REVIEW ARTICLE

Semi-clathrate hydrate based carbon dioxide capture and separation techniques

  • Lijuan Gu 1,2 ,
  • Hailong Lu , 1,2
Expand
  • 1. Beijing International Center for Gas Hydrate, School of Earth and Space Sciences, Peking University, Beijing 100871, China
  • 2. National Engineering Research Center for Gas Hydrate Exploration and Development, Guangzhou 511466, China
hlu@pku.edu.cn

Received date: 02 Nov 2022

Revised date: 05 Apr 2023

Accepted date: 19 May 2023

Copyright

2023 Higher Education Press

Highlights

● Structural and thermodynamical properties of semi-clathrate hydrate are summarized.

● Properties of quaternary salts and gas mixture hydrate are summarized.

● Challenges persist in the application of semi-clathrate hydrates for carbon capture and separation.

Abstract

CO2 is considered as the main contributor to global warming, and hydrate enclathration is an efficient way for carbon capture and separation (CCS). Semi-clathrate hydrate (SCH) is a type of clathrate hydrate capable of encaging CO2 molecules under mild temperature and pressure conditions. SCH has numerous unique advantages, including high thermal stability, selective absorption of gas molecules with proper size and recyclable, making it a promising candidate for CCS. While SCH based CCS technology is in the developing stage and great efforts have to be conducted to improve the performance that is determined by their thermodynamical and structural properties. This review summarizes and compares the thermodynamic and structural properties of SCH and quaternary salt hydrates with gas mixtures to be captured and separated. Based on the description of the physical properties of SCH and hydrate of quaternary salts with gas mixture, the CO2 capture and separation from fuel gas, flue gas and biogas with SCH are reviewed. The review focuses on the use of tetra-n-butyl ammonium halide and tetra-n-butyl phosphonium halide, which are the current application hotspots. This review aims to provide guidance for the future applications of SCH.

Cite this article

Lijuan Gu , Hailong Lu . Semi-clathrate hydrate based carbon dioxide capture and separation techniques[J]. Frontiers of Environmental Science & Engineering, 2023 , 17(12) : 144 . DOI: 10.1007/s11783-023-1744-7

1 Introduction

Fossil fuels currently serve as the primary energy source for humanity, yet their combustion results in most of the greenhouse gas emission (Klara and Srivastava, 2002). From 1997 to 2020, carbon emissions produced by burning fossil fuels and manufacturing cement have almost doubled from 5.0 GtC/yr to 9.5 GtC/yr (Friedlingstein et al., 2022). With concerns on the climate change, the reduction of carbon emission has been highly valued. As climate change concerns intensify, reducing carbon emissions has become a crucial priority. The Intergovernmental Panel on Climate Change (IPCC) report highlights the need to limit global warming to 1.5 °C, requiring the peak of carbon dioxide emissions by 2030 and striving for carbon neutrality by 2060 (Yang et al., 2022). CO2 is considered as the main contributor to the global warming and its emission have increased continuously results from human activities. Carbon dioxide (CO2) is believed to be the primary contributor to global warming, with human activities causing a continuous increase in CO2 emissions. According to NASA’s climate change website, CO2 concentration in January 2023 was 419 ppm, about 33% higher than in 1958, and the rate of increase is accelerating. CO2 capture and separation is essential to reduce greenhouse gas emission to achieve the net-zero carbon emission goal (Davis et al., 2018).
Several CO2 capture methods have been developed, including absorption (Zhuang et al., 2016), adsorption (Ben-Mansour et al., 2016; Chen et al., 2013), membranes (Kárászová et al., 2020; Favre, 2022), cryogenics (Font-Palma et al., 2021), microbial/Algal systems and so on (Wilberforce et al., 2019; Wang et al., 2020a). Compared with the above-mentioned methods, hydrate-based CO2 capture and separation a promising technology for large-scale applications due to its high gas storage capacity, ease of recycling, safety, and cost-effectiveness (Wang et al., 2020a). Clathrate hydrates are crystalline compounds formed by hydrogen-bonded water cages enclosing guest molecules or ions (Jeffrey and McMullan, 1967). Depending on the interaction forces between the host cage and the guest molecules, clathrate hydrate can be divided into several types. Gas hydrates are “true” clathrate hydrates in which the guest molecules with appropriate sizes stabilize the structures through the van der Waals interaction with the water cages. In peralkylonium salt hydrates, the onions participate in the host framework construction and hydrogen-bonded with the water molecules (Dyadin and Udachin, 1987). The cations fill the voids of the framework and ionically interact with the water-ionic polyhedral framework. Jeffrey categorized this type of hydrate as the semi-clathrate hydrate (SCH) since the nature of the host lattice is not necessarily changed (Jeffrey and McMullan, 1967). Semi-clathrate hydrates are stable under ambient pressure and above the ice point, even above room temperature (Dyadin and Udachin, 1984; 1987), due to the ionic and van der Waals interactions between guest cations and the water-ion polyhedral structure.
Semi-clathrate hydrates can be utilized for various applications, including gas separation (Sun et al., 2011a; Zhong and Englezos, 2012; Zhong et al., 2018), H2 storage (Chapoy et al., 2007; Deschamps and Dalmazzone, 2010; Veluswamy et al., 2014) and air conditioning system (Stoporev et al., 2020; Rakkappan et al., 2021; Kim et al., 2022; Kim et al., 2023). Using semi-clathrate hydrates for carbon capture and storage (CCS) has several benefits, including mild operating conditions, recyclability, and high selectivity for CO2. Captured CO2 can be abandoned as solid clathrate hydrate, while the purified methane and hydrogen obtained from fuel and biogas can be used as clean energy sources. The salts used to form semi-clathrate hydrates can be produced on a large scale and at low cost. Furthermore, the salt solution can be recycled as no salts are consumed during the CO2 capture and release process. Semi-clathrate hydrates can significantly simplify CCS facilities and reduce the compression cost of gas hydrate formation, which is the primary cost of CO2 capture.
Various ionic substances can form semi-clathrate hydrates, such as tetra-n-butyl ammonium (TBA), tetra-n-butyl phosphonium (TBP) and tetra-iso-amyl ammonium (TiAA) for the cation and halide, hydroxide and carboxylate for the anion (Dyadin and Udachin, 1984;1987). TBA bromide (TBAB) (Oyama et al., 2005; Shimada et al., 2005a; 2005b; Miwa et al., 2018), TBA chloride (TBAC) (Haruo, 1982; Sato et al., 2013; Kida et al., 2019), TBA floride (TBAF) (Udachin and Lipkowski, 2002; Rodionova et al., 2008), TBP bromide (TBPB) (Muromachi et al., 2014a), and TBP chloride (TBPC) (Sakamoto et al., 2011) are the application hotspots because of the high melting temperatures around 280 K of their hydrates and will be the main content of this paper. For purpose of simplicity, we call these salts the quaternary salts.
Although CCS technology based on semi-clathrate hydrates is still in the development stage, significant efforts are underway to improve its performance. For example, selecting a proper quaternary salt can reduce hydrate formation pressure and increase the amount of captured gas. A better understanding of the phase behavior and structural properties of semi-clathrate systems is also necessary for the design of CO2 capture processes based on hydrate technology. This review will focus on the fundamental properties of SCH and their applications in CO2 capture and separation. The structure and thermal stability of SCH as well as the ternary hydrate of gas mixture and quaternary salts will be described. Finally, we will summarize the application of semi-clathrate hydrates in CO2 capture and separation of fuel gas, flue gas, and biogas, analyze the challenges and future directions in this field, and draw a conclusion.

2 Structure of semi-clathrate hydrate with gas mixture

The clathrate hydrate of quaternary salts is a kind of complex semi-clathrate hydrate with combined cavities to accommodate the large guest cautions. The typical cavities of quaternary salts hydrate include the pentagonal dodecahedron (D), tetrakaidecahedron (T) and pentakaidecahedron (P) (Dyadin and Udachin, 1987). As shown in Tab.1, the unit cells of the idealized host frameworks can be described as 6T·2D·46H2O (CS-I), 4P·16T·10D·172H2O (TS-I), 2P·2T·3D·40H2O (HS-I) and their super lattices. The pentagonal dodecahedron, i.e., the D cavity, is the only polyhedron that is common to all the clathrate hydrates. The host framework is formed based on D cavity with other polyhedrons through vertex linking or face sharing (Dyadin and Udachin, 1987). However, for a certain hydrate, stable structures with different host lattices or several crystal structures with the same host lattice are possible. Detailed crystal structure determination is prerequisite to characterize the clathrate hydrate of quaternary salts.
Tab.1 Hydration numbers in the cages in the semi-clathrate hydrates (Davidson, 1973; Muromachi and Takeya, 2017)
Structure Orthorhombic Tetragonal Cubic
512 cages (A) 6 10 2
51262 cages (B) 4 16 6
51263 cages (C) 4 4 0
water molecules (D) 80 172 40
The host cage of SCH is formed by the anion of the semi-clathrate promoter and the water molecules through hydrogen bond, in which the cation of the semi-clathrate promoter as the guest molecule is hydrophobically included. This section summarizes and compares the cage structure and related unit cell parameters of semi-clathrate hydrates formed by tetra-n-butyl ammonium halide and tetra-n-butyl phosphonium halide. Additionally, the crystal structure of quaternary salts and gas double hydrate is also described to provide a better understanding of the performance of SCH in CCS.

2.1 Structure of semi-clathrate hydrate

2.1.1 TBAB

Under certain conditions, polyhydrates can form in the TBAB-H2O system, where the Br ion is bonded with water molecules to incorporate the TBA cation. There are five TBAB SCH from the literature, including one orthorhombic hydrate TBAB·38H2O (Oyama et al., 2005; Shimada et al., 2005b), three tetragonal hydrates with different stoichiometry (24, 26 and 32) (Gaponenko et al., 1984; Rodionova et al., 2013), and a trigonal hydrate with small hydration number of 21 /3 (Lipkowski et al., 2002). The structural parameters of TBAB semi-clathrate hydrates are summarized in Tab.2. By rapidly cooling of the TBAB solution, another metastable hydrate of cubic structure I was discovered from the X-ray diffraction pattern (Stoporev et al., 2021). However, the structural parameters could not be resolved because of the poor quality of the powder diffraction patterns.
Tab.2 Structural parameters of TBAB semi-clathrate hydrate
Hydrate composition Crystal system Space group Unit cell (Å) Ref.
TBAB∙21/3H2O Trigonal R3c a = 16.609(1)c = 38.853(1) Lipkowski et al. (2002)
TBAB∙24H2O Monoclinic C2/m a = 28.5(2)b = 16.9(1)c = 16.5 (1) β = 125 ° Gaponenko et al. (1984)
TBAB∙26H2O Tetragonal P4/mmm a = 23.9(2)c = 50.8(5) Gaponenko et al. (1984)
TBAB∙32H2O Tetragonal P4/m a = 33.4(3)c = 12.7(1) Gaponenko et al. (1984)
TBAB∙32H2O Tetragonal a = 23.65c = 12.50 McMullan and Jeffrey (1959)
TBAB∙36H2O Orthorhombic Pmmm a = 21.3(2)b = 12.9(1)c = 12.1(1) Gaponenko et al. (1984)
TBAB∙38H2O Orthorhombic Pmma a = 21.060(5)b = 12.643(4)c = 12.018(8) Shimada et al. (2005b)
So far, only the crystal structure of TBAB·38H2O and TBAB∙2 1/ 3H2O were solved. The ideal unit cell of TBAB∙38H2O is 4T·4P·6D·76H2O, with TBA cation located in the four large cages, i.e., 2T·2P cage (Shimada et al., 2005b). As shown in Fig.1(a), the host framework is constructed by Br atoms and water molecules through hydrogen bonds, which is distorted from the idealized water framework because the bond length between Br and oxygen atom is larger than that between two oxygen atoms. The TBA cautions locate at the center of four cages and part of the cage structure is broken, as shown by the dotted lines in the figure. The small D cages are empty and can encage small gas molecules in gas separation and storage applications. There are two types of D cages with distinct symmetry in the crystal lattice, namely the highly distorted DA cages and more regular cages DB (Muromachi et al., 2014b). The chemical formula of the empty host framework can be written as TBAB∙38H2O∙DA∙2DB. The DA cages are highly distorted because the water molecular forming the DA cages are displaced with the TBA+ and pushed inward by the Br. Therefore, it is more suitable to incorporate the bar-shaped CO2 molecule. The single crystal of TBPB SCH was determined to have the same structure with TBAB∙38H2O, while no polymorphism was found (Muromachi et al., 2014a). Similar to the orthogonal TBAB SCH, two types of D cages were found in TBPB SCH that can be written as TBPB∙38H2O∙DA∙2DB. The D cages in TBPB SCH show larger distortion than those in TBAB SCH because the bond length of C–P is larger than that of C–N.
Fig.1 Structure of SCH: (a) TBAB∙38H2O (Shimada et al., 2005b); (b) TBPB∙38H2O (Muromachi et al., 2014a); (c) TBAF∙29.7H2O (Komarov et al., 2007); (d) TBAF∙5.5H2O (Udachin and Lipkowski, 2002).

Full size|PPT slide

In the concentrated solution of TBAB and water binary system, TBAB∙21/3H2O can crystallize in the trigonal space group R3c (Lipkowski et al., 2002). The structure of the water-anionic cluster consists of four tetragonal faces and two pentagonal faces. Unlike the traditional host framework of semi-clathrate hydrate, the water-anionic clusters can transform into guests of host cavity formed by six cross-like tetra-n-butyl ammonium cations. This means that, in the TBAB and water binary system, water molecules can both function as host and guest, depending on the concentration of the system.
It has been found that a tetragonal structure is typical for n-butyl compositions with hydration number of 32 (McMullan and Jeffrey, 1959). Despite differences in the unit cell parameters of TBAB∙32H2O obtained in previous studies (McMullan and Jeffrey, 1959; Gaponenko et al., 1984), its detailed structure has not yet been constructed. The crystallographic parameters of all the tetragonal TBAB hydrate were determined with slightly differences, as shown in Tab.2. Rodionova et al. (2013) built structural models based on the tetragonal structure for crystal samples prepared with different concentrations of the solution. However, the structural stoichiometry was not available. A particular difficulty was encountered when growing single crystals of TBAB∙24H2O, and it only succeeded three times during years of attempts (Gaponenko et al., 1984). Resolving the crystal structure of these metastable phase of TBAB hydrate is very challenging. It is difficult to keep the metastable state to grow single crystals with proper quality, which can be easily affected by the environment during hydrate growth.

2.1.2 Other quaternary salts

All polyhydrates of tetra-n-butyl ammonium chlorine (TBAC) have a tetragonal structure with a space group of P42/m, as shown in Tab.3 (Rodionova et al., 2010). The cation of quaternary ammonium salts (QAS) is located in either the 4T cavities or 3T and 1P cavity of the fused cages. The chlorine ions are hydrophilically included in the hydrate framework.
Tab.3 Structural parameters of other quaternary salts semi-clathrate hydrate
Hydrate composition Crystal system Space group Unit cell (Å) Ref.
TBAC∙24 H2O Tetragonal P42/m a = 23.608±0.028c = 12.561±0.003 Rodionova et al. (2010)
TBAC∙30H2O a = 23.733±0.013c = 12.513±0.001
TBAC∙32H2O a = 23.737±0.011c = 12.492±0.001
TBAF∙5.5H2O Monoclinic C2/c a = 16.590(3) b = 17.040(3) c = 16.510(3) β = 103.18(3)° Udachin & Lipkowski (2002)
TBAF∙29.7H2O Cubic I4¯3d a = 24.375(3) Komarov et al. (2007)
TBAF∙32.8H2O Tetragonal P42/m a = 23.52(1) c = 12.30(1) Rodionova et al. (2008)
TBPB∙38H2O Orthorhombic Pmma a = 21.065(5) b = 12.657 (3) c = 11.992(1) Muromachi et al. (2014)
TBPC∙38H2O Tetragonal a =23.7 c = 12.5 Sakamoto et al. (2011)
The structures of clathrate hydrates formed in the TBAF-water binary system have all been characterized (McMullan et al., 1963; Udachin and Lipkowski, 2002). The hydration number depends on salt concentrations, with TBAF∙29.7H2O and TBAF∙32.8H2O hydrates formed under relatively dilute solutions and TBAF∙5.5H2O hydrate found at very high salt concentration (Udachin and Lipkowski, 2002). As illustrated in Fig.1(c), TBAF∙29.7H2O hydrate exhibits a cubic structure and the host framework is described as a distorted version of the idealized CS-I water framework with unit cell of 6T·2D·46H2O (Komarov et al., 2007). The four T-cavities are fused to accommodate the giant cations, and the symmetry of the idealized water lattice is distorted to have a space group of I4¯3d. The structure of TBAF∙32.8H2O is similar to that of the polyhyrates of TBAC.
The TBAF∙5.5H2O semi-clathrate hydrate has a monoclinic structure with space group of C2/c (Udachin and Lipkowski, 2002). As shown in Fig.1(d), the polyhedron formed by water molecules and the fluorine ions consists of two square faces and four pentagonal faces. The water-anionic clusters are connected with each other to form a chain surrounded by the cations. Similar to TBAB hydrate, the water molecules can function as either the guest or the host, depending on the concentration of solution. The unit volume of TBAF∙5.5H2O is the smallest among all the quaternary salts hydrates, and the cavity in this hydrate is the smallest. As a result, this structure is not suitable for gas capture and separation.
Compared to quaternary ammonia salts, clathrate hydrate of quaternary phosphonium salts (QPS) are less investigated. Only the molecular structure of the tetrabutylphosphonium bromide (TBPB) has been resolved (Muromachi et al., 2014a). As shown in Fig.1(b), TBPB forms a simple hydrate of TBPB∙38H2O with an orthorhombic structure. Similar to TBAB∙38H2O, the tetrabutylphosphonium cation occupies four large cages of 2T∙2P. However, the different bond length between C-P of the TBP ion with 1.73 and C–N of the TBA ion with 1.53 contributes to the different thermal stability of the TBPB and TBAB semi-clathrate hydrate. The distorted D cage in TBPB hydrate is more compressed than that in TBAB, although the volumes of these cages are nearly the same, which corresponds to the different gas separation performance for specific gas molecules.
Similar to TBAC, the structure of TBPC is tetragonal with approximate molecular size (Sakamoto et al., 2011). However, the detailed structure parameters of TBAC hydrate are not available. The structure of TBPF have not been reported to date.
Among all the aforementioned polyhydrate structures, the tetragonal structure exists in all three QAS hydrates. The structure type is related to the salt concentration. In the TBAB-H2O binary system, the orthogonal TBAB ∙38H2O forms at dilute TBAB solution. As the salt concentration increases, the tetragonal structure appears until the water molecule transforms into a guest that enclosed by the TBA caution to form the trigonal structure. In the TBAF-H2O binary system, structural transition occurs with an increase in concentration from the tetragonal to the cubic structure, and the water molecules transforms into guests at very high salt concentration. However, because the structural information of QPS is rather limited, the relationship between salt concentration and structure is not straightforward.

2.2 Structure of semi-clathrate hydrate with gas mixture

In quaternary salts hydrates, the host framework is formed through the hydrophilic interaction between the salt onion and water molecules. The guest cation occupies the space created by the four fused cages, leaving the small D cages vacant that can incorporate small molecules like CO2, H2, CH4, N2, etc. The guest gas molecules can be incorporated into the SCH under mild conditions, even milder than the quaternary salt hydrate because guest molecules can stabilize the SCH. Therefore, quaternary salts have been extensively studied for carbon dioxide gas capture and separation. To achieve better performance in these applications, the structure of quaternary salts and hydrate of quaternary salts with gas mixture to be captured and separated should be characterized.
TBAB is the most extensively studied quaternary salt. It can form polyhydrates with trigonal, tetragonal and orthorhombic structures. In addition to the trigonal TBAB·21/3H2O, the other structures are relatively stable under ambient temperature. It has been found that the guest gas molecules prefer to occupy the empty cages of orthorhombic structure (Muromachi et al., 2014b; 2016a; 2016b). This may be because among all the SCH structures, the orthorhombic one has the most D cages per water molecule (Muromachi and Takeya, 2017), thus offering superior gas capacity compared to the others.
In the hydrate structures of TBAB, as well as those formed with tetra-n-butyl ammonium and tetra-n-butyl phosphonium salts, there are two distinct types of D cages with significantly different shapes. The strongly distorted D cage is referred to as DA, while the relatively regular cages are labeled DB and further divided into DB1 and DB2 (Muromachi et al., 2016b). The structures of double hydrate formed with TBAB and CO2/CH4/N2 molecules were characterized and summarized in Tab.4. CO2 gas molecules had an asymmetrical distribution among D cages, with an occupancy of 0.867 in the strongly distorted cages and an occupancy of 0.49 in the regular D cages (Muromachi et al., 2014b). The symmetry of CO2 and TBAB double hydrate was lowered to Imma, and the unit cell in the direction of b-axis was doubled compared to the Pmma structure.
Tab.4 Summary of TBAB and gas binary hydrate crystal parameters
TBAB + CO2 TBAB + CH4 TBAB + N2
TBAB Concentration (wt%) 0.1 0.1 0.2
PT conditions 1.08 MPa, 282.65 K 2 MPa, 284.5 K 5.8 MPa, 283.9 K
Empirical formula TBAB·38H2O·1.85CO2 TBAB·38H2O·2.16CH4 TBAB·38H2O·1.5N2
Crystal system, space group Orthorhombic, Imma Orthorhombic, Pmma Orthorhombic, Pmma
Unit cell dimensions (Å) a = 21.0197(7)b = 25.2728(8)c = 12.0096(4) a = 21.0329(15)b = 12.5972(9)c = 12.0333(8) a = 21.035(4)b = 12.635(3)c = 12.021(2)
DA cage occupancy 0.867 0.174 0
DB cage occupancy 0.490 0.991 0.75
Average occupancy 0.616 0.719 0.5
Ref. Muromachi et al. (2014b) Muromachi et al. (2016b) Muromachi et al. (2016a)
In contrast to CO2, the occupancy of CH4 in DA and DB was 0.174 and 0.991, respectively (Muromachi et al., 2016b). The structurally regular DB cage hold CH4 gas more tightly than the DA cage with low symmetry. This finding provides guidance for CO2 separation applications from CH4 gas by employing ionic salts that form SCH with more asymmetric cages, thus the CO2 separation performance can be significantly improved. As described below, the CO2 selectivity of TBAC SCH is superior to that of TBAB SCH (Hashimoto et al., 2017b). The unit cell of tetragonal TBAC SCH can be described as 4T3P∙T4∙4DL∙4DM∙2DN (Muromachi et al., 2022). Among all the D cages, the DL cage can encapsulate water and chloride while rejecting CO2 gas, and the remaining distorted DM and DN cages can encage CO2 gas with slightly lower gas occupancy than DA cages in TBAB SCH. There are more distorted D cages in TBAC SCH that show superior CO2 selectivity than TBAB SCH.
N2 gas molecules were observed to only occupy the regular D cages and not the highly distorted D cages in TBAB semi-clathrate hydrate with concentration of 20 wt% (Muromachi et al., 2016a). Additionally, the phase properties of TBAB and N2 mixed hydrate were found to be different under various pressure conditions. It was observed that when the N2 gas pressure exceeded a certain level, it could be incorporated into the TBAB hydrate. In the study by Muromachi et al. (2016a), this critical pressure was found to be 4 MPa. Therefore, to perform CO2 separation with TBAB from flue gas, it is important to determine the critical pressure to prevent N2 enclathration in hydrate cages. It should be noted that at a concentration of 30 wt%, a tetragonal hydrate similar to TBAB·26H2O was found to form in the TBAB-N2-H2O system (Jin et al., 2019), which suggests that a high concentration of quaternary salt may not be desirable for gas separation.
The structure of quaternary salts SCH with H2 have not yet been determined. Raman spectroscopy showed that H2 molecules only occupy the small D cages of TBAB SCH, and their cage occupancy is independent of TBAB concentrations (Hashimoto et al., 2006; 2008). The inclusion of gas molecule in TBAB SCH from CO2/H2 gas mixture was found to be affected by the gas composition. In the case of gas mixture with 40 mol% CO2, CO2 gas molecule was found both in the 51262 and 51263 cages. However, in the case of 18 CO2/82% H2 and 10 CO2/90% H2, CO2 gas molecules were only incorporated into the 51263 cages. In all cases, no H2 was detected in the small cages. A similar phenomenon was discovered in stimulated biogas (45%CO2/55%CH4 gas mixture) by Raman spectroscopy (Xia et al., 2016). CO2 was found to enter both the 51262 and 51263 cages, while CH4 was only incorporated into the 512 cages.
Single crystals of TBAB, TBAC, TBPB and TBPC with 15% CO2/85% N2 gas mixture were formed and characterized, as summarized in Tab.5 (Hashimoto et al., 2017b). The mixed hydrate of TBAB + 15%CO2 + 85%N2 had the same structure as TBAB, and the CO2 mixed hydrate had an orthorhombic structure with space group of Imma, with symmetry lowered compared to TBAB·38H2O semi-clathrate hydrate. This demonstrates that the symmetry lowering is dominated by CO2, even with a mole fraction of only 15% in the gas mixture. The clathrate of TBPC and the gas mixture also had an orthorhombic structure with space group of Cmmm. Due to the poor quality of the single crystals, the TBPB and gas mixture hydrate was probably hexagonal or orthorhombic. Like the TBAB semi-clathrate hydrate, the hydrate after incorporating gas mixtures remained tetragonal.
Tab.5 Summary of quaternary salts and CO2/N2 mixed hydrate crystal parameters (Hashimoto et al., 2017b)
TBAB + CO2 + N2 TBAC + CO2 + N2 TBAF + CO2 + N2 TBPB + CO2 + N2 TBPC + CO2 + N2
Concentration (wt%) 0.2 0.2 0.2 0.2 0.2
PT conditions 5.04 MPa, 284.2 K 5.03 MPa, 287.2 K 298.6K, 3.01MPa 4.93 MPa, 285.2 K 5.09 MPa, 285.2 K
Crystal system, space group Orthorhombic, Imma Tetragonal,P24 /m Tetragonal,P24 /m Hexagonal (Possibly orthorhombic) Orthorhombic, Cmmm
Unit cell dimensions (Å) a = 21.419(4)b = 25.833(5)c = 12.218(2) a = 23.870(3)c = 12.497(3) a = 23.301 (3)c = 12.179 (2) a = 12.0602(17)c = 12.585(3) a = 12.036(2)b = 21.145(4)c = 12.685(3)
TBAF was found to better stabilize CO2 hydrate compared to other quaternary salts mentioned in this paper. The structure of TBAF and 12% CO2/N2 mixed hydrate was characterized to explain its gas separation performance (Hashimoto et al., 2020). Polyhydrates of TBAF semi-clathrate hydrate can form at different concentrations, including cubic, tetragonal and monoclinic (McMullan et al., 1963; Udachin and Lipkowski, 2002). After incorporating the CO2/N2 gas mixture into the TBAF semi-clathrate hydrate, only columnar-shaped crystals with tetragonal structure were found. There were ten D cages per unit cell, while only two D cages per TBAF were available for gas capture. Therefore, TBAF captured less gas than orthorhombic semi-clathrate hydrates with three D cages per salt, like TBAB, TBPB and TBPC.

3 Thermodynamical properties of semi-clathrate hydrate with gas mixture

The thermodynamic properties of semi-clathrate hydrates have been extensively studied as they are essential parameters for practical applications. One unique advantage of the semi-clathrate hydrate over gas hydrates in different application scenarios is that they can be formed at ambient pressure and relatively high temperature. Phase equilibrium diagrams, which represent the thermodynamic properties of semi-clathrate hydrates, can provide a technical foundation for industrial applications and also aid in the study of their dynamic properties. In this section, four methods for measuring the phase equilibrium conditions of semi-clathrate hydrate are introduced, including visual observation method, differential scanning calorimetry, pressure search method and the Schreinemaker’s method. Additionally, the thermodynamical properties of clathrate hydrate of quaternary salts and the mixed hydrate of quaternary salts with gases and gas mixtures are summarized.

3.1 Phase equilibrium condition measurement

The visual observation apparatus provides a simple means of measuring the equilibrium temperature of the SCH with melting temperature typically above the point (Suginaka et al., 2012). As shown in Fig.2(a), the dissociation and nucleation process can be directly observed through an optical window made of transparent glass or sapphire glass using an optical microscope (Sakamoto et al., 2011). To perform the measurement, a solution of the guest substance is filled into a glass tube, and the temperature is controlled by circulating a cooling liquid. SCH is formed by lowering the system temperature and then incrementally increase it in steps of ΔT. At each temperature step, the system temperature is held for several hours to achieve equilibrium. If SCH remains stable, the temperature is increased further, otherwise if the hydrate dissociates at a temperature of T, the equilibrium temperature of this SCH will be TΔT. This technique has the advantage of high accuracy, but is only suitable for measuring the equilibrium temperature above the ice point, as the optical apparatus is unable to distinguish between hydrate and ice. Additionally, this method is time-consuming, taking days to determine the equilibrium condition in a single run.
Fig.2 Schematic diagram of (a) the visual observation apparatus (Sakamoto et al., 2011), (b) isochoric pressure-search technique.

Full size|PPT slide

Ye and Zhang (2014a) used a different method based on visual observation to determine phase equilibrium, but with a different heating protocol. Instead of step heating after hydrate formation, the system temperature is continuously increased at a constant speed of 0.1 K/h. When the hydrate is about to disappear, the heating rate is reduced to one-third of the initial value until hydrate completely dissociates. The corresponding temperature and pressure are considered as the phase equilibrium data. The continuous heating protocol reduces the required time for each measurement. According to the comparative study of step heating and continuous heating protocols (Tohidi et al., 2000), the error associated with continuous heating is a function of the heating rate. Since the heating rate is slow, the measurement is reliable, and the accuracy is consolidated by the phase equilibrium data of nitrogen gas which is consistent with the literature.
The visual observation apparatus typically operates at ambient pressure and can be used to investigate the phase equilibrium condition of SCH. However, in some applications that use SCH as gas hydrate promotor, the phase equilibrium determination requires relatively high pressure. In such cases, differential scanning calorimetry (DSC) is an effective technique that can determine stability conditions and provide the dissociation enthalpy, which is an important parameter in phase-changing material applications (Kharrat and Dalmazzone, 2003; Deschamps and Dalmazzone, 2009; Mayoufi et al., 2011). DSC is more efficient, but its measurement accuracy is compromised. The confidence limits are determined by the calibration results of the melting point of pure samples with known melting points, such as ice. In practical conditions, these confidence limits can be as high as ±0.4 °C (Mayoufi et al., 2010), causing systematic deviation between measurement results by different groups. Additionally, the equilibrium temperature is deemed as the onset on the DSC curve (Mayoufi et al., 2011), which is lower than the visual observation method that corresponds to the point at which all the hydrates disappear.
Different types of SCH with similar equilibrium temperatures can form in the quaternary salt-water binary system under certain temperature and ambient pressure. These SCHs are difficult to distinguish by visual observation, and their corresponding peak are always mixed together on DSC curves. Oyama et al. (2005) proposed two measuring methods to individually measure the thermophysical properties of type A and type B TBAB hydrate. The visual observation method was used to determine the lower melting temperature, while the autoclave system was adopted to obtain the higher melting temperature (Oyama et al., 2003). During the measurement, SCH was first formed under relatively low temperature and the system temperature was then raised at a constant heating rate, say 2 °C/h. If no phase changes occur, the temperature of the sample fluids will increase with the same rate. Because the hydrate dissociation is endothermic, the inflection point of the temperature curve of the sample fluids corresponds to the phase change points, i.e., the higher melting points. Oyama et al. (2003) successfully determined the congruent melting point of both type A and type B TBAB hydrate, which is 12 and 9.9 °C, respectively. This result is consistent with the results measured by visual observation method, which can only determine the congruent melting point of the type A TBAB hydrate (Shimada et al., 2003).
The isochoric pressure-search technique is commonly used to measure the phase equilibrium conditions for gas and quaternary salt solution mixtures. As shown in Fig.2(b), the system is initially cooled down, and at a certain point, the pressure drops abruptly due to hydrate formation. The temperature is then stabilized for a few hours to ensure the complete hydrate formation. Then the system is heated stepwise (Arjmandi et al., 2007; Chapoy et al., 2010) or continuously (Joshi et al., 2013; Sánchez-Mora et al., 2019) until a change in the slope formed between the temperature and pressure conditions is observed. The inflection point, labeled with a red pentagram in Fig.2(b), is the dissociation point.
Schreinemakers’ method is a powerful technique for deriving the phase diagram, yet it has been largely overlooked in the field of chemical engineering. One of the most remarkable features of this method is that it can derive the P-T diagram of a given system based solely on knowledge of the chemical compositions, even when little or no experimental data are available. However, only a few studies have taken advantage of this method to obtain phase equilibrium diagrams (Aladko and Dyadin, 1996). A comprehensive discussion on the use of Schreinemaker’s geometric approach to obtain the complete phase diagram for hydrates is presented by Mehta et al. (1996).

3.2 Thermodynamical properties of semi-clathrate hydrate

3.2.1 TBAB

The equilibrium temperature of TBAB under ambient pressure has been extensively studied by numerous research groups. In the TBAB-water binary system, the guest molecules can stabilize different frameworks under various conditions, as reported in previous studies (Gaponenko et al., 1984; Rodionova et al., 2013; Oshima et al., 2018). However, some studies have reported slightly deviating results in their measurements, as shown in Fig.3, due to various factors such as differences in measurement techniques, the volume of the salt solution used (Tohidi et al., 1994) and polymorphism.
Fig.3 (a) Equilibrium temperature of TBAB SCH under ambient pressure with no consideration of polyhydrates (Sun et al., 2008; Deschamps and Dalmazzone, 2009; Sato et al., 2013; Kobori et al., 2015a; Wang and Dennis, 2015); (b) Equilibrium temperature of Type A and Type B TBAB SCH (Darbouret et al., 2005; Oyama et al., 2005; Shimada et al., 2005a; Hashimoto et al., 2008; Ma et al., 2010).

Full size|PPT slide

It is known that hydrate formation may not be repeatable, as it depends on factors such as cooling rate, water history and subcooling degree.Therefore, the phase equlibrium curve is typically obtained during hydrate dissociation. Wang and Dennis (2015) used a step cooling procedure to determine the formation temperature, which resulted in a noticeable from the other measurements. The equilibrium temperature determined by DSC refers to the onset of hydrate dissociation (Deschamps and Dalmazzone, 2009), which may be lower than that observed by optical observation (Sun et al., 2008; Sato et al., 2013; Kobori et al., 2015a), corresponding to almost complete hydrate dissociation temperature. However, the results obtained through visual observation methods deviate from each other, possibly due to differences in the volume of salt solution used. The more of the fluid, the longer time should be kept to allow the system in equilibrium. The volumes of salt solution used in Sato et al. (2013) and Kobori et al. (2015a) were 4.5 and 0.7 cm3, respectively, and their results are consistent with each other. Sun et al. used a larger volume of 80 cm3, and their equilibrium temperature was approximately 0.15–2.85 K lower than the others, with a linearly decrease in temperature difference with increasing TBAB solution.
Fig.3(a) displays the phase equilibrium behavior of TBAB as a function of its mass fraction, which exhibits an initial increase followed by a slow decrease at high concentrations. The highest temperature in this curve corresponds to the congruent melting temperature (Oyama et al., 2005), which has been reported to occur at a concentration range of 35–40 wt% in Deschamps and Dalmazzone (2009) and Kobori et al. (2015a). However, Sato et al. (2013) have shown that this range is narrower and falls between 36 and 37 wt%.
Tab.2 summarizes the various polyhydrates of TBAB, with TBAB∙26H2O and TBAB∙38H2O being the most easily formed, also known as type A and type B hydrates (Shimada et al., 2003; 2005a). The thermodynamical properties of these two hydrates are depicted in Fig.3(b), which has been reported by several studies (Darbouret et al., 2005; Oyama et al., 2005; Shimada et al., 2005a; Hashimoto et al., 2008; Ma et al., 2010). Oyama et al. (2005) combined visual observation and isochoric continuous heating method to obtain the phase equilibrium temperature. Ma et al. (2010) used DSC to record the melting temperature of TBAB SCH with a heating rate of 0.5 °C/min. As shown in Fig.3(b), when the TBAB concentration is below 19 wt%, type B TBAB SCH is typically more stable than type A hydrate, a result that is better informed at solution concentrations above 20 wt%. The congruent melting points are approximately 40 wt% for type A and 32 wt% for type B, with melting temperature of 285.1 and 282.9 K, respectively. These findings are consistent with the literature (Gaponenko et al., 1984). The congruent melting point of 36 wt% and 37 wt% reported by Sato et al. (2013) may indicate the coexistence of these two types of TBAB SCH.

3.2.2 Other quaternary salts

The equilibrium temperature of TBAC has also been reported with some inconsistency. As shown in Fig.4, the measurements by Sun et al. (2011b) are 1.1–1.3 K lower in all concentration than those obtained by Nakayama (1987) and Sato et al. (2013). While both Sun et al. (2011b) and Sato et al. (2013) used visual observation technique, the fluid volume used was 100 and 4.5 cm3, respectively. TBAC is known to have three structures, and their melting points and stoichiometry have been reported in the literature and are summarized in Tab.6. Unlike TBAB, where different structures have markedly different melting points, all three types of TBAC SCH have tetragonal structures with slightly different melting points. The congruent melting points determined from Fig.4 are found to be between the most stable TBAC·30H2O at 15.2 °C and the metastable phase of TBAC·24H2O at 14.7 °C.
Tab.6 Stoichiometry, melting point and measurement method of TBAC SCH
Stoichiometry m.p. (°C) Measurement method Ref.
TBAC·(32.2±0.4)H2O 15.0 DSC Rodionova et al. (2010)
TBAC·(32.21±0.28)H2O 15.1 Schreinemaker’s method Aladko & Dyadin (1996)
TBAC·30H2O 15 Visual observation Nakayama (1982)
TBAC·30H2Oa 15.2 DSC Nakayama (1987)
TBAC·(29.7±0.4)H2O 15.1 DSC Rodionova et al. 2010)
TBAC·(29.4±0.29)H2O 15.1 Schreinemaker’s method Aladko & Dyadin (1996)
TBAC·(24.8±0.3)H2O 14.9 DSC Rodionova et al. (2010)
TBAC·(24.11±0.16)H2O 14.7 Schreinemaker’s method Aladko & Dyadin (1996)

Note: a) Hydration number is around 30.

As shown in Fig.4, the equilibrium temperature curves of TBAF coincide with each other (Sakamoto et al., 2008; Lee et al., 2010; Mohammadi et al., 2013a). The melting points are consistently reported to be 27.6 °C. This can be attributed to the fact that the polyhydrates of TBAF, namely TBAF·32H2O and TBAF·28H2O, have similar melting point of 27.2 °C (Dyadin et al., 1977) and 27.7 °C (Komarov et al., 2007).
As shown in Fig.4(b), there are discrepancies among the reported equilibrium data of TBPB from literature that may be attrbuted to the measurement techniques used. Suginaka et al. (2012) employed optical observation methods, while the other studies used DSC. The differences among the DSC results originate from the different dissociation points, as shown in Fig.5. Zhang et al. (2013) used the peak temperature Tpeak, which better fit the data from Suginaka et al. (2012) than Tonset, which is used to determine the dissociation temperature of the isothermal phase transition. The corrected peak temperature (TGEFTA) (Höhne et al., 1990) and the end temperature (Tend) were used to estimate the non-isothermal phase transition temperature. Lin et al. (2013) employed two DSC procedures, namely dynamic and stepwise schemes, to estimate the accuracy of different temperatures. The results show that Tend is more accurate than TGEFTA. Mayoufi et al. (2011) considered TGEFTA as the dissociation temperature, which resulted in lower equlibrium data than those obtained from Sales Silva et al. (2016) using Tend. Therefore, the discrepancies among the reported equilibrium data of TBPB may be attributed to differences in the measurement techniques used, as well as the dissociation points and temperature values considered.
Fig.5 Definition of characteristic peak temperature of DSC endothermic peaks, modified from (Sales Silva et al., 2016).

Full size|PPT slide

The two reported equilibrium data of TBPC are consistent with each other, as both studies utilized the visual observation method (Sakamoto et al., 2011; Ye and Zhang, 2014a). Compared with TBPB, TBPC has better thermal stability, with a melting point of 11.3 °C, while the melting point for TBPB is 9.25 °C.

3.2.3 Comparison

Despite differences in the measurement techniques employed, the phase equilibrium temperature of quaternary salt semi-clathrate hydrates initially increases with concentration and then decreases. The maximum phase equilibrium temperature, also known as the congruent melting temperature (CMT), corresponds to a concentration of approximately 35 wt%. If a salt solution with a higher concentration is used, the excess ions will act as hydrate inhibitors and reduce the thermal stability of the hydrate.
The congruent points of different quaternary salts SCH are summarized in Fig.6. The stability of these hydrates are determined by the compatibility of the hydrophobic cations in the framework structures built of water molecules together with an anion (Dyadin and Udachin, 1984). The hydrophobic interaction plays a critical role in the formation process of SCH, and it depends on the size of the molecule’s hydrophobic part. As shown in the figure, with the same anion, the thermal stability of QAS SCH is higher than that of QPS SCH. This can be attributed to the different cations used, as the bond length of C–P in quaternary phosphonium salt is 1.73 Å, which is larger than the bond length of C–N (1.53 Å) in quaternary ammonium salt. As a result, the C–P bond induces larger distortion to the water molecule, which destabilizes the hydrate structure (Muromachi et al., 2014a).
Fig.6 The congruent melting point of different quaternary salts (Nakayama, 1987; Sakamoto et al., 2008; 2011; Suginaka et al., 2012; Sato et al., 2013).

Full size|PPT slide

The hydrophilic interaction between water molecules and anion influences the formation conditions of the quaternary salts SCH. As illustrated in the figure, the CMP decreases with onion radius in both the QAS and QPS. The comparison data are summarized in Tab.7. The spatial complementarity between the guest molecules and the host cages, as well as the hydrophilic interaction between the anion and water molecules, contribute to the thermal stability of quaternary salt semi-clathrate hydrate (Aladko et al., 2003). In the case of QAS clathrates, the thermal stability decreases from fluorine to chloride to bromine with increase radius. Tetra-n-butyl-ammonium iodine cannot form a hydrate structure (Dyadin and Udachin, 1987), which is believed to be caused by the different distortions induced in the water molecule when halides form host cages with water molecules (Nakayama, 1983; Aladko et al., 2003). The hydrogen bond length of the host cages formed from fluorine ion with water molecule is 2.791 Å (Kobori et al., 2015b), which is very close to the hydrogen bond length of 2.8 Å in water. On the other hand, other halides with larger radii, such as chloride and bromide ions, induce larger distortions with hydrogen bond lengths of 3.1 Å (Lundgren and Olovsson, 1967) and 3.4 Å (Lundgren and Olovsson, 1968), resulting in less stable hydrates.
Tab.7 The relation between the congruent melting points and radius of the anion
Quaternary salt CMP (K) Concentration Radius of anion (Å) Ref.
TBAB 285.9 0.36–0.37 1.96 Sato et al. (2013)
TBAC 288.35 0.35 1.81 Nakayama (1987); Oshima et al. (2020)
TBAF 300.75 0.34 1.15 Sakamoto et al. (2008)
TBPB 282.4 0.35 1.96 Suginaka et al. (2012)
TBPC 283.45 0.36 1.81 Sakamoto et al. (2011)

3.3 Thermodynamical properties of semi-clathrate hydrate with gas mixture

In practical CO2 capture and sequestration applications, the operation conditions depend on the thermodynamical properties of carbon dioxide and quaternary salt double hydrate. It is expected that CO2 will form binary hydrate with quaternary salt (Kobori et al., 2015a), while the gas to be separated from CO2 is kept in the gas phase. The operation conditions need to be chosen to ensure the formation of CO2 and quaternary salts hydrate, while avoiding the formation of ternary hydrate formation of gas mixtures and quaternary salts. Therefore, the equilibrium condition of CO2 and quaternary salts hydrate, as well as that of the ternary hydrate, need to be determined.
The equilibrium conditions for the binary hydrates of QAS/QPS and CO2 are summarized in Fig.7 and Fig.8. Acting as hydrate promoters, quaternary salts can greatly alleviate the formation condition of the CO2 gas hydrate. Additionally, CO2 gas enters the empty cages of the QAS SCH, making the double hydrate more stable than QAS SCH (Sun et al., 2008). As shown in Fig.7, the CO2-TBAB double hydrate forms at a pressure as low as 0.4 MPa and temperature of 280.2 K with a TBAB concentration of 0.05. The equilibrium temperature first increases with the concentration of the quaternary salts (Arjmandi et al., 2007; Lin et al., 2008; Deschamps and Dalmazzone, 2009; Li et al., 2010a; Joshi et al., 2013; Chazallon et al., 2014; Ye and Zhang, 2014a; Sánchez-Mora et al., 2019). However, when the salt concentration exceeds a certain value, the stability effect reduces with concentration. In the case of TBAB, the critical concentration is near 40 wt% (Ye and Zhang, 2012) and it is 35 wt% for TBAC (Ye and Zhang, 2014a) and 33 wt% for TBAF (Adisasmito et al., 1991). This behavior is similar to that of QAS SCH forming at atmospheric pressure. TBAF can better stabilize the CO2 hydrate than TBAB and TBAC.
Fig.8 Phase equilibrium condition of CO2 and QPS binary hydrates: (a) TBAB + CO2; (b) TBAC + CO2; (c) TBAF + CO2 (Zhang et al., 2013; Xie et al., 2021; Suginaka et al., 2013; Ye and Zhang, 2014b; Momeni et al., 2020).

Full size|PPT slide

As shown in Fig.8, the stabilization of CO2 with QPS also increase first with concentration and then decrease above a certain value. The critical concentration of TBPB and TBPC is both near 35 wt%. The equilibrium condition of CO2 and TBPB is slightly milder than that of CO2 and TBPC with the same salt concentration. It has been demonstrated that TBAB and TBPB with stoichiometric concentration of TBAB∙26H2O and TBPB∙32H2O have a competent stabilization effect on CO2 hydrates (Mayoufi et al., 2010). As described in the previous section, the equilibrium temperature of TBAC is higher than that of TBPC, due to the smaller distortion to the water molecule from the C–N bond. However, after incorporation of CO2, this behavior partly changes. When the mass fraction w = 0.05, 0.10 and 0.20, the equilibrium temperature of CO2 + TBAC is higher than that of CO2 + TBPC double hydrates at relatively low pressure, while it is the opposite at higher pressure (Ye and Zhang, 2014a). To explain these findings, the morphology evolution of TBAC and TBPC with and without CO2 was compared. The morphology of TBAC with and without CO2 was different at dilute solution. TBAC hydrate initially appeared as a needle-like shape and then evolved into a columnar shape, while TBAC hydrate with CO2 initially showed two quite different morphologies of needle-like and hexagonal plate shape and changed into a columnar shape after a short period. The morphology of TBPC with and without CO2 was the same, showing similar morphology to TBAC hydrate. Thus, the structural variation of TBAC after incorporation of CO2 gas may contribute to the thermodynamical difference between TBAC and TBPC hydrate with CO2.
Therefore, from the perspective of operation conditions in CO2 gas capture and sequestration, TBAF with concentration of 34 wt% is preferred due to the greatest stability enhancement it provides. However, there are also other factors that affect the efficiency of a hydrate-based CO2 removal process, such as the amount of gas entrapped in a given mass of semi-clathrate hydrate and the gas selectivity of the gas mixture. These factors must also be considered when selecting the optimal operating conditions for CO2 gas capture and sequestration.
The phase equilibrium conditions of CO2/H2 mixture and quaternary salt hydrate are summarized in Fig.9 (Li et al., 2010b; Kim et al., 2011; Mohammadi et al., 2013b; Park et al., 2013; Babu et al., 2014c). CO2 hydrate (Adisasmito et al., 1991) has better thermal stability than H2 hydrate (Chapoy et al., 2010), thus the thermal stability of the gas mixture increases with the proportion of CO2. The addition of quaternary salts can shift the phase equilibrium of CO2/H2 mixture hydrate to higher temperature and lower pressure, with TBAF having the highest enhancement of the thermal stability. The thermal stability improvement also increases with the concentration of the quaternary salts. However, this trend has an inflection point that occurs at a TBAB concentration between 3 and 3.7 mol%, similar to the phase equilibrium condition of TBAB hydrate. At TBAB concentration higher than 3.7 mol%, the excess amount of TBAB molecules remains as free ions and inhabits hydrate formation. The inflection point was not discovered in the other quaternary salts, as the measurement data were limited. However, it can be estimated that the existence of this inflection point based on the phase equilibrium curve of the quaternary salts.
Fig.9 Phase equilibrium condition of CO2/H2 mixture and quaternary salt hydrate: (a) TBAB + CO2 + H2 (Li et al., 2010b; Kim et al., 2011; Mohammadi et al., 2013b; Park et al., 2013); (b) TBAF + CO2 + H2 (Park et al., 2013); (c) TBANO3 + CO2 + H2 (Babu et al., 2014c).

Full size|PPT slide

The thermodynamical properties of CO2/N2 mixture and quaternary salt hydrate have been collected in Fig.10 (Duc et al., 2007; Deschamps and Dalmazzone, 2009; Meysel et al., 2011; Belandria et al., 2012; Majumdar et al., 2012; Mohammadi et al., 2012; Hashimoto et al., 2017a; 2017b; Sánchez-Mora et al., 2019; Hashimoto et al., 2020; Kim et al., 2020). The equilibrium pressure of N2 hydrate is much higher than the CO2 hydrate (Sánchez-Mora et al., 2019), thus the CO2 in the gas mixture can increase the thermal stability of gas mixture hydrate. The more CO2 there is, the higher the thermal stability. The enhancement of quaternary salt to gas hydrate is similar to that of CO2/H2 gas mixtures. However, since higher concentrations of quaternary salts have not been measured in the literature, the inflection point indicating the optimum concentration have not been determined. It is estimated to have the same trend as in the case of CO2/H2 gas mixtures. The thermodynamic stability of tetra-iso-amyl ammonium bromide (TiAAB) semi-clathrates with CO2/N2 mixture was also investigated (Majumdar et al., 2012). Since TiAAB has even better thermal stability than TBAF, the ternary hydrate of TiAAB + CO2 + N2 has the highest thermal stability among all the other quaternary salts shown in the figure.
The phase equilibrium curve of CO2/CH4 mixture and quaternary salt hydrate were obtained to design the separation of CO2 from the biogas (Deschamps and Dalmazzone, 2009; Acosta et al., 2011; Fan et al., 2013; Mohammadi et al., 2013b; Li et al., 2017; Zang and Liang, 2017; Yan et al., 2019). As illustrated in Fig.11, the thermal stability of CO2 and CH4 gas mixtures slightly increases with the proportion of CO2. When the concentration of the TBAB solution is as low as 1 wt%, the promotion effect of TBAB on the gas hydrate is negligible. Significant improvements in the thermal stability of gas mixture hydrates occur when the concentration of quaternary salt increases to 4.3 wt%. The operating pressure can be greatly lowered, achieving 2MPa at ambient temperature using TBAB with a concentration of 40 wt%.
Fig.11 Phase equilibrium condition of CO2/CH4 mixture and quaternary salt hydrate with different gas compositions: (a) CO2 composition 60%, blue curve; CO2 composition 50%, black curve; (b) CO2 composition 33%, blue curve; CO2 composition 40%, black curve (Deschamps and Dalmazzone, 2009; Acosta et al., 2011; Fan et al., 2013; Mohammadi et al., 2013b; Li et al., 2017; Zang and Liang, 2017; Yan et al., 2019).

Full size|PPT slide

4 Application in carbon dioxide capture and separation

Carbon dioxide is considered the main contributor to global warming, and as human societies develop, the demand for energy is increasing. Conventional fossil fuels are currently the dominant energy resources, and CO2 emissions during the fossil fuel combustion are the primary causes of the CO2 concentrations in the atmosphere. CO2 capture can be applied either before or after fuel combustion, namely, pre-combustion (Babu et al., 2015) and post-combustion (Wang and Song, 2020). Pre-combustion involves capturing CO2 from fuel gas generated by Integrated Coal Gasification Cycle (IGCC) plants, which typically containing 40% CO2 and 60% H2. Post-combustion deals with CO2 capture from flue gas, consisting of approximately 15%–20% CO2 and 5% O2, with the remaining gas being N2.
Hydrate formation is a feasible way for capturing CO2 from nitrogen, hydrogen, or other waste gases from coal combustion and biogases, among others (Ma et al., 2016). Howerver, the stringent conditions of relatively high pressure and low temperature required to form CO2 hydrate hinder its application. The presence of quaternary ammonium salts can shift the equilibrium condition by formation a double hydrate with CO2 to milder conditions of higher temperature and lower pressure, in which the CO2 molecules enter the empty cages of the hydrate. In this case, CO2 capture and separation from other gases consume less energy and show great application potential. This section summarizes the performance of CO2 capture from fuel gas, flue gas and biogas using various quaternary salts.

4.1 Mechanism of CCS with SCH

Essentially, the principle of CCS with quaternary salts is to take advantage of the thermodynamical difference between the salt with CO2 and the other gas that needs to be separated. As shown in Fig.12, the binary hydrate of TBAB and CO2 exhibits higher thermal stability than that of TBAB and H2/N2/CH4 hydrates. Therefore, CO2 molecules can be enclathrated in the hydrate cage, leaving the gas molecules to be separated in the gas phase. The purified H2 from fuel gas can be further utilized as clean energy.
Fig.12 Phase equilibrium condition of TBAB and H2/N2/CH4/CO2 hydrate with mass concentration of 0.1 (Arjmandi et al., 2007; Mohammadi et al., 2011; Joshi et al., 2013; Jin et al., 2019).

Full size|PPT slide

The phase equilibrium curve of CO2 hydrate and CH4 hydrate are pretty close to each other, with CO2 having slightly better thermal stability than CH4 (Aladko et al., 2003; Lee et al., 2012). Therefore, to separate CO2 from CO2/CH4 gas mixtures, the optimal operation conditions exist in the gap between the P–T curve of CO2 and CH4 gas hydrates that corresponds to CO2 hydrate formation while CH4 remains in the gas phase (Babu et al., 2014b). Delicate control of the operation conditions is required in the practical applications.

4.2 Performance parameters of CCS

First, the parameters that characterize carbon dioxide capture and separation performance are described. The number of moles of gas consumed at any time during hydrate formation can be calculated by the following equation:
Δn= n0 nt.
In some cases, the normalized gas uptake is defined as:
Δnnor=Δn nw,
where nw is the number of moles of the water utilized in the experiment. Δ nnor is related to the conversion of water into gas hydrates or semi-clathrate hydrates.
The split fraction of carbon dioxide, i.e., CO2 recovery, in gaseous and hydrate phase is calculated as follows (Linga et al., 2007a):
S.Fr.=nCO2 HnC O2 Feed,
where nCO2 H is the number of moles of CO2 in hydrate phase at the end of the experiment and nCO2 Feed is the number of moles of CO2 in the initial feed gas.
Another important parameter is the separation factor that is defined as (Linga et al., 2007a):
S.F.= nCO2 H/ n CO2 gas nXH/ n Xgas,
where X denote the other gas from the feed gas mixture that will be separated from CO2, nXH is the number of moles of X in hydrate phase, nCO2 gas and nXgas are the number of moles of CO2 and X in the gas phase at the end of experiment.

4.3 CO2 capture and separation from fuel gas

The gas mixture from IGCC power plants typically contain 40% CO2 and 60% H2. CO2 can be captured for further disposal, while the resulting H2 can be utilized as clean energy. It should be noted that quaternary salts are also promising candidates for hydrogen storage (Chapoy et al., 2007; Hashimoto et al., 2008), although this is beyond the scope of this work. Quaternary salts are potential promoters for CO2 capture from fuel gas through hydrate formation, and their performance are summarized in Tab.8. After one stage separation with the quaternary salt solution, the CO2 concentration in the hydrate slurry can be enriched to more than 90%. Meanwhile, the operation pressure can be moderated to less than 5 MPa. Since the fuel gas pressure from the ICGC plants is 2.5–5 MPa (Park et al., 2013), there is no need to compress the gas mixture, which can significantly reduce the cost of the CO2 capture process.
Tab.8 Performance of quaternary salts for CO2 capture from fuel gas
Gas mixture Solution Concentration (mol) P (MPa) T (K) Reactor volume (mL) Aqueous volume (mL) Induction time (min) n (mol) CO2 in hydrate S.F. S.Fr. Ref.
CO2(39.2)/H2(60.8) H2O 7.5 273.7 323 150 5.3a 0.095 86.5 98.7 0.42 Linga et al. (2007b)
CO2(40)/H2(60) H2O 8 276 250 80 0.02c 95 Park et al. (2013)
CO2(40.2)/H2(59.8) TBAB 0.29% 3.9 273.5 527 260 1 0.05 86 15.7 0.41 Gholinezhad et al. (2011)
CO2(40)/H2(60) TBAB 0.29% 3.08 274.15 1350 180 10.2 0.133 96.4 0.145 Xu et al. (2013)
CO2(40)/H2(60)b TBAB 0.29% 6 279.2 142 53 20.1 0.01c 95.2 49.5 Babu et al. (2014a)
CO2(40.4)/H2(59.6) TBAB 0.29% 5 277.4 863 300 19 82 0.64 Horii & Ohmura (2018)
CO2(39.2)/H2(60.8) TBAB 0.29% 3 278.15 336 180 0.105 96.85 136.08 0.67 Li et al. (2011)
CO2(40.1)/H2(59.9) TBAB 1% 3 280.15 25.53 0.03 89.07 25.99 0.241 Kim et al. (2011)
CO2(40)/H2(60) TBAB 0.6% 8 283.8 250 80 0.0025c 95 Park et al. (2013)
CO2(80.5)/H2(19.5) TBAB 0.29% 3.31 276.7 527 260 96 15 0.35 Gholinezhad et al. (2011)
CO2(18)/H2(82) TBAB 0.29% 3.25 274.15 1350 180 21 0.101 92.8 0.163 Xu et al. (2013)
CO2(10)/H2(90) TBAB 0.29% 4.52 274.15 1350 180 36.4 0.067 96.4 0.192 Xu et al. (2013)
CO2(40)/H2(60) TBAF 3.3% 8 288 250 80 0.0035c 95 Park et al. (2013)
CO2(40.6)/H2(59.4) TBAF 3.38% 4 298 142 53 167.56 0.008 93.1 Zheng et al. (2017)
CO2(40)/H2(60)b TBANO3 1% 6 274.2 142 53 0.55 0.0132c 93.75 32.21 Babu et al. (2014c)
CO2(35)/H2(65) TBAC 2.7% 3.028 288.2 100 30 12 0.011 96 61 0.29 Muromachi (2021)
CO2(34.5)/H2(65.5) TBPB 2.2% 3.032 284.1 100 30 28 0.019 95 26 0.45 Muromachi (2021)
CO2(34.7)/H2(65.3) TBPC 2.5% 2.982 284 100 30 17 0.014 97 23 0.34 Muromachi (2021)

Note: a) Memory water of 3 h after hydrate decomposition is used; b) Batch experiments were conducted; c) Normalized gas uptake.

The CO2 separation from fuel gas involves two phases: gas dissolution and hydrate enclathration. Among all the parameters that affect the CO2 separation performance, the induction time and the volume ratio of the gas mixture and the aqueous solution mainly contribute to the gas dissolution stage. The hydrate enclathration process is largely affected by the pressure and temperature conditions, the selected quaternary salt and its concentration. The S.F. of different quaternary salts in fuel gas separation is summarized in Fig.13, and the performance comparison among different quaternary salts for CO2 separation from fuel gas is summarized in Tab.8.
Fig.13 S.F. of different quaternary salts in fuel gas separation (Gholinezhad et al., 2011; Kim et al., 2011; Li et al., 2011; Babu et al., 2014a; 2014c; Muromachi, 2021).

Full size|PPT slide

TBAB is the mostly studied QAS in this application area. Li et al. (2011) carried out a comparative study of different TBAB concentrations ranging from 0.14 to 1 mol% with the same driving force and found that the critical concentration with the optimum CO2 separation performance was 0.29 mol%. The gas uptake first increased with concentration until 0.29 mol% and then decreased, as did the CO2 concentration in the hydrate slurry. A high separation factor of 136.08 was obtained, which was the highest among all the quaternary salts in the literature. However, the results from Kim et al. (2011) showed that the critical concentration was 1 mol%, which was different from the aforementioned experiments. The S.F. and S.Fr. in 1 mol% TBAB were 25.99 and 0.241, respectively. The experimental temperature in the latter study was approximately 1 K higher than that in study of Li et al. (2011). In another experiment carried out by Babu et al. (2014a), TBAB solution with concentration of 0.5, 1, 1.5, 2 and 3 mol% were compared. The best performance was found in the 0.3 mol% TBAB solution, which is similar to study of Li et al. (2011), while the maximum separation factor in their experiments was 49.5.
Generally speaking, in case of a dilute TBAB solution, such as 0.29 or 0.3 mol%, the introduction time is much longer than in other cases, allowing more gas to dissolve into the solution. The solubility of CO2 is much larger than that of H2, resulting in the formation of more CO2 and TBAB binary hydrate. TBAB solutions with high concentrations can degrade the CO2 separation performance. First, the equilibrium temperature of TBAB and H2 binary hydrate increases with concentration (Li et al., 2010b), and more H2 tends to enter into SCH structure under certain temperature conditions. Secondly, TBAB SCH forms under higher concentrations before the gas dissolves into the liquid, and the agglomeration of TBAB hydrate hinders gas dissolution. This has been verified by visual observation of CO2 separation experiments (Zheng et al., 2017).
Polymorphism is one of the key factors that influence gas capture properties. As is known, TBAB forms five polyhydrates under different conditions (Gaponenko et al., 1984; Lipkowski et al., 2002). It has been found that two types of TBAB hydrates, namely, tetragonal type A and orthorhombic type B hydrate, are preferentially formed (Shimada et al., 2003). In the presence of CO2, type B hydrate is preferred over the type A structure (Muromachi et al., 2014b). The different gas capacity of these two types of TBAB semi-clathrate hydrates has been confirmed and quantified (Zhou and Liang, 2019). CO2 adsorption experiments were conducted with type A TBAB hydrate particles that were formed in advance and carefully weighed to quantify the volume ratio of CO2 capture. Based on in situ Raman spectroscopy and powder X-ray diffraction, nCO2·TBAB·26H2O transformed into nCO2·TBAB·38H2O and TBAB·21 /3H2O under pressure of 2 MPa. The volume ratio of CO2 in the standard state absorbed by TBAB·26H2O hydrate particles changed from 10 v/v at 1 MPa to 67 v/v at 2 MPa. The trapped CO2 molecules destroyed the TBAB·26H2O hydrate structure, inducing the formation of a new structure that was more compatible to CO2 molecules. Muromachi et al. observed irregular CO2 capture with a TBAB aqueous solution under different subcooling temperatures, the CO2 capture amount under subcooling of 3.3–4.5 K was three times higher than that under a subcooling temperature of 2.2–3.3 K (Muromachi, 2021). Since the gas capacity of type A TBAB hydrate is less than that of type B TBAB hydrate, the phases observed under different subcooling conditions are probably type A and B TBAB hydrate, respectively.
By using a one-stage separation with TBAB solution, the optimum separation factor and the split fraction of TBAB can achieve 136.08 and 0.67. This means that 67% of CO2 can be recovered, and hydrate slurries with more than 97 mol% CO2 can be obtained, indicating the applicability of TBAB in this area. However, under normal conditions, a certain amount of CO2 still exists in the residual gas after single-stage separation, and multi-stage separation, where the gas mixture from the previous stage is used as the feeding gas should be conducted. With two-stage separation, a composition of 80.5 mol% CO2 in the first stage can be enriched to more than 95 mol% using TBAB with a concentration of 0.29 mol% (Gholinezhad et al., 2011). To simulate the multi-stage process, an initial gas mixture of 90% H2/10% CO2 was prepared for separation, and only 3.6 mol% of CO2 was left in the residual gas after the separation (Xu et al., 2013).
Since TBAF can better stabilize the CO2 hydrate than TBAB (Li et al., 2010a), the CO2 separation performance of TBAF were investigated (Park et al., 2013; Zheng et al., 2017). Based on the kinetics of H2-TBAF and CO2-TBAF binary hydrate, Trueba et al. indicated the feasibility of using TBAF for the purification of flue gases (Trueba et al., 2013). However, it was found that the TBAF may not be applicable for CO2 separation due to its poor performance. The gas uptake was much less than that of TBAB solution (Park et al., 2013; Zheng et al., 2017). This is because about 79.5% of the 512 cages in TBAF hydrate framework are filled with water (Komarov et al., 2007; Manakov et al., 2011), which limit the amount of gas that can be absorbed to form binary hydrate. The influence of pressure and temperature conditions, i.e., the driving force of TBAF and CO2 binary hydrate on CO2 separation performance, has been studied (Zheng et al., 2017). The experiments were conducted with TBAF at the stoichiometric concentration of 3.38 mol% under pressure of 2, 4, 6 MPa and temperature of 298, 292 and 286 K. The induction time at 298 K was the highest among the three temperatures because more gas dissolved into the liquid solution. The dissolved gas also increased with pressure due to higher gas solubility and fugacity. However, the mole concentration of CO2 in hydrate slurry in 6 MPa is lower than 4 MPa because at higher pressure, more H2 molecules occupy the hydrate cages.
Tetra-n-butyl ammonium nitrate (TBANO3) has also been evaluated as a promoter for precombustion capture of CO2 (Babu et al., 2014b; Babu et al., 2014c). With 1 mol% TBANO3, the molar concentration of CO2 in the hydrate slurry achieved 93%, with a normalized gas uptake of 0.0132 and a separation factor of 32.21, which was comparable to that of TBAB measured with the same apparatus (Babu et al., 2014a). A subsequent study was conducted to investigate the influence of operation temperature and pressure on the gas separation performance. The gas uptake capacity of TBANO3 was found to be higher than that of TBAB or TBAF at comparable driving forces (Babu et al., 2014b). However, it should be noted that the experiments were conducted with a batch reactor, which was a closed system. The gas uptake was largely affected by the initial feed gas and the liquid (Wang et al., 2020a) and thus was not a proper metric to evaluate the separation performance in this case.
To achieve better CO2 gas selectivity, it would be ideal if the clathrate hydrate of quaternary salts could reject H2 molecules. This was found to be the case for TBPB and TBPC semi-clathrate hydrate (Muromachi, 2020). The conclusion was based on the equilibrium condition of quaternary salts solutions and gas mixture. The equilibrium temperature of TBAB/TBAC + CO2 + H2 hydrate was higher than that of the TBAB/TBAC + CO2 hydrate with identical CO2 fugacity, demonstrating the H2 inclusion in the hydrate framework that increases its thermal stability. Evidence of H2 gas entering into the 512 cages of TBAB was provided by 1H nuclear magnetic resonance (NMR) measurements (Park et al., 2013). The equilibrium temperature of TBPB and TBPC SCH with gas mixtures was lower than that with only CO2 gas, showing their rejection of H2 gas. It was found that TBPB was favorable in terms of gas storage capacity (Deschamps and Dalmazzone, 2010; Mayoufi et al., 2010). Although TBPC can better stabilize CO2 hydrate than TBPB (Sakamoto et al., 2011; Iino et al., 2014), TBPB was considered a superior promoter for CO2 capture from fuel gas in terms of CO2 separation performance and simple phase behavior (Muromachi, 2021).
All the quaternary salts solutions employed as promoters in fuel gas separation in the literature have resulted in low separation factors and high gas consumption (Park et al., 2013). Therefore, efforts to find the proper quaternary salts as promoters are still necessary. Several factors should be considered for their application. The structure of QAS/QPS SCH is the most important factor that limits gas capacity. Among all the structures described in the previous section, the orthorhombic structure with the most D cages per quaternary salt, shows the highest gas capacity. Better thermal stability is preferred to lower separation cost. The polymorphism of the QAS/QPS should be avoided because it requires delicate control of the operation conditions, which is challenging in practice.

4.4 CO2 capture and separation from flue gas

The task of capturing flue gas from power plants using hydrates is to separate CO2 from a CO2/N2 mixture at ambient pressure, in which the CO2 molar content is approximately 15%–20%. As shown in Fig.14, with the same initial CO2 composition, different quaternary salts have similar separation performance in terms of the CO2 composition in the hydrate slurry. The obtained CO2 composition in the hydrate increases with its initial composition, while it is still lower than that of CO2–H2 gas mixture after one stage of separation. In fuel gas separation by quaternary salt solutions, the CO2 composition in the hydrate slurry after one stage of separation is typically larger than 90%. The difference is caused by several factors. First of all, the CO2 concentration in the flue gas is between 3%–15% and it is 40% in the fuel gas. Higher initial CO2 concentration induces a higher CO2 concentration in the SCH. Meanwhile, as shown in Tab.9, the diameter of N2 is similar to that of a CO2 molecule, and it can be incorporated in the semi-clathrate cages when the pressure is above a certain value (Muromachi et al., 2016a). However, H2 can rarely incorporate into the hydrate cages (Muromachi, 2021).
Tab.9 Diameters of different gas molecules
Gas molecule H2 N2 CH4 H2S CO2
Diameter (nm) 0.27 0.41 0.44 0.46 0.51
The details of the CO2 capture performance from flue gas of QAS/QPS in Fig.14 are shown sequentially in Tab.10. Except for the extraordinarily high separation factor of 118.7 obtained from separation of a CO2-rich gas mixture of 66.5% with TBAB solution, the S.F. of QAS/QPS range from 4 to 30. The extra-high S.F. was achieved by TBAB solution with a mass concentration of 0.112, and the gas mixture to be separated is rich in CO2 with 66.5%. The CO2 in the hydrate slurry after separation is as high as 98.9%. With the same initial CO2 composition of approximately 12% and the same mass concentration of 0.2, TBAC has the highest S.F. of 10.2 while TBPC has the lowest S.F. value of 4.
Tab.10 Summary of the literature data for flue gas separation measurements
Salts Initial composition (%) Concentration (wt%) P (MPa) T (K) ΔT (K) nCO2 H S.F. S.Fr. Ref.
TBAB 13.3 0.2 1 281.2 3.1 25.8 7.2 Hashimoto et al. (2017a)
15.24 0.32 1 282.2 3.5 76.1 21.9 Hashimoto et al. (2017b)
16.6 0.05 4.03 277.65 36.53 9.82 0.53 Fan et al. (2009)
20 0.05 2 277 55.5 9.8 0.54 Komatsu et al. (2019)
31.7 0.218 5.2 288.6 90.9 29.34 Herri et al. (2014)
37.4 0.05 2 285.2 78.9 17.6 0.68 Duc et al. (2007)
66.5 0.112 2.5 283.4 98.9 118.7 Herri et al. (2014)
TBAC 12.5 0.2 5 286.2 2.5 51.8 10.2 Hashimoto et al. (2017a)
20 0.345 3 284.7 5 60 8 Kim and Seo (2015)
TBAF 12 0.2 1.01 295.2 37.4 5.2 0.4 Hashimoto et al. (2020)
16.6 0.04 2.46 277.65 56.55 36.98 0.56 Fan et al. (2009)
20 0.338 3 284.7 5 56 Kim and Seo (2015)
TBPB 12.1 0.2 5.02 284.2 3.6 31.2 4.8 Hashimoto et al. (2017a)
17 0.05 3.5 277.5 61.71 13.25 Ye and Zhang (2014b)
61.71 0.05 3 277.5 91.28 16.23 Ye and Zhang (2014b)
TBPC 12 0.2 5 284.2 3.7 28.6 4 Hashimoto et al. (2017a)
The structures of the QAS/QPS semi-clathrate hydrates play an important role in their CO2 capture and separation performance. According to the comparative study of the CO2 separation from 15% CO2/N2 gas mixture by TBAB, TBAC, TBPB and TBPC semi-clathrate hydrate (Hashimoto et al., 2017b), the CO2 selectivity of TBAC with a tetragonal structure was the highest, while the total amount of captured CO2 by TBAC hydrate was the least, corresponding to 60%–90% of the other hydrates. With polycrystalline structures of orthorhombic and tetragonal, TBAB hydrate had moderate performance and was superior in gas selectivity than TBPB and TBPC hydrate with an orthorhombic structure. The type A TBAB hydrate with a tetragonal structure had better CO2 selectivity than type B TBAB hydrate with an orthorhombic structure (Rodriguez et al., 2020).
The different gas capacity among different hydrate structures can be explained with the hydration numbers in different structures, which are summarized in Tab.1 (Muromachi and Takeya, 2017). Since guest gas molecules are trapped in the small cages of the SCH, the smallest ratio of the number of water molecules with the 512 cages in the orthorhombic hydrate has the highest gas capacity. The cubic structure with the highest ratio is assumed to have the poorest gas capture capability. This prediction was consistent with the experimental study of flue gas separation with TBAF hydrate with a cubic structure (Kim and Seo, 2015; Hashimoto et al., 2020).
Based on CO2 gas separation experiments with TBAB, TBAC and TBAF under the same driving force of 5 K, it was concluded that the CO2 selectivity was independent of the types of QAS used (Kim and Seo, 2015). The initial gas mixture with 20% CO2 was enriched to approximately 60% CO2 in all QAS hydrate slurry. The difference between Kim and Seo (2015) and Hashimoto et al. (2017b) may result from different experimental methods. In the former case, a semi-batch operation with continuous gas feeding was utilized, while an isolated reactor was adopted in the latter. Kim et al. adopted a batch mode where the reactor was isolated, thus the driving force in the isolated reactor kept decreasing during the gas separation process, which may degrade the separation performance (Zhong and Englezos, 2012; Hashimoto et al., 2017a).
The experimental conditions and the concentration of quaternary salts may influence the performance of CO2 capture and separation. The pressure of the gas mixture was found to significantly alter the gas separation with TBAB and TBAF solution (Fan et al., 2009). The S.F. first increased and then decreased with pressure because more N2 accompanied with CO2 was incorporated into the hydrate cages. The CO2 recovery remained almost unchanged with pressure. The gas separation performance for TBAB hydrate was found to be better at a pressure of 1 MPa than at high pressure of 2 and 3 MPa (Hashimoto et al., 2017a). They also found that a low temperature was beneficial to engage CO2 in TBAF semi-clathrate hydrate. Dense TBAF aqueous solution reduced both the CO2 selectivity and gas capacity, and this tendency was also found for TBAB, TBAC and TBPB hydrates (Hashimoto et al., 2020). Moreover, at higher mass concentration, polymorphism of the SCH may appear and show inconsistent gas separation performance (Hashimoto et al., 2017a; 2020). The volume of the reactor and quaternary salt solution also affects gas separation performance, which was quantified by the gas liquid volume ratio Rv (Li et al., 2009). It was found that a lower Rv corresponded to a higher ratio of gas to hydrate, which is beneficial to CO2 gas separation.
To improve the practical application of QAS/QPS semi-clathrate hydrates for CO2 capture, research should be focused on finding appropriate quaternary salts and determining the best physical conditions to capture as much CO2 gas as possible from gas mixtures. By optimizing the experimental conditions, such as pressure, temperature, mass concentration, and gas-liquid volume ratio, the CO2 selectivity and gas capacity of QAS/QPS semi-clathrate hydrates can be enhanced. Moreover, the development of new quaternary salts with better gas separation performance and thermal stability is also necessary for the practical application of QAS/QPS semi-clathrate hydrates in CO2 capture.

4.5 CO2 capture and separation from biogas

Biogas, which is usually used as a fuel, contains methane, carbon dioxide and small proportion of H2S gas. To avoid the generation of corrosive sulfurous acid, H2S needs to be removed (Kamata et al., 2005). Additionally, CO2 should be separated to purify the natural gas. Results have shown that H2S can be captured along with CO2 in the same process, and the CO2 separation from CO2/CH4 gas mixture should be considered accordingly (Castellani et al., 2014). The proportion of CO2 in biogas generally ranges from 30%–40%.
The separation of CO2 from CH4 is challenging from a thermodynamical perspective because their phase equilibrium conditions are pretty close to each other, as shown in Fig.12. On the other hand, CO2 and CH4 molecules have similar size and both tend to occupy the small cages in quaternary salts SCH (Lee et al., 2012; Tang et al., 2013). The incorporation of CH4 or CO2 molecules can both improve the equilibrium temperature of the quaternary salts. For instance, the equilibrium temperature of TBAB at a mass concentration of 20% is 283.2 K (Sato et al., 2013), it is 287.55 K at a CH4 gas pressure of 1.421 MPa and 286.9 K at a CO2 gas pressure of 1.52 MPa, respectively. For single CH4 and CO2 gases, the equilibrium temperature is 259.1 K at 1.648 MPa and 274.3 K at 1.42 MPa (Adisasmito et al., 1991). In other words, the temperature difference between TBAB/CH4 and TBAB/CO2 is lowered compared to that between CH4 and CO2 hydrate. The different gas occupancy of CH4 and CO2 in TBAB hydrates may contribute to this phenomenon. As is known, CH4 prefer to occupy the regular distorted DB cages (Muromachi et al., 2016b), while CO2 occupies the highly distorted DA cages (Muromachi et al., 2014b). The chemical composition of the crystal unit cell was determined to be TBAB·38H2O·2.14 CH4 at 284.5 K and 1.92 MPa. It is TBAB·38H2O·1.79 CO2 at a similar temperature and pressure condition of 287.3 K and 1.6 MPa. This means that more CH4 enters into the hydrate cage, thus showing a better stabilization effect on TBAB hydrate.
Despite the difficulties, the separation of CH4 from CO2 with quaternary salts is feasible and attempts have been made to find the best salt and optimal conditions (Lee et al., 2011; Fan et al., 2013; Li et al., 2015; Fan et al., 2016; Wang et al., 2016; Xia et al., 2016; Li et al., 2017; Yue et al., 2018; Li et al., 2019; Yan et al., 2019; Wang et al., 2020b). The investigations have been summarized in Tab.11. All quaternary salts can significantly improve the stability of CO2 and CH4 hydrate, as shown in Fig.11 (Deschamps and Dalmazzone, 2009; Acosta et al., 2011; Fan et al., 2013; Mohammadi et al., 2013a; Li et al., 2017; Zang and Liang, 2017; Yan et al., 2019). Lee et al. measured the stability of the ternary CO2 + TBAB + H2O and CH4 + TBAB + H2O system with different salt concentrations of 0.6, 3.7 and 7.7 mol%, and the highest stabilization effect was observed at a concentration of 3.7 mol% (Lee et al., 2011). This concentration corresponded to the stoichiometric concentration of type A TBAB hydrate (Oyama et al., 2005). Yan et al. (2019) investigated the thermodynamical properties of TBAB + CO2 + CH4 mixed hydrates at TBAB concentrations of 0.29, 0.62, 1.38 and 2.57 mol%. The thermal stability of the mixed hydrate increased with the TBAB concentration. Therefore, there is an optimum TBAB concentrations to stabilize CO2/CH4 hydrate, above which the extra ions from the salt will inhibit the hydrate formation.
Tab.11 CO2 gas separation performance with quaternary salts from biogas
Gases Promoters PT condition Separation performance Ref.
33% CO2/CH4 TBAB(0.05 wt%) + [BMIm]BF4 3 MPa278 K Methane in the residual gas increased and the time required to reach balance was significantly shortened by addition of [BMIm]BF4. The highest CH4 concentration in the residual gas phase was 84.0%, maximum CO2 separation factor of 10.3. Li et al. (2015)
33% CO2/CH4 TBAB(0.1, 0.293, 0.9) mol% 1.14 MPa281.3 K TBAB concentration of 0.293 mol% has the optimum separation performance with maximum gas uptake, CH4 fraction in the residual gas phase, CO2 separation factor, and CH4 separation and recovery factor. Fan et al. (2016)
40% CO2/CH4 TBPB(5–33.2)wt% 2.8 MPa278.1–284.2 K Maximum CO2 separation factor at 33.2 wt% TBPB solution was (31.7 ± 3.3), 8.9 times of 1% THF. CO2 recovery was 0.41 and the CO2 concentration in hydrate achieved 70.8%. Li et al. (2017)
45% CO2/CH4 DMSO + THF/TBAB 2.5 MPa283.25-285.96 K DMSO will increase the solubility of CO2 in solution and CO2 separation factor increased from 40 with only TBAB to 59.22 with TBAB and DMSO. Xia et al. (2016)
33% CO2/CH4 TBAB + [C8min] BF4 4 MPa276.15–279.15 K The combination of TBAB and [C8min] BF4 could increase the mole friction of CH4 in residual gas, while the CH4 in hydrate also increased, and the recovery of CH4 from biogas decreases. Yue et al. (2018)
40% CO2/CH4 TBAB 2.8 MPa, 278.8 K CO2 separation factor is 85.5, CO2 recovery is 0.344, and CO2 concentration in hydrate is 96.5% Li et al. (2019)
TBPB 2.8 MPa, 278 K CO2 separation factor is 33.8, CO2 recovery is 0.293, and CO2 concentration in hydrate is 92.1%
40% CO2/CH4 TBAB(0.29, 0.62, 1.38, 2.57) mol % 2.8 MPa280.1–284.8 K 0.62 mol% TBAB gave highest gas consumption and highest CO2 recovery of 0.502. The highest gas separation factor of 36.5 and the CO2 concentration in hydrate was 73% achieved at concentration of 2.57 mol%. Wang et al. (2020b)
The optimum concentration of quaternary salt that corresponds to the highest thermal stability of mixed quaternary salt and CH4/CO2 mixed hydrate may not be consistent with that providing the optimal separation performance. The optimal separation performance with maximum gas uptake, CH4 fraction of 93.52 mol% in the residual gas phase, CO2 separation factor of 42.17, and CH4 separation and recovery factor of 1.43 was found at a TBAB concentration of 0.293 mol%, compared to concentrations of 0.1 and 0.9 mol% (Fan et al., 2016). The TBAB concentration of 0.293 mol% was also found to provide the best separation performance in fuel gas separation (Li et al., 2010c). This may be because higher concentrations make no further contribution to CO2 capture since TBAB cations and CO2 both occupy the large cages in semi-clathrate hydrate (Fan et al., 2016). However, different results were obtained in a recent investigation with different TBAB concentrations of 0.29, 0.62, 1.38 and 2.57 mol %. The highest gas consumption and highest CO2 recovery of 0.502 happened at a 0.62 mol% TBAB solution, while the highest gas separation factor of 36.5 and CO2 concentration in hydrate of 73% were achieved at 2.57 mol% (Wang et al., 2020b). This may be because at a concentration of 0.62 mol%, more CH4 was incorporated in the hydrate cages. The occupation of different gas molecules in semi-clathrate cages at the molecular level is essential to interpret the inconsistent experimental observations.
In addition to TBAB, the potential of other quaternary salts has been evaluated. There is a range between the CO2 and CH4 phase equilibrium curve in which CO2 is enclathrated in the hydrate cages while CH4 remains in the gas phase. Quaternary salts can enlarge this range and Fan et al. (2013) found that TBAF induced a larger extent of the range than TBAC and TBAB. Compared to the tetragonal TBAB hydrates, the orthorhombic TBPB semi-clathrate hydrates are believed to have a superior CO2 storage capacity (Muromachi et al., 2014a). Experimental results showed that the thermal stability of TBPB + 40%CO2 + 60%CH4 mixed hydrate increased as the TBPB concentration increased from 5 to 33.2 wt% (Li et al., 2017). The highest CO2 separation factor of 31.7±3.3, CO2 recovery of 0.422 and CO2 fraction of 73.8% in the hydrate phase were achieved at a TBPB concentration of 33.2 wt% and a high temperature of 284.2 K, demonstrating its applicability in CO2 gas separation. However, the same group made a further investigation on the formation kinetics of TBAB/TBPB solution and CO2/CH4 gas mixtures and found that the TBPB semi-clathrate hydrate grew laterally at the gas/liquid interface and quickly covered the interface, hindering mass transfer and gas consumption (Li et al., 2019). Meanwhile, the TBAB semi-clathrate hydrate initially formed around the gas/liquid interface and then sank to the reactor bottom. As a consequence, TBAB had a better CO2 separation performance than TBPB, with a CO2 separation factor as high as 85.5, CO2 recovery of 0.352, and CO2 concentration in hydrate is 96.5%.
The combination of quaternary salts and other additives has been proposed for CO2 separation from CH4. 1-butyl-3-methylimidazolium tetrafluoroborate ([BMIm]BF4), which may increase the CO2 solubility (Iarikov et al., 2011) was adopted together with 5 wt% TBAB solution for 33 mol% CO2/67.0 mol% CH4 gas separation (Li et al., 2015). Results demonstrated that the additional additive increased the methane fraction in the residual gas and significantly shortened the time required to reach balance. However, the highest CO2 gas separation factor was only 10.3. Another additive, dimethyl sulfoxide (DMSO), which can also enhance the solubility of CO2 (Xia et al., 2012), was proposed to separate simulated biogas of 45.0 mol% CO2/CH4 binary mixture with TBAB (Xia et al., 2016). The promotion effect of DMSO was verified with improved CO2 fraction in the hydrate phase and decreased CO2 fraction in the residual gas, thus increasing the CO2 separation factor from 40 with only TBAB to 59.22 with TBAB and DMSO. Another homologous compound of omidazolium-based ionic liquid, 1-octyl-3-methylimidazolium tetrafluorborate ([C8min] BF4), was applied for the first time in simulative biogas (37.0 mol%CO2/CH4) separation accompanied with TBAB (Yue et al., 2018). Although the addition of [C8min] BF4 increased the mole fraction of CH4 in the residual gas, the CH4 in the hydrate also increased and the recovery of CH4 from biogas decreased.

5 Challenges and prospects

Currently, there is a lack of structural information of SCH and hydrate of CO2 and quaternary salts under different conditions (Rodionova et al., 2022). The polymorphism of SCH further complicates this situation, as it is difficult to maintain the metastable state to grow high-quality single crystal. The environment during hydrate growth can easily affect crystal growth, and incorporating gas molecules into SCH may induce structure transformations that are operation condition-dependent. For example, Zhou et al. found that different types of TBAB hydrate were formed under various supercooling conditions, with markedly different gas storage capacity (Zhou and Liang, 2019). However, optimizing the CCS process by manipulating this process is challenging. Therefore, it is crucial to have a clear understanding of the type of hydrate and the gas content that will form during a given process when implementing gas separation processes using SCH.
The kinetics of semi-clathrate hydrate formation play a critical role in the CCS process. One problem related to quaternary salts hydrate is the supercooling effect, whereby the salts crystallize at a temperature much lower than their equilibrium temperature and the corresponding temperature difference is called the degree of supercooling. For TBAB hydrate, the maximum allowable degree of supercooling can be as high as 17.7 K (Sugahara and Machida, 2017). Memory effect, a phenomenon in which recrystallization occurs at a lower degree of supercooling or within a shorter induction time, has been proposed to suppress supercooling (Oshima et al., 2010; Machida et al., 2018; 2020). While the lifetime of the memory effect sharply decreases when the temperature is 2 K higher than the equilibrium temperature of TBAB hydrate. Nucleators, or nucleating agents, have also been proposed to reduce the degree of supercooling. Halloysite clay nanotubes were tested as nucleating agents of TBAB hydrate, and adding 1 wt% of halloysite increased the onset temperature of hydrate crystallization by 2 °C (Stoporev et al., 2020).
Another common challenge in hydrate related applications is the slow kinetics of formation. Surfactants are commonly used as promoters for hydrate formation, and they have been extensively studied in this area (Mohammadi et al., 2018; Zheng et al., 2018; Xie et al., 2021). In addition to surfactants, numerous studies have been conducted to enhance hydrate formation by physically increasing the surface area available through reactor medium such as sand packs, silica gels, foams, nanoparticles, hydrogels (Stoporev et al., 2020) and so on (Linga and Clarke, 2017). The kinetics of CO2 hydrate formation are also affected by the driving force, stirring speed, agitator geometry, initial volume of water and vessel size (Liu et al., 2022).
Since the cation of quaternary salts occupy most of the cage space, the low gas capacity of SCH is a major obstacle in CCS applications. Selecting an appropriate quaternary salt to reduce hydrate formation pressure and increase the amount of captured gas is key to CCS applications. Besides quaternary ammonium and phosphonium salts with halide anions, the potential of halogen-free salts has attracted much attention (Shimada et al., 2019). Quaternary salts with carboxylic acid anions in the guest compounds, which are more environment friendly than halide anions, have been studied (Muromachi et al., 2015; Yamauchi et al., 2017a; 2017b; Shimada et al., 2018; Koyama et al., 2020; Miyamoto et al., 2020). Previous research on semi-clathrate hydrate based on carboxylic acid anions mainly focused on their potential applications as thermal energy storage materials, and their phase equilibrium conditions and dissociation heat have been determined. The melting points of this group of semi-clathrate hydrate are between 7 and 22 °C, which are compatible with the SCH of quaternary salts with halide anions. Moreover, the hydroxycarboxylates that are formed by adding a hydroxy group to carboxylates can also form semi-clathrate hydrates. The properties of hydroxycarboxylates based SCH can be tailored by crystal engineering technique to modify the onions with hydroxy groups (Muromachi and Takeya, 2018; 2019). However, their performance on CO2 gas capture and sequestration have not yet been investigated.

6 Conclusions

This paper reviews the structure and thermodynamical properties of SCH and SCH with inclusion of the gas molecules for separation. Quaternary salts have the ability to incorporate CO2 gas molecules at mild pressure and temperature conditions, thus reducing the costs associated with gas mixture compression. Among all the quaternary salts considered in this review, TBAF has been found to better stabilize CO2 gas molecules, although its gas storage capacity is limited because about 79.5% of the 512 cages are occupied by water molecules. SCH with an orthogonal structure is superior in terms of gas storage capacity because the ratio of water molecules with the 512 cages is smallest. Polymorphism plays an important role to the performance of CCS with SCH, and precise manipulation of hydrate structure is currently difficult. The performance of CCS is also influenced by the operation conditions, the reactor volume, the quaternary salt concentrations. The main obstacles of the CCS with SCH are low gas storage capacity and low formation kinetics. Efforts should be conducted to find the most appropriate quaternary salt to achieve better gas storage capacity as well as high thermal stability. Ways to improve the kinetic behaviors of CO2 and quaternary salt hydrate formation are necessary to scale up its industrial application.

Acknowledgements

This work was funded by the financial support from the China Geological Survey (No. DD20230063) and the Guangdong Major Project of Basic and Applied Basic Research (No. 2020B0301030003).

Conflict of Interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.
1
Acosta H Y, Bishnoi P R, Clarke M A. (2011). Experimental measurements of the thermodynamic equilibrium conditions of tetra-n-butylammonium bromide semiclathrates formed from synthetic landfill gases. Journal of Chemical & Engineering Data, 56(1): 69–73

DOI

2
Adisasmito S, Frank R J III, Sloan E D Jr. (1991). Hydrates of carbon dioxide and methane mixtures. Journal of Chemical & Engineering Data, 36(1): 68–71

DOI

3
Aladko L S, Dyadin Y A. (1996). Clathrate formation in the Bu4NClNH4ClH2O system. Mendeleev Communications, 6(5): 198–200

DOI

4
AladkoL SDyadin Y ARodionovaT VTerekhovaI S (2003). Effect of size and shape of cations and anions on clathrate formation in the system: halogenides of quaterly ammonium bases and water. Journal of Molecular Liquids, 106(2−3): 229−238

5
Arjmandi M, Chapoy A, Tohidi B. (2007). Equilibrium data of hydrogen, methane, nitrogen, carbon dioxide, and natural gas in semi-clathrate hydrates of tetrabutyl ammonium bromide. Journal of Chemical & Engineering Data, 52(6): 2153–2158

DOI

6
Babu P, Chin W I, Kumar R, Linga P. (2014a). Systematic evaluation of tetra-n-butylammonium bromide (TBAB) for carbon dioxide capture employing the clathrate process. Industrial & Engineering Chemistry Research, 53(12): 4878–4887

DOI

7
Babu P, Datta S, Kumar R, Linga P. (2014b). Impact of experimental pressure and temperature on semiclathrate hydrate formation for pre-combustion capture of CO2 using tetra-n-butyl ammonium nitrate. Energy, 78: 458–464

DOI

8
Babu P, Linga P, Kumar R, Englezos P. (2015). A review of the hydrate based gas separation (HBGS) process for carbon dioxide pre-combustion capture. Energy, 85: 261–279

DOI

9
Babu P, Yao M, Datta S, Kumar R, Linga P. (2014c). Thermodynamic and kinetic verification of tetra-n-butylammonium nitrate (TBANO3) as a promoter for the clathrate process applicable to precombustion carbon dioxide capture. Environmental Science & Technology, 48(6): 3550–3558

DOI

10
BelandriaVMohammadi A HEslamimaneshARichonDSánchez-Mora M FGalicia-LunaL A (2012). Phase equilibrium measurements for semi-clathrate hydrates of the (CO2+N2+tetra-n-butylammonium bromide) aqueous solution systems: Part 2. Fluid Phase Equilibria, 322–323: 105–112

11
Ben-Mansour R, Habib M A, Bamidele O E, Basha M, Qasem N A A, Peedikakkal A, Laoui T, Ali M. (2016). Carbon capture by physical adsorption: materials, experimental investigations and numerical modeling and simulations: a review. Applied Energy, 161: 225–255

DOI

12
Castellani B, Rossi F, Filipponi M, Nicolini A. (2014). Hydrate-based removal of carbon dioxide and hydrogen sulphide from biogas mixtures: experimental investigation and energy evaluations. Biomass and Bioenergy, 70: 330–338

DOI

13
Chapoy A, Anderson R, Tohidi B. (2007). Low-pressure molecular hydrogen storage in semi-clathrate hydrates of quaternary ammonium compounds. Journal of the American Chemical Society, 129(4): 746–747

DOI

14
Chapoy A, Gholinezhad J, Tohidi B. (2010). Experimental clathrate dissociations for the hydrogen+water and hydrogen+tetrabutylammonium bromide+water systems. Journal of Chemical & Engineering Data, 55(11): 5323–5327

DOI

15
Chazallon B, Ziskind M, Carpentier Y, Focsa C. (2014). CO2 capture using semi-clathrates of quaternary ammonium salt: structure change induced by CO2 and N2 enclathration. Journal of Physical Chemistry B, 118(47): 13440–13452

DOI

16
Chen Z, Deng S, Wei H, Wang B, Huang J, Yu G. (2013). Activated carbons and amine-modified materials for carbon dioxide capture: a review. Frontiers of Environmental Science & Engineering., 7(3): 326–340

DOI

17
Darbouret M, Cournil M, Herri J M. (2005). Rheological study of TBAB hydrate slurries as secondary two-phase refrigerants. International Journal of Refrigeration, 28(5): 663–671

DOI

18
DavidsonD W (1973). Water: A Comprehensive Treatise. New York: Plenum Press

19
Davis S J, Lewis N S, Shaner M, Aggarwal S, Arent D, Azevedo I L, Benson S M, Bradley T, Brouwer J, Chiang Y M. . (2018). Net-zero emissions energy systems. Science, 360(6396): eaas9793

DOI

20
Deschamps J, Dalmazzone D. (2009). Dissociation enthalpies and phase equilibrium for TBAB semi-clathrate hydrates of N2, CO2, N2+CO2 and CH4+CO2. Journal of Thermal Analysis and Calorimetry, 98(1): 113–118

DOI

21
Deschamps J, Dalmazzone D. (2010). Hydrogen storage in semiclathrate hydrates of tetrabutyl ammonium chloride and tetrabutyl phosphonium bromide. Journal of Chemical & Engineering Data, 55(9): 3395–3399

DOI

22
Duc N H, Chauvy F, Herri J M. (2007). CO2 capture by hydrate crystallization: a potential solution for gas emission of steelmaking industry. Energy Conversion and Management, 48(4): 1313–1322

DOI

23
Dyadin Y A, Terekhova I S, Polyanskaya T M, Aladko L S. (1977). Clathrate hydrates of tetrabutylammonium fluoride and oxalate. Journal of Structural Chemistry, 17(4): 566–571

DOI

24
Dyadin Y A, Udachin K A. (1984). Clathrate formation in water-peralkylonium salts systems. Journal of Inclusion Phenomena, 2(1): 61–72

25
Dyadin Y A, Udachin K A. (1987). Clathrate polyhydrates of peralkylonium salts and their analogs. Journal of Structural Chemistry, 28(3): 394–432

DOI

26
Fan S, Li Q, Nie J, Lang X, Wen Y, Wang Y. (2013). Semiclathrate hydrate phase equilibrium for CO2/CH4 gas mixtures in the presence of tetrabutylammonium halide (bromide, chloride, or fluoride). Journal of Chemical & Engineering Data, 58(11): 3137–3141

DOI

27
Fan S, Li S, Wang J, Lang X, Wang Y. (2009). Efficient capture of CO2 from simulated flue gas by formation of TBAB or TBAF semiclathrate hydrates. Energy & Fuels, 23(8): 4202–4208

DOI

28
Fan S, Long X, Lang X, Wang Y, Chen J. (2016). CO2 capture from CH4/CO2 mixture gas with tetra-n-butylammonium bromide semi-clathrate hydrate through a pressure recovery method. Energy & Fuels, 30(10): 8529–8534

DOI

29
Favre E. (2022). Membrane separation processes and post-combustion carbon capture: state of the art and prospects. Membranes (Basel), 12(9): 884

DOI

30
Font-Palma C, Cann D, Udemu C. (2021). Review of cryogenic carbon capture innovations and their potential applications. C, 7(3): 730058

DOI

31
Friedlingstein P, Jones M W, O’Sullivan M, Andrew R M, Bakker D C E, Hauck J, Le Quéré C, Peters G P, Peters W, Pongratz J. . (2022). Global carbon budget 2021. Earth System Science Data, 14(4): 1917–2005

DOI

32
Gaponenko L A, Solodovnikov S F, Dyadin Y A, Aladko L S, Polyanskaya T M. (1984). Crystallographic study of tetra-n-butylammonium bromide polyhydrates. Journal of Structural Chemistry, 25(1): 157–159

DOI

33
Gholinezhad J, Chapoy A, Tohidi B. (2011). Separation and capture of carbon dioxide from CO2/H2 syngas mixture using semi-clathrate hydrates. Chemical Engineering Research & Design, 89(9): 1747–1751

DOI

34
Nakayama H. (1982). Hydrates of organic compounds. VI. Heats of fusion and of solution of quaternary ammonium halide clathrate hydrates. Bulletin of the Chemical Society of Japan, 55(2): 389–393

DOI

35
Nakayama H. (1983). Hydrates of organic compounds. VII. The effect of anions on the formation of clathrate hydrates of tetrabutyl ammonium salts. Bulletin of the Chemical Society of Japan, 56(3): 877–880

DOI

36
NakayamaH (1987). Hydrates of organic compounds. XI. Determination of the melting point and hydration numbers of the clathrate-like hydrate of tetrabutyl ammonium chloride by differential scanning calorimetry. Bulletin of the Chemical Society of Japan, 60(3): 839−843

37
Hashimoto H, Ozeki H, Yamamoto Y, Muromachi S. (2020). CO2 capture from flue gas based on tetra-n-butylammonium fluoride hydrates at near ambient temperature. ACS Omega, 5(13): 7115–7123

DOI

38
Hashimoto H, Yamaguchi T, Kinoshita T, Muromachi S. (2017a). Gas separation of flue gas by tetra-n-butylammonium bromide hydrates under moderate pressure conditions. Energy, 129: 292–298

DOI

39
Hashimoto H, Yamaguchi T, Ozeki H, Muromachi S. (2017b). Structure-driven CO2 selectivity and gas capacity of ionic clathrate hydrates. Scientific Reports, 7(1): 17216

DOI

40
Hashimoto S, Murayama S, Sugahara T, Sato H, Ohgaki K. (2006). Thermodynamic and Raman spectroscopic studies on H2+tetrahydrofuran+water and H2+tetra-n-butyl ammonium bromide+water mixtures containing gas hydrates. Chemical Engineering Science, 61(24): 7884–7888

DOI

41
Hashimoto S, Sugahara T, Moritoki M, Sato H, Ohgaki K. (2008). Thermodynamic stability of hydrogen + tetra-n-butyl ammonium bromide mixed gas hydrate in nonstoichiometric aqueous solutions. Chemical Engineering Science, 63(4): 1092–1097

DOI

42
Herri J M, Bouchemoua A, Kwaterski M, Brântuas P, Galfré A, Bouillot B, Douzet J, Ouabbas Y, Cameirao A. (2014). Enhanced selectivity of the separation of CO2 from N2 during crystallization of semi-clathrates from quaternary ammonium solutions. Oil & Gas Science and Technology, 69(5): 947–968

43
Höhne G W H, Cammenga H K, Eysel W, Gmelin E, Hemminger W. (1990). The temperature calibration of scanning calorimeters. Thermochimica Acta, 160(1): 1–12

DOI

44
Horii S, Ohmura R. (2018). Continuous separation of CO2 from a H2+CO2 gas mixture using clathrate hydrate. Applied Energy, 225: 78–84

DOI

45
Iarikov D D, Hacarlioglu P, Oyama S T. (2011). Supported room temperature ionic liquid membranes for CO2/CH4 separation. Chemical Engineering Journal, 166(1): 401–406

DOI

46
Iino K, Sakakibara Y, Suginaka T, Ohmura R. (2014). Phase equilibria for the ionic semiclathrate hydrate formed with tetrabutyl phosphonium chloride plus CO2, CH4, or N2. Journal of Chemical Thermodynamics, 71: 133–136

DOI

47
Mohammadi A, Pakzad M, Mohammadi A H, Jahangiri A. (2018). Kinetics of (TBAF+CO2) semi-clathrate hydrate formation in the presence and absence of SDS. Petroleum Science, 15(2): 375–384

DOI

48
Jeffrey G A, McMullan R K. (1967). The clathrate hydrates. Progress in Inorganic Chemistry, 8: 43–108

49
Jin Y, Kida M, Nagao J. (2019). Crystal phase conditions of semiclathrate hydrates in nitrogen–tetra-n-butylammonium bromide–water systems below 1 MPa. Journal of Chemical & Engineering Data, 64(6): 2843–2848

DOI

50
Joshi A, Sangwai J S, Das K, Sami N A. (2013). Experimental investigations on the phase equilibrium of semiclathrate hydrates of carbon dioxide in TBAB with small amount of surfactant. International Journal of Energy and Environmental Engineering, 4(1): 11

DOI

51
Kamata Y, Yamakoshi Y, Ebinuma T, Oyama H, Shimada W, Narita H. (2005). Hydrogen sulfide separation using tetra-n-butyl ammonium bromide semi-clathrate (TBAB). Energy Fuels, 19(4): 1717–1722

DOI

52
KárászováM ZachBPetrusová ZČervenkaVBobákMŠyc MIzákP (2020). Post-combustion carbon capture by membrane separation. Separation and Purification Technology, 238: 116448 (Review)

53
Kharrat M, Dalmazzone D. (2003). Experimental determination of stability conditions of methane hydrate in aqueous calcium chloride solutions using high pressure differential scanning calorimetry. Journal of Chemical Thermodynamics, 35(9): 1489–1505

DOI

54
Kida M, Jin Y, Nagao J. (2019). Changes in the 13C NMR spectra of tetra-n-butylammonium chloride by clathrate hydration. Chemical Physics, 522: 233–237

DOI

55
Kim H, Zheng J, Yin Z, Babu P, Kumar S, Tee J, Linga P. (2023). Semi-clathrate hydrate slurry as a cold energy storage and transport medium: rheological study, energy analysis and enhancement by amino acid. Energy, 264: 126226

DOI

56
Kim H, Zheng J, Yin Z, Kumar S, Tee J, Seo Y, Linga P. (2022). An electrical resistivity-based method for measuring semi-clathrate hydrate formation kinetics: application for cold storage and transport. Applied Energy, 308: 118397

DOI

57
Kim S, Ko G, Kim K S, Seo Y. (2020). Phase equilibria of tetra-iso-amylammonium bromide (TiAAB) semiclathrates with CO2, N2, or CO2+N2. Journal of Chemical Thermodynamics, 142: 106024

DOI

58
Kim S, Seo Y. (2015). Semiclathrate-based CO2 capture from flue gas mixtures: an experimental approach with thermodynamic and Raman spectroscopic analyses. Applied Energy, 154: 987–994

DOI

59
Kim S M, Lee J D, Lee H J, Lee E K, Kim Y. (2011). Gas hydrate formation method to capture the carbon dioxide for pre-combustion process in IGCC plant. International Journal of Hydrogen Energy, 36(1): 1115–1121

DOI

60
Klara S M, Srivastava R D. (2002). U.S. DOE integrated collaborative technology development program for CO2 separation and capture. Environment and Progress, 21(4): 247–253

DOI

61
Kobori T, Muromachi S, Ohmura R. (2015a). Phase equilibrium for ionic semiclathrate hydrates formed in the system of water + tetra-n-butylammonium bromide pressurized with carbon dioxide. Journal of Chemical & Engineering Data, 60(2): 299–303

DOI

62
Kobori T, Muromachi S, Yamasaki T, Takeya S, Yamamoto Y, Alavi S, Ohmura R. (2015b). Phase behavior and structural characterization of ionic clathrate hydrate formed with tetra-n-butylphosphonium hydroxide: discovery of primitive crystal structure. Crystal Growth & Design, 15(8): 3862–3867

DOI

63
KomarovV YRodionova T VTerekhovaI SKuratievaN V (2007). The cubic superstructure-I of tetrabutylammonium fluoride (C4H9)4NF·29.7H2O clathrate hydrate. Journal of Inclusion Phenomena and Macrocyclic Chemistry, 59(1-2): 11-15

64
Komatsu H, Maruyama K, Yamagiwa K, Tajima H. (2019). Separation processes for carbon dioxide capture with semi-clathrate hydrate slurry based on phase equilibria of CO2+N2+tetra-n-butylammonium bromide+water systems. Chemical Engineering Research & Design, 150: 289–298

DOI

65
Koyama R, Hotta A, Ohmura R. (2020). Equilibrium temperature and dissociation heat of tetrabutylphosphonium acrylate (TBPAc) ionic semi-clathrate hydrate as a medium for the hydrate-based thermal energy storage system. Journal of Chemical Thermodynamics, 144: 106088

DOI

66
Lee S, Lee Y, Park S, Kim Y, Lee J D, Seo Y. (2012). Thermodynamic and spectroscopic identification of guest gas enclathration in the double tetra-n-butylammonium fluoride semiclathrates. Journal of Physical Chemistry B, 116(30): 9075–9081

DOI

67
Lee S, Lee Y, Park S, Seo Y. (2010). Phase Equilibria of semiclathrate hydrate for nitrogen in the presence of tetra-n-butylammonium bromide and fluoride. Journal of Chemical & Engineering Data, 55(12): 5883–5886

DOI

68
Lee S, Park S, Lee Y, Lee J, Lee H, Seo Y. (2011). Guest gas enclathration in semiclathrates of tetra-n-butylammonium bromide: stability condition and spectroscopic analysis. Langmuir, 27(17): 10597–10603

DOI

69
Li Q, Fan S, Wang Y, Lang X, Chen J. (2015). CO2 removal from biogas based on hydrate formation with tetra-n-butylammonium bromide solution in the presence of 1-butyl-3-methylimidazolium tetrafluoroborate. Energy & Fuels, 29(5): 3143–3148

DOI

70
Li S, Fan S, Wang J, Lang X, Liang D. (2009). CO2 capture from binary mixture via forming hydrate with the help of tetra-n-butylammonium bromide. Journal of Natural Gas Chemistry, 18(1): 15–20

DOI

71
Li S, Fan S, Wang J, Lang X, Wang Y. (2010a). Semiclathrate hydrate phase equilibria for CO2 in the presence of tetra-n-butyl ammonium halide (bromide, chloride, or fluoride). Journal of Chemical & Engineering Data, 55(9): 3212–3215

DOI

72
Li X S, Xia Z M, Chen Z Y, Wu H J. (2011). Precombustion capture of carbon dioxide and hydrogen with a one-stage hydrate/membrane process in the presence of tetra-n-butylammonium bromide (TBAB). Energy & Fuels, 25(3): 1302–1309

DOI

73
Li X S, Xia Z M, Chen Z Y, Yan K F, Li G, Wu H J. (2010b). Equilibrium hydrate formation conditions for the mixtures of CO2 + H2 + tetrabutyl ammonium bromide. Journal of Chemical & Engineering Data, 55(6): 2180–2184

DOI

74
Li X S, Xia Z M, Chen Z Y, Yan K F, Li G, Wu H J. (2010c). Gas hydrate formation process for capture of carbon dioxide from fuel gas mixture. Industrial & Engineering Chemistry Research, 49(22): 11614–11619

DOI

75
Li Z, Zhong D L, Lu Y Y, Yan J, Zou Z L. (2017). Preferential enclathration of CO2 into tetra-n-butylphosphonium bromide semiclathrate hydrate in moderate operating conditions: application for CO2 capture from shale gas. Applied Energy, 199: 370–381

DOI

76
Li Z, Zhong D L, Zheng W Y, Yan J, Lu Y Y, Yi D T. (2019). Morphology and kinetic investigation of TBAB/TBPB semiclathrate hydrates formed with a CO2+CH4 gas mixture. Journal of Crystal Growth, 511: 79–88

DOI

77
Lin W, Dalmazzone D, Fürst W, Delahaye A, Fournaison L, Clain P. (2013). Accurate DSC measurement of the phase transition temperature in the TBPB–water system. Journal of Chemical Thermodynamics, 61: 132–137

DOI

78
LinWDelahayeA FournaisonL (2008). Phase equilibrium and dissociation enthalpy for semi-clathrate hydrate of CO2+TBAB. Fluid Phase Equilibria, 264(1–2): 220–227

79
Linga P, Clarke M A. (2017). A review of reactor designs and materials employed for increasing the rate of gas hydrate formation. Energy & Fuels, 31(1): 1–13

DOI

80
Linga P, Kumar R, Englezos P. (2007a). The clathrate hydrate process for post and pre-combustion capture of carbon dioxide. Journal of Hazardous Materials, 149(3): 625–629

DOI

81
Linga P, Kumar R, Englezos P. (2007b). Gas hydrate formation from hydrogen/carbon dioxide and nitrogen/carbon dioxide gas mixtures. Chemical Engineering Science, 62(16): 4268–4276

DOI

82
LipkowskiJKomarov V YRodionovaT VDyadinY AAladkoL S (2002). The structure of tetrabutylammonium bromide hydrate (C4H9)4NBr·21/3H2O. Journal of Supramolecular Chemistry, 2(4–5): 435–439

83
Liu F P, Li A R, Qing S L, Luo Z D, Ma Y L. (2022). Formation kinetics, mechanism of CO2 hydrate and its applications. Renewable & Sustainable Energy Reviews, 159: 112221

DOI

84
Lundgren J O, Olovsson I. (1967). Hydrogen-bond studies. XVI. The crystal structure of chloride trihydrate. Acta Crystallographica, 23(6): 971–976

DOI

85
Lundgren J O, Olovsson I. (1968). Hydrogen-bond studies. XXX. The crystal structure of hydrogen bromide tetrahydrate, (H7O3)+(H9O4)+2Br·−+H2O. Journal of Chemical Physics, 49(3): 1068–1074

DOI

86
Ma Z W, Zhang P, Bao H S, Deng S. (2016). Review of fundamental properties of CO2 hydrates and CO2 capture and separation using hydration method. Renewable & Sustainable Energy Reviews, 53: 1273–1302

DOI

87
MaZ WZhang PWangR ZFuruiSXiG N (2010). Forced flow and convective melting heat transfer of clathrate hydrate slurry in tubes. International Journal of Heat and Mass Transfer, 53(19–20): 3745–3757

88
Machida H, Sugahara T, Hirasawa I. (2018). Memory effect in tetra-n-butyl ammonium bromide semiclathrate hydrate reformation: the existence of solution structures after hydrate decomposition. CrystEngComm, 20(24): 3328–3334

DOI

89
Machida H, Sugahara T, Masunaga H, Hirasawa I. (2020). Calorimetric and small-angle X-ray scattering studies on the memory effect in the tetra-n-butylammonium bromide semiclathrate hydrate system. Journal of Crystal Growth, 533: 125476

DOI

90
Majumdar A, Maini B, Bishnoi P R, Clarke M A. (2012). Three-phase equilibrium conditions of TiAAB semiclathrates formed from N2, CO2, and their mixtures. Journal of Chemical & Engineering Data, 57(8): 2322–2327

DOI

91
ManakovARodionova TTerekhovaIKomarovVBurdinA SizikovA (2011). Structural and physico-chemical studies of ionic clathrate hydrates of tetrabutyland tetraisoamylammonium salts. In: Proceedings of the 7th International Conference on Gas Hydrates, Edinburgh, Scotland, United Kingdom, July 17–21

92
Mayoufi N, Dalmazzone D, Fürst W, Delahaye A, Fournaison L. (2010). CO2 enclathration in hydrates of peralkyl-(ammonium/phosphonium) salts: stability conditions and dissociation enthalpies. Journal of Chemical & Engineering Data, 55(3): 1271–1275

DOI

93
Mayoufi N, Dalmazzone D, Delahaye A, Clain P, Fournaison L, Fürst W. (2011). Experimental data on phase behavior of simple tetrabutylphosphonium bromide (TBPB) and mixed CO2+TBPB semiclathrate hydrates. Journal of Chemical & Engineering Data, 56(6): 2987–2993

DOI

94
McMullan R, Jeffrey G A. (1959). Hydrates of the tetra n-butyl and tetra i-amyl quaternary ammonium salts. Journal of Chemical Physics, 31(5): 1231–1234

DOI

95
McMullan R K, Bonamico M, Jeffrey G A. (1963). Polyhedral clathrate hydrates. V. Structure of the tetra-n-butyl ammonium fluoride hydrate. Journal of Chemical Physics, 39(12): 3295–3310

DOI

96
Mehta A P, Makogon T Y, Burruss R C, Wendlandt R F, Sloan E D. (1996). A composite phase diagram of structure H hydrates using Schreinemakers’ geometric approach. Fluid Phase Equilibria, 121(1–2): 141–165

DOI

97
Meysel P, Oellrich L, Raj Bishnoi P, Clarke M A. (2011). Experimental investigation of incipient equilibrium conditions for the formation of semi-clathrate hydrates from quaternary mixtures of (CO2+N2+TBAB+H2O). Journal of Chemical Thermodynamics, 43(10): 1475–1479

DOI

98
MiwaYMatsumura KTakeyaKTaniA (2018). THz-TDS Study on Tetrabutylammonium Bromide Hydrate, 43rd International Conference on Infrared, Millimeter, and Terahertz Waves (IRMMW-THz), Nagoya, Japan, September 9–14

99
Miyamoto T, Koyama R, Kurokawa N, Hotta A, Alavi S, Ohmura R. (2020). Thermophysical property measurements of tetrabutylphosphonium oxalate (TBPOx) ionic semiclathrate hydrate as a media for the thermal energy storage system. Frontiers in Chemistry, 8: 547

DOI

100
Mohammadi A, Manteghian M, Mohammadi A H. (2013a). Dissociation data of semiclathrate hydrates for the systems of tetra-n-butylammonium fluoride (TBAF)+methane+water, TBAF+carbon dioxide+water, and TBAF+nitrogen+water. Journal of Chemical & Engineering Data, 58(12): 3545–3550

DOI

101
Mohammadi A H, Eslamimanesh A, Belandria V, Richon D. (2011). Phase equilibria of semiclathrate hydrates of CO2, N2, CH4, or H2+tetra-n-butylammonium bromide aqueous solution. Journal of Chemical & Engineering Data, 56(10): 3855–3865

DOI

102
Mohammadi A H, Eslamimanesh A, Belandria V, Richon D, Naidoo P, Ramjugernath D. (2012). Phase equilibrium measurements for semi-clathrate hydrates of the (CO2+N2+tetra-n-butylammonium bromide) aqueous solution system. Journal of Chemical Thermodynamics, 46: 57–61

DOI

103
Mohammadi A H, Eslamimanesh A, Richon D. (2013b). Semi-clathrate hydrate phase equilibrium measurements for the CO2+H2/CH4+tetra-n-butylammonium bromide aqueous solution system. Chemical Engineering Science, 94: 284–290

DOI

104
Momeni K, Jomekian A, Bazooyar B. (2020). Semi-clathrate hydrate phase equilibria of carbon dioxide in presence of tetra-n-butyl-ammonium chloride (TBAC): experimental measurements and thermodynamic modeling. Fluid Phase Equilibria, 508: 112445

DOI

105
Muromachi S. (2020). Phase equilibrium data for semiclathrate hydrates formed with tetra-n-butylammonium (bromide or chloride) and tetra-n-butylphosphonium (bromide or chloride) under hydrogen+carbon dioxide pressure. Fluid Phase Equilibria, 506: 112389

DOI

106
Muromachi S. (2021). CO2 capture properties of semiclathrate hydrates formed with tetra-n-butylammonium and tetra-n-butylphosphonium salts from H2+CO2 mixed gas. Energy, 223: 120015

DOI

107
Muromachi S, Hashimoto H, Maekawa T, Takeya S, Yamamoto Y. (2016a). Phase equilibrium and characterization of ionic clathrate hydrates formed with tetra-n-butylammonium bromide and nitrogen gas. Fluid Phase Equilibria, 413: 249–253

DOI

108
Muromachi S, Kida M, Takeya S, Yamamoto Y, Ohmura R. (2015). Characterization of the ionic clathrate hydrate of tetra-n-butylammonium acrylate. Canadian Journal of Chemistry, 93(9): 954–959

DOI

109
Muromachi S, Takeya S. (2017). Gas-containing semiclathrate hydrate formation by tetra-n-butylammonium carboxylates: acrylate and butyrate. Fluid Phase Equilibria, 441: 59–63

DOI

110
Muromachi S, Takeya S. (2018). Design of thermophysical properties of semiclathrate hydrates formed by tetra-n-butylammonium hydroxybutyrate. Industrial & Engineering Chemistry Research, 57(8): 3059–3064

DOI

111
Muromachi S, Takeya S. (2019). Thermodynamic properties and crystallographic characterization of semiclathrate hydrates formed with tetra-n-butylammonium glycolate. ACS Omega, 4(4): 7317–7322

DOI

112
Muromachi S, Takeya S, Alavi S, Ripmeester J A. (2022). Structural CO2 capture preference of semiclathrate hydrate formed with tetra-n-butylammonium chloride. CrystEngComm, 24(24): 4366–4371

DOI

113
Muromachi S, Takeya S, Yamamoto Y, Ohmura R. (2014a). Characterization of tetra-n-butylphosphonium bromide semiclathrate hydrate by crystal structure analysis. CrystEngComm, 16(10): 2056–2060

DOI

114
Muromachi S, Udachin K A, Alavi S, Ohmura R, Ripmeester J A. (2016b). Selective occupancy of methane by cage symmetry in TBAB ionic clathrate hydrate. Chemical Communications (Cambridge), 52(32): 5621–5624

DOI

115
Muromachi S, Udachin K A, Shin K, Alavi S, Moudrakovski I L, Ohmura R, Ripmeester J A. (2014b). Guest-induced symmetry lowering of an ionic clathrate material for carbon capture. Chemical Communications (Cambridge), 50(78): 11476–11479

DOI

116
Oshima M, Jin Y, Kida M, Nagao J. (2020). Thermodynamic and crystallographic properties depending on hydration numbers in tetra-n-butylammonium chloride semiclathrate hydrates. Journal of Chemical Thermodynamics, 142: 106004

DOI

117
Oshima M, Kida M, Nagao J. (2018). Hydration numbers and thermal properties of tetra-n-butylammonium bromide semiclathrate hydrates determined by ion chromatography and differential scanning calorimetry. Journal of Chemical Thermodynamics, 123: 32–37

DOI

118
Oshima M, Shimada W, Hashimoto S, Tani A, Ohgaki K. (2010). Memory effect on semi-clathrate hydrate formation: a case study of tetragonal tetra-n-butyl ammonium bromide hydrate. Chemical Engineering Science, 65(20): 5442–5446

DOI

119
OyamaHEbinuma TShimadaWTakeyaSNagao JUchidaTNaritaH (2003). An experimental study of gas-hydrate formation by measuring viscosity and infrared spectra. Canadian Journal of Physics, 81(1–2): 485–492

120
OyamaHShimada WEbinumaTKamataYTakeyaS UchidaTNagao JNaritaH (2005). Phase diagram, latent heat, and specific heat of TBAB semiclathrate hydrate crystals. Fluid Phase Equilibria, 234(1–2): 131–135

121
Park S, Lee S, Lee Y, Seo Y. (2013). CO2 capture from simulated fuel gas mixtures using semiclathrate hydrates formed by quaternary ammonium salts. Environmental Science & Technology, 47(13): 7571–7577

DOI

122
Rakkappan S R, Sivan S, Praveen B, Naarendharan M, Sai Sudhir P. (2021). Thermal property, charging and discharging characteristics study on tetra-n-butyl ammonium bromide semi-clathrate hydrates for air-conditioning cold storage and secondary refrigerant applications. Journal of Chemical Thermodynamics, 153: 106275

DOI

123
Rodionova T, Komarov V, Villevald G, Aladko L, Karpova T, Manakov A. (2010). Calorimetric and structural studies of tetrabutylammonium chloride ionic clathrate hydrates. Journal of Physical Chemistry B, 114(36): 11838–11846

DOI

124
Rodionova T V, Komarov V Y, Villevald G V, Karpova T D, Kuratieva N V, Manakov A Y. (2013). Calorimetric and structural studies of tetrabutylammonium bromide ionic clathrate hydrates. Journal of Physical Chemistry B, 117(36): 10677–10685

DOI

125
RodionovaT VManakovA YSteninY G VillevaldG VKarpovaT D (2008). The heats of fusion of tetrabutylammonium fluoride ionic clathrate hydrates. Journal of Inclusion Phenomena and Macrocyclic Chemistry, 61(1–2): 107–111

126
Rodionova T V, Terekhova I S, Manakov A Y. (2022). Ionic clathrate hydrates of tetraalkylammonium/phosphonium salts: structures, properties, some applications, and perspectives. Energy & Fuels, 36(18): 10458–10477

DOI

127
Rodriguez C T, Le Q D, Focsa C, Pirim C, Chazallon B. (2020). Influence of crystallization parameters on guest selectivity and structures in a CO2-based separation process using TBAB semi-clathrate hydrates. Chemical Engineering Journal, 382: 122867

DOI

128
Sakamoto H, Sato K, Shiraiwa K, Takeya S, Nakajima M, Ohmura R. (2011). Synthesis, characterization and thermal-property measurements of ionic semi-clathrate hydrates formed with tetrabutylphosphonium chloride and tetrabutylammonium acrylate. RSC Advances, 1(2): 315–322

DOI

129
Sakamoto J, Hashimoto S, Tsuda T, Sugahara T, Inoue Y, Ohgaki K. (2008). Thermodynamic and Raman spectroscopic studies on hydrogen+tetra-n-butyl ammonium fluoride semi-clathrate hydrates. Chemical Engineering Science, 63(24): 5789–5794

DOI

130
Sales Silva L P, Dalmazzone D, Stambouli M, Lesort A L, Arpentinier P, Trueba A, Fürst W. (2016). Phase equilibria of semi-clathrate hydrates of tetra-n-butyl phosphonium bromide at atmospheric pressure and in presence of CH4 and CO2+CH4. Fluid Phase Equilibria, 413: 28–35

DOI

131
Sánchez-Mora M F, Galicia-Luna L A, Pimentel-Rodas A, Mohammadi A H. (2019). Experimental Determination of gas hydrates dissociation conditions in CO2/N2+ethanol/1-propanol/TBAB/TBAF+water systems. Journal of Chemical & Engineering Data, 64(2): 763–770

DOI

132
Sato K, Tokutomi H, Ohmura R. (2013). Phase equilibrium of ionic semiclathrate hydrates formed with tetrabutylammonium bromide and tetrabutylammonium chloride. Fluid Phase Equilibria, 337: 115–118

DOI

133
Shimada J, Shimada M, Sugahara T, Tsunashima K. (2019). Phase equilibrium relations of tetra-n-butylphosphonium propionate and butyrate semiclathrate hydrates. Fluid Phase Equilibria, 485: 61–66

DOI

134
Shimada J, Shimada M, Sugahara T, Tsunashima K, Tani A, Tsuchida Y, Matsumiya M. (2018). Phase equilibrium relations of semiclathrate hydrates based on tetra-n-butylphosphonium formate, acetate, and lactate. Journal of Chemical & Engineering Data, 63(9): 3615–3620

DOI

135
ShimadaWEbinuma TOyamaHKamataYNaritaH (2005a). Free-growth forms and growth kinetics of tetra-n-butyl ammonium bromide semi-clathrate hydrate crystals. Journal of Crystal Growth, 274(1–2): 246–250

136
Shimada W, Ebinuma T, Oyama H, Kamata Y, Takeya S, Uchida T, Nagao J, Narita H. (2003). Separation of gas molecule using tetra-n-butyl ammonium bromide semi-clathrate hydrate crystals. Japanese Journal of Applied Physics, 42(2A): L129–131

137
Shimada W, Shiro M, Kondo H, Takeya S, Oyama H, Ebinuma T, Narita H. (2005b). Tetra-n-butylammonium bromide-water (1/38). Acta Crystallographica. Section C, Crystal Structure Communications, 61(2): o65–o66

DOI

138
Stoporev A, Mendgaziev R, Artemova M, Semenov A, Novikov A, Kiiamov A, Emelianov D, Rodionova T, Fakhrullin R, Shchukin D. (2020). Ionic clathrate hydrates loaded into a cryogel–halloysite clay composite for cold storage. Applied Clay Science, 191: 105618

DOI

139
Stoporev A S, Kiiamov A G, Varfolomeev M A, Rodionova T V, Manakov A Y. (2021). Metastable ionic cubic structure I clathrate hydrate formed with tetra-n-butylammonium bromide. Mendeleev Communications, 31(1): 17–19

DOI

140
Sugahara T, Machida H. (2017). Dissociation and nucleation of tetra-n-butyl ammonium bromide semi-clathrate hydrates at high pressures. Journal of Chemical & Engineering Data, 62(9): 2721–2725

DOI

141
Suginaka T, Sakamoto H, Iino K, Sakakibara Y, Ohmura R. (2013). Phase equilibrium for ionic semiclathrate hydrate formed with CO2, CH4, or N2 plus tetrabutylphosphonium bromide. Fluid Phase Equilibria, 344: 108–111

DOI

142
Suginaka T, Sakamoto H, Iino K, Takeya S, Nakajima M, Ohmura R. (2012). Thermodynamic properties of ionic semiclathrate hydrate formed with tetrabutylphosphonium bromide. Fluid Phase Equilibria, 317: 25–28

DOI

143
Sun Q, Guo X, Liu A, Liu B, Huo Y, Chen G. (2011a). Experimental study on the separation of CH4 and N2 via hydrate formation in TBAB solution. Industrial & Engineering Chemistry Research, 50(4): 2284–2288

DOI

144
Sun Z G, Jiang C M, Xie N L. (2008). Hydrate equilibrium conditions for tetra-n-butyl ammonium bromide. Journal of Chemical & Engineering Data, 53(10): 2375–2377

DOI

145
Sun Z G, Liu C G, Zhou B, Xu L Z. (2011b). Phase equilibrium and latent heat of tetra-n-butylammonium chloride semi-clathrate hydrate. Journal of Chemical & Engineering Data, 56(8): 3416–3418

DOI

146
Tang J, Zeng D, Wang C, Chen Y, He L, Cai N. (2013). Study on the influence of SDS and THF on hydrate-based gas separation performance. Chemical Engineering Research & Design, 91(9): 1777–1782

DOI

147
Tohidi B, Burgass R, Danesh A, Todd A. (1994). Experimental study on the causes of disagreements in methane hydrate dissociation data. Annals of the New York Academy of Sciences, 715(1): 532–534

DOI

148
Tohidi B, Burgass R W, Danesh A, Østergaard K K, Todd A C. (2000). Improving the accuracy of gas hydrate dissociation point measurements. Annals of the New York Academy of Sciences, 912(1): 924–931

DOI

149
Trueba A T, Radović I R, Zevenbergen J F, Peters C J, Kroon M C. (2013). Kinetic measurements and in situ Raman spectroscopy study of the formation of TBAF semi-hydrates with hydrogen and carbon dioxide. International Journal of Hydrogen Energy, 38(18): 7326–7334

DOI

150
UdachinK ALipkowski J (2002). Water-fluorine chains in (n-Bu)4NF·5.5H2O hydrate. Journal of Supramolecular Chemistry, 2(4-5): 449–451

151
Veluswamy H P, Chin W I, Linga P. (2014). Clathrate hydrates for hydrogen storage: the impact of tetrahydrofuran, tetra-n-butylammonium bromide and cyclopentane as promoters on the macroscopic kinetics. International Journal of Hydrogen Energy, 39(28): 16234–16243

DOI

152
Wang F, Fu S, Guo G, Jia Z Z, Luo S J, Guo R B. (2016). Experimental study on hydrate-based CO2 removal from CH4/CO2 mixture. Energy, 104: 76–84

DOI

153
Wang X, Dennis M. (2015). An experimental study on the formation behavior of single and binary hydrates of TBAB, TBAF and TBPB for cold storage air conditioning applications. Chemical Engineering Science, 137: 938–946

DOI

154
Wang X, Song C. (2020). Carbon capture from flue gas and the atmosphere: a perspective. Frontiers in Energy Research, 8: 560849

DOI

155
Wang X, Zhang F, Lipiński W. (2020a). Research progress and challenges in hydrate-based carbon dioxide capture applications. Applied Energy, 269: 114928

DOI

156
Wang Y, Zhong D L, Li Z, Li J B. (2020b). Application of tetra-n-butylammonium bromide semi-clathrate hydrate for CO2 capture from unconventional natural gases. Energy, 197: 117209

DOI

157
Wilberforce T, Baroutaji A, Soudan B, Al-Alami A H, Olabi A G. (2019). Outlook of carbon capture technology and challenges. Science of the Total Environment, 657: 56–72

DOI

158
Xia Z M, Chen Z Y, Li X S, Zhang Y, Yan K F, Lv Q N, Xu C G, Cai J. (2012). Thermodynamic equilibrium conditions for simulated landfill gas hydrate formation in aqueous solutions of additives. Journal of Chemical & Engineering Data, 57(11): 3290–3295

DOI

159
Xia Z M, Li X S, Chen Z Y, Li G, Yan K F, Xu C G, Lv Q N, Cai J. (2016). Hydrate-based CO2 capture and CH4 purification from simulated biogas with synergic additives based on gas solvent. Applied Energy, 162: 1153–1159

DOI

160
Xie F M, Li X Y, Zhong D L, Englezos P, Lu G X. (2021). A Calorimetric study on the phase behavior of tetra-n-butyl phosphonium bromide+CO2 semiclathrate hydrate and evaluation of CO2 consumption—impact of a surfactant. Journal of Chemical & Engineering Data, 66(11): 4228–4235

DOI

161
Xu C G, Zhang S H, Cai J, Chen Z Y, Li X S. (2013). CO2 (carbon dioxide) separation from CO2–H2 (hydrogen) gas mixtures by gas hydrates in TBAB (tetra-n-butyl ammonium bromide) solution and Raman spectroscopic analysis. Energy, 59: 719–725

DOI

162
Yamauchi Y, Arai Y, Yamasaki T, Endo F, Hotta A, Ohmura R. (2017a). Phase equilibrium temperature and dissociation heat of ionic semiclathrate hydrate formed with tetrabutylammonium butyrate. Fluid Phase Equilibria, 441: 54–58

DOI

163
Yamauchi Y, Yamasaki T, Endo F, Hotta A, Ohmura R. (2017b). Thermodynamic properties of ionic semiclathrate hydrate formed with tetrabutylammonium propionate. Chemical Engineering & Technology, 40(10): 1810–1816

DOI

164
Yan J, Lu Y Y, Zhong D L, Qing S L. (2019). Insights into the phase behaviour of tetra-n-butyl ammonium bromide semi-clathrates formed with CO2, (CO2+CH4) using high-pressure DSC. Journal of Chemical Thermodynamics, 137: 101–107

DOI

165
Yang H, Huang X, Hu J L, Thompson R, Roger J F. (2022). Achievements, challenges and global implications of China’s carbon neutral pledge. Frontiers of Environmental Science & Engineering, 16(8): 111

DOI

166
Ye N, Zhang P. (2012). Equilibrium data and morphology of tetra-n-butyl ammonium bromide semiclathrate hydrate with carbon dioxide. Journal of Chemical & Engineering Data, 57(5): 1557–1562

DOI

167
Ye N, Zhang P. (2014a). Phase equilibrium and morphology characteristics of hydrates formed by tetra-n-butyl ammonium chloride and tetra-n-butylphosphonium chloride with and without CO2. Fluid Phase Equilibria, 361: 208–214

DOI

168
Ye N, Zhang P. (2014b). Phase equilibrium conditions and carbon dioxide separation efficiency of tetra-n-butylphosphonium bromide hydrate. Journal of Chemical & Engineering Data, 59(9): 2920–2926

DOI

169
Yue G, Liu A, Sun Q, Li X, Lan W, Yang L, Guo X. (2018). The combination of 1-octyl-3-methylimidazolium tetrafluorborate with TBAB or THF on CO2 hydrate formation and CH4 separation from biogas. Chinese Journal of Chemical Engineering, 26(12): 2495–2502

DOI

170
Zang X, Liang D. (2017). Phase equilibrium data for semiclathrate hydrate of synthesized binary CO2/CH4 gas mixture in tetra-n-butylammonium bromide aqueous solution. Journal of Chemical & Engineering Data, 62(2): 851–856

DOI

171
Zhang P, Ye N, Zhu H, Xiao X. (2013). Hydrate equilibrium conditions of tetra-n-butylphosphonium bromide+carbon dioxide and the crystal morphologies. Journal of Chemical & Engineering Data, 58(6): 1781–1786

DOI

172
Zheng J, Bhatnagar K, Khurana M, Zhang P, Zhang B Y, Linga P. (2018). Semiclathrate based CO2 capture from fuel gas mixture at ambient temperature: effect of concentrations of tetra-n-butylammonium fluoride (TBAF) and kinetic additives. Applied Energy, 217: 377–389

DOI

173
Zheng J, Zhang P, Linga P. (2017). Semiclathrate hydrate process for pre-combustion capture of CO2 at near ambient temperatures. Applied Energy, 194: 267–278

DOI

174
Zhong D, Englezos P. (2012). Methane separation from coal mine methane gas by tetra-n-butylammonium bromide semiclathrate hydrate formation. Energy & Fuels, 26(4): 2098–2106

DOI

175
Zhong D L, Wang W C, Zou Z L, Lu Y Y, Yan J, Ding K. (2018). Investigation on methane recovery from low-concentration coal mine gas by tetra-n-butyl ammonium chloride semiclathrate hydrate formation. Applied Energy, 227: 686–693

DOI

176
Zhou X, Liang D. (2019). Enhanced performance on CO2 adsorption and release induced by structural transition that occurred in TBAB·26H2O hydrates. Chemical Engineering Journal, 378(12): 122128

DOI

177
Zhuang Q, Clements B, Dai J, Carrigan L. (2016). Ten years of research on phase separation absorbents for carbon capture: achievements and next steps. International Journal of Greenhouse Gas Control, 52: 449–460

DOI

Outlines

/