Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

The Application of Green Solvents in Separation Processes
The Application of Green Solvents in Separation Processes
The Application of Green Solvents in Separation Processes
Ebook1,062 pages33 hours

The Application of Green Solvents in Separation Processes

Rating: 4 out of 5 stars

4/5

()

Read preview

About this ebook

The Application of Green Solvents in Separation Processes features a logical progression of a wide range of topics and methods, beginning with an overview of green solvents, covering everything from water and organic solvents, to ionic liquids, switchable solvents, eutectic mixtures, supercritical fluids, gas-expanded solvents, and more.

In addition, the book outlines green extraction techniques, such as green membrane extraction, ultrasound-assisted extraction, and surfactant-mediated extraction techniques. Green sampling and sample preparation techniques are then explored, followed by green analytical separations, including green gas and liquid capillary chromatography, counter current chromatography, supercritical fluid chromatography, capillary electrophoresis, and other electrical separations.

Applications of green chemistry techniques that are relevant for a broad range of scientific and technological areas are covered, including the benefits and challenges associated with their application.

  • Provides insights into recent advances in greener extraction and separation processes
  • Gives an understanding of alternatives to harmful solvents commonly used in extraction and separation processes, as well as advanced techniques for such processes
  • Written by a multidisciplinary group of internationally recognized scientists
LanguageEnglish
Release dateFeb 28, 2017
ISBN9780128054437
The Application of Green Solvents in Separation Processes

Related to The Application of Green Solvents in Separation Processes

Related ebooks

Chemistry For You

View More

Related articles

Reviews for The Application of Green Solvents in Separation Processes

Rating: 4.166666666666667 out of 5 stars
4/5

6 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    The Application of Green Solvents in Separation Processes - Francisco Pena-Pereira

    Poland

    Section I

    Introduction

    Outline

    Chapter 1 Initial Considerations

    Chapter 1

    Initial Considerations

    Francisco Pena-Pereira¹ and Marek Tobiszewski²,    ¹University of Vigo, Vigo, Spain,    ²Gdańsk University of Technology (GUT), Gdańsk, Poland

    Abstract

    Chemical processes have made use of a wide number of volatile organic solvents due to their efficiency in countless unit operations. The increasing knowledge on the environmental, health, and safety hazards associated to their use has significantly contributed to the search for greener alternatives. The primary purpose of this book is to provide an overview of alternative solvents with reduced issues that have successfully replaced harmful organic solvents in chemical processes and methodologies. In this introductory chapter, initial considerations on the applicability of alternative solvents reported in the scientific literature toward greener chemical processes are provided. Besides, a brief overview of greener extraction and separation techniques described along the book is presented herein.

    Keywords

    Extraction processes; green chemistry; green solvents; microextraction; separation techniques; sustainability

    1.1 The Need to Use Solvents

    Solvents are ubiquitous auxiliary substances mainly used in paints, coatings, adhesives, cleaners, and cosmetics, but also in chemical and pharmaceutical processes. A worldwide solvent consumption of about 30 million metric tons per year has been recently estimated [1]. From them, alcohols such as methanol, ethanol, n-butanol, and iso-propanol are the more typically used solvents, with an annual consumption of 6.5 million tons in accordance with a recent market study [2]. Aromatics, ketones, esters, and ethers are also widely used. The global solvent consumption has continuously increased in the last years, and this trend seems likely to persist and even intensify in the future. The expected increase in the short term has been mainly attributed to the growing demand for solvents by the emerging market economies, especially to meet the requirements in construction and automobile industries.

    Solvents are also incorporated in many fields of chemistry. They seldom play the leading role in chemical activities. Chemists are more interested in the minor amounts of compounds that are dissolved or obtained in them. It is very common that solvents build the great majority of the mass of chemicals used in chemical processes and they are often out of direct chemists’ attention. This is especially valid in case of fine chemicals and pharmaceutical industries, where great mass of solvent is utilized per mass of product [3]. Chemical synthesis and separation processes are two main areas involving organic solvents. Thus, solvents allow carrying out countless chemical reactions, isolation and purification of target compounds, formation of azeotropes for separation, or temperature control, among other applications, with excellent performance. Solventless processes are nowadays favored, and alternative approaches are introduced to reduce the environmental impact and safety issues of volatile organic solvents.

    1.2 Traditional Solvents

    Organic solvents show a paramount role in a wide range of scientific and technological applications. However, several problems can be ascribed to the use of most of traditional solvents. Apart from the great amounts of solvents used, there is high risk connected to the application of solvents of very high concern. They originate from various classes of chemicals, so they are characterized by very diverse problems they may cause. Historically, halogenated solvents were willingly applied because of their excellent properties to dissolve many organic compounds. However, several of them are currently under regulation, and their application is only allowed for those specific activities where appropriate alternatives are not available. Other frequently used solvents are aliphatic or aromatic hydrocarbons. All of these compounds are characterized by the benefit of poor water solubility, what gives the possibility for easy separation of these solvents from aqueous fractions. Another issue related to the utilization of traditional organic solvents is their toxicity, especially concerning hydrocarbons and chlorinated solvents. These are characterized by high oral and inhalation toxicities. They are often characterized by chronic negative effects, such as mutagenicity, teratogenicity, and some of the solvents are listed as carcinogens [4].

    In addition, a wide range of organic solvents show environmental persistence. While polar solvents are readily degradable in the aquatic environment via hydrolysis reaction and biodegradation, other solvents such as chlorinated volatile organic compounds (VOCs) are hardly degradable, with their aquatic environment half-lives reaching decades. Present in the atmosphere, solvents can contribute to global warming, ozone depletion, or the occurrence of tropospheric ozone.

    Because of their high volatility, several organic solvents can also cause risk to their users. Their occupational exposure cannot be neglected and they proved to cause acute and chronic threats. What is more, some of them are difficult to handle as they are flammable and explosive.

    Important efforts have been made to control the production and consumption of hazardous organic solvents in the last decades. For instance, the International Labor Organization (ILO) adopted, in 1971, the Benzene Convention, regarding the protection against hazards arising from benzene [5]. Accordingly, the ILO established that harmless or less harmful substitute products should be used instead of benzene or products containing benzene whenever possible. In addition, the Montreal Protocol on Substances that Deplete the Ozone Layer set, in 1989, a list of ozone-depleting substances with the aim of controlling their production and consumption [6]. Fig. 1.1 shows a summary of proposed phaseout schedules for developed and developing countries under the Montreal Protocol that include several halogenated solvents typically used in chemical processes. Furthermore, The Registration, Evaluation, Authorization and Restriction of Chemicals (REACH) aimed to achieve a minimization of significant adverse effects on human health and the environment in both production and use of chemicals by 2020 [7]. REACH regulations pursued to rule out dangerous substances or progressively replacing substances of concern by less dangerous substances where suitable alternatives are available. Several commonly used solvents are included among a large number of substances classified as carcinogenic, mutagenic, or toxic to reproduction that are controlled by REACH with specific conditions of restriction.

    Figure 1.1 Summary of control measures under the Montreal Protocol: (A) Chlorofluorocarbons (CFC-11, CFC-12, CFC-113, CFC-114, CFC-115); (B) halons (halon 1211, halon 1301, halon 2402); (C) other fully halogenated CFCs (CFC-13, CFC-111, CFC-112, CFC-211, CFC-212, CFC-214, CFC-215, CFC-216, CFC-217); (D) carbon tetrachloride; (E) methyl chloroform; (F) methyl bromide; (G) hydrochlorofluorocarbons (HCFCs, consumption); (H) HCFCs (production) [6]. Control measures established for developed countries are represented by black lines, while those established for developing countries are shown in red.

    Regulations have been implemented to decrease the consumption of solvents showing important issues or, in the case of substances of very high concern, cease of their use when they can cause serious and irreversible harm to human health and the environment. Regulations show an important role in order to minimize the deleterious effects of solvent consumption. For instance, nonmethane VOC emissions have experienced a decrease of 3.4% per year in the European Union in the period 2000–13 [8]. Even though the corresponding data (7 million tons of nonmethane VOC emissions in 2013) is still improvable, it can be considered that the implementation of stricter regulations and the control of solvent use and emissions have contributed to the reduction of VOC consumption in Europe.

    Regrettably, hazardous organic solvents are yet used in a wide range of scientific and technological activities, thus contributing to the generation of significant amounts of wastes. Thus, activities such as purification of wastes and vapor streams, removal of contaminants, or monitoring of environmental pollutants significantly contribute to further pollution due to the organic solvents used in the corresponding processes. Bearing in mind the environmental, health, and safety issues associated with the use of many conventional organic solvents, the development of greener alternatives is becoming increasingly important. Another important aspect regarding organic solvents is that they are mainly derived from fossil-based feedstocks. Thus, the possibility of obtaining valuable solvents from biomass waste without making use of additional solvents derived from crude oil is a challenging research area that has been explored in the last years with highly promising results [9,10]. Production of green solvents from renewable raw materials is therefore of utmost interest to develop more sustainable chemical processes and methodologies.

    All the above mentioned issues related to the application of traditional solvents are good reasons for searching greener solutions. In this sense, the inception of green chemistry paved the way for a more careful and conscientious use of solvents. Regarding extraction and separation processes, several aspects should be considered to select a given solvent, including not only the environmental, health, and safety issues of considered solvents but, also importantly, metrological and economic aspects. Thus, the extraction and/or separation performance, energy demand, possibility of solvent recovery, and recycling, as well as solvent compatibility with analytical instrumentation (in the case of analytical method development), should be carefully assessed.

    1.3 Green Solvents

    In the last decades, scientists have paid increasing attention to the adverse effects of reagents and solvents used in chemical processes. The introduction of the 12 principles of green chemistry represented a turning point toward a reduction of environmental, safety, and health hazards associated to conventional chemical processes [11,12]. The development and application of greener solvents is currently a hot topic in a variety of scientific and technological areas. Green solvents can be defined as those solvents that display reduced health, safety, and environmental issues and a reduced life cycle impact [13–15]. Fig. 1.2 shows a number of alternative solvents that fulfill, to a greater or lesser extent, this definition and are described in depth in this book.

    Figure 1.2 Green solvents used in extraction and separation processes.

    Two main measures have been considered to make chemical processes greener when solventless approaches are not feasible. The more conservative one, yet challenging, involves a significant reduction of organic solvent consumption in a given chemical process. More desirably, the second measure involves a replacement of harmful solvents by greener alternatives. Needless to say that the combination of both measures represent the most advantageous approach. Significant efforts have been made in both directions. Substitution of a harmful solvent by a more benign alternative is, of course, not trivial in many cases, as novel challenges and obstacles can arise due to the different physicochemical properties of the solvents considered.

    Water can be considered as the greenest solvent from the wide list of substances traditionally used in the chemical industry. It is environmentally benign, nonflammable, and easily obtainable on a large scale with high purity. Remarkably, its physicochemical properties can be tuned by temperature and pressure. Thus, water can even be used for extraction of nonpolar compounds with no need for additional organic modifiers [16,17]. However, water has limited applicability in chemical industry and the need for searching other green solvents is considered very urgent [14]. In some cases, conventional solvents with reduced health, environmental, and safety issues have been recommended. For instance, some analytical methods involving extractions with harmful solvents have been revised so as alternative, less harmful, solvents are nowadays recommended. For instance, USEPA methods 413.1 and 413.2 for testing oil and grease in water were replaced by the EPA methods 1664A and 1664B a decade ago. In practice, Fluorocarbon-113, an ozone-depleting substance controlled by the Montreal Protocol [18], was replaced by n-hexane in these methods [19]. In this vein, the recent efforts made toward the development of solvent selection guides allows choosing appropriate sets of solvents for a given application taking into account their environmental, health, and safety issues, potential to be recycled, possibility to be obtained from renewable feedstocks, etc. [20–22]. Thus, organic solvents can be relatively easily found for problem solving, where polar solvents can be applied. In such cases alcohols, esters, or carboxylic acids are frequently incorporated. This is especially true in the case of conventional oxygenated solvents that can be prepared from renewable feedstocks (e.g., acetone, methyl ethyl ketone, methanol, ethyl acetate, and 2-methyltetrahydrofuran) [23].

    Recently, there is a trend to search for and apply bio-based solvents [24]. A sustainable conversion of biomass to high value-added chemicals is a challenging research area. Biorefineries are systems that produce chemicals, including solvents from a given plant feedstock, usually seaweed or wastes from certain food industrial processes and agriculture [25]. Bio-based solvents are increasingly of interest for scientific and technological applications. A number of platform molecules, such as lactic acid, levulinic acid, furfural, 5-hydroxymethylfurfural, or γ-valerolactone [26], have been used for the synthesis of bio-based solvents. Production of benign solvents from renewable feedstock is in agreement with green chemistry principles. A detailed assessment of their environmental impact is, however, required since their generation may require more energetic input than obtaining solvents from fossils.

    Supramolecular solvents represent appealing solvents that consist in nanostructured solvents formed from minute amounts of amphiphile molecules by self-assembly and coacervation [27,28]. The selection of amphiphiles significantly affects the physicochemical properties of the obtained supramolecular solvents and, in turn, their applicability. Thus, supramolecular solvents can be considered as tunable solvents with potential applicability for extraction of compounds over a wide polarity range.

    As many of the environmental problems with traditional solvents are related to their volatility, ionic liquids have gained much attention of chemical society. Their negligible vapor pressures, ability to dissolve many compounds and possibility to recover them, good thermal stability, low flammability, and other desirable features make them attractive green solvents [29,30]. The great amount of cation–anion combinations, offers the chance to adjust their physicochemical properties to chemists’ requirements. However, it turned out that certain ionic liquids are even more toxic to aquatic organisms than traditional solvents, and their environmental inertness has been questioned [31]. Their synthesis and application thus requires more detailed assessment procedures.

    Deep eutectic solvents (DESs) are another class of solvents that are concerned green [32–34]. They are obtained by using, at least, two components that behave as hydrogen bond donor and acceptor, respectively. DESs have similar physicochemical parameters to ionic liquids but are also characterized by certain advantages over them. The most important are reduced toxicity, tenability, and ease of synthesis with high efficiency. Remarkably, they can be obtained from renewable feedstocks such as choline chloride and carbohydrates or amino acids.

    Switchable solvents, also known as smart solvent systems, are appealing solvents that reversibly change their physicochemical properties under an external stimulus [35,36]. The addition or removal of a gas (e.g., CO2) or a temperature change gives rise to a reversible reaction that enables the formation of solvents with switchable properties. The switching behavior of these solvents can be exploited for solute purification in separation processes and facilitates their recycling, thus minimizing waste generation.

    Finally, supercritical fluids and gas-expanded solvents have also significantly contributed to greening a wide range of separation processes [37–39]. A number of solvents have been used under supercritical conditions. From them, CO2 is the most commonly used supercritical fluid since it is safe and environmentally benign, shows a low critical temperature, high diffusivity, and affordability. Furthermore, solvent-free compounds of interest can be easily obtained after separation by means of supercritical CO2. Among the drawbacks of CO2 as supercritical fluid, it can be highlighted its reduced polarity, which makes necessary the use of additional organic modifiers for separation of highly polar compounds. Gas-expanded solvents are mixed solvents consisting in organic solvents dissolved in compressed gases. Among the different possible compressible gases, CO2 is the most commonly used. Gas-expanded solvents are highly tunable by simple pressure variations, showing intermediate properties between liquid solvents and supercritical fluids. A range of applications, including chemical reactions, separations, or advanced materials, have been reported in the literature based on the use of these tunable solvent systems.

    As briefly introduced in this section, an increasing number of green solvents have been reported in the literature. While several of them are well established in different separation processes, even at industrial scale, more recent alternative solvents are still in their initial stages. Design, preparation, characterization, and application of novel solvents in scientific and technological activities thus represent a challenging area of research. Greener solvents briefly described earlier have been discussed in depth in Section II, Green Solvents, of the book.

    1.4 Greener Extraction Techniques

    Extraction techniques allow the isolation, purification, and/or enrichment of target compounds, sample cleanup, and transfer of relevant molecules into a more convenient phase for further processing. Various extraction techniques have been reported in the literature for pretreatment of liquid, gaseous, and solid matrices. Conventional extraction techniques such as liquid–liquid extraction and Soxhlet extraction have been extensively used, for instance, to the isolation of high value-added products, removal of pollutants, and analytical method development [40,41]. However, these extraction techniques are nowadays considered inefficient, taking into account the extended times needed to extract target compounds with the required extraction efficiency, and the reduced enrichment yields typically achieved. More importantly from the point of view of green chemistry, they make use of large solvent volumes per sample and, as a result, large amounts of wastes are generated [42]. Several examples of extraction procedures based on these conventional techniques can be found in the literature, for instance in reference methods of analysis. Greener extraction techniques have been incorporated in the last years with the aim of improving extraction capability while minimizing and even avoiding solvent consumption. In addition, the application of microwaves and ultrasounds allows a reduction of solvent volumes consumed and energy savings due to shortened extraction times and more efficient energy delivery [43–46]. These systems, commonly considered as clean energies, are widely used in organic synthesis, food industry and analytical chemistry.

    Application of membranes has also been helpful in the reduction of organic solvents volumes [47,48]. Such systems are developed for industrial and analytical purposes. The selectivity achieved by means of the various types of membranes is one of their main advantages in separation processes.

    Supercritical fluid extraction allows for selective extraction of compounds from various media. The most frequently used CO2 in its supercritical state is nontoxic, nonflammable, and its supercritical point is relatively easy to be reached [38,49]. Adding modifiers to the supercritical fluid allows the adjustment of its polarity, what extends its applicability even to the extraction of polar compounds.

    Extraction techniques have enabled the development of greener methodologies and processes. Section III, Green Extraction Techniques, of the book is devoted to green extraction techniques.

    1.5 Green Sampling and Sample Preparation

    The techniques described in Section III, Green Extraction Techniques, of this book are applied in chemical industry and in the smaller scale for analytical determinations. However, analytical chemistry has its own, specific ways for greening separation processes, mainly by elimination or significant reduction of solvent volume.

    There are strategies applied at sample collection stage that determine the environmental impact of an analytical procedure at its further stages. The classical example is sample collection with passive sampling followed by thermal desorption of analytes [50]. Similar strategies, implemented at sample collection stage, leading to greening of the whole procedure are based on microextraction techniques. These techniques allow for separation of analytes even from very complex biological matrices [51]. Sample pretreatment is a key step of the analytical process required to improve both the sensitivity and selectivity of a given analytical methodology that commonly involves relatively large amounts of organic solvents. Thus, significant improvements have been performed to make greener this critical step. The important efforts performed in the last two decades toward miniaturization of conventional sample preparation techniques such as liquid–liquid extraction and solid-phase extraction resulted in the introduction of solventless or solvent-minimized sample preparation approaches [52–54]. Solid-phase microextraction and liquid-phase microextraction are nowadays well-established techniques that enable the enrichment of target compounds prior to their determination by the more appropriate analytical instrumentation. In spite of the evident slowness in the implementation of microextraction techniques in standard method selection, the number of reference methods involving micronized analytical systems is expected to increase in the near future bearing in mind their advantageous conditions in terms of metrology, economy, and environmental impact.

    Section IV, Green Sampling and Sample Preparation Techniques, of the book is devoted to sampling and sample preparation techniques that have significantly contributed to greening these important steps of the analytical process.

    1.6 Solvents for Analytical Separations

    Chromatographic and electrophoretic separation techniques are widely used for separation and determination of target compounds in matrices of different complexity. A large number of reference methods of analysis involve separation techniques coupled with appropriate detectors. Nonchromatographic methods can sometimes be used for determination of relevant analytes with a negligible consumption of solvents [55]. Nevertheless, analytical separations can be unavoidable or even more convenient than nonchromatographic methods in certain cases. An issue that cannot be omitted, however, is the application of solvents in separation techniques, especially in the case of liquid chromatography (LC). Separation techniques importantly contribute to the total solvent consumption in analytical laboratories and large amounts of liquid wastes are consequently produced. The magnitude of this activity is far from being negligible. In fact, it has been estimated that an amount of wastes of around 150,000 tons could be generated by LC systems on a yearly basis [15,56]. While efforts have been mainly devoted to solvent reduction, reuse, and recycling, recently, much attention is given to apply greener solvents as mobile phases instead of conventional solvents such as acetonitrile or dichloromethane [57–59]. It should be highlighted that apart from their ability to enable successful separations of a great variety of relevant compounds, the compatibility of alternative solvents with typical detection systems used in combination with analytical separation techniques is also of paramount importance and should therefore be demonstrated. Miniaturization of chromatographic systems, as well as reduction of packed columns dimensions or application of capillary columns in LC are efforts to the decreasing of mobile phases consumption.

    Alternative techniques to LC allow for separation of compounds without application of organic solvents. Supercritical fluid chromatography [60], gas chromatography [56], and capillary electrophoresis [61] are considered to be green separation techniques, with their own advantages and disadvantages, but they all apply mobile phases that are more benign in their nature. The combination of these greener analytical separation techniques with solventless sample preparation approaches provides excellent possibilities for determination of target compounds at trace and ultra-trace levels while fulfilling the principles of green chemistry. Green aspects of analytical separation techniques are discussed in Section V, Green Analytical Separations, of the book.

    1.7 Concluding Remarks

    As described in the earlier sections, the design, synthesis and application of solvents with reduced environmental, health, and safety hazards represent a fundamental area of research toward sustainable chemical processes. This book is aimed to provide an up-to-date view on the application of green solvents in extraction and separation techniques. Important aspects regarding more benign solvents are addressed. In addition, challenging aspects derived from the use of greener solvents as substitutes of harmful solvents in extraction and separation processes are discussed. Furthermore, relevant scientific and technological applications are also provided in the book. From the table of contents of the book it can be easily seen that it is substantially focused on analytical chemistry. The analytical perspective has not been presented yet in the application of green solvents in separation processes. At the same time we wanted to keep the book interesting for nonanalytical audience.

    Acknowledgments

    At this point we would like to express our gratitude for the hard work and dedication of all contributors that concurred to the development of this book.

    F. Pena-Pereira acknowledges Xunta de Galicia for financial support as a postdoctoral researcher of the I2C program.

    References

    1. E. Linak, S.N. Bizzari, Global Solvents: Opportunities for Greener Solvents, 2013.

    2. Ceresana, Market Study: Solvents, third ed. <http://www.ceresana.com/en/marketstudies/chemicals/solvents/>, 2014 (accessed 23.08.16).

    3. Sheldon RA. The E factor: fifteen years on. Green Chem. 2007;9:1273–1283.

    4. Huang B, Lei C, Wei C, Zeng G. Chlorinated volatile organic compounds (Cl-VOCs) in environment—sources, potential human health impacts, and current remediation technologies. Environ Int. 2014;71:118–138.

    5. International Labour Organization. Benzene convention: convention concerning protection against hazards of poisoning arising from benzene. <http://www.ilo.org/dyn/normlex/en/f?p=NORMLEXPUB:12100:0::NO::P12100_ILO_CODE:C136>, 1971 (accessed 24.07.16).

    6. United Nations Environmental Programme (UNEP), Handbook for the Montreal Protocol on Substances That Deplete the Ozone Layer, ninth ed., Nairobi, 2012.

    7. Regulation (EC) No 1907/2006 of the European Parliament and of the Council of 18 December 2006, Off. J. Eur. Union, 2006, L396/1–L396/849.

    8. Eurostat Sustainable development in the European Union—2015 monitoring report of the EU Sustainable Development Strategy, 2015.

    9. Tuck CO, Pérez E, Horváth IT, Sheldon RA, Poliakoff M. Valorization of biomass: deriving more value from waste. Science. 2012;337:695–699.

    10. Gu Y, Jérôme F. Bio-based solvents: an emerging generation of fluids for the design of eco-efficient processes in catalysis and organic chemistry. Chem Soc Rev. 2013;42.

    11. Anastas PT, Warner JC. Green Chemistry: Theory and Practice New York, NY: Oxford University Press; 1998.

    12. Anastas P, Eghbali N. Green chemistry: principles and practice. Chem Soc Rev. 2010;39:301–312.

    13. Nelson WM. Green solvents for chemistry: perspectives and practice Oxford: Oxford University Press; 2003.

    14. Capello C, Fischer U, Hungerbühler K. What is a green solvent? A comprehensive framework for the environmental assessment of solvents. Green Chem. 2007;9:927–934.

    15. Pena-Pereira F, Kloskowski A, Namieśnik J. Perspectives on the replacement of harmful organic solvents in analytical methodologies: a generation of eco-friendly alternatives. Green Chem. 2015;17:3687–3705.

    16. Kronholm J, Hartonen K, Riekkola ML. Analytical extractions with water at elevated temperatures and pressures. Trends Anal Chem. 2007;26:396–412.

    17. Smith RM. Superheated water chromatography—a green technology for the future. J Chromatogr A. 2008;1184:441–455.

    18. The Montreal Protocol on substances that deplete the ozone layer. 2000 (accessed 1.2.17).

    19. E.P.A., Method 1664, Revision A: n-hexane extractable material (HEM; oil and grease) and silica gel treated n-hexane extractable material (SGT-HEM; non-polar material) by extraction and gravimetry, Washington, DC, 1999.

    20. Henderson RK, Jiménez-González C, Constable DJC, et al. Expanding GSK’s solvent selection guide—embedding sustainability into solvent selection starting at medicinal chemistry. Green Chem. 2011;13:854–862.

    21. Tobiszewski M, Tsakovski S, Simeonov V, Namiésnik J, Pena-Pereira F. Solvent selection guide based on chemometrics and multicriteria decision analysis. Green Chem. 2015;17:4773–4785.

    22. Prat D, Hayler J, Wells A. A survey of solvent selection guides. Green Chem. 2014;16:4546–4551.

    23. Dunn PJ. The importance of green chemistry in process research and development. Chem Soc Rev. 2012;41:1452–1461.

    24. Li Z, Smith KH, Stevens GW. The use of environmentally sustainable bio-derived solvents in solvent extraction applications—a review. Chin J Chem Eng. 2016;24:215–220.

    25. Esposito D, Antonietti M. Redefining biorefinery: the search for unconventional building blocks for materials. Chem Soc Rev. 2015;44:5821–5835.

    26. Gallezot P. Conversion of biomass to selected chemical products. Chem Soc Rev. 2012;41:1538–1558.

    27. Ballesteros-Gómez A, Sicilia MD, Rubio S. Supramolecular solvents in the extraction of organic compounds A review. Anal Chim Acta. 2010;677:108–130.

    28. López-Jiménez FJ, Lunar ML, Sicilia M, Rubio S. Supramolecular solvents in the analytical process. Encycl Anal Chem. 2014;1–16 http://dx.doi.org/10.1002/9780470027318.a9396.

    29. Rogers RD, Seddon KR. Ionic liquids—solvents of the future? Science. 2003;302:792–793.

    30. Cevasco G, Chiappe C. Are ionic liquids a proper solution to current environmental challenges? Green Chem. 2014;16:2375–2385.

    31. Cvjetko Bubalo M, Radošević K, Radojčić Redovniković I, Halambek J, Gaurina Srček V. A brief overview of the potential environmental hazards of ionic liquids. Ecotoxicol Environ Saf. 2014;99:1–12.

    32. Khandelwal S, Tailor YK, Kumar M. Deep eutectic solvents (DESs) as eco-friendly and sustainable solvent/catalyst systems in organic transformations. J Mol Liq. 2016;215:345–386.

    33. Zhang Q, De Oliveira Vigier K, Royer S, Jérôme F. Deep eutectic solvents: syntheses, properties and applications. Chem Soc Rev. 2012;41:7108–7146.

    34. Francisco M, van den Bruinhorst A, Kroon MC. Low-transition-temperature mixtures (LTTMs): a new generation of designer solvents. Angew Chem Int Ed Engl. 2013;52:3074–3085.

    35. Jessop PG, Heldebrant DJ, Li X, Eckertt CA, Liotta CL. Green chemistry: reversible nonpolar-to-polar solvent. Nature. 2005;436:1102.

    36. Pollet P, Eckertabc CA, Liotta CL. Switchable solvents. Chem Sci. 2011;2:609–614.

    37. Smith RM. Supercritical fluids in separation science—the dreams, the reality and the future. J Chromatogr A. 1999;856:83–115.

    38. da Silva RPFF, Rocha-Santos TAP, Duarte AC. Supercritical fluid extraction of bioactive compounds. Trends Anal Chem. 2016;76:40–51.

    39. Jessop PG, Subramaniam B. Gas-expanded liquids. Chem Rev. 2007;107:2666–2694.

    40. Raynie DE. Modern extraction techniques. Anal Chem. 2006;78:3997–4003.

    41. Luque de Castro MD, Priego-Capote F. Soxhlet extraction: past and present panacea. J Chromatogr A. 2010;1217:2383–2389.

    42. Garrigues S, Armenta S, Guardia MD La. Green strategies for decontamination of analytical wastes. Trends Anal Chem. 2010;29:592–601.

    43. Camel V. Microwave-assisted solvent extraction of environmental samples. Trends Anal Chem. 2000;19:229–248.

    44. Jacotet-Navarro M, Rombaut N, Deslis S, et al. Towards a dry bio-refinery without solvents or added water using microwaves and ultrasound for total valorization of fruit and vegetable by-products. Green Chem. 2016;18:3106–3115.

    45. Chemat F, Rombaut N, Sicaire A-G, Meullemiestre A, Fabiano-Tixier A-S, Abert-Vian M. Ultrasound assisted extraction of food and natural products Mechanisms, techniques, combinations, protocols and applications A review. Ultrason Sonochem. 2017;34:540–560.

    46. Koubaa M, Mhemdi H, Barba FJ, Roohinejad S, Greiner R, Vorobiev E. Oilseed treatment by ultrasounds and microwaves to improve oil yield and quality: an overview. Food Res Int. 2016;85:59–66.

    47. Kubáň P, Boček P. Micro-electromembrane extraction across free liquid membranes instrumentation and basic principles. , J Chromatogr A. 2014;1346:25–33.

    48. Figoli A, Marino T, Simone S, et al. Towards non-toxic solvents for membrane preparation: a review. Green Chem. 2014;16:4034–4059.

    49. Luque de Castro MD, Jiménez-Carmona MM. Where is supercritical fluid extraction going? Trends Anal Chem. 2000;19:223–228.

    50. Charriau A, Lissalde S, Poulier G, Mazzella N, Buzier R, Guibaud G. Overview of the Chemcatcher® for the passive sampling of various pollutants in aquatic environments Part A: Principles, calibration, preparation and analysis of the sampler. Talanta. 2016;148:556–571.

    51. Souza Silva EA, Risticevic S, Pawliszyn J. Recent trends in SPME concerning sorbent materials, configurations and in vivo applications. Trends Anal Chem. 2013;43:24–36.

    52. Pawliszyn J. Handbook of Solid Phase Microextraction Beijing: Chemical Industry Press; 2009.

    53. Kokosa JM, Przyjazny A, Jeannot MA. Solvent Microextraction, Theory and Practice Horboken, NJ: Wiley; 2009.

    54. Pena-Pereira F, ed. Miniaturization in Sample Preparation. Berlin: De Gruyter Open; 2014.

    55. Gonzalvez A, Cervera ML, Armenta S, de la Guardia M. A review of non-chromatographic methods for speciation analysis. Anal Chim Acta. 2009;636:129–157.

    56. Płotka J, Tobiszewski M, Sulej AM, Kupska M, Górecki T, Namieśnik J. Green chromatography. J Chromatogr A. 2013;1307:1–20.

    57. Chardon FM, Blaquiere N, Castanedo GM, Koenig SG. Development of a tripartite solvent blend for sustainable chromatography. Green Chem. 2014;16:4102–4105.

    58. Peterson EA, Dillon B, Raheem I, et al. Sustainable chromatography (an oxymoron?). Green Chem. 2014;16:4060–4075.

    59. Welch CJ, Wu N, Biba M, et al. Greening analytical chromatography. Trends Anal Chem. 2010;29:667–680.

    60. Dispas A, Lebrun P, Sassiat P, et al. Innovative green supercritical fluid chromatography development for the determination of polar compounds. J Chromatogr A. 2012;1256:253–260.

    61. Koel M. Do we need green analytical chemistry? Green Chem. 2016;18:923–931.

    Section II

    Green Solvents

    Outline

    Chapter 2 Water as the First Choice Green Solvent

    Chapter 3 A Systematic Approach to Green Solvent Selection, Design, and Verification

    Chapter 4 Bio-Based Molecular Solvents

    Chapter 5 Supramolecular Solvents for Green Chemistry

    Chapter 6 Ionic Liquids, Switchable Solvents, and Eutectic Mixtures

    Chapter 7 Supercritical Fluids and Gas-Expanded Liquids

    Chapter 2

    Water as the First Choice Green Solvent

    Kari Hartonen and Marja-Liisa Riekkola,    University of Helsinki, Helsinki, Finland

    Abstract

    Water is the most environmentally friendly solvent to be considered for green separation processes. Thanks to its good and tunable physicochemical properties, its use for separation, and especially for extraction and chromatography, has been remarkably expanded. Although water has been most frequently used mixed with organic solvents in numerous applications carried out by many separation techniques at ambient or slightly elevated temperatures, the main goal of this chapter is to describe its exploitation alone in extractions and chromatographic separations at high temperatures. Material and method development together with the most recent applications are presented. Extraction applications using high-temperature water are mainly focused on studies related to biorefineries, bioactive compounds, energy technology, amino acids, and essential oils. Liquid chromatography with water at high temperature has been most commonly applied to analysis of pharmaceuticals, cosmetics, and food.

    Keywords

    Water; green solvent; high temperature; extraction; liquid chromatography; applications

    2.1 Introduction

    During the last decade, environmentally friendly (green) solvents have attracted continuing and increasing interest in clean chemical processes of biorefineries [1,2], chemical reactions/synthesis [3], energy technology [4,5], and chemical separations [6]. In-depth research on different green solvents, such as gas expanded liquids, ionic liquids, supercritical fluids, and water, has greatly complemented these studies [7]. Although cold or room temperature water has also many interesting applications, for example, in the extraction of bioactive compounds [8–10], this chapter will mainly focus on the use of water alone as a green solvent in different chemical separations at elevated temperatures. Namely water, mixed with organic solvents is also widely employed in many separation techniques and processes to enhance solvation of less polar analytes. Special attention here is given to most recent developments and applications.

    2.1.1 Why to Use Water as a Solvent?

    Molecular structure of water and its capability to form hydrogen bonds (and hydrogen-bonded network) result in unique physicochemical properties making water an interesting solvent and vital substance for life. Ionic and polar compounds are readily dissolved in water at ambient condition and less polar compounds at elevated temperatures. Water is also nontoxic, nonflammable, cheap, and widely available in pure form. In addition, it exists in liquid state at relatively high temperatures (0–100°C) at atmospheric pressure. Thanks to its good solvation properties, water plays an important role in the transport of many nutrients and other substances in the environment, enabling various biological processes. Due to its high heat capacity, it also stabilizes temperatures of organisms and in various regions of the earth. Based on all these facts, it is no wonder that water is the most natural choice for the solvent. However, aqueous waste generated need to be cleaned to guarantee the quantity and quality of our water resources also in the future.

    2.1.2 Water—The Most Green Choice

    Water is definitely the safest and the most environmentally friendly solvent that can be considered for use as solvent in different, widely studied, and developed green separation processes. It is favorably assessed in recent solvent selection guides [11,12]. In addition to aqueous separations, water is also popular medium and reagent in synthesis/reactions [3,13], oxidation of waste materials [14,15], and conversion of biomass into fuels and chemicals [4,5,16].

    2.2 Solvent Properties of Water

    Water can most likely solubilize more different kinds of compounds than any other solvent. Its solvent properties can be characterized by its dipole moment, relative permittivity εr, which is a measure of solute solvent interactions, and by its polarizability π*. For water at ambient conditions (25°C), all these three parameters have high values 1.85, 78.5, and 1.12, respectively [17]. In addition, for better solvent characterization, one should include hydrogen bond donating (α=1.2 at 25°C) and accepting (β=0.37 at 25°C) abilities included in solvatochromic Kamlet-Taft parameters [18]. The values listed above explain why water is such a good solvent for ionic and polar compounds. However, it should be noted that at elevated temperatures, it becomes less polar being an alternative to organic solvents.

    As can be seen from Table 2.1 and Fig. 2.1, relative permittivity (dielectric constant), polarizability, and degree of hydrogen bonding in water are decreased when temperature of water is increased. At the same time, its viscosity and surface tension are decreased. These large changes result in remarkable changes in solvent properties of water that are very unique compared to any other solvent. Basically, water polarity is decreased with temperature, allowing the fine-tuning of solvent properties of water for each case by just adjusting temperature.

    Table 2.1

    Chemical and Physical Properties of Liquid Water at Different Temperatures and at 10 MPa Pressure

    Calculated With NIST/ASME Steam Properties Database Version 2.01.

    aAt saturation pressure.

    bAt 0.1 MPa or at saturation pressure.

    Figure 2.1 Polarizability π* of water (A) and hydrogen bond donating acidity α (B) as a function of temperature. Reprinted with permission from J. Lu, J.S. Brown, E.C. Boughner, C.L. Liotta, C.A. Eckert, Solvatochromic characterization of near-critical water as a benign reaction medium, Ind. Eng. Chem. Res. 41 (2002) 2835–2841 [22]. Copyright © 2002 American Chemical Society.

    Separations using water at elevated temperatures can be divided into different categories: (1) hot or high temperature (HT) water from ambient to 100°C, (2) pressurized hot (also subcritical or superheated) water (PHW) from 100°C to 374°C, and (3) supercritical water (SCW). This classification is not very strict since the term HT is also often used at temperatures above 100°C. Temperature plays a dominant role in the determination of the solvent properties of water, and pressure has only a minor effect. However, because pressure affects greatly to water properties when the change of state from gas to liquid or vice versa occurs, the users should distinguish whether the water is in the state of liquid or steam.

    For separations carried out at different temperatures in water, it is vital to have information on different physicochemical properties of analytes. With increasing temperature, changes in selectivity and solute transfer thermodynamics were observed in reversed-phase liquid chromatography (RPLC) using PHW due to disruption of hydrogen-bonding network [23]. Binary diffusion coefficients in PHW were determined for phenolic compounds by Srinivas et al. [24]. Moreover, it is essential to know analyte solubilities in PHW that can be easily determined experimentally, for example, via correlation with retention factor k by liquid chromatography utilizing PHW as eluent [25]. Compared to conventional separation carried out in aqueous–organic solvent eluent, extrapolation to pure water is thus avoided in this approach. Recent solubility data in PHW are available also for parabens [26] and for several bioactive compounds [27]. Static permittivities [28], viscosities [29], and vapor pressures [30] of pure solvents and binary mixtures at HTs have been determined to estimate their suitability for high-temperature liquid chromatography (HT-LC).

    When performing separations using HT water, it is crucial to have information on thermal stability or possible degradation of the analytes. At elevated temperatures, the reactivity (hydrolysis, oxidation) in water also increases; thus, causing problems in material durability of the separation system (stationary phase, column, tubings, extraction vessel, etc.).

    Extraction of thermolabile compounds (polyphenols) has been successfully carried out by dynamic extraction at 110°C and high flow-rate of 4 mL/min leading to about 90% recoveries (residence time and degradation were minimized) [31]. Mixture of water, ethanol, and formic acid (94:5:1, v/v/v) was used as extraction solvent, and recoveries were much better than in batch (static) mode or with conventional methanol extraction. Several stability/degradation studies have also been performed with water at HTs. Co et al. studied degradation of antioxidants during their pressurized hot-water extraction (PHWE) from birch bark [32]. Stabilities of preservatives used in cosmetics were studied in PHW at 100–200°C by Kapalavavi et al. [33]. They noticed that PHW up to 150°C can be applied for the chromatographic separation of these preservatives. HT (up to 150°C) and pH stability of stationary phases used in HT-LC were evaluated by Haun et al. [34]. They concluded that ethylene-bridged hybrid (BEH) technology provides still the most temperature and pH stable silica-based columns. Thermal stabilities of thiazide and other diuretics in PHW were also studied during chromatographic separation [35]. In some cases, the degradation does not cause any problems or might be even desired especially if PHW is used for soil remediation to extract, e.g., explosives [36] and polyaromatic hydrocarbons (PAHs) [37], in purification process where mycotoxins are removed in beverage production [38] or when new valuable (e.g., bioactive) compounds are formed during extraction or processing with PHW [39,40].

    2.3 Separation Techniques Utilizing Water

    As well known, water is conventional solvent in many separation techniques. Typically it is used in capillary electrophoretic (CE) separations. As described in more detail in Chapter 17, Capillary Electrophoresis as a Green Alternative Separation Technique, only two examples are given here where CE have been applied to off-line analysis of amines and amino acids from PHW extracts [41] and alkaloids from PHW extracts of Sophora flavescens Ait. [42]. Aqueous phase separations are also most common ones in field-flow-fractionation for particles, polymers and macromolecules [43,44], and many distillation (hydrodistillation and steam distillation) processes are done in water, for example, for the purification or isolation of compounds or oils from natural products (Fig. 2.2), and in some studies distillation is compared to extraction with PHW [45,46]. PHW as steam state has also been successfully used for crude oil distillation where the parameters of the distillation system were optimized and modeled [47]. Techniques, such as extraction and liquid chromatography, that fully utilize the changes in solvent properties of water when temperature is increased are described in more detail in the following subchapters.

    Figure 2.2 Green extraction and distillation processes for isolation of essential oils from plant materials. Reprinted with permission from A. Filly, A.S. Fabiano-Tixier, C. Louis, X. Fernandez, Water as a green solvent combined with different techniques for extraction of essential oil from lavender flowers, C. R. Chim. 19 (2016) 707–717. Copyright © 2016 Elsevier Masson SAS. All rights reserved.

    2.3.1 Extraction at HTs

    Extraction of hydrophobic organic compounds with organic solvent can easily be replaced by PHWE (also known as subcritical water extraction, SWE or superheated water extraction). In addition to changes in solvent properties (discussed in Section 2.2) of water with temperature, initial thermal desorption and diffusivity of the analytes from sample matrix are enhanced, thus speeding up the extraction process. PHWE is of great interest today as a green treatment technology for the biomass to isolate/produce different products and chemicals (biorefinery concept). Usually with biomass, the extraction temperatures for valuable, bioactive and thermolabile compounds have been 100–200°C [31,48,49], whereas temperatures higher than 200°C are less frequently used and when employed, the extracted analytes have been thermally stable [50–52]. SCW (supercritical fluid extraction is discussed in Chapter 7: Supercritical Fluids and Gas Expanded Liquids) is even more rarely used in extraction due to high critical temperature of water. High reactivity (corrosion) of SCW can also limit its applications, although corrosion can be minimized by tuning the density in SCW. The corrosion is worse at PHW just below critical temperature despite of density [53]. However, interesting separation system for desalination was introduced by Lean et al. [54], where SCW was utilized to precipitate salts and other type of particles and waste material in treated water following their separation in spiral separator. Both PHW at higher temperatures and SCW are suitable for various reactions, oxidation, and separation processes, where degradation of the compounds or biomolecules is not a problem or even desired (i.e., reactive extraction) as mentioned already earlier in Section 2.2. In reactive PHWE, often some catalysts or other additives are included. SCW can also be used as a pretreatment of the biomass or waste material to be processed, like for the recovery of metals from printed circuit board waste [15].

    After the PHWE, the temperature of the extract is often reduced resulting in decreased solubilities of organic molecules. This can be benefit when combining PHWE, e.g., with some microextraction techniques to recover and concentrate analytes for the analysis. Solid-phase microextraction (SPME) as a solventless technique is well suited for this purpose [55] and can be used for direct or head space sampling. Automated on-fiber derivatization has also been successfully used in PHWE-SPME-GC-MS/MS of sewage sludge [56]. Similarly, stir bar sorptive extraction (SBSE) has been useful after PHWE for aqueous and solid environmental matrices [57,58]. Liquid-phase microextraction has also been utilized for the preconcentration of PHWE extract prior to analysis either by hollow fibers [59], via Vortex-assisted liquid–liquid microextraction [60] or using dispersive liquid–liquid microextraction [61]. Additionally, monolithic capillary [62] and molecularly imprinted polymer (MIP) [63] have been used for trapping of the extracted analytes in PHWE. PHW has also been exploited for desorption in online SPE-HPLC [64], desorption of 1,1-bis-(4-chlorophenyl)-2,2-dichloroethene from nanofiber sorbents [65], and for efficient removal of MIP template molecules during preparation of MIP material with minimal template bleeding, as shown in Fig. 2.3 [66]. Membranes can also be used in conjunction with PHWE before the analysis, as in PHWE-miniaturized membrane-assisted solvent extraction of sediments [67] (membrane extraction is covered in Chapter 8: Green Membrane Extraction).

    Figure 2.3 Absorbance of quercetin in each washing solution at 10 min intervals for the different washing techniques (A), at 2 h intervals for the Soxhlet and ultrasound (B), and template bleeding after washing of the MIP material (C). Reprinted with permission from B.S. Batlokwa, J. Mokgadi, T. Nyokong, N. Torto, Optimal template removal from molecularly imprinted polymers by pressurized hot water extraction, Chromatographia 73 (2011) 589–593. Copyright © 2011 Springer-Verlag.

    PHWE can be accelerated in several ways, such as by assisting extraction with microwaves [68,69], ionic liquids [70], enzymes [71], addition of ethanol [72] or phosphate, and sonication [73]. PHWE has provided better or equivalent extraction yields when compared to many other techniques (Fig. 2.4), like hydrodistillation [45], sonication, refluxing and leaching [74], Soxhlet [46], solvent extraction with methanol or ethanol [49], and accelerated solvent extraction (ASE) [61].

    Figure 2.4 Yield of alkaloids in S. flavescens Ait. (A), and flavanols in green tea leaves (B) with different extraction techniques. PHWE (SCWE) in panel (A) was done at 100°C and in panel (B) at 150°C (SWE), ASE with ethanol at 100°C. EC, epicatechin; ECG, epicatechin gallate; EGCG, epigallocatechin gallate; Ct, catechin; CG, catechin gallate; and GCG, gallocatechin gallate. Reprinted with permission from H. Wang, Y. Lu, J. Chen, J. Li, S. Liu, Subcritical water extraction of alkaloids in Sophora flavescens Ait. and determination by capillary electrophoresis with field-amplified sample stacking, J. Pharm. Biomed. Anal. 58 (2012) 146–151 (copyright © 2012 Elsevier) and from M.-J. Ko, C.-I. Cheigh, M.-S. Chung, Optimization of subcritical water extraction of flavanols from green tea leaves, J. Agric. Food Chem. 62 (2014) 6828–6833. (copyright © 2014 American Chemical Society).

    2.3.2 Liquid Chromatography at HTs

    RPLC separation can be carried out faster with lower back pressure by increasing column temperature. Generally, mild temperatures up to 100°C have been used in separations. What is more interesting, more sustainable separations can be achieved by replacing concentration gradient (organic solvent) with HT or temperature gradient. Instead of adding methanol or acetonitrile into water, the same effect to elution strength can be obtained by raising the temperature while at the same time using pure water as eluent. To achieve polarity changes needed, temperatures from 100°C to 200°C are often required. The upper limit of the HT separation is determined by the thermal stabilities of the analytes (discussed earlier) and the stability of the column stationary phase material. Usually, temperatures around 200°C can be tolerated by zirconium oxide, polymeric or carbon-based phases in addition to some other specific stable phases [75–78]. HT-LC, subcritical water chromatography and superheated water chromatography are terms generally used for HT-LC separations. In this chapter, the term HT-LC is preferred and used from now on. In addition to RPLC-type separations, also ion-exchange chromatography (IEC) separations have been conducted at HTs [79].

    HT-LC has been successfully used with increasing number of different detectors. Inductively coupled plasma–mass spectrometry (ICP-MS) and inductively coupled plasma-atomic emission spectrometry (ICP-AES; Fig. 2.5) have been utilized after HT-LC to detect metals, organometals, and organic compounds from various sample matrices [80–82]. Evaporative light scattering detector (ELSD) has also been used after HT-LC for carbohydrates in food analysis where fast 3 min separations were achieved at 150°C [83]. In the same study, temperature gradient from 25°C to 175°C was also employed. ELSD provided fivefold increase in response when temperature was increased from 30°C to 180°C in HT-LC [84]. At the same time, much smaller temperature dependence was noticed for corona-charged aerosol detector. HT-LC with electrospray ionization-MS (ESI-MS) [85] and isotope ratio-MS (IRMS) [86,87] have been utilized for the analysis of pharmaceuticals, drug metabolites, and steroids, for example. As flame ionization detector (FID) has earlier been frequently employed as a universal HT-LC detector, only some attempts have recently been made to improve LC-FID interface design [88].

    Figure 2.5 HT-LC-ICP-AES system with pneumatic nebulization. Reprinted with permission from A. Terol, E. Paredes, S.E. Maestre, S. Prats, J.L. Todolí, High-temperature liquid chromatography inductively coupled plasma atomic emission spectrometry hyphenation for the combined organic and inorganic analysis of foodstuffs, J. Chromatogr. A 1217 (2010) 6195–6202. Copyright © 2010 Elsevier.

    The performance of stationary phases at HT is critical for the successful HT-LC. With monolithic column, excellent column performance was maintained after continuous use at over 200°C during 1000 column volumes [89]. Special bonded hybrid column was used for the separation of steroids [90] and zirconia-based columns (often used in HT-LC) were utilized for parabens from 100°C to 180°C [75]. Applicability of new stationary phases, including poly(2-hydroxyethylmethacrylate-N-methacryloyl-(L)-histidine-methyl-ester) (NA-PHEMAH) polymeric phase [77] and monodisperse porous poly(glycidylmethacrylate-co-ethylenedimethacrylate) microspheres [91], has been studied. NA-PHEMAH was shown to be stable over 500 h at 150°C and could be used even at 200°C with pure water or few percent methanol in water. Polymeric microspheres were stable at 150°C during 160 injections and chromatographic peaks were found symmetrical with column efficiency up to

    Enjoying the preview?
    Page 1 of 1