2.1. Alcohol Oxidase and Peroxidase-Catalyzed Transformations
A recent study has shown the use of different enzymatic preparations to selectively oxidize 5-hydroxymethylfurfural (HMF,
Scheme 1), a natural product, to 2,5-diformylfuran (DFF), 5-hydroxymethyl-2-furancarboxylic acid (HMFCA), 5-formyl-2-furancarboxylic acid (FFCA), and 2,5-furandicarboxylic acid (FDCA) [
21]. These oxidized derivatives are precursors of important compounds. For instance, DFF and FDCA have antifungal and anti-
Pneumocystis carinii activities. Also, HMFCA is an important monomer for synthesizing polyesters. Alcohol oxidase (AO) from
Candida boidinii oxidized HMF and showed a high activity compared to the other alcohol oxidases. AO was able to oxidize HMF to DFF in 41% yield. In addition, galactose oxidase (GO) from
Dactylium dendroides was also able to perform this transformation in excellent yields (~90%) in the presence of catalase and horseradish peroxidase. Xanthine oxidase (XO) from
Escherichia coli was able to selectively oxidize the formyl group located in the HMF substrate and gave HMFCA in 94% yield after only seven hours (
Scheme 1). In addition, this XO-mediated process had many advantages, as it used air instead of the toxic H
2O
2 to achieve the oxidation, it did not use harmful organic solvents, it was highly selective, and therefore, no by-products were observed [
21].
This research team also used three different laccases applied to the oxidation of HMF to FFCA in the presence of the mediator 2,2,6,6-tetramethyl-1-piperidinyloxy radical (TEMPO). At the beginning of the reaction, DFF appeared as an intermediate and was then transformed to FFCA that was accumulated along the time. The laccase from Panus conchatus gave the highest yield of FFCA (82%) after 96 h. The laccase from Trametes versicolor gave a 68% yield of FFCA after 48 h, and the laccase from Flammulina velutipes afforded a 70% yield of FFCA after 72 h. For the synthesis of FDCA, a sequential oxidation protocol was achieved, combining GO and lipase from Candida antarctica type B (CAL-B). In a first step, GO oxidized HMF to DFF (75% conversion) as previously described after 48 h. After extraction with ethyl acetate (EtOAc), t-butanol, CAL-B and H2O2 were added to the reaction medium to convert DFF to FDCA. An excellent yield of FDCA in this second step was obtained after 24 h (88%).
In the same study, the successful separation of HMF from DFF with the help of DES was also described after the enzymatic transformations, because DES have affinity towards hydrogen bond-containing molecules, and they enable to dissolve those molecules. The authors were able to separate DFF from HMF using three types of DES. The composition of these solvents was: choline chloride and glycerol (ChCl:Gly 1:2 mol/mol), choline chloride and urea (ChCl:Ur 1:2 mol/mol), and choline chloride and xylitol (ChCl:Xyl 1:1 mol/mol). They were used to extract HMF from an ethyl acetate solution that contained both DFF and HMF. The researchers discovered that DES composed of ChCl:Gly and ChCl:Ur were the ones that selectively extracted HMF from the mixture. When this mixture was extracted three-times using ChCl:Gly DES, the purity of DFF increased to 97% from the original purity of 76%.
In another investigation conducted by Yang et al., they tested the effect of 24 DES and 21 NADES as co-solvents on the biotransformation of isoeugenol into vanillin using
Lysinibacillus fusiformis CGMCC1347 whole cells. It was discovered that those eutectic solvents (1% v/v) enhanced the bioconversion because it made the bacterial cell membrane permeable to the lipophilic substrate (
Figure 1). Among them, the DES composed of choline acetate (ChAc) were usually better than the ones composed of ChCl in terms of product yields (up to 142%) regarding the control reaction without DES [
22]. The whole cells were immobilized on poly(vinyl alcohol)-alginate beads, and this biotransformation could be repeated up to 13 times in the presence of choline chloride:galactose (ChCl:Gal 5:2 mol/mol, 20% v/v), maintaining their activity to a similar extent, offering this system a promising design for further developments.
In a study conducted by Wu et al. [
23], it was found that DES composed of ChCl and different HBDs (Ur, Gly, acetamide, and EG) at different molar ratios promoted more efficiently the activity of horseradish peroxidase (HRP) compared to DES which was composed of choline acetate. In that study, twenty-four DES were synthesized to see their effect on HRP activity, and they were found to have a stabilizing effect on HRP, especially those when the molar ratio of the salt was higher than that of the HBD. The authors also identified through spectroscopic studies that DES were able to enhance the α-helix conformations thus providing a more relaxed tertiary structure of the enzyme, hence improving its activity. They concluded that DES can be versatile solvents and found that the hydrogen bonding network that is provided by DES prevents this solvent from dissociation in aqueous solutions at least at concentrations of 0.5 M.
Luna-Bárcenas and co-workers in 2016 have used two types of DES to conduct an enzyme-mediated free radical polymerization of acrylamide. One DES was made of ChCl:Ur and the other was made of ChCl:Gly (1:2 molar ratio). When the DES was used at higher volumes than the phosphate buffer solution, the catalytic activity of the HRP diminished drastically. However, the thermal stability of the enzyme was improved. The enzyme was still able to achieve the initiation of the acrylamide free radical polymerization even at 80% v/v concentrations of DES. Also, the research group was able to synthesize polyacrylamide in ChCl:Gly at 4 °C because it has a low freezing point. The combination of HRP, H
2O
2 and pentane-2,4-dione as initiator was responsible to promote the reaction leading to polyacrylamide synthesis in aqueous medium. The free radical polymerization occurred also at room temperature and at 50 °C. HRP was partially denatured in ChCl:Ur because the heme group was extracted into the solution and this caused a decrease in the enzymatic activity at higher DES concentrations. The concentrations of ChCl:Ur and ChCl:Gly were fixed to 80% v/v when used in the reaction media for acrylamide polymerization to protect the hydrogen bonding network in the hydrated medium while decreasing the viscosity of these solvents. Also, a homogenous mixture was obtained throughout the process [
24].
Another study conducted by Papadopolou et al. [
25] investigated the effect of DES made of ChCl or ethylammonium chloride (EAC) and different HBDs such as Ur or Gly at different molar ratios, on the catalytic activities of cytochrome c (cyt c) and HRP. This research team showed that the activity of the peroxidase depended on the nature of both the ammonium salt and the HBD. In addition to that, the study confirmed that EAC-derived DES stabilized cyt c and enhanced its capabilities for the degradation of an industrial dye compared to the eutectic mixture containing ChCl, suggesting that these eco-friendly solvents could be an interesting medium for biocatalytic reactions with applications in industry. When they compared the peroxidase activity in pure buffer and in DES-based media, it was concluded that the neoteric media increased the activity of the enzyme. In the case of choline chloride-based DES, when ChCl:Ur was used in the reaction medium at a concentration of 30% v/v, the activity of cyt c was enhanced up to 8-fold times compared to the one in buffer, while using EAC:Ur at 50% v/v it was possible to increase its activity up to 200-fold times. An explanation for this effect could be the higher viscosity of ChCl-based DES, increasing the mass transfer limitations of the substrates. Also, the HBD affected the activity of cyt c. When Ur was used, its activity was largely increased compared to the ones observed for DES composed of Gly and EG. The thermostability of cyt c could also be enhanced using DES as reaction media. Hence, when the enzyme was incubated with buffer or in the presence of 30% v/v of different DES, while cyt c lost 35% of its activity after incubation in buffer at 40 °C after 24 h, it remained perfectly stable when it was incubated in ChCl:Ur or EAC:Ur media. As an application, the decolorization of a dye, pinacyanol chloride, in different DES was tested using cyt c. The authors could immobilize the enzyme on Celite and perform the decolorization of this compound up to four cycles observing excellent activities using EAC:Ur and EAC:EG at 50% v/v [
25].
2.2. Deep Eutectic Solvents and Alcohol Dehydrogenases (ADHs)
ADHs are probably the most used enzymes corresponding to the family of oxidoreductases for synthetic purposes. They catalyze the reversible transformation between alcohols and carbonyl compounds through redox processes. These reactions are mediated by a nicotinamide cofactor, namely NAD or NADP, which is responsible of the electron transfer from or to the substrate [
26]. Apart from using free, isolated enzymes, the setup of whole cell-mediated transformations enables simple and cheaper procedures. Especially interesting is their use applied to the cofactor-dependent reactions, due to the fact that whole cell microorganisms can provide it on its own and avoid the external addition of these expensive molecules. As eco-friendly and cheap solvents, DES have been selected as good solvent candidates for ADH-mediated protocols. As the subsequent studies will show, DES present many stabilizing effects on biocatalytic reactions involving ADHs from different organisms.
ADHs are extremely important enzymes due to many reasons. For instance, in the human body they are the main enzymes responsible for digesting and degrading the alcohol that is consumed. These enzymes are usually present in the human liver and they detoxify ethanol converting it into acetate that is later used by cells. Moreover, ADHs from yeast are used to convert glucose to ethanol to make alcoholic beverages [
27]. Also, from a synthetic point of view, ADHs from microorganisms are used to convert ketones into chiral alcohols that can be precursors and building blocks of some important drugs such as cholesterol lowering drugs, anti-arthritic, and antibacterial agents. It is worthy to mention that these enzymes reduce ketones selectively according to the Prelog’s rule, which can be used to predict the stereopreference of alcohol dehydrogenase-catalyzed carbonyl reductions. Hence, ADHs usually provide the hydride through the
Re face of this prochiral moiety. In that case, the ADH follows the Prelog’s rule; on the contrary, the biocatalyst will show anti-Prelog selectivity [
28].
As a first example, Domínguez de María and co-workers demonstrated that DES could be an appropriate media for whole cell catalysis, achieving the bioreduction of ethyl acetoacetate (
1a) with Baker’s yeast [
29]. This study showed that DES can be efficient inhibiting oxidoreductases. For instance, alcohol dehydrogenases with (
S)-stereopreference were inhibited by ChCl:Gly (1:2 mol/mol). When adding different proportions of this DES into water, it was observed the inversion of the stereoselectivity in the bioreduction of
1a using this biocatalyst. Baker’s yeast whole cells showed an excellent stereoselectivity towards the (
S)-enantiomer of
1b in pure water (95%
ee). However, when the reaction was left in a DES medium containing less than 20% volume of water, the product was obtained with high (
R)-enantioselectivity (95%
ee,
Scheme 2). The reason for that was because Baker’s yeast genome contains a mixture of approximately 50 ADHs that can present different selectivities. Therefore, this effect could be explained due to the fact that some (
S)-selective enzymes were inhibited by DES while (
R)-selective ones were more active in this medium. In conclusion, DES were proven to be great solvents for whole cell biocatalysis, being able to influence both activity and selectivity.
In 2015 Bubalo et al. [
30], tested different DES on yeast mediated reduction of
1a and found that the DES that were composed of a sugar derivative had the best biocompatible results. Also, it was found that the amount of water present in the DES influenced the reaction results leading to an increase in the reaction yield when enlarging the percentage of DES. That group tested the effect of different types of DES using the yeast
Saccharomyces cerevisiae (baker’s yeast). They found that the reaction yield was influenced by the type of HBD (e.g., glucose, fructose, or glycerol) and by the amount of water added to the prepared DES solution. It was concluded that the aqueous solutions of DES (50% v/v of water) gave the highest reaction yield (93%). The viability of yeast cells in DES containing ChCl and Ur was also investigated, and it was found that it decreased in this type of solvent. The increased osmotic pressure on yeast cells grown in the urea type of DES led to a decrease in the viability of the cells. As a result, water diffused out of those cells. However, DES containing a sugar or glycerol demonstrated a yeast viability of 76–99% after 24 h of inoculation of the yeast cells. DES containing a sugar or glycerol provided a pH around 4.5, while the DES containing urea afforded a pH superior to 8. Since the preferred pH range for
S. cerevisiae growth is from 4 until 6, and as mentioned before the pH for DES containing sugars was 4.5, the best bioreduction yields were observed in these media. These results confirm that pH values and cell viability in bioreduction reactions involving DES can be an important effect, strongly influencing the enzymatic results. Glucose and other sugars such as fructose can also play an important role in the cofactor regeneration necessary for these processes [
30].
In a subsequent study done by Xu et al. [
31], the authors used different DES to study the enantioselective oxidation of racemic 1-(4-methoxyphenyl)ethanol (MOPE,
2a) employing whole cells of
Acetobacter sp. CCTCC M209061 and acetone to recycle the oxidized nicotinamide cofactor. They found out that the DES composed of ChCl:Gly (1:2 mol/mol) at 10% (v/v) afforded the best results because it improved the permeability of the cells membrane. It also improved the stability of the enzyme(s) involved in the process. Therefore, it was possible to obtain the remaining (
S)-
2a (
Figure 2) at 50% conversion in enantiopure form at a substrate concentration of 55 mM, higher than the one used in plain buffer (30 mM). Enantiopure (
S)-MOPE can be used to make cycloalkyl[
b]indoles which treat allergic responses.
In another study developed by the same group, it was demonstrated that DES ChCl:Ur (1:2 mol/mol) was able to improve the asymmetric reduction of 3-chloropropiophenone to (
S)-3-chloro-1-phenylpropan-1-ol (
3a,
Figure 2) catalyzed by
Acetobacter sp. CCTCC M209061 whole cells immobilized on PVA-sodium sulfate using glucose to recycle the cofactor [
32]. Among the different DES studied, ChCl:Ur increased the permeability of the bacterial cells as it was confirmed by flow cytometry. By augmenting it, the biocatalyst(s) could bind better the ketone substrate to convert it into the corresponding enantiopure alcohol. Hence, using 5% v/v of DES at 10 mM substrate concentration, the product yield was 82.3% and the
ee >99%. Later, utilizing a mixture of this DES with a water-immiscible ionic liquid (1-butyl-3-methylimidazolium hexafluorophosphate) in a biphasic system, it was possible to further increase the productivity of this transformation to 1.87 mmol/L.h.
In 2015, Müller et al. tested the effect of DES on the activity and stereoselectivity of ADHs from
Ralstonia sp. (RasADH),
Thermoanaerobacter ethanolicus (TeSADH) and horse liver (HLADH), which were overexpressed on
E. coli. They used aliphatic and aromatic ketones as substrates and different DES proportions. It was found that these solvents could enhance the selectivity of the ADHs, maintaining a good activity even at high DES concentrations. In addition, the best solvent was the mixture ChCl:Gly (1:2 mol/mol) and buffer at a proportion of 80:20 v/v. The three ADHs that were used showed excellent activities when different DES-buffer proportions were utilized with the substrates: octan-2-one (for TeSADH), benzaldehyde (for HLADH), and propiophenone (for RasADH). In another set of experiments, the performance of RasADH with propiophenone as a substrate was tested. When the ancillary cosubstrates (ethanol, propan-1-ol, and propan-2-ol) were tested in the DES-aqueous-media mixtures, conversion remained high (>80%) and
ee towards the (
S)-alcohol were enhanced up to >95%. For other substrates a similar trend was observed, obtaining high
ee for some aromatic substrates [
33].
A research study led by Xu et al. explored different DES compositions and their effects in the reduction of octan-2-one using whole cells of
Acetobacter pasteurianus GIM1.158 to obtain (
R)-octan-2-ol (
4a,
Figure 2). They confirmed that DES composed of ChCl:EG (1:2 mol/mol) offered the best results in this transformation. Moreover, when combining this DES with an imidazolium based water-immiscible IL, the productivity of this bioreduction could be highly improved. Thus, using a mixture of ChCl:EG (32% v/v) and 1-butyl-3-methylimidazolium hexafluorophosphate (C
4MIM•PF
6, 20% v/v), in the presence of propan-2-ol as a hydrogen donor, 2-octanone (1.5 M) could be reduced at 90% conversion providing the enantiopure (
R)-alcohol. Authors explained that DES increased the cell membrane permeability and kept the cells stable in the reaction system. In the case of the biphasic system, those solvents improved the substrate consumption by the cells because the second phase acted as a substrate reservoir, diminishing substrate or product inhibition and leading to a good yield of the desired product (
R)-octan-2-ol [
34].
Ethyl (
S)-4-chloro-3-hydroxybutanoate ((
S)-CHBE,
5a,
Figure 2) is known to be a precursor for drugs like atorvastatin calcium. In a study conducted by Dai and co-workers [
35], (
S)-CHBE was produced from the corresponding ketone using recombinant whole cells of
E. coli CCZU-T15. Thus, when using ChCl:Gly (1:2 mol/mol) at 12.5% v/v in the presence of surfactant Tween-80 and L-glutamine, the substrate could be efficiently transformed (>93%) into the enantiopure alcohol at very high substrate concentrations (>2 M). L-Glutamine participated in the biosynthesis of the nicotinamide cofactors and could promote the biocatalytic activity, so no external addition of NAD was necessary. Another reason behind the high production of the product (
S)-
5a, was that the solvents and the surfactant increased the membrane permeability of the cells helping the substrate to bind to the enzyme(s) involved. This study confirmed that DES containing glycerol were appropriate solvents for the bioconversions that used recombinant bacterial whole cells.
A study performed by Panic et al. used NADES as solvents for plant cells biocatalysis. They were able to reduce 1-(3,4-dimethylphenyl)ethanone to the corresponding enantioenriched alcohol using carrot root. In addition, the hydrolysis of (±)-1-phenylethyl acetate by carrot root was also achieved in a eutectic mixture based on ChCl. Interestingly, although conversions were lower than in plain water, inversion of the stereoselectivity was obtained when increasing the percentage of DES. The NADES that were used in this study were made at different molar ratios of choline chloride and glucose, xylose, xylitol, glycerol, or ethylene glycol. In addition, the pH of the water solutions varied from 3.3 to 7.1. Different selectivities were obtained when the water content was changed in the NADES. Thus, usually when employing percentages of 30% v/v of the eutectic solvent made of choline chloride and glucose, the (
R)-enantioenriched alcohol was obtained, while when increasing it up to 80% v/v, the (
S)-antipode was predominantly observed. The reactions were made using up to 80 mM substrate concentration. This is another example that shows how the use of DES in combination with multienzymatic systems can greatly improve the selectivity observed in conventional media [
36].
In 2016 Wei et al. were able to selectively oxidize racemic
2a (MOPE) with
Acetobacter sp. CCTCC M209061 cells more efficiently by the addition of DES in a two-phase system. In this research, the authors used two different systems in the biocatalytic asymmetric oxidation of MOPE to get (
S)-
2a. Water-immiscible organic solvents and ILs such as 1-butyl-3-methylimidazolium hexafluorophosphate were used as a second phase. Also, they added the DES ChCl:Gly (1:2 mol/mol) to the biphasic system to make the biocatalytic reaction more efficient. When the biocatalytic reaction occurred in the C
4MIM•PF
6/buffer medium, a substrate concentration of 65 mM could be converted at the maximum theoretical conversion value of 50.5%, affording the remaining enantiomer with an
ee value of >99% after 10 h. When ChCl:Gly (10% v/v) was added to the aqueous phase, this enhanced the biocatalytic oxidation rate of the substrate. Hence, the initial reaction rate increased to 124.0 µmol/min, and the reaction occurred successfully only after 7 h with 51.3% conversion at 80 mM substrate concentration. As it was described, when the concentration of
2a was enhanced, the initial rate increased continuously until having a substrate concentration of 80 mM. The highest initial reaction rate was attributed to the improved effect that DES had on the cell membrane permeability. The immobilized bacterial cells were able to keep 72% of their initial activity even after 9 batches of reuse in the C
4MIM•PF
6/ChCl:Gly‒buffer system. The researchers also used different systems to explore the transformation of MOPE to (
S)-
2a using the
Acetobacter sp. cells in a 500-mL preparative scale. The asymmetric oxidation of the racemic substrate in the presence of DES was better compared to the one that occurred in the aqueous buffer-IL system, obtaining the desired alcohol in an enantiopure form [
37].
In another investigation performed by Vitale et al., they explored the ability of baker’s yeast to reduce a series of ketones (
Table 2) using DES as main solvents [
38]. In those experiments, different eutectic mixtures were studied but the best results were obtained with the one composed of ChCl:Gly (1:2 mol/mol). The researchers proved that the amount of water and DES which were added to the reaction mixture was very important even for reversing the selectivity. Hence, the
anti-Prelog reduction of different ketones could be achieved when ChCl:Gly DES was used with the whole cell biocatalyst due to the inhibition of the (
S)-oxidoreductases present in it. In the case of phenylacetone (
6a), the corresponding alcohol was produced with a high stereoselectivity in pure water, giving (
S)-
6b in 96%
ee, while in DES‒aqueous mixtures up to 90% w/w of DES, the formation of the
R-enantiomer was clearly favored (up to 96%
ee). When water was used at 40% w/w, a racemic mixture of the alcohol was observed and there was no bioconversion when pure DES was employed in the reaction mixture. This team also proved that the outcome of the results depended on the composition of DES. The final stereoselectivity was obtained as a result of an interplay of whole cells “solvation” and the selective “inhibition” of some enzymes based on the nature of DES components. The DES that contained ChCl:glucose (2:1 mol/mol) and ChCl:Ur (1:2 mol/mol) deactivated the biocatalyst when used with 10% w/w water. Also, when water was replaced by ChCl:D-fructose DES mixture (3:2 w/w), this did not result in the conversion of phenylacetone to 1-phenylpropan-2-ol after five days. However, when this eutectic solvent was used with 40% w/w water, (
S)-1-phenylpropan-2-ol was obtained with an
ee value of 78% and the conversion was 31% after five days. This proved that the Baker’s yeast cells required a certain amount of water in order to show activity. It must be mentioned that the chiral (
S)- and (
R)-1-phenylpropan-2-ol (
6b) derivatives are used in the preparation of neuroprotective drugs.
When the aqueous‒ChCl:Gly eutectic mixtures (10:90 w/w) were used with arylpropanone derivatives bearing electron-withdrawing groups (F or CF
3), the chiral enantioenriched
R-configured alcohols were obtained (> 40%
ee) although at low extent. Thus, when 1-(4-fluorophenyl)propan-2-one (
7a) was incubated at 37°C in water with Baker’s yeast, this substrate was reduced to (
S)-
7b with 82% conversion in 94%
ee. However, the (
R)-enantiomer was preferentially attained when this substrate was dissolved in a DES‒water (80:20 w/w) mixture in 14% conversion and 82%
ee. When the amount of water was reduced to 10% w/w, the
ee of (
R)-
7b increased to 90% and the conversion was 10%. Also, 1-(4-(trifluoromethyl)phenyl)propan-2-one (
8a) was reduced to (
S)-
8b with very high conversion (90%) and
ee (98%) in plain water while to (
R)-
8b (15% conv. and 40%
ee) when using a mixture ChCl:Gly/water 90:10 w/w. With 1,1,1-trifluoro-3-phenylpropan-2-one (
9a), a racemic mixture of the alcohol
9b was achieved in water and the
ee increased to 24% to form the (
R)-enantiomer with a 12% yield utilizing the same DES at 90% w/w. When an additional methylene group existed between the aryl and the carbonyl group, the stereopreference did not change in the eutectic mixtures, obtaining the corresponding
S-enantiomers as in the aqueous medium but with lower conversions and selectivities. Unfortunately, an aromatic and a heteroaromatic ketone were slightly or not reduced using baker’s yeast in the absence or presence of DES [
38].
In a study conducted by Maczka et al. they used seven yeast strains to transform 3-acetyldihydrofuran-2(3
H)-one (
10a) and get access preferentially to the
anti-stereoisomers of the corresponding alcohol
10b (
Scheme 3) through a dynamic kinetic resolution. The authors also investigated the effect of the medium components on the efficiency and stereoselectivity of this biotransformation. Different strains of
Yarrowia reduced successfully
10a into
anti-
10b. Particularly, two strains of
Y. lipolytica performed this transformation with very high
ee. Hence, after one day, optically pure
anti-(3
S,1’
R)-
10b was produced using
Y. lipolytica P26A in a medium composed of yeast extract (1% w/w), peptone (2% w/w), and glucose (2% w/w). On the other hand,
Candida viswanathi AM120 was able to transform
10a to a mixture of two isomers:
anti-(3
R,1’
S)- and (3
S,1’
R)-
10b with a trace amount of the
syn-(3
S,1’
S)-enantiomer. Then, it was decided to add small quantities of different organic solvents to study their influence on the reductions mediated by the both yeast strains after 1 day. Thus, ethanol, glycerol, hexane and isopropanol were added to the medium (5-20% v/v) to enhance the solubility of the substrate. The alcohols could also regenerate the nicotinamide cofactors NADH or NADPH needed by the ADHs. In the case
C. viswanathi, a reduction in the enantiomeric excess of the
anti-(3
R,1’
S)-
10b diastereoisomer was usually observed. Also, a large decrease in the conversion of the product was seen when ethanol was added to the reaction with
Y. lipolytica P26A. In addition, the
syn isomers were also detected when large proportions of the organic solvent were added. The use of glycerol or hexane up to 10% v/v did not have an effect on the reaction. Due to the good results achieved with glycerol, the authors confirmed that when adding a DES composed of ChCl:Gly (1:2 mol/mol), the substrate could be completely reduced after seven days with
C. viswanathi AM120 cells in a buffer with a pH of 7. The DES also increased the stereoselectivity in the transformation. The addition of 10% v/v of DES on the resting cells resulted in the formation of
anti-(3
R,1’
S)-
10b with 85% diastereomeric excess and the
ee was 76%, higher than in the previous cases in the absence of any additive or in the presence of other organic solvents. The researchers tried other strains such as
Hansenula anomala C2 and
Saccharomyces pombe C1, but they observed negligible or very low conversions [
39].
Cicco et al. performed a research, where the asymmetric bioreduction of a series of ketones was achieved by using purified ketoreductases in the presence of DES as a part of a cascade protocol. These enzymes showed excellent catalytic performances and stabilities in mixtures made of DES and aqueous buffer. The most successful bioreductions occurred in the media containing ChCl:Gly (1:2 mol/mol)-buffer and ChCl:sorbitol (1:1 mol/mol)-buffer mixtures. It was demonstrated that when there was a high amount of DES in the mixture, a better enantioselectivity towards the formation of the secondary alcohol was obtained. Based on these results, they suggested that the DES-buffer mixtures could be used in a chemoenzymatic cascade process and could be run in both sequential and in concurrent modes: first the isomerization of racemic allylic alcohols, which is catalyzed by a ruthenium catalyst to afford the saturated ketone, and then this step was coupled with an asymmetric enzymatic reduction (
Table 3). Thus, a racemic mixture of the allylic alcohol was converted to an enantiopure saturated secondary alcohol (
R or
S) without using isolation and purification steps of the ketone intermediate [
40].
As a first step, it was explored the bioreduction of propiophenone to 1-phenylpropan-1-ol in different DES-buffer mixtures from 50 to 80% (w/w) using a set of ten commercial KREDs. The four different DES were composed of ChCl and Gly, Ur or lactic acid at a molar ratio of 1:2, and sorbitol at a molar ratio 1:1. Most of the KREDs used gave poor conversions in the presence of only ChCl, and these enzymes were inactive in the DES composed of Ur or lactic acid even at 50% w/w DES. However, the DES that contained sorbitol or Gly as a HBD and ChCl as a HBA allowed better conversions at 50% w/w. More than half of KREDs were still active in the presence of 80% w/w of both DES, staying especially very active in ChCl:Sorb (conv. >80%), particularly KRED-P2-C11. Since it is known that 16 out of 24 KREDs are derived from the short-chain dehydrogenase of Lactobacillus kefiri (LKADH), this enzyme was also tried under these conditions, showing a conversion for this ketone that varied from 80% to >99% in both DES at 50% and 80% w/w. Hence, ChCl:Gly and ChCl:Sorb mixtures were used as co-solvents in the reduction of different methyl and ethyl ketones. KRED-P2-C11 was an active enzyme in the presence of 80% w/w DES giving excellent conversion results for all tested ketones except for 1-(naphthalen-1-yl)propan-1-one. In some cases, the stereoselectivity was improved by increasing DES percentage. For instance, KRED-P2-C11 enhanced significantly the ee from 78% to >99% for 1-phenylpropan-1-ol and it increased the ee from 54% to >99% for 4-phenylbutan-2-ol when comparing the reaction in plain aqueous buffer. They also explored the stability of KRED-P2-C11 in DES-buffer media at different temperatures. For this, the researchers performed the reduction of propiophenone with ChCl:Gly-buffer 80:20 (w/w) at 30 °C, 40 °C, and 50 °C. While the temperature was increased up to 40 °C, the bioreduction occurred more rapidly than at 30 °C, at 50 °C the enzyme partially lost its activity.
As a final goal, the researchers tried to test whether the excellent activity of KRED in the reduction of ketones in the neoteric solvent could be used in a one pot process combining a metal-catalyzed isomerization of different allylic alcohols with the asymmetric KRED-mediated reduction of the obtained ketones (
Table 3). Firstly, the team focused on a sequential one-pot two-step methodology (entries 1–8). The metal-catalyzed isomerization was investigated in ChCl:Gly DES-buffer 50:50 w/w mixtures at 50 °C using 5 mol% of a ruthenium complex as a catalyst. When the isomerization was complete, the KRED and NADP
+ were added without isolating the intermediate ketone. Then the mixture was stirred for 24 h at 30 °C. Different secondary alcohols were obtained with excellent to quantitative conversions and high to excellent isolated yields (60–95%) and
ee values (93–99%). These results suggested that the KREDs were compatible with the reaction medium which came from the metal-catalyzed step, and that the impact of the Ru(IV) catalyst on the enzymatic performance was negligible. KRED enzymes got partially deactivated when the metal-catalyzed isomerization ran in a concurrent manner coupled with the bioreduction process in pure aqueous buffer, thus leading to low conversions into the allylic alcohols. However, when a mixture of DES-buffer 80:20 (w/w) was applied (entries 9–13) much better results were achieved. For instance, when 1-phenylprop-2-en-1-ol (
11a) was incubated at 40 °C and 250 rpm in this medium containing both catalysts, after 24 h the substrate was converted to (
R)-1-phenylpropan-1-ol (
12a) with 90% conversion and >99%
ee. Also, other racemic allylic alcohols (
11b‒
d) were successfully transformed to the enantiopure saturated (
R)-alcohols
12b‒
d with good to excellent conversions (68–96%) [
40].