You are on page 1of 5

DOI: 10.1002/slct.

201800204 Communications

1
2 z Organic & Supramolecular Chemistry
3
4
5
Direct Formation of Amides from Carboxylic Acids and
6 Amines Catalyzed by Niobium(V) Oxalate Hydrate
7
8 Zhihui Wang,*[a] Xinyu Bao,[a] Mengying Xu,[a] Zihao Deng,[a] Yongshun Han,[a] and
9 Ningfeng Wang[b]
10
11
A wide range of metal salts and oxides were screened to test variation of metal types and possible activity promotion
12
the catalytic effect in the formation of amides. A water- accompanied by organic ligands.[9] Therefore, the investigation
13
compatible catalyst, Niobium(V) oxalate hydrate, proved the on the metal-based systems, promoted by ligands, may lead to
14
ability to catalyze the direct formation of amides from a possibility of a highly efficient catalytic system of the direct
15
carboxylic acids and amines with a low catalyst loading of 1% amide formation.
16
in a 1:1 ratio of acids and amines, with the water removal The direct amide formation procedure, as Scheme 1 shows,
17
through azeotropic distillation of toluene/water mixtures, has some advantages: 1) direct use of carboxylic acids without
18
which showed a good atomic economic efficiency. Both
19
aromatic and aliphatic carboxylic acids worked well with
20
primary aliphatic amines in the reaction, however, aniline, as an
21
example of primary aromatic amines, needed more than one
22
equivalent amount to complete the amide formation.
23
24 Scheme 1. Direct amide formation from carboxylic acids and amines,
25 illustrated by the case of secondary amides.
26
27 Introduction
28
Amide bond has a large range of existence in polymers like the oxidation level change or activation,[10] meanwhile the
29
proteins or polyamides, and small organic compounds, natural stable existence of carboxylic acids and large amount of natural
30
or man-made.[1] It is noteworthy that amides play an important carboxylic acids such as natural fatty acids, which means a big
31
role in pharmaceutical field because of the 25% appearance of convenience from the synthetic perspective; 2) avoidance of
32
known drugs.[2] In spite of the established synthetic methods of stoichiometric usage of toxic coupling reagents, like DCC,
33
amides from activated carboxylic acids, such as acyl chloride or HATU, DEPBT et cetera, which have already been widely used in
34
a mixed or symmetrical anhydride, isolated or not,[3] a lot of the synthesis of pharmaceuticals;[11] 3) giving the only by-
35
efforts have been made to develop new ways to build up product of water, which means a clean and an atom-efficient
36
amide bonds,[4] among which the catalytic amide formation way.
37
directly from carboxylic acids and amines draws a large amount
38
of attention because of the potential value in synthetic aspect,
39
such as peptide synthesis.[5] Many catalytic ways have been Results and Discussion
40
developed in the direct amide synthesis from non-activated
41 Screening of metal-based catalysts
carboxylic acids and amines, for example, enzymatic systems,[6]
42
boron-based catalysts[7] and metal-based catalysts.[8] The metal- The reaction between benzoic acid and phenethylamine was
43
based systems, especially homogeneous metal-catalyzed proto- chosen to test the catalytic reactivity of metal-based com-
44
cols show significant importance, because of the wide range of pounds, during which water was supposed to be removed
45
from reaction system by azeotropic distillation together with
46
toluene. A lot of compounds were tested, such as Cp2ZrCl2,
47
[a] Dr. Z. Wang, X. Bao, M. Xu, Z. Deng, Y. Han Nb2O5, NbCl5, CuCl, CuBr, CuI, Cu2O, CuSO4.5H2O, CuBr2, CuCl2,
48 College of Science Fe, FeCl3, FeCl2.4H2O, Pd(OAc)2, Fe2O3, AgNO3, ZnBr2, CeCl3.7H2O,
49 State Key Laboratory of Heavy Oil Processing
Zn, LiCl, Mn(acac)3, MnCl2.4H2O, Mn(OAc)2.4H2O, MnO2,
50 China University of Petroleum (Beijing)
102249 Beijing (China) LaCl3.7H2O, SnCl2, Co(OAc)2.4H2O, Co(acac)2, CoCl2, NiCl2.6H2O,
51
E-mail: wangzhihui@cup.edu.cn TiO2, V(acac)3, ZrCl4, AlCl3, SnCl4, among which AlCl3, FeCl3,
52 [b] N. Wang Cp2ZrCl2, SnCl4 and NbCl5, those five metal chlorides, showed
53 Pipeline Marketing Company
higher reactivity, above 10% GC yield, than the others.
54 Petrochina Pipeline Company
065000 Langfang (China) In our screening result, NbCl5 showed a good catalytic
55
Supporting information for this article is available on the WWW under reactivity, with a Niobium of + 5 valence. Interestingly, when
56
https://doi.org/10.1002/slct.201800204 benzoic acid was firstly mixed and stirred with NbCl5 in toluene
57

ChemistrySelect 2018, 3, 2599 – 2603 2599  2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Wiley VCH
1809 / 107472
Freitag, 02.03.2018
[S. 2599/2603] 1
Communications
before addition of phenethylamine, the reaction gave nearly Nb2O5, calcined from niobic acid at T = 500 oC for 3 h, can
1
the same yield as the reaction of mixing benzoic acid, catalyze the amide formation reaction efficiently.[12] This differ-
2
phenethylamine, NbCl5 in toluene at the beginning. Never- ence may result from the following points: 1) Nb2O5 in our
3
theless, when phenethylamine was firstly mixed and stirred screening process was used directly from bottle, as 100 mesh
4
with NbCl5 in toluene before addition of benzoic acid, the power, which had been purchased from Alfa aesar company
5
reaction gave a relatively low GC yield of 36%, which means without any further treatment; 2) Nb2O5 in the literature was
6
the interaction between NbCl5 and benzoic acid may have newly achieved by calcination from niobic acid, which led to a
7
some connection to the true catalytic species. So in this NbCl5 very large surface area of 54 m2g 1; 3) the load amount in our
8
involved amide formation reaction, benzoic acid may have dual process was 5 mol-% of benzoic acid, which was different from
9
roles, one as carboxylic acid supplier and one as part of the the corresponding amount in literature, 50 mg Nb2O5 per mmol
10
catalytic species. Since niobium(V) chloride is easily hydrolyzed of carboxylic acid, about 19 mol-% of carboxylic acid.[12a]
11
in the reaction condition, even in air, it cannot play as the true
12
catalytic species, and besides, it is not easily handled in
13 Optimization of reaction conditions
weighing and transfer. Based on the inference above, a
14
combination of multivalent metal center and peripheral ligand, The optimization process of direct amidation (Table 2), modeled
15
like carboxylate ion, may lead to a possible catalytic system for by benzoic acid and phenethylamine, showed that 1% of
16
the direct amide formation. Eventually, niobium(V) oxalate niobium(V) oxalate hydrate could catalyze the reaction to have
17
hydrate (Nb(HC2O4)5.xH2O), a combination of niobium(V) ion 88% isolated yield within 18 hours (entry 6). Higher amounts of
18
and oxalate, seems to meet this need, which gave a GC yield of niobium(V) oxalate hydrate were found to accelerate reaction
19
40% in the screening experiment and is stable to air or water, speed (entries 1–3), while lower amounts of niobium(V) oxalate
20
in spite of that some other commercial available niobium hydrate led to a slower reaction (entries 4–5). Extended reaction
21
compounds, such as LiNbO3, MgNb2O6, Nb4C5, Nb3N5, didn’t time, from 16 h to 24 h, didn’t show obvious promotion in yield
22
show much activity (not beyond 5% of GC yield under the (entries 6–8). Despite the azeotropic distillation of mixture of
23
same condition). These results, mentioned above, are illustrated toluene and water to remove the water in the amide formation
24
in Table 1. process, we also tested the way of predried 4 Å molecular
25
Additionally, in our screening process, the commercial sieves to absorb water from reaction system, but it led to lower
26
available Nb2O5 didn’t show remarkable catalytic reactivity (7% yield than azeotropic method with toluene.
27
of GC yield), although, according to Shimizu group’s report,
28
29 Scope of amide formation reaction catalyzed by niobium(V)
30 oxalate hydrate
Table 1. Screening of metal-based catalysts for amide formation.[a]
31
With the optimized conditions, we then investigated the scope
32
of amide formation reaction with niobium(V) oxalate hydrate as
33
catalyst. In the case of nonaromatic amines, the amide
34
formation reaction went smoothly to afford amide products in
35
moderate or high yield (Table 3). For instance, when reacted
36 Entry Catalyst GC yield [%]
with phenethylamine (2 a), aromatic carboxylic acids worked
37
1 AlCl3 18
38 2 FeCl3 11
39 3 Cp2ZrCl2 10
40 4 SnCl4 22 Table 2. Screening of the reaction conditions for amide formation reaction
5 NbCl5 56 catalyzed by Niobium(V) oxalate hydrate.[a]
41
6[b] NbCl5 55
42 7[c] NbCl5 36
43 8 Nb(HC2O4)5.xH2O 40
44 9 LiNbO3 3
10 MgNb2O6 4
45
11 Nb2O5[d] 7
46 12 Nb4C5 5 Entry Nb(HC2O4)5.xH2O Reaction time Yield[b]
47 13 Nb3N5 4 (mol-%) [h] [%]
48 14 none 2
1 1% 10 75
49 [a] Reaction conditions: benzoic acid (1.0 mmol), phenethylamine 2 5% 10 80
50 (1.0 mmol), catalyst (5 mol-%), toluene (10 ml), and n-dodecane (1.0 mmol), 3 8% 10 85
51 reflux, 10 h. [b] Benzoic acid (1.0 mmol), NbCl5 (5 mol-%), n-dodecane 4 0.1% 18 43
(1.0 mmol) and toluene was stirred first under 25 oC for 1 h before the 5 0.5% 18 83
52 addition of Phenethylamine (1.0 mmol), then reflux for 10 h. [c] 6 1% 18 88
53 Phenethylamine (1.0 mmol), NbCl5 (5 mol-%), n-dodecane (1.0 mmol) and 7 1% 16 87
54 toluene was stirred first under 25 oC for 1 h before the addition of Benzoic 8 1% 24 90
55 acid (1.0 mmol), then reflux for 10 h. [d] Nb2O5 was purchased from Alfa
aesar company, as 100 mesh power with lot number of A28Z021, and was [a] Reaction conditions: benzoic acid (1.0 mmol), phenethylamine
56 used directly from bottle. (1.0 mmol), catalyst, toluene (10 ml), reflux. [b] Isolated yield
57

ChemistrySelect 2018, 3, 2599 – 2603 2600  2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Wiley VCH
1809 / 107472
Freitag, 02.03.2018
[S. 2600/2603] 1
Communications

1 Table 3. Amide formation reaction with nonaromatic amines catalysed by Table 4. Amide formation reaction with aniline catalysed by Niobium(V)
Niobium(V) oxalate hydrate.[a] oxalate hydrate.[a]
2
3
4
5
6
7
Entry R1 Acid/aniline Product Yield [%][b]
8 Entry R1 R2 Product Yield [%][b]
(mol/mol)
9 1 Ph Ph(CH2)2 3a 88(4)
1 Ph 1:1 3s 40
10 2 2-MeC6H4 Ph(CH2)2 3b 87(1)
2 Ph 1:2 3s 53
11 3 3-MeC6H4 Ph(CH2)2 3c 86(6)
3 Ph 1:3 3s 67
4 4-MeC6H4 Ph(CH2)2 3d 87(6)
12 4 Ph 1:4 3s 99(1)
5 4-CF3C6H4 Ph(CH2)2 3e 84(2)
13 5 Ph 1:5 3s 91
6 4-ClC6H4 Ph(CH2)2 3f 82(1)
6 4-MeC6H4 1:4 3t 68(2)
14 7 4-OMeC6H4 Ph(CH2)2 3g 80(4)
7 4-CF3C6H4 1:4 3u 98(1)
15 8 2-Furanyl Ph(CH2)2 3h 57(4)
8 Cy 1:4 3v 74(2)
9 2-Pyridyl Ph(CH2)2 3i 68(20)
16
10 CH3(CH2)4 Ph(CH2)2 3j 86(41) [a] Reaction conditions: carboxylic acid (2.0 mmol), aniline (2.0 mmol),
17 11 Cy Ph(CH2)2 3k 84(21) Nb(HC2O4)5.xH2O (1 mol-%), toluene (20 ml), reflux, 18 h. [b] Isolated yield.
18 12 t-Bu Ph(CH2)2 3l 78(1) The yields of control experiments (without Nb(HC2O4)5.xH2O) are in
19 13 1-Adamantyl Ph(CH2)2 3m 83(6) parentheses.
14 Ph PhCH2 3n 96(6)
20
15 4-MeC6H4 PhCH2 3o 90(5)
21 16 4-CF3C6H4 PhCH2 3p 83(5)
22 17 Cy PhCH2 3q 97(19)
18 Ph 2-Pyridylmethyl 3r 88(1) Then we tried to extend the amide formation reaction
23
catalyzed by niobium(V) oxalate hydrate to other amines, like
24 [a] Reaction conditions: carboxylic acid (2.0 mmol), amine (2.0 mmol),
aniline, with low nucleophilic reactivity (Table 4). Aniline
25 Nb(HC2O4)5.xH2O (1 mol-%), toluene (20 ml), reflux, 18 h. [b] Isolated yield.
The yields of control experiments (without Nb(HC2O4)5.xH2O) are in seemed to have poor reactivity under the same condition
26
parentheses. (entry 1). With increased amount, aniline could react with
27
benzoic acid to give corresponding amide smoothly, and 4
28
times amount of aniline gave the isolated yield of 99%. If used
29
in more than 4 times, like 5 times, aniline gave a decreased
30
smoothly to form corresponding amides. The position of yield of amide product (entries 2–5). Under this condition, para
31
methyl group, which has an electro-donating inductive effect, substituted benzoic acids could react with aniline to form
32
in phenyl moiety (ortho, meta, para to the carboxylic acid amide products, in which methyl group, an electro-donating
33
group) didn’t seem to have much influence to the yield inductive group, seemed to retard the reaction while trifluor-
34
(entries 2–4), while benzoic acid derivatives having a para omethyl group, an electro-withdrawing group, worked well
35
substituent, such as -CF3, -Cl, -OMe, having an electro-with- (entries 6,7). Cyclohexanecarboxylic acid reacted with aniline as
36
drawing inductive effect, gave a relatively lower yield (en- well to give corresponding amide 3 v in a yield of 74% (entry 8).
37
tries 5–7). Heteroaromatic carboxylic acids, 2-furanyl or 2- Here, in the amide formation reaction with excessive aniline, a
38
pyridyl carboxylic acid, afforded moderate yields of correspond- relatively weaker nucleophile than nonaromatic amines, the
39
ing amide products (entries 8–9). Aliphatic carboxylic acids acidity of carboxylic acids seem to have an important influence
40
worked well to give amide products, and the steric difference on the reaction.
41
between those carboxylic acids didn’t play a big role in this In the area of multivalent metal based catalytic amide
42
reaction (10-13). When reacted with benzylamine, aromatic or formation from non-activated carboxylic acids, there are four
43
aliphatic carboxylic acids worked well to give the amide typical systems which have been well developed, such as
44
products in high yields, except 4-trifluoromethylbenzoic acid ZrCl4,[8b,g] ZrCp2Cl2,[8a] Ti(Oi-Pr)4[8e] and Nb2O5.[12a] Aniline has been
45
(entries 14–17). Benzoic acid could react with 2-pyridylmethyl- found a problematic substrate in the catalytic amidation
46
amine to give corresponding amide product 3 r in an isolated reaction, in the case of ZrCl4, ZrCp2Cl2 and Ti(Oi-Pr)4, except
47
yield of 88%, which shows the existence of pyridine part in Nb2O5 which was prepared from calcination of niobic acid.
48
amine doesn’t retard the reaction as much as the being of Interestingly, niobium(V) oxalate hydrate didn’t show the
49
pyridine structure in carboxylic acid (entries 9, 18). reactivity in amidation with aniline, in the case of N-phenyl-
50
Under the same reaction conditions mentioned above, we benzamide, as effective as Nb2O5, and needed excessive
51
explored the reaction between benzoic acid and 3,5-dimethyl- amount of aniline to make sure the conversion of carboxylic
52
piperidine (mixture of cis and trans) in a 1:1 mole ratio, and the acids. For aromatic carboxylic acids, such as benzoic acid, ZrCl4
53
reaction gave no isolated amide product (less than 1% isolated and Ti(Oi-Pr)4 didn’t work as effectively as ZrCp2Cl2, while well
54
yield), which shows that the steric hindrance of amine may play prepared Nb2O5 could catalyze the amidation reaction of
55
an important role in this reaction. benzoic acid even with less reactive aniline, and niobium(V)
56
oxalate hydrate showed the high efficiency in many kinds of
57

ChemistrySelect 2018, 3, 2599 – 2603 2601  2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Wiley VCH
1809 / 107472
Freitag, 02.03.2018
[S. 2601/2603] 1
Communications
aromatic carboxylic acids as well as aliphatic acids in the amide amines, can’t attack Nb(V)-mixed carboxylic anhydride as easily
1
formation reactions. as other aliphatic amines, so aniline needs more than one
2
equivalent to complete the reaction. Since the ligand exchange
3
step of carboxylic acid and oxalic acid exists in an equilibrium
4 Explanation of mechanism
reaction, the structure of carboxylic acid doesn’t seem to be as
5
Furthermore, we investigated the reaction between oxalic acid important as the influence of nucleophilicity of amines. Oxalic
6
and phenethylamine in a 1:2 mole ratio under the same acid, which has more than one acid moiety, uses one
7
conditions mentioned above, and no expected amide product carboxylate to connect Nb(V) and leaves another carboxylate
8
was isolated (less than 1%), which means that oxalic acid is free to block amines to attack the Nb(V)-connected carboxylate,
9
inert under our reaction conditions. Therefore, oxalic acid so oxalic acid can survive in the reaction as a ligand but not a
10
doesn’t react well with amines added in the reaction and reactant.
11
doesn’t give notable corresponding amide product as by-
12
products or contaminants. We also noticed that the catalyst,
13 Conclusions
niobium(V) oxalate hydrate, was dissolvable in the reaction
14
system to form a clear solution. So, oxalic acid can work as a In summary, an easy-handled process of direct amide formation
15
ligand in this amide formation reaction, to disperse and reaction catalyzed by niobium(V) oxalate hydrate has been
16
stabilize niobium(V) in the reaction. developed to form secondary amides from carboxylic acids and
17
So a plausible mechanism is proposed (Scheme 2). The amines, during which water, the only by-product, can be
18
carboxylic acid might replace one of the oxalate in the niobium removed through the azeotropic distillation with toluene under
19
the boiling point of toluene. This reaction condition is workable
20
for variable carboxylic acids, aromatic or aliphatic, and primary
21
amines, aromatic or aliphatic. After all, the ligand assistant
22
metal-based catalytic amide formation has been rarely reported
23
according to the scope of our literature search. So, it is
24
noteworthy that oxalate works as a ligand to Nb(V) but not a
25
reactant in the amide formation reaction, because of its
26
inertness to the nucleophilic amine attack, which may afford a
27
new clue to how to design suitable ligands for the catalytic
28
amide formation reaction.
29
30
31 Supporting Information Summary
32
General: All chemicals were commercially available. 1H NMR
33
and 13C NMR spectra were recorded at 400 and 100 MHz,
34
respectively, with a Bruker Ascend 400 MHz NMR spectrometer
35 Scheme 2. A plausible mechanism for the amide formation catalysed by
Nb(HC2O4)5.xH2O.
in CDCl3 with TMS (tetramethylsilane) as an internal standard.
36
Mass spectra (EI) were performed with an Agilent GC-MS 7890B
37
5977 system and an Agilent 6110 Quadrupole LC/MS system.
38
Melting point was made by capillary melting point apparatus.
39
(V) oxalate, as a ligand exchange step, with the release of one Column chromatography was performed with silica gel (200-
40
oxalic acid to form the combination of carboxylate and 300 mesh), and toluene was used directly without distillation.
41
niobium(V), as an activated carboxylic acid like ester, acyl
42
chloride or anhydride. Then one amine attacks the carbonyl
43 Acknowledgements
group of this niobium(V) activated carboxylic acid to give
44
amide as product and (HC2O4)4NbOH. The latter reacts with We are grateful to the National Natural Science Foundation of
45
oxalic acid, as another ligand exchange step, to give H2O as China (No. 21202204) and Science Foundation of China University
46
another product and niobium(V) oxalate as the recovered of Petroleum, Beijing (No. YJRC-2011-13, No. JCXK-2011-04, No.
47
catalyst to continue the catalytic cycle. The amine might not 2462015YQ0601) for financial support of this research. Petrochina
48
directly attack the carbonyl group of non-activated carboxylic Pipeline Company is acknowledged as well.
49
acid, but when Nb(V) connects with carboxylate to form the
50
resulting salt, a Nb(V) involved mixed anhydride has been
51 Conflict of Interest
formed, in which the carbonyl group can be attacked by the
52
amine. The authors declare no conflict of interest.
53
The attack of amine to the Nb(V) activated carboxylic acid
54
plays an important role in the amide formation, so the Keywords: Amide synthesis · Niobium(V) oxalate hydrate ·
55
nucleophilicity of amines is crucial. As mentioned above, Catalysis · Carboxylic acids · Amines
56
aniline, with a relatively lower nucleophilicity than aliphatic
57

ChemistrySelect 2018, 3, 2599 – 2603 2602  2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Wiley VCH
1809 / 107472
Freitag, 02.03.2018
[S. 2602/2603] 1
Communications
[1] a) A. Ravve in Principles of Polymer Chemistry, 3rd ed., Springer Science + 47, 2876–2879; f) H. Charville, D. Jackson, G. Hodges, A. Whiting, Chem.
1 Business Media, LLC, 2012, pp. 430–446, 547–556; b) E. Valeur, M. Commun. 2010, 46, 1813–1823.
2 Bradley, Chem. Soc. Rev. 2009, 38, 606–631. [8] a) C. L. Allen, A. R. Chhatwal, J. M. J. Williams, Chem. Commun. 2012, 48,
3 [2] a) A. K. Ghose, V. N. Viswanadhan, J. J. Wendoloski, J. Comb. Chem. 1999, 666–668; b) H. Lundberg, F. Tinnis, H. Adolfsson, Chem. Eur. J. 2012, 18,
1, 55–68; b) C. A. G. N. Montalbetti, V. Falque, Tetrahedron 2005, 61, 3822–3826; c) F. Tinnis, H. Lundberg, H. Adolfsson, Adv. Synth. Catal.
4
10827–10852. 2012, 354, 2531–2536; d) A. C. Shekhar, A. R. Kumar, G. Sathaiah, V. L.
5 [3] A. El-Faham, F. Albericio, Chem. Rev. 2011, 111, 6557–6602. Paul, M. Sridhar, P. S. Rao, Tetrahedron Lett. 2009, 50, 7099–7101; e) H.
6 [4] a) R. M. de Figueiredo, J.-S. Suppo, J.-M. Campagne, Chem. Rev. 2016, Lundberg, F. Tinnis, H. Adolfsson, Synlett 2012, 23, 2201–2204; f) Y.
7 116, 12029–12122; b) V. R. Pattabiraman, J. W. Bode, Nature 2011, 480, Terada, N. Ieda, K. Komura, Y. Sugi, Synthesis 2008, 2318–2320; g) F.
471–479; c) B. Shen, D. M. Makley, J. N. Johnston, Nature 2010, 465, Tinnis, H. Lundberg, T. Kivijrvi, H. Adolfsson, Org. Synth. 2015, 92, 227–
8
1027–1033; d) C. Gunanathan, Y. Ben-David, D. Milstein, Science 2007, 236.
9 317, 790–792; e) C. L. Allen, J. M. J. Williams, Chem. Soc. Rev. 2011, 40, [9] R. J. Lundgren, M. Stradiotto in Ligand Design in Metal Chemistry:
10 3405–3415; f) S. D. Sarkar, A, Studer, Org. Lett. 2010, 12, 1992–1995; Reactivity and Catalysis (ed.: M. Stradiotto, R. J. Lundgren), John Wiley &
11 g) J. W. Bode, S. S. Sohn, J. Am. Chem. Soc. 2007, 129, 13798–13799; Sons Ltd, Chichester, UK, 2016, pp. 1–12
h) D. G. Gusev, ACS Catal. 2017, 7, 6656–6662; i) J.-S. Suppo, R. M. de [10] J. B. Hendrickson, J. Am. Chem. Soc. 1975, 97, 5784–5800.
12
Figueiredo, J.-M. Campagne, Org. Synth. 2015, 92, 296–308; j) V. Pappula, [11] a) D. J. C. Constable, P. J. Dunn, J. D. Hayler, G. R. Humphrey, J. L. Leazer,
13 C. Ravi, S. Samanta, S. Adimurthy, ChemistrySelect 2017, 2, 5887–5890; Jr., R. J. Linderman, K. Lorenz, J. Manley, B. A. Pearlman, A. Wells, A. Zaks,
14 k) T. B. Halima, J. K. Vandavasi, M. Shkoor, S. G. Newman, ACS Catal. 2017, T. Y. Zhang, Green Chem. 2007, 9, 411–420; b) J. R. Dunetz, J. Magano,
15 7, 2176–2180; l) A. Ojeda-Porras, D. Gamba-Snchez, J. Org. Chem. 2016, G. A. Weisenburger, Org. Process Res. Dev. 2016, 20, 140–177.
81, 11548–11555. [12] a) M. A. Ali, S. M. A. H. Siddiki, W. Onodera, K. Kon, K.-i. Shimizu,
16
[5] H. Lundberg, F. Tinnis, N. Selander, H. Adolfsson, Chem. Soc. Rev. 2014, ChemCatChem 2015, 7, 3555–3561; b) M. A. Ali, S. M. A. H. Siddiki, K.
17 43, 2714–2742. Kon, K.-i. Shimizu, ChemCatChem 2015, 7, 2705–2710; c) M. A. Ali,
18 [6] a) M. J. J. Litjens, A. J. J. Straathof, J. A. Jongejan, J. J. Heijnen, Tetrahe- S. M. A. H. Siddiki, K. Kon, J. Hasegawa, K.-i. Shimizu, Chem. Eur. J. 2014,
19 dron 1999, 55, 12411–12418; b) R. V. Ulijn, B. BaragaÇa, P. J. Halling, S. L. 20, 14256–14260; d) M. A. Ali, S. K. Moromi, A. S. Touchy, K.-i. Shimizu,
Flitsch, J. Am. Chem. Soc. 2002, 124, 10988–10989; c) M. A. P. J. Hacking, ChemCatChem 2016, 8, 891–894; e) R. Rodrigues, D. Mandelli, N. S.
20
H. Akkus, F. van Rantwijk, R. A. Sheldon, Biotechnol. Bioeng. 2000, 68, 84– Gonçalves, P. P. Pescarmona, W. A. Carvalho, J. Mol. Catal. A: Chem. 2016,
21 91. 422, 122–130.
22 [7] a) T. Marcelli, Angew. Chem. Int. Ed. 2010, 49, 6840–6843; b) K. Ishihara,
23 Tetrahedron 2009, 65, 1085–1109; c) K. Ishihara, S. Ohara, H. Yamamoto,
J. Org. Chem. 1996, 61, 4196–4197; d) P. Tang, Org. Synth. 2005, 81, 262– Submitted: January 23, 2018
24
272; e) R. M. Al-Zoubi, O. Marion, D. G. Hall, Angew. Chem. Int. Ed. 2008, Accepted: February 13, 2018
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57

ChemistrySelect 2018, 3, 2599 – 2603 2603  2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Wiley VCH
1809 / 107472
Freitag, 02.03.2018
[S. 2603/2603] 1

You might also like