Вы находитесь на странице: 1из 30

Solvents for Sustainable Chemical Processes

A Seminar Submitted to
Institute of Chemical Technology
Mumbai

Submitted by
Ameya Nitin Kulkarni
B. Chem. Engg.
2016-2017

BNT
Index

1. Introduction ………………………………………………..………………………….1

2. Scope and Objective

2.1 Scope ……………………………………………………………………………...2

2.2 Objective ………………………………………………………………………….2

3. Work done and Literature Survey

3.1 Switchable Solvents …………………………………………………….…….…...3

3.1.1 Sulfolenes …………………………………………………………………...3

3.1.2 Reversible Ionic Liquids ……………………………………….……………6

3.2 Water at Elevated Temperature (WET) ………………………………………… 10

3.3 Supercritical Fluids (SCFs) ………………………………………………...……12

3.3.1 Supercritical Carbon-Dioxide ………………….……..……………………14

3.4 Gas-Expanded Liquids (GXLs) and Organic Aqueous Tunable Solvents


(OATS)……………………………………………………….…………………. 18

3.4.1 Gas-Expanded Liquids ……………………………………………….……18

3.4.2 Organic Aqueous Tunable Solvents……………………………………......20

4. Industrial Relevance – Case Studies ………………………………………………...24

5. Conclusion …………………………………………………………………………...25

6. Nomenclature ………………………………………………………………………..26

7. References …………………………….…………………..……………………...….27
Introduction

1. Introduction
Before venturing into solvents for sustainable processes, let us get clarity on what is a
sustainable process or sustainable development? The Brundtland Commission of the United
Nations (BC-UN) presented the following definition of sustainable development:
“development that meets the needs of the present without compromising the ability of future
generations to meet their own needs.” This definition implies that we not only have to
concentrate on green processes but also on their applicability in today’s industry from an
economic point of view. Today’s chemical industry has developed certain parameters which
are able to predict the life-cycle or “greenness” of a particular process (carbon efficiency,
mass efficiency, E-factor, etc.) which is reported by R. A. Sheldon et al., 2000. A better way
would be to incorporate green technology into the early stages of process development rather
than remediation which can prove to be expensive and not very effective.
We encounter solvents in almost all wet chemical processes. Also, they contribute
majorly to all industry waste streams. Here, we deal with “green and sustainable solvents”
which can be defined as solvents that address environmental issues, contribute to the
optimization of the overall process, and are cost-effective. Every chemical process has certain
constraints such as operating conditions, type of catalyst employed, method of solvent
recycling etc. Thus, we cannot envisage a common universal green solvent but we can make
an attempt to generate a portfolio of solvents which can offer enough flexibility to the end
user.
Qualitatively, a solvent is defined as any liquid, gas, solid, gas-expanded liquid or
supercritical fluid in which a solute is dissolved, either partially or completely. Solvents have
countless applications and often play multiple roles in chemical processes. For instance, they
bring reactants together to facilitate reaction and provide a means of temperature control for
both endothermic and exothermic transformations. In academic laboratories, the scale of
reaction is often small enough that solvent recovery is not usually considered. In industry,
however, the choice of solvent system is dependent upon its ability to facilitate a particular
reaction, promote a facile separation of product, and potentially be recovered and recycled.
When multiple solvents are employed in a single process (solvents are often changed for a
given reaction or isolation/purification step) the associated cost and waste stream issues are
compounded. A more holistic approach to solvent choice is envisioned. Solvent systems
whose properties can be modified at will are to be designed in process systems.

Solvents for Sustainable Chemical Processes 1


Scope and Objective

2. Scope and Objective of the Seminar


2.1 Scope:

After going through research articles and performing a literature survey, it was found
that the following solvents have a potential to be used in the industry from the sustainable and
economically advantageous point of view. Thus, the scope of this survey is limited to:

1) Switchable solvents, specifically sulfolenes and reversible ionic liquids (RevILs)


2) Water at elevated temperature (WET)
3) Supercritical fluids (SCF’s), with emphasis on supercritical CO2
4) Gas-expanded liquids (GXL’s) and organic-aqueous tunable solvents (OATS)

2.2 Objective:
There are certain challenges present today with the use of current solvents. The issues
today are mainly with regard to solving the problems associated with the use of solvents in
chemical reactions, in product separations and to make repeated use of solvents i.e. their
recycling. If we try to address each of these issues separately but solving them in such a way
that they complement each other, the overall configuration will indeed lean towards the
sustainable path. The following article tries to do a comprehensive survey of possible
solvents which can be used in today’s times without hampering the process and without
introducing drastic changes in the process at the same time ensuring sustainability and
economic viability.
This comprehensive survey deals with not only modes of synthesis for these solvents
but also immediate possible applications to the real world industries and properties of these
solvents (both physical and chemical). While discussing these solvents or rather solvent
systems, attempt has been made to compare them to normal gases and liquids (fluids)
presently in use as solvents in the industry.

So, the following critical review deals with physicochemical properties and
applications of: (1) Switchable solvents, specifically sulfolenes and reversible ionic liquids
(RevILs). (2) Water at elevated temperature (WET), (3) Supercritical fluids (SCF’s), with
emphasis on supercritical CO2, (4) gas-expanded liquids (GXL’s) and organic-aqueous
tunable solvents (OATS)

Solvents for Sustainable Chemical Processes 2


Work done and Literature Survey

3. Work done and Literature Survey


3.1 Switchable Solvents:
A switchable solvent is a unique kind of solvent that can be modified using the two operating
variables namely temperature and/or pressure for optimizing reactions and separations. The
two classes of switchable solvents that will be discussed are: (1) sulfolenes such as piperylene
sulfone (PS) — a volatile and recyclable DMSO substitute and (2) one and two component
ionic liquids–solvent systems that can be reversibly switched between a molecular liquid and
an ionic liquid.

3.1.1 Sulfolenes – DMSO Substitutes:


DMSO ideally would be a good solvent due to its ability to dissolve both organic and
inorganic (partially or completely) solutes. Separation of the solvent becomes difficult due to
its high boiling point (189 °C at 51.7 Torr). Thus, separation is tried by employing phase
separation with addition of water. Thus, recycling of DMSO being expensive is not even
considered today as reported by P.J. Dunn et al., 2010. It can be seen below that for most of
the properties like dipole moment, dielectric constants are similar for both solvents. In
addition, the solvatochromic parameters are also comparable although PS has a lower β value
(0.46) compared to DMSO (0.76) suggesting a lower hydrogen bond donor ability which was
reported by D. Vinci et al., 2007.

Table 1: Physical Properties of DMSO and Piperylene sulfone


Property DMSO Piperylene Sulphone
Boiling Point (°C) 189 (34/3 Torr) 85 (7 Torr)
Melting Point (°C) 16-19 -12
Dipole Moment (D) 4.27 5.32
α 0 0
β 0.76 0.46
Π* 1.00 0.87
ET (30) (kJ/mol) 189 189
ε 46.7 42.5

Reverse chelotropic reactions of sulfolenes to form their corresponding diene and sulfur
dioxide are known. Thermal treatment of a sulfolene reverses the reaction (retrocheletropic
reaction) with the reformation of the conjugated diene and sulfur dioxide. Experimentally,
synthesis of piperylene sulfone from trans-1, 3-pentadiene (trans-piperylene) and sulfur
dioxide in the presence of a radical inhibitor was reported. (L.R. Drake et al., 1946) (Fig.1)

Solvents for Sustainable Chemical Processes 3


Work done and Literature Survey

Fig 1: Synthesis of Piperylene sulfone


PS can undergo a retrocheletropic process—a property which can facilitate product isolation
and enable solvent regeneration (recycling). Indeed, at temperatures greater than 100 °C
piperylene sulfone cleanly and efficiently decomposes back to gaseous trans-1, 3-pentadiene
(B.P. 42 °C) and sulfur dioxide. (B.P. −10 °C) (Fig. 2) The products of this decomposition
can be collected and subsequently be allowed to react in order to reform the piperylene
sulfone solvent.

Fig 2: Reversible decomposition of piperylene sulfone into transpiperylene and sulfur


dioxide.
Reactions in Piperylene Sulfone:
A series of nucleophilic substitution reactions were conducted in the two solvents. Vinci et al.
reported the rates of reaction of benzyl chloride with a variety of anionic nucleophiles in both
piperylene sulfone and DMSO at 40 °C. Most of the anionic nucleophilic salts are only
partially soluble in the solvents. Thus, the reaction system is heterogeneous and is shown in
Fig 3. The results are shown in Table 2.

Fig 3: Reaction of Benzyl Chloride with Anionic Nucleophiles

Solvents for Sustainable Chemical Processes 4


Work done and Literature Survey

Table 2: Second order rate constants for nucleophilic displacement reactions in


DMSO, DMSO with 3% water, PS, and PS with 1% water at 40 °C
DMSO (3%
Nucleophile DMSO H2O) Piperylene Sulfone Piperylene Sulfone (1% H2O)
KTA >1800 >1800 >1800 >1800
NaPDTC >1800 >1800 >1800 >1800
1.4
KSCN (±0.1) 1.7 (±0.1) 2.1 (±0.1) 2.3 (±0.2)
3.4
KOAc (±0.1) 11.0 (±0.1) 0.013 (±0.004) 0.19 (±0.01)
5.8
KCN (±0.8) 17 (±5) No Rxn 0.15 (±0.01)
CsN3 69 (±1) 16.7 (±0.9) 2.4 (±0.9) 5.8 (±0.9)
22.7
CsOAc (±0.6) 16.4 (±0.2) 0.35 (± 0.04) 0.35 (± 0.06)

The rates in PS are slower but the yields are comparable. Surprisingly, no reaction was
observed between benzyl chloride and potassium cyanide in anhydrous PS. Only in the
presence of trace amounts of water (0.1 wt %) did the reaction proceed. As a result of this
observation, reaction rates were screened in the presence of trace quantities of water. In PS
and DMSO alike, it was speculated that the addition of water increased the mass transfer and
possibly the solubility of the salts in the organic phase, thus increasing the overall reaction
rates.
Piperylene sulfone was investigated by Jiang et al. as a recyclable green replacement
for DMSO in room temperature copper-catalyzed aerobic oxidation of alcohols. Very high
turnover frequency (over 31 h−1) for the room temperature aerobic oxidation of alcohols in
piperylene sulfone was observed. The immiscibility between piperylene sulfone and n-
pentane facilitated product isolation as well as recovery and reuse of the catalytic system.
Various copper salts were examined as the pre-catalyst in the presence of DMAP (4-
dimethylaminopyridine) and acetamido-TEMPO for the oxidation of 4-methoxybenzyl
alcohol to 4-methoxy benzaldehyde (Fig. 4)

Fig 4: Oxidation of 4-methoxybenzyl alcohol in piperylene sulfone.

Solvents for Sustainable Chemical Processes 5


Work done and Literature Survey

Copper (I) bromide was shown to be optimal, yielding the highest turnover frequency (33.3
h−1). Furthermore, using H-NMR analysis, the authors could not detect any over-oxidized
product (4-methoxybenzoic acid). The recyclability of the optimized catalytic system (copper
bromide, CuBr, acetamido-TEMPO, and DMAP) for aerobic oxidation of benzyl alcohol in
piperylene sulfone was also reported. The schematic of the overall process is shown in Fig. 5.

Fig 5: Process schematic for the PS-mediated oxidations of alcohols.


The product, benzaldehyde, was extracted with n-pentane in 98% yield. The recovery and
reuse of the catalytic/solvent system was easily accomplished because piperylene sulfone and
the catalytic components (CuBr, acetamidoTEMPO, and DMAP) were insoluble in n-
pentane. After the extraction of benzaldehyde from the reaction mixture with n-pentane,
another reaction cycle was started by simply adding fresh benzyl alcohol. The catalytic
systems proved to be readily recyclable for two additional runs with only a slight drop in
activity, 88% and 82% yield respectively. Thus, the PS phase containing the catalysts can be
recycled as long as the catalyst activity remains high (Recycle Stream (1). Once the catalyst
system is no longer functioning effectively the piperylene sulfone can be decomposed into
piperylene and SO2. The PS then can be reformed and recycled (Recycle stream (2) and the
catalyst residue disposed. This contrasts with a DMSO mediated process where both DMSO
and the catalyst system would have to be disposed.

3.1.2 Reversible Ionic Liquids:


Salts with melting points near or below ambient temperature are classically called ionic
liquids (ILs). Some advantageous properties like negligible vapor pressure, non flammability,
a wide liquid temperature range, excellent solvent properties, high thermal conductivity, and

Solvents for Sustainable Chemical Processes 6


Work done and Literature Survey

immiscibility with many organic solvents were reported by Zhao et al, 2005. By changing
structures such as cationic portion of the salt, the anionic portion, and the subsequent pairing
of specific cations with specific anions, properties of ionic liquids can be varied.

Fig 6: Structure of BMIM-PF6 & BMIM-BF4.


Two-Component RevIL’s:
In contrast to the classical ionic liquids, the ionic liquid formed from the reaction of low
boiling dimethylamine with gaseous carbon dioxide (N, N-dimethylammonium N′,N′-
dimethylcarbamate (DIMCARB) can actually be “distilled.” U.P. Kreher et al., 2003 reported
that under the influence of heat, the liquid ionic salt is reverting back to dimethylamine and
carbon dioxide and then, upon collection, reforming the liquid ionic salt at ambient
temperatures.

Fig 7: Two-component RevILs.


The molecular liquids are composed of 1, 8-diazabicycloundec-7-ene (DBU) (top) or
N,N,N’,N’-tetramethyl-N’’-butylguanidine (TMBG) (bottom) and alcohol (ROH). Under
CO2 pressure, a reversible switch forms ionic liquids [DBUH] + [RCO3] − and [TMBGH] +
[RCO3] −, respectively. In contrast with classical ILs, however, RevILs can be switched back
and forth from a molecular liquid (non-ionic liquid like an organic amine) to an ionic liquid
upon the reversible reaction with CO2 (Fig. 7). Hence, reversible ionic liquids can, in

Solvents for Sustainable Chemical Processes 7


Work done and Literature Survey

principle, present a solution to the issues of product separation and solvent purification
encountered by classical ionic liquids.

Fig 8: Claisen–Schmidt condensation of 2-butanone and benzaldehyde.


R.Hart et al., 2010 investigated the Claisen–Schmidt reaction between butanone and
benzaldehyde in TMBG to form the terminal enone (major product), the internal enone
(minor product) (Fig.8). TMBG acted as solvent and base catalyst. After reaction completion,
methanol and octane were added to the reaction mixture. Subsequent addition of carbon
dioxide promoted the formation of the RevIL resulting in a liquid–liquid phase split. The
enone products selectively partitioned in the octane phase and were easily separated by
decantation. The ionic liquid phase was reversed to TMBG by heating and subsequently
reused in another reaction. Three cycles were successfully conducted with isolated yields 34,
32, and 34% of enones (95% selectivity toward the terminal enone). Partial conversions and
drying step were necessary to eliminate undesired polycondensation products and avoid
formation of guanidinium carbonate ionic species, respectively.

Fig 9: Claisen Palladium catalyzed Heck-type coupling reaction.


Hart et al. also reported Heck couplings in the RevIL from DBU–hexanol (Fig. 9). Heck
reactions inherently result in the formation of HX (X = halide; Fig. 9) which must be
neutralized by a base to recover the catalyst. Therefore, stoichiometric amounts of inorganic
halide salt are formed as a byproduct. RevILs offered the capability to couple reaction with a
two-stage separation: switching from the RevIL to the molecular liquid induced precipitation

Solvents for Sustainable Chemical Processes 8


Work done and Literature Survey

of unwanted inorganic salt by-products allowing for recycling of both solvent and the
palladium catalyst (which remained soluble in the molecular liquid).

Single-Component RevIL’s
In 2009, Blasucci et al. reported the first single-component reversible ionic liquid based on
trialkoxy- and trialkylsilylpropylamine structures. The application was specifically for crude
oil recovery.

Fig 10: Silylated amines system R = alkoxy or alkyls.


Upon reaction with CO2, the amine forms ionic liquids composed of the corresponding
carbamate anion and ammonium cation pairs (Fig. 10). It is thought that the presence of the
silicon is critical for achieving a liquid ionic material.

Fig 11: Silylated amines tested for post-combustion CO2 capture


The trialkylsilylpropylamines based RevILs have proven to be particularly attractive for
energy-conscious CO2 capture strategies as reported by A. L. Rohan et al., 2012. (Fig 11) For
these applications, the RevILs behave in a dual capacity: (1) the selective and efficient
reaction of CO2 with the molecular liquid to form the ionic liquid and (2) the enhanced
capacities offered by the physical absorption of the CO2 into the resulting ionic liquid. The
efficiency of a particular silylamine to effectively capturing CO2 has been shown to be
strongly dependent on its structure. For example, the influence of the length of the alkyl
tether (between the silyl and amine groups) or the position of one or more methyl groups on
the silylamine (tether or reactive amine groups) has significant effect on the CO2 capture
properties such as reaction enthalpy, viscosity, and reversal temperature. While selection of a
CO2 capture solvent is dependent on the specifications of its end application, the authors
identified two silylamine solvents, 1-(aminomethyl)triethylsilane and (trans)- 3-(triethylsilyl)
prop-2-en-1-amine, as candidates for pilot-scale investigations. Overall, these systems can (1)

Solvents for Sustainable Chemical Processes 9


Work done and Literature Survey

operate neat, which provide an energy incentive (limiting the energy penalty associated with
aqueous amine systems), (2) capture CO2 efficiently via chemical reaction and physical
adsorption means, (3) release CO2 and regenerate the silylamine solvent for recycling and (4)
enable a process-controlled viscosity since the viscosity is a function of the extent of
conversion of the starting amine to the ionic liquid.

3.2 Water at Elevated Temperatures


C. L. Liotta et al., 2007 reported that over the range of temperatures from ambient to 275 °C
(1) the density decreases from 1 g cc−1 to 0.75 g cc−1, (2) the dielectric constant decreases
from 78 to approximately 20 (Fig. 13), and (3) the auto-ionization constant, KW, increases
from 10−14 to 10−11 (Fig. 12).

Fig 12: Dissociation constant of water as a function of temperature from 0–350 °C.

Fig 13: Dielectric constant of water as a function of temperature from 0–500 °C.

Solvents for Sustainable Chemical Processes 10


Work done and Literature Survey

The physiochemical properties of water at elevated temperature are thus considerably


different from water at ambient temperature. Thus, as the temperature increases, the solubility
of organic molecules in water increases. The physiochemical properties of water now begin
to resemble organic solvents. For instance, at saturation pressure, a gradual change of
dielectric constant is observed from 78 at ambient temperature to approximately 50 at 125 °C,
40 at 175 °C and 20 at near critical conditions (300 °C). Each of these temperatures reflects
the dielectric constant of DMSO (47), acetonitrile (37.5), and acetone (21), respectively.
The dielectric constant of the medium resembles typical organic solvents when reaction is
conducted under homogeneous conditions at higher temperatures. By cooling the reaction
mixture, we can separate the organic products from the water solvent system easily. It should
be pointed out that the reduced dielectric constant of water at elevated temperatures, while
increasing the solubility of non-polar organics, reduces the solubility of inorganic salts.
J. S. Brown et. al., reported that for applications as a reaction medium, a tradeoff exists for
the desired solubility of ionic and nonionic species.
Water at elevated temperatures has also been investigated as a direct alternative to organic
solvents. Without increasing the temperature by a lot (i.e. moderate temperatures of around
100-200 °C) chemical transformations can be carried out when one or more components in
the reaction system are hydrophilic.
Significant success has been reported when water was used as the sole solvent at
temperatures between 100 and 200 °C for the coupling of phenylboronic acids and various
haloaromatics (Fig. 14).

Fig 14: Suzuki coupling reaction of substituted halobenzene and phenylboronic.


In some cases, the addition of a co-solvent such as an alcohol and/or the use of microwave
(MW) radiation for fast heating were also explored. Leadbeater and Marco, 2002 and 2003
reported coupling yields up to 98% in water at 150 °C with conventional and with MW
heating. The catalyst used was palladium acetate; no ligands were necessary. The procedure
did include sodium carbonate as the base as well as a phase transfer catalyst (tetra-n-
butylammonium bromide).

Solvents for Sustainable Chemical Processes 11


Work done and Literature Survey

An increase in the dissociation constant as shown above, has led to development of


successful strategies in acid-base transformations. There have been studies in algae
hydrothermal treatment in order to produce biomass. In this context, chemical reactions, both
catalyzed and uncatalyzed, in water near or above its critical point is conducive to convert
biomass into hydrogen, methane and/or liquid fuels. By changing the reaction temperature,
protecting groups such as esters, ethers, amides and carbamates can be selectively and
efficiently removed as reported by J. Wang et al., 2009.
Water at elevated temperature can provide a vehicle for conducting a homogeneous
reaction and allow for a facile product separation by returning to room temperature
accompanied by a phase split. Other advantages include replacing environmentally
undesirable catalysts, eliminating unwanted byproducts, improved selectivity, and
elimination of mass transfer limitations commonly present in heterogeneous systems.

3.3 Supercritical Fluids:


The first presence of supercritical fluids as an alternative to the current in-use solvents
was seen way back in the 1970’s. That means supercritical fluid research is has been around
for at least 40 years and has proved to have a wide range of applications. A fluid having its
pressure and temperature slightly higher than its critical pressure and temperature is termed
as a supercritical fluid. The liquid starts becoming less dense and the vapor starts becoming
more and denser as we go on increasing pressure and temperature. At the critical point they
converge to ultimately become identical. All the special properties of SCF’s investigated
hereafter are in the region of high compressibility. One important point to note is that an SCF
should have an accessible critical point for it to have industrial applicability. E.g. Water
(Pc=22.1 MPa and Tc=373 °C) also can qualify as an SCF but has very limited use. On the
other hand, scCO2 has been the most studied SCF (Pc=4.9 Mpa and T c=32 °C). Another
common SCF with a readily accessible critical point is ethane (Pc 4.9 MPa; Tc 32 °C). Unlike
CO2, however, its high flammability imposes a significant safety constraint. A limitation on
its use as an SCF is its extremely high flammability.

Properties:
The main property which holds SCF as a probable replacement for today’s solvents is
its easy manipulation of physicochemical properties for tunability i.e. achieving required
results. The tunability manifests itself in the form of either pressure variations or addition of

Solvents for Sustainable Chemical Processes 12


Work done and Literature Survey

any co-solvents. SCF have an advantage over traditional solvents because of this very reason
of their tunable properties e.g. diffusivity, transport ability etc.
In the plot presented below (Fig. 15), it can be seen that a SCF offers tremendous
improvement in solubility for what are small changes in pressure. It can also be seen that at
low pressures the behavior of a SCF is similar to that of an ideal gas but at higher pressures
there is a definite enhancement in solubility (even up to 5 orders of magnitude).
The solubility in a SCF can also be manipulated by adding small amounts of a co-solvent
such as acetone or ethanol. It can be seen that there is an increase in selectivity upon addition
of methanol as co-solvent (only 1% methanol) (Fig. 16). The hydrogen proton undergoes
hydrogen bonding with the nitrogen atom in acridine as hence there is solubility
enhancement.

Fig 15: Solubility enhancement with SCF over ideal gas.

Fig 16: Co-solvent effect on selectivity of anthracene and acridine.

Solvents for Sustainable Chemical Processes 13


Work done and Literature Survey

3.3.1 Supercritical Carbon-Dioxide:


Synthesis:
ScCO2 as a solvent is relatively poor. The major requirement to be fulfilled by a solvent is its
ability to reasonable solubility of reactants in the solvent system and to make them react with
each other. If the reactants are of the ionic type, then this poses a problem with regard to
SCF. In this case, phase transfer catalysis is employed for reactions conducted in scCO2 and
scEthane. For instance, benzyl chloride was reacted with solid potassium cyanide in the
presence of the phase transfer catalyst (PTC) and tetra-n-heptylammonium bromide (THAB),
to form phenylacetonitrile (Fig 17). If PTC was not used, only 5% product was formed in 48
hours. On the other hand, in the presence of 10 mol% catalyst, 100% yield of product was
obtained in 10 hours.

Fig 17: Substitution of benzyl chloride with potassium cyanide to form phenylacetonitrile in
scCO2 (top),
Formation of α-phenyl-α-ethylcyanoacetate during the alkylation of phenylacetonitrile with
scCO2 (bottom).
Unlike the classical phase transfer catalysis mechanism, the phase transfer catalyst has
limited solubility in the scCO2 and resides almost exclusively on the surface of the cyanide
salt as reported by K. Chandler et al., 1998. The experimental results suggest that the actual
substitution process takes place on the surface of the ionic solid (Fig 18) within a region rich
in both cyanide ion and catalyst. Indeed, the addition of 5 mol% acetone to the scCO2, while
increasing the solubility of the catalyst in the fluid phase, actually decreased the overall rate
of reaction—an observation consistent with the reaction taking place on the surface of the
solid.

Solvents for Sustainable Chemical Processes 14


Work done and Literature Survey

Fig 18: Mechanism of PTC reaction in scCO2.


Also, Suzuki-Mayaura coupling reaction study with SCF has also been carries out. This
method employs use of packed bed reactors for the reaction as reported by G. A. Leeke et al.,
2006. It is specifically useful when the reactants are soluble in scCO2 but the catalyst is not.
Specifically the coupling of p-tolylboronic acid and iodobenzene to form 4-methylbiphenyl in
continuous mode was conducted in scCO2. Pd(II)EnCat and tetra-n-butylammonium
methoxide (Bu4NOMe) were used as solid supported catalyst and base, respectively. (Fig.
19)

Fig 19: Reaction scheme for Suzuki–Miyaura reaction of p-methyl boronic and iodobenzene
in Packed Bed Reactor (PBR).
A conversion of 81% at 100 °C and 16.6 MPa was achieved when 10% methanol was added
as a co-solvent. When the co-solvent ratio was decreased to 1.5%, the conversion fell to 21%.
This clearly highlights a key role of the co-solvent in improving the solubility of reaction
partners in supercritical fluids of low polarity. One more important aspect which a green
solvent should have is recyclability. This is very easy in the case of scCO2 as we only need to
depressurize it and it returns into the gas state. This not only helps for solvent recovery but
also for product separation from solvent.

Solvents for Sustainable Chemical Processes 15


Work done and Literature Survey

Similar synergetic benefits have been reported for the metathesis of 1-octene to 7-
tetradecene catalyzed by Re2O7 supported on γ-Al2O3. In this case, M. Selva et al., 2012
reported the continuous metathesis reaction of 1-octene in scCO2 to produce 7-tetradecene
with a conversion of 75% and a selectivity of 95%. (Fig. 20)

Fig 20: Metathesis reaction of 1-octene to 7-tetradecene in scCO2.


Although comparable initial conversion (65%) and selectivity (90%) were obtained when n-
hexane was used as solvent, the selectivity decreased to 0% (in the outlet stream) after 15
min. In contrast, when scCO2 was employed, the selectivity remained unchanged for 40 min,
after which it gradually decreased to 40% at 90 min. The decrease of reactivity over time was
attributed to the coating of the catalyst bed with insoluble side-products. Washing the
catalytic bed with acetone and scCO2 allowed for the recovery of the initial selectivity of
95%.
R. A. Bourne et al., 2009 and J.F.B. Hall et al., 2013 reported the replacement of traditional
processes for oxidation reactions by continuous flow processes in scCO2 as oxygen is
completely soluble in non-flammable CO2. For example, the platinum catalyzed oxidations of
1- and 2-octanol to 1-octanal and 2-octanone, respectively, were performed in scCO2 with
oxygen as the oxidizing reagent. (Fig 21)

Fig 21: Oxidation of 2-octanol and 1-octanol in scCO2.


The reaction was carried out in flow mode (packed bed reactor) at 10 MPa. The 1-octanal and
2-octanone were each obtained in 75% yield at 150 °C and 135 °C, respectively. Importantly,

Solvents for Sustainable Chemical Processes 16


Work done and Literature Survey

the 1-octanal was selectively produced—no octanoic acid (over-oxidation product) was
observed and no mass losses were detected.
Applications:
One field of the future which is expected to grow exponentially is nanotechnology
and nanotechnology-enabled products. Thus, widespread use of green solvents will have a
huge impact on the nanotechnology industry. The unique properties of scCO2, such as the
high diffusivity and modest solvation of solutes, are beneficial for the processing of advanced
materials.
It is seen that pharmaceutical industries are interested in the formation of micron and
sub-micron sized particles and scCO2 helps by introducing the rapid expansion of
supercritical solutions (RESS) and supercritical antisolvent (SAS) precipitation methods.
Casciato et al., 2012 reported the scCO2 mediated synthesis and deposition of silver
nanoparticles (AgNPs) on silicon and glass surfaces for application in surface-enhanced
Raman spectroscopy (SERS). The aim of the study was to tune particle size, vary density and
surface coverage onto the support. For this, they explored the mechanisms and parameters
affecting heterogeneous deposition of AgNPs.

Fig 22: Structure of the precursor 1,1,1,5,5,5-hexafluoroacetylacetonate cyclooctadiene and


particle size distribution of silver nanoparticles deposited from scCO2 onto a HCl treated
silicon surface.
For example, the histogram plotting fraction versus AgNPs diameter (deposited on HCl-
treated silicon support) as a function of temperature (60–180 °C) is shown in Fig. 22. The
mean particle size of the nanoparticles decreases from 196, to 104, and to 27 nm as the
temperature increases from 60, 120, 180 °C, respectively. Overall, their results are consistent
with the deposition process being reaction rate limited (reduction of the silver precursor with

Solvents for Sustainable Chemical Processes 17


Work done and Literature Survey

hydrogen) rather than being limited by the transport of the precursor to the surface. This is in
part due to the high solubility of the silver precursor in scCO2 and contrasts with other,
traditional systems like chemical vapor deposition that are often transport-limited. The two
variables (temperature and surface) can be efficiently adjusted to tune the AgNPs mean size,
density and surface coverage.

3.4 Gas-Expanded Liquids and Organic Aqueous Tunable Solvents

3.4.1 Gas-Expanded Liquids:


A. M. Scurto et al., 2009 studied gas-expanded liquids and came up with an informal
definition. A mixture of an organic solvent with a gas (most commonly CO2) at moderate
pressures of around 3-8 MPa is termed as a gas-expanded liquid. Alcohols, ketones, ethers
and many other solvents are highly miscible with CO2 and that too at much lower pressures
than those required by SCF. When CO2 dissolution into organic liquids (e.g. with benzene,
cyclohexane, decane, 2-propanol, acetonitrile and methanol) takes place, it leads to volume
expansion as shown in Fig. 23

Fig 23: Volume expansion of organic liquids with CO2


Solvent polarity is an important parameter to be considered here. If we are able to tune the
polarity of the solvent it enables us to vary the solubility of different solutes to a different
extent. The broad impact of GXL (Gas-expanded liquids) is basically due to this ability of
being able to tune solvent polarity.
In contrast to scCO2, typical organics such as acetone and THF are relatively good
solvents with dielectric constants substantially greater than CO2. With the addition of CO2 to

Solvents for Sustainable Chemical Processes 18


Work done and Literature Survey

these solvents, the dielectric constant of the resulting media can be tuned to any value
between that of the pure organic solvent and that of pure CO2. Thus a wide range of solvent
properties are easily accessible with GXLs.
E.g. the solubility of phenanthrene in acetone at 25 °C is plotted as a function of CO2
pressure. (Fig 24) The point designated in red is pure CO2—a liquid just below its critical
point. The mol fraction of phenanthrene in acetone is reduced from 0.17 in pure acetone to
0.0044 at 50.6 bar (5.06 MPa) which was reported by D.J. Dixon et. al., 1991. Thus, by
simply adjusting CO2 pressure, the polarity of a solvent system can be tuned for a wide
variety of processes.

Fig 24: Solubility of Phenanthrene in gas expanded acetone as a function of CO2 pressure.
The addition of CO2 also leads to the formation of alkyl-carbonic acids as reported by J. L.
Gohres et al., 2009 (Fig. 25) Due to this the acid-base properties of the solvent are modified.
This can be exploited when working on in-situ reversible acid catalysts. Some examples of
the above can be hydration of pinenes, synthesis of diazonium salts and formation of ketals.

Fig 25: In situ formation of alkyl carbonic acids in CO2 expanded alcohols.
In situ acid catalysis using methylcarbonic acid for the formation of ketals in gas expanded
methanol was studied as shown in the reaction. (Fig 26)

Fig 26: Ketal formation in gas-expanded alcohol.

Solvents for Sustainable Chemical Processes 19


Work done and Literature Survey

Rate enhancement studies were reported by X. Xie et al., 2004. (Fig 27) In the case of the
acid-catalyzed formation of the dimethylketal of cyclohexanone, a rate enhancement of 135%
was observed in CO2 expanded methanol at 3.5 MPa and 50 °C compare to methanol alone.
At higher pressures the rate enhancement decreased. This could be attributed to the lower
polarity of the solvent due to the high presence of CO2 in methanol.
When the reaction was run at lower temperatures, the maximum enhancement factor occurred
at lower pressures. Thus, at 40 °C, the reaction rate increased 100% at 2 MPa; while at 25 °C
a maximum enhancement of 85% was observed at 1.5 MPa. Thus, at lower temperatures,
there wasn’t significant enhancement in the reaction rate.

Fig 27: Enhancement of rate constant for cyclohexanone Ketal formation in gas expanded
methanol. Symbols key: (○) 25 °C, (●) 40 °C and (△) 50 °C.

3.4.2 Organic aqueous tunable solvents (OATS):


For most reactions, we require a homogenous reaction system. This is difficult to achieve
while dealing with hydrophilic catalyst and hydrophobic substrates in order to form
hydrophobic products. If a process can be developed which can induce a phase split
separating organic and aqueous components, it would be advantageous. The hydrophilic
catalyst would then be located in the aqueous phase and the hydrophobic product in the
organic phase. As the product and the catalyst are already separated, it enables easy solvent
recycling.

Solvents for Sustainable Chemical Processes 20


Work done and Literature Survey

Fig 28: Schematic of an OATS-mediated process.


The general processing schematic in Fig 28 highlights the three key features of OATS
processes:
(1) The reaction is performed homogeneously (reaction efficiency)
(2) The product separation/isolation is conducted heterogeneously (optimizing product
recovery)
(3) The organic solvent, catalyst-bearing aqueous phase and CO2 can all be recycled.
This schematic also demonstrates the relationship between OATS and the principle of a
green/sustainability process. Only one feed stream is entering and only one product stream is
exiting. All other streams are recycled, maximizing efficiency and minimizing waste stream.

Fig 29: Hydroformylation of 1-octene to form 1-nonanal (linear product) and 2-methyloctanal
(branched product).
Industrially, the hydroformylation of propylene to form butyraldehyde is performed via a
water/organic biphasic process. As the reaction is still taking place in the aqueous phase;
some solubility of the alkene in the aqueous phase is therefore necessary. While the solubility
of propylene in water (∼200 ppm) is indeed sufficient from an industrial standpoint, the
negligible water solubility of longer-chain olefins prohibits such biphasic processes. This is
exactly where the OATS technology could provide a competitive and sustainable alternative.
Hydroformylation reactions of 1-octene (Fig. 25) and p-methylstyrene (an analog for an
ibuprofen precursor) were successfully conducted using OATS technology.

Solvents for Sustainable Chemical Processes 21


Work done and Literature Survey

The solubility of the 1-octene in water is less than 3 ppm. However, it is increased more than
10,000-fold by the addition of tetrahydrofuran (THF) as cosolvent. A 70/30 (vol/vol) THF–
water mixture produces a viable homogeneous system for conducting the hydroformylation
of 1-octene. Specifically, the reaction of 1-octene with syn-gas in the presence of a rhodium
catalyst produced the two isomeric aldehyde products, 1-nonanal (linear product) and 2-
methyloctanal (branched product) as reported by J. P. Hallett et al., 2008. Three ligands with
very different hydrophilic character were employed: triphenylphosphine (TPP),
triphenylphosphine monosulfonate sodium salt (TPPMS) and triphenylphosphine trisulfonate
sodium salt (TPPTS). The partition behavior of the ligands, 1-octene and the aldehyde
products as a function of CO2 pressure was reported. Overall, TPPMS ligand emerged as the
optimum choice. The reaction performed efficiently. The aldehyde products were obtained in
85% yield at 80 °C with only 4% isomerization of the 1-octene to internal alkenes. And in
contrast with the TPP based catalyst, the hydrophilic TPPMS catalyst was easily separated,
recovered and recycled three times upon phase splitting with the addition of CO2. The TOFs
were consistent from cycle to cycle, ranging from 51 to 47 to 54 h−1, indicative of the
retention of catalytic activity. Importantly, no rhodium leaching was detected in the organic
phase. (Using atomic absorption spectroscopy with a detection limit of <1 ppm)

Applications:
Nanoparticle Processing
Methods like reverse micelle and precipitation are used traditionally for particle preparation
when it comes to nanoparticles. The principle for nanoparticle synthesis is that they are
synthesized and stabilized in solution (dispersed) by capping agents usually molecules like
dodecanethiol. Post-synthesis, the separation and/or deposition of the nanoparticles from
solution is achieved by adding an anti-solvent. The role of the anti-solvent is to alter the
solubility of the capped nanoparticles, effectively lowering the dispersibility of the ligand-
coated nanoparticles in solution and causing their precipitation. This process must be well
controlled to obtain ordered and uniform deposition onto the chosen surface. In practice, the
deposition (or dewetting) step can be challenging with liquid solvents, owing essentially to
capillary forces and high surface tension at the liquid–vapor interface. In contrast, CO2
pressure is an effective, recyclable, and inexpensive means to tune the composition and
physicochemical properties (density, viscosity, diffusivity, solvent strength, and surface
tension) of organic solvents. Precisely, GXLs will offer marked benefits compare to

Solvents for Sustainable Chemical Processes 22


Work done and Literature Survey

traditional methods to: (1) minimize wetting and surface tensions effects, (2) allow for a
precise control over the deposition step simply by controlling pressure, (3) improve
commercial viability—in contrast with SCF mediated processes, no fluorinated (CO2- philic)
ligands are necessary, moderate pressure and higher particles concentration are possible, (4)
minimize wastestreams.
The effectiveness of GXLs-mediated process to improve the deposition of dodecanethiol-
stabilized silver nanoparticle (AgNP) in hexane on carbon grid was effectively demonstrated
by M. C. McLeod et al., 2005. The dodecane thiol-stabilized AgNP were prepared by
literature reported methods (arrested precipitation) and then re-dispersed into chloroform or
hexane for subsequent deposition. TEM images of the deposited AgNP obtained from
evaporation of hexane in air show local disorder, agglomeration of the nanoparticles and
percolating networks. In contrast, the AgNP particles deposited via the CO2-expanded hexane
method are closely packed with no major defects. Expanding the organic solvent with CO2
enables the controlled deposition of nanoparticles leading to evenly dispersed nanoparticles
over a wide surface area, free of unwanted features such as local high concentration and
random interconnected particle network. Since the deposition of nanoparticles from a CO2
expanded liquids is size-dependent; experimental protocols have been developed and reported
by K. M. Hurst et al., 2009 to selectively fractionate a polydispersed suspension into several
fractions with narrower size distributions at scales ranging from 1 mL to 20 mL of
nanoparticle suspensions.

Solvents for Sustainable Chemical Processes 23


Industrial Relevance

4. Industrial Relevance – Case Studies


Following are 2 examples of the industrial applications of green solvents. These examples
highlight the relevance of this study and the immediate applications to the industry.

BASF’s nucleophilic HCl:


The chlorination of alcohols requires reactants that do not produce water as a by-product,
such as COCl2, SOCl2 or PCl3, etc. This is because the water produced as a by-product of the
reaction forces the equilibrium back towards the starting alcohol. When the starting material
is a diol, a number of possible partially chlorinated and ether by-products are formed.
However, these are toxic, difficult to handle and environmentally damaging. BASF has
recently commercialized an ionic liquid process for nucleophilic substitutions for the
conversion of alcohols to halogenoalkanes that allows HCl to be used as the chlorinating
agent as reported by Stegmann V et. al., 2005. When used in the chlorination of 1,4-
butanediol this yields the dichloride without the formation of by-products.
In the nucleophilic HCl process, HCl is dissolved in a chloride ionic liquid, forming an
[HCl2]− salt. This salt is the chlorinating agent. However, this does not explain why the water
produced in the reaction no longer causes a problem. Spectroscopic investigations of water in
ionic liquids show that it can interact very strongly with the ionic liquid’s ions, particularly
when the anion of the ionic liquid is a strong hydrogen bond acceptor, as is Cl−. These
interactions lead to ionic liquids being able to stabilize water-sensitive solutes or prevent
water from reacting with a solute. This behaviour is only possible when the ionic liquid is dry
and the water level must be below 25 mol% for the reaction to be successful. The
introduction of this process has led to the elimination of the highly toxic gas COCl 2, with the
attendant savings that derive from not needing to put in place the necessary engineering
controls to handle it safely.

Decaffeination:
In the latter half of the twentieth century, health concerns over the effects of caffeine led to
increased demand for decaffeinated coffee. Early forms of decaffeinated coffee were
produced by caffeine extraction with dichloromethane. The direct decaffeination of green
coffee beans occurs before their roasting, which removed the DCM from the beans to levels
of a few ppm. It was not the environmental concern that led to the replacement of this
process. Alongside caffeine the DCM also removed important flavour components of the

Solvents for Sustainable Chemical Processes 24


Industrial Relevance

coffee, giving a poor quality product. This led to a number of other less environmentally
concerning solvents being used for coffee decaffeination, but with the commercial driver
being the search for a better product.
Ethyl acetate is an environmentally preferred solvent used for coffee bean decaffeination by
Morrison LR et. al., 1984. Water has also been used to commercially decaffeinate coffee in
the Swiss Water process.
Supercritical CO2 (sc-CO2) decaffeination is also often described as solvent-free and has been
reported and patented by Zosel K, 1974. The green coffee beans are wetted and then the sc-
CO2 is used to extract the caffeine. The sc-CO2 process is much more selective for the
removal of caffeine than any of the other processes, leading to a high-quality product without
the need for the additional steps to isolate it that are required for other methods. The start-up
costs for an sc-CO2 decaffeination plant are higher than those of the other methods, but the
economic viability of the sc-CO2 process is enhanced because the caffeine is a saleable co-
product, particularly as it can be labelled as ‘natural’, for use in products such as cosmetics
and so-called ‘energy’ drinks for which this label can carry a premium. Sc-CO2 processing
has become a widely used method in the food industry, such as in the decaffeination of tea,
the removal of fat to produce low-fat varieties, the removal of alcohol to produce low alcohol
beers and wines and the removal of pesticides from rice and the extraction of flavours and
fragrance compounds.

5. Conclusion
The environmental concerns that surround the use of solvents for chemicals processing will
ensure that this remains an active area for research for some time to come. Each system
explained in the review offers a unique set of properties for varied applications. For each
application given, however, the common thrust is to enable greener and sustainable solutions
for chemical processes. The industrial examples also demonstrate that the implementation of
the concept of sustainability in the production and use of chemicals and chemical products
requires that chemicals processing must be both environmentally and commercially
sustainable. No one traditional solvent, or for that matter any of the solvent systems discussed
in this review, can be universally successful for all processes and separation. Innovative
solutions to expand the present portfolio of sustainable and green solvents are an on-going
process.

Solvents for Sustainable Chemical Processes 25


Nomenclature

6. Nomenclature

1. BR-UN : Brundtland Commission of the United Nations


2. DBU : 1, 8-diazabicycloundec-7-ene
3. DCM : dichloromethane
4. DIMCARB : N, N-dimethylammonium N′,N′- dimethylcarbamate
5. DMAP : 4-dimethylaminopyridine
6. DMSO : Dimethyl Sulfoxide
7. GXL : Gas-Expanded Liquid
8. H-NMR : Hydrogen Nuclear Magnetic Resonance
9. MW : Microwave
10. OATS : Organic Aqueous Tunable Solvent
11. PTC : Phase Transfer Catalyst
12. PS : Piperylene Sulfone
13. SCF : Supercritical Fluid
14. THAB : tetra-n-heptylammonium bromide
15. TMBG : N,N,N’,N’-tetramethyl-N’’-butylguanidine
16. TOF : Turnover Frequency
17. TPP : triphenylphosphine
18. TPPMS : triphenylphosphine monosulfonate sodium salt
19. TPPTS : triphenylphosphine trisulfonate sodium salt
20. WET : Water at Elevated Temperatures

Solvents for Sustainable Chemical Processes 26


References

7. References

1. A. L. Rohan, J. R. Switzer, K. M. Flack, R. J. Hart, S. Sivaswamy, E. J. Biddinger, M.


Talreja, M. Verma, S. Faltermeier, P. T. Nielsen, P. Pollet, G. F. Schuette, C. A.
Eckert and C. L. Liotta, ChemSusChem, 5, 2181–2187 (2012)
2. A. M. Scurto, K. Hutchenson and B. Subramaniam, Gas Expanded Liquids and near-
Critical Media: Green Chemistry and Engineering, vol. 1006, pp. 3–37. (2009)
3. Antonia Pérez de los Ríos, Angel Irabien, Frank Hollmann, and Francisco José
Hernández Fernández, Hindawi Publishing Corporation, Journal of Chemistry,1-2
(2013)
4. Brundtland CG. Our common future. Oxford, UK: World Commission on
Environment and Development, Oxford University Press. (1987)
5. D. J. Dixon and K. P. Johnston, AIChE Journal, 37, 1441– 1449. (1991)
6. D. Vinci, M. Donaldson, J. P. Hallett, E. A. John, P. Pollet, C. A. Thomas, J. D.
Grilly, P. G. Jessop, C. L. Liotta and C. A. Eckert, Chem. Commun., 0, 1427–1429.
(2007)
7. G. A. Leeke, B. Al-Duri, J. P. K. Seville, C. J. Smith, C. K. Y. Lee, A. B. Holmes and
I. F. McConvey, Org. Process Res. Dev., 11, 144–148 (2006)
8. H. Zhao, S. Q. Xia and P. S. Ma, J. Chem. Technol. Biotechnol., 80, 1089–1096.
(2005)
9. J. F. B. Hall, R. A. Bourne, X. Han, J. H. Earley, M. Poliakoff and M. W. George,
Green Chem., 15, 177–180. (2013)
10. J. L. Gohres, A. T. Marin, J. Lu, C. L. Liotta and C. A. Eckert, Ind. Eng. Chem. Res.,
48, 1302–1306. (2009)
11. J. P. Hallett, J. W. Ford, R. S. Jones, P. Pollet, C. A. Thomas, C. L. Liotta and C. A.
Eckert, Ind. Eng. Chem. Res., 47, 2585–2589 (2008)
12. J. S. Brown, J. P. Hallett, D. Bush and C. A. Eckert, J. Chem. Eng. Data, 45, 846–850
(2000)
13. J. Wang, Y. L. Liang and J. Qu, Chem. Commun., 5144–5146. (2009)
14. Karem Shanab, Cathanna Neudorfer, Eva Schirmer and Helmut Spreitzera, Current
Organic Chemistry, 17, 1179-1187 (2013)
15. K. Chandler, C. W. Culp, D. R. Lamb, C. L. Liotta and C. A. Eckert, Ind. Eng. Chem.
Res., 37, 3252–3259. (1998)

Solvents for Sustainable Chemical Processes 27


References

16. K. M. Hurst, C. B. Roberts and W. R. Ashurst, Nanotechnology, 20. (2009)


17. L. R. Drake, S. C. Stowe and A. M. Partansky, J. Am. Chem. Soc., 68, 2521–2524.
(1946)
18. Marina Cvjetko Bubalo, Senka Vidovic, Ivana Radojci, Redovnikovi and Stela Jokic,
J Chem Technol Biotechnol; 90: 1631–1639 (2015)
19. M. C. McLeod, C. L. Kitchens and C. B. Roberts, Langmuir, 21, 2414–2418. (2005)
20. M. Casciato, G. Levitin, D. Hess and M. Grover, J. Nanopart. Res., 14, 1–15. (2012)
21. Morrison LR, Phillips JH. Accelerated decaffeination process. Patent no. EP
0114426B1. (1984)
22. M. Selva, S. Guidi, A. Perosa, M. Signoretto, P. Licence and T. Maschmeyer, Green
Chem., 14, 2727–2737. (2012)
23. N. E. Leadbeater and M. Marco, J. Org. Chem., 68, 888–892. (2003)
24. N. E. Leadbeater and M. Marco, Org. Lett., 4, 2973– 2976. (2002)
25. N. Jiang, D. Vinci, C. L. Liotta, C. A. Eckert and A. J. Ragauskas, Ind. Eng. Chem.
Res., 47, 627–631. (2007)
26. P. J. Dunn, A. S. Wells and M. T. Williams, Green Chemistry in the Pharmaceutical
Industry, Wiley-VCH Verlag GmbH & Co. KGaA (2010)
27. Richard K. Henderson, Concepcion Jimenez-Gonzalez, David J. C. Constable, Sarah
R. Alston, Graham G. A. Inglis, Gail Fisher, James Sherwood, Steve P. Binksa and
Alan D. Curzons, Green Chem., 13, 854 (2011)
28. R. A. Bourne, X. Han, M. Poliakoff and M. W. George, Angew. Chem., Int. Ed., 48,
5322–5325. (2009)
29. R. A. Sheldon, C. R. Acad. Sci., Ser. IIc: Chim., 3, 541– 551. (2000)
30. R. Hart, P. Pollet, D. J. Hahne, E. John, V. Llopis-Mestre, V. Blasucci, H.
Huttenhower, W. Leitner, C. A. Eckert and C. L. Liotta, Tetrahedron, 66, 1082–1090.
(2010)
31. Stegmann V, Massonne K., Method for producing haloalkanes from alcohols. World
Patent No. WO 2005026089, Canadian Patent no. CAN 142:338152. (2005)
32. X. Xie, C. L. Liotta and C. A. Eckert, Ind. Eng. Chem. Res., 43, 2605–2609. (2004)
33. Zosel K., Process for recovering caffeine. US Patent no. 3,806,619. (1974)

Solvents for Sustainable Chemical Processes 28

Вам также может понравиться