Next Article in Journal
Recovery of Lithium from Industrial Li-Containing Wastewater Using Fluidized-Bed Homogeneous Granulation Technology
Previous Article in Journal
Numerical Modeling of Electron Beam Cold Hearth Melting for the Cold Hearth
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Laboratory Experiments and Geochemical Modeling of Gas–Water–Rock Interactions for a CO2 Storage Pilot Project in a Carbonate Reservoir in the Czech Republic

1
Faculty of Mining and Geology, VSB-Technical University of Ostrava, 17. Listopadu 2172/15, Poruba, 70800 Ostrava, Czech Republic
2
Faculty of Mining, Safety Engineering and Industrial Automation, Silesian University of Technology, 2 Akademicka St., 44-100 Gliwice, Poland
*
Authors to whom correspondence should be addressed.
Minerals 2024, 14(6), 602; https://doi.org/10.3390/min14060602
Submission received: 5 April 2024 / Revised: 4 June 2024 / Accepted: 6 June 2024 / Published: 8 June 2024
(This article belongs to the Section Mineral Geochemistry and Geochronology)

Abstract

:
The aim of this study was to characterize the influence of CO2 in geological structures on mineralogical changes in rocks and assess the sequestration capacity in mineral form and solution of a potential pilot storage site in the Czech Republic. Rock samples from a dolomite reservoir and the overburden level, as well as the corresponding pore water, were used. The most important chemical process occurring in the reservoir rock is the dissolution of carbonate minerals and feldspars during the injection of CO2 into the structure, which increases the porosity of the structure by approximately 0.25 percentage points and affects the sequestration capacity of the reservoir rock. According to the results of geochemical modeling, the secondary carbonate minerals (dolomite, siderite, and ephemeral dawsonite) were present only during the first 50 years of storage, and the porosity at this stage decreased by 1.20 pp. In the caprocks, the decomposition of K-feldspar and calcite resulted in an increase in porosity by 0.15 percentage points at the injection stage only, while no changes in porosity were noted during storage. This suggests that their insulation efficiency can be maintained during the injection and post-injection periods. However, further experimental research is needed to support this observation. The results of this study indicate that the analyzed formation has a low potential for CO2 sequestration in mineral form and solution over 10,000 years of storage, amounting to 5.50 kg CO2/m3 for reservoir rocks (4.37 kg CO2/m3 in mineral form and 1.13 kg CO2/m3 in dissolved form) and 3.22 kg CO2/m3 for caprock rocks (3.01 kg CO2/m3 in mineral form and 0.21 kg CO2/m3 in dissolved form). These values are lower than in the case of the depleted Brodské oil field, which is a porous reservoir located in the Moravian part of the Vienna Basin.

1. Introduction

One of the current major environmental and climatic issues is the need to reduce atmospheric CO2 concentrations. One method to achieve this aim is carbon capture and storage (CCS), which represents a technology based on CO2 injection and its safe underground storage. In 2023, the global average atmospheric carbon dioxide was 419.3 ppm, which set a new record high. The increase between 2022 and 2023 was 2.8 ppm, which marked the 12th year in a row where the amount of carbon dioxide in the atmosphere increased by more than 2 ppm [1]. Deep geological storage in porous rock formations is considered the most appropriate strategy for CO2 sequestration [2,3,4] and the injectivity of storage rocks is a key technical and economic issue for CCS projects [5]. The viability of CO2 injection depends mainly on the porosity and permeability of the storage rocks. Further consequences of the CO2 interactions with host rock such as dissolution–precipitation of minerals are also important [6,7].
Geological storage of CO2 is possible through several physicochemical mechanisms, one of which is mineral trapping. These processes evolve over time after CO2 injection and are dominated by structural, stratigraphic, or hydrodynamic capture early in the project. These mechanisms are primarily governed by the following physical processes: fluid flow in the liquid and gas phase under pressure and gravity, capillary pressure effects, and heat flow via conduction, convection, and diffusion. The transport of aqueous and gaseous substances via advection and molecular diffusion is considered in both the liquid and gaseous phases. After the CO2 injection is complete, a number of capture mechanisms become important. CO2 can be partially trapped through residual capture as the plume moves away from the wellbore. The gas also mixes with and dissolves in the formation water at the leading and trailing edges of the plume (solubility trapping). The dissociation of CO2 dissolved in the formation water creates acidity that reacts with the minerals in the formation and can dissolve fast-reacting carbonate minerals (if present) in the acidified zone surrounding the injection well, leading to an increase in dissolved bicarbonate (called entrapment).
Theoretical, modeling, and experimental studies have been carried out to investigate CO2–water interactions from rock formations under deep storage conditions. Most of these experimental studies were designed to simulate the injection of CO2 mixed with formation water into the rocks at p–T conditions of a deep storage environment. The results of many of these experiments indicate an increase in the porosity/permeability of the storage rock caused by the partial dissolution of the carbonate minerals (mainly calcite) [4,8,9,10,11]. On the other hand, another set of experiments showed that porosity decreases due to initial carbonate dissolution followed by secondary precipitation/mineralization with porosity changes [10,12,13,14,15,16].
The essence of the research activity described in this article was to carry out model studies and laboratory experiments for the implementation of a pilot CO2 storage project—the first in Central and Eastern Europe—in a mature, depleting, hydrocarbon reservoir in the southeastern part of the Czechia.
The Žarošice field was discovered in 2001. The field is currently considered “mature”, with a declining production trend, and consideration is therefore given to further use of the geological structure, which forms the current hydrocarbon field, to implement a pilot project of geological storage of carbon dioxide.
Research methods included data collection and quality control from core samples, water samples taken from the deposit, laboratory tests and experiments, computer modeling, and numerical synthesis of results. In the experimental stage, a high-pressure reactor was used to test the interactions of reservoir and caprocks from selected core samples with formation water and CO2. The samples cover key types of fluid and rock chemical compositions. Geochemical modeling was performed using The Geochemist’s Workbench 11 software, in particular, the React module. The results of the reaction modeling were compared with experimental data from the tests in the reaction chamber. Comparing the results of both procedures is important to ensure the use of the most appropriate methods for predicting the evolution of the environment of a potential CO2 storage site.

2. Methodologies and the Initial State Study

Models involving kinetic transport through porous media and thermodynamic issues of multiphase systems are very useful in forecasting the impact of CO2 storage on the changes in the rock environment at the storage site. These models are based on information on the petrophysical, mineralogical, and hydrogeological properties of porous media, hydrochemical analysis of fluid composition, pressure and temperature values, and kinetic parameters.

2.1. Geological Setting

The area of interest, representing the surroundings of the structure referred to as “Zar-3” (Žarošice hydrocarbon field), covering an area of 94.5 km2, is located in the southeastern part of the Czech Republic. The wider surroundings of the Žarošice oil and gas field include the contact zone between the Western European foreland plate (Western European Platform) and the Western Carpathian thrust belt in the territory of southern Moravia (south-eastern part of the Czech Republic) [17].
This work is focused on the study of rocks whose geological development in the studied locality began in the Mesozoic period.
The Mesozoic to Cenozoic Alpine cycle in southern Moravia began with continental rifting in the Early Jurassic and deposition of synrift detrital deposits, followed by marine transgression and deposition of mixed carbonate and siliciclastic rocks of the passive margins. The carbonate platforms occupied the shallow marginal zone and the uplifted rift blocks, whereas the organic-rich marly facies (Mikulov marls) formed in a deeper basinal environment. During the Late Cretaceous to early Paleogene, the European foreland in Moravia was uplifted and deeply eroded. Rivers cut the deep Nesvačilka and Vranovice valleys that turned into submarine canyons more than 1500 m deep, which were gradually filled with the Paleogene detrital organic-rich deposits [18,19,20].
In the early Late Cretaceous, the Carpathian passive margins were converted into a foreland-type synorogenic basin marked by deep-water turbiditic flysch sedimentation, followed by a late orogenic and post-orogenic shallow marine and continental molasse sedimentation. During the late phases of the Alpine orogeny, in the late Paleogene and early Miocene, the flysch sequences were deformed and thrust over the European foreland, including the inner zone of the Neogene foredeep [17]. The Carpathian thrust belt in Moravia is composed of Late Jurassic to early Miocene synorogenic sequences of the Flysch belt (Figure 1). The Neogene foredeep, except for the innermost zones adjacent to the front of the thrust belt, remained undeformed. Superimposed on the thrust belt is the Neogene Vienna Basin.
Žarosice field is a buried Jurassic patch reef, which was subsequently dolomitized and faulted and then partially eroded from the west. The seal is provided by two different caprocks, Jurassic Marl of Mikulov Fm. in the eastern part of the field and Paleogene claystone in the western and northern parts described as Nesvačilka Fm. The reservoir rock mainly comprises heavily fractured dolomite of the Vranovice Fm. and, in the southern part near the base, dolomitic sandstones of the Nikolčice Fm. Beds (Figure 1), both of Jurassic age. The primary porosity of the dolomite is around 1–3%. Much more important is the secondary porosity, which is based on log data reaching 2–9% (locally, even more). The permeability of the reservoirs is 193–630 mD, which was estimated based on hydrodynamic tests [22].
The most promising exploration targets proved to be the basal sandstones and the carbonates of the Jurassic, occurring as isolated erosional relicts on the northern slope of the Nesvačilka graben. These relicts (buried hills) are sealed mainly by impermeable shales and mudstones of the paleovalley fill. The isolated bodies of turbiditic and channelized sandstones enclosed within the Paleogene fill of the Nesvačilka paleovalley represent the secondary exploration play.
Geochemical studies [23,24,25] suggest that the hydrocarbons in the Nesvačilka graben area were predominantly sourced from Jurassic, organic-rich Mikulov marls. The Paleogene shales and mudstones may be considered as additional source rocks, especially for gas.

2.2. Hydrogeological, Mineralogical, and Reservoir Characteristics of the Storage Site

Data and outputs from laboratory experiments carried out in the laboratories of the Department of Geological Engineering at VSB—Technical University of Ostrava and data received from partners within the project were used in this study.
The reservoir rocks represented by dolomite from the Vranovice fm. (labeled ZA4A) were the focus of this research (Table 1). The Vranovice fm. lies at a depth of about 1500 m and is formed by an Upper Jurassic light gray dolomite about 100 million years old.
The caprock is represented by the Nesvačilka fm. rock sample (ZA4) (Table 2). The Nesvačilka fm. lies at a depth of approximately 1500 m and is formed by dark gray shale mudstone from the Middle Paleogene.
The composition of the pore water of the reservoir formation was determined based on the analysis of a sample from the Vranovice formation from the Žarošice 4A well (Table 3). The water sample from the aquifer was taken by the project partner (MND a.s.) from a depth of 1530 m and chemical analysis of the formation water was carried out in the laboratories of the VSB—Technical University of Ostrava.
To model changes in the rock environment, it was also necessary to take into account the CO2 phase, which becomes supercritical under the conditions of the bearing horizon: temperature 53 °C; pressure 150 bar (determined on site by the project partner MND a.s.). The mineralogical composition and porosity of the studied rock samples are presented in Table 4.
The hydrogeological and reservoir parameters of the Vranovice formation (Table 5) are important for understanding and predicting changes in the storage area. During the extraction of hydrocarbons from the studied deposit, it was found that the vertical permeability, kz, is greater than the horizontal permeability, kR, which results from geological processes occurring after the diagenesis of the dolomite body. The studied reservoir rock shows typical fractured permeability.

2.3. Methods

The methods selected to investigate the impact of CO2 injection on changes in the rock environment at the studied site are based on information on petrophysical, mineralogical, and hydrogeological characteristics of porous media, hydrochemical analysis of the liquid composition, pressure and temperature values of the deposit, and kinetic parameters.

2.3.1. XRD Analysis and Porosity Determination

The mineralogical composition of rock samples was determined using XRD analysis by means of a Bruker AXS D8 θ-θ powder diffractometer. The experiments were conducted at room temperature in the Bragg–Brentan parafocusing geometry using CoKα wavelength (λ = 1.7903 Å, U = 34 kV, I = 30 mA). The porosity of the reservoir and caprock formations was initially assessed by the project partner (MND a.s.), based on acoustic well logging. However, for the purposes of this work, it was precisely determined using mercury porosimetry (Autopore 9220 Micro-metrics Injection Porosimeter, Micromeritics Instrument Corporation, Norcross, GA, USA) on cylindrical samples cut from drill cores. The samples from the ZA4A and ZA4 wells were analyzed before the experiments in the reaction chamber and after the subsequent experimental intervals.

2.3.2. Experiments in the Reaction Chamber

The experiments in the reaction chamber were conducted in two phases at the VSB—Technical University of Ostrava. The first phase involved only reservoir rock experiments and the second phase involved only caprock experiments. The experiments were carried out in the reaction chamber in rock-formation water–gas systems at temperatures below 53 °C and a pressure of 150 bars. The experiments lasted 30, 90, 120, and 180 days. During reservoir rock and caprock testing, the tested rocks had their own separate cells for a defined time interval. The formation water from the Žarošice 4A well was used as the reaction solution (see Table 3).
The reaction chamber is a cylindrical vessel welded from a shell made of stainless steel and DN150 PN160 flanges with an outer diameter of 169 mm and a length of 955 mm. The lids (blind flanges DN150 PN160) are also made of stainless steel. The apparatus is divided into two parts with two reaction cells. The inner diameter of the reaction cells is 141 mm and the length of the cells is 368 mm. The insides of the cells have a protective coating and the cells are further lined. The lining is inert to CO2. On the side of the casing, two pins are welded for attaching the pressure vessel in the bearings, which enables the positioning and rocking movement provided by the rocking mechanism.
The highest operating pressure is 200 bar. The range of operating temperatures is 5–80 °C. The heating for the reaction chamber is provided by an external heating coil. The operating medium can be reservoir water, hydrocarbons, CO2, or nitrogen. The dynamic conditions of the geohydrodynamic system are simulated by changing the position of the reaction chamber with two swings per day (this simulation method was used when solving the issue of CO2 geosequestration). Changes in position are secured using a swinging mechanism driven by an electric motor.

2.3.3. Geochemical Modeling

The geochemical modeling in this study focused on gas–water–rock interactions rather than on tracking the evolution of changes in CO2 capture in individual zones of the storage site. For this purpose, kinetic modeling was performed in order to determine mineralogical transformations and identify chemical reactions occurring in the gas–water–rock systems, changes in porosity, and to estimate the mineral and solution sequestration capacity. This approach also allowed for a partial comparison of the results of batch experiments with the results of reaction kinetics modeling.
Simulations were performed using Geochemist Workbench 11 (GWB) [26] and the thermodynamic database “thermo.dat”. The software used did not provide dynamic modeling in which it would be possible to track the evolution of changes in CO2 capture in individual zones of the storage site. Equilibrium models were aimed at assessing changes in the chemical composition of pore water in conditions of chemical equilibrium in the rock–water–CO2 system.
Kinetic modeling enabled the evaluation of hydrogeochemical changes in the studied formation in two stages: the first stage reproduced the immediate changes in the aquifer and caprock impacted by CO2 injection, and the second stage reproduced the long-term effects of sequestration.
This method further allowed for the assessment of the volume and amount of precipitated or dissolved mineral phases and their impact on the porosity of rocks in relation to the prediction of possible CO2-related hazards.
The geochemical modeling was performed assuming the temperature and formation pressure relevant to the depth of the modeled environments, which are given in Table 5. The appropriate value of carbon dioxide fugacity fCO2 75 bar was calculated based on the measured well pressure of 150 bar using the model in [27]. For the purpose of the simulation, the chemical compositions of the formation water in the reservoir and the caprock were obtained via equilibration of the formation water (Table 3) using the mineral assemblage typical for the modeled environments (Table 4).
The following kinetic dissolution/precipitation rate equation derived from the transition state theory [28] was used in simplified form in the calculations:
r k = A S k T 1 Q K
where rk is the reaction rate [mol·s−1] (dissolution at rk > 0, precipitation at rk < 0), AS is the mineral’s surface area [cm2], kT is the rate constant [mol·cm−2·s−1] at temperature T, Q is the activity product [-], and K is the equilibrium reaction for the dissolution reaction [-].
According to the above equation, a given mineral precipitates when it is supersaturated or dissolves when it is undersaturated at a rate proportional to its rate constant and the surface area. The Arrhenius law expresses the dependence of the rate constant kT on temperature T:
k T = k 25 exp E A R 1 T 1 298,15
where k25 is the rate constant at 25 °C [mol·m−2·s−1], EA is the activation energy [J·mol−1], R is the gas constant (8.3143 J·K−1·mol−1), and T is the absolute temperature (K).
The kinetic rate constants for the minerals involved in modeled reactions (Table 6) were taken from the authors of [29]. The values of the specific surface areas calculated according to the spherical model in [30] are presented in Table 6.

3. Results and Discussion

Immediately after injection, CO2 diffuses in its supercritical form and partially dissolves into the groundwater. This leads to acidification of the primary pore water, causes changes in the mineralogical composition of the reservoir and caprock, and may affect the well cement. All these reactions may pose potential risks in the CO2 injection process in the reservoir structure and may lead to CO2 leakage from the storage to the surface. Therefore, it is necessary to understand the mechanisms of these reactions and quantify and evaluate them to ensure the safety and sustainability of CO2 storage.

3.1. Equilibrium Model

The reaction of CO2 with groundwater and its effects on potential storage is of utmost importance for the process of mineral sequestration because only the dissolved form of CO2 can react with the rock. CO2 solubility is a function of the temperature, pressure, and salinity of the solution, with CO2 solubility decreasing with increases in the temperature and molarity of the solution, while CO2 solubility increases with increasing pressure. The dissolution simulation was carried out until the thermodynamic equilibrium was reached and the solution was saturated with CO2. Simulations were performed using Geochemist Workbench 11 (GWB) [26] and the thermodynamic database “thermo.dat”.
Based on the input parameters used in the model (brackish water Žarošice 4A, Table 3), we report the theoretical acidification of brackish water Žarošice 4A as a result of CO2 injection and calculated chemistry changes (Table 7). The HCO3 ion is the dominant carbonate component in the solution at alkaline pH. When the pH drops to acidic values, the bicarbonate ions decrease in solution and carbonic acid, and CO2(aq) becomes the dominant carbonate component [31].
The dissolution of CO2 in the solution results in a decrease in pH and an enrichment of carbonate ions according to the following reactions:
CO2(g) ↔ CO2(aq)
CO2(aq) + H2O ↔ H2CO3
H2CO3 ↔ HCO3 + H+
HCO3 ↔ CO32− + H+
After the reaction of brackish water Žarošice 4A with spCO2, there was an increase in CO2(aq) to a value of 11,584.62 mg/L as a result of the dissolution of CO2(g) according to Equation (3) under the reservoir conditions. This process acidified the water, lowering the pH to a value of 4.2.
We observed a slight decrease in the concentration of sodium and sulfate ions; on the contrary, the concentration of chlorides increased slightly. During the reaction with spCO2, Na+ cations reacted with sulfates and bicarbonates to form NaSO4 and NaHCO3 complexes. The most important change during CO2 injection into the pore environment of the collector saturated with formation water was an increase in the content of dissolved CO2(aq) in the solution to a value of 11,584.62 mg/L and a decrease in pH to a value of 4.2. Therefore, the concentration of bicarbonates, which participated in the buffering of the system during CO2 injection, dropped significantly.

3.2. Experiments in the Reaction Chamber

The mineralogical composition of all rock samples after the experimental intervals was determined using XRD analysis. Although data on mineral composition evolution derived from the XRD method have certain limitations, their presentation in this work shows additional possibilities for predicting mineral changes in the host rock as a result of CO2 injection.
After the end of the experiments, the reaction solutions were analyzed in the laboratories of the Institute of Environmental Engineering at VSB—TUO in Ostrava (Czech Republic). The chemical composition of the formation water is shown in Table 8.
The following graphs (Figure 2) show the mineralogical changes in the rock samples that occurred during the experiments in the reaction chamber.
Mineralogical changes in the reservoir rock over 30 days are shown in Figure 2a. As shown in the figure, the dolomite content decreased. The formation water after 30 days of experiments in the reaction chamber was depleted in magnesium but enriched with calcium and carbonate ions. A decrease in magnesium in the solution indicates the precipitation of ankerite. This is also confirmed by the decrease in iron concentration. Calcite precipitates because the dissolution of dolomite releases calcium and bicarbonate ions. Albite and muscovite remained stable throughout the experiment. Sodium concentration dropped significantly after 30 days of CO2 injection, with no indication of chemical binding of the sodium ion in another mineral phase. The concentration of potassium in the solution increased slightly, although no precipitation or dissolution of muscovite was shown in the XRD analysis results. However, there was a slight decrease in the content of microcline, which also releases potassium. Quartz and kaolinite slightly increased, and minerals from the chlorite group were completely dissolved. Due to the dissolution of chlorite, the concentration of manganese increased in the solution.
Mineralogical changes in the reservoir rock from the injection of CO2 over 90 days are shown in Figure 2b. Dolomite, kaolinite, and microcline dissolved only slightly, while chlorite was completely dissolved. This effect led to an increase in the concentrations of calcium, magnesium, iron, manganese, and carbonate ions. The potassium concentration increased only very slightly. Calcite precipitated due to the dissolution of dolomite and ankerite, and also due to the dissolution of minerals from the chlorite group. The content of albite increased after 90 days of the experiment and, simultaneously, the concentration of sodium in the solution decreased. Muscovite remained stable throughout the experiment.
Mineralogical changes in the reservoir rock over 120 days are shown in Figure 2c. An important phenomenon observed after 120 days of the experiment was the increase in dolomite content. This is related to the dissolution of ankerite during this experiment. Calcium, magnesium, iron, and carbonate ions are released into the solution. This also indicates the precipitation of calcite, although in a much lower content than in the previous experiments. Quartz, kaolinite, and albite dissolved (the latter is evidenced by an increased sodium concentration in the solution). Muscovite and microcline precipitated, but the potassium concentration increased significantly. This may indicate the existence of another mineral as a source of potassium, which was not revealed in the XRD analysis results, e.g., potassium feldspar or minerals from the potassium salts group (sylvite and carnallite). Chlorite was completely dissolved.
Mineralogical changes in the reservoir rock over 180 days are shown in Figure 2d. Only dolomite and muscovite were dissolved in this case. Calcium, magnesium, potassium, and carbonate ions were released into the solution. Dissolution of dolomite allows the precipitation of calcite and ankerite. In this experiment, chlorite also precipitated, so iron was consumed from the solution. Hence, another mineral (e.g., accessory pyrite) can be considered as the source of iron, which is needed for ankerite precipitation. Quartz, kaolinite, microcline, and albite slightly increased. The concentration of potassium in the solution after 180 days of the experiment increased due to muscovite dissolution, unlike sodium, which decreased due to the precipitation of albite. A slight increase in pH also contributed to the precipitation of carbonates.
Figure 3 presents the mineralogical changes in the ZA4 caprock samples that occurred during the experiments.
Mineralogical changes in the cap rock over 30 days are shown in Figure 3a. The content of CaFe–dolomite (ankerite) decreased. After 30 days of experiments, the formation water in the reaction chamber was depleted in magnesium but enriched with calcium and carbonate ions. The dissolution of ankerite is confirmed by the increase in the concentration of iron in solution. Some iron was consumed from the reservoir water solution during the precipitation of pyrite, whose concentration increased slightly after 30 days of the experiment. Albite remained stable throughout the experiment, but sodium concentration very slightly increased after 30 days of CO2 injection, which indicates another source of sodium ions in some mineral phases. The concentration of potassium in the solution increased slightly, but the precipitation of muscovite was demonstrated by the XRD analysis results. There was a slight increase in the content of microcline, which also fixes potassium. The slightly higher potassium concentration after 30 days of experiment in the reaction chamber indicates an initial rapid dissolution of microcline and muscovite, enrichment of the solution with K+ ion, and subsequent secondary precipitation of muscovite. The concentration of manganese increased in the solution and minerals from the chlorite group were precipitated. The amount of manganese in solution indicates another source of this element in the caprock sample. Quartz increased slightly.
Mineralogical changes in the cap rock from the injection of CO2 over 90 days are shown in Figure 3b. CaFe–dolomite and microcline dissolved. Quartz, muscovite, and pyrite were precipitated. These processes led to an increase in the calcium, iron, manganese, and carbonate ions concentrations in the solution. The potassium and magnesium concentrations decreased only very slightly. Calcite precipitated due to the dissolution of ankerite. Albite and minerals from the chlorite group remained stable throughout the experiment, but sodium concentration decreased. This indicates the fixation of sodium ions into other mineral phases, such as muscovite.
Mineralogical changes in the reservoir rock over 120 days are shown in Figure 3c. The decrease in CaFe–dolomite content was lower compared with experiments lasting 30 days. Calcium and carbonate ions were released into the solution, but secondary calcite was not detected in the case of this sample. The solution did not have suitable thermodynamic conditions for its precipitation. The concentration of iron ions was the biggest in this phase of the experiment due to the ankerite dissolution. Some iron was consumed from the solution during the precipitation of pyrite and minerals from the chlorite group, whose concentration increased slightly after 120 days of the experiment. Albite, muscovite, chlorite, and microcline precipitated very slightly. This was reflected by increased potassium concentrations in the solution, but this did not correlate with the very low sodium concentrations. Quartz was dissolved.
Mineralogical changes in the cap rock over 180 days are shown in Figure 3d. Only CaFe–dolomite was dissolved in this case. Calcium and carbonate ions were released into the solution again. The concentration of potassium, sodium, and magnesium in the solution after 180 days of the experiment decreased by precipitation of most of the defined mineral phases.
Secondary precipitation of most defined mineral phases is a favorable phenomenon in cap rock. It leads to a decrease in their permeability. However, precipitation of secondary phases is not a positive phenomenon in reservoir rock where decreasing permeability reduces the sequestration capacity of the reservoir rock.

3.3. Results of and Discussion on Kinetic Modeling

The kinetic modeling of mineralogical changes in caprock and reservoir rock was carried out by means of the GWB React package in two stages. The first stage, performed in sliding fugacity mode, represents short-term changes in the system under the influence of CO2 injection over a period of 100 days, during which the gas fugacity rises in the system to the assumed value of fCO2 75 bar. The second stage reproduces the long-term effects of sequestration over a period of 10,000 years.

3.3.1. Stage 1—Caprock Sample ZA4

At the first stage of the CO2 injection, lasting for 100 days, gas fugacity (fCO2) increased to 75 bar (corresponding to a reservoir pressure of 150 bar). In effect, the CO2 concentration in solution, CO2(aq), increased from 0.004 mol/kg to 1.384 mol/kg, and HCO3 increased from 0.034 mol/kg to 0.094 mol/kg. Due to the dissolution of primary minerals, the porosity increased slightly by 0.15 percentage points, from 4.79 to 4.94. At the same time, the pH dropped from 7.1 to 4.9 due to the hydrochemical balance between the species in solution (Figure 4).
The mechanisms of mineral sequestration of CO2 are activated mainly as a result of Reaction (7), which involves the decomposition of primary K-feldspar, in which dawsonite is formed. The dissolution of primary calcite (Equation (8)) leads to an increase in bicarbonate and calcium concentrations in the pore solution (Figure 5).
Na+ + KAlSi3O8+ CO2(g) + H2O = NaAlCO3(OH)2 + K+ + 3SiO2
CaCO3 + CO2(g) + H2O = Ca++ + 2HCO3

3.3.2. Stage 2—Caprock Sample ZA4

In this stage, gas fugacity (fCO2) dropped to 0.2 bar, the CO2 concentration in solution (CO2(aq)) dropped to about 0.0003 mol/kg, and HCO3 decreased to 0.011 mol/kg. The porosity decreased slightly to 4.92. At the same time, the pH increased above the primary value to 7.7 (Figure 6).
Mineral sequestration of CO2 is mainly associated with the formation of dolomite, which precipitates due to the decomposition of ankerite: Equation (9). The formation of dawsonite (Equation (10)) plays a minor role in only the first 20 years of the storage period (Figure 7 and Figure 8). It should be noted, however, that in reality, the processes of dissolution of ankerite and crystallization of dolomite will proceed at a similar, but not identical, rate. For modeling purposes, the rate constant for ankerite was assigned based on dolomite and for this reason, the formation of secondary minerals occurs effectively under conditions close to local equilibrium.
CaMg0.3Fe0.7(CO3)2 + 2.24CO2(g) + 0.14Mg5Al2Si3O10(OH)8 + 0.56H2O = 2.24HCO3 + CaMg(CO3)2 + 0.7Fe2+ + 0.28Al3+ + 0.42SiO2(aq)
CaCO3 + CO2(g) + 0.4HCO3 + 0.2Mg5Al2Si3O10(OH)8 + 0.4Na+ = CaMg(CO3)2 + 0.6H2O + 0.4NaAlCO3(OH)2 + 0.6SiO2(aq)

3.3.3. Stage 1—Reservoir Rock Sample ZA4A

During stage 1, gas fugacity (fCO2) increased to 75 bar, the CO2 concentration in solution (CO2(aq)) changed from 0.004 mol/kg to 1.396 mol/kg, and HCO3 increased from 0.033 mol/kg to 0.062 mol/kg. The porosity increased by 0.0023, from 8.15 to 8.38. At the same time, the pH dropped from 7.1 to 4.8 (Figure 9).
During the injection stage, changes in the mineral composition (Figure 10) follow Equation (7) of the K-feldspar dissolution, in which dawsonite is produced. Ankerite dissolution is also important, causing siderite and dolomite to precipitate (Equation (11)):
CaMg0.3Fe0.7(CO3)2 + 0.7CO2(g) + 0.7H2O = 0.7Ca2+ + 0.7FeCO3 + 1.4HCO3 + 0.3CaMg(CO3)2

3.3.4. Stage 2—Reservoir Rock Sample ZA4A

During stage 2, gas fugacity (fCO2) dropped to 0.06 bar, the CO2 concentration in solution (CO2(aq)) decreased to 0.0011 mol/kg, and HCO3 decreased to 0.013 mol/kg. The porosity decreased to 7.23. At the same time, the pH increased above the primary value to 7.2 (Figure 11).
During the storage stage (Figure 12), the mineral sequestration of CO2 takes place in three steps in which the ankerite dissolution plays the main role:
(1)
In the first step, dawsonite formation occurs due to the dissolution of albite (Equation (12)) and the precipitation of dolomite and siderite (Equation (11)).
NaAlSi3O8 + CO2(g) + H2O = NaAlCO3(OH)2 + 3SiO2
(2)
In the second step, dawsonite is dissolved and dolomite and siderite are precipitated (Equation (13)).
CaMg0.3Fe0.7(CO3)2 + 1.54CO2(g) + 0.14Mg5Al2Si3O10(OH)8 = 0.84HCO3 + CaMg(CO3)2 + 0.7FeCO3 + 0.14H2O + 0.28Al3+ + 0.42SiO2(aq)
(3)
Dolomite precipitates due to ankerite dissolution in the third step (Equation (14)).
CaMg0.3Fe0.7(CO3)2 + 1.53CO2(g) + 0.14Mg5Al2Si3O10(OH)8 + 0.55H2O + 0.29NaAlSi3O8 + 0.17O2(aq) =
1.53HCO3 + CaMg(CO3)2 + 0.35Mg0.165Fe2Al0.33Si3.67O10(OH)2 + 0.45Al3+ + 0.17Na+

3.3.5. Sequestration Capacity in Mineral Forms and Solution

The use of the reaction path model (GWB React) only allowed for the calculation of mineral and solution trapping; therefore, no assessment of the hydrodynamic trapping capacity was made. However, the mechanism of mineral trapping can be considered the most important because it ensures permanent trapping of CO2.
The estimated mineral and solution sequestration capacities refer to closed systems in which, after CO2 fugacity reaches the level assumed for the end of the injection stage, only the transformation of CO2 aqueous phases into the mineral phases can occur. This type of approach allows for an approximate assessment of the size and mutual proportions of the effectiveness of the two trapping processes analyzed.
Based on kinetic modeling, changes in the volume of primary and secondary minerals during simulated reactions were calculated, which demonstrated an effect on the porosity of the analyzed formations. In addition, the amounts of dissolved or precipitated carbonates were determined and, based on their balance, the amount of sequestered CO2 was calculated for each cubic meter of formation.
The chemical composition of pore water, especially the concentrations of aqueous species (e.g., HCO3, CO2(aq), CO32−, and NaHCO3) obtained based on simulations, enabled an assessment of the amount of carbon dioxide trapped in the solution. This assessment was calculated based on the volume of pore space in a 1 m3 rock formation multiplied by the amount of carbon dioxide contained in each component. During the injection stage lasting 100 days, it was determined that CO2 trapping in solution was more effective than the mineral trapping mechanism (Table 9).
Both analyzed formations are characterized by a low total CO2 sequestration potential in mineral form and solution for 10,000 years of storage, amounting to 3.22 and 5.50 kg CO2/m3 for the ZA4 caprock and ZA4A reservoir rocks, respectively. These values are lower than the ones obtained in simulations for other considered storage formations [32,33,34,35].
The sequestration capacity values calculated in our research are also low compared with the results from the depleted Brodské oil field (Middle Badenian), which is also located in the Moravian part of the Vienna Basin. Capacities calculated for the latter location were 13.22 kg CO2/m3 for the reservoir and 5.07 kg CO2/m3 for the caprock, respectively [36].
During 100 days of injection, solution trapping predominated, while in the storage stage, mineral trapping showed a significantly greater potential (Table 9). The minerals responsible for mineral sequestration in each of the rocks are shown in Table 10.

4. Summary and Conclusions

Experiments and geochemical modeling of gas–water–rock interactions were carried out in order to characterize the impact of CO2 on mineralogical changes in reservoir rocks and caprock and assess the sequestration capacity of a potential storage site in Czechia. The results of experimental research and geochemical modeling led to the following conclusions.
At the CO2 injection stage, reservoir rocks with a significant share of dolomite and ankerite and a small percentage of feldspars showed a slight increase in porosity by 0.25 pp. due to the decomposition of ankerite and feldspar. The secondary carbonate minerals of the storage stage observed in this study were dolomite, siderite, and ephemeral dawsonite, which were present in only the first 50 years of storage; porosity decreased by 1.20 pp at this stage.
In the caprocks, with a high proportion of quartz, microcline, and ankerite, the decomposition of K-feldspar and calcite was responsible for the increase in porosity by 0.15 pp at the injection stage. Therefore, it can be concluded that the increase in porosity of the caprock is of little importance and corresponds to only 3% of the initial value, which has practically no effect on the sealing properties. During the storage stage, the decomposition of ankerite and calcite resulted in the precipitation of secondary dolomite; moreover, mineral changes had almost no effect on porosity.
The reservoir formation and caprocks have a low total CO2 sequestration potential in mineral form and solution over 10,000 years of storage, amounting to 5.50 kg CO2/m3 for reservoir formation (4.37 kg CO2/m3 in mineral form and 1.13 kg CO2/m3 in dissolved form) and 3.22 kg CO2/m3 for caprock (3.01 kg CO2/m3 in mineral form and 0.21 kg CO2/m3 in dissolved form), respectively. These values may not be encouraging, especially when compared with the results of the depleted Brodské oil field, which is also located in the Moravian part of the Vienna Basin.
During the 100-day injection period, solution trapping dominates, while during the storage stage, mineral trapping has a greater potential.
The ability to store CO2 in the aquifer depends on many parameters, such as the composition of the rock matrix, porosity, permeability, pressure, and salinity. These parameters have a significant impact on sequestration processes. Therefore, it should be noted that most of these parameters evolve over time. Moreover, in the model used in this study, it was not possible to take into account these changes, e.g., the specific surface of minerals or changes in reservoir temperature due to the gas injection.

Author Contributions

Conceptualization, M.L., K.L. and M.K.; methodology, M.L., K.L., M.K. and D.M.; software, K.L. and M.L.; validation, M.L. and K.L.; formal analysis, D.M., K.L. and M.L.; investigation, M.L., K.L. and M.K.; resources, M.L., K.L., M.K. and M.V.; data curation, M.L. and K.L.; writing—original draft preparation, M.L., K.L. and M.K.; writing—review and editing, K.L. and M.L.; visualization, M.L. and M.V.; supervision, K.L.; project administration, M.K. and M.L.; funding acquisition, M.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by project SGS No. SP2023/080 and SP2022/125, Faculty of Mining and Geology, VSB—Technical University Ostrava, Czech Republic, and by EEA—Norway Grants and the Technology Agency of the Czech Republic (TA ČR) within the CO2-SPICER project (No: TO01000112, CO2 Storage Pilot In a Carbonate Reservoir).

Data Availability Statement

Publicly available datasets were analyzed in this study. This data can be found here: https://zenodo.org/communities/co2spicer/records?q=&l=list&p=1&s=20&sort=newest.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Climate Change: Atmospheric Carbon Dioxide. Available online: https://www.climate.gov/news-features/understanding-climate/climate-change-atmospheric-carbon-dioxide (accessed on 24 April 2024).
  2. Benson, S.B.; Cole, D.R. CO2 sequestration in deep sedimentary formations. Elements 2008, 4, 325–331. [Google Scholar] [CrossRef]
  3. Gaus, I. Role and impact of CO2–rock interactions during CO2 storage in sedimentary rocks. Int. J. Greenh. Gas Control 2010, 4, 73–89. [Google Scholar]
  4. Izeg, O.; Demiral, B.; Bertin, H.; Akin, S. CO2 injection into saline carbonate aquifer formations I: Laboratory investigation. Transp. Porous Media 2008, 72, 1–24. [Google Scholar]
  5. Bacci, G.; Korre, A.; Durucan, A. An experimental and numerical investigation into the impact ofdissolution/precipitationmechanisms on CO2 injectivity inthe wellbore and far field regions. Int. J. Greenh. Gas Control 2011, 5, 579–588. [Google Scholar] [CrossRef]
  6. Saeedi, A.; Rezaee, R.; Evans, B.; Clennell, B. Multiphase flow behaviour during CO2 geo-sequestration: Emphasis on the effect of cyclic CO2–brine flooding. J. Pet. Sci. Eng. 2011, 79, 65–85. [Google Scholar] [CrossRef]
  7. Sayegh, S.G.; Krause, F.F.; Girard, M. Rock/fluid interactions of carbonated brines in a sandstone reservoir Pembina Cardium, Alberta, Canada. SPE Form. Eval. 1990, 5, 399–405. [Google Scholar] [CrossRef]
  8. Desbois, G.; Urai, J.; Kukla, P.; Konstanty, J.; Baerle, C. High-resolution 3D fabric and porosity model in a tight gas sandstone reservoir: A new approach to investigate microstructures from mm- to nm-scale combining argon beam cross-sectioning and SEM imaging. J. Pet. Sci. Eng. 2011, 78, 243–257. [Google Scholar] [CrossRef]
  9. Gunter, W.D.; Bachu, S.; Benson, S. The Role of Hydrogeological and Geochemical Trapping in Sedimentary Basins for Secure Geological Storage of Carbon Dioxide; Geological Society: London, UK, 2008; Volume 233, pp. 129–145. [Google Scholar]
  10. Luquot, L.; Gouze, P. Experimental determination of porosity and permeability changes induced injection of CO2 into carbonate rocks. Chem. Geol. 2009, 265, 148–159. [Google Scholar] [CrossRef]
  11. Rochelle, C.A.; Czernichowski-Lauriol, I.; Milodowski, A.E. The Impact of Chemical Reactions on CO2 Storage in Geological Formations: A Brief Review; Geological Society: London, UK, 2004; pp. 87–106. [Google Scholar]
  12. Cailly, B.; Le Thiez, P.; Egermann, P.; Audibert, A.; Vidal-Gilbert, S.; Longaygue, X. Geological storage of CO2: A state-of-the-art of injection processes and technologies. Oil Gas Sci. Technol.–Rev. IFP 2005, 60, 517–525. [Google Scholar] [CrossRef]
  13. Kaszuba, J.P.; Janecky, D.R.; Snow, M.G. Experimental evaluation of mixed fluid reactions between supercritical carbon dioxide and NaCl brine: Relevance to the integrity of a geologic carbon repository. Chem. Geol. 2005, 217, 277–293. [Google Scholar] [CrossRef]
  14. Kaszuba, J.P.; Janecky, D.R.; Snow, M.G. Carbon dioxide reaction processes in a model brine aquifer at 200 °C and 200 bars: Implications for geologic sequestration of carbon. Appl. Geochem. 2003, 18, 1065–1080. [Google Scholar] [CrossRef]
  15. Mito, S.; Xue, Z.; Ohsumi, T. Case study of geochemical reactions atthe Nagaoka CO2 injection site, Japan. Int. J. Greenh. Gas Control 2008, 2, 309–318. [Google Scholar] [CrossRef]
  16. Sterpenich, J.; Sausse, J.; Pironon, J.; Géhin, A.; Hubert, G.; Perfetti, E.; Grgic, D. Experimental ageing of oolitic limestones under CO2 storage conditions: Petrographical and chemical evidence. Chem. Geol. 2009, 265, 99–112. [Google Scholar] [CrossRef]
  17. Kostelnıcek, P.; Ciprys, V.; Berka, J. Examples of recently discovered oil and gas fields in the Carpathian foredeep and in the European foreland plate underneath the Carpathian thrust belt, Czech Republic. In The Carpathians and Their Foreland: Geology and Hydrocarbon Resources: AAPG Memoir; Golonka, J., Picha, F.J., Eds.; The American Association of Petroleum Geologists: St. Tulsa, OK, USA, 2006; Volume 84, pp. 177–189. [Google Scholar] [CrossRef]
  18. Jiricek, R. Die stratigraphische und fazielle Unterteilung der autochtonen Sedimenten des Palaogens and SO Abhangen der Bohmische Masse (in Czech with German summary): Knihovnicka Zemniho plynu a nafty 6b. Misc. Micropaleontol. Hodonin 1987, II/2, 247–314. [Google Scholar]
  19. Picha, F. Ancient submarine canyons of Tethyan continental margins, Czechoslovakia. AAPG Bull. 1979, 63, 67–86. [Google Scholar]
  20. Picha, F. Exploring for hydrocarbons under thrust belts—A challenging new frontier in the Carpathians and elsewhere. AAPG Bull. 1996, 80, 1547–1564. [Google Scholar]
  21. Opletal, V.; Klímová, L.; Pagáč, M.; Prochác, R.; Pereszlényi, M.; Franců, J.; Jirman, P.; Jurenka, L.; Rez, J.; Tveranger, J.; et al. 3D Geological Model of the Storage Complex and Set of Structural Geological Maps. TO01000112-V1, CO2-SPICER—CO2 Storage Pilot In a CarbonatE Reservoir; Project report; Zenodo: Genève, Switzerland, 2022. [Google Scholar]
  22. Durica, D.; Suk, M.; Ciprys, V. Energetické Zdroje–Včera, Dnes a Zítra; Moravské Zemské Múzeum: Brno, Czech Republic, 2010; 165p. [Google Scholar]
  23. Francu, J.; Radke, M.; Schaefer, R.G.; Poelchau, H.S.; Caslavsky, J.; Bohacek, Z. Oil-oil and oil-source rock correlations in the northern Vienna basin and adjacent Carpathian Flysch Zone (Czech and Slovak area). In Oil and Gas in Alpidic Thrustbelts and Basins of Central and Eastern Europe; European Association of Geoscientists and Engineers Special Publication; Wessely, G., Liebl, W., Eds.; Instytut Nafty i Gazu-Państwowy Instytut Badawczy: Kraków, Poland, 1996; Volume 5, pp. 343–353. [Google Scholar]
  24. Krejci, O.; Francu, J.; Poelchau, H.S.; Muller, P.; Stranik, Z. Tectonic evolution and oil and gas generation at the border of the north European platform with the West Carpathians (Czech Republic). In Oil and Gas in Alpidic Thrustbelts and Basins of Central and Eastern Europe; European Association of Geoscientists and Engineers Special Publication; Wessely, G., Liebl, W., Eds.; Instytut Nafty i Gazu-Państwowy Instytut Badawczy: Kraków, Poland, 1996; Volume 5, pp. 177–186. [Google Scholar]
  25. Picha, F.; Peters, K.E. Biomarker oil-to-source rock correlation in the Western Carpathians and their foreland, Czech Republic. Petroleum Geosci. 1998, 4, 289–302. [Google Scholar] [CrossRef]
  26. Bethke, C.M. The Geochemist’s Workbench, Release 6.0—Reference Manual; University of Illinois: Springfield, IL, USA, 2006. [Google Scholar]
  27. Duan, Z.H.; Sun, R.; Zhu, C.; Chou, I.M. An improved model for the calculation of CO2 solubility in aqueous solutions containing Na+, K+, Ca2+, Mg2+, Cl, and SO42−. Mar. Chem. 2006, 98, 131–139. [Google Scholar] [CrossRef]
  28. Lasaga, A.C. Chemical kinetics of water-rock interactions. J. Geophys. Res. 1984, 89, 4009–4025. [Google Scholar] [CrossRef]
  29. Palandri, J.L.; Kharaka, Y.K. A Compilation of Rate Parameters of Water-Mineral Interaction Kinetics for Application to Geochemical Modeling (Open File Report 2004-1068); U.S. Geological Survey: Menlo Park, CA, USA, 2004.
  30. Labus, K.; Bujok, P. CO2 mineral sequestration mechanisms and capacity of saline aquifers of the Upper Silesian Coal Basin (Central Europe)—Modeling and experimental verification. Energy 2011, 36, 4974–4982. [Google Scholar] [CrossRef]
  31. Appelo, C.A.J.; Postma, D. Geochemistry, Groundwater and Pollution, 2nd ed.; A.A. Balkema Publishers: Leiden, The Netherlands, 2005; p. 635. ISBN 04 1536 421 3. [Google Scholar]
  32. Klunk, M.A.; Shah, Z.; Caetano, N.R.; Conceição, R.V.; Wander, P.R.; Dasgupta, S.; Das, M. CO2 sequestration by magnesite mineralisation through interaction of Mg-brine and CO2: Integrated laboratory experiments and computerised geochemical modelling. Int. J. Environ. Stud. 2020, 77, 492–509. [Google Scholar] [CrossRef]
  33. Labus, K.; Tarkowski, R.; Wdowin, M. Assessment of CO2 sequestration capacity based on hydrogeochemical model of Water-Rock-Gas interactions in the potential storage site within the Bełchatów area (Poland). Miner. Resour. Manag. 2010, 26, 69–84. [Google Scholar]
  34. Labus, K. Phenomena at interface of saline aquifer and claystone caprock under conditions of CO2 storage. Ann. Soc. Geol. Pol. 2012, 82, 255–262. [Google Scholar]
  35. Xu, T.; Apps, J.A.; Pruess, K. Reactive geochemical transport simulation to study mineral trapping for CO2 disposal in deep arenaceous formations. J. Geophys. Res. 2003, 108, 2071–2084. [Google Scholar] [CrossRef]
  36. Labus, K.; Bujok, P.; Klempa, M.; Porzer, M.; Matýsek, D. Preliminary geochemical modeling of water-rock-gas interactions controlling CO2 storage in the Badenian Aquifer within Czech Part of Vienna Basin. Environmental Earth Sci. 2016, 75, 1086. [Google Scholar] [CrossRef]
Figure 1. The area of storage formation ZAR-3 site (Žarošice hydrocarbon field) formed by dolomite of the Vranovice Fm. (rock sample ZA4A) and the area of Nesvačilka Fm. formed by Paleogene claystone (rock sample ZA4) and represents tested caprock at the study site: NNW-SSE geological profile with well logs (depth below sea level) [21] (modified by authors).
Figure 1. The area of storage formation ZAR-3 site (Žarošice hydrocarbon field) formed by dolomite of the Vranovice Fm. (rock sample ZA4A) and the area of Nesvačilka Fm. formed by Paleogene claystone (rock sample ZA4) and represents tested caprock at the study site: NNW-SSE geological profile with well logs (depth below sea level) [21] (modified by authors).
Minerals 14 00602 g001
Figure 2. Mineralogical changes in the ZA4A reservoir rock sample after 30 days (a), 90 days (b), 120 days (c), and 180 days (d) from the beginning of CO2 injection. Dol—dolomite, Q—quartz, Calc—calcite, Mu—muscovite, Ka—kaolinite, Mic—microcline, Alb—albite, Chl—chlorite, and Pyr—pyrite.
Figure 2. Mineralogical changes in the ZA4A reservoir rock sample after 30 days (a), 90 days (b), 120 days (c), and 180 days (d) from the beginning of CO2 injection. Dol—dolomite, Q—quartz, Calc—calcite, Mu—muscovite, Ka—kaolinite, Mic—microcline, Alb—albite, Chl—chlorite, and Pyr—pyrite.
Minerals 14 00602 g002
Figure 3. Mineralogical changes in the ZA4 caprock sample over 30 days (a), 90 days (b), 120 days (c), and after 180 days (d) from the beginning of CO2 injection. Dol—dolomite, Q—quartz, Calc—calcite, Mu—muscovite, Ka—kaolinite, Mic—microcline, Alb—albite, Chl—chlorite, and Pyr—pyrite.
Figure 3. Mineralogical changes in the ZA4 caprock sample over 30 days (a), 90 days (b), 120 days (c), and after 180 days (d) from the beginning of CO2 injection. Dol—dolomite, Q—quartz, Calc—calcite, Mu—muscovite, Ka—kaolinite, Mic—microcline, Alb—albite, Chl—chlorite, and Pyr—pyrite.
Minerals 14 00602 g003
Figure 4. Caprock sample ZA4: changes in fCO2, some component concentrations, pH, and porosity at the injection stage.
Figure 4. Caprock sample ZA4: changes in fCO2, some component concentrations, pH, and porosity at the injection stage.
Minerals 14 00602 g004
Figure 5. Caprock sample ZA4: relative changes in mineral composition at the injection stage.
Figure 5. Caprock sample ZA4: relative changes in mineral composition at the injection stage.
Minerals 14 00602 g005
Figure 6. Caprock sample ZA4: changes in fCO2, pH, porosity, and some component concentrations at the storage stage (the lower figure in logarithmic scale).
Figure 6. Caprock sample ZA4: changes in fCO2, pH, porosity, and some component concentrations at the storage stage (the lower figure in logarithmic scale).
Minerals 14 00602 g006
Figure 7. Caprock sample ZA4: relative changes in mineral composition at the storage stage.
Figure 7. Caprock sample ZA4: relative changes in mineral composition at the storage stage.
Minerals 14 00602 g007
Figure 8. Caprock sample ZA4: relative changes in carbon-bearing minerals at the storage stage.
Figure 8. Caprock sample ZA4: relative changes in carbon-bearing minerals at the storage stage.
Minerals 14 00602 g008
Figure 9. Reservoir rock sample ZA4A: fCO2, some component concentrations at the storage stage, pH, and porosity.
Figure 9. Reservoir rock sample ZA4A: fCO2, some component concentrations at the storage stage, pH, and porosity.
Minerals 14 00602 g009
Figure 10. Reservoir rock sample ZA4A: relative changes in mineral composition at the injection stage.
Figure 10. Reservoir rock sample ZA4A: relative changes in mineral composition at the injection stage.
Minerals 14 00602 g010
Figure 11. Reservoir rock sample ZA4A: fCO2, pH, porosity, and some component concentrations at the storage stage (the lower figure in logarithmic scale).
Figure 11. Reservoir rock sample ZA4A: fCO2, pH, porosity, and some component concentrations at the storage stage (the lower figure in logarithmic scale).
Minerals 14 00602 g011
Figure 12. Reservoir rock sample ZA4A: relative changes in mineral composition at the storage stage.
Figure 12. Reservoir rock sample ZA4A: relative changes in mineral composition at the storage stage.
Minerals 14 00602 g012
Table 1. Basic data for the samples of reservoir rocks.
Table 1. Basic data for the samples of reservoir rocks.
BoreholeCoreBoxTrue Vertical Depth (m)StratigraphyLithostratigraphyLithology
ZA4A331528.5–1537.5Upper JurassicVranovice fm.Light gray dolomite
ZA4A421571.0–1578.8Upper JurassicVranovice fm.Light gray dolomite
Table 2. Basic data for the sample of caprock.
Table 2. Basic data for the sample of caprock.
BoreholeCoreBoxTrue Vertical Depth (m)StratigraphyLithostratigraphyLithology
ZA4151569.0–1577.0Middle PaleogeneNesvačilka fm.Dark gray shaly siltstone
Table 3. Physicochemical parameters of formation water from the Žarošice 4A well.
Table 3. Physicochemical parameters of formation water from the Žarošice 4A well.
Formation Water Žarošice 4A
Ca2+111.0mg/L
Mg2+107.0mg/L
Na+8570.0mg/L
K+252.0mg/L
Cl12,300.0mg/L
NH4+32.6mg/L
Br54.0mg/L
I43.0mg/L
HCO32580.0mg/L
SO42−535.0mg/L
Li+3.2mg/L
Sr2+19.2mg/L
Mn2+0.0mg/L
Fetotal0.4mg/L
B3+43.0mg/L
TDS24,590.0mg/L
Other parameters
T53°C
pH7.1-
Depth1530Meter
TDS—total dissolved substances; T—temperature.
Table 4. Mineralogical composition and porosity of the rock samples.
Table 4. Mineralogical composition and porosity of the rock samples.
BoreholeDolAnkQCalcMuKaMicAlbChlPyrPorosity
Reservoir ZA4A(wt%)(%)
55.0417.954.010.002.883.127.667.921.42--
(vol. %)(%)
50.6016.563.680.002.763.237.367.391.12-8.00
Caprock
ZA4
(wt%)(%)
-18.1057.100.304.00-16.681.181.141.50-
(vol. %)(%)
-17.2554.420.293.81-15.901.121.091.434.70
Dol—dolomite, Ank—ankerite, Q—quartz, Calc—calcite, Mu—muscovite, Ka—kaolinite, Mic—microcline, Alb—albite, Chl—chlorite, and Pyr—pyrite.
Table 5. Values of CO2 density and viscosity and hydrogeological and reservoir parameters of light gray dolomite reservoir rock (measured by the project partner MND a.s).
Table 5. Values of CO2 density and viscosity and hydrogeological and reservoir parameters of light gray dolomite reservoir rock (measured by the project partner MND a.s).
Parameters ValueUnit
Thickness200m
Porosity8.00%
Horizontal permeability, kR500mD
Vertical permeability, kz1600mD
Pressure P 116bar
Temperature T 53°C
Maxsalinity S 72 g/L
CO2 density195 kg/m3
CO2 viscosity1.5 × 10−4Pa.s
Table 6. Kinetic rate constants and specific surface areas applied in the models.
Table 6. Kinetic rate constants and specific surface areas applied in the models.
MineralKinetic Rate k25
(mol/cm2·s−1)
Specific Surface
(cm2/g)
Dolomite7.760 × 10−10161.1
Ankerite7.760 × 10−10146.6
Quartz1.023 × 10−18113.3
Calcite3.311 ×10−8170.3
Muscovite1.413 × 10−16212.0
Kaolinite4.898 × 10−161156.0
K-feldspar8.710 × 10−15180.5
Albite6.918 × 10−15229.0
Clinochlore7.762 × 10−161118.0
Pyrite3.020 × 10−12149.7
Table 7. Physicochemical parameters of original formation water from the Žarošice 4A well and model-calculated acidification of formation water Žarošice 4A.
Table 7. Physicochemical parameters of original formation water from the Žarošice 4A well and model-calculated acidification of formation water Žarošice 4A.
Original Formation Water Žarošice 4AAcidified Formation Water Žarošice 4A
Ca2+111.0111.58mg/L
Mg2+107.0108.03mg/L
Na+8570.08565.23mg/L
K+252.0257.45mg/L
Cl12,300.012,314.67mg/L
NH4+32.632.64mg/L
Br54.054.15mg/L
I43.042.76mg/L
HCO3/CO2(aq)2580.011,584.62mg/L
SO42−535.0528.89mg/L
Li+3.23.24mg/L
Sr2+19.218.91mg/L
Mn2+0.01.45mg/L
Fe2+0.40.28mg/L
B3+43.041.20mg/L
TDS24,590.024,590.29mg/L
Other parameters
T5353°C
pH7.14.2-
Ι0.280.25mol/L
Log PCO2 = 2.176.
Table 8. Chemical analysis of reaction solutions from experiments with reservoir and caprocks in the reaction chamber.
Table 8. Chemical analysis of reaction solutions from experiments with reservoir and caprocks in the reaction chamber.
ParameterpHρ at 15 °CCond.SO4ClHCO3NH4KMgNaCaLiSrMnFeB
Units(-)(g/cm3)(mS/cm)(mg/L)
Input—MND analysis7.11.036.5535.012,300.02580.032.6252.0107.08570.0111.03.219.20.00.4-
BoreholeDays of experiment
Reservoir
ZA4A
306.71.036.9640.013,250.62702.39.1257.099.64570285.02.817.00.90.0145.8
906.91.035.3660.012,727.63440.415.8256.0100.06260315.02.913.61.194.943.0
1207.21.037.2550.013,046.73225.719.4463.094.88750348.02.816.30.718.940.8
1806.81.038.0540.012,444.03660.019.1334.092.76770702.02.713.40.460.841.6
Caprock
ZA4
307.11.037.3380.025,455.32840.210.8279.0103.08740250.02.915.20.23.547.8
907.71.037.9500.013,188.52852.42.7229.087.28100253.02.818.50.725.047.4
1206.91.035.4550.012,266.72976.814.8289.095.25680274.0-0.10.740.538.8
1806.81.035.6600.011,664.03042.714.7215.093.65380266.0-0.10.423.539.6
Table 9. Calculated sequestration capacity.
Table 9. Calculated sequestration capacity.
SampleSequestration Capacity (kg/m3)
t = 100 dt = 10,000 y
ZA4 caprock1. Mineral0.183.01
2. In solution3.040.21
sum (1 + 2)3.223.22
ZA4A reservoir1. Mineral0.224.37
2. In solution5.281.13
sum (1 + 2)5.505.50
Table 10. Minerals responsible for mineral sequestration of CO2.
Table 10. Minerals responsible for mineral sequestration of CO2.
SampleStage IStage II
DissolutionPrecipitationDissolutionPrecipitation
ZA4 caprockCalciteDawsoniteAnkerite
Calcite
Dawsonite Dolomite
ZA4A reservoirAnkeriteDawsonite Siderite dolomiteAnkeriteSiderite dolomite
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Licbinska, M.; Labus, K.; Klempa, M.; Matysek, D.; Vasek, M. Laboratory Experiments and Geochemical Modeling of Gas–Water–Rock Interactions for a CO2 Storage Pilot Project in a Carbonate Reservoir in the Czech Republic. Minerals 2024, 14, 602. https://doi.org/10.3390/min14060602

AMA Style

Licbinska M, Labus K, Klempa M, Matysek D, Vasek M. Laboratory Experiments and Geochemical Modeling of Gas–Water–Rock Interactions for a CO2 Storage Pilot Project in a Carbonate Reservoir in the Czech Republic. Minerals. 2024; 14(6):602. https://doi.org/10.3390/min14060602

Chicago/Turabian Style

Licbinska, Monika, Krzysztof Labus, Martin Klempa, Dalibor Matysek, and Milan Vasek. 2024. "Laboratory Experiments and Geochemical Modeling of Gas–Water–Rock Interactions for a CO2 Storage Pilot Project in a Carbonate Reservoir in the Czech Republic" Minerals 14, no. 6: 602. https://doi.org/10.3390/min14060602

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop