You are on page 1of 13

Edited with the trial version of

Foxit Advanced PDF Editor


To remove this notice, visit:
www.foxitsoftware.com/shopping

ChapTEr 1

Introduction to High-pressure
Solvent Systems
ThomaS m. aTTarD*a aND aNDrEw J. hUNT*b
a
Green Chemistry Centre of Excellence, Department of Chemistry, The
University of York, heslington, York Yo10 5DD, UK; bmaterials Chemistry
research Center, Department of Chemistry, Faculty of Science, Khon Kaen
University, Khon Kaen, 40002, Thailand
*E-mail: thomas.attard@york.ac.uk, andrew@kku.ac.th

1.1   Green Chemistry
Sustainable development, i.e. meeting the present needs without affecting
future generations' ability to meet their own requirements, has been a major
worldwide issue for several decades, having been addressed at the 1987
Brundtland Commission (United Nations Commission on Environment and
Development).1
Two key aspects, arising from sustainable development, that are valid from
a chemical, engineering and energy point of view are: (i) the rate at which the
current generation can consume fossil fuels and (ii) the quantities of waste
that the environment can support on a sustainable basis.
The Earth has a natural capacity to cope with the waste and pollution
produced by society and, when this capacity is surpassed, unsustainabil-
ity results.1,2 Furthermore, it is becoming increasingly clear that the rate


Green Chemistry Series No. 57
Supercritical and Other High-pressure Solvent Systems: For Extraction, Reaction and Material
Processing
Edited by Andrew J. Hunt and Thomas M. Attard
© The Royal Society of Chemistry 2018
Published by the Royal Society of Chemistry, www.rsc.org

1
Edited with the trial version of
Foxit Advanced PDF Editor
To remove this notice, visit:
www.foxitsoftware.com/shopping

2 Chapter 1
at which non-renewable fossil feedstocks, i.e. coal, oil and natural gas, are
being consumed is much higher than the rate at which natural geological
processes can replace them.3 This makes their use, in the long run, unsus-
tainable. Using fossil fuels also leads to the production of high rates of car-
bon dioxide (Co2), rates that are much higher than what can be assimilated
by the environment, something that the scientific community widely accepts
as the leading problem of climate change.3
So as to tackle this sustainability issue, the United States Environmental
protection agency (Epa) coined the term Green Chemistry in the early 1990s,
defined as follows:

“To promote innovative chemical technologies that reduce or eliminate the use
or generation of hazardous substances in the design, manufacture and use of
chemical products.”4

This definition indicates that green chemistry is a philosophy rather than


some new type of chemistry, whereby the practice of chemistry and engineer-
ing should be done in a sustainable manner. with this in mind, paul anastas
and John warner developed the 12 principles of Green Chemistry in 1998,
shown in Figure 1.1.5
one strategy for meeting the principles of green chemistry is to reduce the
use of hazardous organic solvents and to encourage the utilisation of sustain-
able or more environmentally friendly greener solvents. Several of the twelve
principles of green chemistry have a strong connection with the use of super-
critical, superheated or pressurised solvents. The most obvious of which are
number 5 “the use of auxiliaries (e.g. solvents) should be made unnecessary
whenever possible and innocuous when used” and number 3 “synthetic
methodologies should be designed to use and generate substances that pos-
sess little or no toxicity to human health and the environment”. Common
supercritical solvents utilised throughout the research community include
carbon dioxide and water, both of which are relatively non-toxic, abundant
and can be easily recycled.
Legislation surrounding solvent registration, application, disposal and
emission have been introduced and enforced in order to ensure that solvents
are safely used in industry.6 Legislation has effectively led to the reduction in
the use of more toxic solvents via the complete prohibition of their use or by
introducing maximum residual levels. a number of solvent guides have been
developed in order to recommend to industry suitable alternatives to sol-
vents that are deemed undesirable.7–9 The ChEm21 (Chemical manufactur-
ing methods for the 21st Century pharmaceutical Industries) solvent guide,
which ranks the solvent greenness based on a benchmark of existing solvent
selection guides, highlights water as a recommended solvent (Figure 1.2).9
while water is regarded as being favourable, Figure 1.2 demonstrates that
non-polar solvents, such as the hydrocarbons, are all regarded as hazardous
or problematic due to consistently bad safety and environmental scores.9
although supercritical fluids such as carbon dioxide and also gas expanded
liquids were not included in the guide, the authors did highlight that the
Edited with the trial version of
Foxit Advanced PDF Editor
To remove this notice, visit:
www.foxitsoftware.com/shopping

Introduction to High-pressure Solvent Systems 3

Twelve principles of Green Chemistry.3,5 reproduced from ref. 3 with


permission from the royal Society of Chemistry.

project ChEm21 did involve the use of these solvent systems.9 Supercritical
carbon dioxide can be utilised as a greener alternative to traditional non-po-
lar solvents such as the hydrocarbons, e.g. hexane.10 pressurised solvent
systems can improve efficiency and reduce solvent usage. however, super-
critical, superheated or pressurised solvents are consistently criticised for
being energy intensive processes and, as such, this must be considered in
any environmental and economic evaluation of such a process.11

1.2   Supercritical Fluids
The standard definition of a supercritical fluid is any substance that is above
its critical temperature and pressure (Pc, Tc).12 Variations in the temperature
and pressure lead to a change in the physical properties of a substance; a
phenomenon that can be better explained by referring to a phase diagram.13
Edited with the trial version of
Foxit Advanced PDF Editor
To remove this notice, visit:
www.foxitsoftware.com/shopping

4 Chapter 1

ChEm21 solvent selection guide for common solvents.9 reproduced


from ref. 9 with permission from the royal Society of Chemistry.

Figure 1.3 represents a typical P–T phase diagram for a substance. all sub-
stances that are stable are said to have a triple point and a critical point.14
The former refers to the pressure and temperature conditions at which the
solid, liquid and gas phases coexist as they are in equilibrium with each
other. The line that moves from the triple point to the critical point, found
lying between the liquid and gaseous regions, is called the gas–liquid (G–L)
coexistence curve.15 when moving towards the critical point along this G–L
curve, thermal expansion causes the density of the liquid to decrease while
the increase in pressure causes the density of the gas to increase; the den-
sities of the two phases become equivalent at the critical point, making it
no longer possible to distinguish between the liquid and gas phases due to
them having identical properties. Therefore, the critical point, consisting
of the critical temperature and critical pressure, can be described as being
Edited with the trial version of
Foxit Advanced PDF Editor
To remove this notice, visit:
www.foxitsoftware.com/shopping

Introduction to High-pressure Solvent Systems 5

phase diagram (P–T diagram) of a substance.

Critical pressures and temperatures of various compounds.


Compound pressure (bar) Temperature (°C)
ammonia 113.2 132.4
Carbon dioxide 73.8 31.1
methane 46 −82.8
Ethane 48.7 32.2
propane 42.5 96.7
Ethene 50.4 9.2
methanol 80.9 239.5
Ethanol 61.4 240.8
acetone 47 235
Nitrous oxide 33.4 73.5
water 374.2 220.5

the maximum pressure and temperature applied wherein a substance


exists as a liquid and a gas in equilibrium with one another. Beyond the
critical point, there is no longer a distinction between the liquid and gas
phases.13,15 The critical points for various common supercritical fluids can
be found in Table 1.1.
From a macroscopic point of view, the G–L coexistence curve terminates
at the critical point. This is the standard explanation and it is assumed that,
once the critical point is surpassed, the substance acts as a homogeneous
fluid (no longer a biphasic heterogeneous system). however, recent work has
shown that, even under supercritical conditions, there are two regions that
Edited with the trial version of
Foxit Advanced PDF Editor
To remove this notice, visit:
www.foxitsoftware.com/shopping

6 Chapter 1


Thermophysical properties of liquids, SCFs and gases.


Diffusion Coefficient
Fluid Density (kg m−3) (m2 s−1) Viscosity (N s m−2)
Liquid 800–1200 10−8–10−9 10−3–10−2
Supercritical fluid 250–800 10−7–10−8 10−4–10−3
Gas 1–100 10−4–10−5 10−5–10−4
Edited with the trial version of
Foxit Advanced PDF Editor
To remove this notice, visit:
www.foxitsoftware.com/shopping

Introduction to High-pressure Solvent Systems 7


Dobbs and Johnston carried out solubility measurements for different
systems (binary, ternary and quaternary), which consisted of scCo2, a co-
solvent and combinations of solid phases. In the ternary systems composed
of two solutes and the supercritical phase, a proportional increase in one
solute's solubility was observed relative to the second solute's solubility.
This causes an entraining effect, wherein the more soluble solute leads to
an enhancement of the less soluble one. as an example, considering naph-
thalene and phenanthrene, the former has a much higher solubility in Co2
than the latter. however, in a system comprising scCo2, naphthalene and
phenantrene, the solubility of phenanthrene is raised by naphthalene by
75%. Therefore, the additions of certain molecules to SCFs leads to a varia-
tion in the solvent properties.29
Dobbs et al. found that adding small amounts of numerous co-solvents
leads to a noticeable improvement in the non-polar SCF selectiveness for
polar vs non-polar solid compounds.29,30 an enhancement in the solubility of
certain solids (of higher polarity) in scCo2 was observed with the addition of
several mol% of different co-solvents. Upon addition of around 3.5 mol% of
methanol, there was an increase in the solubility of 2-aminobenzoic acid in
scCo2 by 620%.29,30

1.2.2   Supercritical Carbon Dioxide
By far, the most popular SCF is scCo2 as it is an ideal solvent for a variety
of different applications, ranging from extraction and separation (chro-
matography) to reactions, processing of materials and power genera-
tion.31 one can exploit the advantages of near-critical operation at low
temperatures (<35 °C) due to the Co2 relatively low Tc. Co2 has favour-
able health, safety and environmental characteristics, being non-flamma-
ble and demonstrating very low toxicity. Furthermore, it is cheap, widely
available, unregulated and easily recyclable. The relatively high Pc of Co2
(73.8 bar) may be seen as a disadvantage. however, it has become fairly
routine to operate at such pressures in industrial processes like industrial
scCo2 extraction (such as in the decaffeination of coffee and extraction of
hops).31,32
scCo2 has a polarity that is quite similar to the polarities of hexane and
toluene.33,34 The addition of probe molecules, for example reichardt's dye
[2,6-diphenyl-4-(2,4,6-triphenylpyridinio) phenolate], is a practical and
easy way to measure the polarity of a solvent.35,36 Since reichardt's dye is
a zwitterionic molecule, it demonstrates solvatochromic effects as a result
of the ground state of the dye interacting with the solvent.35 The ET(30)
and ETN scales are the empirical scales of solvent polarity associated with
reichardt's dye:
  
28951
ET  30   kcalmol 1   (1.1)
   max
abs
 nm 
Edited with the trial version of
Foxit Advanced PDF Editor
To remove this notice, visit:
www.foxitsoftware.com/shopping

Chapter 1

 ET solvent  ET TMS  
ETN   (1.2)
 ET water ET TMS 

 ET solvent  30.7  
ETN   (1.3)
32.4
  
Kamlet and Taft proposed that the interaction between the phenoxide
oxygen and the solvent in question is around 2/3 of the shift of the maxi-
mum absorption wavelength of reichardt's dye. From ET(30), values of 30.7
kcal mol−1 and 63.1 kcal mol−1 are obtained for tetramethylsilane (TmS)
(highly non-polar solvent) and water (highly polar solvent), respectively.
These values can then be used to normalise the scale such that a dimen-
N N
sionless figure is obtained, whereby ET = 0 for tetramethylsilane and ET =
35
1 for water.
N
as can be seen from Table 1.3, the ET value for scCo2 lies between those of
hexane and toluene. It can therefore be taken as a general rule of thumb that
compounds of low molar mass, which have appreciable vapour pressures,
that are soluble in hexane should also dissolve in scCo2 (this is, however, not
always the case).

1.2.3   Subcritical and Supercritical Water
with chemists and industrialists continuously attempting to develop greener
and more environmentally benign chemical processes, there is ever-growing
attention towards using water close to or above the critical point (374 °C,
220.5 bar) as a medium in chemical processes.37 This is due to the fact that
the replacement of conventional organic solvents with subcritical (near-crit-
ical) or supercritical water in chemical processes can have environmental
advantages as well as prevent pollution. Significant research has been ded-
icated to utilising sub- and supercritical water in a variety of applications,
including organic chemistry, biomass processing, waste treatment, synthetic
fuel production, synthesis of materials and geochemistry.37,38
There are substantial differences between the physical properties of
ambient liquid water and the physical properties of water near the critical

Table 1.3   ETN values for a variety of solvents.


Solvent type ETN values
TmS 0
hexane 0.009
Supercritical carbon dioxide 0.012–0.034
Toluene 0.099
Dichloromethane (DCm) 0.309
methanol 0.762
water 1
Edited with the trial version of
Foxit Advanced PDF Editor
To remove this notice, visit:
www.foxitsoftware.com/shopping

Introduction to High-pressure Solvent Systems 9


38
point. Near-critical water (NCw) has a much lower polarity (dielectric
constant, α and ETN) and there is a significant decrease in the number and
persistence of hydrogen bonds. This leads to much higher solubilities of
organic compounds in NCw (it behaves more like typical organic solvents)
compared to water at room temperature and there is complete miscibility
of organic compounds with supercritical water (SCw). as a general rule of
thumb, the ETN of NCw is similar to that of acetone, while at higher tem-
peratures it has complete miscibility with toluene.39 NCw also has lower
density, surface tension and viscosity compared to ambient liquid water;
however, with the changing temperature, the diffusivity and specific heat
capacity of NCw increases while the hydrogen-bond acceptor ability (β)
remains constant.38
Since the dissociation constant (Kw) increases with the increasing tem-
perature, the concentration of h3o+ (hydronium) and oh− (hydroxide) ions
increases and, as water approaches the critical point, its Kw is three orders
of magnitude higher than that of water at room temperature.39 Therefore,
dense high-temperature water can be used effectively for acid- and base-ca-
talysed organic reactions. In fact, in certain acid-catalysed reactions, the h+
concentration of NCw is high enough that no additional acid needs to be
introduced. however, as the critical point of water is surpassed (becomes
SCw), there is a dramatic decrease in Kw – as an example, at conditions of
600 °C and 253.3 bar, Kw is approximately nine orders of magnitude lower
than that of water at room temperature, making SCw a poor solvent for ionic
chemistry in this low-density, high-temperature medium.38,39
The dielectric constant of water increases when increasing the density
and decreases when increasing the temperature. The typical high ε value
of 80 only takes place at low temperatures in a small region. ε values of
10–25 occur in a large supercritical region at high density; which are sim-
ilar to the ε values of dipolar solvents such as acetone or acetonitrile at
ambient conditions (shown in Figure 1.4).40 ε values of 10–25 are ade-
quately large enough for dissolving and ionising electrolytes while also
allowing for miscibility with non-polar molecules. The dielectric constant
rapidly decreases at low densities resulting in a decrease in the ability to
dissolve and ionise electrolytes. The dielectric constant has a value of six
at the critical point.39
a further difference between SCw and water at room temperature is that
the continuous variation of the properties (including Kw, dielectric constant,
viscosity, etc.) occurs over much larger ranges under supercritical conditions,
which enables the possibility of fine-tuning the properties of the reaction
medium for achieving optimal results by varying the pressure and tempera-
ture conditions.28

1.2.4   Pressurised Solvent Systems
pressurised liquids, including gas expanded liquids, have been utilised as
solvent systems for extractions and as reaction media.41,42
Edited with the trial version of
Foxit Advanced PDF Editor
To remove this notice, visit:
www.foxitsoftware.com/shopping

10 Chapter 1

Diagram illustrating the density, dielectric constant (static) and ion dis-
sociation constant (Kw) of water as a function of the temperature at 300
bar. There is a substantial drop in the Kw of water as T increases and, at
supercritical conditions, it becomes similar to that of a less-polar sol-
vent. reproduced from ref. 40 with permission from the royal Society
of Chemistry.

pressurised liquid systems involve the use of liquid solvents at elevated


temperature and pressure, to enhance the performance as compared to
processes carried out at room temperature and atmospheric pressure.43
Utilising solvents at temperatures above their atmospheric boiling point
enhances the solubility and mass transfer properties, thus resulting in
higher diffusion rates within the system. The elevated pressure maintains
the solvent below its boiling point and in its liquid state. pressurised
liquid extraction (pLE) has been used to reduce extraction times and
improve extraction yields, whilst reducing the solvent consumption.44 a
key advantage of pLE is that solvent mixtures can be tuned to enhance
both the efficiency and selectivity of the extraction. Furthermore, auto-
mated pressurised solvent systems may also protect light or oxygen sensi-
tive compounds.42 pLE technologies have been demonstrated as effective
extraction systems for a range of ‘bioactive compounds’ from foods and
herbal plants.42
a gas-expanded liquid (GXL) is a mixture of a compressible gas dissolved in
an organic solvent. Significant work has been carried out to develop generic
approaches to assess and design gas-expanded liquids.41 The properties of
carbon dioxide-expanded liquids (CXLs) span the range from those of the pure
organic solvent to those of carbon dioxide.45 Variation of the ratio between
the solvent and carbon dioxide can lead to tuneable solvent properties.
Edited with the trial version of
Foxit Advanced PDF Editor
To remove this notice, visit:
www.foxitsoftware.com/shopping

Introduction to High-pressure Solvent Systems 11


These solvent systems have demonstrated promise in applications such as
separations, extractions, reactions, fine particle precipitation, polymer pro-
cessing and other applications.46 Carbon dioxide has a number of advan-
tages for use in such processes, including being inert, widely available, low
cost, non-flammable and easily removed at the end of the process. CXLs
demonstrate high gas miscibility, enhanced transport rates compared to
conventional solvent systems and they require lower pressures compared
to supercritical carbon dioxide.45 The use of such systems also reduces the
volume of organic solvent required, thereby helping to prevent waste in
chemical processes (a key principle of green chemistry).

1.3   Conclusions
pressurised-solvent systems including supercritical fluids, sub-critical flu-
ids and gas-expanded liquids are becoming ever-more important, especially
with the growing legislation restricting the use of conventional organic sol-
vents in industrial chemical processes, as well as the strict guidelines for
minimal solvent residues in consumer products. The use of high-pressure
systems enables the possibility to design clean, sustainable and environmen-
tally friendly processes, as well as generate novel products with unique prop-
erties. The use of supercritical fluids and also pressurised solvents systems is
of global interest for use in extraction, reaction and materials processing. In
the following chapters, examples of how these pressurised-solvent systems
can be utilised in a variety of different processes as effective alternative sol-
vents for obtaining different products will be presented, including the latest
innovative research on the use of supercritical, subcritical fluids and gas-
expanded liquids in extractions, reactions, processing of materials and
power generation.

References
1. Report of the World Commission on Environment and Development: Our
Common Future, http://www.un-documents.net/our-common-future.pdf,
accessed 11th November, 2014.
2. J. h. Clark, Green Chem., 2006, 8, 17–21.
3. r. a. Sheldon, Green Chem., 2016, 18, 3180–3183.
4. m. Lancaster, Green Chemistry: An Introductory Text, royal Society of
Chemistry, UK, 2010.
5. p. T. anastas and J. C. warner, Green chemistry: Theory and Practice, oxford
university press, UK, 2000.
6. F. m. Kerton and r. marriott, Alternative Solvents for Green Chemistry,
royal Society of Chemistry, Cambridge, UK, 2013.
7. r. K. henderson, C. Jimenez-Gonzalez, D. J. C. Constable, S. r. alston, G.
G. a. Inglis, G. Fisher, J. Sherwood, S. p. Binks and a. D. Curzons, Green
Chem., 2011, 13, 854–862.
Edited with the trial version of
Foxit Advanced PDF Editor
To remove this notice, visit:
www.foxitsoftware.com/shopping

12 Chapter 1
8. D. prat, o. pardigon, h.-w. Flemming, S. Letestu, V. Ducandas, p. Isnard,
E. Guntrum, T. Senac, S. ruisseau, p. Cruciani and p. hosek, Org. Process
Res. Dev., 2013, 17, 1517–1525.
9. D. prat, a. wells, J. hayler, h. Sneddon, C. r. mcElroy, S. abou-Shehada
and p. J. Dunn, Green Chem., 2016, 18, 288–296.
10. E. h. K. Sin, r. marriott, a. J. hunt and J. h. Clark, C. R. Chim., 2014, 17,
293–300.
11. T. attard, C. mcElroy and a. hunt, Int. J. Mol. Sci., 2015, 16, 17546.
12. C. a. Eckert, Nature, 1996, 383, 313–318.
13. r. S. oakes, a. a. Clifford and C. m. rayner, J. Chem. Soc., Perkin Trans. 1
1, 2001, 917–941.
14. r. Noyori, Chem. Rev., 1999, 99, 353–354.
15. E. D. ramsey, B. minty and r. Babecki, in Analytical Supercritical Fluid
Extraction Techniques, ed. E. D. ramsey, Springer, Netherlands, 1998, ch.
4, pp. 109–157.
16. G. G. Simeoni, T. Bryk, F. a. Gorelli, m. Krisch, G. ruocco, m. Santoro and
T. Scopigno, Nat. Phys., 2010, 6, 503.
17. p. F. mcmillan and h. E. Stanley, Nat. Phys., 2010, 6, 479.
18. a. Clifford and T. Clifford, Fundamentals of Supercritical Fluids, oxford
University press, 1999.
19. S. K. Kumar and K. p. Johnston, J. Supercrit. Fluids, 1988, 1, 15–22.
20. m. E. paulaitis, V. J. Krukonis, r. T. Kurnik and r. C. reid, Rev. Chem.
Eng., 1983, 1, 179–250.
21. h. Ebeling and E. U. Franck, Ber. Bunsenges. Phys. Chem., 1984, 88,
862–865.
22. K. p. Johnston and C. a. Eckert, AIChE J., 1981, 27, 773–779.
23. G. a. m. Diepen and F. E. C. Scheffer, J. Phys. Chem., 1953, 57, 575–577.
24. r. T. Kurnik, S. J. holla and r. C. reid, J. Chem. Eng. Data, 1981, 26, 47–51.
25. w. Leitner, Nature, 2000, 405, 129–130.
26. r. L. mendes, B. p. Nobre, m. T. Cardoso, a. p. pereira and a. F. palavra,
Inorg. Chim. Acta, 2003, 356, 328–334.
27. m. Fattori, N. r. Bulley and a. meisen, J. Agric. Food Chem., 1987, 35,
739–743.
28. a. Baiker, Chem. Rev., 1999, 99, 453–474.
29. J. m. Dobbs and K. p. Johnston, Ind. Eng. Chem. Res., 1987, 26, 1476–1482.
30. J. m. Dobbs, J. m. wong, r. J. Lahiere and K. p. Johnston, Ind. Eng. Chem.
Res., 1987, 26, 56–65.
31. B. Subramaniam, r. a. rajewski and K. Snavely, J. Pharm. Sci., 1997, 86,
885–890.
32. m. a. mchugh and V. J. Krukonis, Supercritical Fluid Extraction: Principles
and Practice, Butterworth-heinemann, 1994.
33. w. w. Christie and X. han, Lipid Analysis: Isolation, Separation, Identifica-
tion and Lipidomic Analysis, pJ Barnes & associates, 2010.
34. m. m. r. de melo, a. J. D. Silvestre and C. m. Silva, J. Supercrit. Fluids,
2014, 92, 115–176.
Edited with the trial version of
Foxit Advanced PDF Editor
To remove this notice, visit:
www.foxitsoftware.com/shopping

Introduction to High-pressure Solvent Systems 13


35. C. reichardt, Chem. Rev., 1994, 94, 2319–2358.
36. K. a. Fletcher, I. a. Storey, a. E. hendricks, S. pandey and S. pandey, Green
Chem., 2001, 3, 210–215.
37. p. E. Savage, Chem. Rev., 1999, 99, 603–622.
38. a. Loppinet-Serani, C. aymonier and F. Cansell, J. Chem. Technol. Biotech-
nol., 2010, 85, 583–589.
39. h. weingärtner and E. U. Franck, Angew. Chem., Int. Ed., 2005, 44,
2672–2692.
40. a. a. peterson, F. Vogel, r. p. Lachance, m. Froling, J. m. J. antal and J. w.
Tester, Energy Environ. Sci., 2008, 1, 32–65.
41. E. Siougkrou, a. Galindo and C. S. adjiman, Chem. Eng. Sci., 2014, 115,
19–30.
42. a. mustafa and C. Turner, Anal. Chim. Acta, 2011, 703, 8–18.
43. V. Camel, Analyst, 2001, 126, 1182–1193.
44. a. p. D. F. machado, J. L. pasquel-reátegui, G. F. Barbero and J. martínez,
Food Res. Int., 2015, 77, 675–683.
45. p. G. Jessop and B. Subramaniam, Chem. Rev., 2007, 107, 2666–2694.
46. a. m. Scurto, K. hutchenson and B. Subramaniam, Gas-expanded Liquids
and Near-critical Media, american Chemical Society, 2009, vol. 1006, ch.
1, pp. 3–37.

You might also like