You are on page 1of 31

J.

of Supercritical Fluids 43 (2007) 150180

Review article

A review of ionic liquids towards supercritical uid applications


Seda Keskin, Defne Kayrak-Talay, U ur Akman , Oner Hortacsu g
Department of Chemical Engineering, Bo azi i University, Bebek 34342, Istanbul, Turkey g c Received 2 August 2006; received in revised form 8 May 2007; accepted 29 May 2007

Abstract Ionic liquids (ILs), considered to be a relatively recent magical chemical due their unique properties, have a large variety of applications in all areas of the chemical industries. The areas of application include electrolyte in batteries, lubricants, plasticizers, solvents and catalysis in synthesis, matrices for mass spectroscopy, solvents to manufacture nano-materials, extraction, gas absorption agents, etc. Non-volatility and nonammability are their common characteristics giving them an advantageous edge in various applications. This common advantage, when considered with the possibility of tuning the chemical and physical properties of ILs by changing anioncation combination is a great opportunity to obtain task-specic ILs for a multitude of specic applications. There are numerous studies in the related literature concerning the unique properties, preparation methods, and different applications of ILs in the literature. In this review, a general description of ILs and historical background are given; basic properties of ILs such as solvent properties, polarity, toxicology, air and moisture stability are discussed; structure of ILs, cation, anion types and synthesis methods in the related literature are briey summarized. However, the main focus of this paper is how ILs may be used in the chemicals processing industries. Thus, the main application areas are searched and the basic applications such as solvent replacement, purication of gases, homogenous and heterogeneous catalysis, biological reactions media and removal of metal ions are discussed in detail. Not only the advantages of ILs but also the essential challenges and potentials for using ILs in the chemical industries are also addressed. ILs have become the partner of scCO2 in many applications and most of the reported studies in the literature focus on the interaction of these two green solvents, i.e. ILs and scCO2 . The chemistry of the ILs has been reviewed in numerous papers earlier. Therefore, the major purpose of this review paper is to provide an overview for the specic chemical and physical properties of ILs and to investigate ILscCO2 systems in some detail. Recovery of solutes from ILs with CO2 , separation of ILs from organic solvents by CO2 , high-pressure phase behavior of ILscCO2 systems, solubility of ILs in CO2 phase, and the interaction of the ILscCO2 system at molecular level are also included. 2007 Elsevier B.V. All rights reserved.
Keywords: Ionic liquids; Supercritical carbon dioxide; Review

Contents
1. 2. 3. General description of ILs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . History of ILs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Basic properties of ILs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.1. Solvent properties of ILs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2. Polarity of ILs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3. Toxicology of ILs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.4. Air and moisture stability of ILs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Structure and synthesis of ILs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.1. Anions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2. Cations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.3. Synthesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Major applications suggested for ILs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.1. Solvent replacement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151 151 151 153 154 154 155 155 156 156 156 156 157

4.

5.

Corresponding author. Tel.: +90 212 3596867; fax: +90 212 2872460. E-mail address: akman@boun.edu.tr (U. Akman).

0896-8446/$ see front matter 2007 Elsevier B.V. All rights reserved. doi:10.1016/j.supu.2007.05.013

S. Keskin et al. / J. of Supercritical Fluids 43 (2007) 150180

151

6. 7.

8.

5.2. Purication of gases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3. Homogenous and heterogeneous catalysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.4. Biological reactions media . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.5. Removing of metal ions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Challenges of ILs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ILs and scCO2 systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.1. High-pressure phase behavior of ILCO2 systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.1.1. The [bmim][PF6 ]CO2 system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.1.2. Other ILCO2 systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.2. IL solubility in CO2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.3. ILCO2 interaction at the molecular level. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.4. Solute recovery from ILs with scCO2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.5. Other applications of ILscCO2 systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

157 158 159 159 160 161 163 163 166 167 168 168 170 175 176 176

1. General description of ILs Ionic liquids (ILs) have been accepted as a new green chemical revolution which excited both the academia and the chemical industries. This new chemical group can reduce the use of hazardous and polluting organic solvents due to their unique characteristics as well as taking part in various new syntheses. The terms room temperature ionic liquid (RTIL), nonaqueous ionic liquid, molten salt, liquid organic salt and fused salt have all been used to describe these salts in the liquid phase [1]. ILs are known as salts that are liquid at room temperature in contrast to high-temperature molten salts. They have a unique array of physico-chemical properties which make them suitable in numerous applications in which conventional organic solvents are not sufciently effective or not applicable. Short [2] pointed out in 1980, that there were only a few patent applications for ILs, in 2000, the number of patent applications increased to 100, and nally by 2004, there were more than 800. This is a clear indication of the high afnity of the academia and industry to the ILs. 2. History of ILs ILs have been known for a long time, but their extensive use as solvents in chemical processes for synthesis and catalysis has recently become signicant. Welton [1] reported that ILs are not new, and some of the ILs such as [EtNH3 ][NO3 ] was rst described in 1914 [3]. The earliest IL in the literature was created intentionally in 1970s for nuclear warheads batteries [4]. During 1940s, aluminum chloride-based molten salts were utilized for electroplating at temperatures of hundreds of degrees Celsius. In the early 1970s, Wilkes tried to develop better batteries for nuclear warheads and space probes which required molten salts to operate [4]. These molten salts were hot enough to damage the nearby materials. Therefore, the chemists searched for salts which remain liquid at lower temperatures and eventually they identied one which is liquid at room temperature. Wilkes and his colleagues continued to improve their ILs for use as battery electrolytes and then a small community of researchers began

to make ILs and test their properties [5,6]. In the late 1990s, ILs became one of the most promising chemicals as solvents. The rst ILs, such as organo-aluminate ILs, have limited range of applications because they were unstable to air and water. Furthermore, these ILs were not inert towards various organic compounds [7]. After the rst reports on the synthesis and applications of air stable ILs such as 1-n-butyl-3-methlyimidazolium tetrauoroborate ([bmim][BF4 ]) and 1-n-butyl-3-methlyimidazolium hexauorophosphate ([bmim][PF6 ]), the number of air and water stable ILs has started to increase rapidly [7]. Recently, researchers have discovered that ILs are more than just green solvents and they have found several applications such as replacing them with volatile organic solvents, making new materials, conducting heat effectively, supporting enzyme-catalyzed reactions, hosting a variety of catalysts, purication of gases, homogenous and heterogeneous catalysis, biological reactions media and removal of metal ions [4]. Some of the basic physical properties of ILs such as density and viscosity are still being evaluated by the researchers since the study of the IL is a relatively young eld [8]. The number of research on ILs and their specic applications is increasing rapidly in the literature. For example, the cation 1-n-ethyl-3methylimidazolium has been the most widely studied until 2001, and nowadays, 1-3-dialkyl imidazolium salts are the most popularly used and investigated class of ILs. For the future of ILs, the aim of research is the commercialization of ILs in order to use them as solvents, reagents, catalysts and materials in large-scale chemical applications. 3. Basic properties of ILs ILs are made of positively and negatively charged ions, whereas water and organic solvents, such as toluene and dichloromethane, are made of molecules. The structure of ILs is similar to the table salt such as sodium chloride which contains crystals made of positive sodium ions and negative chlorine ions, not molecules. While, salts do not melt below 800 C, most of ILs remain liquid at room temperature. The melting points of sodium

152

S. Keskin et al. / J. of Supercritical Fluids 43 (2007) 150180

chloride and lithium chloride are known as 801 and 614 C, respectively. Since these conventional molten salts exhibit high melting points, their use as solvents in applications is severely limited. However, RTILs are liquid generally up to 200 C. ILs have a wide liquidus ranges. The adopted upper melting temperature limit for the classication as IL is known as 100 C and higher melting ion systems are generally referred as molten salts. Researchers explained that ILs remain liquid at room temperature due to the reason that their ions do not pack well [9]. Combination of bulky and asymmetrical cations and evenly shaped anions form a regular structure namely a liquid phase. The low melting points of ILs are a result of the chemical composition. The combination of larger asymmetric organic cation and smaller inorganic counterparts lower the lattice energy and hence the melting point of the resulting ionic medium. In some cases, even the anions are relatively large and play a role in lowering the melting point [10]. Most widely used ILs and their structures are given in Table 1. As solvents, ILs posses several advantages over conventional organic solvents, which make them environmentally compatible [1,4,8,1015]: ILs have the ability to dissolve many different organic, inorganic and organometallic materials. ILs are highly polar. ILs consist of loosely coordinating bulky ions. ILs do not evaporate since they have very low vapor pressures. ILs are thermally stable, approximately up to 300 C. Most of ILs have a liquid window of up to 200 C which enables wide kinetic control. ILs have high thermal conductivity and a large electrochemical window. ILs are immiscible with many organic solvents. ILs are nonaqueous polar alternatives for phase transfer processes. The solvent properties of ILs can be tuned for a specic application by varying the anion cation combinations. Generally, the above statements are valid for the most commonly used ILs. However, one should note that there are many ILs containing different anions and cations and their properties cover a vast range. Therefore, the above statements should not be generalized for all existing ILs and for those designed in the future. ILs exhibit the ability to dissolve a wide variety of materials including salts, fats, proteins, amino acids, surfactants, sugars and polysaccharides. ILs have very powerful solvent properties such that they can dissolve a wide range of organic molecules, including crude oil, inks, plastics, and even DNA [9]. Two important groups of ILs are those based on imidazolium and pyridinium cations with PF6 and BF4 anions [13,14]. Figs. 1 and 2 illustrate the imidazolium and pyridinium derivatives of ILs and their possible anions which are extensively investigated in literature. ILs tend not to give off vapors in contrast to traditional organic solvents such as benzene, acetone, and toluene. The

Fig. 1. Imidazolium derivatives of ILs (www.sigmaaldrich.com).

vapor pressures of the ILs are extremely low and are considered as negligible. For example, Kabo et al. [16] gave the vapor pressure of [bmim][PF6 ] at 298.15 K as 1011 Pa. ILs are introduced as green solvents because unlike the volatile organic compounds (VOCs) they replace, many of these compounds have negligible vapor pressure, they are not explosive and it may be feasible to recycle and repeatedly reuse them. It is more convenient to work with ILs in the laboratory since the non-evaporating properties of ILs eliminate the hazardous exposure and air pollution problems. ILs are also known as designer solvents since they give the opportunity to tune their specic properties for a particular need. The researchers can design a specic IL by choosing negatively charged small anions and positively charged large cations, and these specic ILs may be utilized to dissolve a certain chemical or to extract a certain material from a solution. The ne-tuning of the structure provides tailor-designed properties to satisfy the specic application requirements. The physical and chemical properties of ILs are varied by changing the alkyl chain length on the cation and the anion. For example, Huddleston et al. [17] concluded that density of ILs increases with a decrease in the alkyl chain length on the cation and an increase in the molecular weight of the anion. Although ILs are studied by a great number of research groups, there are still many questions that scientist are not able to answer. For example, one of the basic rules of chemistry like dissolves like is seem to be broken by some ILs: Nonpolar benzene is up to 50% soluble (by volume) in polar tetrachloroaluminate-based ILs [9]. Therefore, studies on why ILs are able to dissolve uncharged covalent molecules are continuing. Until recently, ILs have been considered to be scarce but it is now known that many salts form liquids at or close to room temperature. There are literally billions of different structures that may form an IL. The composition and the specic properties of these liquids depend on the type of cation and anion in the IL structure. By combining various kinds of cation and anion structures, it is estimated that 1018 ILs can be designed [18,19].

Fig. 2. Pyridinium derivatives of ILs (www.sigmaaldrich.com).

S. Keskin et al. / J. of Supercritical Fluids 43 (2007) 150180 Table 1 Most widely used ILs, their structures and short names Ionic liquid Structure Short name

153

1-Butyl-3-methylimidazolium tetrauoroborate

[bmim][BF4]

1-Butyl-3-methylimidazolium triate

[bmim][TfO]

1-Butyl-3-methylimidazolium methide

[bmim][methide]

1-Butyl-3-methylimidazolium dicyanamide

[bmim][DCA]

1-Butyl-3-methylimidazolium hexauorophosphate

[bmim][PF6 ]

1-Butyl-3-methylimidazolium nitrate

[bmim][NO3 ]

1-Butyl-3-methylimidazolium bis(triuoromethylsulfony1) imide

[bmim][Tf2 N]

l-Hexyl-3-methylimidazolium bis(triuoromethylsulfonyl) imide; R = C6 H17

[hmim][Tf2 N]

l-Octyl-3-methylimidazolium bis(triuoromethylsulfonyl) imide; R = C8 H17

[omim][Tf2 N]

2,3-Dimethyl-1-hexylimidazolium bis(triuoromethylsulfonyl) imide

[hmmim][Tf2 N]

Reprinted with permission from [121]. Copyright 2004 American Chemical Society.

3.1. Solvent properties of ILs Both the chemical industry and academia search for alternative solvents to meet the cleaner technology requirements

since the most widely used solvents are volatile and damaging. ILs are good solvents for a wide range of substances; organic, inorganic, organometallic compounds, bio-molecules and metal ions. They are usually composed of poorly coordinat-

154

S. Keskin et al. / J. of Supercritical Fluids 43 (2007) 150180

ing ions which makes them highly polar but non-coordinating solvents. ILs are immiscible with most of the organic solvents, thus they provide a nonaqueous, polar alternative for two-phase systems [19]. Furthermore, ILs which are not miscible with water can be used as immiscible polar phases with water. Although all other conventional solvents evaporate to the atmosphere, ILs do not evaporate and their nonvolatility gives an opportunity to utilize them in highvacuum systems. The negligible volatility is the basic property which characterizes them as green solvents. Considering potential as solvents, ILs can easily replace other conventional organic solvents which are used in large quantities in chemicals processing industries to eliminate major environmental problems. Many chemical reactions are carried out in conventional solvents. Upon the completion of reaction, chemical products must be taken out of the solvent. There are several techniques to recover a product from a solvent: For example, water-soluble compounds may be extracted with water; distillation may be used for chemicals with high vapor pressures. On the other hand, for the chemicals with low vapor pressures, distillation must be performed at low pressures, which may not be economical. In addition to this, there are some chemicals that can decompose as a result of heating, such as pharmaceuticals. Therefore, ILs seem to be potentially good solvents for many chemical reactions in the cases where distillation is not practical, or water insoluble or thermally sensitive products are the components of a chemical reaction. Although, ILs are not considered to be distilled due to their low volatility, Earle et al. [20] showed that many ionic liquids, especially bistriamide ILs, can be distilled at 200300 C and low pressure without decomposition. It was once more understood that there is a long way for total investigation of the properties of ILs. The authors suggested that the possibility of IL distillation introduced a new method for IL purication, and also new application areas (such as isolation of highly soluble products by hightemperature crystallization) could emerge. But, distillation still cannot be applied when heat-labile products are encountered in ILs. In most chemical applications, extraction is used for separation since it is an energy efcient technique. Generally, extraction consists of two immiscible phases such as an organic phase and an aqueous phase. Many organic solvents used in extractions are known with their ammable and toxical properties. In order to improve the safety and environmental friendliness of this conventional technique, ILs may be used as ideal substitutes due to their stability, nonvolatility and adjustable miscibility and polarity [15]. The solvent properties of ILs are mainly determined by the ability of the salt to act as a hydrogen bond donor and/or acceptor and the degree of localization of the charges on the anions [1,21]. Charge distribution on the anions, H-bonding ability, polarity, dispersive interactions are the main factors that inuence the physical properties of ILs [22]. For example, imidazolium-based ILs are highly ordered hydrogen-bonded solvents and they have strong effects on chemical reactions and processes.

3.2. Polarity of ILs Polarity of chemicals is commonly used to classify the solvents. The terms used as polar, nonpolar and apolar are generally related to the values of dielectric constants, dipole moments, polarizabilities. If a solvents has the ability to dissolve and stabilize dipolar or charged solutes, it is dened as a polar solvent. Under this simple denition, ILs are highly polar solvents, but it is not completely true to make such strict conclusions since there ILs can be designed in a vast range. The existences of polar and nonpolar domains, believed to be associated with the unique amphiphilic solvent properties of ILs, are found in the structures of PF6 and BF4 salts [23]. Since polarity is the simplest indicator of solvent strength, researchers compared polarities of ILs and conventional solvents: Carmichael and Seddon [24] showed that 1-alkyl-3-methylimidazolium ILs with anions [PF6 ], [BF4 ], [(CF3 SO2 )2 N], and [NO3 ] are in the same polarity region as 2-aminoethanol and lower than alcohols such as methanol, ethanol and butanol. Aki et al. [25] indicated that [bmim][PF6 ], [C8 mim][PF6 ], [bmim][NO3 ] and [N-bupy][BF4 ] are more polar than acetonitrile and less polar than methanol and these ILs are expected to be at least partially miscible with water. ILs based on [PF6 ] anion is preferred as solvents in most extraction application to form biphasic systems due to their immiscibility with water. 3.3. Toxicology of ILs The green character of ILs has been usually related with their negligible vapor pressure; however their toxicology data have been very limited until now. Several authors [2629] already mentioned this lack of toxicological data in the literature [30]. Although ILs will not evaporate and thus will not cause air pollution, it does not mean that they will not harm the environment if they enter. Most of ILs are water soluble and they may enter the aquatic environment by accidental spills or efuents. The most commonly used ILs [bmim][PF6 ] and [bmim][BF4 ] are known to decompose in the presence of water and as a result hydrouoric and phosphoric acids are formed [31]. Therefore, both toxicity and ecotoxicity information which provide metabolism and degradability of ILs are also required to label them as green solvents or investigate their environmental impact. The ecotoxicological studies performed to understand the effects of different ILs on enzymatic activities, cells and microorganisms are utilized to obtain LC50 levels (lethal concentration). Decreasing LC50 values indicate higher toxicities according to the toxicity classes of Hodge and Sterner scale (1956) [32]. This scale indicates that the LC50 value (in terms of mg/L) of 10 or less shows that the chemical is extremely toxic, LC50 value between 10 and 100 shows that chemical is highly toxic, LC50 value between 100 and 1000 shows that chemical is slightly toxic, and nally LC50 value between 1000 and 10,000 means that chemical is practically nontoxic. The impact of ILs on aquatic ecosystems is highly important since some of ILs have a high solubility in water. Maginn [33] provided the LC50 levels for two imidazolium-based ILs with

S. Keskin et al. / J. of Supercritical Fluids 43 (2007) 150180 Table 2 LC50 values for certain solvent [33] Compound [bmim][PF6 ] [bmim][BF4 ] Acetone Dichloromethane Toluene Benzene Chlorobenzene Phenol Ammonia Chlorine LC50 (mg/L) 250300 225275 30,642 310 60313 203 586 5 0.534.94 0.028

155

The environmental persistence [34,40] of commonly used ILs, along with their possible toxicity should be taken into account. There are still very few results about the (eco)toxicological effect of ILs and they can be evaluated more satisfactorily as green solvents or not after more data on the subject will be provided. The possible toxic and non-biodegradable nature of the existing ILs also led to the development of new types of nontoxic and biodegradable ILs [4046]. 3.4. Air and moisture stability of ILs The stability of ILs is crucial for optimum performance. Many of ILs are both air and moisture stable, some are even hydrophobic. On the other hand, most imidazolium and ammonium salts are hydrophilic and if they are used in open vessels, hydration will certainly occur. The hydrophobicity of an IL increases with increasing length of the alkyl chain [25]. Despite their wide spread usage, ILs containing PF6 and BF4 have been reported to decompose in the presence of water, giving off HF. Wasserscheid et al. [47] pointed out that ILs containing halogen anions generally show poor stability in water, and also give off toxic and corrosive species such as HF or HCl. Therefore, they suggest the use of halogen-free and relatively hydrolysis-stable anions such as octylsulfate-compounds. The degree to which this hydration is a problem depends on the application. For instance, small amounts of highly reactive species which are used as catalysts may be deactivated by even very small amounts of water. For this kind of application, ILs must be handled under an inert atmosphere. Moreover, the solutes used may be sensitive for air or moisture, thus an inert atmosphere is required for the ILsolute systems. The interaction between water and ILs and their degree of hydroscopic character are strongly dependent on anions. The amount of absorbed water is highest in the BF4 and lowest in PF6 [48]. However, Tf2 N is much more stable in the presence of water as well as having the advantage of an increased hydrophobic character. ILs immiscible with water tend to absorb water from the atmosphere. The infra-red (IR) studies of Cammarata et al. [31] demonstrated that the water molecules absorbed from the air are mostly present in the free state, bonded via H-bonding with the PF6 and BF4 anions. The presence of water may have dramatic effect on IL reactivity. Since water is present in all ILs, they are usually utilized after a moderate drying process. The new ILs synthesized are more stable than the old halogenoaluminate systems. Certain ILs incorporating 1-3dialkyl imidazolium cations are generally more resistant than traditional solvents under harsh process conditions, such as those occurring in oxidation, photolysis and radiation processes [10]. 4. Structure and synthesis of ILs There are a great number of different cation and anion combinations to synthesize IL. Different types of ILs give an opportunity to modify the physical and chemical properties of the IL. The most widely used cations are imidazolium, pyridinium, phosphonium and ammonium. The properties of ILs

Daphnia magna, common fresh water crustaceans. Due to the reason that D. magna are lter feeders at the base of the aquatic food chain, their responses to ILs are essential to understand how these new solvents may impact an environmental ecosystem. As ILs, 1-n-butyl 3-methylimidazolium cation with PF6 and BF6 anions are used and the results are tabulated in Table 2: These two ILs are as toxic to Daphnia as benzene and even far more toxic than acetone, but much less toxic than ammonia, chlorine, phenol, etc. Wells and Coombe [34] also provided the results of freshwater ecotoxicity tests of some common ILs with imidazolium, ammonium, phosphonium and pyridinium cations on invertebrate D. magna and the green alga Pseudokirchneriella subcapitata (formerly known as Selenastrumcapricornutum). The results were reported using medium effective concentration (EC50 ) values. The toxicity values of the most toxic IL were four orders of magnitude more than the least toxic IL. There was a relation between the order of toxicity and alkyl side chain length of the cation. For alkyl methylimidazolium ILs with C4 side chain constituents showed moderate toxicity, whereas the C12 , C16 , and C18 species were very highly toxic to both organisms under investigation. Pyridinium, phosphonium, and ammonium species with C4 side chain constituents had also only moderate toxicity, whereas C6 and longer side chains showed signicant increases in toxicity. It was shown that the least toxic ionic liquids ecotoxicity were comparable to hydrocarbons such as toluene and xylene. The most toxic ionic liquids are many orders of magnitude more acutely ecotoxic than organic solvents such as methanol, tert-butyl methyl ether, acetonitrile, and dichloromethane. The authors also emphasized that simple acute ecotoxicity measurements did not enough to fully characterize the full impact of a solvent released to the environment but were only part of the environmental impact assessment. There were also some studies on investigation of toxicity of ILs performed on animals such as the nematode model organism (Caenorhabditis elegans) [35], freshwater pulmonate snails (Physa acuta) [36], Fischer 344 rats [37] and zebra sh (Danio rerio) [38]. One of the most important points that must be taken into account during the toxicological study of the ILs is to pay attention to the purity of the IL studied. Therefore, different authors [35,39] attached signicant importance to proper analyzing techniques [30]. ILs are introduced under the concept of green chemistry in all research papers due to their nonvolatile nature.

156

S. Keskin et al. / J. of Supercritical Fluids 43 (2007) 150180

Fluorinated anions tend to be expensive and in response to cost and safety concerns new ILs with non-uorous ions have been introduced. In the synthesis of these ILs, anions are derived from inexpensive bulk chemicals. Alkylsulfate anions are the most popular non-uorous anions due to their nontoxic and biodegradable structures. The rst commercially available IL for which toxicology data are available contains alkylsulfate anion (methosulfate) [50]. 4.2. Cations
Fig. 3. Most commonly used cation structures and possible anion types [50].

are determined by mutual t of cation and anion, size, geometry, and charge distribution. Among the similar class of salts, small changes in types of ions inuence the physico-chemical properties. The overall properties of ILs result from the composite properties of the cations and anions and include those that are superacidic, basic, hydrophilic, water miscible, water immiscible and hydrophobic. Usually, the anion controls the water miscibility, but the cation also has an inuence on the hydrophobicity or hydrogen bonding ability [49]. The structures of most commonly used cations and some possible anion types are tabulated in Fig. 3 [50]. 4.1. Anions The properties of ILs are determined by the anion type. The introduction of different anions results in an increasing number of alternative ILs with various properties. There are two different types of IL anions: ILs containing uorous anions such as PF6 , BF4 , CF3 SO3 , (CF3 SO3 )2 N and ILs with non-uorous anions such as AlCl4 . In designing ILs, uorous anions are usually used. The most popular anions consist of chloride, nitrate, acetate, hexauorophosphate and tetrauoroborate [13]. The most widely investigated ILs are the ones with anions PF6 and BF4 . Especially PF6 is the most prominent anion used in IL research. Since the anion chemistry has a large effect on the properties of IL, although the cations are the same, there are signicant differences between ILs with different anions. For example IL with 1-n-butyl-3-methylimidazolium cation and PF6 anion is immiscible with water, whereas IL with same cation and BF4 anion is water soluble. This example represents the designer solvent property of ILs: different ion pairs determine physical and chemical properties of the liquid. By changing the anion the hydrophobicity, viscosity, density and solvation of the IL system may be changed [8]. Although PF6 and BF4 are the two anion types that are utilized in most of IL applications, they have an important disadvantage: these two anions may decompose when heated in the presence of water and liberate HF. After the researchers realized the production of HF in the presence of water, the bonding style of anion was altered and uorous anions inert to hydrolysis were used. The uorine of the anion is bonded to carbon and CF bond becomes inert to hydrolysis. In this way, ILs such as CF3 SO3 and (CF3 SO3 )2 N are produced [50].

The cation of IL is generally a bulk organic structure with low symmetry. Most ILs are based on ammonium, sulfonium, phosphonium, imidazolium, pyridinium, picolinium, pyrrolidinium, thiazolium, oxazolium and pyrazolium cations. The research mainly focuses on RTILs composed of asymmetric N,N-dialkylimidazolium cations associated with a variety of anions. 1-n-butyl-3-methylimidazolium and 1-n-ethyl-3methylimidazolium are the most investigated structures of this class. Chiappe and Pieraccini [18] indicated that the melting points of the most ILs are uncertain since ILs undergo considerable supercooling. Therefore, by examining the properties of a series of imidazolium cation based ILs, it has been concluded that as the size and asymmetry of the cation increases, the melting point decreases. Further, an increase in the branching on the alkyl chain increases the melting point. The melting point of ILs is essential because it represents the lower limit of the liquidity and with thermal stability it denes the interval of temperatures within which it is possible to use ILs as solvents [18]. 4.3. Synthesis There are three basic methods to synthesize ILs: metathesis reactions, acidbase neutralization, direct combination [1]. Many alkylammonium halides are commercially available; they can also be prepared simply by the metathesis reaction of the appropriate halogenoalkane and amine. Pyridinium and imidazolium halides are also synthesized by metathesis reaction. On the other hand, monoalkyllammonium nitrate salts are best prepared by the neutralization of aqueous solutions of the amine with nitric acid. After neutralization reactions, ILs are processed under vacuum to remove the excess water [1]. Tetraalkylammonium sulfonates are also prepared by mixing sufonic acid and tetraalkylammonium hydroxide [51]. In order to obtain pure IL, products are dissolved in an organic solvent such as acetonitrile and treated with activated carbon, and the organic solvent is removed under vacuum. The nal method for the synthesis of ILs is the direct combination of halide salt with a metal halide. Halogenoaluminate and chlorocuprate ILs are prepared by this method. The synthesis methods of ILs have been given in numerous articles [5256]. 5. Major applications suggested for ILs The research areas on ILs are growing very rapidly and the potential application areas of ILs are numerous. The unique

S. Keskin et al. / J. of Supercritical Fluids 43 (2007) 150180

157

Fig. 4. Major application areas of ILs.

chemical and physical properties of ILs bring about several application areas including reaction and synthesis media. The application areas of ILs can be expressed as solvents for organic, organometallic synthesis and catalysis; electrolytes in electrochemistry, in fuel and solar cells; lubricants; stationary phases for chromatography; matrices for mass spectrometry; supports for the immobilization of enzymes; in separation technologies; as liquid crystals; templates for synthesis nano-materials and materials for tissue preservation; in preparation of polymergel catalytic membranes; in generation of high conductivity materials [7]. Fig. 4 represents the major applications suggested for ILs and these essential applications are discussed in detail below. 5.1. Solvent replacement A majority of common solvents have potential health hazards although they are extensively utilized. For example, approximately half of 189 hazardous air pollutants regulated by Clean Air Act Amendment of U.S. (1990) are VOCs including solvents such as dichloromethane [57]. The VOCs are the workhorses of industrial chemistry in the pharmaceutical and petrochemical areas. The use of VOCs by these industries can be assessed using the Sheldon E-factor. This factor is responsible to measure process by-products as a proportion of production on the mass basis. Researchers investigated how VOC use is distributed across the chemical industry and found that the value of E-factor is between 25 and 100 for pharmaceuticals industries with a production of 10 to 103 t/year although oil rening industries with a production of 106 to 108 t/year have an E-factor of 0.1 [9] (adapted from [58]). These values suggest that pharmaceuticals industries use inefcient and dirty processes although on smaller scale as compared to the oil rening industries. The oil and bulk chemicals industries which are commonly regarded as dirty are apparently remarkably waste conscious when the E-factors are analyzed [9]. As the introduction of cleaner technologies has become a major concern throughout both industry and academia, the search for the alternative solvents has become a high priority. Therefore, environmentally friendly ILs can easily replace the hazardous VOCs in large scale to reduce E-factors. ILs are able to dissolve a variety of solutes. They can be used instead of traditional solvents in liquidliquid extractions where hydrophobic molecules such as simple benzene derivatives will partition to the IL phase. Huddleston et al. [17] showed that [bmim][PF6 ] could be used to extract aromatic compounds from water. Fadeev and Meagher [59] demonstrated that two imidazolium ILs with PF6 anion could be used for the extraction of butanol from aqueous fermentation broths. Selvan et al. [60]

used ILs for the extraction of aromatics from aromatic/alkane mixtures, whereas Letcher et al. [61] used ILs for the extraction of alcohols from alcohol/alkane mixtures. Moreover, binary temperaturecomposition curves of ILs with alcohols, alkanes, aromatics and water; ternary temperaturecomposition curves of ILs with alcohols and water; solubilities of some organics and water in ILs are all investigated by various groups to completely benet from the solvent properties of ILs [6264]. Arce et al. [65] studied essential oil terpenless by extraction using organic solvents or ILs. Citrus essential oil is simulated as a mixture of limonene and linalool and 2-butene-1,4-diol and ethylene glycol are used as solvents. They choose 1ethyl-3-methylimidazolium methanesulfonate as the IL and liquidliquid equilibria data for the ternary systems are reported. Arce et al. [65] concluded that IL presents the highest selectivity but close to the other organic solvents and they reported that the results for solute distribution ratio depend on the concentration range of extraction. 5.2. Purication of gases Reliable information on the solubility of gases in ILs is needed for the design and operation of any possible processes involving IL. Processes using ILs to purify gas streams were developed after solubilities of various gases in ILs were reported by some researchers [6668]. These experimental studies show that some gases, especially CO2 is highly soluble in ILs. The simulations performed explain that the anion of the IL is responsible for high gas solubility. With this property ILs, can be replaced as solvents in reactions involving gaseous species. Anthony et al. [69] investigated solubility of nine different gases up to 13 bar: carbon dioxide, ethylene, ethane, methane, argon, oxygen, carbon monoxide, hydrogen, and nitrogen in [bmim][PF6 ]. These gases were chosen for several reasons: CO2 solubility is important due to the possibility of using scCO2 to extract solutes from ILs; ethylene, hydrogen, carbon monoxide, and oxygen are reactants in several types of reactions studied in IL such as hydroformylations, hydrogenations, and oxidations. Due to the nonvolatile nature of IL, the gas solubilities in IL were measured using a gravimetric technique, usually with a microbalance. The study of Anthony et al. [69] showed that CO2 has the highest solubility and strongest interactions with [bmim][PF6 ], followed by ethylene and ethane. Argon and oxygen had very low solubilities and immeasurably weak interactions. Fig. 5 demonstrates the solubility of various gases (CO2 , C2 H4 , C2 H6 , CH4 , Ar, O2 ) in [bmim][PF6 ] at 25 C and at different pressures. Except for CO2 , all gases remained in the Henrys law regime up to 13 bar. However, CO2 showed some nonlinearity, indicating some degree of saturation. Henrys constants of these gases in various organic solvents and in [bmim][PF6 ] were compared and the results showed that the gases that are less soluble in the IL are less soluble in the other solvents as well. However, CO2 is more soluble in the IL than in the other solvents. The relatively high solubility of CO2 was explained as a result of its large quadrapole moment. The solubility of CO2 in [bmim][PF6 ] at different temperatures is demonstrated in Fig. 6.

158

S. Keskin et al. / J. of Supercritical Fluids 43 (2007) 150180

Fig. 5. Solubility of various gases in [bmim][PF6 ] at 25 C. (Reprinted with permission from [69]. Copyright 2002 American Chemical Society)

Camper et al. [66] measured the solubility of CO2 and C2 H4 in [bmim][PF6 ], [emim][Tf2 N], [emim][CF3 SO3 ] (ethylmethylimidazolium triuoromethanesulfone), [emim][dca] (ethylmethylimidazolium dicyanamide) and [thtdp][Cl] (trihexyltetradecylphosphonium chloride) to demonstrate that the regular solution theory can be used to model the gas solubilities in RTILs at low pressures and studied the effects on pressure and the temperature on the solubility of gases in RTILs. The previous works; Blanchard et al. [70] and Anthony et al. [71] related that the solubility of the gases in ILs to the intermolecular interactions between the anion of the IL and the gas. On the other hand, Camper et al. [66] indicated that at low pressures, the solubility of CO2 and C2 H4 may be explained using the regular solution theory without considering the intermolecular interactions between the anion of the IL and the gas. At higher pressures, regular solution theory is limited and Camper et al. [66] attributed this limitation to the dominant entropic effects. Recently, the results of solubility of hydrogen in [bmim][PF6 ] was presented for temperatures from 313 to 373 K and pressures up to 9 MPa. The results demonstrated that the solubility of hydrogen in [bmim][PF6 ] is low and increases slightly with

temperature [72]. Since ILs can dissolve certain gaseous species, they may be used in conventional gas absorption applications. The nonvolatility of ILs prevent any cross contamination of the gas stream by the solvent during the process. Moreover, regeneration of the solvent may be performed easily by a simple ash or distillation to remove the gas from the solvent without any cross contamination. The other advantages of ILs as separating agents are no solvent loss and no air pollution. Currently, researchers are interested in examining the potential of ILs for the separation of CO2 from ue gases emitted from fossilfuel combustion operations [73]. ILs may also be utilized as supported liquid membranes. In conventional membranes, gas dissolves in liquid but then the liquid in which the gas dissolved evaporates rendering the membrane useless [50]. Due to the nonvolatility of ILs, they can be immobilized on a support and used in supported liquid membranes. ILs are also used for storage and delivery of hazardous specialty gases such as phosphine (PH3 ), arsine (AsH3 ) and stibine (SbH3 ). GASGUARD Sub-Atmospheric Systems supply the major ion implant gases: AsH3 , boron triuoride (BF3 ), enrich boron triuoride (11 BF3 ) and PH3 sub-atmospherically [74]. The system is combined with gas supply technologies for the delivery of the gases when needed. In the complexed gas technology, the desired gases (BF3 and PH3 ) are chemically bond to ILs subatmospherically, then pulling the vacuum on the ILgas complex provides the mechanism to evolve high purity gas, similar to desorbing a gas from active carbon. 5.3. Homogenous and heterogeneous catalysis One of the most important targets of modern chemistry is to combine the advantages of both homogenous and heterogeneous catalysis [75]. Greater selectivity is generally observed in homogenous catalysis compared to its heterogeneous counterparts, but separation of the catalyst from the product stream or from the extract stream causes a problem [8]. ILs offer the advantages of both homogenous and heterogeneous catalysts with their two main characteristics: A selected IL may be immiscible with the reactants and products, but on the other hand the IL may also dissolve the catalysts. ILs combine the advantages of a solid for immobilizing the catalyst, and the advantages of a liquid for allowing the catalyst to move freely [76]. Brennecke and Maginn [8] indicated that the ionic nature of the IL also gives an opportunity to control reaction chemistry, either by participating in the reaction or stabilizing the highly polar or ionic transition states. ILs have an active role in chemical reactions and catalysis. Some of the examples where ILs are utilized are: reactions of aromatic rings; clean polymerization [77]; Friedel Crafts alkylation [78]; reduction of aromatic rings [79]; carbonylation [80]; halogenation [81]; oxidation [82]; nitration [83]; sulfonation [84]; solvents for transition metal catalysis; immobilization of charged cationic transition metal catalysis in IL phase without need for special ligands [85]; in situ catalysis directly in IL rather than aqueous catalysis followed by extraction of products from solution: this process eliminates washing steps, minimizes losses of catalysis and enhances purity of the products [86].

Fig. 6. Solubility of CO2 in [bmim][PF6 ] at different temperatures. (Reprinted with permission from [69]. Copyright 2002 American Chemical Society)

S. Keskin et al. / J. of Supercritical Fluids 43 (2007) 150180

159

Many applications of ILs in catalytic reactions can be found in various articles in the literature [1,12,85,8789]. Holbrey and Seddon [90] described many of the catalytic processes which use low temperature ILs as reaction media and indicated that the classical transition metal catalyzed hydrogenation, hydroformylation, isomerization, dimerization and coupling reactions can be performed in IL solvents. In their review, Holbrey and Seddon [90] concluded that ILs may be used as effective solvents and catalysts for clean chemical reactions instead of the volatile organic solvents. Brennecke and Maginn [8], concluded that ILs have been used successfully for hydrogenations, hydroformylations, isomerizations, dimerizations, alkylations, Diels-Alder reactions and Heck and Suzuki coupling reactions, and in general researchers have concluded that the reaction rates and selectivities are as good or better in ILs than in conventional organic solvents. The catalytic hydrogenation of cyclohexene using rhodium-based homogenous catalysts [91] and hydrogenation of olens using ruthenium and cobalt-based homogenous catalyst [92] in various ILs are studied and the results indicated that there is a certain increase in the reaction rates and selectivity compared to the other normal liquid solvents. Lagrost et al. [12] used immidazolium and ammoniumbased ILs ([emim][NTf2 ], [bmim][NTf2 ], [bmim][PF6 ], [(C8 H17 )3 NCH3 ][NTf2 ]) as reaction media for different types of electrochemical reactions and investigated the oxidation of organic molecules (anthracene, naphthalene, durene, 1,4dithiafulvene and veratrole) in ILs. Their results suggest that the nature of investigated mechanisms is almost unchanged in ILs as compared with the conventional organic media although the structure of molecular solvents and ILs are expected to be quite different. Lagrost et al. [12] also concluded that the diffusion coefcient of the organic compounds are about 100 times smaller than those in conventional media as expected from the lower viscosity of RTILs versus organic solvents. The positive results of this study demonstrated that ILs can be used as a new media for organic electrochemistry [12]. 5.4. Biological reactions media ILs are used in biological reactions such as the synthesis of pharmaceuticals due to the stability of enzymes in ILs, and in separation processes such as the extraction of amino acids [15]. IL biphasic systems are used to separate many biologically important molecules such as carbohydrates, organic acids including lactic acid [93,94]. Carbohydrates are renewable and inexpensive sources of energy and raw material for the chemical industry. The underivatized carbohydrates are not soluble in most of the conventional solvents although they are soluble in water. Their insolubility in most solvents prevents the transformation of carbohydrates. Therefore, the ability of ILs to dissolve carbohydrates enables transformation possibilities [15,93,95,96]. Lau et al. [95] studied the alcoholysis, ammoniolysis, and perhydrolysis reactions by Candida antarctica lipase catalysis using the [bmim[[PF6 ] and [bmim[[BF4 ] as reaction media.

Reaction rates were generally comparable with, or better than, those observed in organic media. Park and Kazlauskas et al. [96] studied the acetylation of 1-phenylethanol catalyzed by lipase from Pseudomonas cepacia (PCL) in several ILs and the reaction was as fast and as enantioselective in ILs as in toluene. They also investigated the acetylation of glucose catalyzed by lipase B from C. antarctica (CALB) and found that the transformation was more regioselective in ionic liquids because glucose is up to one hundred times more soluble in ionic liquids. Liu et al. [93] stated that carbohydrates are only sparingly soluble in common organic solvents as well as in weakly coordinating ionic liquids, such as [bmim][BF4 ]. They found that ILs that contain the dicyanamide anion could dissolve approx. 200 g L1 of glucose, sucrose, lactose and cyclodextrin and the esterication of sucrose with dodecanoic acid in [bmim][dca] could be performed with CALB. Swatloski et al. [97] showed that ILs can also be used as non-derivatizing solvents for cellulose, the most abundant biorenewable material. Cellulose, which is insoluble in water and in most of the common organic solvents, has many derivitized products in many applications of the ber, paper, and polymer industries. ILs incorporating anions which are strong hydrogen bond acceptors are most effective solvents for cellulose, whereas ILs containing non-coordinating anions including PF6 and BF4 are not effective. Furthermore, Przybysz et al. [98] examined the inuence of ILs on a cellulose product, paper and found that the wettability of paper is improved, whereas the strength decreased as a result of weakening of cellulose hydrogen bonds. Finally, Pfruender et al. [99] tested the water immiscible ILs namely; [bmim][PF6 ], [bmim][Tf2 N] and [oma][Tf2 N] (methyltrioctylammonium bistriuoromethanesulfonylimide) for their biocompatibility towards Escherichia coli and Saccharomyces cerevisiae. The results of this study showed that these water immiscible ILs do not damage microbial cells and therefore one can utilize these water immiscible ILs as substrate reservoirs and in situ product extracting agents for biphasic whole cell biocatalytic processes. Generally, toxic organic solvents have been used as substrate reservoirs and with this study it is shown that water immiscible ILs may be used as biocompatible solvents for microbial biotransformations. The experimental results demonstrated that there is an increase of chemical yield from <50% to 8090% in simple batch processes and (R)-1-(4chlorophenyl) ethanol was produced at a higher initial reaction rate in the biphasic system (>50 M s1 L1 ) compared to the aqueous system [99]. Although ILs are known by their highly viscous characteristics, good mass transfer rates were obtained in their study. 5.5. Removing of metal ions Dai et al. [100] studied the effects of ILs (with PF6 and Tf2 N anions) on improving the ability of crown ethers to remove metal ions from aqueous solutions. Strontium nitrate, a ssion product for which there is no available extraction technique for its removal from radioactive waste sites, was used in this study.

160

S. Keskin et al. / J. of Supercritical Fluids 43 (2007) 150180

Visser et al. [57] designed and synthesized several ILs to remove cadmium and mercury from contaminated water. The hydrophobic ILs come into contact with contaminated water and they snatch the metal ions out of water. Task-specic ionic liquids (TSIL) concept is introduced in order to synthesize ILs with desired properties to extract metal ions. Visser et al. [57] produced TSIL cations by appending different functional groups (namely thiother, urea and thiourea) to imidazolium cations. These IL cations can be considered as a new IL class, or a novel class of IL extractants. Synthesized TSIL cations were combined with PF6 anion and used alone or in a mixture with [bmim][PF6 ]. The results of the study gave signicant distribution ratios for mercury and cadmium in liquidliquid separations and minimized the reliance on traditional organic solvents for this process. Davis [101] gave a detailed analysis and related information on TSILs. In traditional solvent extraction technologies, adding extractants that reside quantitatively in the extracting phase increases the metal ion partitioning to the more hydrophobic phase. The added extractant molecules dehydrate the metal ions and provide a more hydrophobic environment enabling their transport to the extracting phase [50]. In TSILs, attaching a metal ion coordinating group directly to the imidazolium cation makes the extractant an integral part of the hydrophobic phase and in this way the chance for IL loss to the aqueous phase is reduced. Therefore, TSILs act as the hydrophobic solvent and the extractant at the same time [50]. However, the cost of TSIL is generally high. In order to eliminate this drawback, TSILs may be added to the mixtures of less expensive ILs. Furthermore, Davis [101] and Zhao et al. [15] stated in their recent reviews that TSILs are not limited to extraction processes; they can be also used as versatile solvents in organic catalysts, solid phase synthesis and even in production of liquid Teon. The extraction of radioactive metals (lanthanides and actinides) has particular industrial signicance among IL extraction of metal ions for the handling of nuclear materials [15]. The behaviors of uranium species in various ILs were investigated in early studies [102105]. Recently, researchers have focused on the fundamental understanding of ILs in nuclear chemistry such as radiochemical stability of ILs [106]. Recently, Nakashima et al. [107] examined the feasibility of extracting of rare earth metals into ILs from aqueous solutions and stripping of metal ions from ILs into an aqueous phase by complexing agents. They successfully accomplishing to recycle the extracting IL phase. In this study, octyl(phenyl)-N,Ndiisobutylcarbamoylmethyl phosphine oxide (CMPO) dissolved in [bmim][PF6 ] showed an extremely high extraction ability and selectivity of metal ions as compared to in an ordinary diluent, n-dodecane. The results of this study indicate that ILs are a promising medium for actinide and ssion product separation. In the literature, various studies were performed to extract metal ions using ILs [108113]. Different metal ions including alkali, alkaline earth metals, heavy metals and radioactive metals are researched by using different ILs. Generally, the side chain of the IL on the cation is varied and the effect of structure of the

IL on the extraction efciency of the metal ions is investigated. The side chain of the cation inuences the hydrophobic character of the IL and thus the partition coefcient of the metal ions is affected. Visser et al. [108] extracted Na+ , Cs+ using [Cn mim][PF6 ] (n = 4, 6, 8); Chun et al. [109] investigated extraction of other alkali metals such as Li+ , K+ , Rb+ using [Cn mim][PF6 ] (n = 49); Luo et al. [112,113] studied the extraction of Na+ , K+ , Cs+ ions using [Cn mim][Tf2 N] (n = 2, 4, 6, 8). Not only the alkali metals but also extraction of alkaline earth metals were studied by various groups: Visser et al. [108] removed Sr2+ using [Cn mim][PF6 ] (n = 4,6,8); Bartsch et al. [110] studied the removal of Mg2+ , Ca2+ , Sr2+ , Ba2+ ; Luo et al. [112,113] utilized [Cn mim][Tf2 N] (n = 2, 4, 6, 8) to extract Sr2+ . The extraction of heavy and radioactive metals such as Cu2+ , Ag+ , Pb2+ , Zn2+ , Cd2+ , Hg2+ were studied by using [Cn mim][PF6 ] (n = 49) and TSILs [110,111,114]. 6. Challenges of ILs The unique properties of ILs and the ability to design their properties by choice of anion, cation and substituents create many more processing options, alternative to the ones with conventional solvents. However, high cost, lack of physical property and toxicity data restrict the advantageous use of ILs as process chemicals and processing aids at the present. The challenges in the use of ILs must be also addressed as well as their advantages. The major challenge is the cost. A kilogram of IL costed about 30,000-fold greater than a common organic solvent such as acetone. Renner [9] reported that this cost could be reduced to approximately 1000-fold greater depending on the composition of IL and the scale of production. Wagner and Uerdingen [115] anticipated that the price of cation systems based on imidazole will be in the range of D 50100 kg1 , if larger quantities of ILs are produced. The price can be lowered even below D 25 kg1 if ILs are prepared with cheaper cation sources on a ton scale. Another estimation was done by Wassersheid and Haumann [116]. They expected that for bulk ionic liquids choosing proper (relatively cheap) cations and anions lead to prices approximately D 30 l1 for production rates of multi-ton. Moreover, scientists emphasized that the price of the ILs may look still discouraging however, the essential factor is the price to performance ratio. If the performance of an IL is extremely higher than that of the material (solvent) it aims to replace, less amounts of the IL may be needed for a given specic job [2], thus totally or partially overcoming the price disadvantage. The second problem is associated with the manufacturing method of ILs. In manufacturing ILs environmental issues also need to be tackled since some VOCs are used to manufacture ILs. Recently, some advanced methods have been developed in the solventless syntheses of ILs. For example, 1-alkyl3-methylimidazolium halides have been synthesized in open containers in a microwave oven without any VOCs by Varma and Namboodiri at the Environmental Protection Agency of U.S., 2001 [9].

S. Keskin et al. / J. of Supercritical Fluids 43 (2007) 150180

161

Researchers need to nd alternative ways to recycle ILs due to the reason that many processes for cleaning up ILs involve washing with water or VOCs which creates another waste stream. This problem has been solved by adopting supercritical extraction technologies to recover the dissolved organic compounds from ILs or using membrane separation processes. However, there are other solid matrixes which adsorb some part of the ILs. Thus, a second rinse would be required and this would create an aqueous waste stream that contains ILs. Despite this disadvantage there may be some cleaning applications where ILs would be attractive [8]. Incomplete physico-chemical data are another challenge for the application of ILs. At the present most available data are focused on bulk physical properties such as viscosity, density and phase transitions. Relatively little is known about the microscopic physical properties of ILs. After these properties are investigated properly and the inuence of ILs on chemical reaction rates is found, new ILs with precisely tailored properties can be synthesized. It is extremely important to obtain reliable thermophysical data and transport properties of ILs in order to make them available for many applications and to design IL-based processes efciently. Harris et al. [117] measured the viscosities of two members of one of the most commonly studied IL groups, that are based on imidazolium cations; [omim][PF6 ] and [omim][BF4 ] between 0 and 80 C and at pressures to 176 MPa ([omim][PF6 ]) and 224 MPa ([omim][BF4 ]) with a falling body viscometer and densities between 0 and 90 C at atmospheric pressure. The bulk physical properties of most widely used ILs at wider temperature and pressure ranges are essential. Another barrier to the large-scale application of ILs arises from their high viscosities. The viscosities of ILs are higher than most organic solvents and water, usually similar to viscosity of oils. This high viscosity may be responsible to produce a reduction in the rate of many organic reactions and even a reduction in the diffusion rates of species. Also, handling of ILs with high viscosities is difcult however; increasing temperature, changing anioncation combinations may yield ILs with lower viscosities. To overcome mass transfer limitations in gasIL systems resulting from high viscosity reactions using ILs may be run at high pressures and in efcient gasliquid contacting equipment. In chemical processing, pharmaceuticals, ne chemicals, petroleum rening, metal rening, polymer processing, pulp and paper, and textiles where a nonvolatile liquid with a wide liquidus range could work better, ILs are the best choice however, the challenges of turning ILs into useful and environmentally benign uids must be overcome. 7. ILs and scCO2 systems Green chemistry, also known as sustainable chemistry, describes the search for reducing or even eliminating the use of substances in the production of chemical products and reactions which are hazardous to human health and environment. The goal of green chemistry is to create a cleaner and more

sustainable chemistry and it has received more and more attention in recent years. Green chemistry searches for alternative, environmentally friendly reaction media as compared to the traditional organic solvents and at the same time aims at increased reaction rates, lower reaction temperatures as well higher selectivities. The ideal situation for a safe and green chemical process is using no solvent, however most of the chemical processes depend on solvents. Some of these solvents are soluble in water and therefore they must be stripped from water before it leaves the process not only for ecological but also for economic reasons. Solvents must be recovered for recycle and reuse for an economically viable process. Water, peruorinated hydrocarbons and supercritical uids (SCFs) are alternative solvents which may be used in green chemistry. Among these, the most promising elements of green chemistry are ILs and scCO2 . The low volatility of ILs is the key property that makes them green solvents. However, this advantage also causes a problem for product separation and recovery [15]. Several techniques for solute recovery from ILs exist: volatile products can be extracted from IL by distillation or simply by evaporation. However, nonvolatile or thermo-sensitive products cannot be separated from ILs with these methods. ILs exhibiting immiscibility with water can be extracted with water to separate water-soluble solutes from IL into the aqueous phase; but this method is not suitable for hydrophilic ILs [17]. Of course, organic solvents such as hexane and toluene may be effective to recover the products from IL but this approach obviously compromises the ultimate goal of green technologies [15]. Furthermore, the cross contamination between the phases presents another problem. Finally, another green solvent is discovered which solves all the problems and recovers various kind of solutes from ILs without cross contamination: supercritical uids (SCFs). SCFs are compounds which are above their critical temperature and pressure and they can be manipulated from gas like to liquid like densities due to their unusual properties near the critical point. They are commercially viable solvents in several applications such as dry cleaning and polymer impregnation. scCO2 is the most widely used SCF as a result of nontoxic and non-ammable characteristics. scCO2 has low critical temperature and pressure and it is not expensive. The advantages of using SCFs as extraction medium include low cost, nontoxic nature, recoverability and ease of separation from the products. SCFs have been adapted for product recovery from ILs and supercritical uid extraction (SCFE) is shown to be a viable technique with the additional benets of environmental sustainability and pure product recovery [118]. Among the SCFs, an inexpensive and readily available one, scCO2 has become a partner of IL and two environmentally benign solvents are utilized together in several applications. The volatile and nonpolar scCO2 forms different two-phase systems with nonvolatile and polar ILs. The product recovery process with these systems is based on the principle that scCO2 is soluble in ILs, but ILs are not soluble in scCO2 [70]. Since most of the organic compounds are soluble in scCO2 , with the high solubility of scCO2 in ILs, these products are transferred from the IL to the supercritical phase.

162 Table 3 ILgas systems phase behaviors System [bmim][PF6 ]CO2 [C8 -mim][PF6 ]CO2 [C8 -mim][BF4 ]CO2 [bmim][NO3 ]CO2 [emim][EtSO4 ]CO2 [N-bupy][BF4 ]CO2 [bmim][PF6 ]CO2 [bmim][PF6 ]C2 H4 [bmim][PF6 ]C2 H6 [bmim][PF6 ]CH4 [bmim][PF6 ]Ar [bmim][PF6 ]O2 [bmim][PF6 ]CO [bmim][PF6 ]N2 [bmim][PF6 ]H2 [bmim][PF6 ]CO2 [bmim][BF4]CO2 [bmim][BF4]O2 Temperature 40, 50, and 60 C

S. Keskin et al. / J. of Supercritical Fluids 43 (2007) 150180

Pressure Up to 93 bar

Findings As pressure increases, solubility of CO2 in the IL increases Solubility of CO2 in IL-rich phase decreases with temperature CO2 solubility depends on the nature of the anion and cation The solubility of CO2 in IL-rich phase is highest for ILs with uorinated anions The general trend of the phase behavior is almost identical for all ILs Water and carbon dioxide exhibited the strongest interactions and the Highest solubilities in [bmim][PF6 ], followed by ethylene, ethane, and methane Argon and oxygen both had very low solubilities and essentially no interactions with the IL

Reference [70]

10, 25, and 50 C

Up to 13 bar

[69]

20, 40, 60, 80, 100, and 120 C 3070 C

Up to 9.7 MPa Atmospheric pressure

Total pressure increases linearly with increasing amount of CO2 in IL CO2 is found to be one order of magnitude more soluble in the IL than O2 The solubility of CO2 in the IL decreases with temperature The solubility of O2 in the IL slightly increases with temperature Similar results are obtained in [bmim][PF6 ] Solubility of CO2 in ten different IL is reported The solubility of CO2 is strongly dependent on the choice of the anion Increasing the alkyl chain length, increases the solubility of CO2 in IL All of the ILs expand a relatively small amount when CO2 is added

[119] [131]

[bmim][PF6 ]CO2 [bmim][BF4 ]CO2 [bmim][TfO]CO2 [bmim][NO3 ]CO2 [bmim][methide]CO2 [bmim][DCA]CO2 [bmim][Tf2 N]CO2 [hmim][Tf2 N]CO2 [omim][Tf2 N]CO2 [hmmim][Tf2 N]CO2 [emim][PF6 ]CHF3

25, 40, and 60 C

Up to 150 bar

[121]

36.1594.35 C

1.651.6 MPa

The solubility of supercritical CHF3 in [emim][PF6 ] is very high At low CHF3 concentrations (mole fraction <0.5), the equilibrium pressure increases almost linearly with CHF3 concentration, whereas further increase in CHF3 concentration causes a sharp increase in the equilibrium pressure The Peng-Robinson EoS is capable of describing the experimental bubble point data of the system satisfactorily and qualitatively predicting the solubility of the ionic liquid in supercritical CHF3 The high-pressure phase behavior of [emim][PF6 ]CHF3 system was completely different from that of [bmim][PF6 ]CO2 CO2 is more soluble in [bmim][PF6] than in [emim][PF6 ] The general phase behavior of [emim][PF6]CO2 and [bmim][PF6]CO2 systems are found to be completely similar CO2 is more soluble in [bmim][PF6 ] than in [emim][PF6 ] The phase behaviors of the systems [emim][PF6 ]CO2 and [emim][PF6 ]CHF3 are different CHF3 is more soluble in the IL than CO2 at higher pressures and the IL is also more soluble in supercritical CHF3 than in supercritical CO2 CO2 is more soluble in [hmim][PF6] than in [emim][PF6 ] The general phase behaviors of [emim][PF6]CO2 and [hmim][PF6]CO2 are similar The solubility of the IL in scCO2 phase is very low and cannot be detected A group contribution form of a non-random lattice uid model is applied to predict the solubility of CO2 in ILs

[142]

[emim][PF6 ]CO2

3593 C

1.4997.10 MPa

[129]

[hmim][PF6 ]CO2

25.1690.43 C

0.6494.60 MPa

[130]

[bmim][PF6 ]CO2 [C6 mim][PF6 ]CO2 [emim][BF4 ]CO2 [C6 mim][BF4 ]CO2 [emim][Tf2 N]CO2 [C6 mim][Tf2 N]CO2

25 C

Up to 1 MPa

[123]

S. Keskin et al. / J. of Supercritical Fluids 43 (2007) 150180 Table 3 (Continued ) System [bmim][PF6 ]CO2 [emim][PF6 ]CO2 [hmim][PF6 ]CO2 [hmim][BF4 ]CO2 [hmim][PF6 ]CO2 [bmim][BF4 ]CO2 Temperature 4090 C Pressure Up to 97 MPa Findings CO2 has good solubilities in these ILs at lower pressures There is a linear relationship between the alkyl chain length and solubility of CO2 CO2 was found to be more soluble in [hmim][PF6 ] than in [hmim][BF4 ] CO2 has a high solubility in [bmim][BF4 ] at lower pressures, but the solubility decreases dramatically at higher pressures The phase behavior of the system [bmim][BF4 ]CO2 shows similarities with the phase behavior of the system [hmim][BF4 ]CO2 CO2 is more soluble in [hmim][BF4 ] than in [bmim][BF4 ] High solubilities of CO2 for low CO2 mole fractions (mole fraction < 0.6) are attained at relatively low pressure, whereas for mole fraction >0.6, the pressure needed for completely dissolving the CO2 increases drastically Increase in temperature slightly decreases the solubility of CO2 The CO2 solubility increases in the IL with increasing chain length of the alkyl group Suggested for further study to investigate the effect of the anion on the solubility of CO2 .when the cation is xed as [omim]

163

Reference [13]

2095 C 5.3295.07 C

0.54100 MPa 0.58767.62 MPa

[133] [132]

[omim][BF4 ]CO2

29.8589.95 C

0.1100 MPa

[139]

7.1. High-pressure phase behavior of ILCO2 systems Preliminary works have shown that scCO2 extraction is a viable method for solute recovery from an IL. However, the knowledge of phase behavior of ILCO2 systems is an essential aspect of this methodology. scCO2 dissolution in the IL phase is not only necessary for contact with the solute but it also reduces the viscosity of the IL and therefore enhancing the mass transfer process. Early studies of ILCO2 phase behavior indicated that these systems are very unusual biphasic systems. No measurable amount of [bmim][PF6 ] was soluble in the CO2 -rich phase, although a large amount of CO2 dissolved in the IL-rich phase, reducing the viscosity of IL [70]. Blanchard and Brennecke [118] concluded that the system remained as two distinct phases even under pressures up to 400 bar. Therefore, high-pressure phase behavior of [bmim][PF6 ]CO2 is totally different from that of any ordinary organic liquidCO2 systems. This different phase behavior is the key phenomena which makes extraction of solutes from IL with CO2 attractive. Table 3 summarizes the phase behavior studies performed for ILgas systems, demonstrates the type of IL and gases used in these studies, the experimental conditions, the basic ndings and the related references. 7.1.1. The [bmim][PF6 ]CO2 system The phase behaviors of ILscCO2 systems are studied very extensively in the literature for a better understanding of the processes involving both IL and scCO2 . Since [bmim][PF6 ] is the most widely studied IL in the literature, many researchers studied the high-pressure phase behavior of the [bmim][PF6 ]CO2 system [13,69,70,119122]. Blanchard et al. [70] measured the high-pressure vaporliquid phase behavior of [bmim][PF6 ]CO2 system by using two different apparatus sets: a static high-pressure phase equilibrium apparatus and a dynamic ow apparatus. In

the static high-pressure vaporliquid equilibrium apparatus, a glass cell was loaded with a known amount of IL sample and known amounts of CO2 were metered into the cell while the sample within was vigorously stirred to ensure equilibrium. With the assumption of pure CO2 vapor phase, the composition of the IL-rich phase was calculated by knowing the amount of CO2 added to the cell. At the end of the equilibration period, CO2 was completely removed from IL phase upon depressurization. The same group also used a dynamic apparatus, i.e. a high-pressure extractor to determine the solubility of the IL in the CO2 phase. The detailed description of these apparatus sets and experimental procedures can be found in the literature [70]. Different experimental set-ups were used by other researchers: A schematic diagram of a general ILscCO2 experimental apparatus, which was used to measure the solubility of CO2 in IL ([bmim][PF6 ]) is given by Kim et al. [123]. Shiett and Yokozeki [124] measured the gas solubility and diffusivity of CO2 in [bmim][PF6 ] using a gravimetric microbalance for which the details of the experimental set-up is given in the related reference. The solubility of CO2 in [bmim][PF6 ] was determined at 40, 50 and 60 C and pressures up to 93 bar [70]. As the pressure increases, the solubility of CO2 in the IL-rich phase increases dramatically and the solubility value reaches a mole fraction of 0.72 at 40 C and 93 bar. A general rule suggests that an increase in temperature results with a decrease in the solubility of gases in liquids. As expected, the solubility of CO2 in [bmim][PF6 ] rich phase decreases with temperature. However, they noticed that the temperature dependence of the solubility is quite small in this temperature and pressure range. Another crucial point is the effect of large degree of CO2 solubility on the viscosity of IL. The viscosity of IL decreases when a certain amount of CO2 is dissolved in IL and this effect can easily be observed by the reduced drag on the stirring magnet when a static highpressure vaporliquid equilibrium set-up is used. This reduction in viscosity of the liquid facilitates the solution process.

164

S. Keskin et al. / J. of Supercritical Fluids 43 (2007) 150180

Fig. 7. Total pressure versus molality of gas for [bmim][PF6 ]CO2 system. (Reprinted with permission from [119]. Copyright 2003 American Chemical Society)

Kamps et al. [119] presented the solubility of CO2 in [bmim][PF6 ] for temperatures 293393 K in 20 K intervals and pressures up to about 9.7 MPa. The total pressure is plotted versus the stoichiometric molality of the gas (number of moles per kilogram of the IL) in Fig. 7. The total pressure increases linearly with increasing amount of the gas in IL. The solubility data represented by Kamps et al. [119] differ from the previously reported solubility data of Blanchard et al. [70]. The comparison of experimental data of two studies is given in Fig. 8. There are several [bmim][PF6 ]CO2 high-pressure vaporliquid equilibrium data sets available in the literature. However, these sets differ from each other considerably in the values they report for similar conditions. The reason of differing solubility data reported may be due to the small amounts of water dissolved in the IL sample used. For example, Blanchard et al. [125] presented a solubility data of CO2 in [bmim][PF6 ] which is different than the data reported by the same group in 2001. In the rst study, this group used [bmim][PF6 ] which was saturated with water at 22 C, containing 2.3 wt.% water. In the second study, they used [bmim][PF6 ]

Fig. 9. [bmim][PF6 ]CO2 liquid phase compositions for dried and wet IL samples at 40 C. (Reprinted with permission from [70]. Copyright 2001 American Chemical Society)

Fig. 8. Comparison of experimental data of Blanchard et al. [70] ( , , and ) and Kamps et al. [119] ( and ) for [bmim][PF6 ]CO2 system. (Reprinted with permission from [119]. Copyright 2003 American Chemical Society)

which was dried to approximately 0.15 wt.% water. Drying of IL has a signicant effect on the phase behavior with CO2 . Thus, Blanchard et al. [70] showed that the solubility of CO2 in ILs was decreased in the presence of water. In Fig. 9, [bmim][PF6 ]CO2 liquid phase compositions are given for dried and wet IL samples. In order to observe the effect of water impurity in IL, phase behaviors of two IL samples (dry and wet) with CO2 were compared. The effect of water impurity in IL is signicant at 57 bar. For dried IL sample, the mole fraction of CO2 is 0.54, whereas for the wet (water saturated) IL sample it is only 0.13. The effect of water in IL may be explained by CO2 -phobic nature of water. Even at high pressures, mutual solubilities of water and CO2 are very low [126]. Another point is the formation of carbonic acid from the reaction of CO2 with water that can result in a reduction of the aqueous phase pH to as little as 2.80 [127]. Rubero and Baldelli [128] investigated gasliquid interface of imidazolium ILs using surface-sensitive vibrational spectroscopy sum frequency generation. The results indicated that when the IL is dry, the cation is oriented with the imidazolium ring parallel to the surface plane for both hydrophilic and hydrophobic ILs. But the cation reorients itself with respect to the surface for the hydrophobic liquid when water is added, while the orientation in the hydrophilic liquid is unaffected. After the inuence of water is noticed, researchers working on ILCO2 solubility and equilibrium have started to dry and degas all ILs under vacuum at room temperature for several days prior to use. After ILs are dried and degassed, the water contents are estimated by the Karl Fischer analysis before solubility data are taken. Measurements show that the most widely studied IL; [bmim][PF6 ] absorbs a couple wt.% water when left to the atmosphere. The estimated water content of [bmim][PF6 ] after drying was approximately 0.15 wt.% water as measured by Karl Fischer analysis [70]. A number of high and low-pressure [bmim][PF6 ]CO2 solubility studies have appeared in the literature. Although consistent

S. Keskin et al. / J. of Supercritical Fluids 43 (2007) 150180

165

results have been established for low-pressure solubility data of CO2 in [bmim][PF6 ], there are large discrepancies among highpressure solubility data of several researchers [13,70,119121]. These large solubility differences in the literature are most probably due to the differences in the purities of the ILs used. Since the behaviors of ILCO2 systems are different from other organic liquidCO2 systems, full phase diagrams of ILCO2 systems are investigated. Blanchard and co-workers [125] found two-phase immiscibility regions with three cloud point measurements of 1.31, 4.92 and 7.15 mole% IL mixtures with the balance being CO2 . Although, these experiments were conducted with water-saturated ILs, qualitatively a similar behavior is expected with dried samples. Blanchard et al. [70] gave a qualitative phase behavior of [bmim][PF6 ]CO2 system over a wide pressure range. They noticed that the phase behavior where a large miscibility gap exists even at extremely high pressures. Blanchard et al. [70] reported and referred that as a complementary work, McHugh and co-workers studied [bmim][PF6 ]CO2 phase behavior at higher pressures up to 3100 bar and found two distinct phases at all conditions. The existence of large immiscibility gap even at very high pressures is not expected for organic liquidCO2 systems and the existence of two distinct phases is explained by the following discussion: At high-pressures density of pure CO2 phase increases but since the liquid phase does not expand, the two phases will never become identical and a mixture critical point will never be reached. Therefore, the ILCO2 system remains as two phases even at very high pressures, although the CO2 solubility is quite high, the mixtures never become a single phase [70]. Anthony et al. [69] reported the solubilities and Henrys constants of different gases (carbon dioxide, ethylene, ethane, methane, argon, oxygen, carbon monoxide, hydrogen, and nitrogen) in [bmim][PF6 ] and showed that CO2 has the highest solubility and strong interaction with [bmim][PF6 ]. The solubility data of CO2 in [bmim][PF6 ] is in good agreement with the published results of Blanchard et al. [70] although different techniques were used in these studies. Furthermore, Baltus et al. [68] reported that Henrys constants for Kamps et al. [119] data are in reasonable agreement with those obtained by Anthony et al. [69]. Aki et al. [121] studied the high-pressure phase behavior of CO2 in imidazolium-based ILs and compared the phase behavior of the system [bmim][PF6 ]CO2 with the other solubility data present in the literature. The solubility results of CO2 in [bmim][PF6 ] at 25 C measured by Aki et al. [121] agreed remarkably well with the solubility results of Anthony et al. [69] at low pressures and with the solubility results of Kamps et al. [119]. Aki et al. [121] investigated the solubility of CO2 in [bmim][PF6 ] at 40 C and compared the results with the previous studies of Kamps et al. [119], Blanchard et al. [70] and Liu et al. [120]. As expected the agreement between the data points is good at low pressures but the discrepancy is obvious at high pressures. Aki et al. [121] explained that in their previous work [70], they were not aware of the various impurities and degradation products that were present in the samples they used. There-

fore, they attributed the difference between their study [121] and the previous study of the same group [70] to the purity of the IL. At 40 C and at all pressures, there is a very good agreement within the solubility values reported by Aki et al. [121] and Liu et al. [120]. However, this statement is not correct for the reported solubility data of Aki et al. [121] and Kamps et al. [119]. The results of two studies agree at low pressures, but at high pressures, there is a signicant difference: At about 43 bar, the solubility of CO2 in [bmim][PF6 ] was measured as 0.43 (mole fraction) by Aki et al. [121], however, at the same point Kamps et al. [119] reported the solubility of CO2 as 0.38. By considering the studies mentioned above, it may thus be concluded that the solubility of CO2 in one of the most widely studied IL, [bmim][PF6 ], varies among different groups in the literature and is not a good agreement especially at higher pressures. The solubility of CO2 in [bmim][PF6 ] was experimentally studied at 298.15 K and up to 1.0 MPa by Kim et al. [123]. A group contribution form of a non-random latticeuid model (GC-NLF) was applied to predict solubility of CO2 in [bmim][PF6 ]. They used the solubility data of Kamps et al. [119] for a wider pressure range for the group parameter determination. Comparisons of calculated solubility data with Kamps et al. data [119] for [bmim][PF6 ] demonstrated that the method applied is fairly accurate except for regions close to the critical conditions of CO2 . Kim et al. [123] also compared the calculated values of solubility of CO2 in different ILs ([emim][PF6 ], [bmim][PF6 ], and [C6 mim][PF6 ]) with the experimental solubility data reported by other groups for the same ILs. The agreements are generally good up to 10 MPa pressure, however, further comparisons for higher pressure shows some degree of discrepancy between the calculated solubility data of Kim et al. [123] and the experimental solubility data of Shariati and Peters [129,130]. Finally, Shariati and Peters [13] studied the comparison of the phase behavior of [bmim][PF6 ]scCO2 system with the other studies present in the literature. The solubility of CO2 in [bmim][PF6 ] was determined by measuring the bubble point pressure of the binary system at different temperatures for several isopleths and pressures up to 97 MPa. The solubility data of CO2 in [bmim][PF6 ] at 323.15 K was compared with that of Blanchard et al. [70] and Anthony et al. [71] and the results of Shariati and Peters [13] were in a good agreement with those of Anthony et al. [71]. The solubility data taken at 333.15 K was compared with that of Blanchard et al. [70], Kamps et al. [119], Liu et al. [120]. Although the experimental methods were completely different, there is also a good agreement between the results of Shariati and Peters [13] and Kamps et al. [119] at a temperature of 333.15 K. The solubility data of Blanchard et al. [70] and Liu et al. [120] show greater deviations from the data of Shariati and Peters [13] especially at higher pressures. Shariati and Peters [13] reported the existence of the three-phase equilibrium liquidliquidvapor (LLV). As the other studies demonstrated, this study also indicated that CO2 has a high solubility in [bmim][PF6 ] and there is a linear relationship between the alkyl chain length and the solubility of CO2 .

166

S. Keskin et al. / J. of Supercritical Fluids 43 (2007) 150180

7.1.2. Other ILCO2 systems High-pressure phase behavior of different types of ILs are similar to that of [bmim][PF6 ]. Blanchard et al. [70] investigated the high-pressure phase behavior of CO2 with six different ILs: 1-n-butyl-3-methylimidazolium hexauorophosphate ([bmim][PF6 ]), 1-n-octyl-3 methylimidazolium hexauorophosphate ([C8-mim][PF6]), 1-n-octyl3-methylimidazolium tetrauoroborate ([C8 -mim][BF4 ]), 1n-butyl-3-methylimidazolium nitrate ([bmim][NO3 ]), 1-ethyl3-methylimidazolium ethyl sulfate ([emim][EtSO4 ]), and N-butylpyridinium tetrauoroborate ([N-bupy][BF4 ]). They investigated the solubility of CO2 in different ILs at 40, 50, 60 C and pressures up to 93 bar. The focus of the work was to develop an insight into the physical interaction between CO2 and ILs with different cationanion congurations. The solubility of CO2 in the IL-rich phase was greatest for ILs with uorinated anions, following the trend of [bmim][PF6 ] and [C8 -mim][PF6 ] > [C8 -mim][BF4 ] > [Nbupy][BF4 ] > [bmim][NO3 ] > [emim][EtSO4 ] [70]. The solubility of CO2 was found to be greatest in ILs with PF6 , anion and the IL that exhibited the lowest solubility of CO2 was [emim][EtSO4 ]. The solubility of CO2 in ILs increases with increasing pressure but the exact amount of CO2 dissolved in the liquid phase varies signicantly. In the study of Blanchard et al. [70], it is found that at 70 bar the solubility of CO2 in [emim][EtSO4 ] was 0.36 (mole fraction) whereas, it was 0.63 in [C8 -mim][PF6 ]. Although there are some numerical differences in the mole fraction of CO2 dissolved, the general trend of the phase behavior is nearly identical for all ILs. The solubility of CO2 in the IL-rich phase changes slightly with the temperature as in the case of [bmim][PF6 ]CO2 system. Since the qualitative phase behavior of almost all ILs seems similar, it can easily be concluded that scCO2 can be used to recover solutes not only from [bmim][PF6 ] but also from all kind of ILs. In order to understand the effect of the anion on the phase behavior of ILCO2 systems, two pairs of ILs with the same cations were compared: [C8 mim][PF6 ][C8 mim][BF4 ] and [bmim][PF6 ][bmim][NO3 ]. Changing the anion from [PF6 ] to [BF4 ] in the [C8 mim] salts, results in an approximately 8% decrease in CO2 solubility at 40 C over the range of pressures studied by Blanchard et al. [70]. The solubility of CO2 in [bmim][NO3 ] is about 25% less than in [bmim][PF6 ]. Experimental and molecular simulation studies found that the anions of IL dominate the interaction with CO2 , with the cation playing a secondary role. Cadena et al. [122] pointed out that the changes in the imidazolium cation involving alkyl groups have relatively little inuence on the solubility of CO2 in IL. [bmim][BF4 ] is one of the popular ILs which has been very widely used in most of the studies: Husson-Borg et al. [131] reported the solubility of CO2 in [bmim][BF4 ] as a function of temperature between 303 and 343 K and at atmospheric pressure. This group used a new type of experimental apparatus based on a saturation method. The equilibrium cell is specially designed for viscous solvents like the IL, and an appropriate gasliquid contact is obtained by good agitation. Kroon et al. [132] studied the phase behavior of different

IL-supercritical uid systems including [bmim][BF4 ]. They investigated the phase behavior of [bmim][PF6 ]CO2 binary system experimentally and reported its bubble point pressures for CO2 concentrations between 10.22 and 60.17 mole% and in a temperatures range of 278.47368.22 K. They found that CO2 has a high solubility at lower pressures, but the solubility decreases dramatically at higher pressures. The experimental results for the [bmim][BF4 ]CO2 system were compared with the available phase behavior data of the binary system 1-hexyl-3-methylimidazolium tetrauoroborate ([hmim][BF4 ])CO2 [133] to investigate the effect of the alkyl group length on the phase behavior. The results showed that a larger alkyl group led to lower bubble-point pressures and, therefore, to higher solubilities of CO2 in the imidazolium-based ionic liquid. Thus, CO2 was more soluble in [hmim][BF4] than in [bmim][BF4 ]. Shariati and Peters [129] studied the phase behavior of binary system [emim][PF6 ]CO2 experimentally by measuring its bubble point pressures at temperatures and pressure ranges of 308.14366.03 K and 1.4997.10 MPa, and compared it with [bmim][PF6 ]CO2 system to understand the effect of the length of alkyl chain on the solubility of CO2 . The solubility of CO2 in [bmim][PF6 ] is higher than in [emim][PF6 ]. Since the butyl group in [bmim][PF6 ] is bulkier than the ethyl group in [emim][PF6 ], Shariati and Peters [129] emphasized that CO2 can dissolve better in [bmim][PF6 ] than in [emim][PF6 ], the latter component being denser. Moreover, it was observed that the experimentally determined phase behavior of the [emim][PF6 ]CO2 is similar to the phase behavior of [bmim][PF6 ]CO2 at 333.15 K. Aki et al. [121] also presented the solubility of CO2 in ten different imidazolium-based ILs at 25, 40, and 60 C and pressures to 150 bar. They concluded that the solubility of CO2 in imidazolium-based ILs increases with increasing pressure and decreases with increasing temperature for all the ILs investigated. Furthermore, Aki et al. [121] investigated the inuence of the different anions, namely; dicyanamide ([DCA]), nitrate ([NO3 ]), tetrauoroborate ([BF4 ]), hexauorophosphate ([PF6 ]), trifuoromethanesulfonate ([TfO]), bis(triuoromethylsulfonyl) imide ([Tf2 N]), and tris(triuoromethylsulfonyl)methide ([methide]). The results of their study showed that the solubility of CO2 is strongly dependent on the choice of anion. Solubility measurements of Anthony et al. [69] and Cadena et al. [122]; spectroscopic studies of Kazarian et al. [134] also all showed that the solubility of CO2 in ILs depend on the anion, especially the interaction between the anion of the IL and CO2 . CO2 is the least soluble in the two ILs with non-uorinated anions, [NO3 ] and [DCA], and it has the highest solubility in ILs with anions containing uoroalkyl groups, [TfO], [Tf2 N], and [methide]. Aki et al. [121] attributed the high solubility of CO2 in ILs to the anions containing uoroalkyl groups as a result of favorable interactions between CO2 and the uoroalkyl substituents on the anion since uoroalkyl groups are known as CO2 -philic ones [135137]. Aki et al. [121] also compared the solubility of CO2 in [bmim][Tf2 N], [hmim][Tf2 N], and [omim][Tf2 N] at 25, 40, and

S. Keskin et al. / J. of Supercritical Fluids 43 (2007) 150180

167

60 C to investigate the inuence of cation alkyl chain length. The solubility of CO2 increases with an increase in the alkyl chain length at all pressures, with the increase being more apparent at higher pressures. This result of Aki et al. [121] is consistent with the results of previous studies of Shariati and Peters [129,130]. Furthermore, Blanchard et al. [70] reported that the CO2 solubility increases when the alkyl chain length was increased from butyl to octyl for ILs containing [PF6 ]. It is known that the densities of the imidazolium-based ILs decrease as the alkyl chain length increases [49,138]. Therefore, imidazolium-based ILs with longer alkyl chains have greater free volume and thus greater solubility of CO2 is expected in these ILs. Therefore, the higher solubility of CO2 in ILs with longer cation alkyl chains may be attributed to the entropic rather than enthalpic arguments [121]. Finally, it may be concluded that it is possible to increase the solubility of CO2 in ILs by increasing the alkyl chain length on the cation. Constantini et al. [133] investigated the phase behavior of a binary mixture of 1-hexyl-3-methylimidazolium tetrauoroborate ([hmim][BF4 ])scCO2 system and compared it with the experimental data of the binary system of 1-hexyl3-methylimidazolium hexauoroborate ([hmim][PF6 ])scCO2 system, in order to demonstrate the anion effect. Although the phase behaviors of the binary systems are similar, the solubility of the scCO2 is found to be higher in [hmim][PF6 ] than in [hmim][BF4 ]. This higher solubility may be explained with the greater interaction between CO2 and the [PF6 ] anion, although the former is denser than the latter [133]. In another study [139], the high-pressure phase behavior of the binary system 1-octyl-3-methylimidazolium tetrauoroborate ([omim][BF4 ])CO2 was studied in the liquid-phase CO2 mole fraction range 0.10.75, and in the pressure and temperature range of 0.1100 MPa and 303363 K. They compared their experimental data with the data given by Blanchard et al. [70]. Although there was a discrepancy between two studies for CO2 mole fractions higher than 0.6, they both suggested a peculiar behavior of a rapid pressure increase at higher CO2 mole fractions [139]. Kroon et al. [140] presented the phase behavior of several ILCO2 binary systems experimentally, and nally developed an equation of state (EoS) to predict the phase behavior of ILCO2 systems based on the truncated perturbed chain polar statistical associating uid theory (tPC-PSAFT) EoS. The EoS was used to describe the CO2 solubility in several 1-alkyl3-methylimidazolium-based ionic liquids with different alkyl chain lengths within a pressure and CO2 mole fraction range of 0100 MPa and 075%, respectively. The binary interaction parameter was tted to VL equilibrium data. In general when a gas is dissolved in a liquid phase, dilation of the liquid occurs. ILs show only slight dilations in volume with CO2 dissolution. This behavior is different than normal liquidCO2 systems. In fact, the dissolution of CO2 in liquids to expand them and reduce their solvent strength is a well-known phenomena and it is the basis of gas anti-solvent (GAS) process to precipitate solutes from liquids [141]. However, ILs do not follow this trend due to the strong Coulombic forces associated with the ionic nature of ILs. The lack of expansion of ILs seems to derive from the fact that dissolved CO2 does not greatly affect the

strength of interionic interactions in IL. Therefore, the ILCO2 system is extremely different from other organic liquidCO2 systems with the large immiscibility region and lack of dilation of the liquid phase [70]. Although there are not so many available studies concerning the IL/SCF other than CO2 systems, for the completeness of the material they must be mentioned. Shariati and Peters [142] investigated the high-pressure phase behavior of [emim][PF6 ]supercritical uoroform, the experimental conditions and the results of this study is given on Table 3. 7.2. IL solubility in CO2 The majority of the research focus on the solubility of CO2 in IL, on the other hand the concentration of IL in CO2 -rich phase is also important. A dynamic apparatus with a high-pressure cell, was used to measure the solubility of [bmim][PF6 ] in the CO2 -rich phase [70]. The solubility of [bmim][PF6 ] in CO2 was determined at 40 C and 137.9 bar by owing 0.5866 mole of CO2 through a cartridge loaded with [bmim][PF6 ]. UVvisible analysis gave no appreciable IL absorption peak, indicating [bmim][PF6 ] solubility of less than 5 107 (mole fraction of [bmim][PF6 ]) of in the CO2 phase. Due to the fact that no measurable IL dissolves in CO2 , a solute dissolved in an IL can be easily recovered with scCO2 without any cross contamination. The lack of solubility of IL in the CO2 phase can be attributed to two reasons: extremely low vapor pressure of IL and the inability of CO2 to adequately solvate ions in the gaseous phase. Although the solubility of IL in scCO2 is extremely low and not measurable, in industrial applications, the scCO2 phase may contain some other components such as reactants, products which may act as cosolvents to enhance the ability of scCO2 to dissolve IL signicantly. In this case, the amount of IL dissolved in CO2 -rich phase may not be negligible under some conditions. In order to decide on the conditions to avoid cross contamination, solubility of IL in scCO2 /organic compound mixtures must be known. Wu et al. [143] conducted the rst study in the literature to observe the effect of organic compounds in scCO2 on the solubility of an IL in CO2 phase. The solubility of [bmim][PF6 ] in scCO2 , and in scCO2 ethanol, scCO2 acetone, scCO2 n-hexane mixtures was investigated quantitatively. The solubility of IL in scCO2 is extremely low as the earlier studies suggested, however by addition of ethanol and acetone, the solubility of IL increases dramatically as the concentration of the organic compounds in scCO2 exceeds 10 mole%. This enhancement of the IL solubility by addition of ethanol and acetone results mainly from strong interaction of the two compounds with the IL due to their strong polarity. The polarity of acetone is stronger than ethanol; therefore, the solubility of IL is higher in the scCO2 acetone system than in the scCO2 ethanol system. Since n-hexane is a nonpolar substance its inuence on the solubility of IL in scCO2 phase is very limited. A further study of Wu et al. [144] also showed that the ability of cosolvents to increase the solubility of ILs ([bmim][PF6 ] and [bmim][BF4 ]) in scCO2 follows the order: acetonitrile > acetone > methanol > ethanol > n-hexane. This order is same with the order of dipole moments of cosolvents. With this

168 Table 4 IL solubility in CO2 System [bmim][PF6 ]CO2

S. Keskin et al. / J. of Supercritical Fluids 43 (2007) 150180

Temperature 40 C

Pressure 137.9 bar

Findings The solubility of [bmim][PF6] in scCO2 is extremely low scCO2 may be used to recover solutes dissolved in IL without any cross contamination By addition of ethanol and acetone, the solubility of the IL in scCO2 phase enhances

Reference [70]

[bmim][PF6 ]CO2 [bmim][PF6 ]CO2 ethanol [bmim][PF6 ]CO2 acetone [bmim][PF6 ]CO2 n-hexane [bmim][PF6 ]CO2 [bmim][PF6 ]CO2 acetonitrile [bmim][PF6 ]CO2 methanol [bmim][PF6 ]CO2 ethanoln-hexane [bmim][PF6 ]CO2 acetonen-hexane [bmim][PF6 ]CO2 ethanolacetone [bmim][BF4 ]CO2 [bmim][BF4 ]CO2 ethanol

40 and 55 C

1215 MPa

[143]

40 C

15 MPa

By the addition of organic cosolvents, the solubility of ILs in scCO2 phase may be enhanced signicantly

[144]

discussion, it is clear that the polarity of the organic compound is a dominant factor in inuencing solubility. With this study Wu et al. [144] emphasized that the amount of IL dissolved in scCO2 rich phase may be signicant if the system contains sufciently polar organic compounds in sufcient concentrations. Table 4 shows the studies of Blanchard et al. [70], Wu et al. [143] and Wu et al. [144], the components of their systems, experimental conditions and basic results. The phase behavior of the ILCO2 methanol system and the viscosity of the mixtures were studied previously by Liu et al. [120] at different conditions. The phase behavior of ILCO2 water system was investigated by Zhang et al. [145] and nally this group studied the phase behavior of [bmim][PF6 ]CO2 acetone system in detail at 313.15 K over a wide pressure range. Zhang et al. [146] determined the distribution coefcients of the components between different phases and found that CO2 distribution coefcient decreases with increasing pressure while the acetone distribution coefcient increased with pressure. 7.3. ILCO2 interaction at the molecular level An in situ attenuated total reectanceinfra red (ATRIR) study of CO2 dissolved in two ILs ([bmim][PF6 ] and [bmim][BF4 ]) at high pressures has demonstrated the effects of anionic species of the ILs on the molecular state of the dissolved CO2 . Kazarian et al. [147] showed that CO2 forms weak Lewis acidbase complexes with the anions in [bmim][PF6 ] and [bmim][BF4 ]. Furthermore, they demonstrated that this interaction is stronger with [bmim][BF4 ]. BF4 acts as a stronger Lewis base towards CO2 than PF6 . In addition to this, as the size of anion increases the strength of interaction decreases. However, the solubility data of various group indicated that CO2 has a higher solubility in [bmim][PF6 ] than in [bmim][BF4 ]. Thus, the strength of these interactions cannot be solely responsible for the solubility of CO2 in these ILs, and presumably, a free volume contribution in IL plays a signicant role [148]. The strength of the anioncation interactions in IL affects the avail-

able free volume and one can expect that a weaker interacting anion leads to more free volume being available. ILs are generally distinguished with their low melting points. Another in situ ATRIR spectroscopic study of Kazarian et al. [134] showed that high-pressure CO2 reduces the melting temperature of ILs. This possibility of reducing melting temperature of ILs further under high pressure CO2 provides a new opportunity to use ILs as solvents at milder temperatures. Kazarian et al. [134] investigated the effect of CO2 pressure on the phase behavior of 1-hexadecyl-3-imidazolium hexauorophosphate. The data presented indicated that 70 bar CO2 reduces the melting point of this IL from 75 to 50 C. This result was assigned to a weak Lewis acidbase type interaction between anion and CO2 , with the PF bonds perpendicular to O C C axis thereby reducing the rather stronger interactions between the PF bonds and the cations. CO2 disrupts the cationanion and the tail-tail interactions in IL rather than simply playing an impurity role in the mechanism of induced melting. Preliminary results obtained by Kazarian et al. [134] also indicated that melting temperatures of other analogous ILs decrease by high pressure CO2 allowing their use in many applications at mild temperatures. 7.4. Solute recovery from ILs with scCO2 Researchers found that nonvolatile organic compounds can be extracted from ILs using scCO2, which is widely used to extract large organic compounds with minimal pollution. Blanchard et al. [125] showed that CO2 can be used to extract naphthalene, a low volatility model solute, from an IL. They synthesized [bmim][PF6 ] which was stable in the presence of water and oxygen. The model compound naphthalene was readily soluble in [bmim][PF6 ] (maximum solubility of 0.30 mole fraction at 40 C) and in CO2 . Their study investigated the phase behavior of [bmim][PF6 ] with CO2 , as well as with naphthalene and nally that of the [bmim][PF6 ]CO2 naphthalene ternary system. The results showed that CO2 -rich phase was not signicantly contaminated by IL, as would be expected during contact of CO2 with any conventional organic solvent. Blanchard et al.

S. Keskin et al. / J. of Supercritical Fluids 43 (2007) 150180

169

[125] demonstrated that CO2 is highly soluble in [bmim][PF6 ] reaching a mole fraction of 0.6 at 8 MPa. After extracting IL with CO2 at 13.8 MPa and 40 C there was no detectable [bmim][PF6 ] in the extract, indicating that the solubility is less than 105 mole fraction. In contrast to this result, a mixture of CO2 with conventional organic liquid results in signicant solubility of the liquid in the CO2 phase. A mixture of 0.12 mole fraction naphthalene in [bmim][PF6 ] was extracted with CO2 at 13.8 MPa and 40 C with recoveries of 9496% and therefore it was concluded that it is possible to quantitatively extract an organic solute having a reasonable molecular weight from IL using scCO2 without any cross contamination. Furthermore, the dissolution of scCO2 in IL was completely reversible and pure IL remained after the extraction of naphthalene and depressurization [125]. The experimental work of Blanchard and Brennecke [118], which showed that a wide variety of solutes can be extracted from [bmim][PF6 ] with scCO2 , with recovery rates greater than 95%, provided a signicant step to visualize the partnership established by the ILCO2 system. One of the essential problems of ILs, namely product recovery, was solved by applying the SCFE technique. Hexane and benzene were chosen as the roots to which numerous substituents groups were added to explore the effect of chemical structure on the solubility and extractability of an organic solute in and from an IL. The substituent groups represented were halogen, alcohol, ether, amide, ketone, carboxylic acid, ester, and aldehyde with a wide range of polarity. The extraction experiments were conducted at 40 C and 138 bar. The authors noted that for some organic solutes, scCO2 extraction achieved greater than 98% recoveries before the extraction test was terminated. These high recovery rates clearly indicated that although the ionic nature of IL might lead to an interaction with solute, it did not limit the extent of reaction. Benzene and chlorobenzene which exhibited phase immiscibility with [bmim][PF6 ] required the least amount of CO2 for recovery. Phenols, benzoic acid and benzamide which are solids at room temperature, required the largest amount of CO2 . Figs. 10 and 11 illustrate % recovery of solutes as a function of molar ratio of CO2 passed through the extractor to organic solute loaded in the reactor (solute dipole moments are also given in these gures in

Fig. 11. Extraction of aliphatic solutes from [bmim][PF6 ] with scCO2 at 40 C and 138 bar. (Reprinted with permission from [118]. Copyright 2001 American Chemical Society)

terms of Debye). In this paper they showed that there was a relation between the dipole moments of solutes and the amount of CO2 required for extraction, concluding that the organics with a dipole moment of zero (such as benzene, hexane, cyclohexane, etc.) are easily extracted compared to solutes with nonzero dipole moments. Fig. 12 illustrates the number of CO2 per mole of organic solute as a function of dipole moment. Studies on the solubilities of organics in [bmim][PF6 ] were carried out under ambient conditions, 22 C and 0.98 bar [118], where ILsolute mixtures were stirred in closed containers to avoid contamination with air and water vapor. Their results demonstrated that organics with the potential for strong intermolecular interactions, those with a large dipole moment for example, generally exhibited complete miscibility or a large degree of solubility in [bmim][PF6 ]. Solubilities of the solid organic solutes were considerably less than the solubilities of the liquid organics, with the exception of the phenol. The authors

Fig. 10. Extraction of aromatic solutes from [bmim][PF6 ] with scCO2 at 40 C and 138 bar. (Reprinted with permission from [118]. Copyright 2001 American Chemical Society)

Fig. 12. Effect of solute dipole moment on ease of extraction of [bmim][PF6 ] with scCO2 at 40 C and 138 bar. (Reprinted with permission from [118]. Copyright 2001 American Chemical Society)

170

S. Keskin et al. / J. of Supercritical Fluids 43 (2007) 150180

concluded that the solubilities of the benzene-based organics were signicantly greater than those of their hexane-based counterparts. Solubility measurement results indicated that aromatic compounds are more soluble in [bmim][PF6 ] than in non-aromatic compounds of equivalent molecular weight and polarity. Distribution coefcient (K), may be dened as the ratio of solute mole fractions in the supercritical and in the IL phases, respectively. Since ILs do not dissolve in scCO2 appreciably, supercritical phase is essentially CO2 and organic solute. The distribution coefcient is an important thermodynamic property to comprehend the ILscCO2 solute systems. Theoretically, one can anticipate the trend of distribution coefcient between the IL and the scCO2 phases by considering the volatility and polarity characteristics of the solute. For example, a solute with a high volatility and low polarity will have a large afnity for CO2 , whereas a solute with high polarity and aromaticity will have a large afnity for the IL-rich liquid phase. Therefore, solutes with high polarities give a small K value due to high afnity for IL and low afnity for CO2 . Conversely, nonpolar solutes give large K values as a result of high afnity for CO2 . With this discussion, it is clear that the compounds that have high afnity for CO2 can be more easily extracted from the IL mixture. The phase behaviors of the organic soluteCO2 binary systems also affect the ease of extraction of a compound from ILs with scCO2 . Investigating the liquid phase data at low pressures, a measure of afnity of compounds for CO2 was determined, and also that compounds in which CO2 readily dissolves at low pressures have a greater attraction for CO2 [118]. Furthermore, the mole ratio of CO2 to solute for 95% recovery is determined as 1840 for CO2 cyclohexane binary system [149] however, the same ratio is as high as 20,300 for CO2 acetophenone system [150]. Using scCO2 to recover products from IL is a good alternative if the products are thermally sensible or nonvolatile. A high boiling point solute can also be recovered from IL with scCO2 , where distillation is not a desirable option. Blanchard and Brennecke [118] extracted a high-boiling point (230 C) organic liquid namely 1,4-butanediol from [bmim][PF6 ]. This high-boiling point solute followed the same solubility and extractability trend as the other liquid solutes. Therefore, SCFE technology is readily applicable for the recovery of various kinds of compounds from ILs. Scurto et al. [151] studied the use of scCO2 as a separation switch for ILorganic mixtures. They demonstrated that the solutions of methanol and 3-butyl-1-methyl-imidazolium hexauorophosphate ([C4 mim][PF6 ]) can be induced to form three phases in the presence of scCO2 . Although the original solution is quite dilute in IL, application of scCO2 induces the formation of an additional liquid phase which is rich in IL. This study showed that there is an alternative way that CO2 may be used to separate ILs from organic compounds and the extraction of organic materials can be achieved without any IL cross contamination in the recovered product. In their following study, Scurto et al. [152] demonstrated that separation of hydrophobic and hydrophilic imidazolium-based ILs from aqueous solutions by the application of scCO2 is possible and [bmim][PF6 ] is sepa-

rated from an IL-saturated aqueous solution at 293 K and at a CO2 pressure of 4.9 MPa. The result of the mentioned studies presented that organic compoundILscCO2 mixture has a complex phase behavior. According to the results of their research Scurto et al. [151] emphasized that during the ILscCO2 reaction studies in which larger amounts of organic reactant and products are present, one must be aware of the possible formation of the additional liquid phases that might contain only part of the components necessary for the desired reaction. After showing ILs could be recovered from methanol and water using CO2 -induced separation, the same research group [153] investigated the factors that control the vaporliquidliquid equilibrium in ILorganic compoundCO2 ternary systems via studying on several homogeneous ILorganic compound mixtures. The experiments were conducted at 40 C. The results showed that the lowest critical endpoint pressure (LCEP) was dependent on the choice of both organic, IL and the initial concentration of IL in the organic. The K-point pressure was however independent of the type of IL, was identical with the organic compoundCO2 mixture critical point. Najdanovic-Visak et al. [154] investigated the vaporliquid equilibrium of ternary (1-butanolwaterCO2 ) and quaternary ([C4 mim][NTf2 ]1-butanolwaterCO2 ) systems. The demixing pressures of both mixtures were strongly controlled by the water concentration. The studies concerning the solute recovery from ILs by scCO2 , the system components, experimental conditions, major results of the studies and related references are given in Table 5. 7.5. Other applications of ILscCO2 systems Brennecke and Maginn [8] discussed the potential industrial applications of ILs in many areas such as catalytic reactions, liquidliquid extractions, gas separations etc. After the ILscCO2 systems and the advantages of these systems are realized, a number of studies have been done for ILCO2 biphasic systems: It is shown that a desired solute may be extracted from an IL using scCO2 without any cross contamination [118,125]. The use of scCO2 to separate ILs from their organic solvents [151]; the addition of CO2 to separate hydrophobic and hydrophilic imidazolium-based ILs from aqueous solutions [152] have all been important applications of ILscCO2 systems. Dzyuba and Bartsch [155] demonstrated the recent applications of room temperature ILscCO2 systems in metal catalyzed organic reactions and enzyme-catalyzed transformations. The solubility or stability of organometallic or enzymatic catalysts in ILs and their negligible solubility in scCO2 is the basic advantage of ILscCO2 systems. Several groups have studied the ILscCO2 reaction systems [156168]. In these studies, mostly IL was used as reaction media and scCO2 was used as transport media for reactants and products. Cole-Hamilton et al. [169] summarized the continuous ow homogeneous catalysis using ILscCO2 biphasic systems in detail according reaction types. Also, Gordon and Leitner [170] mentioned some of ILscCO2 biphasic reaction systems. In the following paragraphs, ILscCO2 biphasic reaction systems are summarized.

S. Keskin et al. / J. of Supercritical Fluids 43 (2007) 150180 Table 5 Recovery of organic compounds from IL by scCO2 System [bmim][PF6 ]CO2 Solute Naphthalene Temperature 2540 C Pressure Up to 40 MPa Findings CO2 is highly soluble in [bmim][PF6] Two phases are not completely miscible CO2 may be used to extract naphthalene, model solute, from an IL The dissolution of CO2 in IL is completely reversible: pure IL remains after extraction of naphthalene and depressurization CO2 can extract a wide variety of organic solutes from an IL IL contamination in the recovered product is eliminated by using CO2 All organic solutes exhibited recovery greater than 95% Intermolecular interactions between organic solutes and [bmim][PF6] have an effect on the solubility of solute in IL, but these interactions do not limit the degree of recovery CO2 can extract a wide variety of organic solutes from an IL IL contamination in the recovered product is eliminated by using CO2 All organic solutes exhibited recovery greater than 95% Intermolecular interactions between organic solutes and [bmim][PF6] have an effect on the solubility of solute in IL, but these interactions do not limit the degree of recovery

171

Reference [125]

[bmim][PF6 ]CO2

Benzene Chlorobenzene Phenol Anisole Aniline Acetophenone Benzoic acid Methyl benzoate Benzamide Benzaldehyde Hexane 1-Chlorohexane 1-Hexanol Butyl ethyl ether Cyclohexane 2-Hexanone Hexanoic acid Methylpentanoate Hexanamide 1,4-Butanediol

40 C

138 bar

[118]

Sellin et al. [156] and Webb et al. [157] studied the hydroformylation of alkenes in ILscCO2 biphasic reaction media and described the continuous ow homogeneous catalysis in ILscCO2 biphasic system. They dissolved the catalyst in ionic liquid and used scCO2 as the transport medium for the substrates and products. In Sellin et al. [156], they demonstrated the hydroformylation of several alkenes such as 1-hexene, 1nonene and 1-octene in ILscCO2 biphasic mixture. First, the hydroformylation of 1-hexene was studied using triphenylphosphite as the rhodium-based ligand ([Rh2 (OAc)4 ]/P(OPh3 )) in [bmim][PF6 ] using scCO2 . The hydroformylation of 1-hexene was also performed in [bmim][PF6 ] without scCO2 . The results showed that addition of scCO2 to the reaction mixture lowers the conversion from >9940%, but the selectivity and linear to branched (l:b) ratio were enhanced from 15.7 to 83.5% and 2.4 to 6.1, respectively. The success of this system encouraged the authors to carry out similar reactions of 1-hexene and 1-nonene with repetitive uses of the same catalyst to search for the possibility of continuous ow homogeneous catalysis. They found that the catalyst retained its activity and selectivity for only three to three runs, so in order to eliminate this drawback, the ligandcatalyst system was changed. [Ph2 P(C6 H4 SO3 )][bmim] was used together with [Rh2 (OAc)4 ] as the catalyst precursor for the hydroformylation of 1-nonene in the [[bmim][PF6 ]scCO2 biphasic system. The products were ushed from the reactor with scCO2 . The activity of this cata-

lyst system remained high for 12 runs with an acceptable l:b ratio. However, after the ninth run Rh leaching became important which was attributed to ligand oxidation. Contamination with air during the many openings of the reactor may cause the oxidation, which will be eliminated during continuous ow. Finally, Sellin et al. [156] demonstrated the continuous hydroformylation of 1-octene using [PhP(C6 H4 SO3 )2 ][pmim]2 and [Rh2 (OAc)4 ] dissolved in [bmim][PF6 ]. The reactants and the products were transported into and out of the reaction medium via scCO2 . The total pressure, temperature and reaction time were 200 bar, 100 C and 33 h. The results showed that the catalyst was stable at least between 8 and 10 h reaction time, no ligand oxidation occurred and the l:b ratio of the product aldehydes was 3.1. In another study, Webb et al. [157] investigated the hydroformylation 1-dodecene which was representative for hydroformylation of relatively low volatility alkenes. The reactions were catalyzed by either Rh/[prmim][Ph2 P(3-C6 H4 SO3 )] or Rh/[prmim][TPPMS]. The latter one was used for its easiness in crystallization and purication. They performed optimization reactions in order to obtain higher rates. They investigated the effect of the ionic liquid, substrate ow rate, temperature, gas composition on the rate of hydroformylation of 1-octene. Among several ionic liquids used ([bmim][PF6 ] [bmim][NTf2 ], [octmim][PF6 ], [octmim][NTf2 ], [decmim][NTf2 ]), [octmim]NTf2 was found to be the most effective in terms of high conversion rates attained (>80%).

172

S. Keskin et al. / J. of Supercritical Fluids 43 (2007) 150180

The authors found that the catalyst turnover frequency (TOF), which indicates productivity, could be >500 h1 as the substrate (1-octene) ow rate was increased (>0.3 mL/min) at 200 bar and 100 C. They concluded that continuous ow homogeneous catalysis in ILscCO2 biphasic system can be used for hydroformylation of long chain alkenes at rates comparable with the ones found in commercial systems. In the latest study conducted by Cole-Hamilton and co-workers [158] on hydroformylation of alkenes, they achieved rapid hydroformylation of 1-octene (rates up to 800 h1 ) with the catalyst remaining stable for at least 40 h and with very low rhodium leaching levels (0.5 ppm). A new system, in which the substrate, reacting gases and products dissolved in scCO2 and were owing over a xed bed supported ionic liquid phase (SILP) catalyst, was introduced. The reactants in scCO2 (1-octene, CO and H2 ) owed upwards through a tubular reactor containing a catalyst composed of [prmim][Ph2 P(3-C6 H4 SO3 )], [Rh(acac)(CO)2 ] and [octmim][Tf2 N] supported on silica gel at 100 C, and 100 bar. The authors claimed that this new system provided excellent diffusion of the substrate and gases to the catalyst surface, excellent solubility of the substrates and gases within the supported ionic liquid and extraction of heavy products that might otherwise foul the catalyst by lling the pores. There were two other studies on SILP catalysis [171,172], but both studies were conducted in batch mode. Wang et al. [171] described the synthesis of cyclic carbonates from CO2 and epoxides over silica-supported quaternary ammonium salts and Ciriminna et al. [172] studied the aerobic oxidation of alcohols over a perruthenate catalyst. Brown et al. [159] studied the asymmetric hydrogenation of tiglic acid catalyzed by Ru(O2 CMe)2 ((R)tolBINAP) in [bmim]PF6 with addition of water as cosolvent. In this study, scCO2 was not involved in the reaction, but it was used to recover the products from the IL after the reaction time was over. The results were evaluated in terms of enantioselectivity and conversion. The catalyst/ionic liquid solution was reused repeatedly (ve cycles) without signicant loss of enantioselectivity or conversion. In another work of the same group [160], a number of different solvents were studied in order to evaluate the most effective system for the enantioselective hydrogenation of , -unsaturated acids. Different solvent systems comprised scCO2 , different ILs, ILs with cosolvents and also CO2 expanded ionic liquids. They studied two types of substrates, namely class I (atropic acid) and class II (tiglic acid) substrates. Class I substrates were hydrogenated in high enantioselectivity at high H2 concentration whereas Class II substrates were hydrogenated in high enantioselectivity at low H2 concentration. For both substrates, Ru(O2 CMe)2 (RtolBINAP) was used as the catalyst. Atropic acid was hydrogenated with the highest enantioselectivity (92%) in methanol (50 bar H2 pressure). Lower enantioselectivity values were obtained for reactions in ILs ([bmim][PF6 ]32%, [bmim][BF4 ]15%, [emim][O3 SCF3 ]25%, [emim][N(O2 SCF3 )2 ]31%, [dmpim][N(O2 SCF3 )2 ]39%) and ILcosolvent systems [bmim][PF6 ]PrOH33%, ([bmim][PF6 ]toluene19%, [bmim][PF6 ]MeOH54%) at 50 bar H2 pressure. [bmim] [PF6 ], [bmim][PF6 ]toluene, [bmim][PF6 ]PrOH, [bmim]

[PF6 ]MeOH systems were expanded by CO2 (50 bar CO2 ). Expansion caused increases in enantioselectivity consistent with greater H2 solubility and mass transfer rate in all mediums except for [bmim][PF6 ]toluene. Tiglic acid was hydrogenated with the highest enantioselectivity (95%) in [emim]N(O2 SCF3 )2 (5 bar H2 ). Reasonably high enantioselectivity values were also obtained for other ionic liquids, as the authors expected, because of low H2 concentration ([dmpim][N(O2 SCF3 )2 ]93%, [bmim][PF6 ]93%, [mbpy][BF4 ]88%, [bmim][BF4 ] 88%, [emim][O3 SCF3 ]84%). ILcosolvent systems ([bmim][PF6 ]toluene, [bmim][PF6 ]PrOH) were less selective than hydrogenation in IL ([bmim][PF6 ]) alone. Tiglic acid hydrogenations were also conducted in [bmim][PF6 ], [bmim][PF6 ]toluene, [bmim][PF6 ]PrOH expanded systems. In these CO2 (70bar) expanded systems, decreases in enantioselectivity were observed when compared to non-expanded systems. The decrease in enantioselectivity for [bmim][PF6 ] system was from 93% (non-expanded) to 85% (expanded). Another example of hydrogenation reaction in ILscCO2 biphasic system was reported by Liu et al. [161]. IL phase was used to immobilize the transition metal catalyst, and scCO2 phase was used to recover the products. They examined the hydrogenation of 1-decene and cyclohexene using Wilkinsons catalyst RhCl(PPh3 )3 , and hydrogenation of carbon dioxide in the presence of dialkylamines to form N,N-dialkylformamides using RuCl2 (dppe)2 in [bmim][PF6 ]scCO2 biphasic system. 98% conversion was attained for hydrogenation of 1-decene at 48 bar H2 and 207 bar total pressure (TOF: 410 h1 ) at the end of 1 h reaction time. Hydrogenation of cyclohexene under the same experimental conditions mentioned above proceeded more slowly (82% conversion after 2 h, 96% conversion after 3 h). They also performed four repetitive batch runs for hydrogenation of 1-decene, and demonstrated the efcient catalyst recycling via immobilization in ionic liquid. The hydrogenation of 1-decene and cyclohexene were also done with [bmim][PF6 ]n-hexane biphasic system. The results showed that there was no reactivity advantage for CO2 over n-hexane for simple hydrogenation reactions. The conversion and selectivity were much higher for hydrogenation of carbon dioxide in [bmim][PF6 ]scCO2 biphasic medium starting with the di-n-propylformamide (80 C, 276 bar, 5 h), when they were compared to the conversion and selectivity values obtained only in scCO2 for less bulky diethylamine and n-propylamine. B ssmann et al. [162] also utilized the benets of IL/CO2 o biphasic system to immobilize the organometallic catalyst in IL phase and recover the product without harming the catalyst. They investigated the continuous ow system for the hydrovinylation of styrene with Wilkes catalyst in IL/CO2 biphasic system. They initially identied the suitable ILs that would allow the activation of the catalyst. It was found that the activation strongly depended on the ILs anion type. The reaction conversion rates for the anions BARF (BARF: tetrakis[3,5-bis(triuoromethyl)phenyl]borate), Al[OC(CF3 )2 Ph]4 , Tf2 N and BF4 were 100, 90.5, 69.9 and 39.6, respectively for [emim] cation. It was also noted that the activation of the catalyst was not just a simple anion-exchange reaction, and the specic environment of the ionic solvent sys-

S. Keskin et al. / J. of Supercritical Fluids 43 (2007) 150180

173

tem seemed to activate the catalyst. When [emim] and [4-mbp] (4-mbp: 1-n-butyl-4-methylpyridinium) were used as different cations with the same anion, it was reported that at comparable conversions, higher enantioselectivity values were found with [4-mbp][BF4 ] and [4-mbp][Tf2 N] than the corresponding [emim] salts. The continuous ow styrene hydrovinylation was conducted in [emim][Tf2 N]CO2 at 80 bar. The catalyst was stable over a reaction time of 61 h and enantioslectivity decreased slightly at that period, while products were extracted with compressed CO2 . Tkatchenko et al. [163] studied the palladium catalyzed dimerization of methyl methacrylate at 83 C and 200 bar in [bmim][BF4 ]scCO2 biphasic system. The selectivity (>98%) was identical to that of the monobasic system. The turnover frequency and turnover numbers were comparable for both monophasic and biphasic systems, increasing with increasing substrate to palladium ratios in the IL phase. They concluded that, as the CO2 -rich phase acted as a substrate and product reservoir, there was a possibility for the reaction to be conducted under continuous feed and product recovery conditions with greener solvents. Hou et al. [164] investigated the oxidation of 1-hexene by molecular oxygen in [bmim][BF6 ]scCO2 biphasic system as well as in [bmim][BF6 ], scCO2 , and in the absence of solvent with catalysts PdCl2 and CuCl2 . The results showed that the selectivity to the desired product 2-hexanone was much higher when the reaction was done in [bmim][BF6 ]scCO2 biphasic system (125 bar, 333.2 K and 17 h of reaction time). Additionally, the catalyst was more stable in biphasic system than it was in scCO2 only. Kawanami et al. [165] performed the chemical xation of CO2 to cyclic carbonates in a ILscCO2 biphasic system. They reported that synthesis of propylene carbonate from propylene oxide and carbon dioxide in [omim][BF4 ]scCO2 biphasic system (14 MPa, 100 C) was achieved with nearly 100% yield and selectivity within 5 min and TOF value was 77 times higher than those so far reported. They also observed that the yield was increased as the alkyl chain length of the ILs cation was also increased from C2 to C8 . Moreover, different epoxide substrates having phenyl substituted groups and alkyl side chain groups were examined for the synthesis of the corresponding carbonates in [omim][BF4 ]scCO2 biphasic system at 14 MPa and 100 C. Gao et al. [166] studied the transesterication between isoamyl acetate and ethanol in scCO2 , [bmim][BF6 ] and [bmim][BF6 ]scCO2 system using p-xylenesulfonic acid (p-TSA) as the catalyst. The results showed that the equilibrium conversion in [bmim][BF6 ]scCO2 media was lower than those observed in scCO2 or [bmim][BF6 ]. An interesting application of ILscCO2 biphasic system in synthesis was electro-oxidation of benzyl alcohol to benzaldehyde in an electrochemical cell [167]. The reaction was carried out at 318.2 K and up to 10.3 MPa, and two ILs, 1-butyl-3-methylimidazolium tetrauoroborate ([bmim][BF4 ]) and 1-butyl-3-methylimidazolium hexauorophosphate ([bmim][PF6 ]), were used as the solvents and electrolytes. [bmim][BF4 ] was a better medium for the electro-oxidation of benzyl alcohol. The Faradic efciency (FE) and selectivity of benzaldehyde increased as the pressure increased up to 9.3 MPa, whereas the FE decreased as the pressure was increased further. The authors noted that the products

could be easily recovered from the IL by using scCO2 extraction after the electrolysis, and the IL could be reused. Yoon et al. [168] studied Heck coupling of iodobenzene with olens in [bmim][PF6 ] catalyzed by PdCl2 /Et3 N. They did not incorporate scCO2 in the reaction, it was used after the reaction for product recovery. ILscCO2 biphasic systems were also used for enzyme catalysis [173179]. Lozano et al. [173174] investigated the synthesis of butyl butyrate from vinyl butyrate and 1-butanol, and the kinetic resolution of rac-1-phenylethanol with vinyl propionate by transesterication. They used both free and immobilized C. antarctica lipase B (CALB) in IL ([emim][Tf2 N] and [bmim][Tf2 N]) as catalyst for continuous biphasic biocatalysis. CO2 was used as transport medium for reactants and products. The synthetic activity of the enzyme in ILCO2 system was tested through operation/storage cycles. Operation period (4 h) was followed by a storage period (20 h) of the enzymeIL system under dry conditions at room temperature. The continuous synthesis of butyl butyrate from vinyl butyrate and 1-butanol by transesterication was studied at 12.5 and 15 MPa at 40, 50 and 100 C in [bmim][Tf2 N]CO2 system [173]. The specic activity and the selectivity were enhanced as the temperature increased, the selectivity was high (>95%) in all cases giving higher than 50% conversions. The activity decay for the repetitive uses of the free enzymeIL system was the highest at high temperature. The continuous synthesis of (R)-1-phenylethyl propionate from the kinetic resolution of rac1-phenylethanol with vinyl propionate catalyzed by free CALB dissolved in [emim][Tf2 N] and [bmim][Tf2 N] was also investigated at 15 MPa, 50 and 100 C [173]. The results showed that the selectivity of the reaction, the activity and the halflife of the enzymeIL system were lower for this reaction than those observed for butyl butyrate synthesis. However, high enantioselectivity (>99.9%) was exhibited by the enzyme. As the reaction temperature increased, the selectivity was increased but the specic activity decreased. In another publication of Lozano et al. [174], the continuous kinetic resolution of rac1-phenylethanol in ILCO2 biphasic system was investigated at 120 and 150 C and 10 MPa. Both free and immobilized CALB dispersed in [emim][Tf2 N] and [bmim][Tf2 N] were used as catalyst. [emim][Tf2 N] was a better IL in terms of obtaining higher initial synthetic activity and longer half-life time of the free enzymeIL system. Immobilized enzymeIL system had longer half-life time and higher synthetic activity compared to free enzymeIL system. No activity losses during successive operation cycles were observed for immobilized enzyme at 120 C and 10 MPa. Additionally, at the same operating conditions (120 C, 10 MPa), in immobilized CALB[emim][Tf2 N] system, high selectivity values (>98%) were always obtained for successive operation cycles, whereas in free CALB[emim][Tf2 N] system, the selectivity increased from 36 to 98.5% as the number of operation cycles increased. Reetz et al. [175] studied the acylation of 1-octanol by vinyl acetate (batch and continuous mode) and kinetic resolution of 1-phenylethanol (batch mode), which were catalyzed by CALB in [bmim][BTA]CO2 system (BTA: bis(triuoromethanesulfonamide)). The acylation of 1-octanol

174

S. Keskin et al. / J. of Supercritical Fluids 43 (2007) 150180

by vinyl acetate in continuous mode was performed at 10.5 MPa and 45 C. A total yield of 93.9% was obtained after 24 h. Reetz et al. [180] introduced a new method for enantiomer separation utilizing biocatalytic kinetic resolution and SCF extraction using an ILscCO2 system. Application of this new method for separation of racemic secondary alcohols by CALB catalysis and CO2 extraction was represented. They converted several alcohols (2-octanol, 1-phenyl-ethanol, 3-methyl-2-butanol, 1-(2-phenylethyl) ethanol, and 1-(2-napthyl) ethanol) enantioselectively to corresponding acetates and laureates by immobilized CALB suspended in [bmim][BTA], and separate (S)-alcohol from the product (R)-ester via CO2 . Vinyl laureate was found to make the ester less soluble in CO2 than the alcohol, which allows for efcient separation, therefore it was used as the acylation agent. In that work, both batch and continuous modes of operation were studied. In batch mode, when vinyl laureate was used as the acylation agent, in the early fractions of isobaric extraction, the alcohol was extracted with high selectivity, whereas the lauryl ester was obtained with high purity in the later fractions. In order to speed up the extraction for the ester-rich fraction, they suggested increasing the CO2 density after most of the alcohol has been isolated. This procedure was successful for the separation of several alcohols and lauryl esters. Additionally, they include a separation chamber leading to a two-step extraction procedure. First, extraction was done at 60 C and 105 bar, then in the separation chamber the pressure was reduced to 80 bar and then the CO2 was vented through cryo-traps. By this procedure 66% of the theoretical amount of the alcohol under investigation was isolated with an enantiomeric purity >99.9% and less than 0.5% contamination with its corresponding lauryl esters. Then the pressure of reactor and the extraction chamber was increased to 200 bar and 89% of the theoretical amount of lauryl ester was isolated with an enantiomeric purity >99.9% and less than 1% contamination with alcohol. Kinetic resolution of rac-1-phenylethanol was used to demonstrate the continuous process. The separation after extraction was provided through two separation chambers with two steps of density reduction. After 112 h operation, 81% of the theoretical amount of the (S)1-phenylethanol was isolated with an enantiomeric purity >97% and less than 0.1% contamination with its corresponding lauryl ester. Lozano et al. [176] studied the synthesis of glycidyl esters from rac-glycidol catalyzed with free and immobilized forms of lipases from C. antarctica (CALA and CALB) and Mucor miehei (MML) in toluene, IL and ILscCO2 (40, 50 C and 100, 150 bar). Four different ILs were used: [emim][NTf2 ], [bmim][PF6 ], [bmim][NTf2 ], and trioctylmethylammonium triimide ([troma][Tf2 N]). Using ILs instead of a classical organic solvent (toluene) and the increase in the alkyl chain length of the acylation agent had both positive effect on the enzyme activity and when these effects combined, the synthetic activity can be enhanced 95 times. CALA and MML favored the synthesis of R-glycidyl esters, whereas CALB favored the synthesis of Sglycidyl esters. CALB showed the highest activity among other enzymes. CO2 was used to transport the substrates and products. The synthetic activity of the free and immobilized lipases decreased in ILscCO2 biphasic system, but the enantioselec-

tivity remained unchanged with respect to the values obtained in only IL reaction media. CALB catalyzed ester synthesis in ILscCO2 biphasic systems with ve different ILs ((3-hydroxypropyl)-trimethylammonium bis(triuoromethylsulfonyl) imide [C3 OHtma] [NTf2 ]; (3-cyanopropyl)-trimethylammonium bis(triuoromethylsulfonyl) imide [C3 CNtma][NTf2 ]; butyl-trimethylammonium bis(triuoromethylsulfonyl) imide, [C4 tma][NTf2 ]; (5cyanopentyl)-trimethylammonium bis(triuoromethylsulfonyl) imide [C5 CNtma][NTf2 ]; hexyl-trimethylammonium bis(triuoromethylsulfonyl) imide, [C6 tma][NTf2 ]) were studied by Lozano et al. [177]. The suitability of these ILs as enzymatic reaction media was tested for the kinetic resolution of rac-phenylethanol by transesterication with vinyl propionate, and all of the tested ILs were found to be suitable media for enzyme catalysis. The synthetic activities and stabilities of the enzyme were determined in these ve ILs. Then, the performance of CALB catalysis in ILscCO2 biphasic system in continuous operation was tested for [C4 tma][NTf2 ]scCO2 and [C3 CNtma][NTf2 ]scCO2 systems at 50 C and 10 MPa for the synthesis of several short-chain alkyl esters (butyl acetate (BA), butyl propionate (BP), butyl butyrate (BB), hexyl propionate (HP), hexyl butyrate (HB), and octyl propionate (OP)), by transesterication from the respective vinyl alkyl esters with alkyl-1-ols. The results showed that the synthetic activity of CALB in [C4 tma][NTf2 ]scCO2 system was higher than that in [C3 CNtma][NTf2 ]scCO2 , even though the opposite result was obtained for activity values in pure IL. The authors concluded that rate-limiting parameters (synthetic activity and mass transfer phenomena between IL and scCO2 phases) were related with the solubility parameter of the ILs alkyl chain and reagents. Hernandez et al. [178] presented the synthesis of butyl propionate in scCO2 and ILscCO2 using a recirculating enzymatic membrane reactor, in which -alumina membranes were immobilized with CALB. The reactor was tested in only scCO2 at 50 C and 80 bar. In the second part of the study, the immobilized enzyme was coated with three different ILs, i.e. [bmim][PF6 ], 1-butyl-2,3-dimethylimidazolium hexauorophosphate [bdimim][PF6 ], [omim][PF6 ]. The selectivity was 95% when only scCO2 medium was utilized, and it increased to greater than 99.5% when ILscCO2 biphasic system was used. Lozano et al. [179] investigated the importance of the supporting material for the activity of immobilized CALB in ILhexane and ILscCO2 biphasic systems at controlled water activity. For this purpose they immobilized a commercial solution of free CALB onto twelve different silica supports, modied with specic side chains (e.g. alkyl, amino, carboxylic, nitrile, etc.) by adsorption. Both biphasic media was tested for the kinetic resolution of rac-phenylethanol by transesterication with vinyl propionate. CALB activity and stability was tested ILhexane and ILscCO2 biphasic systems with two different ILs: [btma][NTf2 ] and [toma][NTf2 ] (btma: butyltrimethylammonium, toma: trioctyl-methylammonium). The highest synthetic activity of immobilized enzyme was obtained for butyl-silica derivative, however for all supports, the selectivity for the reaction was higher than 94% except for the quaternary ammonium-Si support and the enantiomeric excess was greater

S. Keskin et al. / J. of Supercritical Fluids 43 (2007) 150180

175

Fig. 13. A process suggested for cleaning of soils using IL and scCO2 [181].

than 99.9%. The immobilized activity decreased 10 times in ILhexane compared to that in only hexane, but half-life times were enhanced by up to six times. The synthetic activity of immobilized CALB increased by six times in ILscCO2 compared to that in hexane. Recently, the potential of ILs to dissolve soil contaminants at ambient conditions and the ability of supercritical carbon dioxide (scCO2 ) to recover these contaminants from IL extracts were mutually utilized to clean contaminated soils [181]. Naphthalene was used as the model component to represent a group of soil contaminants, and 1-n-butyl 3-methylimidazolium hexauorophosphate ([bmim][PF6 ]) was used as the IL. The results demonstrated that soil contaminated with naphthalene was cleaned using [bmim][PF6 ], and the amount of naphthalene remaining in the soil was below the allowable contamination limit. This study was the rst in the literature which investigates the soil/model-contaminant/IL/scCO2 system. On the basis of the ndings a process owsheet for the IL extraction of contaminated soils and continuous scCO2 extraction of the contaminants from IL extracts was suggested and is given in Fig. 13.15

8. Summary This work aims to summarize and discuss the information present in the literature about ILs and ILCO2 systems. A great number of researchers investigated the high-pressure phase behavior of ILCO2 systems and concluded that CO2 is highly soluble in most ILs, and ILs are not measurably soluble in scCO2 . The effects of pressure, temperature, nature of the anion and the alkyl chain length of the cation on the solubility of CO2 reported by various studies are discussed and summarized here. Volatile and nonpolar scCO2 has become a good partner of nonvolatile and polar IL and this new system with its unique properties have been utilized to extract organic compounds from ILs using scCO2 . ILs are receiving more and more attention every day both in academic research and commercial applications and they seem as good replacements for volatile organic solvents. However, there is a discussion about the greenness of the ILs due to their incomplete physical, chemical and toxicological data. Although, there are some question marks related to the specic character-

176

S. Keskin et al. / J. of Supercritical Fluids 43 (2007) 150180 [17] J.G. Huddleston, H.D. Willauer, R.P. Swatloski, A.E. Visser, R.D. Rogers, Room temperature ionic liquids as novel media for clean liquidliquid extraction, Chem. Commun. (1998) 17651766. [18] C. Chiappe, D. Pieraccini, Review commentary; ionic liquids: solvent properties and organic reactivity, J. Phys. Org. Chem. 18 (2005) 275297. [19] D.C. Donata, F. Marida, H. Migen, University of Torino, http://lem. ch.unito.it/didattica/infochimica/Liquidi%20Ionici/Denition.html (accessed 20 June, 2006). [20] M.J. Earle, J.M.S.S. Esperanca, M.A. Gilea, J.N.C. Lopes, L.P.N. Rebelo, J.W. Magee, K.R. Seddon, J.A. Widegren, The distillation and volatility of ionic liquids, Nature 72 (2006) 13911398. [21] J. Dupont, C.S. Consorti, J. Spencer, Room-temperature molten salts: neoteric green solvents for chemical reactions and processes, J. Braz. Chem. Soc. 11 (2000) 337344. [22] F. Mutelet, V. Butet, J.N. Jaubert, Application of inverse gas chromatography and regular solution theory for characterization of ionic liquids, Ind. Eng. Chem. Res. 44 (2005) 41204127. [23] P. K lle, R. Dronskowski, Synthesis, crystal structures and electrical cono ductivities of the ionic liquid compounds butyl dimethylimidazolium tetrauoroborate, hexauoroborate and hexauoroantimonate, Eur. J. Inorg. Chem. (2004) 23132320. [24] A.J. Carmichael, K.R. Seddon, Polarity study of some 1-alkyl3-methylimidazolium ambient-temperature ionic liquids with the solvatochromic dye, Nile red, J. Phys. Org. Chem. 13 (2000) 591595. [25] S.N.V.K. Aki, J.F. Brennecke, A. Samanta, How polar are room temperature ionic liquids? Chem. Commun. (2001) 413414. [26] R.A. Sheldon, R.M. Lau, F. Sorgedrager, K. van Rantwijk, R. Seddon, Biocatalysis in ionic liquids, Green Chem. 4 (2002) 147151. [27] B. Jastorff, R. St rmann, J. Ranke, K. M lter, F. Stock, B. Oberheitmann, o o W. Hoffmann, J. Hoffmann, M. N chter, B. Ondruschka, J. Filser, How u sustainable are ionic liquids? Structureactivity relationships and biological testing as important elements for sustainability evaluation, Green Chem. 5 (2003) 136142. [28] D.R. MacFarlane, Ionic liquids symposium, Aust. J. Chem. 57 (2004) 111112. [29] J.S. Wilkes, Properties of ionic liquid solvents for catalysis, Mol. Catal. A-Chem. 214 (2004) 1117. [30] Merck Chemicals, http://pb.merck.de/servlet/PB/menu/1341610/index. html (accessed 26 June, 2006). [31] L. Cammarata, S.G. Kazarian, P.A. Salter, T. Welton, Molecular states of water in room temperature ionic liquids, Phys. Chem. Chem. Phys. 23 (2001) 51925200. [32] H.C. Hodges, J.H. Sterner, Combined tabulation of toxicity classes, in: W.S. Spector (Ed.), Handbook of Toxicilogy, 1, W.B. Saunders Company, Philadelphia, 1956. [33] E.J. Maginn, Research: Ionic Liquids, www.nd.edu/ed/Research/ IL toxicology.html (accessed 19 April, 2007). [34] A.S. Wells, V.T. Coombe, On the freshwater ecotoxicity and biodegradation properties of some common ionic liquids, Org. Process Res. Dev. 10 (2006) 794798. [35] R.P. Swatloski, J.D. Holbrey, S.B. Memon, G.A. Caldwell, K.A. Caldwell, R.D. Rogers, Using Caenorhabditis elegans to probe toxicity of 1-alkyl-3methylimidazolium chloride based ionic liquids, Chem. Commun. (2004) 668669. [36] R.J. Bernott, E.E. Kennedy, G.A. Lamberti, Effects of ionic liquids on the survival, movement, and feeding behavior of the freshwater snail, Physa acuta, Environ. Toxicol. Chem. 24 (2005) 17591765. [37] T.D. Landry, K. Brooks, D. Poche, M. Woolhiser, Acute toxicity prole of 1-butyl-3-methylimidazolium chloride, Bull. Environ. Contam. Toxicol. 74 (2005) 559565. [38] C. Pretti, C. Chiappe, D. Pieraccini, M. Gregori, F. Abramo, G. Monnia, L. Intorre, Acute toxicity of ionic liquids to the zebra sh (Danio rerio), Green Chem. 8 (2006) 238240. [39] P. Stepnowski, A. M ller, P. Behrend, J. Ranke, J. Hoffmann, B. Jastorff, u Reversed-phase liquid chromatographic method fort the determination of selected room-temperature ionic liquid cations, J. Chromatogr. A 993 (2003) 173178.

istics of ILs, it seems that most of the researchers will continue to work with this new solvent, the developments of new applications utilizing ILs will increase rapidly and the number of publications will rise exponentially in the future. The present costs of the ILs are quite prohibitive in many probable commercial applications. However, there are hopes that in the near future, the cost/benet gures of the ILs will bring economic viability to their more common use. Acknowledgements The nancial supports provided by Bo azici University g ITAK Research Foundation via the Project No. 06A502 and TUB (The Scientic and Technical Research Council of Turkey) via the Project No. 104M185 are gratefully acknowledged. The authors thank the three anonymous reviewers for their detailed and insightful comments and suggestions, which signicantly improved the manuscript. References
[1] T. Welton, Room temperature ionic liquids: Solvents for synthesis and catalysis, Chem. Rev. 99 (1999) 20712084. [2] P.L. Short, Out of the ivory tower, Chem. Eng. News 84 (2006) 1521. [3] S. Sugden, H. Wilkins, The parachor and chemical constitution. Part. Fused metals and salts, J. Chem. Soc. (1929) 12911298. [4] J. Gorman, Faster, better, cleaner? New liquids take aim at old-fashioned chemistry, Sci. News 160 (2001) 156158. [5] J.S. Wilkes, J.A. Levisky, R.A. Wilson, C.L. Hussey, Dialkylimidazolium chloroaluminate melts, a new class of room-temperature ionic liquids for electrochemistry, spectroscopy and synthesis, Inorg. Chem. 21 (1982) 12631264. [6] A.A. Fannin, D.A. Floreani, L.A. King, J.S. Landers, B.J. Piersma, D.J. Stech, R.L. Vaughn, J.S. Wilkes, J.L. Williams, Properties of 1,3dialkylimidazolium chloride aluminium chloride ionic liquids. Part 2, J. Phys. Chem. 88 (1984) 26142627. [7] J. Dupont, On the solid, liquid and solution structural organization of imidazolium ionic liquids, J. Braz. Chem. Soc. 15 (2004) 341 350. [8] J.F. Brennecke, E.J. Maginn, Ionic liquids: innovative uids for chemical processing, AIChE J. 47 (2001) 23842388. [9] R. Renner, Ionic liquids: an industrial cleanup solution, Environ. Sci. Technol. 35 (2001) 410A413A. [10] Q. Yang, D.D. Dionysiou, Photolytic degradation of chlorinated phenols in room temperature ionic liquids, J. Photochem. Photobiol. A: Chem. 165 (2004) 229240. [11] K.R. Seddon, Room-temperature ionic liquids: neoteric solvents for clean catalysis, Kinet. Catal. 37 (1996) 693697. [12] C. Lagrost, D. Carrie, M. Vaultier, P. Hapiot, Reactivities of some elec trogenerated organic cation radicals in room-temperature ionic liquids: toward an alternative to volatile organic solvents? J. Phys. Chem. A 107 (2003) 745752. [13] A. Shariati, C.J. Peters, High pressure phase equilibria of systems with ionic liquids, J. Supercrit. Fluids 34 (2005) 171176. [14] A. Shariati, K. Gutkowski, C.J. Peters, Comparison of the phase behavior of some selected binary systems with ionic liquids, AIChE J. 51 (2005) 15321540. [15] H. Zhao, S. Xia, P. Ma, Review: use of ionic liquids as green solvents for extractions, J. Chem. Technol. Biotechnol. 80 (2005) 10891096. [16] G.J. Kabo, A.V. Blokhin, A. Paulechka, U. Ya, A.G. Kabo, M.P. Shymanovich, J.V. Magee, Thermodynamic properties of 1-butyl-3methylimidazolium hexauorophosphate in the condensed state, J. Chem. Eng. Data 49 (2004) 453461.

S. Keskin et al. / J. of Supercritical Fluids 43 (2007) 150180 [40] N. Gathergood, M.T. Garcia, P.J. Scammells, Biodegradable ionic liquids. Part I. Concept, preliminary targets and evaluation, Green Chem. 6 (2004) 166175. [41] N. Gathergood, P.J. Scammells, Design and preparation of roomtemperature ionic liquids containing biodegradable side chains, Aust. J. Chem. 55 (2002) 557560. [42] E.B. Carter, S.L. Culver, P.A. Fox, R.D. Goode, I. Ntai, M.D. Tickell, R.K. Traylor, N.W. Hoffman, J.H. Davis Jr., Sweet success: ionic liquids derived from non-nutritive sweeteners, Chem. Commun. (2004) 630631. [43] M.T. Garcia, N. Gathergood, P.J. Scammells, Biodegradable ionic liquids. Part II. Effect of the anion and toxicology, Green. Chem. 7 (2005) 914. [44] G.-H. Tao, L. He, N. Sun, Y. Kou, New generation ionic liquids: cations derived from amino acids, Chem. Commun. (2005) 35623564. [45] N. Gathergood, P.J. Scammells, M.T. Garcia, Biodegradable ionic liquids. Part III. The rst readily biodegradable ionic liquids, Green Chem. 8 (2006) 156160. [46] G.-H. Tao, L. He, W.-S. Liu, L. Xu, W. Xiong, T. Wang, Y. Kou, Preparation, characterization and application of amino acid-based green ionic liquids, Green Chem. 8 (2006) 639646. [47] P. Wasserscheid, R. van Hal, A. B smann, 1-n-Butyl-3-methylimio dazolium ([bmim]) octylsulfatean even greener ionic liquid, Green Chem. 4 (2002) 400404. [48] C.D. Tran, S.H.D.P. Lacerda, D. Oliveira, Absorption of water by room temperature ILs: effect of anions on concentration and state of water, Soc. Appl. Spectrosc. 57 (2003) 152157. [49] J.G. Huddleston, A.E. Visser, W.M. Reichert, H.D. Willauer, G.A. Broker, R.D. Rogers, Characterization and comparison of hydrophilic and hydrophobic room temperature ionic liquids incorporating the imidazolium cation, Green Chem. 3 (2001) 156164. [50] D.C. Donata, F. Marida, H. Migen, University of Torino, http://lem.ch. unito.it/didattica/infochimica/Liquidi%20Ionici/Composition.html (accessed 20 June, 2006). [51] S.K. Poole, P.H. Shetty, C.F. Poole, Chromatographic and spectroscopic studies of the solvent properties of a new series of room-temperature liquid tetraalkylammonium sulfonates, Anal. Chim. Acta 218 (1989) 241264. [52] J.S. Wilkes, M.J. Zaworotko, Air and water stable 1-ethyl-3methylimidazolium based ionic liquids, J. Chem. Soc. Chem. Commun. (1992) 965967. [53] J.D. Holbrey, R.D. Rogers, Clean synthesis of 1,3-dialkylimidazolium ionic liquids, Abstr. Pap. Am. Chem. Soc. 224 (2002) U612. [54] J.D. Holbrey, W.M. Reichert, R.P. Swatloski, G.A. Broker, W.R. Pitner, K.R. Seddon, R.D. Rogers, Efcient, halide free synthesis of new low cost ionic liquids: 1,3-dialkylimidazolium salts containing methyl- and ethyl-sulfate anions, Green Chem. 4 (2002) 407413. [55] Y.R. Mirzaei, B. Twamley, J.M. Shreeve, Syntheses of 1-alkyl-1,2,4triazoles and the formation of quaternary 1-alkyl-4-polyuoroalkyl1,2,4-triazolium salts leading to ionic liquids, J. Org. Chem. 67 (2002) 93409345. [56] W.L. Bao, Z.M. Wang, Y.X. Li, Synthesis of chiral ionic liquids from natural amino acids, J. Org. Chem. 68 (2003) 591593. [57] A.E. Visser, R.P. Swatloski, W.M. Reischert, R. Mayton, S. Sheff, A. Wierzbicki, J.H. Davis, R.D. Rogers, Task-specic ionic liquids incorporating novel cations for the coordination and extraction of Hg2+ and Cd2+ : synthesis, characterization, and extraction studies, Environ. Sci. Technol. 36 (2002) 25232529. [58] R.A. Sheldon, in: M.P.C. Weijnen, A.A.H. Drienkenburg (Eds.), Precision Process Technology: Perspectives for Pollution Prevention, Kluwer, Dordrecht, 1993, p. 125. [59] A.G. Fadeev, M.M. Meagher, Opportunities for ionic liquids in recovery of biofuels, Chem. Commun. (2001) 295296. [60] M.S. Selvan, M.D. McKinley, R.H. Dubois, J.L. Atwood, Liquidliquid equilibria for toluene + heptane + 1-ethyl-3-methylimidazolium triiodide and toluene + heptane + 1-butyl-3-methylimidazolium triiodide, J. Chem. Eng. Data 45 (2000) 841845. [61] T.M. Letcher, N. Deenadayalu, B. Soko, D. Ramjugernath, P.K. Naicker, Ternary liquidliquid equilibria for mixtures of 1-methyl-3octylimidazolium chloride + an alkanol + an alkane at 298.2 K and 1 bar, J. Chem. Eng. Data 48 (2003) 904907.

177

[62] W. Wu, K.N. Marsh, A.V. Deev, J.A. Boxall, Liquidliquid equilibria of room temperature ionic liquids and butanol, J. Chem. Eng. Data 48 (2003) 486491. [63] M. Wagner, O. Stanga, W. Schroer, Corresponding states analysis of the critical points in binary solutions of room temperature ionic liquids, Phys. Chem. Chem. Phys. 5 (2003) 39433950. [64] U. Domanska, A. Marciniak, Solubility of 1-alkyl-3-methylimidazolium hexauorophosphate in hydrocarbons, J. Chem. Eng. Data 48 (2003) 451456. [65] A. Arce, A. Marchiaro, O. Rodriguez, A. Soto, Essential oil terpenless by extraction using organic solvents or ionic liquids, AIChE J. 52 (2006) 20892097. [66] D. Camper, P. Scovazzo, C. Koval, R. Noble, Gas solubilities in roomtemperature ionic liquids, Ind. Eng. Chem. Res. 43 (2004) 30493054. [67] P. Scovazzo, D. Camper, J. Kieft, J. Poshusta, C. Koval, R. Noble, Regular solution theory and CO2 gas solubility in room-temperature ionic liquids, Ind. Eng. Chem. Res. 43 (2004) 68556860. [68] R.E. Baltus, B.H. Culbertson, S. Dai, H.M. Luo, D.W. DePaoli, Lowpressure solubility of carbon dioxide in room-temperature ionic liquids measured with a quartz crystal microbalance, J. Phys. Chem. B 108 (2004) 721727. [69] J.L. Anthony, E.J. Maginn, J.F. Brennecke, Solubilities and thermodynamic properties of gases in the ionic liquid 1-n-butyl-3methylimidazolium hexauorophosphate, J. Phys. Chem. B 106 (2002) 73157320. [70] L.A. Blanchard, Z. Gu, J.F. Brennecke, High pressure phase behavior of ionic liquid/CO2 systems, J. Phys. Chem. B 105 (2001) 2437 2444. [71] J.L. Anthony, E.J. Maginn, J.F. Brennecke, Solution thermodynamics of imidazolium-based ionic liquids and water, J. Phys. Chem. B 105 (2001) 1094210949. [72] J. Kumelan, A.P.-S. Kamps, D. Tuma, G. Maurer, Solubility of H2 in the ionic liquid [bmim][PF6 ], J. Chem. Eng. Data 51 (2006) 1114. [73] S. Dai, D. DePaoli, M. Dietz, J. Mays, J. McFarlane, W. Steele, Technical summaries on ionic liquids in chemical processing, Chemical Industry Vision 2020 Technology Partnership Workshop, 2003. [74] Air Products and Chemicals Inc., www.airproducts.com/NR/rdonlyres/ 9D919AAD-DE4F-4614-B3C4-E923C6BF786E/0/GasguardSAS.pdf, 2005. [75] J. Dupont, G.S. Fonseca, A.P. Umpierre, P.F.P. Fichtner, S.R. Teixeira, Transition-metal nanoparticles in imidazolium ionic liquids: recycable catalysts for biphasic hydrogenation reactions, J. Am. Chem. Soc. 124 (2002) 42284229. [76] R. Morland, American Chemical Society Division of Industrial and Engineering Chemistry, in: Proceedings of the 221st American Chemical Society National Meeting, San Diego, 2001. [77] I.I. Abuadoun, Photoinitiated cationic polymerization by imidazolium salts, Abstr. Pap. Am. Chem. Soc. 198 (1989) p.96. [78] P. Wasserscheid, Potential to apply ionic liquids in industry: exemplied for the use as solvents in industrial applications as homogenous catalysis, in: Green Industrial Applications of Ionic Liquids, NATO Science Series II: Mathematics, Physics and Chemistry 92, Kluwer Academic Publishers, Dordecht, 2003, pp. 2947. [79] E. Ota, Some aromatic reactions using AlCl3 -rich molten salts., J. Electrochem. Soc. 134 (1987), C512. [80] T. Nemeth, L. Bricker, C.J. Holmgren, S.J. Monson, E. Lyle, Direct carbonylation of parafns using an ionic liquid catalyst, US Patent 628828 (2001). [81] M.J. Earle, S.P. Katdare, Oxidative halogenation of aromatic compounds in the presence of an ionic liquid, World Patent WO 2002030852 (2002). [82] M.J. Earle, S.P. Katdare, Oxidation of alkylaromatics in the presence of ionic liquids, World Patent WO 2002030862 (2002). [83] M.J. Earle, S.P. Katdare, Aromatic nitration reactions in ionic liquids, World Patent WO 2002030865 (2002). [84] M.J. Earle, S.P. Katdare, Aromatic sulfonation reactions conducted in the presence of ionic liquids, World Patent WO 2002030878 (2002). [85] H. Olivier-Bourbigou, L. Magna, Ionic liquids: Perspectives for organic and catalytic reactions, J. Mol. Catal. A 182183 (2002) 419437.

178

S. Keskin et al. / J. of Supercritical Fluids 43 (2007) 150180 tion by crown ethers in room temperature ionic liquids, Ind. Eng. Chem. Res. 39 (2000) 35963604. S. Chun, S.V. Dzyuba, R.A. Bartsch, Inuence of structural variation in room temperature ionic liquids on the selectivity and efciency of competitive alkali metal salt extraction by a crown ether, Anal. Chem. 73 (2001) 37373741. R.A. Bartsch, S. Chun, S.V. Dzyuba, Ionic Liquids as Novel Diluents for Solvent Extraction of Metal Salts by Crown Ethers, Ionic Liquids Industrial Applications for Green Chemistry, American Chemical Society, Washington, DC, 2002, pp. 5868. G.-T. Wei, Z. Yang, C.J. Chen, Room temperature ionic liquid as a novel medium for liquid/liquid extraction of metal ions, Anal. Chim. Acta 488 (2003) 183192. H. Luo, S. Dai, P.V. Bonnesen, A.C.I. Buchanan, J.D. Holbrey, N.J. Bridges, R.D. Rogers, Extraction of cesium ions from aqueous solutions using calix[4]arene-bis(tert-octylbenzocrown-6) in ionic liquids, Anal. Chem. 76 (2004) 30783083. H. Luo, S. Dai, P.V. Bonnesen, Solvent extraction of Sr2+ and Cs+ based on room temperature ionic liquids containing monoaza-substituted crown ethers, Anal. Chem. 76 (2004) 27732779. A.E. Visser, R.P. Swatloski, W.M. Reischert, R. Mayton, S. Sheff, A. Wierzbicki, J.H. Davis, R.D. Rogers, Task-specic ionic liquids for the extraction of metal ions from aqueous solutions, Chem. Commun. (2001) 135136. M. Wagner, M. Uerdingen, in: B. Cornils, W.A. Herrmann, I.T. Horvath, W. Leitner, S. Mecking, H. Olivier-Bourbigou, D. Vogt (Eds.), Multiphase Homogeneous Catalysis, Wiley-VCH, Weinheim, 2005. P. Wassersheid, M. Haumann, in: D.J. Cole-Hamilton, R.P. Tooze (Eds.), Catalyst Separation, Recovery and Recycling: Chemistry and Process Design, Springer, Dordrecht, 2006. K.R. Harris, M. Kanakubo, L.A. Woolf, Temperature and pressure dependence of the viscosity of the ionic liquids: 1-methyl-3-octylimidazolium hexauorophosphate and 1-methyl-3-octylimidazolium tetrauoroborate, J. Chem. Eng. Data 51 (2006) 11611167. L.A. Blanchard, J.F. Brennecke, Recovery of organic products from ionic liquids using supercritical carbon dioxide, Ind. Eng. Chem. Res. 40 (2001) 287292. A.P.-S. Kamps, D. Tuma, J. Xia, G. Maurer, Solubility of CO2 in the ionic liquid [bmim][PF6 ], J. Chem. Eng. Data 48 (2003) 746749. Z.M. Liu, W.Z. Wu, B.X. Han, Z.X. Dong, G.Y. Zhao, J.Q. Wang, T. Jiang, G.Y. Yang, Study on the phase behaviors, viscosities, and thermodynamic properties of CO2 [C4 mim][PF6 ]methanol system at elevated pressures, Chem. Eur. J. 9 (2003) 38973903. S.N.V.K. Aki, B.R. Mellein, E.M. Saurer, J.F. Brennecke, High-pressure phase behavior of carbon dioxide with imidazolium-based ionic liquids, J. Phys. Chem. B 108 (2004) 2035520365. C. Cadena, J.L. Anthony, J.K. Shah, T.I. Morrow, J.F. Brennecke, E.J. Maginn, Why is CO2 so soluble in imidazolium-based ionic liquids? J. Am. Chem. Soc. 126 (2004) 53005308. Y.S. Kim, W.Y. Choi, J.H. Jang, K.-P. Yoo, C.S. Lee, Solubility measurement and prediction of carbon dioxide in ionic liquids, Fluid Phase Equilib. 228229 (2005) 439445. M.B. Shiett, A. Yokozeki, Solubilities and diffusivities of carbon dioxide in ionic liquids:[bmim][PF6 ] and [bmim][BF4 ], J. Am. Chem. Soc. 44 (2005) 44534464. L.A. Blanchard, D. Hancu, E.J. Beckman, J.F. Brennecke, Green processing using ionic liquid and carbon dioxide, Nature 399 (1999) 2829. M.B. King, A. Mubarak, J.D. Kin, T.R. Bott, The mutual solubilities of water with supercritical and liquid carbon dioxides, J. Supercrit. Fluids 5 (1992) 296302. K.L. Toews, R.M. Scholl, N.G. Smart, pH-dening equilibrium between water and supercritical CO2 : inuence on SFE of organics and metal chelates, Anal. Chem. 67 (1995) 40404043. S.R. Rubero, S. Baldelli, Inuence of water on the surface of hydrophilic and hydrophobic room temperature ILs, J. Am. Chem. Soc. 126 (2004) 1178811789. A. Shariati, C.J. Peters, High pressure phase behavior of systems with ionic liquids: II. The binary system carbon dioxide + 1-ethyl-3-

[86] B. Cornils, W.A. Herrmann, Concepts in homogenous catalysis: the industrial view, J. Catal. 216 (2003) 2331. [87] C.J. Adams, M.J. Earle, G. Roberts, K.R. Seddon, Friedel-Crafts reactions in room temperature ionic liquids, Chem. Commun. (1998) 20972098. [88] M.J. Earle, K.R. Seddon, Ionic liquids: green solvents for future, Pure Appl. Chem. 72 (2000) 13911398. [89] C.M. Gordon, New developments in catalysis using ionic liquids, Appl. Catal. A 222 (2001) 101117. [90] J.D. Holbrey, K.R. Seddon, Review: ionic liquids, Clean Prod. Process. (1999) 223236. [91] P.A.Z. Suarez, J.E.L. Dullis, S. Einloft, R.F. de Souza, J. Dupont, The use of ionic liquids in two phase catalytic hydrogenation reaction by rhodium complexes, Polyhedron 15 (1996) 12171219. [92] P.A.Z. Suarez, J.E.L. Dullis, S. Einloft, R.F. de Souza, J. Dupont, Twophase catalytic hydrogenation of olens by Ru(II) and Co(II) complexes dissolved in 1-n-butyl-3-methylimidazolium tetrauoroborate ionic liquid, Inorg. Chim. Acta 255 (1997) 207209. [93] Q. Liu, M.H.A. Janssen, F. van Rantwijk, R.A. Sheldon, Room temperature ionic liquids that dissolve carbohydrates in high concentrations, Green Chem. 7 (2005) 3942. [94] M. Matsumoto, K. Mochiduki, K. Fukunishi, K. Kondo, Extraction of organic acids using imidazolium-based ionic liquids and their toxicity to Lactobacillus rhamnosus, Sep. Purif. Technol. 40 (2004) 97101. [95] R.M. Lau, F. van Rantwijk, K.R. Seddon, R.A. Sheldon, Lipase-catalyzed reactions in ionic liquids, Org. Lett. 2 (2000) 41894191. [96] S. Park, R.J. Kazlauskas, Improved preparation and use of roomtemperature ionic liquids in lipase-catalyzed enantioand regioselective acylations, J. Org. Chem. 66 (2001) 83958401. [97] R.P. Swatloski, S.K. Scott, J.D. Holbrey, R.D. Rogers, Dissolution of cellulose with ionic liquids, J. Am. Chem. Soc. 124 (2002) 4974 4975. [98] K. Przybysz, E. Drzewinska, A. Stanislawska, A.W. Robak, Ionic liquids and paper, Ind. Eng. Chem. Res. 44 (2005) 45994604. [99] H. Pfruender, R. Jones, D.W. Botz, Water immiscible ionic liquids as solvents for whole cell biocatalysis, J. Biotechnol. 124 (2006) 182190. [100] S. Dai, Y.H. Ju, C.E. Barnes, Solvent extraction of strontium nitrate by a crown ether using room temperature ionic liquids, J. Chem. Soc. Dalton Trans. (1999) 12011202. [101] J.H.J. Davis, Task-specic ionic liquids, Chem. Lett. 33 (2004) 10721077. [102] R. De Waele, L. Heerman, W. DOlieslager, Electrochemistry of uranium(IV) in acidic AlCl3 + N-(n-butyl)pyridinium chloride roomtemperature molten salts, J. Electroanal. Chem. 142 (1982) 137146. [103] L. Heerman, R. De Waele, W. DOlieslager, Electrochemistry and spectroscopy of uranium in basic AlCl3 + N-(n-butyl) pyridinium chloride room temperature molten salts, J. Electroanal. Chem. 193 (1985) 289294. [104] P.B. Hitchcock, T.J. Mohammed, K.R. Seddon, J.A. Zora, C.L. Hussey, E.H. Ward, 1-Methyl-3-ethylimidazolium hexachlorouranate(IV) and 1-methyl-3-ethylimidazolium tetrachlorodioxo-uranate(VI): synthesis, structure, and electrochemistry in a room temperature ionic liquid, Inorg. Chim. Acta 113 (1986) L25L26. [105] S. Dai, Y.S. Shin, L.M. Toth, C.E. Barnes, Comparative UVVis studies of uranyl chloride complex in two basic ambienttemperature melt systems: the observation of spectral and thermodynamic variations induced via hydrogen bonding, Inorg. Chem. 36 (1997) 49004902. [106] G.M.N. Baston, A.E. Bradley, T. Gorman, I. Hamblett, C. Hardacre, J.E. Hatter, M.J.F. Healy, B. Hodgson, R. Lewin, K.V. Lovell, G.W.A. Newton, M. Nieuwenhuyzen, W.R. Pitner, D.W. Rooney, D. Sanders, K.R. Seddon, H.E. Simms, R.C. Thied, Ionic Liquids for the Nuclear Industry: A Radiochemical, Structural and Electrochemical Investigation, Ionic Liquids Industrial Applications for Green Chemistry, American Chemical Society, Washington, DC, 2002, pp. 162177. [107] K. Nakashima, F. Kubota, T. Maruyama, M. Goto, Feasibility of ionic liquids as alternative separation media for industrial solvent extraction processes, Ind. Eng. Chem. Res. 44 (2005) 43684372. [108] A.E. Visser, R.P. Swatloski, W.M. Reischert, S.T. Grifn, R.D. Rogers, Traditional extractants in nontraditional solvents: groups 1 and 2 extrac-

[109]

[110]

[111]

[112]

[113]

[114]

[115]

[116]

[117]

[118]

[119] [120]

[121]

[122]

[123]

[124]

[125] [126]

[127]

[128]

[129]

S. Keskin et al. / J. of Supercritical Fluids 43 (2007) 150180 methylimidazolium hexauorophosphate, J. Supercrit. Fluids 29 (2004) 4348. A. Shariati, C.J. Peters, High pressure phase behavior of systems with ionic liquids. Part III. The binary system carbon dioxide + 1-hexyl-3methylimidazolium hexauorophosphate, J. Supercrit. Fluids 30 (2004) 139144. P. Husson-Borg, V. Majer, M.F.C. Gomes, Solubilities of oxygen and carbon dioxide in butyl methyl imidazolium tetrauoroborate as a function of temperature and at pressures close to atmospheric pressure, J. Chem. Eng. Data 48 (2003) 480485. M.C. Kroon, A. Shariati, M. Costantini, J. van Spronsen, G.-J. Witkamp, R.A. Sheldon, C.J. Peters, High-pressure phase behavior of systems with ionic liquids. Part V. The binary system carbon dioxide + 1-butyl3-methylimidazolium tetrauoroborate, J. Chem. Eng. Data 50 (2005) 173176. M. Constantini, V.A. Toussaint, A. Shariati, C.J. Peters, I. Kikic, Highpressure phase behavior of systems with ionic liquids. Part IV. Binary system carbon dioxide + 1-hexyl-3-methylimidazolium tetrauoroborate, J. Chem. Eng. Data 50 (2005) 5255. S.G. Kazarian, N. Sakellarios, C.M. Gordon, High pressure CO2 -induced reduction of the melting temperature of ionic liquids, Chem. Commun. (2002) 13141315. P. Diep, K.D. Jordan, J.K. Johnson, E.J. Beekman, CO2 uorocarbon and CO2 hydrocarbon interactions from rst-principles calculations, J. Phys. Chem. A 102 (1998) 22312236. M.F.C. Gomes, A.A.H. Padua, Interactions of carbon dioxide with liquid uorocarbons, J. Phys. Chem. B 107 (2003) 1402014024. C.R. Yonker, B.J. Palmer, Investigation of CO2 /uorine interactions through the intermolecular effects on the 1 H and 19 F shielding of CH3F and CHF3 at various temperatures and pressures, J. Phys. Chem. A 105 (2001) 308314. K.R. Seddon, A. Stark, M.J. Torres, in: M.A. Abraham, L. Moens (Eds.), Clean Solvents, American Chemical Society, Washington, DC, ACS Symp. Ser. 819 (2002) p. 3449. K.I. Gutkowski, A. Shariati, C.J. Peters, High-pressure phase behavior of the binary ionic liquid system 1-octyl-3-methylimidazolium tetrauoroborate + carbon dioxide, J. Supercrit. Fluids 39 (2006) 187191. M.C. Kroon, E.K. Karakatsani, I.G. Economou, G.-J. Witkamp, C.J. Peters, Modeling of the carbon dioxide solubility in imidazolium-based ionic liquids with the tPC-PSAFT equation of state, J. Phys. Chem. B 110 (2006) 92629269. J.C. De la Fuente Badilla, C. Peters, J. de Swaan Arons, Volume expansion in relation to the gasantisolvent process, J. Supercrit. Fluids 17 (2000) 1323. A. Shariati, C.J. Peters, High-pressure phase behavior of systems with ionic liquids: measurements and modeling of the binary system uoroform + 1-ethyl-3-methylimidazolium hexauorophosphate, J. Supercrit. Fluids 25 (2003) 109117. W. Wu, J. Zhang, B. Han, J. Chen, Z. Liu, T. Jiang, J. He, W. Li, Solubility of room-temperature ionic liquid in supercritical CO2 with and without organic compounds, Chem. Commun. (2003) 14121413. W. Wu, W. Li, B. Han, T. Jiang, D. Shen, Z. Zhang, D. Sun, B. Wang, Effect of organic cosolvents on the solubility of ionic liquids in scCO2 , J. Chem. Eng. Data 49 (2004) 15971601. Z.F. Zhang, W.Z. Wu, H.X. Gao, B.X. Han, B. Wang, Y. Huang, Triphase behavior of ionic liquidwaterCO2 system at elevated pressures, Phys. Chem. Chem. Phys. 6 (2004) 50515055. Z.F. Zhang, W.Z. Wu, B. Wang, J. Chen, D. Shen, B. Han, High-pressure phase behavior of CO2 /acetone/ionic liquid system, J. Supercrit. Fluids 40 (2007) 16. S.G. Kazarian, B.J. Briscoe, T. Welton, Combining ionic liquids and supercritical uids: in situ ATRIR study of CO2 dissolved in two ionic liquids at high pressures, Chem. Commun. (2000) 20472048. Z. Gu, L.A. Blanchard, D. Hancu, E.J. Beckman, J.F. Brennecke, Environmentally-benign ionic liquid/CO2 biphasic system for separations and reactions, in: Proceedings of 5th International Symposium on Supercritical Fluids, Atlanta, USA, 2000.

179

[130]

[131]

[132]

[133]

[134]

[135]

[136] [137]

[138]

[139]

[140]

[141]

[142]

[143]

[144]

[145]

[146]

[147]

[148]

[149] T. Al-Sahhof, R.S. Al-Ameeri, S.E. Hamam, Bubble point measurements for the ternary system: CO2 , cyclohexane and naphthalene, Fluid Phase Equilib. 53 (1989) 3137. [150] W.L. Weng, M.J. Lee, Vaporliquid equilibrium for binary systems containing a heavy liquid and a dense uid, Ind. Eng. Chem. Res. 31 (1992) 27692773. [151] A.M. Scurto, S.N.V.K. Aki, J.F. Brennecke, CO2 as a separation switch for ionic liquid/organic mixtures, J. Am. Chem. Soc. 124 (2002) 1027610277. [152] A.M. Scurto, S.N.V.K. Aki, J.F. Brennecke, Carbon dioxide induced separation of ionic liquids and water, Chem. Commun. (2003) 572573. [153] S.N.V.K. Aki, A.M. Scurto, J.F. Brennecke, Ternary phase behavior of ionic liquid (IL)organicCO2 systems, Ind. Eng. Chem. Res. 45 (2006) 55745585. [154] V. Najdanovic-Visak, L.P.N. Rebelo, M.N. Da Ponte, Liquidliquid behavior of ionic liquid1-butanolwater and high pressure CO2 -induced phase changes, Green Chem. 7 (2005) 443450. [155] S.V. Dzyuba, R.A. Bartsch, Recent advances in applications of roomtemperature ionic liquid/supercritical CO2 systems, Angew. Chem. Int. Ed. 42 (2003) 148150. [156] M.F. Sellin, P.B. Webb, D.J. Cole-Hamilton, Continuous ow homogeneous catalysis: hydroformylation of alkenes in supercritical uidionic liquid biphasic mixtures, Chem. Commun. (2001) 781782. [157] P.B. Webb, M.F. Sellin, T.E. Kunene, S. Williamson, A.M.Z. Slawin, D.J. Cole-Hamilton, Continuous ow hydroformylation of alkenes in supercritical uidionic liquid biphasic systems, J. Am. Chem. Soc. 125 (2003) 1557715588. [158] U. Hintermair, G. Zhao, C.C. Santini, M.J. Muldoona, D.J. ColeHamilton, Supported ionic liquid phase catalysis with supercritical ow, Chem. Commun. (2007) 14621464. [159] R.A. Brown, P. Pollet, E. McKoon, C.A. Eckert, C.L. Liotta, P.G. Jessop, Asymmetric hydrogenation and catalyst recycling using ionic liquid and supercritical carbon dioxide, J. Am. Chem. Soc. 123 (2001) 12541255. [160] P.G. Jessop, R.R. Stanley, R.A. Brown, C.A. Eckert, C.L. Liotta, T.T. Ngo, P. Pollet, Neoteric solvents for asymmetric hydrogenation: supercritical uids, ionic liquids, and expanded ionic liquids, Green Chem. 5 (2003) 123128. [161] F. Liu, M.B. Abrams, R.T. Baker, W. Tumas, Phase-separable catalysis using room temperature ionic liquids and supercritical carbon dioxide, Chem. Commun. (2001) 433434. [162] A. B smann, G. Franci` , E. Janssen, M. Solinas, W. Leitner, P. Wassero o scheid, Activation, tuning, and immobilization of homogeneous catalysts in an ionic liquid/compressed CO2 continuous-ow system, Angew. Chem. Int. Ed. 40 (2001) 26972699. [163] D. Ballivet-Tkatchenko, M. Picquet, M. Solinas, G. Franci` , P. Wassero scheid, W. Leitner, Acrylate dimerisation under ionic liquidsupercritical carbon dioxide conditions, Green Chem. 5 (2003) 232235. [164] Z. Hou, B. Han, L. Gao, T. Jiang, Z. Liu, Y. Chang, X. Zhang, J. He, Wacker oxidation of 1-hexene in 1-n-butyl-3-methylimidazolium hexauorophosphate ([bmim][PF6 ]), supercritical (SC) CO2 , and SC CO2 /[bmim][PF6 ] mixed solvent, New J. Chem. 26 (2002) 12461248. [165] H. Kawanami, A. Sasaki, K. Matsui, Y. Ikushima, A rapid and effective synthesis of propylene carbonate using a supercritical CO2 ionic liquid system, Chem. Commun. (2003) 896897. [166] L. Gao, T. Jiang, G. Zhao, T. Mu, W. Wu, Z. Hou, B. Han, Transesterication between isoamyl acetate and ethanol in supercritical CO2 , ionic liquid, and their mixture, J. Supercrit. Fluids 29 (2004) 107111. [167] G. Zhao, T. Jiang, W. Wu, B. Han, Z. Liu, H. Gao, Electro-oxidation of benzyl alcohol in a biphasic system consisting of supercritical CO2 and ionic liquids, J. Phys. Chem. B 108 (2004) 1305213057. [168] B. Yoon, C.H. Yen, S. Mekki, S. Wherland, C.M. Wai, Effect of water on the heck reactions catalyzed by recyclable palladium chloride in ionic liquids coupled with supercritical CO2 extraction, Ind. Eng. Chem. Res. 45 (2006) 44334435. [169] D.J. Cole-Hamilton, T.E. Kunene, P.B. Webb, in: B. Cornils, W.A. Herrmann, I.T. Horvath, W. Leitner, S. Mecking, H. Olivier-Bourbigou, D. Vogt (Eds.), Multiphase Homogeneous Catalysis, Wiley-VCH, Weinheim, 2005.

180

S. Keskin et al. / J. of Supercritical Fluids 43 (2007) 150180 [176] P. Lozano, T. de Diego, S. Gmouh, M. Vaultier, J.L. Iborra, Synthesis of glycidyl esters catalyzed by lipases in ionic liquids and supercritical carbon dioxide, J. Mol. Catal. A: Chem. 214 (2004) 113119. [177] P. Lozano, T. de Diego, D. Carri , M. Vaultier, J.L. Iborra, Criteria to e design green enzymatic processes in ionic liquid/supercritical carbon dioxide systems, Biotechnol. Prog. 20 (2004) 661669. [178] F.J. Hern ndez, A.P. de los Ros, D. G mez, M. Rubio, G. Vllora, A a o new recirculating enzymatic membrane for ester synthesis in ionic liquid/supercritical carbon dioxide biphasic systems, Appl. Catal. B 67 (2006) 121126. [179] P. Lozano, T. de Diego, T. Sauer, M. Vaultier, S. Gmouh, J.L. Iborra, On the importance of the supporting material for activity of immobilized Candida antartica lipase B in ionic liquid/hexane and ionic liquid/supercritical carbon dioxide biphasic media, J. Supercrit. Fluids 40 (2007) 93100. [180] M.T. Reetz, W. Wiesenh fer, G. Franci` , W. Leitner, Continuous ow o o enzymatic kinetic resolution and enantiomer separation using ionic liquid/supercritical carbon dioxide media, Adv. Syn. Catal. 345 (2003) 12211228. [181] S. Keskin, Cleaning of contaminated soils using ILs and supercritical carbon dioxide, M.S. Thesis, Bo azici University, g Istanbul, Turkey, 2006.

[170] C.M. Gordon, W. Leitner, in: D.J. Cole-Hamilton, R.P. Tooze (Eds.), Catalyst Separation, Recovery and Recycling: Chemistry and Process Design, Springer, Dordrecht, 2006. [171] J.-Q. Wang, D.-L. Kong, J.-Y. Chen, F. Cai, L.-N. He, Synthesis of cyclic carbonates from epoxides and carbon dioxide over silica-supported quaternary ammonium salts under supercritical conditions, J. Mol. Catal. A: Chem. 249 (2006) 143148. [172] R. Ciriminna, P. Hesemann, J.J.E. Moreau, M. Carraro, S. Campestrini, M. Pagliaro, Aerobic oxidation of alcohols in carbon dioxide with silicasupported ionic liquids doped with perruthenate, Chem. Eur. J. 12 (2006) 52205224. [173] P. Lozano, T. de Diego, D. Carri , M. Vaultier, J.L. Iborra, Continuous e green biocatalytic processes using ionic liquids and supercritical carbon dioxide, Chem. Commun. (2002) 692693. [174] P. Lozano, T. de Diego, D. Carri , M. Vaultier, J.L. Iborra, Lipase catalysis e in ionic liquids and supercritical carbon dioxide at 150 C, Biotechnol. Prog. 19 (2003) 380382. [175] M.T. Reetz, W. Wiesenh fer, G. Franci` , W. Leitner, Biocatalysis in o o ionic liquids: batchwise and continuous ow processes using supercritical carbon dioxide as the mobile phase, Chem. Commun. (2002) 992993.

You might also like