at a temperature and pressure of 250 °C and 11 MPa).
9
To evaluate the reactive behaviour of CO2(Sc) –brine-caprock under physical-chemical conditions (Kaszuba et al., 2005)
Reacted brine- Inductively coupled Intact caprock mineralogy plasma atomic emission • 65% illite, 27% quartz, 5% feldspar and 3% kaolinite. spectroscopy (ICP –OES) and ICP- Caprock-CO2(Sc)-brine tests MS • Precipitation of magnesite and siderite (after 45 days at a temperature and Reacted solid fractions- SEM, EDS pressure of 200 ºC and 20 MPa).
To identify the dissolution and Reacted brine- ICP-MS, pH precipitation characteristics of measurements phlogopite, a clay mineral in Reacted solid fractions- atomic caprocks (Shao et al., 2010) force microscopy (AFM) analysis, SEM-EDX, XRD and X-ray photoelectron spectroscopy (XPS)
Mineralogy • Phlogopite sample Phlogopite -CO2(Sc)-brine tests • Phlogopite dissolution and precipitation of amorphous silica and kaolinite as secondary minerals (after 6 days reaction at a pressure of 10.3 MPa and temperature of 95 ºC).
10
MIDDLE EARLY
180
JURASSIC
160
200
ALBIAN
Coolyena Group
Osbourne Formation
Warnbro Group
Leederville Formation South Perth Shale
Seal
Reservoir
Shore
Source
APTIAN BARREMIAN HAUTERIVIAN VALANGINIAN BERRIASIAN
TITHONIAN KIMMERIDGIAN OXFORDIAN CALLOVIAN BATHONIAN BAJOCIAN AALENIAN TOARCIAN PLEINSBACHIAN SINEMURIAN
Gage sandstone
Parmelia Group
Yarragadee Formation
Cadda Formation Cockleshell
Cattamarra Coal Measures
Gully fromation
Eneabba formation
HETTANGIAN
240
LATE EARLYMIDDLE
220
TRIASSIC
RHAET IAN
NORIAN CARNIAN
Myalup Member Lesueur Sandstone
LADINIAN ANISIAN
Wonnerup Member
SCYTHIAN
Sabina Sandstone
DZHULFIAN
UFIMIAN
ARTINSKIAN SAKMARIAN ASSELIAN
PRECAMBRIAN
192
Willespie Formation
MIDIAN
KAZANIAN KUNGURIAN
EARLY
280
PERMIAN
260
LATE
CHANGHSINGIAN
CONTINENTAL DEPOSITS
LATE LATE
140
STRATIGRAPHIC UNITS
CENOMANIAN
NEOCOMIAN
120
EARLY
100
AGE
CRETACEOUS
Ma
Sue Group
Redgate Coal Measures Ashbrook Sandstone Rosabrook Coal Measures Woodynook Sandstone Mosswood Formation Basement
Figure 1: Stratigraphy of the Perth basin (Backhouse and Crostella, 2000)
11
193
Figure 2: The location of four Harvey wells in South-west Hub geo-sequestration project
194 Temperature (ºC) 0
20
40
60
80
Depth (m)
0 500
Harvey-1
1000
Harvey-2
1500
Harvey-3
2000
Harvey-4
2500 3000
195 196
3500
Figure 3: Temperature of different Harvey wells (ODIN_Reservoir_Consultants, 2015)
197 198 199 200
12
201
Table 3: Physical and mineralogical properties of mudstone obtained from Harvey-2
202
well Rock Type
Mudstone
Colour
Dark grey
Mineralogical composition
Quartz – 32%, Goethite- 6%, Orthoclase -9%, Illite -20%, Kaolinite -6%, Chlorite -13%, Plagioclase -12% and other minerals in minor percentages (pyrite and montmorillonite)
Mean porosity
9.5 (+/- 4.8) %
Mean permeability
0.7
Grain density
2.64 g/cm3
203 204
2.2 Experimental set-up and procedures
205
A reaction chamber was utilized to achieve the steady-state reaction of the brine-caprock-CO2
206
system, as shown in Figure 4, which consists of reactor cell, thermal unit, gas inlet and outlet.
207
The maximum temperature and pressure that can be applied in the chamber is 100 ºC and 12
208
MPa, respectively. The reactor volume capacity is 1.25 L, and its inner diameter is 100 mm.
209
Gas flowing from the cylinder is pressurized up to desired pressure using a Teledyne ISCO
210
500D syringe pump and injected to the reactor cell. The interior pressure of the chamber is
211
measured using a GS4200 pressure transducer. In addition, the whole apparatus including the
212
chamber, tubing (316 stainless Steel), valves and o-rings (perfluoroelastomer polymer) is
213
made up of non-corrosive materials which do not react with CO2 and acidic medium to avoid
214
gas leaking and corrosion of the set-up. First, the crushed mudstone sample (75-425 µm) was
215
placed in the chamber and brine was poured into it. As identified using a formation tester in
216
the field, 4.5% (NaCl) salinity brine was used for this test series. The crushed rock sample
217
was stirred well in the synthetic brine to enhance dissolution. The reaction chamber was then
218
heated to 40 ºC, which is considered to be the temperature in the field using a heat band and
219
CO2 was injected into it under 10 MPa pressure. The temperature of 40 ºC and pressure of 10
220
MPa were maintained throughout the test series, and it was confirmed by monitoring the
221
temperature and pressure time to time using an infrared thermometer and pressure gauge
222
respectively. Thus, the crushed mudstone sample reacted under supercritical CO2, because the 13
223
used temperature (40 ºC) and pressures (10 MPa) are higher than the critical temperature
224
(31.1 ºC) and pressure (7.39 MPa) of CO2. Moreover, the purity level of injected CO2 is
225
99.9 %, and its moisture is less than 100 ppm. In different time intervals (2, 5, 9, 18 and 37
226
weeks), the solid and liquid samples were collected after cooling and releasing the pressure in
227
the reaction cell since a drain valve for sampling is absent in the current reaction chamber
228
(it’s an experimental limitation here). The brine samples collected after each reaction period
229
were filtered using 0.45 µm filter paper to separate the solution from the crushed rock sample.
230
The collected solid samples were analysed using XRD, SEM and EDX tests. The pH and
231
alkalinity measurements were taken for each collected solution. In addition, ICP-OES tests
232
were conducted to find the elemental concentration.
233
Figure 4: Schematic diagram of the reaction chamber used for caprock-brine-CO2
234
interaction
235
2.2.1
X-ray diffraction (XRD)
236
Powder XRD testing was conducted using the Bruker D8 cobalt XRD instrument at Monash
237
University, which consists of an X-ray source, an X-ray detector and the sample holder
238
(Figure 5a). For the current test, the X-ray diffraction patterns were recorded in the range of
239
5°≤ 2θ ≤70° and the minerals were quantified using the crystallography open database in the 14
240
DIFFRAC EVA 4.3 software using a semi-quantitative method. Accordingly, the intact
241
sample consisted of a high clay content with illite, chlorite and kaolin at around 42%.
242
Plagioclase is a mixture of anorthite and albite. The halite deposition on the CO2-reacted
243
caprock samples was considered to be an experimental artefact because the sample was not
244
washed with deionized water before the experiment.
245
2.2.2
Scanning Electron microscopy (SEM) analysis
246
The intact and reacted mudstone (crushed) samples were scanned under a FEI Nova
247
NanoSEM 450 FEGSEM (Figure 5b) which is available at the Monash Centre for Electron
248
Microscopy (MCEM) to identify the mineral structure changes of rock samples which were
249
collected after 2, 5, 9, 18 and 37 weeks reaction. Since the intact and reacted powder samples
250
were non-conductive, they were coated with Iridium (Ir), a highly conductive metal, to avoid
251
the accumulation of static charges on the specimen surface or simply ‘surface charging’
252
which can cause problems such as reducing the landing energy of incident electrons (Kim et
253
al., 2010). In order to identify the elements in the caprock samples, EDX analysis also was
254
conducted. a).
b).
255
Figure 5: a). Bruker D8 cobalt XRD instrument at Monash University, and b). FEI Nova
256
NanoSEM 450 FEGSEM
257
2.2.3
Finding alkalinity of reacted brine samples
258
The brine samples were titrated with H2SO4 to find the total alkalinity. For that purpose, the
259
alkalinity test kit (model AL-DT) was used, which consists of a digital titrator, standard
260
titration cartridges (H2SO4), a phenolphthalein indicator powder pillow and a Bromcresol
261
green-methyl red indicator powder pillow. The brine sample volume, titration cartridge type
262
according to the N factor and the digital multiplier were selected based on prior estimation of
263
the alkalinity range. First, the phenolphthalein indicator was added to the sample, but no
264
colour change could be observed. Second, the green-methyl red indicator was added to the 15
265
brine sample and titrated with H2SO4 acid (cartridge 1.6N) until the green turned to a light
266
pink end point. The number of digits recorded was used to calculate the alkalinity, as shown
267
in Eq.1. Figure 6 shows titration using the alkalinity test kit and pH measurement.
268
Digits required × digit multiplier = mg/l CaCO3 total alkalinity
269
2.2.1
(1)
Inductively coupled plasma optical emission spectrometry (ICP-OES)
270
The ICP-OES tests were conducted in the PerkinElmer Flagship facility at Monash
271
University, Melbourne (Figure 7) for each solution collected at different time intervals to
272
identify the chemical elements. The calibration standard curves were derived using internal
273
standards and the element concentrations of solutions were found. a).
274
b).
Figure 6: a) Titration of solution using alkalinity test kit, and b) pH measurement
16
275
Figure 7: ICP-OES set-up and the real-time colour viewing of ignited plasma using
276
plasma camera
277
2.2.2
278
2.2.5.1
Geo-chemical modelling Calculation of saturation indices
279
Geo-chemical modelling was built with PHREEQC software version 3.4.0 using the
280
LLNL.dat database. It is a computer program written in C programming language, which can
281
be used to perform different geochemical calculations, such as finding saturation indices (SIs),
282
advective-transport and inverse modelling (Parkhurst, 1995). SIs were calculated for each
283
reacted sample at different time intervals, depending upon the measured elemental
284
composition of the saturated solution. Hence, the ability of the solution to dissolve or
285
precipitate mineral types confirms the chemical reactivity of caprock minerals in the presence
286
of CO2 and brine. Generally, the SI is positive when the mineral tends to precipitate, negative
287
during mineral dissolution and zero when the mineral and solution are at chemical
288
equilibrium. SI can be calculated using the chemical activities of dissolved ions of minerals
289
such that Ion Activation Products (IAPs) and the solubility product (Ksp), as shown in Eq.2.
290
= log ( )
291
(2)
2.2.5.2
Batch geo-chemical modelling
292
Batch geochemical modelling estimates the possible geochemical reactions due to caprock-
293
brine-CO2 interactions. Since there is no flow in batch calculations, the dimensions are zero.
294
Here, the dissolution and precipitation reactions were calculated based on mineralogy of
295
caprock, CO2 solubility, initial brine properties and kinetic rates. The kinetic calculation was
296
run for mudstone which was reacted with CO2 and brine for 37 weeks at 40 °C and 10 MPa.
297
The simulation consists of two steps as equilibrating brine with caprock minerals at 40 °C
298
and then, pressurizing the system with CO2 up to 10 MPa. The model was simulated for 37
299
weeks to find the chemico-mineralogical reactions of caprock with time. In order to calculate
300
the mineral dissolution and precipitation rate, a general rate equation (Eq 3) derived using
301
transition state theory was used (Alemu et al., 2011; Helgeson et al., 1984).
302
= () ( ) [1 − "$%]
#
$
(3)
17
303
Where ' represents the mineral index and is dissolution (positive value)/precipitation
304
rate (negative value) of mineral which is measured in mol m-2 sec-1. denotes the reactive
305
surface area, () is the rate constant which depends upon temperature, refers to proton
306
activity, ( is the order of the reaction with respect to protons, ) represents equilibrium
307
constant for mineral water reaction and * depicts the corresponding ion activity product. The
308
rate constants, () at any temperature can be calculated using the rate constants at 25 ºC,
309
+, (mol m-2 sec-1) incorporating Arrhenius law as shown in Eq 4 (Lasaga, 1984).
310
() = +, exp [
311
Where denotes the universal gas constant (8.314 J mol-1 K-1), 9: is the activation energy (J
312
mol-1) and is the temperature in Kelvin. The kinetic data, +, obtained from different
313
literature was converted into 40 ºC and then the model was simulated for 37 weeks. Due to
314
non-reliability of kinetic rate constants, the model was run for several times incorporating
315
data published in different studies to find most accurate values (trial and error method). The
316
most suitable kinetic rate constants for each mineral were selected after comparing the
317
elemental concentration of reacted solution obtained from laboratory and model results. Table
318
4 shows the selected kinetic rate constants and other data used to calculate the
319
dissolution/precipitation rates of each caprock mineral under CO2 and brine saturation. Here,
320
the specific surface areas of minerals were calculated assuming the spherical shape with a
321
diameter of 2 µm for all clay minerals and 63 µm for the rest of the minerals.
012 3
4
4
](5 − +67.4,)
322
(4)
Table 4: Data used in batch modelling Minerals
Mass
Rate log25
Rate log40
Ea (kJ
(mol/kg
(mol m-2
(mol m-2
mol-1)
of
-1
n
-1
sec )
sec )
1×10-14
5.45×10-14
Surface
Reference
area
for kinetic
2
-1
(m g )
rates
0.036
(Gherardi
water) Quartz
4.26
87.70
0
et al., 2007; Palandri and Kharaka., 2004) Illite
0.47
3.98×10-13
9.68×10-13
46
0.1
0.461
(Köhler et
18
al., 2003) Chlorite
0.19
7.76×10-12
3.34×10-11
88
0.5
1.140
(Palandri and Kharaka., 2004)
Anorthite
0.17
1×10-8
1.38×10-8
16.6
1.41
0.035
(Palandri and Kharaka., 2004)
Albite
0.18
1×10-12
3.51×10-12
65
0.5
0.036
(Palandri and Kharaka., 2004)
K-feldspar
0.26
8.71×10-11
2.37×10-10
51.7
0.5
0.0373
(Palandri and Kharaka., 2004)
Kaolinite
0.19
4.90×10-12
1.75×10-11
65.9
0.2
1.16
(Palandri and Kharaka., 2004)
Goethite
0.54
1.15×10-13
3.42×10-8
86
0.5
0.022
(Palandri and Kharaka., 2004)
Pyrite
0.07
4×10-11
1.35×10-10
62.76
0.5
0.036
(Xu et al., 2005)
Montmorillo nite
0.02
1.7×10-13
3.34×10-13
35
0.5
0.045
(Gherardi et al., 2007; Palandri and Kharaka., 2004) 19
323 324 325
3. Results 3.1 Fluid chemistry
326
The ions (Si, Al, Fe, Ca, Mg and K) which were absent in the initial NaCl solution were
327
found at 2, 5, 9, 18 and 37 weeks due to the dissolution of reacting minerals, as shown in
328
Figure 8. According to the ICP-OES results, the concentration of Na+ ions in the solution did
329
not change significantly with time compared to the initial solution concentration. However,
330
Ca2+, Mg2+ and K+ ions were significantly released to the solution during chemical reactions.
331
The concentration of Si in the solution reached 10.67 ppm after a 2-week reaction, and it was
332
reduced by 46% when the reaction was continued for 5 weeks and doubled again after 9
333
weeks compared to 5th week. After that, the Si ion concentration gradually reduced up to 3.87
334
ppm after a 37-week reaction with CO2. In addition, the Al concentration which appeared at
335
the 2nd week as 0.78 ppm was approximately doubled after 9 weeks and then reduced by up
336
to 0.62 ppm in the solution after reacting for 37 weeks. The release of Fe ions was very
337
similar to the Al ion concentration in the solution, possibly because of secondary mineral
338
precipitation such as aluminosilicate minerals or transitional products with Fe ions. 100000
Concentration Log(ppm)
10000 Si
1000
Al Fe
100
Ca 10
Mg K
1
Na
0.1 0
10
20
30
40
Reaction time (weeks)
339 340
Figure 8: Changes in ion concentrations of the solution with reaction time
341
The predicted model and experimental results for pH and alkalinity values are depicted in Figure
342
9. In the presence of CO2, alkalinity continues to increase throughout the reaction time, such that
343
the alkalinity at the 37th week was around 5 times greater than that at the 2nd week. There is a 20
344
good relationship between the model and experimental alkalinity values (relative error 2%). On
345
the other hand, the pH value of the solution at the beginning was around 4.0 and it gradually
346
increased up to 5.0, which can be considered as a 25% increment. Although the model result
347
deviated from the experimental values (relative error 24 %), both show an increment with time
348
due to the consumption of H+ in other chemical reactions. 6
4
3 4
pH
2.5
3
pH using model
2
2
pH
1.5 1
Alkalinity (meq/l)
1
Alkalinity (meq/l)
3.5
5
0.5 Alkalinity (meq/l) using model 0
0 0
5
10
15
20
25
30
35
40
Weeks
349 350
Figure 9: Laboratory and model results for pH and alkalinity measurements as a
351
function of reaction time
352
3.2 Reaction of minerals in the mudstone sample
353
The mineralogical composition of intact mudstone obtained from XRD results shows that the
354
major minerals are quartz, illite, orthoclase, plagioclase and chlorite. In addition, goethite,
355
pyrite, montmorillonite and kaolinite are present in minor percentages (Vatankhah). Figure
356
10 provides the mineral composition variation of intact and reacted mudstone over 37 weeks.
357
Accordingly, the mineral composition of reacted mudstone changed, but no significant
358
mineral variation could be observed in quartz, illite, pyrite, albite and montmorillonite. All
359
the other minerals decreased with time, with the exception of kaolinite. Kaolinite dissolved in
360
the first 5 weeks and then its content almost doubled at the end of the 37th week compared to
361
the intact sample. However, the XRD results were obtained using semi-quantitative analysis
362
with a ±2% error. In addition, SEM and EDX results are shown in Figure 11. Using the EDX
363
results, the chemical characterization of the sample was identified, and the minerals present
364
were verified.
21
Mineral Composition (%)
35 30 25 20 15
decrease
Intact
increase
2 weeks
decrease
10
5 weeks
5
9 weeks
0
18 weeks 37 weeks
Minerals
365
Figure 10: XRD results of intact and reacted mudstone after reactivity tests a).
b).
22
366
Figure 11a) SEM results of Intact, and b) two-week reacted mudstone samples c).
d).
367
Figure 11c) SEM results of five-week, and d) nine-week reacted mudstone samples
23
e).
f).
368
Figure 11e) SEM results of 18-week, and f) 37-week reacted mudstone samples
369
3.3 Geochemical modelling
370
3.3.1
Calculation of saturation indices
371
In order to find the dissolution, precipitation and equilibrium state of minerals in the reacted
372
samples, saturation indices (SIs) were calculated using the brine compositions obtained from
373
the ICP-OES tests. The predicted SIs of the PHREEQC model results are shown in Table 5.
374
According to the results, the mudstone sample reacted with CO2 was under-saturated with
375
respect to K-feldspar, illite, chlorite, montmorillonite, anorthite and albite for 37 weeks while
24
376
the solution was saturated with respect to gibbsite and kaolinite after 9 weeks reaction with
377
CO2. However, chlorite, anorthite and illite have higher saturation indices compared to other
378
minerals, which suggests that their dissolution rates are higher. Since the SIs of goethite and
379
quartz are very close to zero, these minerals are stable (in equilibrium) in the presence of
380
weak carbonic acid at low temperatures. In addition, gibbsite mineral appeared in the latter
381
part of the test series.
382
Table 5: Saturation indices for major minerals involved in chemical reactions Mineral
383
2 weeks
5 weeks
9 weeks
18 weeks
37 weeks
Quartz
0.02
-0.25
-0.01
-0.39
-0.42
Goethite
2.24
0.83
0.27
-1.52
-0.9
K-Feldspar
-7.19
-7.3
-2.59
-1.57
-1.51
Illite
-12.21
-11.86
-2.95
-0.17
-0.17
Kaolinite
-6.39
-6.05
0.52
2.7
2.6
Chlorite
-28.33
-27.37
-20.91
-17.84
-11.97
Montmorillonite
-9.5
-9.61
-2.58
-1.02
-1.08
Anorthite
-21.37
-20.65
-12.1
-8.7
-8.57
Albite
-7.5
-7.73
-3.01
-2.05
-2.14
Gibbsite
-
-
-2.73
1.13
1.33
3.3.2
Batch geo-chemical modelling
384
The element composition of the solution with reaction time was estimated using batch geo-
385
chemical (kinetic reaction) model, as shown in Figure 12. There is a good relationship
386
between model and laboratory results except for Al and Fe ions which do not match with the
387
measured concentration patterns over reaction time. In the model results of the solution
388
concentration, the changes occur rapidly in the first few weeks and afterwards, the ion
389
releasing rate is very small.
25
Al modelled
100,000
Fe modelled Si modelled
Concentration Log (ppm)
10,000
Mg modelled K modelled
1,000
Ca modelled Na modelled
100
Al experimental Fe experimental
10
Si experimental Mg experimental
1 0 0
10
20
30
40
K experimental Ca experimental
Reaction time (weeks)
Na experimental
390 391
Figure 12: The comparison between experimental and modelled results of ion
392
concentrations in CO2-mudstone-brine reacted solution.
393 394
4. Discussion 4.1 Dissolution of primary minerals in caprock
395
The results of reactivity tests conducted on caprock samples obtained from the South-west
396
Hub geo-sequestration project provide a qualitative understanding of caprock-brine-CO2
397
interactions in deep saline aquifers. Since the reaction temperature is quite low (40 ºC), rapid
398
geo-chemical reactions were not observed in the test series. According to the results, almost
399
all the mineral content is reduced in the presence of carbonic acid due to dissolution, with the
400
exception of kaolinite. Kaolinite content was increased after 5-week reaction due to
401
kaolinization (Knut, 1984). The change in clay mineral content was less than that of chlorite,
402
k-feldspar and anorthite mineral dissolution. Quartz was stable and did not show any change
403
with ScCO2 at 40 °C. In addition, kaolinite content increased compared with the original
404
amount and new minerals precipitated with elements of Si, Al and Fe. Therefore, Fe, Al and
405
Si ion concentration in the solution reduced with reaction time, as shown by the ICP-OES
406
results (Figure 8). Thus, it is important to investigate the net pore volume change due to geo-
407
chemical reactions because high porosity can increase the back migration of injected CO2,
408
which is a negative point for CO2 sequestration.
26
409
CO2 tends to dissolve in brine forming carbonic acid, which is considered to be a weak acid
410
(Ni et al., 2014) and it easily decomposes into bicarbonate ions, as shown in Eq. 5 (Leonenko
411
and Keith, 2008). Due to the rapid release of H+ ions into the solution at the beginning, the
412
initial pH was around 4.0, and it increased with time due to the exchange of H+ ions with
413
caprock minerals such as K-feldspar, goethite, anorthite and chlorite. In addition, the
414
alkalinity which measures the acid-buffering capacity increased rapidly during the reaction
415
period due to the high HCO3- concentration in the solution. According to Reuss et al. (1987),
416
the alkalinity of the solution does not increase because of the HCO3- ions until an exchange
417
reaction or silicate weathering consumes the H+ ions. Thus, we can conclude that in these
418
reaction series, such kind of exchange reactions have taken place as alkalinity increased with
419
time.
420
0 ;<+(:=) + ?+ <(@) ↔?+ ;
(5)
421
Diffusion and advection are the main mechanisms used by injected CO2 to saturate the
422
caprock in deep saline aquifers. Therefore, the caprock-CO2 reactions are considered as rock-
423
dominated reactions where the fluid flow is reduced (Kaszuba et al., 2013). Of the minerals
424
present in the selected caprock, the chlorite mineral is more reactive in the presence of
425
carbonic acid. The dissolution rate of chlorite at 40 °C and pH 4 is around 3.34 × 10-11
426
mol/m-2sec-1 (Black et al., 2015). Chlorite dissolution in the presence of CO2 causes
427
continuous release of Al, Mg and Fe, as shown in Eq 6, verifying the ICP-OES, XRD and
428
PHREEQC model results. Therefore, chloride is more reactive than clay minerals and has
429
mineral reactivity ranging from 10-14 to 10-10 mol/m-2sec-1 (Du et al., 2018). In addition, the
430
mineral reactivity of anorthite and albite at 40 °C and pH 4 are approximately 1.38×10-8 and
431
3.51×10-12 mol/m-2sec-1, respectively (Black et al., 2015; Oelkers and Schott, 1995; Palandri
432
and Kharaka., 2004). Therefore, anorthite dissolves (Eq. 7) faster than albite (Eq 8) releasing
433
Ca, Si and Al, but albite dissolves comparatively more rapidly than quartz (Welch and
434
Ullman, 1996), and the dissolution rate is approximately 5.45 × 10-14 mol/m-2sec-1 in the
435
solid quartz phase (Black et al., 2015; Gherardi et al., 2007). According to the ICP-OES
436
results, the Ca2+ ion concentration increased with time, implying continuous anorthite
437
dissolution throughout the reaction time. The dissolution of anorthite is also consistent with
438
the XRD and PHREEQC model results. Moreover, Amrhein and Suarez (1992) conducted
439
experiments to investigate the long-term (4.5 years) reactivity of anorthite under controlled
440
CO2 conditions and found that the final reaction rates are 200 times slower than the short-
27
441
term dissolution. Therefore, it is crucial to identify the long-term reaction rates in order to
442
predict caprock reactivity over thousands of years.
443
Moreover, according to the PHREEQC model results, illite dissolves slowly, and the mineral
444
reactivity is around 9.68 × 10-13 mol/m-2sec-1 at 40 °C and pH 4 (Black et al., 2015; Köhler et
445
al., 2003) as shown in Eq. 9, releasing K+ and Al3+ ions. Although such a mineral content
446
change was not observed in the XRD results, the percentage variation was so low that it could
447
be considered as a quantification error. Similar patterns were observed for montmorillonite
448
and goethite minerals. According to the XRD results, goethite dissolves with the presence of
449
H+ ions in the solution, which can be described as shown in Eq. 10 (Cornell et al., 1976).
450
Therefore, it can be considered as one of the sources of Fe3+ ions in the solution. However,
451
there is no compatibility between the model results and the XRD results in a few cases. The
452
geo-chemical model results can vary from the actual results due to limitations in the
453
PHREEQC model, such as treating CO2 as an ideal gas and non-reliability of kinetic rate
454
constants. Hence, it is essential to implement the model with new modifications such as
455
introducing a CO2 solubility model and building the model with the most accurate kinetic rate
456
constants.
457
[I A BCD.6 EF.G HF.4 HF.4 + 0.1H + + A., EF., ]<4F ()7 + 2? → 4.9BC + 1.2E
458
0.1H A + 3.5?D I
(6)
459
;E+ I+ <7 + 8? → ;+ + 2E A + 2I<+(:=) + 4?+ <
(7)
460
QEIA <7 + ? + + ?+ < → Q + 2?D [I
(8)
461
)F.G, E4.G, [IA., EF., <4F ]()+ + 14? → 0.75) 2.25E A + 3.5?D I
(9)
462
H< + 3? → H A + ?+ <
(10)
463
The brine solutions obtained at different time intervals were under-saturated with respect to
464
K-feldspar (from the 2nd week to the 37th week), or in other words, k-feldspar tended to
465
dissolve in the presence of carbonic acid at 40 °C. Moreover, according to the XRD test
466
results, the reacted solid minerals had a much lower percentage of K-feldspar at the end of
467
the reaction period. According to Du et al. (2018), such reactions can make micro-fractures
468
and round, oval and square dissolution pores in the mineral surfaces due to the pressure and
469
weak carbonic acid. As a result, the porosity and permeability reduce with time increasing
470
CO2 leakage risk which is a disadvantage for CO2 sequestration.
6
28
471
4.2 Precipitation of secondary minerals
472
Generally, the dissolution/precipitation of kaolinite can be studied by observing Si and Al ion
473
concentrations in the solution and usually the release of Si ions is greater than Al ions, as
474
shown in Figure 8 (Polzer and Hem, 1965; Wieland and Stumm, 1992; Xiong et al., 2009).
475
According to the ICP-OES results, the Al and Si ion concentrations increased up to the 9th
476
week due to the dissolution of kaolinite, as given in Eq. 11 (Polzer and Hem, 1965; Wieland
477
and Stumm, 1992) and later, the solution became supersaturated with respect to kaolinite.
478
This is supported by the results of the geo-chemical model. According to Nagy et al. (1991),
479
the kinetic rates of kaolinite during dissolution and precipitation at pH 3 and 40 °C are -2.02
480
× 10-11 and 2.29× 10-12 mol/m-2sec-1. Since kaolinite already existed in the caprock sample,
481
new formation of kaolinite could not be observed very clearly in the SEM images. However,
482
the kaolinite percentage almost doubled compared with the intact sample in the XRD test
483
results, possibly because of the dissolution of K-feldspar and the precipitation of kaolinite on
484
the caprock sample, as shown in Eq. 12. Therefore, at low temperatures in deep saline
485
aquifers, K-feldspar acts as one of the reactive minerals among silicate minerals which can be
486
subjected to kaolinization. In addition, according to the ICP-OES test results, K+ ion
487
concentration increases during the reaction time, because K-feldspar releases K+ ions and
488
forms kaolinite. Therefore, the saturation indices calculated by PHREEQC are compatible
489
with the mineral changes found in the reacted solid minerals. The same phenomenon was
490
observed in several other studies too (Wollast, 1967; Yuan et al., 2019). A number of studies
491
have been conducted to investigate the alteration of K-feldspar at elevated temperatures and
492
pressures (hydro-thermal reactions), but the behaviour of K-feldspar at low temperatures and
493
pressures is different from that at high temperatures and pressures because the rates and the
494
products are totally different in the reactions. According to Simpson et al. (1979) and Taylor
495
H. (1979), kaolinization, chloritization and hematization occur at low temperatures (<200 °C);
496
thus these kind of reactions cannot be observed at very high temperatures (>200 °C).
497
According to Yuan et al. (2019), when temperature and pressure are greater than 100 °C and
498
300 bars respectively, k-feldspar and kaolinite react and form illite and quartz instead of
499
kaolinization. Therefore, Wollast (1967) carried out experiments at room temperature and
500
pressure to identify K-feldspar behaviour and found that the release of Si and Al to the
501
solution and the exchange of H+ for K+ are the dominant observations for K-feldspar in the
502
presence of carbonic acid. As a result, a high K+ ion release was observed in the ICP-OES
503
results as shown in Figure 8. 29
504
The dissolution of kaolinite in the first stages of the reaction increases the porosity of the
505
caprock, which would enhance the permeability and back-migration of injected CO2 (She et
506
al., 2016). This is a disadvantage for caprock integrity. However, the porosity and
507
permeability decrease due to the re-precipitation of kaolinite as a secondary mineral after 9th
508
week. These phenomena have been observed in different studies (Nagy et al., 1991; Nagy and
509
Lasaga, 1993; She et al., 2016), and it is crucial to recognize the net gain in porosity due to
510
dissolution and precipitation to identify the permeability of caprock. During kaolinization,
511
one volume of K-feldspar forms approximately 60% kaolinite and 40% silicic acid (Knut,
512
1984). However, significant kaolinite precipitation cannot be seen under these conditions
513
because kaolinite formation in solutions with high K+ concentration and lower H+
514
concentration at lower temperatures is very slow. Once the solution becomes supersaturated
515
with amorphous silica, H4SiO4 precipitates on the caprock, reducing its porosity (Wollast,
516
1967; Yoshiro et al., 1974) which is an advantage for CO2 sequestration.
517
E+ I+ <, ()D + 6? ↔ 2E A + 2?D I
518
2)EIA <7(UVWXUY@:YZ) + 2? + 9?+ < ↔ E+ I+ <, ()D([:U@\\WZ) + 4?D I
519
At the beginning, the pH value was low (around 4.0) due to ?+ ;
520
rapid release of Al, Si and Fe ions could be observed as a result of mineral dissolutions such
521
as kaolinite, k-feldspar, goethite and chlorite. However, Si, Fe and Al ion concentrations were
522
reduced (due to kaolinization and precipitation of H4SiO4) after 9th week. Another reason
523
may be the formation of Si(OH)4, gibbsite and Fe(OH)3 colloidal precipitations in the solution
524
(Wei and Taian, 1998). According to Ni et al. (2014), the dissolution rate of Si was maximum
525
at 45 ºC and Si(OH)4 was formed, reducing the Si concentration. In addition, the presence of
526
Al ions in the solution led to the precipitation of gibbsite, resulting in a decrease of Al ion
527
concentration in the solution. This result complies with the geo-chemical model. Therefore,
528
secondary precipitations of minerals in the clay matrix leads to porosity reduction, decreasing
529
the diffusion and reaction rate, which is a positive characteristic for caprock integrity in long-
530
term scenarios.
531
5.
532
The results of reactivity tests conducted on caprock samples obtained from the South-west
533
Hub geo-sequestration project provide a qualitative and quantitative understanding of
534
caprock-brine-CO2 interactions in deep saline aquifers. In this reactivity test series, in situ
(11)
Conclusion
30
535
conditions at a depth of 979 m (40 °C and 10 MPa) were provided for 37 weeks to recognize
536
possible field geo-chemical reactions in the caprock, rather than conducting hydro-thermal
537
experiments which give unreliable results due to the absence of such high pressures and
538
temperatures in the field. The following key findings were made:
539•
Injection of CO2 can cause a reduction of pH in brine up to 4.0. In the beginning, the ion
540
release rate is accelerated due to high H+ activity. However, with time, the H+ concentration
541
reduces because different chemical reactions consume it. As a result, the ion release rate
542
gradually decreases. Thus, the mineral changes in caprock become significant in geological
543
time scale so that, the long term caprock integrity should be evaluated.
544•
The caprock minerals react with injected CO2 causing precipitations and dissolutions
545
depending upon aquifer temperature and pressure. In the current study, considerable
546
dissolution was observed in chlorite, k-feldspar and anorthite minerals, in contrast minerals
547
such as quartz are not reactive in the presence of carbonic acid at low temperatures (40 °C).
548
In addition, the kaolinite content increases compared to the original amount and new minerals
549
precipitate with elements of Si, Al and Fe at aquifer temperature and pressure decreasing pore
550
volume of caprock. Thus, net pore volume change should be evaluated after the reaction time
551
period to identify whether the caprock is at risk or not.
552•
In the beginning, the dissolution of minerals is very significant, and this enhances the pore
553
volume of the caprock, and eventually increases the porosity, diffusion and permeability.
554
This is a negative point for caprock integrity as it enhances the back- migration of injected
555
CO2. However, due to the precipitation of secondary minerals in the rock pore structure such
556
as kaolinite, the pore volume can be reduced to some extent, which can be considered as a
557
benefit for caprock in long-term scenarios. Thus, the net change of porosity due to dissolution
558
and precipitation of caprock is crucial to predict the ability of CO2 to flow through caprock.
559•
Due to high heterogeneity nature of caprock mineralogy, the geo-chemical reactions differ
560
from site to site, and it is a challenge to predict the chemical reactivity of caprock in different
561
aquifers. The geo-chemical reactions occurring in the caprock with the presence of CO2 is
562
site-specific and it is essential to investigate for the long-term integrity of CO2 sequestration
563
process. Therefore, the prediction of reactivity incorporating heterogeneity of caprock using a
564
conceptual geological model is highly recommended as future work.
31
565
6. Author contribution
566
Jayasekara: Investigation, Writing - Original draft, Software. Ranjith: Conceptualization.
567
Wanniarachchi: Methodology. Rathnaweera: Writing - Review and Editing. Van Gent:
568
Resources.
569
7. Declarations of interest
570
The authors declare no conflicts of interest.
571
8. Acknowledgement
572
Authors would like to acknowledge the West Australian Department of Mines, Industry
573
Regulation and Safety for the provision of the sample.
574
9. References
575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603
Alemu, B.L., Aagaard, P., Munz, I.A. and Skurtveit, E., 2011. Caprock interaction with CO2: A laboratory study of reactivity of shale with supercritical CO2 and brine. Applied Geochemistry, 26(12): 1975-1989. Amrhein, C. and Suarez, D.L., 1992. Some factors affecting the dissolution kinetics of anorthite at 25°C. Geochimica et Cosmochimica Acta, 56(5): 1815-1826. Anabaraonye, B.U., Crawshaw, J.P. and Trusler, J.P.M., 2019. Brine chemistry effects in calcite dissolution kinetics at reservoir conditions. Chemical Geology, 509: 92-102. Backhouse, J. and Crostella, A., 2000. Geology and Petroleum Exploration of the Central and Southern Perth Basin, Western Australia. Commonwealth of Australia (Geoscience Australia). Bennion, D.B. and Bachu, S., 2007. Permeability and Relative Permeability Measurements at Reservoir Conditions for CO2-Water Systems in Ultra Low Permeability Confining Caprocks, EUROPEC/EAGE Conference and Exhibition. Society of Petroleum Engineers, London, U.K., pp. 9. Bjørlykke, K., Aagaard, P., Houseknecht, D.W. and Pittman, E.D., 1992. Clay Minerals in North Sea Sandstones, Origin, Diagenesis, and Petrophysics of Clay Minerals in Sandstones. SEPM Society for Sedimentary Geology, pp. 0. Black, J.R., Carroll, S.A. and Haese, R.R., 2015. Rates of mineral dissolution under CO2 storage conditions. Chemical Geology, 399: 134-144. Cornell, R.M., Posner, A.M. and Quirk, J.P., 1976. Kinetics and mechanisms of the acid dissolution of goethite (α-FeOOH). Journal of Inorganic and Nuclear Chemistry, 38(3): 563-567. Credoz, A. et al., 2009. Experimental and modeling study of geochemical reactivity between clayey caprocks and CO2 in geological storage conditions. Energy Procedia, 1(1): 3445-3452. Dávila, G., Cama, J., Luquot, L., Soler, J.M. and Ayora, C., 2017. Experimental and modeling study of the interaction between a crushed marl caprock and CO2-rich solutions under different pressure and temperature conditions. Chemical Geology, 448: 26-42.
32
604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 629 630 631 632 633 634 635 636 637 638 639 640 641 642 643 644 645 646 647 648 649 650 651 652 653
De Silva, G.P.D., Ranjith, P.G. and Perera, M.S.A., 2015. Geochemical aspects of CO2 sequestration in deep saline aquifers: A review. Fuel, 155: 128-143. Du, Y. et al., 2018. Experimental Study of the Reactions of Supercritical CO2 and Minerals in High-Rank Coal under Formation Conditions. Energy & Fuels, 32(2): 1115-1125. Ellis, B. et al., 2011. Deterioration of a fractured carbonate caprock exposed to CO2-acidified brine flow. Greenhouse Gases: Science and Technology, 1(3): 248-260. Fu, Q. et al., 2009. Coupled alkali-feldspar dissolution and secondary mineral precipitation in batch systems: 1. New experiments at 200 °C and 300 bars. Chemical Geology, 258(3): 125-135. Gaus, I., Azaroual, M. and Czernichowski-Lauriol, I., 2005. Reactive transport modelling of the impact of CO2 injection on the clayey cap rock at Sleipner (North Sea). Chemical Geology, 217(3-4): 319-337. Gherardi, F., Xu, T. and Pruess, K., 2007. Numerical modeling of self-limiting and selfenhancing caprock alteration induced by CO2 storage in a depleted gas reservoir. Chemical Geology, 244(1): 103-129. Helgeson, H.C., Murphy, W.M. and Aagaard, P., 1984. Thermodynamic and kinetic constraints on reaction rates among minerals and aqueous solutions. II. Rate constants, effective surface area, and the hydrolysis of feldspar. Geochimica et Cosmochimica Acta, 48(12): 2405-2432. Ilgen, A.G. et al., 2018. Shale-brine-CO2 interactions and the long-term stability of carbonate-rich shale caprock. International Journal of Greenhouse Gas Control, 78: 244-253. Kaszuba, J., Yardley, B. and Andreani, M., 2013. Experimental perspectives of mineral dissolution and precipitation due to carbon dioxide-water-rock interactions. Reviews in Mineralogy and Geochemistry, 77(1): 153-188. Kaszuba, J.P., Janecky, D.R. and Snow, M.G., 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): 1065-1080. Kaszuba, J.P., Janecky, D.R. and Snow, M.G., 2005. Experimental evaluation of mixed fluid reactions between supercritical carbon dioxide and NaCl brine: Relevance to the integrity of a geologic carbon repository. Chemical Geology, 217(3-4): 277-293. Kharaka, Y.K. et al., 2009. Potential environmental issues of CO2 storage in deep saline aquifers: Geochemical results from the Frio-I Brine Pilot test, Texas, USA. Applied Geochemistry, 24(6): 1106-1112. Kim, K.H., Akase, Z., Suzuki, T. and Shindo, D., 2010. Charging Effects on SEM/SIM Contrast of Metal/Insulator System in Various Metallic Coating Conditions. MATERIALS TRANSACTIONS, 51(6): 1080-1083. Kloprogge, J.T., 2017. Chapter 8 - Application of Vibrational Spectroscopy in Clay Minerals Synthesis. In: W.P. Gates, J.T. Kloprogge, J. Madejová and F. Bergaya (Editors), Developments in Clay Science. Elsevier, pp. 222-287. Knut, B., 1984. Clastic Diagenesis. Formation of Secondary Porosity: How Important Is It?: Part 2. Aspects of Porosity Modification, University of Bergen, Norway. Kohler, E., Parra, T. and Vidal, O., 2009. Clayey cap-rock behavior in H2O-CO2 media at low pressure and temperature conditions: an experimental approach. Clays and Clay Minerals, 57(5): 616-637. Köhler, S.J., Dufaud, F. and Oelkers, E.H., 2003. An experimental study of illite dissolution kinetics as a function of ph from 1.4 to 12.4 and temperature from 5 to 50°C. Geochimica et Cosmochimica Acta, 67(19): 3583-3594. Kweon, H. and Deo, M., 2017. The impact of reactive surface area on brine-rock-carbon dioxide reactions in CO2 sequestration. Fuel, 188: 39-49. 33
654 655 656 657 658 659 660 661 662 663 664 665 666 667 668 669 670 671 672 673 674 675 676 677 678 679 680 681 682 683 684 685 686 687 688 689 690 691 692 693 694 695 696 697 698 699 700 701 702
Lasaga, A.C., 1984. Chemical kinetics of water-rock interactions. Journal of Geophysical Research: Solid Earth, 89(B6): 4009-4025. Leonenko, Y. and Keith, D.W., 2008. Reservoir Engineering To Accelerate the Dissolution of CO2 Stored in Aquifers. Environmental Science & Technology, 42(8): 2742-2747. Liu, F. et al., 2012. CO2–brine–caprock interaction: Reactivity experiments on Eau Claire shale and a review of relevant literature. International Journal of Greenhouse Gas Control, 7: 153-167. Nagy, K.L., Blum, A.E. and Lasaga, A.C., 1991. Dissolution and precipitation kinetics of kaolinite at 80°C and pH 3: the dependence on solution saturation state. American Journal of science, 291: 649-686. Nagy, K.L. and Lasaga, A.C., 1993. Simultaneous precipitation kinetics of kaolinite and gibbsite at 80°C and pH 3. Geochimica et Cosmochimica Acta, 57(17): 4329-4335. Nguyen, M.C. et al., 2017. An object-based modeling and sensitivity analysis study in support of CO2 storage in deep saline aquifers at the Shenhua site, Ordos Basin. Geomechanics and Geophysics for Geo-Energy and Geo-Resources, 3(3): 293-314. Ni, X., Li, Q. and Chen, W., 2014. Dissolution kinetics of Si and Al from montmorillonite in carbonic acid solution. International Journal of Coal Science & Technology, 1(1): 3138. ODIN_Reservoir_Consultants, 2015. Petrophysical Interpretation of the Harvey Wells. Oelkers, E.H. and Schott, J., 1995. Experimental study of anorthite dissolution and the relative mechanism of feldspar hydrolysis. Geochimica et Cosmochimica Acta, 59(24): 5039-5053. Olierook, H.K.H. et al., 2014. Facies-based rock properties characterization for CO2 sequestration: GSWA Harvey 1 well, Western Australia. Marine and Petroleum Geology, 50: 83-102. Palandri, J.L. and Kharaka., Y.K., 2004. A compilation of rate parameters of water-mineral interaction kinetics for application to geochemical modeling. US Geol. Surv. Open file Rep.: 1-64. Parkhurst, D.L., 1995. User's guide to PHREEQC, a computer program for speciation, reaction-path, advective-transport, and inverse geochemical calculations. 95-4227. Pearce, J.K. et al., 2019. A combined geochemical and µCT study on the CO2 reactivity of Surat Basin reservoir and cap-rock cores: Porosity changes, mineral dissolution and fines migration. International Journal of Greenhouse Gas Control, 80: 10-24. Polzer, W.L. and Hem, J.D., 1965. The dissolution of kaolinite. Journal of Geophysical Research, 70(24): 6233-6240. Rathnaweera, T.D., Ranjith, P.G. and Perera, M.S.A., 2014. Salinity-dependent strength and stress–strain characteristics of reservoir rocks in deep saline aquifers: An experimental study. Fuel, 122: 1-11. Reuss, J.O., Cosby, B.J. and Wright, R.F., 1987. Chemical processes governing soil and water acidification. Nature, 329(6134): 27-32. Shainberg, I. and Levy, G.J., 2005. FLOCCULATION AND DISPERSION. In: D. Hillel (Editor), Encyclopedia of Soils in the Environment. Elsevier, Oxford, pp. 27-34. Shao, H., Ray, J.R. and Jun, Y.-S., 2010. Dissolution and Precipitation of Clay Minerals under Geologic CO2 Sequestration Conditions: CO2−Brine−Phlogopite Interactions. Environmental Science & Technology, 44(15): 5999-6005. She, M. et al., 2016. Experimental simulation of dissolution law and porosity evolution of carbonate rock. Petroleum Exploration and Development, 43(4): 616-625. Shukla, R., Ranjith, P., Haque, A. and Choi, X., 2010. A review of studies on CO2 sequestration and caprock integrity. Fuel, 89(10): 2651-2664.
34
703 704 705 706 707 708 709 710 711 712 713 714 715 716 717 718 719 720 721 722 723 724 725 726 727 728 729 730 731 732 733 734 735 736 737 738 739 740 741 742 743 744 745 746
Simpson, P.R. et al., 1979. Uranium mineralization and granite magmatism in the British Isles. Philosophical Transactions of the Royal Society of London. Series A, Mathematical and Physical Sciences, 291(1381): 385-412. Song, J. and Zhang, D., 2013. Comprehensive Review of Caprock-Sealing Mechanisms for Geologic Carbon Sequestration. Environmental Science & Technology, 47(1): 9-22. Stalker, L., Varma, S., Van Gent, D., Haworth, J. and Sharma, S., 2013. South West Hub: a carbon capture and storage project AU - Stalker, L. Australian Journal of Earth Sciences, 60(1): 45-58. Stephansson, O., Hudson, J.A. and Jing, L., 2004. Coupled Thermo-Hydro-MechanicalChemical Processes in Geo-Systems - Fundamentals, Modelling, Experiments and Applications. Elsevier. Tarkowski, R., Wdowin, M. and Manecki, M., 2015. Petrophysical examination of CO brine-rock interactions-results of the first stage of long-term experiments in the potential Zaosie Anticline reservoir (central Poland) for CO storage. Environ Monit Assess, 187(1): 4215-4215. Taylor H., P., 1979. Oxygen and hydrogen isotope relationships in hydrothermal mineral deposits. Geochemistry of hydrothermal ore deposits, New York. Vatankhah, A.R., 2011. Approximate Solutions to Complete Elliptic Integrals for Practical Use in Water Engineering. Journal of Hydrologic Engineering, 16(11): 942-945. Wdowin, M., Tarkowski, R. and Franus, W., 2014. Determination of changes in the reservoir and cap rocks of the Chabowo Anticline caused by CO2–brine–rock interactions. International Journal of Coal Geology, 130: 79-88. Wei, L. and Taian, Y.L., 1998. An experimental study on reaction of montmorillonite with mud acid. Oilfield Chem(15(3)): 237-240. Welch, S.A. and Ullman, W.J., 1996. Feldspar dissolution in acidic and organic solutions: Compositional and pH dependence of dissolution rate. Geochimica et Cosmochimica Acta, 60(16): 2939-2948. Wieland, E. and Stumm, W., 1992. Dissolution kinetics of kaolinite in acidic aqueous solutions at 25°C. Geochimica et Cosmochimica Acta, 56(9): 3339-3355. Wollast, R., 1967. Kinetics of the alteration of K-feldspar in buffered solutions at low temperature. Geochimica et Cosmochimica Acta, 31(4): 635-648. Xiao, T., Dai, Z., McPherson, B., Viswanathan, H. and Jia, W., 2017. Reactive transport modeling of arsenic mobilization in shallow groundwater: impacts of CO2 and brine leakage. Geomechanics and Geophysics for Geo-Energy and Geo-Resources, 3(3): 339-350. Xiong, S.C., Liu, W.D., Gong, Y.G. and Xiao, H.M., 2009. Study on reaction between montmorillonite and alkaline flooding agent. J Liaoning Tech Univ 28: 73-75. Xu, T., Apps, J.A. and Pruess, K., 2005. Mineral sequestration of carbon dioxide in a sandstone–shale system. Chemical Geology, 217(3): 295-318. Yoshiro, T., Shinjiro, M., Hiroshi, S. and Hiroshi, H., 1974. Kinetics of alteration of Kfeldspar and its application to the alteration zoning. Geochemical Journal, 8: 1-20. Yuan, G. et al., 2019. A review of feldspar alteration and its geological significance in sedimentary basins: From shallow aquifers to deep hydrocarbon reservoirs. EarthScience Reviews, 191: 114-140.
747
35
Highlights •
Chemical reactivity tests conducted for mudstone providing reservoir conditions
•
Conducted mineral-brine analysis for 37 weeks including geo-chemical modelling
•
Dissolution of minerals such as K-feldspar, anorthite, and chlorite
•
Precipitation of minerals such as kaolinite, gibbsite, amorphous silica and Fe(OH)3
•
Dissolution of minerals is significant which can increase pore volume of caprock
Conflict of Interests We declare that there is no conflict of interests.