Next Article in Journal
Microcrack Porosity Estimation Based on Rock Physics Templates: A Case Study in Sichuan Basin, China
Previous Article in Journal
Solid Digestate—Physicochemical and Thermal Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Measurement of Solubility of CO2 in NaCl, CaCl2, MgCl2 and MgCl2 + CaCl2 Brines at Temperatures from 298 to 373 K and Pressures up to 20 MPa Using the Potentiometric Titration Method

by
Bo Liu
1,
Barham Sabir Mahmood
2,
Erfan Mohammadian
1,3,*,
Abbas Khaksar Manshad
4,
Nor Roslina Rosli
5,* and
Mehdi Ostadhassan
1
1
Key Laboratory of Continental Shale Hydrocarbon Accumulation and Efficient Development, Ministry of Education, Northeast Petroleum University, Daqing 163318, China
2
Department of Petroleum, Faculty of Engineering, Koya University, Koya KOY45, Kurdistan Region—F.R., Iraq
3
Department of Petroleum and Natural Gas Engineering, Cyprus International University, Via Mersin 10, Haspolat-Nicosia 99258, Turkish Republic of Northern Cyprus, Turkey
4
Department of Petroleum Engineering, Abadan Faculty of Petroleum Engineering, Petroleum University of Technology (PUT), Abadan, Iran
5
School of Chemical Engineering, College of Engineering, Universiti Teknologi MARA, Shah Alam 40450, Selangor, Malaysia
*
Authors to whom correspondence should be addressed.
Submission received: 4 August 2021 / Revised: 12 September 2021 / Accepted: 28 September 2021 / Published: 2 November 2021

Abstract

:
Understanding the carbon dioxide (CO2) solubility in formation brines is of great importance to several industrial applications, including CO2 sequestration and some CO2 capture technologies, as well as CO2-based enhanced hydrocarbon recovery methods. Despite years of study, there are few literature data on CO2 solubility for the low salinity range. Thus, in this study, the solubility of CO2 in distilled water and aqueous ionic solutions of NaCl, MgCl2, CaCl2 and MgCl2 + CaCl2 were obtained in a low salinity range (0–15,000 ppm) at temperatures from 298–373 K and pressures up to 20 MPa using an accurate and unconventional method called potentiometric titration. An experimental data set of 553 data points was collected using this method. The results of the experiments demonstrate that increasing pressure increases the solubility of CO2 in various brines, whereas increasing temperature and salinity reduces the solubility. The role of different ions in changing the solubility is elaborated through a detailed discussion on the salting-out effect of different ionic solutions. To verify the experimental results of this research, the solubility points obtained by the potentiometric titration method were compared to some of the well-established experimental and analytical data from the literature and a very good agreement with those was obtained.

1. Introduction

It is unanimously accepted that global warming and its dire consequences have become a serious problem for the whole world. Carbon dioxide (CO2), which accounts for over 62% of all greenhouse gases, has a significant impact on global warming [1]. The primary source of CO2 emissions is anthropogenic activities, as well as deforestation due to land clearing and a number of production and resource extraction processes [2]. The use of fossil fuels as the main source of energy increases the level of CO2 in the atmosphere. According to Apadula et al. [3], CO2 concentration in the atmosphere gradually increases with the growth rate of 2.05 ± 0.03 ppm/year. The key mechanism for mitigating the greenhouse effect is to considerably lower CO2 emissions into the atmosphere. Various methods have been proposed to reduce CO2 concentration in the atmosphere, such as sequestration of CO2 in subsurface formations (mature hydrocarbon reservoirs, coal beds and aquifers), injection to oceans and CO2 capture via mineral carbonation [4]. Among these options, deep saline aquifers are seen as promising storage sites for CO2, as they can serve as large storage capacities and are common throughout the world. The technological and economic feasibility of CO2 sequestration in aquifers is proven via a number of experimental and theoretical studies. However, the detailed mechanisms of sequestration of CO2 to aquifer parts of mature hydrocarbon files and saline aquifers are not yet well established. Consequently, many uncertainties remain in terms of the efficiency of different CO2 sequestration methods, as well as the safety of the operation due to the relatively high risk of leakage. The phase behavior of CO2 in contact with the aqueous phase and the solubility of CO2 in the aqueous phase are very important for assessing the effectiveness of this method. Moreover, the influence of reservoir conditions, such as reservoir pressure and temperature, brine composition and salinity, on CO2 dissolution are some of the key factors that must be carefully evaluated when planning any CO2 sequestration project [5]. In addition, the solubility of CO2 in formation brines is of great importance for the application of various CO2-based enhanced oil recovery (EOR) techniques. Accurate measurement of the solubility of CO2 in brine helps to more accurately predict the amount of CO2 available to interact with and mobilize reservoir oil [6]. There are many experimental studies on the solubility of carbon dioxide in deionized water in the literature. The measurement of the solubility in aqueous solutions of NaCl, MgCl2 and CaCl2 is also considered in some of the previous studies [7,8]. However, experimental data for the combination of aqueous salt solutions under conditions of interest for CO2 sequestration are scarce. Table 1 provides a summary of previous experimental studies on the solubility of CO2 in various brines.
Carroll et al. [7] comprehensively studied the solubility of CO2 in water in the low-pressure range. They regressed Henry’s constant equation on the experimental dataset from their previous studies. Later, numerous experimental studies on the solubility of CO2 in pure water, aqueous solutions and seawater were carried out [9,10,11,12,13,14]. Prutton and Savage [15] carried out a detailed study on the solubility of CO2 brines saturated with CaCl2 in a wide range of thermodynamic conditions of salinity, temperature and pressure. However, their results were limited to a maximum temperature of 393 K. Malinin [16] studied the solubility of CO2 in CaCl2-saturated brines at temperatures above those encountered under conditions of interest for CO2 injection/sequestration; in addition, all data provided by this author are limited to one salt molality (1 mol.kg−1). Malinin and Saveleva [17] and Malinin and Kurovskaya [18] studied the solubility of CO2 in an aqueous solution of CaCl2 throughout a wide temperature and salinity range, but all tests were conducted at a low pressure, 4.795 MPa. Liu et al. [19] studied the solubility of CO2 in CaCl2 solutions at low temperatures (318 K) and Bastami et al. [20] studied the solubility of CO2 in CaCl2 solutions with two different salinities (1.9 and 4.8 mol.kg−1), at temperatures up to 375 K. Zhao et al. [21] investigated CO2 solubility in 0.33–2 mol.kg−1 NaCl brine at temperatures of 323, 373 and 423 K, but only at a pressure of 15 MPa. A volumetric technique was utilized to test solubility in all three studies.
Apart from the experimental approaches, various theoretical methods were used to estimate the solubility of CO2 at different conditions. Gilbert et al. [22] estimated CO2 solubility in Bravo Dome and two other brines using a different correlation. Based on Pitzer’s electrolyte theory, Shi and Mao [23] constructed a model to estimate CO2 solubility in aqueous NaCl. Venkatraman et al. [24] proposed a method for estimating the solubility of CO2 in various salts, including NaCl, CaCl2 and KCl. Menad et al. [25] implemented a neural network with a radial basis function that was improved using various optimization algorithms to determine the solubility of CO2 in brine. Mohammadian et al. [26] accurately estimated CO2 solubility in NaCl and distilled brine using a data-driven approach (extreme learning machine). At temperatures as high as 473 K and pressures as high as 50 MPa, Tong et al. [27] developed a synthetic approach based on the quantitative determination of solvent masses and the visual observation of phase transitions. In a mixed NaCl/KCl brine, CO2 solubility data are shown at 14 points of state, 36 points in CaCl2 and 38 points in MgCl2. The findings greatly broaden the range of conditions in which CO2 solubility in these brines may be determined (temperature, pressure and molality). Furthermore, the results demonstrate that CO2 solubility in CaCl2 and MgCl2 brines of the same molarity are, in fact, very close. Drummond [28] experimentally measured numerous CO2 solubilities in NaCl-saturated brines. The latter is among the most complete experimental database of solubility; however, since several presumptions were used to calculate the solubility, the accuracy of the data is dubious.
The majority of previous studies on CO2 solubility in literature assumed NaCl is the only constituent of formation brines. However, there are many brine formations around the world in which a considerable number of other salts, such as MgCl2 and CaCl2, can be found [22,27]. Furthermore, despite years of prior research on CO2 solubility in ionic liquids over a wide range of pressures, temperatures and salinity (see Table 1), there are still research gaps that need to be filled. While a wide range of salinity has been explored in the literature, evidence on solubility in the low salinity region is scarce, for example, in the range from 0 ppm to 15,000 ppm (from 0 to 0.258 mol.kg−1) brine salinity. The Sabah basin, off the coast of Peninsular Malaysia, contains such geological formations, with a mean salinity of roughly 10,000 ppm (0.17 mol.kg−1) [29]. As a result, one of the goals of this study is to develop a study on solubility applicable to the injection of CO2 into low salinity subsurface formations.
This study uses an unconventional solubility measurement method, i.e., potentiometric titration, to determine the solubility of CO2 in brine. The technique described above is typical in chemical engineering, although it is rarely employed in research on CO2 sequestration/injection into subsurface formations. Furthermore, because there is a scarcity of data in the literature on CO2 solubility in brines with low salinity, in this study, the solubility of CO2 was computed in brines saturated with NaCl, CaCl2, MgCl2 and MgCl2 + CaCl2 in the low salinity range of 0–1.5 wt.%. The salting-out effect, which is a measure of decreasing solubility by increasing salinity, is studied in depth. Although the findings of the current study are mainly aimed at CO2 sequestration projects, such as CO2 injection into subsurface formation, the results could be of importance for other applications, such as CO2 capture technologies [30], CO2 mineral carbonation [31] and food industry [32].
Table 1. Experimental data available in the literature on the solubility of CO2 in deionized water and aqueous solutions of single and mixed salts.
Table 1. Experimental data available in the literature on the solubility of CO2 in deionized water and aqueous solutions of single and mixed salts.
Temperature (K)Pressure (Mpa)Aqueous PhaseExperiment MethodRef.
298–448Up to 18Deionized waterDesigned new analytical apparatus[33]
283–363Up to 13Deionized waterDeveloped high pressure cell; cubic-plus-association and the RKSA-Infochem EOS were used to estimate CO2 solubility[34]
Single salt aqueous solutions
323–373Up to 20NaCl solutionDesigned new customized mixing unit; measuring heat of mixing of a supercritical gas was used to estimate CO2 solubility[35]
323.15–423.15Up to 15NaCl solutionNew PVT cell designed; activity coefficient
osmotic coefficients were estimated from Pitzer’s model to accurately measure CO2 solubility
[21]
323.15–423.15Up to 20NaCl solutionA simple analysis method was developed to obtain solubility points at different pressures and temperatures[36]
323.15–423.15Up to 18NaCl solutionDesigned new analytical apparatus; asymmetric ( γ φ ) approach was used to model the phase behavior of the two systems, with the
Peng–Robinson equation of state and the electrolyte NRTL solution model
[10]
303–33310–20NaCl solutionThe solubility was estimated by measuring the mass of the sample and the pressure of the dissolved gas; an equation was developed to predict CO2 fraction in solution as a function of temperature, pressure and mass fraction[10]
323–4135–40NaCl solutionHigh-pressure PVT apparatus was designed; two models were used in the Eclipse simulator—the correlations of Chang et al. and the Søreide and Whitson EoS model[37]
333.15–373.15Up to 25NaCl solutionUnconventional potentiometric titration method to determine the solubility of CO2[4]
333.15Up to 40NaCl solutionTitration method to determine the solubility of CO2[6]
308–424Up to 40CaCl2 solutionDesigned new analytical apparatus[27]
328.15–375.156.89–20.68CaCl2 solutionHigh-pressure cylinder used to measure CO2 solubility; the modified model was developed by refitting interaction parameters[20]
323–42315CaCl2 solutionA high-pressure cylinder was used to measure CO2 solubility; the fugacity-activity procedure was used for modeling and extended to take into account the effect of different types of salts on the solubility of CO2 at different temperatures, pressures and salt concentrations[21]
333.15Up to 40CaCl2 solutionTitration method to determine the solubility of CO2[6]
308–424Up to 40MgCl2 solutionDesigned new analytical apparatus[22]
Mixed salts aqueous solutions
308–408Up to 40Na+, Ca2+, Mg2+, Cl, HCO3−, Fe2+, SO42−Apparatus based on the static approach was prepared; Duan model and e PR–HV model were used to predict CO2 solubility[38]
308–328Up to 16NaCl + KCl + CaCl2High-pressure cylinder used to measure CO2 solubility; solubility was obtained from the amount of liquid sample and CO2 in the sample.[19]
308–424Up to 40CaCl2 + MgCl2Designed new analytical apparatus[27]
33229Ca2+, Mg2+, Na+, K+, Fe2+, Cl, SO42−PVT apparatus was designed; a correlation in the literature was used to predict the solubility of CO2; a simple method for determining the density of aqueous solutions of CO2 is recommended.[39]
268–2981.0–4.5NaCl + MgCl2 + MgSO4 + CaCl2 + KCl + NaHCO3 + NaBrDistilled the CO2 out of the sample, absorbed it in an excess of standard Ba(OH)2 and back-titrate the excess base[40]
RKSA, Redlich Kwong-Soave equations of state; PVT, Pressure volume temperature; NRTL, Non-random two-liquid model; PR–HV, Peng–Robinson and Huron–Vidal equation of state.

2. Materials and Methods

2.1. Materials

SIGTM provided CO2 (purity > 99.9%), which was used in all of the experiments. The brines of salinity and composition were made with distilled and deionized water (Milli-Q filter) with a resistance of 18.20 ohms. SystermTM provided NaCl, MgCl2 and CaCl2 with a mass fraction purity of 0.99. No further purification or alternation was performed on the chemicals. NaOH and HCL were used with a purity of 99.7 and were purchased from EmsureTM. The reactor was made of stainless steel with a pressure rating of 45 MPa and a temperature rating of 400 K. It also had sufficient resistivity towards corrosive materials it might have come in contact with during the experiments.

2.2. Experimental Methods

Figure 1 depicts a schematic of the experimental setup used in this investigation. A Teledyne ISCO pump, a CO2 bottle and a 0.1 L autoclave reactor with a magnetic stirrer were the key components. An electric heater was used to keep the reactor warm. A dip tube was attached to a floating piston sampler made locally and powered by a medTM syringe pump. An immersion tube was used to sample the CO2-saturated saline solution. An ISCO pump regulated the pressure in each experiment. The reactor was equipped with a temperature jacket with an accuracy of 273.25 K according to the instructions for use of the device. Despite the 0.85 cm thickness of the reactor’s base, a proper “connection” between the magnetic stirrer and the stirrer ball was accomplished, resulting in a well-mixed solution. The brine was poured into the reactor and warmed to the desired temperatures. CO2 was then supplied at the desired pressure into a warmed reactor containing 70 mL of brine. After that, the reactor’s inlet and outlet valves were closed and the solution was stirred for 3 h until it achieved equilibrium. Equilibration times have been observed to range from 10 min to 24 h in previous studies [41]. Thereafter, the bottom valve in the reactor was gently opened to minimize the pressure change in the reactor. An immersion tube was used to transport a sample of CO2-saturated brine from the reactor to the sampling chamber, which had a floating piston. As soon as the sample entered the chamber, it reacted with the 0.5 M NaOH solution, filling half of the chamber.
The sampling cylinder was half-filled with brine that was later removed to let the CO2 saturated sample from the reactor be mixed with the NaOH solution via a syringe pump. Because there was an abundance of NaOH in the solution, it dissolved all types of carbon particles and converted them to carbonates, resulting in no bubble gas [41]. The sample was then titrated with hydrochloric acid (HCl) until it reached the equivalency points. Once the endpoint of the reaction was reached, the volume of titrant was measured and the solubility of CO2 was calculated using Equation (1):
C a = C t   ×   V t   ×   N M a  
where Ca is the analyte concentration (solubility of CO2) in the brine (in mol.kg−1), Ct is the titrant concentration (in mol.L−1), Vt is the volume of titrant (mL), N is the molar ratio of analyte and reactant from a balanced chemical equation and Ma is the mass of the sample to be titrated (grams). The advantage of utilizing the solution mass rather than the solution volume Va, as has been performed in many earlier research studies, is that the mass of the solution is not affected by temperature or pressure. As a result, the calculation of solubility is less ambiguous. The titrant was (0.5 M) HCl, which was utilized to react in a 5 mL sample. The pH of the sample was determined as a function of the titrant volume given and the titration was maintained until the pH reached values below 2; plotting the derivative of HCl volume versus pH yielded equivalence points. In addition to the technical simplicity, another benefit of this method compared to previous methods by which the solubility is measured is that the preservation of the samples inhibits the loss of CO2 because of the degassing during the depressurization phase. Moreover, in contrast to previous research studies, the potentiometric method, unlike a number of previous methods, does not depend on any additional parameters, such as fugacity, density, or volatility, to be able to estimate solubility accurately.
To ensure repeatability and accuracy of the results, several experiments were repeated 3 times. The errors in the measurements were found to be 0.9–7.8%. As expected, the highest error occurred with the measurements near atmospheric pressure (0.1 MPa), regardless of the temperature, salinity and type of brine. The errors were markedly lower when the pressure in the experiments exceeded 0.2 MPa. The reason for this phenomenon was that the very low values of CO2 solubilities close to atmospheric pressure were in the order of a thousandth of mol/kg, as compared to solubility values at higher pressures. Therefore, these low values could not be accurately detected using the experimental method used in this study. It is noteworthy that the focus of our study was the solubility of CO2 in subsurface formations; in those scenarios, CO2 is often injected at high pressures, hence it is in liquid or supercritical fluid state [42]. Therefore, near-atmospheric measurements are of a little significance for the aforementioned applications.

3. Results

Despite the fact that there have been several investigations on CO2 solubility in various liquids, evidence in the low salinity range is limited. As a result, the impact of pressure change on CO2 solubility was investigated under a variety of conditions in the current study. The experiments were carried out at pressures ranging from 1 to 20 MPa and temperatures ranging from 298 to 373 K. Furthermore, the experiments were carried out in a saline solution of different values of salinity (0–15,000 ppm) to confirm the reliability of the results under conditions more representative of CO2 injection to subsurface formations (aquifers and hydrocarbon reservoirs). Likewise, the solubility of CO2 in distilled water was tested under identical temperature and pressure conditions (1–20 MPa, 298–373 K). The outliers in the solubility databank, i.e., data with unusually high or low values, were identified through analyzing the z-score of the data points. The data points with unusually high or low values were treated as outliers and hence removed from the database using the z-score method that was applied with SPSS 18TM. The abnormal solubility data points were mostly caused by the rapid opening of the sampling valve, which resulted in a significant pressure decrease in the solution and subsequent supercritical CO2 breakthrough.

3.1. Impacts of Pressure and Temperature on the Solubility

CO2 solubility in distilled water, NaCl, MgCl2, CaCl2 and MgCl2 + CaCl2 is shown in Figure 2 in four different temperature series, namely, 298, 333, 353 and 373 K, respectively. It can be seen, from the figures, that increasing the pressure increased the solubility of CO2 in the brine, irrespective of the type of ionic solution and temperature. In addition, it is apparent, from the figures, that the pressure dependence of carbon dioxide decreased with increasing the pressure in all temperature series. However, in this study, the points at which the solubility would become entirely unresponsive to pressure were not observed. The same effect (pressure insensitivity at higher pressures) was reported for pressure around 30 MPa by previous researchers who used different solubility measurement methods at different temperatures, pressure and salinities [11,27]; however, as the highest point of solubility measurement in this study was 20 MPa, the point of pressure insensitivity was not observed.
The effect of pressure on solubility can be expressed using Henry’s law of solubility (partial pressures) [43]. According to Henry’s Law, the partial pressure of the gas above the solution determines the solubility of the gas in the water. Because the concentration of molecules in the gas phase increases as pressure increases, the concentration of dissolved gas molecules in the solution at equilibrium also increases. When a gas is introduced to a system that is primarily made up of brine (solvent), some of the gas molecules collide with the liquid’s surface and dissolve. When the concentration of dissolved gas molecules rises to the point where the rate at which gas molecules escape into the gas phase equals the rate at which it dissolves, dynamic equilibrium is reached. As the gas pressure rises, the amount of gas molecules per unit volume rises, increasing the rate at which gas molecules collide with the liquid’s surface and dissolve. The concentration of dissolved gas rises as more gas molecules dissolve at higher pressures, until a new dynamic equilibrium is reached [43].
When it comes to the effect of temperature, it can be seen that, as the temperature rose, the solubility decreased. CO2 solubility in NaCl is 1.421 mol kg−1 at 10.33 MPa and 298 K, whereas it is 1.037 mol kg−1, 0.877 mol kg−1 and 0.788 mol kg−1 at the same pressure and 353 K, 353 K and 373 K, respectively. In other words, there is a decrease in solubility of 27%, 38% and 44.54% as the temperature rises to 333 K, 353 K and 373 K from the initial value of 298 K. Previous researchers have also reported a decrease in the solubility with the increase in the temperature [12,44,45]. Le Chatelier’s law could explain the reduction in solubility at higher temperatures. CO2 dissolves in brine due to the interactions of the molecules of solute with those of solvents. The process of dissolving CO2 in brine is exothermic (ΔH reaction < 0), which implies that heat is produced as new attractive contacts arise as a result of the dissolution process [35]. According to the principle of Le Chatelier, if the system is heated, since this is an exothermic reaction, the system shifts towards the reactant’s side to neutralize the applied stress, which is the rise in the temperature. The kinetic energy of a system increases as the temperature rises. As the temperature rises, this causes a more rapid motion among the molecules and the breakage of intermolecular bonds, allowing molecules to escape to the gas phase from the solution [46]. As a result, independently from pressure, type of ion in the brine, or brine salinity, increasing the temperature decreased the solubility in this set of experiments.

3.2. Effects of Salinity on CO2 Solubility

The solubility of CO2 versus pressures at 0, 1000, 10,000 and 15,000 ppm in NaCl, MgCl2, CaCl2 and MgCl2 + CaCl2 solutions at 298 K is shown in Figure 3. Figure 3 shows the reduction in CO2 solubility in formation brine as the salinity increased at various pressures and temperatures for all types of aqueous solutions employed in the current study. The solubility of CO2 decreased by 1% with an increase in the salinity from 0 to 1000 ppm for brine solutions in the experiments conducted in this work, while a decrease in the solubility of 3–6% was found with a factor of 10 increase in brine concentration (from 1000 to 10,000 ppm). Moreover, increasing the brine concentration from 10,000 to 15,000 ppm resulted in a 4–5% decrease in CO2 solubility. The range of reduction in solubility as the salinity increased is in line with those reported in the literature when pressure and temperature were set in the same range of this study [6,25,47].
The decrease in the solubility can be explained by the fact that when salt ions such as NaCl, MgCl2 and CaCl2 are added to water, they bind water molecules to “solvates”, leaving less water for CO2 to adhere to. In other words, the presence of water molecules in the solvation of ions significantly lowers CO2 molecules’ weak attraction to water/brine and displaces dissolved CO2 from polar water. When solutes such as NaCl, MgCl2 and CaCl2 (or any combination of them) are present, the solubility of CO2 in brine is greatly impacted. In reality, because of the enhanced salting-out effect, the solubility reduces as the salinity rises (the salting-out effect is discussed in detail in the next part of the results). Similar results can be observed in the previous studies in which different solubility measurement methods (such as depressurization) were used to measure CO2 solubility in brines which were significantly more saline than the brine used in this study [37,45].

3.3. Salting-Out Effect

As the concentration of dissolved solids in the brine rises, the salting-out effect reduces CO2 solubility in aqueous solutions (in this case, brine). The effect is significant because it aids in quantifying the decrease in CO2 solubility as salinity rises. Studies on the hydration of ions and the interaction of ions with water molecules have shown that, at a high density, smaller ions tend to bind the molecules of water more effectively, while larger ions with a low charge density bind the water molecules weakly [48,49]. Therefore, high charge density ions have a robust impact on the structure of the water, which governs the ability of the brine to dissolve higher amounts of CO2. The experiments show that, if two single-salt aqueous solutions have the same electrolyte type and share the same anion (e.g., Cl), the cation with a higher charge density (smaller radius and grater charge) has a greater salting-out effect on dissolved CO2 than the cation with a lower charge density (larger radius and lower charge). For instance, Mg2+ has a charge density that is a little higher than that of Ca2+ (they have the same charge number, but Mg2+ has a smaller radius than Ca2+) [50]; hence, the amount of CO2 dissolved in aqueous MgCl2 is less than that in aqueous CaCl2 at the same ionic strength. The solubility of CO2 in an aqueous solution of NaCl follows a similar pattern. Because Na+ has a lower charge density than Mg2+, CO2 is substantially more soluble in aqueous NaCl solutions than in aqueous MgCl2 solutions [21].
Figure 4 depicts the salting out effect in NaCl, MgCl2, CaCl2 and MgCl2 + CaCl2 solutions with concentrations ranging from 1000 to 15,000 ppm, at pressures ranging from 1 to 20 MPa and temperatures of 298 K. Equation (2) was used to calculate the percentage of salting-out effect (S-O%).
S O   % = x Dw x b x Dw 100
where S-O (%) is the salting out percentage, xDw is CO2 solubility in distilled water and xb is CO2 solubility in any brine. Figure 4 shows that, at a 1000 ppm salinity, there was no substantial change in solubility. As a result, in low-salinity brine, the salting-out effect is insignificant. The salting-out effect, on the other hand, increased as the concentration of solids in the brine rose. The maximum percentage of S-O is found in the 15,000-ppm data series, where the effect reached 6–9% for all solutions at lower pressures, whereas the lowest percentage of S-O is found in brine with a concentration of 1000 ppm, where S-O fluctuated between 0.25% and 0.65%. This result is in line with the findings of Tong et al. [27], who found that increasing the temperature enhances the salting-out effect, while increasing the pressure tends to diminish it.

3.4. Comparison of Experimental Results with Previous Studies

The CO2 solubility data points obtained using the potentiometric titration method were compared with literature data obtained well-established methods under similar conditions to examine the accuracy of the findings of the experimental method used in this study, even though the data points obtained under the same conditions of pressure, temperature and, specifically, salinity (low salinity range) were very limited. As shown in Figure 5, the experimental results of this study are in good agreement with those obtained in previous studies using more common solubility measurement methods, such as depressurization or combinations [37] and depressurization [39] methods. In Figure 5, the solid lines indicate a regression line drawn based on the data obtained in the current study and error bars indicate a 5% difference from the experimental data of this study.
The data points obtained by Duan and Sun [12] and Li et al. [39] were measured at 323 K and 332 K, respectively. The difference among the results obtained in this study, Duan and Sun’s and Lee et al.’s, is most likely due to the difference in the experimental temperature in our study, which was 333 K vs. 323 K and 332 K in Duan and Sun and Li et al., respectively. Furthermore, as shown in Figure 5, comparing the measurement data from this investigation with the data from Cruz et al. [6] supports the accuracy of the solubility points obtained in this work. As a result, the potentiometric titration method produces reliable results and could be used as an alternative to some of the previous complex and often expensive methods required to accurately measure CO2 solubility in a condition representative of CO2 injection/sequestration to subsurface geological formations, such as aquifers and mature hydrocarbon fields.
There is a scarcity of experimental data in a similar range of pressure, temperature, salinity and composition, making it impossible to compare the findings of this work with other studies. For two MgCl2 and CaCl2 brines, Figure 6 shows a comparison of the solubility results from this study with solubility estimates from Duan and Sun’s (2003) theoretical model. Figure 7 also illustrates the parity-plot at the same conditions of pressure, temperature, salinity and brine type. The solubility values between the two models are almost perfectly in agreement in both brines, with a coefficient of correlation of more than 99% (R2 > 0.99).

3.5. Field Implications and Recommendations for Further Studies

The impacts of pressure, temperature, salinity and brine composition are elaborated in the previous sections of this study. In geological formations, the salinity increases with depth. Since the solubility reduces with the increase in the salinity (irrespective of the type of salt), it can therefore be concluded that, in deeper geological formations, the contribution of the solubility mechanism weakens. Therefore, from point of view of CO2 solubility mechanisms, shallower formations, or depositories with low salinities (such as hydrocarbon fields in the Sabah basin, offshore Sabah, Malaysia) would be more suitable for CO2 sequestration, as larger amounts of CO2 can be rendered immobile using the solubility mechanism. With regards to temperature, as shown in this study, the increase in temperature reduces the solubility of CO2 in brine. Hence, it could be concluded, although only from the point of view of pressure, that deeper formations are more suitable, as the effects of temperature and salinity render the solubility mechanism less effective. The effect of pressure, on the other hand, is favorable on the solubility mechanisms. Hence, quantification of the impact of each parameter on the solubility needs to be conducted to find the optimum depth for sequestration. On the other hand, for enhanced oil recovery (EOR) applications, lower solubility of CO2 in formation brines is more favorable for CO2-based EOR methods, as the amount of available CO2 to interact with and eventually mobilize the residual oil is higher. Hence, it is recommended, for future studies, to address this issue by proposing a selection criterion for CO2 sequestration, CO2 injection and a combination of the two.
Moreover, in most studies focusing on CO2 sequestration, CO2 is considered to be 100% pure. This, however, is not often the case for practical scenarios in which CO2 could be stemmed from different industries (such as gasification, post-combustion CO2 capture, sour gas processing, or even recycled CO2 from EOR operations) [51]. Moreover, it is not uncommon for natural CO2 from subsurface formation to have associated gases. The CO2 stream may contain several impurities, such as H2S, N2, Ar, etc. [51], and it might be economically and technically viable to consider CO2 injection with no further purification. Therefore, it would be of great industrial importance to be able to study the solubility of CO2 stream containing impurities, so that a more realistic estimation of the solubility mechanisms and, ultimately, sequestration efficiency can be made.

4. Conclusions

Using potentiometric titration, 553 data points of CO2 solubility in various brines (NaCl, MgCl2, CaCl2 and MgCl2 + CaCl2) were obtained at temperatures ranging from 298 to 373 K and pressures up to 20 MPa. In comparison with earlier traditional procedures, the new method is shown to be reproducible and accurate. The findings are in good accord with those of prior studies at the same pressure, temperature, salinity and brine composition ranges (salt type). In terms of pressure, independently from the salinity of the brine composition, it is obvious that increasing the pressure increases CO2 solubility in aqueous solutions. However, the pressure dependence of solubility decreased with the increase in the pressure, although, in this study, the point at which the solubility becomes completely independent from the pressure was not detected. The temperature has a reverse effect on the solubility; as the temperature increased, the solubility was significantly reduced, regardless of the salinity and the composition of the brine. Finally, increasing salinity also negatively affects solubility, though, at a low salinity, the effect was not so noticeable. However, the solubility decreased by over 6% as the salinity of brine increased from 0 to 15,000 ppm. Moreover, the effect of the presence of divalent and monovalent ions in brine is here discussed in detail with reference to the salting-out effect. Under the same conditions of pressure and temperature, both ion charge and ion radius were found to be influential factors in the solubility of CO2 in brine. Lastly, the comparison of the experimental results obtained from the potentiometric titration method, as an unconventional method, with those of more conventional and well-established methods from the literature proved the method to be accurate and reliable.

Supplementary Materials

Author Contributions

Conceptualization, formal analysis, methodology, project administration, supervision, visualization and writing (original draft, review and editing), E.M.; data curation, formal analysis, visualization and software, B.S.M., A.K.M. and B.L.; supervision, validation, writing (review and editing), N.R.R. and M.O. All authors have read and agreed to the published version of the manuscript.

Funding

Fundamental Research Grant Scheme 600-IRMI/FRGS 5/3 (085/2019).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article or Supplementary Materials.

Acknowledgments

This research was made available thanks to the financial support provided by the Ministry of Higher Education through Fundamental Research Grant Scheme 600-IRMI/FRGS 5/3 (085/2019).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Stocker, T.F.; Qin, D.; Plattner, G.-K.; Tignor, M.; Allen, S.K.; Boschung, J.; Nauels, A.; Xia, Y.; Bex, V.; IPCC; et al. Climate Change 2013: The Physical Science Basis; Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change; Cambridge University Press: Cambridge, UK; New York, NY, USA, 2013; p. 1535. [Google Scholar] [CrossRef] [Green Version]
  2. Orr, F.M. Storage of carbon dioxide in geologic formations. J. Pet. Technol. 2004, 56, 90–97. [Google Scholar] [CrossRef]
  3. Apadula, F.; Cassardo, C.; Ferrarese, S.; Heltai, D.; Lanza, A. Thirty Years of Atmospheric CO2 Observations at the Plateau Rosa Station, Italy. Atmosphere 2019, 10, 418. [Google Scholar] [CrossRef] [Green Version]
  4. Mohammadian, E.; Hamidi, H.; Asadullah, M.; Azdarpour, A.; Motamedi, S.; Junin, R. Measurement of CO2 Solubility in NaCl Brine Solutions at Different Temperatures and Pressures Using the Potentiometric Titration Method. J. Chem. Eng. Data 2015, 60, 2042–2049. [Google Scholar] [CrossRef]
  5. Ahmadi, P.; Chapoy, A. CO2 solubility in formation water under sequestration conditions. Fluid Phase Equilibr. 2018, 463, 80–90. [Google Scholar] [CrossRef]
  6. Cruz, J.L.; Contamine, F.; Cézac, P. Experimental CO2 Solubility in Nacl-Cacl2 Brines At 333.15 and 453.15 K Up to 40 Mpa. In Proceedings of the 1st Geoscience & Engineering in Energy Transition Conference, Strasbourg, France, 16–18 November 2020; Volume 2020, pp. 1–5. [Google Scholar] [CrossRef]
  7. Carroll, J.J.; Slupsky, J.D.; Mather, A.E. The Solubility of Carbon-Dioxide in Water at Low-Pressure. J. Phys. Chem. Ref. Data 1991, 20, 1201–1209. [Google Scholar] [CrossRef]
  8. Zheng, D.-Q.; Guo, T.-M.; Knapp, H. Experimental and modeling studies on the solubility of CO2, CHC1F2, CHF3, C2H2F4 and C2H4F2 in water and aqueous NaCl solutions under low pressures. Fluid Phase Equilibr. 1997, 129, 197–209. [Google Scholar] [CrossRef]
  9. Kiepe, J.; Horstmann, S.; Fischer, K.; Gmehling, J. Experimental determination and prediction of gas solubility data for CO2 + H2O mixtures containing NaCl or KCl at temperatures between 313 and 393 K and pressures up to 10 MPa. Ind. Eng. Chem Res. 2002, 41, 4393–4398. [Google Scholar] [CrossRef]
  10. Bando, S.; Takemura, F.; Nishio, M.; Hihara, E.; Akai, M. Solubility of CO2 in aqueous solutions of NaCl at (30 to 60) degrees C and (10 to 20) MPa. J. Chem. Eng. Data 2003, 48, 576–579. [Google Scholar] [CrossRef]
  11. Duan, Z.H.; Sun, R. An improved model calculating CO2 solubility in pure water and aqueous NaCl solutions from 273 to 533 K and from 0 to 2000 bar. Chem. Geol. 2003, 193, 257–271. [Google Scholar] [CrossRef]
  12. 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]
  13. Bermejo, M.D.; Martin, A.; Florusse, L.J.; Peters, C.J.; Cocero, M.J. The influence of Na2SO4 on the CO2 solubility in water at high pressure. Fluid Phase Equilibr. 2005, 238, 220–228. [Google Scholar] [CrossRef]
  14. Chapoy, A.; Mohammadi, A.H.; Chareton, A.; Tohidi, B.; Richon, D. Measurement and modeling of gas solubility and literature review of the properties for the carbon dioxide-water system. Ind. Eng. Chem. Res. 2004, 43, 1794–1802. [Google Scholar] [CrossRef]
  15. Prutton, C.F.; Savage, R.L. The Solubility of Carbon Dioxide in Calcium Chloride-Water Solutions at 75, 100, 120° and High Pressures. J. Am. Chem. Soc. 1945, 67, 1550–1554. [Google Scholar] [CrossRef]
  16. Malinin, S.D. The system water-carbon dioxide at high temperature and pressures. Geokhimiya 1959, 3, 292–306. [Google Scholar]
  17. Malinin, S.D.; Savelyeva, N.I. The solubility of CO2 in NaCl and CaCl2 solutions at 25, 50 and 75 °C under elevated CO2 pressures. Geokhimiya 1972, 6, 643–653. [Google Scholar]
  18. Malinin, S.D.; Kurovskaya, N.A. Solubility of CO2 in chlorides solutions at elevated temperatures and CO2 pressures. Geochem. Int. 1975, 12, 199–201. [Google Scholar]
  19. Liu, Y.H.; Hou, M.Q.; Yang, G.Y.; Han, B.X. Solubility of CO2 in aqueous solutions of NaCl, KCl, CaCl2 and their mixed salts at different temperatures and pressures. J. Supercrit. Fluid 2011, 56, 125–129. [Google Scholar] [CrossRef]
  20. Bastami, A.; Allahgholi, M.; Pourafshary, P. Experimental and modelling study of the solubility of CO2 in various CaCl2 solutions at different temperatures and pressures. Pet. Sci. 2014, 11, 569–577. [Google Scholar] [CrossRef] [Green Version]
  21. Zhao, H.; Fedkin, M.V.; Dilmore, R.M.; Lvov, S.N. Carbon dioxide solubility in aqueous solutions of sodium chloride at geological conditions: Experimental results at 323.15, 373.15, and 423.15K and 150 bar and modeling up to 573.15 K and 2000 bar. Geochim. Cosmochim. Acta 2015, 149, 165–189. [Google Scholar] [CrossRef] [Green Version]
  22. Gilbert, K.; Bennett, P.C.; Wolfe, W.; Zhang, T.; Romanak, K.D. CO2 solubility in aqueous solutions containing Na+, Ca2+, Cl, SO42− and HCO3-: The effects of electrostricted water and ion hydration thermodynamics. Appl. Geochem. 2016, 67, 59–67. [Google Scholar] [CrossRef] [Green Version]
  23. Shi, X.L.; Mao, S.D. An improved model for CO2 solubility in aqueous electrolyte solution containing Na+, K+, Mg2+, Ca2+, Cl- and SO42- under conditions of CO2 capture and sequestration. Chem. Geol. 2017, 463, 12–28. [Google Scholar] [CrossRef]
  24. Venkatraman, A.; Argüelles-Vivas, F.J.; Okuno, R.; Singh, G.; Lake, L.W.; Wheeler, M.F. Modeling Impact of Aqueous Ions on solubility of CO2 and its Implications for Sequestration. In Proceedings of the SPE Annual Technical Conference and Exhibition, Dubai, United Arab Emirates, 26–28 September 2016. [Google Scholar]
  25. Menad, N.A.; Hemmati-Sarapardeh, A.; Varamesh, A.; Shamshirband, S. Predicting solubility of CO2 in brine by advanced machine learning systems: Application to carbon capture and sequestration. J. CO2 Util. 2019, 33, 83–95. [Google Scholar] [CrossRef]
  26. Mohammadian, E.; Motamedi, S.; Shamshirband, S.; Hashim, R.; Junin, R.; Roy, C.; Azdarpour, A. Application of extreme learning machine for prediction of aqueous solubility of carbon dioxide. Environ. Earth Sci. 2016, 75, 215. [Google Scholar] [CrossRef]
  27. Tong, D.L.; Trusler, J.P.M.; Vega-Maza, D. Solubility of CO2 in Aqueous Solutions of CaCl2 or MgCl2 and in a Synthetic Formation Brine at Temperatures up to 423 K and Pressures up to 40 MPa. J. Chem. Eng. Data 2013, 58, 2116–2124. [Google Scholar] [CrossRef]
  28. Drummond, S.E. Boiling and Mixing of Hydrothermal Fluids: Chemical Effects on Mineral Precipitation. Ph.D. Thesis, Pennsylvania State University, University Park, PA, USA, 1981. Available online: https://www.worldcat.org/title/boiling-and-mixing-of-hydrothermal-fluids-chemical-effects-on-mineral-precipitation/oclc/8343728/editions?referer=di&editionsView=true (accessed on 11 September 2021).
  29. Heavysege, R.G. Formation Evaluation Of Fresh Water Shaly Sands Of The Malay Basin, Offshore Malaysia. In Proceedings of the SPWLA 43rd Annual Logging Symposium, Oiso, Japan, 2–5 June 2002. [Google Scholar]
  30. Ramdin, M.; de Loos, T.W.; Vlugt, T.J.H. State-of-the-Art of CO2 Capture with Ionic Liquids. Ind. Eng. Chem. Res. 2012, 51, 8149–8177. [Google Scholar] [CrossRef]
  31. Azdarpour, A.; Asadullah, M.; Mohammadian, E.; Hamidi, H.; Junin, R.; Karaei, M.A. A review on carbon dioxide mineral carbonation through pH-swing process. Chem. Eng. J. 2015, 279, 615–630. [Google Scholar] [CrossRef]
  32. Illera, A.E.; Sanz, M.T.; Beltrán, S.; Melgosa, R. High pressure CO2 solubility in food model solutions and fruit juices. J. Supercrit. Fluids 2019, 143, 120–125. [Google Scholar] [CrossRef]
  33. Hou, S.X.; Maitland, G.C.; Trusler, J.P.M. Measurement and modeling of the phase behavior of the (carbon dioxide plus water) mixture at temperatures from 298.15 K to 448.15 K. J. Supercrit. Fluid 2013, 73, 87–96. [Google Scholar] [CrossRef] [Green Version]
  34. Carvalho, P.J.; Pereira, L.M.C.; Goncalves, N.P.F.; Queimada, A.J.; Coutinho, J.A.P. Carbon dioxide solubility in aqueous solutions of NaCl: Measurements and modeling with electrolyte equations of state. Fluid Phase Equilibr. 2015, 388, 100–106. [Google Scholar] [CrossRef]
  35. Koschel, D.; Coxam, J.Y.; Rodier, L.; Majer, V. Enthalpy and solubility data of CO2 in water and NaCl(aq) at conditions of interest for geological sequestration. Fluid Phase Equilibr. 2006, 247, 107–120. [Google Scholar] [CrossRef]
  36. Messabeb, H.; Contamine, F.; Cézac, P.; Serin, J.P.; Gaucher, E.C. Experimental Measurement of CO2 Solubility in Aqueous NaCl Solution at Temperature from 323.15 to 423.15 K and Pressure of up to 20 MPa. J. Chem. Eng. Data 2016, 61, 3573–3584. [Google Scholar] [CrossRef]
  37. Yan, W.; Huang, S.L.; Stenby, E.H. Measurement and modeling of CO2 solubility in NaCl brine and CO2-saturated NaCl brine density. Int. J. Greenh. Gas. Control. 2011, 5, 1460–1477. [Google Scholar] [CrossRef]
  38. Tang, Y.; Bian, X.Q.; Du, Z.M.; Wang, C.Q. Measurement and prediction model of carbon dioxide solubility in aqueous solutions containing bicarbonate anion. Fluid Phase Equilibr. 2015, 386, 56–64. [Google Scholar] [CrossRef]
  39. Li, Z.W.; Dong, M.Z.; Li, S.L.; Dai, L.M. Densities and solubilities for binary systems of carbon dioxide plus water and carbon dioxide plus brine at 59 degrees C and pressures to 29 MPa. J. Chem. Eng. Data 2004, 49, 1026–1031. [Google Scholar] [CrossRef]
  40. Stewart, P.B.; Munjal, P.K. Solubility of carbon dioxide in pure water, synthetic sea water, and synthetic sea water concentrates at -5.deg. to 25.deg. and 10- to 45-atm. pressure. J. Chem. Eng. Data 1970, 15, 67–71. [Google Scholar] [CrossRef]
  41. Portier, S.; Rochelle, C. Modelling CO2 solubility in pure water and NaCl-type waters from 0 to 300 °C and from 1 to 300 bar: Application to the Utsira Formation at Sleipner. Chem. Geol. 2005, 217, 187–199. [Google Scholar] [CrossRef]
  42. Gilmore, K.A.; Neufeld, J.A.; Bickle, M.J. CO2 Dissolution Trapping Rates in Heterogeneous Porous Media. Geophys. Res. Lett. 2020, 47, e2020GL087001. [Google Scholar] [CrossRef]
  43. Averill, B. Chemistry: Principles, Patterns, and Applications with Student Access Kit for Mastering General Chemistry, 3rd ed.; Prentice Hall: London, UK, 2007. [Google Scholar]
  44. El-Maghraby, R.M.; Pentland, C.H.; Iglauer, S.; Blunt, M.J. A fast method to equilibrate carbon dioxide with brine at high pressure and elevated temperature including solubility measurements. J. Supercrit. Fluid 2012, 62, 55–59. [Google Scholar] [CrossRef] [Green Version]
  45. Wang, L.; Shen, Z.L.; Hu, L.S.; Yu, Q.C. Modeling and measurement of CO2 solubility in salty aqueous solutions and application in the Erdos Basin. Fluid Phase Equilibr. 2014, 377, 45–55. [Google Scholar] [CrossRef]
  46. Zumdahl, S.S.; DeCoste, D.J. Chemical Principles; Cengage Learning: Boston, MA, USA, 2017. [Google Scholar]
  47. Chabab, S.; Théveneau, P.; Corvisier, J.; Coquelet, C.; Paricaud, P.; Houriez, C.; Ahmar, E.E. Thermodynamic study of the CO2—H2O—NaCl system: Measurements of CO2 solubility and modeling of phase equilibria using Soreide and Whitson, electrolyte CPA and SIT models. Int. J. Greenh. Gas. Control. 2019, 91, 102825. [Google Scholar] [CrossRef]
  48. Samoilov, O.Y. A new approach to the study of hydration of ions in aqueous solutions. Discuss. Faraday Soc. 1957, 24, 141–146. [Google Scholar] [CrossRef]
  49. Collins, K.D. Charge density-dependent strength of hydration and biological structure. Biophys J. 1997, 72, 65–76. [Google Scholar] [CrossRef] [Green Version]
  50. Marcus, Y. Ionic-Radii in Aqueous-Solutions. Chem Rev. 1988, 88, 1475–1498. [Google Scholar] [CrossRef]
  51. Wang, J.; Ryan, D.; Anthony, E.J.; Wildgust, N.; Aiken, T. Effects of impurities on CO2 transport, injection and storage. Energy Procedia 2011, 4, 3071–3078. [Google Scholar] [CrossRef] [Green Version]
Figure 1. The experimental setup used for solubility measurements.
Figure 1. The experimental setup used for solubility measurements.
Energies 14 07222 g001
Figure 2. Solubility of carbon dioxide in (●)distilled water and 1000 ppm of (■) NaCl, (●) MgCl2, (▲) CaCl2 and (●) MgCl2 + CaCl2 at (a) 298 °K, (b) 333 °K, (d) 353 °K and (d) 373 °K versus pressure.
Figure 2. Solubility of carbon dioxide in (●)distilled water and 1000 ppm of (■) NaCl, (●) MgCl2, (▲) CaCl2 and (●) MgCl2 + CaCl2 at (a) 298 °K, (b) 333 °K, (d) 353 °K and (d) 373 °K versus pressure.
Energies 14 07222 g002
Figure 3. Solubility of carbon dioxide in distilled water and various salinity of (■) NaCl, (●) MgCl2, (▲) CaCl2 and (●) MgCl2 + CaCl2 at 298 °K versus pressure.
Figure 3. Solubility of carbon dioxide in distilled water and various salinity of (■) NaCl, (●) MgCl2, (▲) CaCl2 and (●) MgCl2 + CaCl2 at 298 °K versus pressure.
Energies 14 07222 g003
Figure 4. Salting-out effect of NaCl, MgCl2, CaCl2 and MgCl2 + CaCl2 brine of (■) 1000 ppm, (▲) 10,000 ppm and (●) 15,000 ppm at 298 °K versus pressure.
Figure 4. Salting-out effect of NaCl, MgCl2, CaCl2 and MgCl2 + CaCl2 brine of (■) 1000 ppm, (▲) 10,000 ppm and (●) 15,000 ppm at 298 °K versus pressure.
Energies 14 07222 g004
Figure 5. Comparison of the results of this study for CO2 solubility (●) with the results of Lara et al. [6] (■), Duan and Sun [11] (▲) and Li et al. [39] (●) in distilled water (zero salinity) at 333 °K. The solid line represents the regression line fitted from the current solubility study.
Figure 5. Comparison of the results of this study for CO2 solubility (●) with the results of Lara et al. [6] (■), Duan and Sun [11] (▲) and Li et al. [39] (●) in distilled water (zero salinity) at 333 °K. The solid line represents the regression line fitted from the current solubility study.
Energies 14 07222 g005
Figure 6. Comparison of Solubility of CO2 in (a) CaCl2 brine and (b) MgCl2 at 298 °K.
Figure 6. Comparison of Solubility of CO2 in (a) CaCl2 brine and (b) MgCl2 at 298 °K.
Energies 14 07222 g006
Figure 7. Parity-plots of solubility values obtained for CaCl2 (a) and MgCl2 (b) brines of this study and those obtained from Duan and Sun (2003) models.
Figure 7. Parity-plots of solubility values obtained for CaCl2 (a) and MgCl2 (b) brines of this study and those obtained from Duan and Sun (2003) models.
Energies 14 07222 g007
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liu, B.; Mahmood, B.S.; Mohammadian, E.; Khaksar Manshad, A.; Rosli, N.R.; Ostadhassan, M. Measurement of Solubility of CO2 in NaCl, CaCl2, MgCl2 and MgCl2 + CaCl2 Brines at Temperatures from 298 to 373 K and Pressures up to 20 MPa Using the Potentiometric Titration Method. Energies 2021, 14, 7222. https://0-doi-org.brum.beds.ac.uk/10.3390/en14217222

AMA Style

Liu B, Mahmood BS, Mohammadian E, Khaksar Manshad A, Rosli NR, Ostadhassan M. Measurement of Solubility of CO2 in NaCl, CaCl2, MgCl2 and MgCl2 + CaCl2 Brines at Temperatures from 298 to 373 K and Pressures up to 20 MPa Using the Potentiometric Titration Method. Energies. 2021; 14(21):7222. https://0-doi-org.brum.beds.ac.uk/10.3390/en14217222

Chicago/Turabian Style

Liu, Bo, Barham Sabir Mahmood, Erfan Mohammadian, Abbas Khaksar Manshad, Nor Roslina Rosli, and Mehdi Ostadhassan. 2021. "Measurement of Solubility of CO2 in NaCl, CaCl2, MgCl2 and MgCl2 + CaCl2 Brines at Temperatures from 298 to 373 K and Pressures up to 20 MPa Using the Potentiometric Titration Method" Energies 14, no. 21: 7222. https://0-doi-org.brum.beds.ac.uk/10.3390/en14217222

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