Selective Recovery of H2S from Gas Mixtures Using a Hydroquinone

May 31, 2017 - Kang et al. measured the phase equilibria of mixed CO2–N2 gas hydrates in which water acts as a host material and proposed a hydrate-...
3 downloads 8 Views 1MB Size
Article pubs.acs.org/jced

Selective Recovery of H2S from Gas Mixtures Using a Hydroquinone Clathrate Jong-Won Lee,† Sang Jun Yoon,*,‡ and Ji-Ho Yoon*,§ †

Department of Environmental Engineering, Kongju National University, 1223-24 Cheonan-Daero, Cheonan-si, Chungnam 31080, Republic of Korea ‡ Clean Fuel Laboratory, Korea Institute of Energy Research, 152 Gajeong-ro, Yuseong-gu, Daejeon 34129, Republic of Korea § Department of Energy and Resources Engineering, Korea Maritime and Ocean University, 727 Taejong-ro, Yeongdo-gu, Busan 49112, Republic of Korea ABSTRACT: The formation and guest enclathration of hydroquinone (HQ) clathrates with gas mixtures were investigated for potential applications to a clathrate-based removal process for sulfides. The HQ clathrate samples were prepared with hydrogen sulfide (H2S)−carbonyl sulfide (COS)−nitrogen (N2) gas mixtures at pressures from 2.0 to 8.0 MPa and ambient temperature. Solid-state 13C NMR and Raman spectroscopy methods were used to identify the crystal structure of the samples and guest enclathration into the clathrate framework. Quantitative information regarding the amount of captured gases was obtained by a combination of the elemental analysis and the numerical integration of the NMR spectra. The results show that H2S can be more concentrated in the solid HQ clathrates than in the gas phase, depending on the formation pressure. In particular, COS did not participate in the enclathration under all conditions, indicating that the molecular size of COS is too large to fit into the cages of the HQ clathrate. The recovery efficiency of H2S was described in terms of the enhancement of the concentration and selectivity of H2S in the HQ clathrates. These results open up the potential for clathrate-based gas separation for sulfur compounds, especially under dilute conditions.

1. INTRODUCTION Combustion is the prevailing mode of fossil energy utilization,1 and coal is one of the principal fossil fuels of electric power generation because it is distributed all over the globe and its reserves are estimated to be sufficient for the next several decades.2 However, growing environmental issues should be resolved to meet stringent regulations related to pollutants from coal combustion power generation. In addition, as the use of gasified energy sources such as natural gas, synthesis gas, and biomass gas increases, the removal of pollutants from the gases has become one of the most important technologies today. Among those pollutants, sulfur compounds are of specific interest because they are toxic and corrosive, resulting in severe catalyst poisoning and damage to equipment and pipelines.3,4 In energy gases, hydrogen sulfide (H2S) is the primary gasphase sulfide, with carbonyl sulfide (COS) forming the secondary sulfide.2 To date, adsorption, absorption, and membrane processes have been adapted to remove such sulfides from gasified coal and other sources.5−8 In addition to the conventional separation technologies, a novel technique using clathrate compounds has also been suggested. Clathrate compounds are formed by the enclathration of guest species into cagelike structures formed by host materials. The clathrate compound can hold large amounts of gas in a unit volume of the solid phase because gaseous guests can be compressed significantly while being enclathrated into the solid phase. In addition, this process can concentrate or © XXXX American Chemical Society

separate a specific component from gas mixtures because the formation conditions vary depending on the guest species used. Kang et al. measured the phase equilibria of mixed CO2−N2 gas hydrates in which water acts as a host material and proposed a hydrate-based CO2 separation process from flue gases.10 In addition, other researchers also have reported hydrate-based gas separation concepts for various alkane/alkene mixtures.11−13 Recently, Lee et al. reported on the dry synthesis of clathrate compounds using hydroquinone (HQ) as a host instead of water, which is solid at room temperature.14 Furthermore, they also used the dry synthesis method to separate gas species from various gas mixtures.15,16 Unlike the gas hydrate with two types of cages, HQ clathrates have only one type of cage acting like a molecular sieve so that selective separation can be achieved for specific molecules from ethylene/ethane and CO2/H2 gas mixtures.16,17 In this study, the potential application of HQ clathrates in a separation/recovery process for sulfides in various energy gases was investigated. To this end, two sulfides, H2S and COS, were selected and reacted with HQ at room temperature and at various pressures. After the reaction, the samples were analyzed by spectroscopic methods and elemental analysis. Both sulfides are known to form the clathrate compounds from experimental Received: March 10, 2017 Accepted: May 19, 2017

A

DOI: 10.1021/acs.jced.7b00248 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

and simulation works in the literature.9,18−20 However, because those works were performed mainly from an academic perspective, pure H2S and COS gases were used, resulting in some differences from practical applications with concentrations of less than 1%. In this regard, the recovery/ enclathration behaviors with HQ clathrates were investigated using a H2S/COS/N2 (2.5/1.5/96 mol %) gas mixture. The concentration of the experimental sour gas selected in this study was based on the industrial maximum H2S concentration of 1.0−1.5% and on the COS concentration being approximately half of the H2S concentration. The experimental and analysis results in this report can be used as fundamental data to design future clathrate-based separation processes. It still needs some work, including examining the reaction kinetics to develop a practical design.

results of the elemental analysis were found to be in good agreement with the theoretical value of pure HQ with deviations of less than ±1.0%. All of the analytical instruments were located in the Analysis Center for Research Advancement of the Korea Advanced Institute for Science and Technology (KAIST, Daejeon, Republic of Korea). Because the clathrate compound forms and exists stably under specific temperature and pressure conditions, the sample could be dissociated after releasing the gas from the high-pressure reactor. To minimize possible modifications or sample deterioration after decomposition, all of the measurements were performed immediately after the sample collection. The experimental procedures are basically the same as described in previous reports.16,17

3. RESULTS AND DISCUSSION Even though both H2S and COS were reported as guest species that form clathrate compounds with HQ as the host, they might show different enclathration behaviors at dilute concentrations. Therefore, the H2S/N2 (5/95 mol %) and COS/N2 (3/97 mol %) gas mixtures were reacted with HQ at 8.0 MPa to determine if they are formers of HQ clathrates. Figure 1 shows the solid-

2. EXPERIMENTAL SECTION HQ with a purity of 99 mol % supplied by Sigma-Aldrich Chemicals Co. was used as a reagent in this study without further purification. The H2S/COS/N2 gas mixtures were supplied by Daemyoung Special Gas. Co. (Republic of Korea). These gases were used for reactions without any further treatment. HQ clathrate compounds were prepared by charging a high-pressure cell with pure HQ. The cell was made with 316 stainless steel with an internal volume of 20 cm3. A reservoir (internal volume of 500 cm3) was attached to the cell to compensate for the pressure drop during the reaction and thus keep the experimental pressure constant throughout the reaction. Before being charged into the reactor, approximately 5.0 g of HQ was ground into a fine powder with a particle size of less than 45 μm to promote reactions between the solid and gas phases. After the loading, the reaction cell was purged and pressurized to the desired pressure at 298 K with the experimental gases. To monitor the pressure in the cell, a digital pressure indicator (Heise, ST-2H) was used. After the formation reaction for 14 days, the pressurized gas was slowly released from the cell, and the samples were collected for further analysis. To identify the clathrate formation and guest enclathration, the collected samples were analyzed by three methods: solidstate 13C nuclear magnetic resonance (NMR), Raman spectroscopy, and elemental analysis. For solid-state 13C NMR, an Agilent DD2 400 MHz spectrometer was used. The 13C NMR spectra were collected at ambient temperature by placing samples within a 1.6 mm HFXY fast magic angle spinning (MAS) probe and recorded at a spinning rate of 20 kHz. The pulse length of the proton was 2 μs, and a phase repetition delay of 10 s under proton decoupling was employed. The downfield carbon resonance peak of adamantane, which was assigned a chemical shift of 38.3 ppm at 300 K, was used as an external chemical shift reference. A dispersive Raman spectrometer (Horiba Jobin Yvon, Lab-RAM ARAMIS model) was equipped with a 460 mm focal length monochromator and an air-cooled charge-coupled device (CCD) detector. The excitation source was an Ar ion laser emitting a 514 nm line with a power of 40 mW. The Raman spectroscopic measurements were conducted at ambient temperature. The compositional contents of C, H, N, S, and O in the HQ samples were determined with an elemental analyzer (Thermo Scientific, Flash 2000 Series). Approximately 3.0 mg of each sample was analyzed with the dynamic flash combustion method with a thermal conductivity detector (TCD). From 10 repeated measurements, the experimental

Figure 1. Solid-state 13C NMR spectra for pure HQ and HQ clathrate samples after reaction with H2S/N2 (5/95 mol %) and COS/N2 (3/97 mol %) gas mixtures. The HQ clathrate samples were prepared for 14 days at 8.0 MPa and ambient temperature. All of the NMR spectra were collected at 298 K and ambient pressure.

state 13C NMR spectra for those samples. As shown in the figure, pure HQ before the reaction (α-form crystal) had two groups of carbon signals (one at around 150 ppm for the hydroxyl-substituted carbon atom and the other at around 120 ppm for the non-substituted carbon atom), while those signals were arranged more clearly after the reaction with the gases. Such changes represent three inequivalent carbons for the ideal centro-symmetric cage symmetry of the HQ clathrate (β-form), which implies the gases are trapped in the cages of the clathrate compound.14,21 Although the formation of HQ clathrates can be identified, it cannot be seen clearly which gas components are accommodated because there was no additional atomic signal other than those of the HQ. The H2S/COS/N2 (2.5/1.5/96 mol %) ternary gas mixture was reacted with HQ at various pressures. Figure 2 shows the solid-state 13C NMR spectra for the HQ samples reacted at 2.0−8.0 MPa. As seen in the figure, the HQ sample was found B

DOI: 10.1021/acs.jced.7b00248 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Figure 2. Solid-state 13C NMR spectra for HQ samples after reaction with H2S/COS/N2 (2.5/1.5/96 mol %) gas mixture at various pressures and ambient temperature for 14 days. All of the NMR spectra were collected at 298 K and ambient pressure.

Figure 3. Raman spectra for HQ samples after reaction with the H2S/ COS/N2 (2.5/1.5/96 mol %) gas mixture at various pressures and ambient temperature for 14 days. All of the Raman spectra were collected at 298 K and ambient pressure.

mainly HQ clathrates (β form) with some residual pure HQ (α form).22,23 Clathration-induced changes in the Raman spectra were clearly shown in the frequency ranges of 1100−1200 and 1500−1700 cm−1 (Figure 4). The two split bands corresponding to the C−H bending mode at 1163 and 1169 cm−1 in the spectra of pure HQ appeared as a single band at 1162 cm−1 in the spectra of HQ clathrates.24,25 Changes in the shape and intensity of unresolved triplet bands at 1601, 1611, and 1625 cm−1, attributed to the C−C stretching vibration mode, are also shown in Figure 4 and compared to the α-form pure HQ. Clearly, the Raman spectra of the HQ clathrate with the gas mixture at 2.0 MPa revealed a small amount of pure HQ, indicating incomplete conversion of the HQ clathrate. This result is consistent with the solid-state 13C NMR results. In addition to the characteristic pattern of the α-form and β-form HQ, additional Raman signals were detected in the spectral range of 2300−2600 cm−1. On the basis of previous reports and the experimental results by elemental analysis, the two Raman bands at 2324 and 2588 cm−1 were assigned to the N−N stretching vibration of N2 molecules and the H−S symmetric stretching vibration of H2S molecules, respectively, captured in the HQ clathrate framework.15,23,26 Notably, the Raman band at 2588 cm−1 for H2S in the HQ clathrates exhibited a substantial red shift from the symmetric vibrational mode (ν1) for H2S in the gas phase, leading to a frequency difference of Δν = 27 cm−1.26 However, no signal for COS was detected in the spectral range from 2500 to 3100 cm−1, as reported previously.27 These results are direct evidence suggesting that HQ captures H2S and N2 gases while excluding COS during clathrate formation. Interestingly, the relative Raman intensity of H2S at 2588 cm−1 increased with an increase in pressure from 2.0 to 8.0 MPa. This implies that H2S molecules preferentially occupy the cages in the HQ clathrate framework when they compete with N2 molecules for the best occupancy. Elemental analysis was also used to identify the captures gases and the captured amount in the HQ clathrate cages formed with the gas mixtures under various conditions. Using the elemental results obtained, the individual cage occupancy of H2S, COS, and N2 and the total cage occupancy of all guests

to be fully converted to the HQ clathrate (β form) after the reaction with the gas mixture at 8.0 MPa. However, when a pressure of 6.0 MPa or less was introduced, small peaks were also detected on both sides of the hydroxyl-substituted carbon signal, which indicate residual HQ (α form). Although such residual HQ was found to increase as the experimental pressure decreased, most of the HQ was converted into the clathrate compound with the enclathration of the gas molecules, which was based on the relatively higher intensity of the β form compared to that of the α form. Because NMR signal intensities are proportional to the number of corresponding atoms, the reaction yield (that is, the relative amounts of the α-form and βform HQ) was calculated from the obtained NMR spectra. First, using the NMR spectra for pure HQ obtained from 10 repeated measurements, the area ratio of the hydroxylsubstituted carbon atom was found to be 1:1.55:0.44 (AAD ≤ ±2.1%) with peak deconvolution.17 Although this value shows some deviation from the theoretical ratio of 2:3:1 representing three distinct hydroxyl carbons in different ratios,21 the conversion yields calculated with these values do not change significantly from those obtained with the theoretical ones (73.97, 86.05, 89.62, and 100% would be obtained with the theoretical ratios instead of 73.82, 85.93, 89.53, and 100% at 2.0, 4.0, 6.0, and 8.0 MPa, respectively). Then, the ratio was used to calculate the peak area for the αform HQ when the mixed α- and β-forms were identified from the 13C NMR spectra. These two peak areas were used to calculate the relative amount of the α form to the β form (or the conversion yield). For the H2S/COS/N2 (2.5/1.5/96 mol %) ternary gas mixture, the conversion yields were found to be 100% at 8.0 MPa and gradually decreased with decreasing pressure, resulting in 74% at 2.0 MPa. High-resolution Raman spectra were also measured to identify the crystal structure of as-prepared samples and guest incorporation in the HQ clathrate framework (Figures 3 and 4). All of the spectra showed patterns similar to those in the previous results.14,22−24 Considering the C−O and C−C stretching modes at around 1200 cm−1 and the strong C−H mode at around 3000 cm−1, the samples were found to be C

DOI: 10.1021/acs.jced.7b00248 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Figure 4. Raman spectra for pure HQ (black lines) and HQ clathrate samples after reaction with the H2S/COS/N2 (2.5/1.5/96 mol %) gas mixture at 2.0 and 8.0 MPa and ambient temperature for 14 days. All of the Raman spectra were collected at 298 K and ambient pressure.

insufficient. Meanwhile, H2S was found to have a higher concentration in the solid clathrate than in the gas phase even at 2.0 MPa, which means that it has a higher enclathration selectivity than does N2. Such selectivity seemed to increase as the experimental pressure was increased, as shown in Figure 6.

(gases) in the HQ clathrate framework were calculated (Figure 5). Because the theoretical formula of the HQ clathrate is 3HQ·

Figure 5. Cage occupancy of gas components in HQ clathrates and the stored volume of gases. The black circle, rectangle, and triangle symbols represent the cage occupancies and stored amount of gases for total gaseous components (H2S + COS + N2), H2S, and COS in the HQ clathrates, respectively.

Figure 6. Concentration enhancement of H2S in HQ clathrates compared to that in the gas phase and the conversion yield of HQ clathrates after reaction with the H2S/COS/N2 (2.5/1.5/96 mol %) gas mixture at various pressures and ambient temperature for 14 days. NMR/EA and Raman stand for the calculated results using the combination of solid-state 13C NMR and elemental analysis and the peak area ratio of Raman bands, respectively.

xgas (0 ≤ x ≤ 1), the calculated occupancies indicate the occupied percentage for one cage formed by three HQ molecules. As shown in the figure, COS was virtually not captured in the cages within the error range. We note that a little occupancy (up to 1.0%) was also detected for the HQ sample prepared with the COS/N2 (3/97 mol %) gas mixture at 8.0 MPa. Therefore, its enclathration preference is thought to be much smaller than that of N2. When formed with the H2S/ COS/N2 (2.5/1.5/96 mol %) ternary gas mixture at 8.0 MPa, the cage occupancy of COS encaged in the HQ clathrate was marginal at less than 0.3%. Because COS was reported to form a hydroquinone clathrate in the literature, 9 the low enclathration observed in this study is mainly attributed to the low partial pressure of COS. Although the relatively larger COS molecules (compared to the other components) would require higher pressure to be enclathrated, they could not be captured in the clathrate cages because the experimental pressure (or partial pressure of COS) used in this study seemed

The concentration enhancement of H2S in the solid phase of the HQ clathrate was found to be 2.2 times higher than that in the gas phase at 2.0 MPa, whereas the values were consistently ∼4−5 times higher than in the gas phase at 4.0 MPa and higher pressures. Interestingly, the total gas occupancy was kept at about 50% at pressures of 4.0 MPa and higher pressures, whereas the conversion yield gradually increased with increasing pressure up to 8.0 MPa. In addition, the cage occupancy of H2S increased in accordance with the pressure increase. This indicates that the enclathration of H2S becomes larger as the formation pressure increases. In other words, the enclathration preference should be in the order of H2S > N2 > D

DOI: 10.1021/acs.jced.7b00248 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

(5) Matson, S. L.; Herrick, C. S.; Ward, W. J. Progress on the Selective Removal of H2S from Gasified Coal Using an Immobilized Liquid Membrane. Ind. Eng. Chem. Process Des. Dev. 1977, 16, 370− 374. (6) Al-Ghawas, H. A.; Sandall, O. C. Simultaneous Absorption of Carbon Dioxide, Carbonyl Sulfide and Hydrogen Sulfide in Aqueous Methyldiethanolamine. Chem. Eng. Sci. 1991, 46, 665−676. (7) Miura, K.; Mae, K.; Inoue, T.; Yoshimi, T.; Nakagawa, H.; Hashimoto, K. Simultaneous Removal of COS and H2S from Coke Oven Gas at Low Temperature by Use of an Iron Oxide. Ind. Eng. Chem. Res. 1992, 31, 415−419. (8) Hinderaker, G.; Sandall, O. C. Absorption of Carbonyl Sulfide in Aqueous Diethanolamine. Chem. Eng. Sci. 2000, 55, 5813−5818. (9) Atwood, J. L., Davies, J. E. D., MacNicol, D. D., Eds. Inclusion Compounds; Academic Press: Orlando, FL, 1984. (10) Kang, S.-P.; Lee, H. Recovery of CO2 from Flue Gas Using Gas Hydrate: Thermodynamic Verification through Phase Equilibrium Measurements. Environ. Sci. Technol. 2000, 34, 4397−4400. (11) Ma, C.-F.; Chen, G.-J.; Wang, F.; Sun, C.-Y.; Guo, T.-M. Hydrate Formation of (CH4+C2H4) and (CH4+C3H6) Gas Mixtures. Fluid Phase Equilib. 2001, 191, 41−47. (12) Sugahara, T.; Makino, T.; Ohgaki, K. Isothermal Phase Equilibria for the Methane+Ethylene Mixed Gas Hydrate System. Fluid Phase Equilib. 2003, 206, 117−126. (13) Kang, S.-P.; Lee, J.-W. Hydrate-Phase Equilibria and 13C NMR Studies of Binary (CH4+C2H4) and (C2H6+C2H4) Hydrates. Ind. Eng. Chem. Res. 2013, 52, 303−308. (14) Lee, J.-W.; Lee, Y.; Takeya, S.; Kawamura, T.; Yamamoto, Y.; Lee, Y.-J.; Yoon, J.-H. Gas-Phase Synthesis and Characterization of CH4-Loaded Hydroquinone Clathrates. J. Phys. Chem. B 2010, 114, 3254−3258. (15) Lee, J.-W.; Yoon, J.-H. Preferential Occupation of CO2 Molecules in Hydroquinone Clathrates Formed from CO2/N2 Gas Mixtures. J. Phys. Chem. C 2011, 115, 22647−22651. (16) Lee, J.-W.; Kang, S.-P.; Yoon, J.-H. Highly Selective Enclathration of Ethylene from Gas Mixtures. J. Phys. Chem. C 2014, 118, 6059−6063. (17) Lee, J.-W.; Poudel, J.; Cha, M.; Yoon, S. J.; Yoon, J.-H. Highly Selective CO2 Extraction from a Mixture of CO2 and H2 Gases Using Hydroquinone Clathrates. Energy Fuels 2016, 30, 7604−7609. (18) Mohammadi, A. H.; Richon, D. Equilibrium Data of Carbonyl Sulfide and Hydrogen Sulfide Clathrate Hydrates. J. Chem. Eng. Data 2009, 54, 2338−2340. (19) Liang, S.; Kusalik, P. G. Crystal Growth Simulations of H2S Hydrate. J. Phys. Chem. B 2010, 114, 9563−9571. (20) Mohammadi-Manesh, H.; Alavi, S.; Woo, T. K.; Najafi, B. Molecular Dynamics Simulation of NMR Powder Lineshapes of Linear Guests in Structure I Clathrate Hydrates. Phys. Chem. Chem. Phys. 2011, 13, 2367−2377. (21) Rdpmeester, J. A. Application of solid state 13C NMR to the study of polymorphs, clathrates and complexes. Chem. Phys. Lett. 1980, 74, 536−538. (22) Kubinyi, M. J.; Keresztury, G. Infrared and Raman spectra of hydroquinone crystalline modifications. Mikrochim. Acta 1997, 14, 525−528. (23) Kim, B.-S.; Lee, Y.; Yoon, J.-H. Pressure-Dependent Release of Guest Molecules Structural Transitions in Hydroquinone Clathrate. J. Phys. Chem. B 2013, 117, 7621−7625. (24) Torré, J.-P.; Coupan, R.; Chabod, M.; Pere, E.; Labat, S.; Khoukh, A.; Brown, R.; Sotiropoulos, J.-M.; Gornitzka, H. CO2− Hydroquinone Clathrate: Synthesis, Purification, Characterization and Crystal Structure. Cryst. Growth Des. 2016, 16, 5330−5338. (25) Park, J.-W.; An, S.; Seo, Y.; Kim, B.-S.; Yoon, J.-H. TemperatureDependent Release of Guest Molecules and Structural Transformation of Hydroquinone Clathrates. J. Phys. Chem. C 2013, 117, 7623−7627. (26) Anderson, A.; Demoor, S.; Hanson, R. C. Raman Study of a New High Pressure Phase of Hydrogen Sulfide. Chem. Phys. Lett. 1987, 140, 471−475.

COS. Moreover, it should be noted that a higher enclathration preference for H2S was observed under dilute circumstances. Consequently, the concentration/separation of H2S from a gas mixture is possible using an HQ clathrate at room temperature with a total pressure of 2.0 MPa or higher. The total gas storage capacity in the clathrate compound was calculated to be 32 L under standard temperature and pressure (STP) conditions per 1 kg of HQ, which reflects 2−4 L of H2S under STP conditions (Figure 5). As described above, it is also important to note that COS was found not to be captured in the clathrate compound. Further investigations including the enclathration behaviors at various temperatures and the reaction kinetics are required to establish a clathrate-based removal process for sulfides.

4. CONCLUSIONS In this study, the application of the HQ clathrate to remove and/or concentrate two sulfides, H2S and COS, was proposed. HQ was reacted with the H2S/COS/N2 gas mixture at various pressures, and the samples were analyzed by solid-state 13C NMR and Raman spectroscopy methods. In addition, elemental analysis was also performed to identify the captured gases and the stored amount. The experimental results show that H2S is concentrated in the solid HQ clathrate (that is, it has a higher enclathration preference than that of the other gases), whereas COS was virtually not captured under all conditions. Although further investigations still need to be performed, the results in this study show the enclathration behaviors of the representative sulfides, H2S and COS, at low concentrations and are expected to provide useful information for the design of future clathrate-based separation processes for sulfides.



AUTHOR INFORMATION

Corresponding Authors

*(S.J.Y.) E-mail: [email protected]. Tel: +82-42-860-3305. Fax: +82-42-860-3134. *(J.-H.Y.) E-mail: [email protected]. Tel: +82-51-410-4684. Fax: +82-51-403-4680. ORCID

Jong-Won Lee: 0000-0002-6771-672X Sang Jun Yoon: 0000-0001-6747-1745 Ji-Ho Yoon: 0000-0002-0234-9367 Funding

This work was supported by a research grant of the Kongju National University in 2015 (no. 2015-0559-01). Notes

The authors declare no competing financial interest.



REFERENCES

(1) Beér, J. M. Combustion Technology Developments in Power Generation in Response to Environmental Challenges. Prog. Energy Combust. Sci. 2000, 26, 301−327. (2) Chauk, S. S.; Agnihotri, R.; Jadhav, R. A.; Misro, S. K.; Fan, L.-S. Kinetics of High-Pressure Removal of Hydrogen Sulfide Using Calcium Oxide Powder. AIChE J. 2000, 46, 1157−1167. (3) Graedel, T. E.; Kammlott, G. W.; Franey, J. P. Carbonyl Sufide: Potential Agent of Atmospheric Sulfur Corrosion. Science 1981, 212, 663−665. (4) Sakanishi, K.; Wu, Z.; Matsumura, A.; Saito, I.; Hanaoka, T.; Minowa, T.; Tada, M.; Iwasaki, T. Simultaneous Removal of H2S and COS Using Activated Carbons and Their Supported Catalysts. Catal. Today 2005, 104, 94−100. E

DOI: 10.1021/acs.jced.7b00248 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

(27) Maki, A. G.; Wells, J. S.; Burkholder, J. B. High-Resolution Measurements of the Bands of Carbonyl Sulfide between 2510 and 3150 cm−1. J. Mol. Spectrosc. 1991, 147, 173−181.

F

DOI: 10.1021/acs.jced.7b00248 J. Chem. Eng. Data XXXX, XXX, XXX−XXX