You are on page 1of 20

Volume 18 Number 16 21 August 2016 Pages 4315–4572

Green
Chemistry
Cutting-edge research for a greener sustainable future
www.rsc.org/greenchem

Themed issue: Molecular Design for Reduced Toxicity


ISSN 1463-9262

PAPER
Nicholas Gathergood, Klaus Kümmerer et al.
On the way to greener ionic liquids: identification of a fully mineralizable
phenylalanine-based ionic liquid
Green Chemistry
PAPER

Synthesis of a series of amino acid derived ionic


Cite this: Green Chem., 2016, 18,
liquids and tertiary amines: green chemistry
4374 metrics including microbial toxicity and
preliminary biodegradation data analysis†
Andrew Jordan,a Annette Haiß,b Marcel Spulak,c Yevgen Karpichev,d
Klaus Kümmerer*b and Nicholas Gathergood*d

A series of L-phenylalanine ionic liquids (ILs), L-tyrosine ILs, tertiary amino analogues and proposed trans-
formation products (PTPs) have been synthesised. Antimicrobial toxicity data, as part of the green chem-
istry metrics evaluation and to supplement preliminary biodegradation studies, was determined for ILs,
tertiary amino analogues and PTPs. Good to very good overall yields (76 to 87%) for the synthesis of 6 ILs
from L-phenylalanine were achieved. A C2-symmetric IL was prepared from TMS-imidazole in a one-pot
two-step method in excellent yield (91%). Synthesis of the L-tyrosine IL derivatives utilised a simple pro-
tection group strategy by using an extra equivalent of the bromoacetyl bromide reagent. Improvements in
the synthesis of the α-bromoamide alkylating reagent from L-phenylalanine were achieved, directed by
green chemistry metric analysis. A solvent switch from dichloromethane to THF is described, however the
yield was 15% lower. Antimicrobial activity testing of L-phenylalanine ILs, L-tyrosine ILs, tertiary amino ana-
logues and PTPs, against 8 bacteria and 12 fungi strains, showed that no compound had a high anti-
microbial activity, apart from an L-proline analogue. In this exceptional case, the highest toxicity (IC95 =
125 and 250 µM) was observed towards the two Gram positive strains Staphylococcus aureus and Staphy-
lococcus epidermidis respectively. High antimicrobial activity was not found for the other bacteria or
fungi strains screened. The limitations of the antimicrobial activity study is discussed in relation to SAR
studies. Preliminary analysis of biodegradation data (Closed Bottle Test, OECD 301D) is presented. The
Received 12th February 2016, pyridinium IL derivative is the preferred green IL of the series based on synthesis, toxicity and biodegrada-
Accepted 7th June 2016
tion considerations. This work is a joint study with Kümmerer and co-workers and the PTPs were selected
DOI: 10.1039/c6gc00415f as target compounds based on concurrent biodegradation studies by the Kümmerer group. For the com-
www.rsc.org/greenchem prehensive biodegradation and transformation product analysis see the accompanying paper.

Introduction efficient synthesis, utilising lower toxicity chemicals (e.g.


reagents,2,3 solvents4 and catalysts5), avoiding non-biodegrad-
The overarching concept of the 12 Principles of Green Chem- able compounds, an iterative approach can ultimately lead to
istry is to promote the development and application of safer significant advances in the field.6–8 A class of chemicals which
chemicals, which will therefore lead to a reduced impact on has been extensively studied as safer and greener replacements
the environment.1 Whether this is due to a cleaner more for a wide range of chemicals is ionic liquids (ILs).9–11 ILs have
been previously defined as salts comprised of a cation and an
anion with melting points <100 °C, however in recent years
a
School of Chemical Sciences, Dublin City University, Glasnevin, Dublin 9, Ireland this arbitrary threshold is less often cited, and classification
b
Institute of Sustainable and Environmental Chemistry, University of Lüneburg,
can be based on chemical structure, functionality, and appli-
Scharnhorststraße 1, D-21335 Lüneburg, Germany.
E-mail: Klaus.Kuemmerer@uni.leuphana.de
cations.12 An attractive feature of ILs is the ability to fine-tune
c
Department of Inorganic and Organic Chemistry, Faculty of Pharmacy, the structure, to tailor the properties for a particular
Heyrovského 1203, Hradec Kralove, Czech Republic, CZ500 05 purpose.9–11 The availability of a series of closely related struc-
d
Department of Chemistry, Chair of Green Chemistry, Tallinn University of tures has led to many research teams determining the
Technology, Akadeemia tee 15, 12618 Tallinn, Estonia.
toxicity13–15 and ecotoxicity16–19 of ILs, performing structure
E-mail: nicholas.gathergood@ttu.ee
† Electronic supplementary information (ESI) available. See DOI: 10.1039/ activity relationship (SAR) studies13,20,21 and creating quanti-
c6gc00415f tative structure activity relationship (QSAR) models.21–23 Bio-

4374 | Green Chem., 2016, 18, 4374–4392 This journal is © The Royal Society of Chemistry 2016
Green Chemistry Paper

city studies and biodegradation analysis. A detailed biodegra-


dation and transformation product evaluation is reported in
the accompanying paper by Haiß et al.39

Results and discussion


The goal of this investigation was to design and develop a green
Fig. 1 One of the first ILs to pass the CO2 Headspace test (ISO 14592) synthesis of phenylalanine ethyl ester derived ILs in tandem
and be classed as readily biodegradable.38 with toxicity and biodegradation studies. Working closely with
Kümmerer’s group from the outset, concurrent synthesis, green
chemistry metrics, toxicity and biodegradation studies40 were
degradation of ILs has also been investigated (Fig. 1, 1),24–26 completed. Through this collaboration, improvements in the
although not as extensively as toxicity studies.14,20,27 Toxicity synthesis of the novel target ILs are potentially attainable, at the
and biodegradation data however is only part of a green chem- same time as their toxicity and biodegradation assessment. The
istry assessment of a chemical (e.g. IL).1 The synthetic route to detailed biodegradation evaluation of compounds presented
prepare the compound can also be considered, and a plethora herein is presented in the previous paper.39
of parameters can be determined,28–30 with the objective of The series of ILs 4–24, Fig. 3–5, were designed with respect
reducing the environmental impact of the process. In 2013, to Boethling’s rules of thumb for designing biodegradable
Gathergood and Connon published a combined microbial toxi- compounds,36 12 Principles of Green Chemistry,1 and the
city, biodegradation, green chemistry metrics and catalyst per- preceding generations of IL research conducted within the
formance study of a series of ILs.31–33 One of the findings from Gathergood research group.15,38,41–44 A selection of common IL
this work was that the ILs studied containing amide group(s) cation headgroups were chosen allowing this study to demon-
exhibited very low biodegradation in the CO2 Headspace test. strate the effect of headgroup on IL synthesis, toxicity and bio-
This was in agreement with previous results for amide contain- degradability, (Fig. 3, 4,15 5–15). L-tyrosine ILs analogues were
ing ILs,33–35 with an exception being 2 and 3 reported by Gath- prepared to examine the effects of the presence of a phenolic
ergood in 2012, (Fig. 2).15 This is in stark contrast to ILs –OH on the amino acid sidechain, (Fig. 4, 16–20). To investi-
containing an ester functional group, where facile breakdown gate how the –OH functionality would affect IL synthesis, tox-
to the carboxylic acid and conversion of the alcohol to CO2 is icity and biodegradability. It is known that the enzyme
widely reported.25,36 However, even with these ester containing phenylalanine hydroxylase (PheOH) converts phenylalanine
ILs, the carboxylic acid transformation product has been shown into tyrosine by an oxidative insertion mechanism42 and there
to persist after a 28 day CBT.37 The development of minerali- exists the possibility that the phenylalanine ILs could be
sable ILs is a significant and challenging research front. converted in vivo to the tyrosine derivatives before undergoing
Herein the authors describe their progress towards develop- biodegradation. The presence of the hydroxyl group is also
ing a green synthesis for biodegradable and mineralisable ILs. expected to increase water solubility of the tyrosine IL com-
The investigation includes synthesis of a series of analogues of pared to the phenylalanine analogue, and microbial activity can
2, green chemistry metrics analysis, preliminary microbial toxi- be compared. Lastly, a series of tertiary amino derivatives,
(Fig. 5, 21–24) were prepared to investigate the influence of a
positive charge due to a quaternary nitrogen on the compounds
examined. The ester chain length (ethyl) and counter-ion
(bromide) were not varied in this study. These were selected as
the focus of this study was the nitrogen headgroup of the IL.
Furthermore, the information obtained from this investigation
is expected to aid the rational design of “next generation” ILs
based on any observed structure activity relationships.
A series of L-phenylalanine ILs (4–15) were prepared in a
high yielding, three step synthesis. The synthesis is outline
below in Scheme 1 and Table 1.
First, Fischer–Speier esterification45 of L-phenylalanine
using thionyl chloride and excess ethanol to produce dry HCl
in situ was completed. The hydrochloride salt intermediate
(25·HCl) was neutralised with base and 2546 isolated in excel-
lent yield. Second, acylation with bromoacetyl bromide gave
the α-bromoamide of L-phenyalanine ethyl ester (26) (see
Fig. 2 Readily biodegradable amino acid ILs.15 Biodegradation values metrics section for optimisation details). Lastly, the α-bromo-
assessed by the CO2 Headspace test ISO 14593. amide alkylating reagent (26) was reacted with an amine (for

This journal is © The Royal Society of Chemistry 2016 Green Chem., 2016, 18, 4374–4392 | 4375
Paper Green Chemistry

Fig. 3 Series of L-phenylalanine ILs (4–15) targets. Charged headgroup study.

example, 1-methylimidazole), to yield the target ILs (4–15) nicotinate in diethyl ether at room temperature after 24 h
(Table 1). Compound 4 has been previously prepared by Gath- failed (entry 4). A fair yield (56%) however was attained for 7
ergood et al. in 98% yield using THF as the solvent,15 however by increasing the temperature, extending the reaction time
a slightly lower yield of 4 was obtained using diethyl ether and changing to a more polar solvent, THF (entry 5). A high
(entry 1). Excellent yields for the pyridinium and 3-methoxy- yield of IL 8 from 4-dimethylaminopyridine in diethyl ether
pyridinium ILs 5 and 6 (entries 2 and 3) however were was observed (entry 6). Aliphatic amines N-methylmorpholine
obtained in diethyl ether. Attempts to prepare IL 7 from ethyl and dimethylaminoethanol gave considerably different yields

4376 | Green Chem., 2016, 18, 4374–4392 This journal is © The Royal Society of Chemistry 2016
Green Chemistry Paper

Fig. 4 Series of L-tyrosine ILs (16–20) targets. Phenol study.

Fig. 5 Series of neutral L-phenylalanine IL analogues (21–24).

for corresponding IL synthesis in diethyl ether. A poor yield were prepared by reductive amination using formaldehyde
(34%) was observed for the morpholinium IL 9 (entry 7), while according to a procedure by Lin et al.,48 followed by Fischer–
10 was isolated in a quantitative yield (entry 8). C2-Symmetric Speier esterification (see ESI, Scheme S1†).46 Proline derivative
IL 11 was prepared by a two-step procedure from N-trimethyl- 31 required ethyl acetate as solvent for synthesis of IL 12 and
silylimidazole.47 The double alkylation proceeded in excellent 13 in a combined yield of 50% (entry 10 vs. 11). Similar results
yield (91%) if after 24 h in diethyl ether, the solvent was were obtained using the D-proline analogue 33 (entry 12). In
changed to ethyl acetate and heated at reflux for 24 h (entry 9). both cases (entries 11 and 12) a 5 : 2 ratio of epimers was
N-Methyl proline ethyl esters (31 and 33, entries 10 and 11) established in the crude product by 1H NMR. However, iso-

This journal is © The Royal Society of Chemistry 2016 Green Chem., 2016, 18, 4374–4392 | 4377
Paper Green Chemistry

Scheme 1 Synthesis of 26 from L-phenylalanine.

Table 1 Alkylation of tertiary amines with 26 to give ILs (4–15)

Entry NR1R2R3 Reaction conditionsa IL Yieldb (%)

1 N-Methylimidazole A 4 89
2 Pyridine A 5 98
3 3-Methoxypyridine A 6 96
4 Ethyl nicotinate A 7 0c
5 Ethyl nicotinate B 7 56
6 4-Dimethylaminopyridine A 8 88
7 N-Methylmorpholine A 9 34
8 Dimethylaminoethanol A 10 99
9 N-Trimethylsilylimidazole C 11 91
10 N-Methyl-L-proline-ethyl ester (31) A 12 0c
13 0c
11 N-Methyl-L-proline-ethyl ester (31) D 12 37
13 13
12 N-Methyl-D-proline-ethyl ester (33) D 14 25
15 34
a
A: Diethyl ether, rt, 24 h; B: THF, reflux, 48 h; C: diethyl ether, rt, 24 h then ethyl acetate, reflux, 24 h, 2 eq. 26; D ethyl acetate, reflux, 48 h.
b
See Table 4 for more detailed green chemistry metric analysis. c No product detected by TLC 10% MeOH : DCM.

lated yields of each epimer after purification by column occurred after 12 h (by TLC), and after acid hydrolysis with
chromatography are shown in Table 1. The absolute stereo- HCl in ethanol, 29 was obtained in an overall yield of 54%.
chemistry of prolinium derivatives (12, 13 and 14, 15) was Potential transesterification problems were avoided by select-
determined by 2-dimensional NMR by utilising the nuclear ing ethanol as the solvent. The final alkylation step in reflux-
overhauser effect (i.e. a NOESY experiment). Very good overall ing ethyl acetate gave ILs 16–20 in good to quantitative yields,
yields (76 to 87%) for the synthesis of ILs 4, 5, 6, 8, 10 and 11 with the highest yields obtained for ILs 17 and 18. IL 19
from L-phenylalanine were achieved. required purification by flash chromatography and a lower iso-
L-Tyrosine ILs (16–20) were synthesised according to the lated yield (71%) was obtained.
route outlined in Scheme 2. Initial attempts to selectively The neutral derivatives (21–24) were synthesised from alky-
acylate the amino group of 27 with bromoacetyl bromide were lating reagent (26) in poor to fair yields (34–56%). We propose
unsuccessful. While a range of phenol protection group strat- that this is due to the higher solubility of the products in
egies (for example, silyl ethers) were considered, the additional either diethyl ether or THF solvent. This effect was also seen
steps required were deemed undesirable. Due to the observed with the tyrosine series above. The compounds 21–24 were
ease of acylation of the phenol group, and expected higher purified by column chromatography and all isolated as liquids
reactivity of the formed phenolic ester vs. amide to hydrolysis, at room temperature (Table 2).
a double acylation of 26 route was proposed. We were pleased Very good to quantitative yields were obtained for the final
to find that quantitative conversion of 27 to intermediate 28 alkylation step in the synthesis of ILs 4, 5, 6, 8, 10, 11, 17 and

4378 | Green Chem., 2016, 18, 4374–4392 This journal is © The Royal Society of Chemistry 2016
Green Chemistry Paper

Scheme 2 General synthetic route employed for L-tyrosine bromide ILs (16–20).

Table 2 Alkylation of secondary amines with 26 to give ILs (21–24)

Entry HNR1R2 Solvent IL Yield (%)

1 N-Trimethylsilylimidazole Diethyl ether 21 55


2 N-Methylaminoethanol Diethyl ether 22 34
3 Morpholine THFa 23 56
b
4 L-Proline-ethyl ester·HCl THF 24 53
a
Reflux. b In situ neutralisation with Na2CO3.

18. The product precipitating from the reaction solvent signifi- nium ring gave an increase in melting point to 53–55 °C (17 vs.
cantly increases the yield compared to examples where chrom- 19) and significantly higher melting points were observed for
atography is required. the 3-methoxypyridinium derivative 18 117–119 °C and the
A wide range of melting points were also observed for the cholinium derivative 20 141–143 °C. The presence of the –OH
ILs synthesised. Under the classical definition of an IL existing group on the tyrosine ring greatly reduces the melting points
as a liquid below 100 °C all phenylalanine examples from of several of the ILs (16, 17 and 19) when compared to their
Fig. 3 (4–15) and Fig. 5 (21–24) with the exception of (8, 9 and L-phenylalanine counterparts. A decrease of 40 °C was observed
11) conform, though (5 and 10) have melting points close to for the imidazolium IL analogues (16 vs. 4), 68 °C for the pyri-
100 °C. A broad range of melting points (30–32 to 141–143 °C) dinium examples (17 vs. 5), and a slight decrease in melting
were also observed for the tyrosine derivatives with the imida- point of 7 °C for ethyl nicotinium analogues (19 vs. 7). Excep-
zolium IL 16 38–40 °C, and pyridinium IL 17 30–32 °C, the tions to this trend are the tyrosine salts, 3-methoxypyridinium
lowest recorded. Introduction of an ester group to the pyridi- IL 18 where an increase in melting point of 45 °C was observed

This journal is © The Royal Society of Chemistry 2016 Green Chem., 2016, 18, 4374–4392 | 4379
Paper Green Chemistry

and cholinium IL 20, with an increase of 40 °C when com- detected by TLC or 1H NMR. The outcome was a reaction
pared to their phenylalanine counterparts, 6 and 10 mixture that could be purified by filtration to remove the
respectively. Na2CO3 followed by aqueous bicarbonate solution washings of
the organic layer to remove any HBr and bromoacetic acid.
Green chemistry metrics Removal of the organic solvent by rotary evaporator furnished
We decided to adopt a two tier approach with the aim of devel- the desired alkylating reagent (26) in high yield (89%) as a
oping a greener synthesis for the ILs described herein. First, a white solid with no further purification required (entry 3). Other
preliminary evaluation of solvent selection would be under- solvents investigated as green replacements for DCM included
taken. This would be based guidelines proposed by Prat et al.4 ethyl acetate and THF. A mixture of decomposition products
and prioritise replacement of toxic solvents and reduce the was observed using ethyl acetate, although a good yield was
volume used while increasing yields. Second, a more detailed obtained of 26 (74%, entry 4) using THF. This solvent offers a
metric analysis would be completed for all reaction steps viable non-chlorinated solvent alternative to DCM, albeit with a
reported herein. This tiered approach to green chemistry lower isolated yield (89% vs. 74%, entry 2 vs. entry 4).
metrics has been proposed before by Clark et al.30 The following metrics were used to evaluate the synthetic
Comparing the reaction solvents used and yields attained methods used in producing the esters, alkylating reagents and
in each of the three main steps in the IL synthesis: (1) ester for- ILs: atom economy,50 E-factor,7 Andraos reaction mass
mation (ethanol), (2) α-bromoamide formation (DCM) and (3) efficiency,51 GSK reaction mass efficiency,52 mass intensity,53
alkylation (diethyl ether, THF and ethyl acetate), the second solvent intensity,53 stoichiometric factor28 and material recov-
step was selected for optimisation. ery parameter.54
The synthesis outlined in Scheme 1 is a modification of It can be seen from the generated metrics, Table 4 and
that employed by Coleman and Gathergood and presents an ESI,† that the final step in the synthesis of nearly all the ILs
overall improvement in the formation of the alkylating reagent had a 100% atom economy, IL 11 (80%) being the only excep-
26 from L-phenylalanine.15 The original method for the syn- tion. This is due to the use of the trimethylsilyl protection
thesis of 26 in the literature describes the reaction of amino group. Atom economy, however, does not take reaction stoi-
acid ester hydrochloride salt with trimethylamine (Table 3, chiometry, solvent use or reagents not incorporated into the
entry 1), followed by addition of bromoacetyl bromide at final product (e.g. auxiliary bases) into account hence
−78 °C. This method results in the formation of coloured additional metrics must be calculated. The stoichiometric
impurities if using TEA, possibly due to the Menschutkin reac- factor for all of the ILs ranges from 0.92–1.00, thus illustrating
tion49 or triethylamine promotion of trace by-products having that less than 10% excess of any one reagent was used in each
partial solubility in the organic layer that was not removed of the reactions. The stoichiometry employed in the synthesis
during an aqueous workup step. Impurities were also present of the ILs leads to a high GSK RME, but only when the reaction
if a heterogeneous base such as Na2CO3 was employed (entry yields are high. Yields above 90% for the ILs show good GSK
2). The neutralisation process of the ester hydrochloride salt RME. Thus, loss in GSK RME for ILs (6, 9 and 12–15) is due to
we suggest was too slow, thus leading to the formation of side lower yields.
products. Purification of 26 by column chromatography was The effect that column chromatography has on the green-
therefore necessary before progressing to the final alkylation ness of a synthetic pathway is highlighted when the E-factor,
step (entries 1 and 2). Using these methods 26 was isolated in solvent intensity and mass intensity of a process is calculated.
yields of 61 and 68% (entries 1 and 2).15 For ILs 4–6, 8–11, 17, 18 and 20, E-factors of 33–150 were
The synthesis route was modified by neutralising the amino observed (IL 16 has an E-factor of 294 even though purification
acid ester·HCl to the free base form, 25, before being used in by chromatography was not required and is due to the lower
the reaction. An excess of Na2CO3 was used as an auxiliary base yield of 83%). However, when IL purification involved column
to remove the HBr formed during the reaction. This pre-neutral- chromatography (7, 12–15 and 19), the E-factors rose consider-
isation step allowed the primary amine group of the amino acid ably (e.g. IL 13, E-factor 4088) and the greenness of the reac-
ester to rapidly and more efficiently react with the bromoacetyl tion based on this metric can be considered considerably
bromide. This significantly reduced the formation of unwanted lower compared to chromatography free trituration methods
side products. The reaction could also be run at room tempera- (e.g. IL 11, E-factor 33). The solvent required for column
ture (cf. −78 °C, entry 1)15 with no formation of side products chromatography also affects the mass and solvent intensity

Table 3 Optimisation of synthesis of 26

Entry Amine Reaction solvent Reaction time (h) Base Column chromatography required Yield 26 (%)

1 25·HCl DCM 5 TEA Y 6815


2 25·HCl DCM 5 Na2CO3 Y 61
3 25 DCM 3 Na2CO3 N 89
4 25 THF 5 Na2CO3 N 74

4380 | Green Chem., 2016, 18, 4374–4392 This journal is © The Royal Society of Chemistry 2016
Green Chemistry Paper

Table 4 Summary of green chemistry metrics for all compounds synthesised (see ESI for worksheet and calculations)

Atom Sheldon GSK Andraos Mass Solvent Stoichiometric Material recovery Yield
Compound economy50 E-factor7 RME52 RME51 intensity53 intensity53 factor28 parameter54 (%)

L-Phe ILs
4 100% 96.7 0.830 0.010 97.7 96.5 0.920 0.012 90
5 100% 93.4 0.964 0.011 94.4 93.3 0.991 0.011 98
6 100% 108.8 0.689 0.009 109.8 108.3 0.922 0.013 75
7 100% 1081.5 0.523 0.001 1082.5 1080.6 0.934 0.002 56
8 100% 120.7 0.812 0.008 121.7 120.5 0.918 0.010 89
9 100% 126.5 0.335 0.008 127.5 124.5 0.991 0.023 34
10 100% 88.1 0.901 0.011 89.1 88.0 0.917 0.012 98
11 80% 33.0 0.718 0.029 34.0 32.6 0.986 0.041 91
12 100% 1416.6 0.335 0.001 1417.6 1414.7 0.921 0.002 37
13 100% 3869.1 0.123 0.001 3870.1 3862.0 0.921 0.002 13
14 100% 1119.0 0.254 0.001 1120.0 1116.1 1.000 0.004 25
15 100% 828.2 0.343 0.001 829.2 826.2 1.000 0.004 34
L-Tyr ILs
16 100% 294.0 0.819 0.003 295.0 293.8 0.979 0.004 84
17 100% 117.1 0.980 0.008 118.1 117.1 0.982 0.009 99
18 100% 120.7 0.971 0.008 121.7 120.7 0.982 0.008 99
19 100% 420.3 0.699 0.002 421.3 419.9 0.986 0.003 71
20 100% 150.7 0.773 0.007 151.7 150.4 0.985 0.009 79
Neutral derivatives
21 66% 1035.4 0.288 0.001 1036.4 1032.9 0.783 0.003 56
22 79% 2754.1 0.242 0.001 2755.1 2750.9 0.923 0.001 33
23 80% 476.6 0.433 0.002 477.6 474.6 0.958 0.005 57
24 53% 770.3 0.266 0.001 771.3 767.6 0.946 0.005 53
Alkylating reagents
26a 59% 705.1 0.307 0.001 709.4 555.4 0.904 0.005 64
26b 49% 84.2 0.297 0.012 85.2 52.2 1.000 0.040 61
c
26 63% 10.1 0.470 0.090 11.1 7.4 0.843 0.192 89
26d 63% 36.7 0.428 0.027 37.7 7.3 0.918 0.062 74
29 50% 55.1 0.281 0.018 56.1 17.9 0.873 0.063 54
Amino acid esters
25 59% 20.8 0.425 0.046 21.8 11.8 0.735 0.108 99
27 71% 3.7 0.620 0.211 4.7 3.1 1.000 0.340 87
31 53% 14.6 0.372 0.064 15.6 12.0 1.000 0.172 70
33 53% 29.5 0.231 0.033 30.5 24.5 1.000 0.142 43
34 64% 5.6 0.610 0.150 6.6 5.0 0.959 0.247 99
Reductive alkylation
30 89% 11.6 0.837 0.080 12.6 11.4 0.980 0.093 98
32 89% 15.0 0.309 0.062 16.0 12.8 0.980 0.200 36

For conditions see a Table 3, entry 1; b Table 3, entry 2; c Table 3, entry 3; d Table 3, entry 4.

metrics in a similar manner and can be seen for ILs 7, 12–15 a generally lower E-factor than calculated for the ILs. The mod-
and 19 (Table 4). Material recovery parameters for all ILs are erate atom economy is due to the use of bromacetyl bromide
relatively low and this is due to only the product mass contri- which is not fully incorporated into the product. HBr is a by-
buting to the material recovery and all other materials are in product of the substitution reaction. The presence of an auxili-
effect waste. ary base also reduces the atom economy. The lower E-factor,
Andraos RME is low for all ILs as it takes into account all which is desirable, is due to the minimum amount of solvent
solvent used in the IL trituration steps and all chemicals that employed during the reaction and purification steps when
are not incorporated into the product are considered waste. compared to the ILs which required multiple triturations.
The diethyl ether and chromatography solvent should there- Amino acid esterification reactions (25, 27, 31, 33 and 34)
fore be recycled in future studies or if this process were to be all proceeded with excellent (e.g. 25, 3.7) to moderate E-factors
scaled up in order to reduce the mass/solvent intensity, (e.g. 31, 40.3) and low solvent and mass intensities. The atom
E-factor, material recovery parameter and Andraos RME. economies are affected by the use of stoichiometric quantities
The metrics generated for the neutral derivatives (21–24) of SOCl2 during the esterification reaction and ranged between
are less desirable as lower yields and a higher consumption of 53–90%.
solvent due to the purification by column chromatography was Lastly, the reductive alkylation of L-proline to give inter-
required. The GSK RME decreases (0.266–0.432) and E-factor mediate 30 can be considered as one of the greenest methods
increases (476.6–2754.1). employed during the synthesis reported here, with high atom
The alkylating reagents (26 and 29) were both synthesised economy (89%), low E-factor (11.6), excellent reaction mass
with moderate atom economy (63% and 50% respectively) and efficiency (0.837) and low solvent and mass intensity (12.6 and

This journal is © The Royal Society of Chemistry 2016 Green Chem., 2016, 18, 4374–4392 | 4381
Paper Green Chemistry

11.4 respectively). The counterpart to this reaction, 32, is less cular weight bases would result in fewer tons of exhausted
green due to the lower yield overall and thus lower GSK RME. base and resulting salts over a longer period. Solvent recycling
Atom economy remains unchanged and the E-factor is only from all aspects of the synthesis reported would dramatically
slightly increased to 15.0. improve the E-factor as the solvents would obviously no longer
During the reaction condition optimisation for alkylating be considered as waste. Material recovery parameters for all
reagent 26 (see Table 3 for optimisation), metric analysis was steps are low due to the inevitable waste produced during pro-
also conducted on each of the reaction conditions (entries cesses that do not have a 100% atom economy. Solvent re-
1–4). The original conditions reported by Coleman et al.15 for cycling would also aid in improving the material recovery
the formation of 26 involving the use of two equivalents of tri- parameter. Solvent intensity can be improved if reducing reac-
ethylamine and purification by column chromatography led to tion solvent can be tolerated while achieving the same yield.
a poor atom economy of 50% and an E-factor of 705.1. The Mass intensity can also be similarly improved.
solvent and mass intensities of the reaction are similarly
affected. Reaction mass efficiency was calculated to 0.307 due Microbial screening
to the moderate yield and excess trimethylamine required in Microbial toxicity screening was performed to principally
the reaction. When the TEA was replaced with Na2CO3 (entry identify any ILs with high activity. This approach enables us to
2), the atom economy was reduced to 49% due to the higher prioritise ILs which do not have high antimicrobial activity to
molecular weight of Na2CO3 compared to that of triethyl- be evaluated first for biodegradation studies. One possibility is
amine. The reaction mass efficiency is also reduced to 0.297. that ILs with high antimicrobial activities may inhibit micro-
The optimised conditions in DCM (entry 3) and THF (entry 4) organisms required for biodegradation to occur. This microbial
both show improved atom economies of 63% and good toxicity assessment represents a starting point for a more com-
E-factors (DCM 10.1, THF – 36.7). The GSK RME is also the prehensive investigation as part of a study to design safer ILs.
highest out of all the trialled conditions as shown in entry 3 Antibacterial activity results for 4–15 are shown in Table 5,
(0.470). Solvent and mass intensities are also low for the pro- 16–20 in Table 6 and 21–24 in Table 7. We were pleased
cedures described in entry 3 and 4 due to the elimination of to observe that all compounds 4–24 did not show high anti-
chromatographic purification. bacterial activity, except 24. In this exceptional case, the
The outlook from the optimisation metrics is that elimin- highest toxicity (IC95 = 125 and 250 µM) were observed towards
ation of column chromatography is essential for low E-factor the two gram positive strains Staphylococcus aureus (SA) and
and high GSK and Andraos RME. Avoiding the use of reagents Staphylococcus epidermidis (SE) respectively. The toxicity test
in excess, if possible, will improve the RME. The use of lower method is only validated up to a concentration of 2000 µM
molecular weight bases can improve atom economy, however it and suits our purpose to identify high activity compounds. We
is important to note that the number of moles of base required submit that IC values determined in the range 125–1000 µM,
will remain the same. This type of optimisation is more rele- although do not represent high antimicrobial activity, are sig-
vant for large scale production where the use of lower mole- nificant. While IC95 values lower than 2000 µM or the solubi-

Table 5 Antibacterial activity results obtained for L-phenylalanine ILs 4–15

IL IC95 (µM)
a
Microorganism Time (h) 4 5 6 7 8 9 10 11 12 13 14 15

Gram positive
SA 24 2000 2000 >500 1000 >500 2000 2000 500 >2000 1000 >2000 1000
48 >2000 >2000 >500 1000 >500 2000 >2000 500 >2000 1000 >2000 1000
MRSA 24 >2000 >2000 >500 1000 >500 2000 >2000 1000 >2000 >1000 >2000 >2000
48 >2000 >2000 >500 1000 >500 2000 >2000 1000 >2000 >1000 >2000 >2000
SE 24 >2000 >2000 >500 1000 >500 >2000 >2000 500 2000 1000 >2000 1000
48 >2000 >2000 >500 1000 >500 >2000 >2000 500 2000 1000 >2000 1000
EF 24 >2000 >2000 >500 >1000 >500 >2000 >2000 1000 >2000 >1000 >2000 >2000
48 >2000 >2000 >500 >1000 >500 >2000 >2000 1000 >2000 >1000 >2000 >2000
Gram negative
EC 24 >2000 >2000 >500 >1000 >500 >2000 >2000 2000 >2000 >1000 >2000 >2000
48 >2000 >2000 >500 >1000 >500 >2000 >2000 2000 >2000 >1000 >2000 >2000
KP 24 >2000 >2000 >500 >1000 >500 >2000 >2000 >2000 >2000 >1000 >2000 >2000
48 >2000 >2000 >500 >1000 >500 >2000 >2000 >2000 >2000 >1000 >2000 >2000
KP-E 24 >2000 >2000 >500 >1000 >500 >2000 >2000 >2000 >2000 >1000 >2000 >2000
48 >2000 >2000 >500 >1000 >500 >2000 >2000 >2000 >2000 >1000 >2000 >2000
PA 24 >2000 >2000 >500 >1000 >500 >2000 >2000 2000 >2000 >1000 >2000 >2000
48 >2000 >2000 >500 >1000 >500 >2000 >2000 2000 >2000 >1000 >2000 >2000
a
SA, Staphylococcus aureus; MRSA, Staphylococcus aureus MRSA; SE, Staphylococcus epidermidis; EF, Enterococcus sp.; EC, Escherichia coli; KP, Kleb-
siella pneumoniae; KP-E, Klebsiella pneumoniae, PA, Pseudomonas aeruginosa.

4382 | Green Chem., 2016, 18, 4374–4392 This journal is © The Royal Society of Chemistry 2016
Green Chemistry Paper

Table 6 Antibacterial activity of L-tyrosine ILs 16–20

IL IC95 (µM)
a
Microorganism Time (h) 16 17 18 19 20

SA 24 1000 >2000 >2000 >2000 2000


48 1000 >2000 >2000 >2000 2000
MRSA 24 2000 2000 >2000 >2000 >2000
48 >2000 2000 >2000 >2000 >2000
SE 24 500 >2000 >2000 >2000 1000
48 1000 >2000 >2000 >2000 2000
EF 24 >2000 >2000 >2000 >2000 >2000
48 >2000 >2000 >2000 >2000 >2000
EC 24 >2000 >2000 >2000 >2000 >2000
48 >2000 >2000 >2000 >2000 >2000
KP 24 >2000 >2000 >2000 >2000 >2000
48 >2000 >2000 >2000 >2000 >2000
KP-E 24 >2000 >2000 >2000 >2000 >2000
48 >2000 >2000 >2000 >2000 >2000
PA 24 >2000 >2000 >2000 >2000 >2000
48 >2000 >2000 >2000 >2000 >2000
a
SA, Staphylococcus aureus; MRSA, Staphylococcus aureus MRSA; SE, Staphylococcus epidermidis; EF, Enterococcus sp.; EC, Escherichia coli; KP, Kleb-
siella pneumoniae; KP-E, Klebsiella pneumoniae, PA, Pseudomonas aeruginosa.

lity limit (e.g. 16) are shown in Tables 5–7, no significant SAR 4 gram positive strains is clearly apparent. In addition, all com-
comparisons can be made due to the similarity of values. It pounds 4–24 do not show high antibacterial activity to the four
would be ill advised to draw conclusions based on comparing gram negative strains.
IC95 values of 500 µM and 2000 µM for selected strains and Antifungal activity results for 4–24 are shown in Table 8.
undertaking a SAR study based on bacterial toxicity. Once Although compounds 9, 11, 12 and 16 have IC50 and IC80
again the only exception would be with compound 24, where values of 2000 µM for some fungal strains, the data overall
there is nearly an order of magnitude difference between the supports the finding that 4–24 do not have high antifungal
IC95 values for SA and the 4 gram negative strains. This, activity. Based on the antibacterial and antifungal activity
however, represents an isolated trend in the context of all the results all compounds were selected for biodegradation
data presented in Tables 5–7. However, the general trend that studies. Compound 24 was included in the biodegradation
compounds 4–23 do not show high antibacterial activity to the study, despite having the highest antimicrobial toxicity
recorded. This decision was based on the combined antibac-
terial and antifungal data, where high toxicity was only
observed for 4 gram positive bacteria strains and not for
Table 7 Antibacterial activity results obtained for neutral amino com-
pounds 21–24
4 gram negative bacteria strains, 9 yeasts and 3 filamentous
fungi strains (Table 8).
IL IC95 (µM) When compared to microbial screening on amino acid
a derived ILs previously carried out by Coleman et al. the most
Microorganism Time (h) 21 22 23 24
active antifungal ILs were comprised of L-valine (35 and 36),
SA 24 1000 1000 2000 125 (Fig. 6), towards the filamentous fungal strains (AC, TM).15
48 1000 1000 2000 125 IC50 values as low as 62.5 µM were determined for L-valine
MRSA 24 1000 2000 2000 500
48 1000 2000 2000 500 ethyl ester analogue 35.
SE 24 1000 500 >2000 250 Compounds presented within this publication (4–24) were
48 1000 500 >2000 250 determined to be less toxic towards the fungal strains
EF 24 >2000 >2000 >2000 >1000
48 >2000 >2000 >2000 >1000 screened. We propose that L-phenylalanine is a preferable
EC 24 >2000 >2000 >2000 >1000 amino acid when considering the synthesis of low toxicity
48 >2000 >2000 >2000 >1000 amino acid ILs.
KP 24 >2000 >2000 >2000 >1000
48 >2000 >2000 >2000 >1000
KP-E 24 >2000 >2000 >2000 >1000
48 >2000 >2000 >2000 >1000 Preliminary biodegradation studies of ionic liquids and
PA 24 >2000 >2000 >2000 >1000 tertiary amino analogues
48 >2000 >2000 >2000 >1000
The biodegradation screening was carried out according to the
a
SA, Staphylococcus aureus; MRSA, Staphylococcus aureus MRSA; SE, Closed Bottle test (CBT).40,55,57 As can be seen from Table 9, of
Staphylococcus epidermidis; EF, Enterococcus sp.; EC, Escherichia coli;
KP, Klebsiella pneumoniae; KP-E, Klebsiella pneumoniae; PA, Pseudomo- the ILs screened, imidazolium IL 4 did not reach the 60% bio-
nas aeruginosa. degradation threshold required to pass the CBT. Pyridinium IL 5,

This journal is © The Royal Society of Chemistry 2016 Green Chem., 2016, 18, 4374–4392 | 4383
Paper Green Chemistry

Table 8 Antifungal activity results for L-phenylalanine, L-tyrosine and 3° amine compounds 4–24

IL MICb (µM)

Microorganisma Time (h) 4–6, 10, 13–15, 17–23 7, 24 8 9 11 12 16

CA1 24 >2000 >1000 >500 >2000 >2000 >2000 >2000


48 >2000 >1000 >500 >2000 >2000 >2000 >2000
CA2 24 >2000 >1000 >500 >2000 >2000 >2000 >2000
48 >2000 >1000 >500 >2000 >2000 >2000 >2000
CP 24 >2000 >1000 >500 >2000 >2000 >2000 2000
48 >2000 >1000 >500 >2000 >2000 >2000 >2000
CK1 24 >2000 >1000 >500 2000 2000 >2000 >2000
48 >2000 >1000 >500 2000 >2000 >2000 >2000
CK2 24 >2000 >1000 >500 2000 >2000 >2000 >2000
48 >2000 >1000 >500 2000 >2000 >2000 >2000
CT 24 >2000 >1000 >500 >2000 2000 >2000 >2000
48 >2000 >1000 >500 >2000 2000 >2000 >2000
CG 24 >2000 >1000 >500 2000 >2000 >2000 >2000
48 >2000 >1000 >500 2000 >2000 >2000 >2000
CL 24 >2000 >1000 >500 >2000 2000 >2000 >2000
48 >2000 >1000 >500 >2000 2000 >2000 >2000
TA 24 >2000 >1000 >500 2000 >2000 >2000 >2000
48 >2000 >1000 >500 2000 >2000 >2000 >2000
AF 24 >2000 >1000 >500 >2000 >2000 >2000 >2000
48 >2000 >1000 >500 >2000 >2000 >2000 >2000
AC 24 >2000 >1000 >500 2000 >2000 >2000 >2000
48 >2000 >1000 >500 >2000 >2000 >2000 >2000
TM 72 >2000 >1000 >500 2000 2000 2000 >2000
120 >2000 >1000 >500 2000 2000 >2000 >2000
a
CA1, Candida albicans; CA2, Candida albicans; CP, Candida parapsilosis; CK1, Candida krusei; CK2, Candida krusei; CT, Candida tropicalis; CG,
Candida glabrata; CL, Candida lusitaniae; TA, Trichosporon asahii; AF, Aspergillus fumigatus; AC, Absidia corymbifera; TM, Trichophyton mentagro-
phytes. b IC50 values were assessed for AF, AC and TM and IC80 for all other strains.

Table 9 Biodegradability and percentage carbon distribution of com-


pounds 4–24, colour coded according to a traffic light classification33

Fig. 6 Previously reported amino acid ILs with antifungal activity (IC50)
by Coleman et al.15

attained the highest biodegradation of 63%, however cannot


be considered readily biodegradable under OECD guidelines
as biodegradation took more than 10 days after degradation
reached 10% ThOD.56,57 This still represents a significant
breakthrough for high biodegradability of halide IL salts. The
addition of an ether group at the 3-position to give the 3-meth-
oxypyridinium IL 6 reduced the biodegradability to 52%. A
similar observation was made for the nicotinate IL 7 and
DMAP derivative 8, although the decrease to 32% and 33%
respectively is more significant. Pyridinium ILs substituted at
the 3 position with ester groups are known to have good bio-
degradation properties24,26 and our result is an exception to
a
this trend. The C2 symmetric imidazolium IL 11 was also Biodegradation data for 5–8, 10–12, 21, 22 and 24 previously reported
in accompanying paper by Haiß et al.39 b One of the two bottles was
shown to be recalcitrant to biodegradation with the lowest leaking. c Not valid (biodegradation values in the replicates differ more
value (17%) recorded. than 20%).

4384 | Green Chem., 2016, 18, 4374–4392 This journal is © The Royal Society of Chemistry 2016
Green Chemistry Paper

The non-aromatic derivatives based on a prolinium (12–15)


scaffold gave poor results in the CBT. A narrow range of values
(21–24%) was determined which suggests that for these four
diastereomers, absolute stereochemistry was not a significant
factor.
The presence of an additional –OH group on the L-tyrosine
derived ILs 16–20 appears not to promote biodegradation.
When comparing L-phenylalanine derivative 4 to L-tyrosine
derivative 16 only a difference of 7% (40 vs. 47%) was deter-
mined. Indeed, all the L-tyrosine analogues of L-phenylalanine
ILs (4–7) and 10 gave similar results within only a 5–13%
Fig. 7 Carbon atom distribution of amino acid IL (4).
difference, which is not large enough to identify specific
trends. This is in contrast to rules of thumb that phenols are
better biodegradable than phenyl groups as also shown for the
plus the ester is considered, then these predicted levels of bio-
antibiotics amoxicillin (with a phenol ring, mineralizable in
degradation suggest that many of the ILs and amino com-
Zahn Wellens test) and piperacillin (even in the ZWT not
pounds (e.g. 4, 5, 10, 11, 16, 17, 20 and 21–23) could achieve
biodegradable).58
>60% biodegradation, based on complete biodegradation of
The similar levels of biodegradability for the L-phenyl-
the aminoacid ester moiety (Table 9). ILs (4–11, and 16–20) are
alanine and L-tyrosine derivatives suggests that the L-phenyl-
observed to have levels of biodegradation greater than only
alanine may not be undergoing conversion into L-tyrosine
ester degradation but less than complete amino acid ester
through enzymatic transformations by hydroxylase enzymes.59
degradation, except 5. However, this rough approximation of
The L-phenylalanine is instead being degraded directly and
carbon distribution and attributable degradation cannot be
may not be entering the homogentisate pathway.60 However it
fully substantiated without a transformation product analysis
cannot be inferred from the data directly and analysis of the
of the CBT. What exactly is breaking down? Is the rate of
metabolic products by a suitable LCMS technique to deter-
biodegradation such that some of the compound in the test
mine the biodegradation pathways and metabolic products
undergoes high levels of mineralisation after a particular bio-
formed is required.39
chemical transformation? Is there some rate limiting step that
None of the tertiary amino derivatives (21–24) CBT results
once completes, allows the IL to biodegrade? Is ester hydro-
can be considered a pass (Table 9). For 21 and 24 valid data
lysis occurring preferentially to amide bond hydrolysis or are
based on two replicates was not obtained in this initial bio-
both happening at the same time? Is there fully intact parent
degradation study. For subsequent CBT investigations and
compound present in the CBT vessel after the 28 day period?
more detailed biodegradation results and discussion see Haiß
To answer these questions a detailed biodegradation assess-
et al.39 22 and 23 did not reach the 60% threshold, within 28
ment, including transformation product identification and
days, and thus does not pass the CBT. However, when the ter-
analysis of the transformation products oxygen demand is
tiary amino derivative 22 was compared to the IL analogue it
required, and is published in the accompanying paper.39
can be seen that tertiary amino derivative 22 undergoes higher
Based on the data presented in Table 9 we cannot state with
levels of biodegradation (+22%) than IL analogue 10. The elev-
absolute certainty that a general trend is biodegradation of all
ated levels of biodegradation suggest that the presence of a
ester groups occurs for 4–24.
quaternary nitrogen atom, substituted with a methyl group,
can impede the biodegradation process, thus tertiary amine 22
can exhibit higher levels of biodegradation. This observation is Transformation product
in agreement with the literature,6 but in stark contrast to the To assist in the identification of transformation products
CBT results for 9 vs. 23 (i.e. not a significant difference; 9, 17% formed during the CBT, synthesis of possible compounds
and 23, 28%) Therefore, a larger sample size of closely related which are consistent with proposed structures based on LCMS
tertiary amino compounds and ILs would need to be examined studies was initiated. While significant progress has been
to identify any general trends. made developing analytical techniques to identify transform-
An assumption which is often made when interpreting bio- ation products there are several benefits to preparing authentic
degradation data of ILs is that ester group hydrolysis is facile samples by total synthesis. This includes unambiguously con-
and complete conversion (e.g. to CO2) under the test con- firming the structure of the transformation product on com-
ditions occurs.25,33 Table 9 includes the percentage biodegra- parison with an authentic sample, increasing the scope of
dation attributed to this process and the percentage carbon synthetic chemistry expertise for the preparation of close
content of the aminoacid ester and ester subunits as exempli- derivatives of compounds of interest. A limitation of LCMS
fied in Fig. 7. studies of biodegradation samples is that due to the low con-
For a number of compounds, the percentage biodegrada- centration used in the test (2–10 mg L−1), and sampling, even
tion observed is greater than that which can be attributed to using preparative HPLC techniques, obtaining sufficient amounts
ester degradation. When the carbon content of the amino acid of the transformation product for further antimicrobial toxicity

This journal is © The Royal Society of Chemistry 2016 Green Chem., 2016, 18, 4374–4392 | 4385
Paper Green Chemistry

Scheme 3 Selective hydrolysis of ethyl ester of 10.

and biodegradation studies is challenging. Our approach was to LiOH mediated base hydrolysis of cholinium IL 10 procedure
establish new synthetic routes to the proposed transformation pro- to furnish PTP 37 was performed in good yield (Scheme 3). We
ducts, provide authentic samples, and due to the sufficient decided not to use this method for examples were two esters
amounts prepared, determine whether these compounds have and an amide were present and the selective hydrolysis of one
high antimicrobial activity. We submit that toxicity data of con- ester was required. Future investigations using enzymatic
firmed transformation products assists in the design of safer methods will search for a green solution to this problem.
chemicals. Of note, due to the collaborative nature of this work, it PTP 40 was synthesised according to a benzyl ester hydroge-
is only by detailed LCMS studies that the structure of the trans- nolysis route via product 39 (Scheme 4). Starting compound
formation products can be confirmed. Therefore, herein we will 3862 underwent a substitution reaction with 26 in ethyl acetate
classify the transformation products (TP) as proposed transform- based on conditions described in Table 1 for the preparation
ation products (PTP) and the reader is directed to the accompany- of L-phenylalanine ILs. Deprotection of the benzyl ester 39
ing paper for the related LCMS studies.39 using Pd/C catalysed hydrogenolysis afforded the carboxylic
The first strategy for the synthesis of proposed transform- acid PTP 40 in 87% yield. Efforts to use the same approach for
ation products (PTP) 37 and 40–46 was based on facile conver- the nicotinium PTPs 41 and 42 failed due to reduction of the
sion of ILs already prepared as part of this investigation. pyridinium ring in the final step. Therefore, a different strategy
Selective hydrolysis of an ester bond in the presence of an avoiding the use of a benzyl protecting group was devised. We
amide is widely reported in peptide synthesis.61 To this end a were pleased to find that sodium nicotinate was a suitable sub-

Scheme 4 Benzyl ester hydrogenolysis route to PTP 40.

4386 | Green Chem., 2016, 18, 4374–4392 This journal is © The Royal Society of Chemistry 2016
Green Chemistry Paper

Scheme 5 PTP 41 synthesis from sodium nicotinate.

Scheme 6 PTP 42 synthesis from L-phenylalanine.

strate for the alkylation reaction with 26 (Scheme 5). This gave
directly the transformation product 41, although the yield was
fair, 37% due to challenging purification by column chromato-
graphy. Although α-bromoamides (e.g. N,N-diethyl) have been
used to N-alkylate nicotinic acid previously,63a albeit in low
yields, this headgroup has not been widely adopted in recent
IL studies.63b The IL 41 was only selected as a target due to the
biodegradation and transformation product joint investigation
with the Kümmerer group.
For the synthesis of PTP 42 a four step method was used
(Scheme 6). L-Phenylalanine was converted to the sodium salt
then reacted with bromoacetyl bromide under biphasic con-
ditions to give the alkylating reagent (2-bromoacetyl)-L-phenyl- Fig. 8 PTPs 43–47.
alanine according to a procedure by Qin et al.63c Treatment
with ethyl nicotinate in refluxing DMF, then HBr gave PTP 42
Table 10 Antibacterial activity results obtained for PTPs 37 and 40–47
in 33% overall yield.
For PTPs (43, 44 and 45) preparation, the route involved IL IC95 (µM)
synthesis of either an ethyl or a methyl ester of the target a
Microorganism Time (h) 37, 40–47 46
heterocycle followed by acid hydrolysis to give the carboxylic
acid PTP in a two-step reaction. The synthetic procedure is as SA 24 >2000 >500
outlined by Pukitis et al. who originally synthesised 43 from 48 >2000 >500
MRSA 24 >2000 >500
dimethylamino pyridine (DMAP).64 PTP 44 and 45 have been
48 >2000 >500
previously reported by Oldfield et al.65 and Deng et al.,37 SE 24 >2000 >500
respectively. PTP 46 was obtained according to the literature 48 >2000 >500
EF 24 >2000 >500
procedure outline by Farfán et al.66 in which methyl-
48 >2000 >500
ethanolamine is reacted with aqueous glyoxal to form a mor- EC 24 >2000 >500
pholine intermediate which undergoes hydrolysis in situ to 48 >2000 >500
KP 24 >2000 >500
give the carboxylic acid compound 46. PTP 47 was commer-
48 >2000 >500
cially available and purchased from TCI Europe (Fig. 8). KP-E 24 >2000 >500
Antimicrobial activity data shows that none of the PTPs (37 48 >2000 >500
PA 24 >2000 >500
and 40–47) have high toxicity to the 20 strains screened
48 >2000 >500
(Tables 10 and 11). This data is consistent with the results for
a
compounds 1–23 in Tables 5–8. ILs (41 and 43) showed the SA, Staphylococcus aureus; MRSA, Staphylococcus aureus MRSA; SE,
Staphylococcus epidermidis; EF, Enterococcus sp.; EC, Escherichia coli;
highest level of fungal toxicity (Table 11). IL 41 had an IC95 of KP, Klebsiella pneumoniae; KP-E, Klebsiella pneumoniae; PA, Pseudomo-
2000 µM towards the two Candida albicans strains whilst IL 43 nas aeruginosa.

This journal is © The Royal Society of Chemistry 2016 Green Chem., 2016, 18, 4374–4392 | 4387
Paper Green Chemistry

Table 11 Antifungal activity results obtained for PTPs 37 and 40–47 Good to very good overall yields (76 to 87%) for the syn-
thesis of ILs 4, 5, 6, 8, 10 and 11 from L-phenylalanine were
IL MICb (µM)
achieved. C2-Symmetric IL 11 was prepared from TMS-imida-
Microorganisma Time (h) 37, 40, 45, 42, 44 41 43 46 zole in a one pot two step method in excellent yield (91%). Syn-
thesis of the L-tyrosine IL derivatives used a simple protection
CA1 24 >2000 2000 >2000 >500
group strategy by using an extra equivalent of the bromoacetyl
48 >2000 2000 >2000 >500
CA2 24 >2000 2000 >2000 >500 bromide reagent. Improvements in the synthesis of α-bromo-
48 >2000 2000 >2000 >500 amide 2615 were made. Reducing the volume of solvent required
CP 24 >2000 >2000 2000 >500
and removing the requirement to run the reaction at low temp-
48 >2000 >2000 2000 >500
CK1 24 >2000 >2000 >2000 >500 erature (−78 °C) lead to a greener method. A procedure repla-
48 >2000 >2000 >2000 >500 cing DCM with THF was also demonstrated, however the yield
CK2 24 >2000 >2000 >2000 >500
decreased by 15%. Green chemistry metrics were determined
48 >2000 >2000 >2000 >500
CT 24 >2000 >2000 >2000 >500 for the synthesis of ILs and their amino derivatives and rec-
48 >2000 >2000 >2000 >500 ommendations for improvements in future studies compiled.
CG 24 >2000 >2000 >2000 >500
We submit that the reactions can be classified into two types.
48 >2000 >2000 >2000 >500
CL 24 >2000 >2000 >2000 >500 First, quantitative to excellent yield has been established and
48 >2000 >2000 >2000 >500 the focus will be on reducing solvent use, using safer solvents
TA 24 >2000 >2000 >2000 >500
and reducing the amount of chemicals required, where poss-
48 >2000 >2000 >2000 >500
AF 24 >2000 >2000 >2000 >500 ible. This will lead to improved outputs for the green chemistry
48 >2000 >2000 >2000 >500 metrics, providing evidence that a greener synthesis has been
AC 24 >2000 >2000 >2000 >500
developed. Second, low to good yield reactions would benefit
48 >2000 >2000 >2000 >500
TM 72 >2000 >2000 2000 >500 from a broader approach to improve the synthesis. Alternative
120 >2000 >2000 2000 >500 green synthetic methodologies need to be trialled in combi-
a
CA1, Candida albicans; CA2, Candida albicans; CP, Candida parapsilo- nation with the aforementioned points. This is the strategy we
sis; CK1, Candida krusei; CK2, Candida krusei; CT, Candida tropicalis; are following for the synthesis of the next generation of ILs as
CG, Candida glabrata; CL, Candida lusitaniae; TA, Trichosporon asahii; part of this body of work.
AF, Aspergillus fumigatus; AC, Absidia corymbifera; TM, Trichophyton
mentagrophytes. b IC50 values were assessed for AF, AC and TM and IC80 Antimicrobial activity testing of L-phenylalanine ILs, L-tyro-
for all other strains. sine ILs, tertiary amino analogues and PTPs, against 8 bacteria
and 12 fungi strains, showed that no compound had a high
antimicrobial toxicity, apart from 24. In this exceptional case,
displayed an IC95 of 2000 µM towards the fungi Candida para- the highest toxicity (IC95 = 125 and 250 µM) were observed
psilosis and Trichophyton mentagrophytes. Even with these ILs towards the two gram positive strains SA and SE respectively.
the general trend is not a high antifungal activity when consid- High antimicrobial activity was not found for the other bac-
ering all 12 strains. teria or fungi strains screened. We were pleased to establish
We submit that this data significantly expands the scope of that the PTPs did not show high toxicity to the microorgan-
compounds in this study which do not show high microbial isms screened. Due to the maximum validated sample con-
toxicity. Not only can this supplement the earlier results centration of the microbial toxicity test (2000 µM, or
(Tables 5–8), assisting the design of greener ILs, but also pro- solubility limit) and the samples not having high toxicity, a
vides fundamental toxicity data of PTPs to aid biodegradation SAR model cannot be established. However, when comparing
studies. the phenylalanine and tyrosine examples with the valine ana-
logue 35, a lower antifungal activity was demonstrated for
4–24.
Preliminary CBT biodegradation results39 for 4–24 gave a
Conclusion wide range of values (17–63%). The highest value, obtained for
pyridinium IL 5, passes the CBT pass 60% threshold, but
A series of L-phenylalanine ILs, L-tyrosine ILs, tertiary amino further biodegradation studies and HRMS analysis were
analogues and PTPs have been synthesised. The PTPs were required before being able to state whether the compound was
selected as target compounds based on concurrent biodegra- fully biodegradable or not.39 The biodegradation data in
dation studies by the Kümmerer group.39 Antimicrobial toxi- Table 9 should not be rationalised according to substructures
city data, as part of the green chemistry metrics evaluation of the ILs and neutral amino compounds (e.g. headgroup,
and to supplement biodegradation studies, was determined aminoacid ester and ester). In particular, stating that every com-
for ILs, tertiary amino analogues and PTPs. We propose that pound undergoes complete conversion of the alcohol
this combined study by the Gathergood and Kümmerer groups fragment of the ester, after a hydrolysis step. Although, indeed
has led to advances in the synthesis and design of low anti- this was an approach our group previously used to account for
microbial toxicity ionic liquids with favourable biodegradation low biodegradation of a series of over 20 ILs.67 In that study,
properties. no compound was found to give a biodegradation value higher

4388 | Green Chem., 2016, 18, 4374–4392 This journal is © The Royal Society of Chemistry 2016
Green Chemistry Paper

than that which can be attributed to ester transformation, so ing an amide group. The conclusion that 5 is the preferred
that assumption was consistent with the data available. This is green IL from this study is only possible due to the close
not the case with compounds 4–24. A far more complex biodegra- collaboration of research groups with expertise in organic
dation process occurs with this series, and based on Table 9 synthesis, biodegradation and toxicity screening. Promoting the
data alone it would be ill-advised to state that ester hydrolysis ‘benign by design’ approach to the development of low toxicity
accounts for the biodegradation values observed. For example, and biodegradable chemicals is our mutual goal.68–70
ILs 12–15 gave 21–23% biodegradation under CBT conditions
with 19% carbon content accounted for by the ethyl fragment
of the ester group. To assume that these results were linked, Experimental
and then design new greener ILs based on this finding, would
Antibacterial activity experimental method
be unadvisable without supporting TP data. One conclusion
that can be made from the data is that the presence of the OH In vitro antibacterial activities71 of the compounds were evalu-
group in tyrosine compared to phenylalanine did not lead to ated on a panel of three ATCC strains (Staphylococcus aureus
an increase in biodegradation. Due to this result, the tyrosine ATCC 6538, Escherichia coli ATCC 8739, Pseudomonas aerugi-
examples were not prioritised for biodegradation LCMS nosa ATCC 9027) and five clinical isolates (Staphylococcus
studies/TP identification and the focus was only the study of aureus MRSA HK5996/08, Staphylococcus epidermidis HK6966/
phenylalanine compounds prepared. 08, Enterococcus sp. HK14365/08, Klebsiella pneumoniae
To assist in the identification of transformation products HK11750/08, Klebsiella pneumoniae ESBL HK14368/08) from
formed during the CBT, synthesis of possible compounds the collection of fungal strains deposited at the Department of
which are consistent with proposed structures based on LCMS Biological and Medical Sciences, Faculty of Pharmacy, Charles
studies was completed. This was by either simple hydrolysis of University, Hradec Králové, Czech Republic. The above-men-
IL targets already in hand (37 from 10), from the stockpile of tioned ATCC strains also served as the quality control strains.
key alkylating reagent (41 from 26) or total synthesis (40 from All the isolates were maintained on Mueller-Hinton agar prior
38; 42 from phenylalanine). We submit that the approach of to being tested. Dimethyl sulfoxide (100%) served as a diluent
the first two methods offers a green solution to the preparation for all compounds; the final concentration did not exceed 2%.
of the required PTPs for this project based on materials and Mueller-Hinton agar (MH, HiMedia, Čadersky-Envitek, Czech
expertise already in the group. It is advisable that future Republic) buffered to pH 7.4 (±0.2) was used as the test
investigations should include crucial TP analysis. Thus, devel- medium. The wells of the microdilution tray contained 200 μL
oping robust green synthetic methods to support these studies of the Mueller-Hinton medium with 2-fold serial dilutions of
is a worthwhile endeavour. Of note also is the case of 41 syn- the compounds (2000 to 0.488 μmol L−1) and 10 μL of inocu-
thesis from sodium nicotinate. The requirement to prepare 41 lum suspension. Inoculum in MH medium was prepared to
as a PTP lead to the discovery that sodium nicotinate is a toler- give a final concentration of 0.5 McFarland scale (1.5 × 108 cfu
ated substrate for the alkylation reaction of 26. This headgroup mL−1). The trays were incubated at 37 °C and MICs were read
can now be included in our ongoing investigations in IL visually after 24 h and 48 h. The MICs were defined as 95%
research. While the microbial toxicity data showed that none inhibition of the growth of control. MICs were determined
of the PTPs have high activity, which was a desirable result, it twice and in duplicate. The deviations from the usually
is the evaluation step which is an important aspect. One poss- obtained values were no higher than the nearest concentration
ible outcome from this microbial toxicity screening was that value up and down the dilution scale.
PTPs with high antimicrobial toxicity were found. Although
the significance of this outcome should not be overstated (only Antifungal activity experimental method
20 microbial strains tested), this is useful data for experts in In vitro antifungal activities of the compounds were evaluated
biodegradation and TP analysis studies. The formation of a TP on a panel of four ATCC strains (Candida albicans ATCC 44859,
which is a potent biocide during biodegradation is obviously Candida albicans ATCC 90028, Candida parapsilosis ATCC
undesirable and all data which can support research to avoid 22019, Candida krusei ATCC 6258) and five clinical isolates of
this eventuality is valuable. yeasts (Candida krusei E28, Candida tropicalis 156, Candida
The preferred green IL from this study was pyridinium com- glabrata 20/I, Candida lusitaniae 2446/I, Trichosporon asahii
pound 5. A short and efficient synthesis of 5 was described, 1188) and filamentous fungi (Aspergillus fumigatus 231, Absidia
with improvements achieved in reducing solvent waste, the corymbifera 272, Trichophyton mentagrophytes 445) from the col-
required for low temperature alleviated and a lower toxicity lection of fungal strains deposited at the Department of Bio-
solvent replacement option described. Green chemistry metrics logical and Medical Sciences, Faculty of Pharmacy, Charles
provide evidence to support the statement that a greener syn- University, Hradec Králové, Czech Republic. Three ATCC
thesis was developed. IL 5 does not have high antimicrobial strains were used as the quality control strains. All the isolates
activity to the strains screened and has excellent biodegradation were maintained on Sabouraud dextrose agar prior to being
properties. Although from the data in this paper we cannot state tested. Minimum inhibitory concentrations (MICs) were deter-
whether 5 is readily biodegradable, passing the 60% threshold mined by modified CLSI standard of microdilution format of
of the CBT test is a significant result for a halide IL contain- the M27-A3 and M38-A2 documents.72,73 Dimethyl sulfoxide

This journal is © The Royal Society of Chemistry 2016 Green Chem., 2016, 18, 4374–4392 | 4389
Paper Green Chemistry

(100%) served as a diluent for all compounds; the final con- 3 X. Liu, S. S. Li, Y. M. Liu and Y. Cao, Chin. J. Catal., 2015,
centration did not exceed 2%. RPMI 1640 (Sevapharma, 36, 1461–1475.
Prague) medium supplemented with L-glutamine and buffered 4 D. Prat, A. Wells, J. Hayler, H. Sneddon, C. R. McElroy,
with 0.165 M morpholinepropanesulfonic acid (Serva) to pH S. Abou-Shehada and P. J. Dunn, Green Chem., 2016, 18,
7.0 by 10 M NaOH was used as the test medium. The wells of 288–296.
the microdilution tray contained 200 μL of the RPMI 5 R. A. Sheldon, I. Arends, G. J. Ten Brink and A. Dijksman,
1640 medium with 2-fold serial dilutions of the compounds Acc. Chem. Res., 2002, 35, 774–781.
(2000 to 0.488 μmol L−1 for the new compounds) and 10 μL of 6 R. S. Boethling, E. Sommer and D. DiFiore, Chem. Rev.,
inoculum suspension. Fungal inoculum in RPMI 1640 was pre- 2007, 107, 2207–2227.
pared to give a final concentration of 5 × 103 ± 0.2 cfu mL−1. 7 R. A. Sheldon, C. R. Acad. Sci., Ser. IIc: Chim., 2000, 3, 541–
The trays were incubated at 35 °C and MICs were read visually 551.
after 24 h and 48 h. The MIC values for the dermatophytic 8 P. Anastas and N. Eghbali, Chem. Soc. Rev., 2010, 39, 301–
strain (T. mentagrophytes) were determined after 72 h and 312.
120 h. The MICs were defined as 80% inhibition (IC80) of the 9 J. P. Hallett and T. Welton, Chem. Rev., 2011, 111, 3508–
growth of control for yeasts and as 50% inhibition (IC50) of the 3576.
growth of control for filamentous fungi. MICs were deter- 10 V. I. Pârvulescu and C. Hardacre, Chem. Rev., 2007, 107,
mined twice and in duplicate. The deviations from the usually 2615–2665.
obtained values were no higher than the nearest concentration 11 N. V. Plechkova and K. R. Seddon, Chem. Soc. Rev., 2008,
value up and down the dilution scale. 37, 123–150.
12 K. Seddon, presented in part at the COIL-6, Jéju, South
Biodegradation method Korea, 2015.
For biodegradation method see Haiß et al.39 13 L. Carson, P. K. W. Chau, M. J. Earle, M. A. Gilea,
B. F. Gilmore, S. P. Gorman, M. T. McCann and
General synthetic methods K. R. Seddon, Green Chem., 2009, 11, 492–497.
Compounds 4,15 25,74 26,15 27,75 30,76 31,77 32,76 3378 and 3479 14 K. M. Docherty and C. F. Kulpa, Green Chem., 2005, 7, 185–
were prepared according to literature methods. Proposed trans- 189.
formation products 43,64 44,65 4537 and 4666 were prepared by 15 D. Coleman, M. Spulak, M. T. Garcia and N. Gathergood,
literature methods. Commercially available PTP 47 was pur- Green Chem., 2012, 14, 1350–1356.
chased from TCI Europe. Experimental procedures and charac- 16 A. S. Wells and V. T. Coombe, Org. Process Res. Dev., 2006,
terisation of all novel compounds, including NMR spectra, are 10, 794–798.
available in the ESI.† 17 S. P. M. Ventura, C. S. Marques, A. A. Rosatella,
C. A. M. Afonso, F. Goncalves and J. A. P. Coutinho, Ecotoxi-
col. Environ. Saf., 2012, 76, 162–168.
18 C. W. Cho, Y. C. Jeon, T. P. T. Pham, K. Vijayaraghavan and
Acknowledgements Y. S. Yun, Ecotoxicol. Environ. Saf., 2008, 71, 166–171.
The authors would like to thank the EPA in Ireland for finan- 19 J. I. Santos, A. M. M. Goncalves, J. L. Pereira, B. F. H.
cial support. This project has also received funding from the T. Figueiredo, F. A. e Silva, J. A. P. Coutinho,
European Union’s Seventh Framework Programme for S. P. M. Ventura and F. Goncalves, Green Chem., 2015, 17,
research, technological development and demonstration under 4657–4668.
grant agreement No. 621364 (TUTIC-Green). The antibacterial 20 B. Jastorff, R. Stormann, J. Ranke, K. Molter, F. Stock,
and antifungal screening was supported by the Czech Science B. Oberheitmann, W. Hoffmann, J. Hoffmann, M. Nuchter,
Foundation ( project No. 15-07332S). We also thank COST B. Ondruschka and J. Filser, Green Chem., 2003, 5, 136–142.
Actions CM1206 and TD1203 for their contribution to this 21 S. Stolte, J. Arning, U. Bottin-Weber, M. Matzke, F. Stock,
study. We acknowledge Hannah Prydderch for her contri- K. Thiele, M. Uerdingen, U. Welz-Biermann, B. Jastorff and
butions to PTP synthesis and Dr Andrew Kellett for his assist- J. Ranke, Green Chem., 2006, 8, 621–629.
ance with the project at Dublin City University. We thank 22 P. Luis, I. Ortiz, R. Aldaco and A. Irabien, Ecotoxicol.
Prof. A. Lapkin at the University of Cambridge and BRITEST Environ. Saf., 2007, 67, 423–429.
Ltd for providing the Green Metrics Lab Book spreadsheet. 23 C. W. Cho, J. Ranke, J. Arning, J. Thoeming, U. Preiss,
C. Jungnickel, M. Diedenhofen, I. Krossing and S. Stolte,
SAR QSAR Environ. Res., 2013, 24, 863–882.
References 24 D. Coleman and N. Gathergood, Chem. Soc. Rev., 2010, 39,
600–637.
1 P. T. Anastas, Green Chemistry: Theory and Practice, ed. 25 S. Stolte, S. Steudte, A. Igartua and P. Stepnowski, Curr.
P. T. Anastas and J. C. Warner, Oxford University Press, Org. Chem., 2011, 15, 1946–1973.
New York, 1998. 26 A. Jordan and N. Gathergood, Chem. Soc. Rev., 2015, 44,
2 P. Tundo and M. Selva, Acc. Chem. Res., 2002, 35, 706–716. 8200–8237.

4390 | Green Chem., 2016, 18, 4374–4392 This journal is © The Royal Society of Chemistry 2016
Green Chemistry Paper

27 M. Petkovic, K. R. Seddon, L. P. N. Rebelo and C. S. Pereira, 53 C. Jiménez-González, D. Constable, A. Curzons and


Chem. Soc. Rev., 2011, 40, 1383–1403. V. Cunningham, Clean Technol. Environ. Policy, 2002, 4, 44–
28 J. Andraos, Org. Process Res. Dev., 2005, 9, 149–163. 53.
29 B. M. Trost, Angew. Chem., Int. Ed. Engl., 1995, 34, 259–281. 54 J. Andraos, Pure Appl. Chem., 2011, 83, 1361–1378.
30 C. R. McElroy, A. Constantinou, L. C. Jones, L. Summerton 55 S. Sasu, J. W. Metzger, M. Kranert and K. Kümmerer, Clean:
and J. H. Clark, Green Chem., 2015, 17, 3111–3121. Soil, Air, Water, 2015, 43, 166–172.
31 L. Myles, R. G. Gore, N. Gathergood and S. J. Connon, 56 OECD, OECD Guideline for Testing of Chemicals, 1992.
Green Chem., 2013, 15, 2740–2746. 57 OECD, Test No. 301: Ready Biodegradability, OECD Publish-
32 R. G. Gore, T.-K.-T. Truong, M. Pour, L. Myles, S. J. Connon ing, 1992.
and N. Gathergood, Green Chem., 2013, 15, 2727–2739. 58 A. Längin, R. Alexy, A. König and K. Kümmerer, Chemo-
33 R. G. Gore, L. Myles, M. Spulak, I. Beadham, T. M. Garcia, sphere, 2009, 75, 347–354.
S. J. Connon and N. Gathergood, Green Chem., 2013, 15, 59 P. F. Fitzpatrick, Annu. Rev. Biochem., 1999, 68, 355–381.
2747–2760. 60 E. Arias-Barrau, E. R. Olivera, J. M. Luengo, C. Fernandez,
34 J. R. Harjani, R. D. Singer, M. T. Garciac and B. Galan, J. L. Garcia, E. Diaz and B. Minambres, J. Bacter-
P. J. Scammells, Green Chem., 2009, 11, 83–90. iol., 2004, 186, 5062–5077.
35 L. Myles, R. Gore, M. Spulak, N. Gathergood and 61 V. M. Krishnamurthy, V. Semetey, P. J. Bracher, N. Shen
S. J. Connon, Green Chem., 2010, 12, 1157–1162. and G. M. Whitesides, J. Am. Chem. Soc., 2007, 129, 1312–
36 A. Jordan and N. Gathergood, Chem. Soc. Rev., 2015, 44, 1320.
8200–8237. 62 P. Zaderenko, M. S. Gil, P. Ballesteros and S. Cerdan, J. Org.
37 Y. Deng, I. Beadham, M. Ghavre, M. F. Costa Gomes, Chem., 1994, 59, 6268–6273.
N. Gathergood, P. Husson, B. Legeret, B. Quilty, 63 (a) B. Potter, J. Dowden, A. Galione, A. Guse and A. Flugel,
M. Sancelme and P. Besse-Hoggan, Green Chem., 2015, 17, US Pat. Appl, US20070105810, 2007; (b) Z. Wang, C. Wang,
1479–1491. W. Bao and T. Ying, J. Chem. Res., 2005, 6, 388–390;
38 N. Gathergood, P. J. Scammells and M. T. Garcia, Green (c) H.-P. Li, S. Li, Z.-D. Wang and L. Qin, J. Korean Chem.
Chem., 2006, 8, 156–160. Soc., 2011, 55, 978–982.
39 A. Haiß, A. Jordan, J. Westphal, E. Logunova, N. Gathergood 64 O. Y. Neilands, E. V. Shebenina and G. G. Pukitis, Int. Arch.
and K. Kümmerer, Green Chem., 2016, DOI: 10.1039/ Allergy Appl. Immunol., 1999, 35, 1443–1450.
C6GC00417B. 65 J. M. Sanders, Y. Song, J. M. W. Chan, Y. Zhang,
40 J. Friedrich, A. Längin and K. Kümmerer, Clean: Soil, Air, S. Jennings, T. Kosztowski, S. Odeh, R. Flessner,
Water, 2013, 41, 251–257. C. Schwerdtfeger, E. Kotsikorou, G. A. Meints, A. O. Gómez,
41 Y. Deng, S. Morrissey, N. Gathergood, A. M. Delort, P. Husson D. González-Pacanowska, A. M. Raker, H. Wang, E. R. van
and M. F. C. Gomes, ChemSusChem, 2010, 3, 377–385. Beek, S. E. Papapoulos, C. T. Morita and E. Oldfield, J. Med.
42 S. Morrissey, B. Pegot, D. Coleman, M. T. Garcia, Chem., 2005, 48, 2957–2963.
D. Ferguson, B. Quilty and N. Gathergood, Green Chem., 66 N. Farfán, L. Cuéllar, J. M. Aceves and R. Contreras, Syn-
2009, 11, 475–483. thesis, 1987, 927–929.
43 M. T. Garcia, N. Gathergood and P. J. Scammells, Green 67 R. G. Gore, L. Myles, M. Spulak, I. Beadham, M. T. Garcia,
Chem., 2005, 7, 9–14. S. J. Connon and N. Gathergood, Green Chem., 2013, 15,
44 N. Gathergood, M. T. Garcia and P. J. Scammells, Green 2747–2760.
Chem., 2004, 6, 166–175. 68 K. Kümmerer, Green Chem., 2007, 9, 899–907.
45 E. Fischer and A. Speier, Ber. Dtsch. Chem. Ges., 1895, 28, 69 I. Beadham, M. Gurbisz and N. Gathergood, Handbook
3252–3258. of Green Chemistry, Chapter 6, Volume 9: Designing
46 E. Kumazaki and H. Nagano, Tetrahedron, 2013, 69, 3486– Safer Chemicals, ed. R. Boethling and A. Voutchkova, Wiley-
3494. VCH Verlag GmbH & Co. KGaA, 1st edn, 2012, pp. 137–
47 Z. Fei, D. Zhao, T. J. Geldbach, R. Scopelliti and P. J. Dyson, 158.
Chem. – Eur. J., 2004, 10, 4886–4893. 70 I. Beadham, M. Gurbisz and N. Gathergood, Handbook of
48 Y. H. L. Nan-horng, R. L. Elliott, M. S. Chorghade, Green Chemistry, Chapter 7, Volume 9: Designing Safer
S. J. Wittenberger, W. H. Bunnelle, B. A. Narayanan, Chemicals, ed. R. Boethling and A. Voutchkova, Wiley-VCH
P. Singam, T. K. J. Esch, D. O. Beer, C. C. Witzig, Verlag GmbH & Co. KGaA, 1st edn, 2012, pp. 159–226.
T. C. Herzig and A. R. Rao, Eur. Pat, EP0717741, 2001. 71 Reference Method for Broth Dilution Antifungal Suscepti-
49 Z. Wang, in Comprehensive Organic Name Reactions and bility Testing of Yeasts. Approved standard. Document
Reagents, John Wiley & Sons, Inc., 2010, vol. 426, pp. M27-A3. Clinical Laboratory Standard Institute, Wayne, PA,
1897–1900. 2008.
50 B. M. Trost, Science, 1991, 254, 1471–1477. 72 Reference Method for Broth Dilution Antifungal Suscepti-
51 J. Andraos, Org. Process Res. Dev., 2006, 10, 212–240. bility Testing of Filamentous Fungi. Approved standard.
52 A. D. Curzons, D. J. C. Constable, D. N. Mortimer and Document M38-A2. Clinical Laboratory Standard Institute,
V. L. Cunningham, Green Chem., 2001, 3, 1–6. Wayne, PA, 2008.

This journal is © The Royal Society of Chemistry 2016 Green Chem., 2016, 18, 4374–4392 | 4391
Paper Green Chemistry

73 Methods for Dilution Antimicrobial Susceptibility Tests for 76 K. D. Daughtry, Y. Xiao, D. Stoner-Ma, E. Cho,
Bacteria that Grow Aerobically. Approved Standard – A. M. Orville, P. Liu and K. N. Allen, J. Am. Chem. Soc.,
Seventh Edition. Document M07-A7. Clinical Laboratory 2012, 134, 2823–2834.
Standard Institute, Wayne, PA, 2006. 77 D. M. Evans, W.O. Pat, 2011051673, 2011.
74 J. Hartwig, J. B. Metternich, N. Nikbin, A. Kirschning and 78 A. M. Likhosherstov, A. S. Lebedeva, V. N. Kulakov,
S. V. Ley, Org. Biomol. Chem., 2014, 12, 3611–3615. A. M. Kritsyn, D. A. Kharkevich and A. P. Skoldinov, Pharm.
75 M. Ousmer, N. A. Braun, C. Bavoux, M. Perrin and Chem. J., 1973, 7, 274–277.
M. A. Ciufolini, J. Am. Chem. Soc., 2001, 123, 7534– 79 R. Almansa, D. Guijarro and M. Yus, Tetrahedron: Asymme-
7538. try, 2007, 18, 2828–2840.

4392 | Green Chem., 2016, 18, 4374–4392 This journal is © The Royal Society of Chemistry 2016

You might also like