Mixtures at 323.2 K and 9.000 MPa - ACS Publications - American

Apr 7, 2017 - Diana H. Bacon,. †. Ronald D. Springer,. ‡. Andrzej Anderko,. ‡ ... Bernard Peter McGrail,. †. Kevin M. Rosso,. † and Herbert ...
0 downloads 0 Views 973KB Size
Article pubs.acs.org/jced

Water Solubility at Saturation for CO2−CH4 Mixtures at 323.2 K and 9.000 MPa John S. Loring,*,† Diana H. Bacon,† Ronald D. Springer,‡ Andrzej Anderko,‡ Smitha Gopinath,§ Catherine M. Yonkofski,† Christopher J. Thompson,† Bernard Peter McGrail,† Kevin M. Rosso,† and Herbert T. Schaef† †

Pacific Northwest National Laboratory, Richland, Washington 99352, United States OLI Systems Inc., 240 Cedar Knolls Road, Suite 301, Morris Plains, New Jersey 07927, United States § Department of Chemical Engineering, Imperial College London, London SW7 2AZ, U.K. ‡

S Supporting Information *

ABSTRACT: The concentration of H2O dissolved in CO2−CH4 supercritical fluids is an important parameter that can control shale permeability, affect CH4 transmissivity, and ultimately impact the efficiency of CO2 enhanced gas recovery (EGR) operations. Here, we use in situ high-pressure infrared (IR) spectroscopic titrations to quantify the solubility of H2O in six CO2−CH4 mixtures, ranging from pure CO2 to pure CH4, at shallow shale reservoir conditions of 323.2 K and 9.000 MPa. Measured concentrations of H2O at saturation increase with increasing mole percent CO2, and our results are in agreement with limited data available in the literature. We use these experimental results to benchmark three current thermodynamic multiphase routines: the mixed-solvent electrolyte (MSE), STOMP-COMP, and the Statistical Associating Fluid Theory for variable range Mie potentials (SAFT γ-Mie) equations of state. Of these models, MSE and STOMP-COMP accurately predict maximum H2O solubilities of the binary CO2−H2O and CH4−H2O systems, and they also reproduce the shape of the water solubility curve as a function of mole percent CO2. Hence, these routines should work well to predict H2O content in reservoir simulations and help to make informed decisions concerning injection strategies for CO2 EGR in shale plays at shallow depths.

1. INTRODUCTION Natural gas production in the United States has increased from 1.6 to 2.0 billion cubic meters per day over the past decade, and continued growth is projected.1 Gas is a cleaner and loweremission fuel than coal, but any type of fossil fuel combustion can increase global CO2 atmospheric concentrations. One unique approach to reducing anthropogenic CO2 emissions involves coupling enhanced gas recovery (EGR) operations in depleted shale gas reservoirs with CO2 storage operations. In CO2-based EGR, carbon dioxide is injected into the subsurface to enhance methane recovery. The premise behind CO2 EGR is that carbon dioxide has a higher affinity for the organic and inorganic components of shales and will therefore displace adsorbed CH4. Potential benefits are that reservoir pressure is maintained, efficiency is improved, and carbon dioxide is stored for the long term, thereby offsetting atmospheric CO2 from methane combustion. At the depths of injection, temperature and pressure conditions dictate that CO2 will exist as a supercritical fluid that can mix with a wide ranging concentrations of CH4 and dissolved H2O. The interactions between shale minerals and dissolved H2O in CO2−CH4 supercritical fluids are not well understood, but this information is needed in simulators to predict optimum injection strategies and maximize natural gas production. © XXXX American Chemical Society

Dissolved H2O concentrations are, in fact, a critical parameter that could affect the efficiency and economic favorability of CO2 EGR in a target shale reservoir. For example, clay minerals partly determine the physical (i.e., permeability, brittleness) and certain chemical properties (i.e., wetting ability, gas adsorption) of shales. Montmorillonites are of particular interest because they swell by the uptake of species (primarily water) in their interlayer. The extent to which expandable clays swell is controlled primarily by the water activity in the contact fluid, and swelling due to H2O intercalation processes could lead to permeability changes that directly impact CH4 transmissivity.2−4 Also, CO2-rich fluids containing dissolved water are highly reactive with respect to divalent metal silicate minerals. Water can adsorb onto these phases, leading to the precipitation of divalent metal carbonates.5−7 These carbonation reactions result in increased solid volume that could plug fractures, reduce porosity and permeability, and lower methane recovery. Dissolved H2O concentrations could also affect competitive CO 2 −CH 4 sorption onto kerogen components of shales. Depending on Received: November 30, 2016 Accepted: March 30, 2017

A

DOI: 10.1021/acs.jced.6b00999 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Table 1. Experimental Mole Percent CO2 Values for the Initially Dry Gas Mixtures Used in this Study as Determined GC Gravimetrically and by GC Analysis (xGr CO2 and xCO2, Respectively), Calculated Densities of Initially Dry Gas Mixtures (ρ), and Experimental Values for H2O Concentrations (CH2O) and Mole Permil (xH2O) at Saturation in these Mixtures at T = 323.2 K and P = 9.000 MPaa b xGr CO2 (%)

c xGC CO2 (%)

ρ (g/mL)d

CH2O (mmol·L−1)e

u(CH2O)e

xH2O (‰)e

u(xH2O)e

0.00 3.01 25.00 49.95 75.00 100.00

na 2.9 24.4 49.7 74.9 na

0.059 0.064 0.091 0.128 0.180 0.285

7.5 7.5 8.5 10.7 14.7 24.2

0.4 0.5 0.3 0.1 0.6 0.4

2.0 2.0 2.14 2.50 3.0 3.74

0.1 0.1 0.06 0.03 0.1 0.06

a

Standard uncertainties u are u(T) = 0.1 K and u(P) = 0.005 MPa. bMeasured by the gas vendor. These mole percents are the compositions of the c initially dry gas mixtures and have standard uncertainties u(xGr CO2) of 0.03. GC measurements were performed by Isotech Laboratories Inc. to verify the gravimetrically measured mole percent CO2 values. Up to 1% N2 + O2 at a ratio consistent with air was detected in some of these samples. We assume that this contamination occurred during the sample preparation or GC analysis. na = not analyzed. These mole percents are the compositions d of the initially dry gas mixtures and have standard uncertainties u(xGC CO2) of 0.2. Densities were calculated using STOMP-COMP and have standard e uncertainties u(ρ) better than 0.001 g/mL. These concentrations are at equilibrium with liquid water. The expanded uncertainties u(CH2O) and u(xH2O) have a coverage factor k = 10 and are based on standard deviation of the mean for replicate titrations.

optics. A calibrated tubing loop (see below) was used to quantitatively meter water into the pressurized IR cell. The cell was jacketed for temperature control by a circulating water bath, and temperature was monitored by a high-pressure, Ktype thermocouple that was calibrated using ice and boiling water. The thermocouple was positioned so that its sensing tip was inside the main fluid cavity of the vessel. The cell temperature varied by ±0.1 K which is within the specifications for the circulating water bath. Cell pressure was monitored by a pressure transducer that was calibrated against a NIST-traceable standard gauge with a precision of better than 0.005 MPa. Further description of the apparatus, cell, spectrometer settings, and the general method for performing titrations are reported elsewhere by Thompson et al. (2014) and will not be repeated here.25 However, two modifications were made in this study. First, the path length of the transmission optics was increased to 40 mm for increased sensitivity to dissolved H2O. Second, the method of water delivery was modified as follows. Prior to each water addition, the pressure in the cell was lowered from 9.000 to 8.950 MPa by a controlled leak. Then, an aliquot of water was injected into the high-pressure cell using anhydrous fluid (appropriate CH4−CO2 mixture) from a syringe pump at 298.0 K and 15.0 MPa. The syringe pump (Teledyne ISCO, model 260D) was jacketed, and its temperature was maintained to within ±0.1 K by a circulating water bath. The pressure of the syringe pump was monitored by a built in pressure transducer that was calibrated against a NIST-traceable standard gauge. The injection was considered complete once the cell pressure returned to 9.000 MPa. The volume of fluid leaked and anhydrous fluid added to the cell was precisely determined from the difference in the volumes of the syringe pump before and after the addition after accounting for the densities of the fluid at 323.2 K and 9.000 MPa and 298.0 K and 15.0 MPa. This net volume was used in our data analysis to account for water loss during leakage and dilution by anhydrous fluid. The analysis of transmission IR data was performed by a least-squares fit of the spectrum of H2O dissolved in the CO2− CH4 mixture of interest, as well as the spectrum of the gas mixture itself to account for small variations (less than 0.04 MPa) in pressure. The wavenumber regions used in the fits for each mole percent CO2 are listed in Table S1. The regions were

oxygen content, kerogen can have both hydrocarbon and water wetting characteristics, and kerogen with high sorbed water content will likely have lower affinity for methane.8,9 In summary, the geochemistry of shales is controlled in part by water activity in CO2−CH4 mixtures. The ability to predict dissolved H2O concentrations in contact fluids may make it possible to tailor injection strategies based on reservoir conditions and maximize CO2−CH4 exchange rates. In this article, we report dissolved H2O concentrations at saturation for six CO2−CH4 mixtures, ranging from pure CO2 to pure CH4, measured at 323.2 K and 9.000 MPa using in situ high-pressure infrared (IR) spectroscopic titrations. Our chosen temperature and pressure correspond to a depth of 900 m assuming a standard hydrostatic gradient of 0.01 MPa/m, a geothermal gradient of 3.8 K per 100 m, and an average surface temperature of 289.2 K. These conditions have relevance to relatively shallow shale reservoirs in the United States suitable for CO2 EGR; for example, the average depths of the Ohio, New Albany, Big Sandy, Fayetteville, and Lewis shale plays are below 1400 m.10−18 We conclude by using our experimental results as a benchmark to test predictions of dissolved H2O concentrations at saturation in CO2−CH4 mixtures at 323.2 K and 9.000 MPa by three models: (1) mixed-solvent electrolyte (MSE)19 (as implemented in the OLI Stream Analyzer20), (2) STOMP-COMP,21 and (3) statistical associating fluid theory for variable range Mie potentials (SAFT γ-Mie).22−24

2. EXPERIMENTAL SECTION The gas mixtures used in this study were purchased from Oxarc Inc. with a purity higher than 99.999%. Their mole percent CO2 values calculated from the masses of CO2 and CH4 measured by this vendor are reported in Table 1. A gas chromatographic (GC) analysis was performed by Isotech Laboratories Inc. (Champaign, IL) to independently verify the gravimetrically determined mole fractions; these verification results are also reported in Table 1. Only deionized (Barnstead NanoPure, Resistivity of 18.2 MΩ·cm) water was used in this study. In situ high-pressure titrations with IR detection were performed at 323.2 K and 9.000 MPa using an automated fluid-delivery apparatus coupled to a high-pressure IR cell with both transmission and attenuated total reflection (ATR) IR B

DOI: 10.1021/acs.jced.6b00999 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Figure 1. Transmission data from IR titrations with water of CO2−CH4 mixtures listed in Table 1 at 323.2 K and 9.000 MPa. The mole percent CO2 values are indicated, and at least one repetition was performed for each gas mixture. These plots show the integrated absorbance of the HOH bending mode of dissolved H2O versus H2O concentration. The intersection of the straight lines fit through the linear and plateau regions of the data is the dissolved H2O concentration at saturation.

integrated absorbances of the transmission IR data reported in this study are of the spectral contribution determined in the least-squares fit from H2O dissolved in the gas mixture in the wavenumber region between 1650 and 2054 cm−1. The volume of the tubing loop that was used to quantitatively meter water into the pressurized IR cell was calibrated for each CO2−CH4 gas mixture using the following procedure. First, an experiment was performed to measure the

chosen based on where dissolved H2O absorbs but is not too intense (i.e., off-scale) and where the CO2 and/or CH4 bands of the gas mixture do not significantly overlap. A linear baseline was also fit for each wavenumber region to correct for baseline drift. The residuals from this fit were almost exclusively the spectrum of water vapor in the spectrometer due to fluctuations in the spectrometer’s vacuum and the humidity of the purge gas provided for the air bearing of the interferometer. The C

DOI: 10.1021/acs.jced.6b00999 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

integrated absorbance of H2O dissolved in a supercritical CO2− CH4 mixture, where the water was manually injected into the high pressure cell using a microliter syringe. The volume of water was typically between 5 and 20 μL and was verified gravimetrically by weighing the syringe before and after the injection. After pressurization, this volume led to an integrated absorbance of dissolved H2O that was between 50 and 60% saturation. Second, another experiment was performed in which water was titrated into the pressurized cell by repeated additions from the tubing loop until the integrated absorbance of the dissolved H2O was larger than that absorbance determined during the experiment using the manual syringe injection. Third, an equivalent number of injections was calculated that was required to achieve the same integrated absorbance of dissolved H2O in the titration experiment as for the manual injection experiment. Finally, the volume of the loop was calculated by dividing the total volume for the loop injections by the number of injections made. The loop volume that was determined in this study was 0.848 ± 0.007 μL and is an average of the individual loop volumes determined for each CO2−CH4 gas mixture. For a titration experiment, the total volume of water added to the cell is the sum of the volumes from n individual additions, and the error in the total volume is

increasing water concentration. Also, the bands of H2O dissolved in the fluid mixtures are not detected in the ATRIR spectra because the effective path length is too short. However, both OH stretching and HOH bending modes of dissolved H2O are easily detected in the longer path length transmission-IR spectra, and these are shown in Figure S2. With increasing mole percent CO2, the rotational fine structure of these bands broadens and collapses into the Q branch due to quadrupole−dipole interactions between CO2 and H2O that hinder rotation.27 Figure S3 shows a representative example of transmission-IR spectra collected during a titration of H2O into one of the fluids, pure CH4. The OH stretching and HOH bending modes of dissolved H2O increase in absorbance with increasing water concentration until the fluid is saturated. At saturation, the absorbance of these bands remains constant despite further increases in water. Figure 1 shows plots of the integrated absorbance of the R branch of the HOH bending mode region (1670 to 2020 cm−1) versus total water concentrations from titrations in all six fluid mixtures. The saturation point for each fluid is the H2O concentration at the intersection of the straight lines that are fit through the data in the linear and plateau regions.25,28,29 These values are listed in Table 1 (in units of both mmol·L−1 and mole fraction) and plotted in Figure 2 (in

n(0.007 μL)2 . It follows that errors in the total volume of water required to reach saturation ranged from ±0.02 μL for pure CH4 to ±0.04 μL for pure CO2. However, this is only one source of uncertainty in the saturation concentrations; others include spectrometer drift and longer-term temperature and pressure fluctuations. The concentrations of water at saturation that we report here are average values from replicate titrations. Errors in these concentrations are based on the variability of these replicates, and we conservatively estimate the error as 10 times the standard deviation.

3. RESULTS AND DISCUSSION The mole percent CO2 values and calculated densities (using STOMP-COMP, see below) for each fluid mixture at 323.2 K and 9.000 MPa are listed in Table 1. The compressibility factor of pure CO2 is about 0.58 that of CH4 at this temperature and pressure;26 hence, the observed increase in density with increasing mole percent CO2 likely reflects stronger intermolecular interactions leading to a greater deviation from an ideal gas for CO2 versus CH4. ATR-IR spectra of the anhydrous fluids are shown in Figure S1 at 323.2 K and 9.000 MPa in the (a) asymmetric CO stretching region of CO2, (b) CO2 bending region of CO2, (c) asymmetric CH stretching region of CH4, and (d) CH bending region of CH4. With increasing mole percent CO2, the asymmetric CO stretching band of CO2 shows a broadening of P and R branch rotational fine structure and collapse into the Q branch. These changes are predominantly due to quadrupole−quadrupole interactions between CO2 molecules that hinder rotational mobility.27 The CO2 bending mode, which shows no obvious rotational fine structure, shifts from 668 in the mixture with 3% CO2 to 665 cm−1 in pure scCO2. The CH stretching and bending modes of CH4, on the other hand, retain their P and R branches and show little change with increasing mole percent CO2. This is likely because of the lack of a dipole or quadrupole in CH4, and thus weak interactions with CO2 or other CH4 molecules. The ATR-IR spectra of the hydrated fluids show no detectable change in the bands of CO2 and CH4 with

Figure 2. Measured dissolved H2O concentration at saturation versus mole percent CO2 for the gas mixtures listed in Table 1 at 323.2 K and 9.000 MPa. Predictions are also shown that are based on the MSE, STOMP-COMP, and SAFT γ-Mie models.

units mmol·L−1) for the six fluid mixtures. Water concentrations at saturation increase nearly linearly with increasing mole percent CO2 between 0% and 50%, but the increase is even steeper beyond 50% CO2 in CH4. The physical reason for this is that the concentration (mol·L−1) of CO2 increases dramatically above 50% CO2 in CH4 at 323.2 K and 9.000 MPa. Because of the relatively strong quadrupole−dipole interactions between CO2 and H2O, increasing concentrations of CO2 lead to greater concentrations of H2O at saturation. Our measured values for the mole fractions of water at saturation in Table 1 can be compared to data in the literature at similar conditions (see Figure S4). Our value of 3.74 ± 0.06‰ for pure CO2 at 323.2 K and 9.000 MPa is between the measured values by Briones et al.30 of 3.64‰ at 323 K and 8.72 MPa and Bamberger et al.31 of 4.1‰ at 323 K and 9.09 MPa, D

DOI: 10.1021/acs.jced.6b00999 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

differ from our newly measured data by ∼9%, while the SAFT γ-Mie model differs by ∼23%. Agreement is better for the CO2−H2O system, where the models differ by only ∼4% for the MSE and SAFT γ-Mie models less than 1% for STOMPCOMP (see Figure 2). Furthermore, in the case of the CO2− H2O system, the solubility of H2O at 323.2 K and 9.000 MPa is close to a maximum as a function of pressure, which somewhat magnifies the uncertainty of both the data and the models. In the ternary CO2−CH4−H2O system, all three models accurately predict the shape of the H2O solubility curve as a function of the composition of the supercritical fluid, and no deterioration of accuracy is observed compared to the binary systems.

and it is in good agreement with the calculated value using the equation of state model by Spycher et al. of 3.54‰.32 We were unable to find literature values for pure CH4 at 323.2 K and 9.000 MPa. However, our value of 2.0 ± 0.1‰ is in reasonable agreement with those reported by Rigby et al.33 of 2.2‰ at 323 K and 6.8 MPa, by Chapoy et al.34 of 1.3‰ at 318 K and 10.01 MPa, and by Yokoyama et al.35 of 2.1‰ at 323 K and 8.0 MPa. Our value is also in the range between values at 9.0 MPa of 0.5‰ at 314 K and 2.9‰ at 366 K by Yarrison et al.,36 but it is lower than values at 323 K reported by Frost et al.37 of 2.65‰ at 8.48 MPa and 2.24‰ at 11.49 MPa. While there are some published studies reporting mole fractions of water at saturation in CO2−CH4 mixtures,38−41 measurements close to our temperature−pressure conditions are sparse. Song et al.40 reported values of 2.8‰ and 3.0‰ at 7.6 and 10.3 MPa, respectively, at 323 K for a mixture of 5.31% CO2 in CH4. Oddly, these concentrations are even higher than our value of 2.14 ± 0.06‰ at 323.2 K and 9.000 MPa for 25% CO2 in CH4. Finally, Al Ghafri et al.38 reported mole fractions of water at saturation in 50% CO2 in CH4 at 323 K of 3.78‰ at 6.035 MPa and 4.11‰ at 10.011 MPa, both of which are higher than our value of 2.50 ± 0.03‰ in this CO2−CH4 mixture at 323.2 K and 9.000 MPa. We calculated concentrations of water at saturation using the MSE, STOMP-COMP, and SAFT-VR-SW models. The MSE model was originally developed for water−electrolyte−nonelectrolyte systems by Wang et al.19 and further extended by Springer et al.42,43 to systems containing common gas components in wide ranges of pressures and temperatures. The model has been designed to reproduce the properties of both aqueous and CO2- or hydrocarbon-rich phases, and it has been parameterized for the binary systems CO2−H2O and CH4−H2O.42,43 Within the MSE framework the standard-state properties are calculated from the Helgeson−Kirkham− Flowers equation of state in which the excess Gibbs energy includes a long-range electrostatic interaction term expressed by a Pitzer−Debye−Hückel equation, a virial coefficient-type term for interactions between ions, and a short-range term for interactions involving neutral molecules.42,43 The STOMPCOMP model21 assumes the multiphase system is composed of noncondensable gases (i.e., CO2 and CH4) and water. The thermophysical properties of water are specified internally, and fluid components may partition into either a single aqueous phase or into two phases (gas-aqueous). The properties and water solubilities of fluid mixtures are calculated using a Peng− Robinson equation of state44 with distinct binary interaction coefficients for each phase45 that have been calibrated to observed gas solubilities in pure water and brine (EOS-COMP, see appendix in Supporting Information).21 The SAFT γ-Mie24 model is an equation of state that provides a continuous description of both liquid and vapor phases. The Helmholtz free energy of each molecule is calculated by considering it to be a chain of fused segments, where the segments interact via the Mie potential (a generalized Lennard-Jones potential). The parameters of the equation of state are state independent and are usually obtained by fitting to vapor−liquid equilibrium data. The parameters of the equation of state used here to predict the properties of the carbon dioxide, water and methane system were developed by Papaioannou et al.23 Figure 2 compares our measured values for the concentrations of water at saturation to those calculated using the MSE, STOMP-COMP, and SAFT γ-Mie models. In the binary CH4−H2O system, both the MSE and STOMP-COMP models

4. CONCLUSIONS We have measured water concentrations at saturation for six CO2−CH4 mixtures at 323.2 K and 9.000 MPa. Water concentrations increase with increasing mole percent CO2 in the supercritical fluid mixtures. This trend is expected, given that CO2 interacts more strongly with H2O than CH4 and that both the concentration of CO2 and the overall fluid density increases with increasing mole percent CO2. We have also used these results to benchmark three thermodynamic models. The MSE, STOMP-COMP, and SAFT γ-Mie models provide reasonable predictions of water solubilities in mixed CO2− CH4 fluids on the basis of parameters that were obtained from binary data, although improvement could be afforded to the SAFT γ-Mie model in fluids with greater CH4 content. Furthermore, the temperature and pressure conditions investigated here are only relevant to CO2 EGR operations in shallow shale plays with depths of close to 900 m. Future studies should focus on conditions pertinent to deeper reservoirs such as Woodford, Marcellus, Eagle Ford, and Barnett shales (1600−2700 m, 348−393 K, 16.0−27.0 MPa) and Monterey, Haynesville, and Mancos shales (3300−4500 m, 413−463 K, 33.0−45.0 MPa).11,12



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.jced.6b00999. Wavenumber regions used in least-squares fits; figures showing the ATR-IR spectra of dry CO2−CH4 mixtures and transmission-IR spectra of H2O dissolved in these mixtures; measured dissolved H2O concentrations (‰) at saturation versus mole percent CO2, including available literature data; measured dissolved H2O concentrations (mmol·L−1) at saturation versus mole percent CO2, including predictions using the SAFT-VRSW model of Miguez et al. (2011);46 appendix providing details of EOS-COMP modeling (PDF)



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Tel.: (509) 371-6743. Fax: (509) 371-6354. ORCID

John S. Loring: 0000-0001-6106-3812 Andrzej Anderko: 0000-0002-1522-4889 Christopher J. Thompson: 0000-0002-9932-3234 Kevin M. Rosso: 0000-0002-8474-7720 E

DOI: 10.1021/acs.jced.6b00999 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

(15) Hill, D. G.; Nelson, C. Gas productive fractured shales: An overview and update. GasTIPS 2000, 6, 4−13. (16) Hosterman, J. W.; Whitlow, S. I. Geological Survey Professional Paper 1298: Clay Mineralogy of Devonian Shales in the Appalachian Basin. 1983; https://pubs.usgs.gov/pp/1298/report.pdf (accessed March 4, 2017). (17) Momentive. Fracline Oilfield Technology Newsletter. Case Histories: Haynesville, Fayetteville, and Bakken Shales. 2011; http:// www.hexionfracline.com/case-histories-haynesville-fayetteville-andbakken-shales (accessed March 04, 2017). (18) Tuttle, M. L. W.; Breit, G. N. Weathering of the New Albany Shale, Kentucky, USA: I. Weathering zones defined by mineralogy and major-element composition. Appl. Geochem. 2009, 24, 1549−1564. (19) Wang, P.; Anderko, A.; Young, R. D. A speciation-based model for mixed-solvent electrolyte systems. Fluid Phase Equilib. 2002, 203, 141−176. (20) OLI Analyzer Studio, Version 9.5; OLISystems: Cedar Knolls, NJ, 2016. (21) Bacon, D. H.; Ramanathan, R.; Schaef, H. T.; McGrail, B. P. Simulating geologic co-sequestration of carbon dioxide and hydrogen sulfide in a basalt formation. Int. J. Greenhouse Gas Control 2014, 21, 165−176. (22) Galindo, A.; Davies, L. A.; Gil-Villegas, A.; Jackson, G. The thermodynamics of mixtures and the corresponding mixing rules in the SAFT-VR approach for potentials of variable range. Mol. Phys. 1998, 93, 241−252. (23) Papaioannou, V.; Calado, F.; Lafitte, T.; Dufal, S.; Sadeqzadeh, M.; Jackson, G.; Adjiman, C. S.; Galindo, A. Application of the SAFTgamma Mie group contribution equation of state to fluids of relevance to the oil and gas industry. Fluid Phase Equilib. 2016, 416, 104−119. (24) Papaioannou, V.; Lafitte, T.; Avendano, C.; Adjiman, C. S.; Jackson, G.; Muller, E. A.; Galindo, A. Group contribution methodology based on the statistical associating fluid theory for heteronuclear molecules formed from Mie segments. J. Chem. Phys. 2014, 140, 054107. (25) Thompson, C. J.; Martin, P. F.; Chen, J.; Benezeth, P.; Schaef, H. T.; Rosso, K. M.; Felmy, A. R.; Loring, J. S. Automated highpressure titration system with in situ infrared spectroscopic detection. Rev. Sci. Instrum. 2014, 85, 044102. (26) Dymond, J. H.; Smith, E. B. The Virial Coefficients of Pure Gasses and Mixtures; Oxford University Press: Oxford, 1980. (27) Bowman, L. E.; Palmer, B. J.; Garrett, B. C.; Fulton, J. L.; Yonker, C. R.; Pfund, D. M.; Wallen, S. L. Infrared and molecular dynamics study of D2O rotational relaxation in supercritical CO2 and Xe. J. Phys. Chem. 1996, 100, 18327−18334. (28) Foltran, S.; Vosper, M. E.; Suleiman, N. B.; Wriglesworth, A.; Ke, J.; Drage, T. C.; Poliakoff, M.; George, M. W. Understanding the solubility of water in carbon capture and storage mixtures: An FTIR spectroscopic study of H2O + CO2 + N2 ternary mixtures. Int. J. Greenhouse Gas Control 2015, 35, 131−137. (29) Jackson, K.; Bowman, L. E.; Fulton, J. L. Water solubility measurements in supercritical fluids and high-pressure liquids using near-infrared spectroscopy. Anal. Chem. 1995, 67, 2368−2372. (30) Briones, J. A.; Mullins, J. C.; Thies, M. C.; Kim, B. U. Ternary Phase-Equilibria for Acetic Acid-Water Mixtures with Supercritical Carbon-Dioxide. Fluid Phase Equilib. 1987, 36, 235−246. (31) Bamberger, A.; Sieder, G.; Maurer, G. High-pressure (vapor plus liquid) equilibrium in binary mixtures of (carbon dioxide plus water or acetic acid) at temperatures from 313 to 353 K. J. Supercrit. Fluids 2000, 17, 97−110. (32) Spycher, N. F.; Reed, M. H. Fugacity Coefficients of H2, CO2, CH4, H2O and of H2O-CO2-CH4 Mixtures - a Virial Equation Treatment for Moderate Pressures and Temperatures Applicable to Calculations of Hydrothermal Boiling. Geochim. Cosmochim. Acta 1988, 52, 739−749. (33) Rigby, M.; Prausnitz, J. M. Solubility of water in compressed nitrogen argon and methane. J. Phys. Chem. 1968, 72, 330−334. (34) Chapoy, A.; Mohammadi, A. H.; Tohidi, B.; Richon, D. Estimation of water content for methane plus water and methane plus

Herbert T. Schaef: 0000-0002-4546-3979 Funding

This material is based upon work supported by the U.S. Department of Energy (DOE), Office of Science, Office of Basic Energy Sciences, Chemical Sciences, Geosciences & Biosciences Division through its Geoscience Program at Pacific Northwest National Laboratory (PNNL). H.T.S. acknowledges support from the DOE Office of Fossil Energy. Development of the IR titration instrumentation for this research was funded through PNNL’s Carbon Sequestration Initiative, a Laboratory Directed Research and Development Program. PNNL is operated for DOE by Battelle Memorial Institute under Contract No. DE-AC06-76RLO-1830. Notes

The authors declare no competing financial interest.



REFERENCES

(1) Natural Gas Overview: Production from 2006 to 2016. http:// www.eia.gov/totalenergy/data/browser/index.php?tbl=T04.01#/?f= A&start=2006&end=2015&charted=3-8-11 (accessed March 4, 2017). (2) Loring, J. S.; Ilton, E. S.; Chen, J.; Thompson, C. J.; Martin, P. F.; Benezeth, P.; Rosso, K. M.; Felmy, A. R.; Schaef, H. T. In Situ Study of CO2 and H2O Partitioning between Na-Montmorillonite and Variably Wet Supercritical Carbon Dioxide. Langmuir 2014, 30, 6120−6128. (3) Schaef, H. T.; Loring, J. S.; Glezakou, V. A.; Miller, Q. R. S.; Chen, J.; Owen, A. T.; Lee, M. S.; Ilton, E. S.; Felmy, A. R.; McGrail, B. P.; Thompson, C. J. Competitive sorption of CO2 and H2O in 2:1 layer phyllosilicates. Geochim. Cosmochim. Acta 2015, 161, 248−257. (4) de Jong, S. M.; Spiers, C. J.; Busch, A. Development of swelling strain in smectite clays through exposure to carbon dioxide. Int. J. Greenhouse Gas Control 2014, 24, 149−161. (5) Loring, J. S.; Chen, J.; Benezeth, P.; Qafoku, O.; Ilton, E. S.; Washton, N. M.; Thompson, C. J.; Martin, P. F.; McGrail, B. P.; Rosso, K. M.; Felmy, A. R.; Schaef, H. T. Evidence for Carbonate Surface Complexation during Forsterite Carbonation in Wet Supercritical Carbon Dioxide. Langmuir 2015, 31, 7533−7543. (6) Loring, J. S.; Thompson, C. J.; Wang, Z. M.; Joly, A. G.; Sklarew, D. S.; Schaef, H. T.; Ilton, E. S.; Rosso, K. M.; Felmy, A. R. In Situ Infrared Spectroscopic Study of Forsterite Carbonation in Wet Supercritical CO2. Environ. Sci. Technol. 2011, 45, 6204−6210. (7) Schaef, H. T.; McGrail, B. P.; Loring, J. L.; Bowden, M. E.; Arey, B. W.; Rosso, K. M. Forsterite [Mg2SiO4)] Carbonation in Wet Supercritical CO2: An in Situ High-Pressure X-ray Diffraction Study. Environ. Sci. Technol. 2013, 47, 174−181. (8) Hu, Y. N.; Devegowda, D.; Striolo, A.; Phan, A. T. V.; Ho, T. A.; Civan, F.; Sigal, R. Microscopic dynamics of water and hydrocarbon in shale-kerogen pores of potentially mixed wettability. SPE J. 2014, 20, 112−124. (9) Zhao, T.; Li, X.; Zhao, H.; Li, M. Molecular simulation of adsorption and thermodynamic properties on type II kerogen: Influence of maturity and moisture content. Fuel 2016, 190, 198−207. (10) Modern Shale Gas Development in the United States: A Primer. 20 09 ; h t t p s : / / e n e r gy . g o v / s i t e s / p r o d / fi le s / 2 01 3/ 03 / f0 / ShaleGasPrimer_Online_4-2009.pdf (accessed March 4, 2017). (11) Review of Emerging Resources: U.S. Shale Gas and Shale Oil Plays. 2011; https://www.eia.gov/analysis/studies/usshalegas/pdf/ usshaleplays.pdf (accessed March 4, 2017). (12) Shale Gas and Shale Oil Resource Assessment Methodology. 2013; http://www.adv-res.com/pdf/00_EIA_ARI_Study_ Methodology%20June_2013_FINAL.pdf (accessed March 4, 2017). (13) Curtis, J. B. Fractured shale-gas systems. AAPG Bull. 2002, 86, 1921−1938. (14) Hexion. Fracline Oilfield Technology Newsletter. What’s Working in the Fayetteville and Barnett Plays? 2009; http:// hexionfracline.com/summer2009/whatsworkfayetbarnett.php (accessed March 4, 2017). F

DOI: 10.1021/acs.jced.6b00999 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

ethane plus n-butane plus water systems using a new sampling device. J. Chem. Eng. Data 2005, 50, 1157−1161. (35) Yokoyama, C.; Wakana, S.; Kaminishi, G.; Takahashi, S. Vapor liquid equilibria in the methane diethylene glycol water-system at 298.15 and 323.15 K. J. Chem. Eng. Data 1988, 33, 274−276. (36) Yarrison, M.; Cox, K. R.; Chapman, W. G. Measurement and modeling of the solubility of water in supercritical methane and ethane from 310 to 477 K and pressures from 3.4 to 110 MPa. Ind. Eng. Chem. Res. 2006, 45, 6770−6777. (37) Frost, M.; Karakatsani, E.; von Solms, N.; Richon, D.; Kontogeorgis, G. M. Vapor-liquid equilibrium of methane with water and methanol. Measurements and modeling. J. Chem. Eng. Data 2014, 59, 961−967. (38) Al Ghafri, S. Z. S.; Forte, E.; Maitland, G. C.; RodriguezHenriquez, J. J.; Trusler, J. P. M. Experimental and modeling study of the phase behavior of (methane + CO2 + water) mixtures. J. Phys. Chem. B 2014, 118, 14461−14478. (39) Chapoy, A.; Burgass, R.; Terrigeol, A.; Coquelet, C. Water content of CO2-rich mixtures: Measurements and modeling using the cubic-plus-association equation of state. J. Nat. Gas Sci. Eng. 2016, 1, 85−97. (40) Song, K. Y.; Kobayashi, R. The water-content of a CO2-rich gasmixture containing 5.31-mole-percent methane along the 3-phase and supercritical conditions. J. Chem. Eng. Data 1990, 35, 320−322. (41) Tabasinejad, F.; Moore, R. G.; Mehta, S. A.; Van Fraassen, K. C.; Barzin, Y.; Rushing, J. A.; Newsham, K. E. Water solubility in supercritical methane, nitrogen, and carbon dioxide: Measurement and modeling from 422 to 483 K and pressures from 3.6 to 134 MPa. Ind. Eng. Chem. Res. 2011, 50, 4029−4041. (42) Springer, R. D.; Wang, P.; Anderko, A. Modeling the properties of H2S - CO2 - salt - water systems in wide ranges of temperature and pressure. SPE J. 2015, 20, 1120−1134. (43) Springer, R. D.; Wang, Z.; Anderko, A.; Wang, P.; Felmy, A. R. A thermodynamic model for predicting mineral reactivity in supercritical carbon dioxide: I. Phase behavior of carbon dioxide − water − chloride salt systems across the H2O-rich to the CO2-rich regions. Chem. Geol. 2012, 322−323, 151−171. (44) Peng, D.; Robinson, D. B. New 2-Constant Equation of State. Ind. Eng. Chem. Fundam. 1976, 15, 59−64. (45) Soreide, I.; Whitson, C. H. Peng-robinson predictions for hydrocarbons, CO2, N2, and H2S with pure water and NaCl brine. Fluid Phase Equilib. 1992, 77, 217−240. (46) Miguez, J. M.; dos Ramos, M. C.; Pineiro, M. M.; Blas, F. J. An examination of theternary methane plus carbon dioxide plus water phase diagram using the SAFT-VR approach. J. Phys. Chem. B 2011, 115, 9604−9617.

G

DOI: 10.1021/acs.jced.6b00999 J. Chem. Eng. Data XXXX, XXX, XXX−XXX