You are on page 1of 8

Send Orders for Reprints to reprints@benthamscience.

net
65
Current Organic Synthesis, 2020, 17, 65-72

RESEARCH ARTICLE
ISSN: 1570-1794
eISSN: 1875-6271

Current
Organic Synthesis
Tetraamminecopper(II) Sulfate Monohydrate in Oxidative Azide-olefin The Journal for Current and In-depth Reviews on Organic Synthesis

Impact
Factor:
1.84

Cyclo-addition and Three-component Click Reaction


BENTHAM
SCIENCE

Jasmin Sultana1 and Diganta Sarma1’*

1
Department of Chemistry, Dibrugarh University, Dibrugarh, India

Abstract: Introduction: An effective Cu-complex, [Cu(NH3) 4SO 4 • H2O] was prepared conveniently from the
inexpensive and easily available starting reagents in a simple route.
Materials and Methods: Excellent reactivity of the catalyst was observed towards two competent click-
ARTICLE HISTORY
cycloadditions: (a) oxidative cycloaddition of azides with electron-poor olefins and (b) one-pot cycloaddition
Received: August 01, 2019 of alkynes with boronic acid and sodium azide under “click-appropriate” conditions.
Revised: November 29, 2019
Accepted: December 22, 2019 Results: No external oxidant, short reaction time, high product yield, wide substrate scope, and aqueous
solvent media make the azide-olefin cycloaddition approach a greener route in contrast to the reported
DOI:
10.2174/1570179417666191223152643 methods.
Conclusion: The newly developed mild, green, and rapid three-component strategy shows product diversity
with superb yields at room temperature by reducing the synthetic process time and using only 1 mol % of the
synthesized copper complex.

Keywords: Ammoniated Cu-complex, OAOC, boronic acid, aqueous media, regioselective, 1,2,3-triazole.

1. INTRODUCTION form an unstable cycloadduct, triazoline which will crush into


The 1,2,3-triazole motif is an affluent scaffold having various products as dictated by reaction conditions [16]. Conversion
incorporation into a vast number of biologically and pharma- of unstable triazoline intermediates into the stable aromatic triazole
derivatives can be achieved by two different paths: (a) eliminative
Current Organic Synthesis

ceutically active molecules [1]. Hydrogen bonding potential to


bioactive targets and high stability to metabolic degradations are the azide-olefin cycloaddition (EAOC), and (b) oxidative azide-olefin
properties that have disclosed it as an attractive and efficient cycloaddition (OAOC), in which elimination and oxidation occur,
respectively, next to the cycloaddition step (Scheme 1).
connection in biological systems. Evolution of “click reactions” [2]
triggered the application of ubiquitous triazole chemistry to several In the area of OAOC approach, a few copper-mediated catalytic
other scientific arena [3], such as material science [4], systems, such as CuI (DIPEA, 1,4-dioxane, O2-balloon, 85 °C)
bioconjugation [5], and polymer chemistry [6]. [17], CuO-nanoparticles (DMF, 90 °C) [18], Cu(OTf)2 (DMF/
Primarily, conventional Huisgen 1,3-dipolar azide-alkyne AcOH, 110 °C) [19], Cu(OAC)2 (DMF, 85 °C) [20] have been re-
cycloaddition was used to access 1,2,3-triazoles as 100% atom ported representing the synthesis of 1,4-disubstituted and 1,4,5-
economical and most straight forward synthetic route [7]. The trisubstituted 1,2,3-triazoles. Nevertheless, some limitations, such
Huisgen method was further replaced by copper-catalyzed azide- as the need for external oxidant, base, long reaction time, harmful
alkyne cycloaddition (CuAAC) and ruthenium-catalyzed azide- organic solvent, etc. are associated with these methods. Therefore,
alkyne cycloaddition (RuAAC) as it was unable to overcome the at this juncture, we have made efforts to develop an efficient,
stumbling blocks, such as high activation energy requirement and economical, and environmentally friendly method by employing
poor regioselectivity [8, 9]. However, CuAAC and RuAAC easily prepared tetraammine copper (II) sulfate complex for OAOC.
methods were found to be selective towards 1,4- and 1,5- Although the organic azides needed for the synthesis of 1,2,3-
disubstituted 1,2,3-triazoles, respectively. In continuation, fully triazoles are generally stable, their handling is not safe because of
substituted 1,2,3-triazoles have also been synthesized by various their deleterious nature. Isolation and purification of polyazides and
synthetic approaches, such as cycloaddition of organic azides with low molecular weight organic azides are also problematic in nature.
internal alkynes [10] and multicomponent reactions of azides with Moreover, the synthetic procedures for aryl azides from their
keto-esters, activated enones, aldehydes, or nitriles [11]. corresponding amine precursors are also not straightforward.
Therefore, the processes involving in situ generated organic azides
As an alternative to AAC strategy, a few methods have been
are highly desirable. The in situ prepared azides react with alkynes,
reported representing the use of organic azides and olefins as
called three-component azide-alkyne cycloaddition and this
substrates such as: (a) dipolar cycloaddition with alkenes bearing a
approach is advantageous in different ways: (a) avoids the waste
leaving group [12], (b) organocatalytic oxidative 1,3-dipolar
creation of the additional steps, (b) avoids the time consumption,
cycloaddition with alkenes [13], (c) Cu-mediated cycloadditions
and (c) minimizes hazards.
with olefins bearing electron-withdrawing functional groups [14].
Pioneer study on this field of research was implemented by L’abbe Use of anilines [21] and aryl halides (bromides and iodides)
and Huisgen [15] with organic azides and electron-poor olefins to [22] as the synthetic precursor of aryl azides have found in the
literature for one-pot synthesis of substituted 1,2,3-triazoles.
However, due to the limited scope of these reactions, we turned our
*
Address correspondence to this author at the Department of Chemistry, Dibrugarh
concentration to the three-component synthesis of 1,4-disubstituted
University, Dibrugarh-786004, Assam, India; E-mails: dsarma22@gmail.com; 1,2,3-triazoles using widely available and less toxic boronic acids
dsarma22@dibruac.in as the precursors of aryl azides. A number of such one-pot three-

1875-6271/20 $65.00+.00 © 2020 Bentham Science Publishers


66 Current Organic Synthesis, 2020, Vol. 17, No. 1 Sultana and Sarma

N3

N
N Elimination [-HX] N Oxidation [-H2] N N
N N N N
EAOC OAOC X
X

Triazoline

= Ar, Bn = COOR, COR, CN etc. X = NO2, OR, OAc etc.

Scheme 1. Eliminative and oxidative azide-olefin cycloaddition.

component methods have been reported by different groups by 2. MATERIALS AND METHODS
using various copper-derivatives. These catalytic systems include
The reagents and solvents were purchased from different
copper supported on Al2O3 [23], CuSO4 supported on chitosan [24],
copper (I)-phosphinite complex [25], Cu(OAc)2 in the presence of commercial sources and used without further purification unless
piperidine [26], Cu(II)–β-cyclodextrin complex (as a nanocatalyst) otherwise is noted. The tetraammine copper (II) complex used in
[27], glutamate-copper catalyst supported on NiFe2O4 [28], Cu(II)- the processes is synthesized by a reported method [33]. The organic
β-cyclodextrin complex supported on Fe3O4 nanoparticle [29], etc. azides used in oxidative azide-olefin cycloaddition reaction were
Most of these methods involve lengthy and laborious methods for also prepared by a previous reported method [34]. Purifications of
the preparation of active Cu-catalysts, long reaction time, and products were carried out by column chromatography using silica
elevated temperature. gel (200-300 mesh). Analytical thin-layer chromatography was
With the increasing command for environmentally favorable performed using silica gel 60F254 plates and visualized with UV
methods, greener conditions for organic synthesis have gained light. 1H NMR (500 MHz), 13C NMR (125 MHz) were recorded on
much attention in the present years. Recently, we developed a bio- a Bruker Advance 500 MHz spectrometer using TMS as an internal
surfactant (sodium deoxycholate)-accelerated one-pot three- standard. For some products, 1H NMR (600 MHz) and 13C NMR
component method for triazole synthesis using Cu(OAc)2 as a (150 MHz) were recorded on a Bruker Advance 600 MHz
catalyst [30]. As an extension of this work, we developed this spectrometer. Chemical shifts are reported in parts per million
method using tetraammine copper sulphate (II) complex (ppm, δ) downfield from residual solvent peaks and coupling
[Cu(NH3)4SO4 • H2O]. constants are reported as Hertz (Hz). Splitting patterns are
This known complex, ammoniated cupric sulfate monohydrate designated as: s = singlet, d = doublet, t = triplet, q = quartet, m =
was prepared by treating concentrated ammonia with saturated multiplet, etc. Melting points were determined on a Büchi B-540
aqueous CuSO4 solution followed by precipitating the complex melting point apparatus.
with ethanol (Fig. 1) [31]. X-ray crystallographic analysis
confirmed the presence of [Cu(NH3)4H2O]2+ unit in the solid-state
of this complex, where the central Cu(II)-ion is bound covalently by 3. EXPERIMENTAL
four ammonia molecules with an average distance of 2.05Å in a 3.1. General Procedure for Oxidative Azide-olefin
near square planar arrangement [32]. Copper ions are linked like a Cycloaddition
chain through the oxygen atoms and these ions and chains are
clutched together via a network of hydrogen bonding. To a mixture of [Cu(NH3)4SO4 • H2O] (10 mol %) and β-
cyclodextrin (5 mol %) in 2 ml of DMSO:H2O (1:1), 1.0 equiv. of
organic azide (0.377 mmol) and 1.5 equiv. of olefin (0.565 mmol)
Concentrated
NH3 were added and allowed to stir at 85 °C for 4 h. After completion of
the reaction, it was extracted with ethyl acetate (2x20 mL), washed
with deionized water, dried over anhydrous sodium sulfate, and
concentrated under vacuum, which on column chromatography
(ethyl acetate:hexanes) afforded the desired product.
CuSO4 in H2O
EtOH (Cooled in
an ice bath) 3.2. General Procedure for One-pot Three-component Cyclo-
addition
To a mixture of Cu-complex (1 mol %), phenylboronic acid
(0.369 mmol), sodium azide (1.107 mmol), phenyl acetylene (0.369
mmol), and K3PO4 (0.738 mmol), 2 mL of DMSO: H2O (1:10)
mixture was added and stirred at room temperature for 1 h. After
[Cu(NH3)4 H2O]SO4 Filtration of Dark purple- completion of the reaction, it was extracted with ethyl acetate (2×10
complex the precipitate blue precipitate mL), washed with brine, dried over anhydrous sodium sulfate, and
concentrated under vacuum, which on column chromatography
Fig. (1). Sequential preparation of ammoniated cupric sulfate monohydrate.
(ethyl acetate: hexanes) afforded the desired product. The products
were then characterized by NMR spectroscopy.
Tetraamminecopper(II) Sulfate Monohydrate in Oxidative Azide-olefin Cycloaddition Current Organic Synthesis, 2020, Vol. 17, No. 1 67

Table 1. Optimization of reaction conditions for OAOC a .

N
N3 O Ph N N
Cu-source, additive
+
O solvent, air, heat, 4 hr OMe
O
1a 2a 3a

Entry Catalyst Additive Solvent Yield of 3a (%)b

1 Cu-complex β-cyclodextrin Toluene 51

2 Cu-complex β-cyclodextrin H2O 65

3 Cu-complex β-cyclodextrin DMSO 70

4 Cu-complex β-cyclodextrin DMSO:H2O (1:2) 80

5 Cu-complex β-cyclodextrin DMSO:H2O (1:1) 91

6 Cu-complex CTAB DMSO:H2O (1:1) 84

7 Cu-complex TBAB DMSO:H2O (1:1) 70

8 Cu-complex CPC DMSO:H2O (1:1) 68

9 CuI β-cyclodextrin DMSO:H2O (1:1) 20

10 CuCN β-cyclodextrin DMSO:H2O (1:1) trace

11 CuBr β-cyclodextrin DMSO:H2O (1:1) trace

12 Cu(OAc)2 β-cyclodextrin DMSO:H2O (1:1) 80

13 Cu-complex β-cyclodextrin DMSO:H2O (1:1) 81c, 0d

14 Cu-complex β-cyclodextrin DMSO:H2O (1:1) 85e, 70f

15 Cu-complex β-cyclodextrin DMSO:H2O (1:1) 81g, 90h


[a]
Reaction conditions: 1a (1.0 equiv., 0.377 mmol), 2a (1.5 equiv., 0.565 mmol), catalyst (10 mol %), and additive (5 mol %) were heated for 4 hrs in 2 ml solvent. [b] Product yields.
[c]
Yield of 3a using 5 mol % of Cu-complex. [d] 0 mol % of Cu-complex was employed. [e] Yield of 3a using 2 mol % of β-cyclodextrin. [f ] No β-cyclodextrin was employed.
[g]
Reaction was carried out at 70 0C. [h] Reaction was carried out at 100 0C.

4. RESULTS AND DISCUSSION in toluene. The prepared Cu-complex in the presence of the phase
Facile reaction kinetics and biocompatibility of the CuAAC transfer catalyst, β-cyclodextrin afforded 51% isolated yield of the
process have made it widely acceptable as a novel and robust way respective 1,4-disubstituted triazole molecule at 85 °C (Table 1
for functionalizing biomolecules. However, the problem of entry 1). Satisfied with that outcome, we optimized the reaction in
cytotoxicity of active copper species has been a tenacious limitation terms of different solvent systems, additives, and Cu-species. The
to carry out these Cu-catalyzed reactions [35]. Therefore, in order solvents, water, and DMSO both were observed to be suitable for
to get out of this problem, Gary R. Abel and coworkers used the our model reaction (Table 1 entries 2 and 3). We decided to
qPCR (quantitative polymerase chain reaction) method and studied perform the reaction in DMSO-H2O mixtures. With the increase in
the oxidative damage to a DNA strand under varying conditions the amount of DMSO in the DMSO-H2O mixtures, the reaction
used for CuAAC reactions [36]. They discovered that the use of yield was observed to increase and the best yield was detected by
DMSO in water (10% mixture) can reduce the extent of oxidative using 1:1 mixture of DMSO in water (Table 1 entries 4 and 5).
damage by two orders of magnitude. Low toxicity and the tendency Fixing DMSO-H2O mixture (1:1) as the best solvent system the
to shield the living cells from the damage mediated by reactive reaction was optimized with respect to some different additives
oxygen species may establish the solvent, DMSO in water to be a (phase transfer catalysts): CTAB (cetyl trimethylammonium
feasible way to protect cells during in vitro and in vivo CuAAC bromide), TBAB (tetra-n-butylammonium bromide), and CPC
bioconjugations. Motivated by this report, here we are trying to use (cetylpyridinium chloride) (Table 1 entries 6- 8). The additive, β-
the DMSO-H2O mixtures as a reliable and useful solvent media for cyclodextrin was found to be superior to the other additives as they
(a) oxidative azide-olefin cycloaddition reaction and (b) one-pot furnished a comparatively higher yield of 3a. Next, the reaction was
three-component method for synthesis of 1,4-disubstituted and analyzed with regard to different copper-salts, such as CuI, CuCN,
1,4,5-trisubstituted 1,2,3-triazoles. CuBr, and Cu(OAc)2 (Table 1 entries 9-12). Very small amounts of
the products were noticed in the presence of CuI, CuCN, and CuBr.
As a practical alternative to alkynes, olefins were chosen to When Cu(OAc)2 was chosen as the catalyst, the yield was increased
access these significant molecules because they are less expensive to 80 %. Thus, among the catalysts tested for the reaction, the
and easy to synthesize. At the outset, the reaction was analyzed by synthesized Cu-complex was selected as the best catalyst as it gave
using phenyl azide and methyl acrylate as a representative substrate the best yield of the product (90 %) under the same reaction
68 Current Organic Synthesis, 2020, Vol. 17, No. 1 Sultana and Sarma

Table 2. Substrate scope of azides with terminal and internal olefinsa, b.

R N
N N
R N3 + X
Cu-complex (10 mol %) X
β-cyclodextrin (5 mol %)
O
DMSO:H2O (1:1) N
R
85 °C, 4hr N N
R N3 +
O
R= Ar, Bn X=COOR, CN
1 2 3

N N OBun
N N OMe N N OEt
N
N N O
O O

3c, 94%
3a, 91% 3b, 93%

N N
N N OMe N N OEt
CN
N
N N
O O

Cl Cl
3d, 79%
3e, 92% 3f, 86%

N OEt N N OEt N N OMe


N
N N N
O O O

F H3C H3C
3g, 82% 3h, 86% 3i, 81%

N N O O2N N O2N
N O N N
N OMe
N N
O
H3C
3k, 84% 3l, 85%
3j, 83%
O2N N
N OBun
- N -
O
3m, 75%
[a]
Reaction conditions: Azide (1 equiv), olefin (1.5 equiv), Cu-complex (10 mol %), and the additive β-cyclodextrin (5 mol %) were heated at 85 0C in DMSO/H20 mixture (2 ml)
under atmospheric condition. [b] Isolated yields of products.

conditions. In addition, catalyst loading also played an important functional groups (chloride and fluoride) at para-positions of the
role in the efficiency of the reaction (Table 1 entry 13). Dropping phenyl azides were also found to be effective to provide the
the quantity of the catalyst from 10 mol % to 5 mol % showed a respective ester-substituted triazole derivatives (Table 2, 3e-g). The
decrease in the yield of the product to 81%. Apparently, no product electron-donating methyl group at the para-position of phenyl azide
was detected in the absence of a catalyst (Table 1 entry 13). A was also observed to be tolerated under the standard reaction
pronounced temperature effect was also noticed (Table 1 entry 17). conditions and good yields of desired triazole products (3h and 3i)
Reducing the temperature to 70 °C afforded a lower yield of were obtained. The cyclic enone, 2-cyclohexen-1-one was also
product (81%), while increasing the temperature to 100 °C showed effectively clicked with 4-Me-phenyl azide and 4-NO2-benzyl azide
no improvement in the yield of the reaction (90%). (3j and 3k). The reactions of nitro-substituted benzyl azide with
After optimizing the reaction conditions, different olefins methyl acrylate (3l) and butyl acrylate (3m) also went well with
(terminal and internal) were screened with both phenyl azides and respective yields of 85% and 75%.
benzyl azides to infer the efficiency of this OAOC methodology. In Next, to demonstrate the scope and applicability of our catalytic
this regard, the reaction was first analyzed by treating aryl azides system, we applied it to a one-pot three-component strategy. The
with an assortment of electron-poor terminal olefins. Cycloaddition phenyl azide and phenylacetylene were allowed to react in the
of phenyl azide with methyl acrylate, ethyl acrylate, and butyl presence of the copper complex, tetra amine copper (II) sulfate in
acrylate provided a very high yield of the corresponding ester the DMSO-H2O mixture (1:10) at room temperature (Table 3, entry
substituted triazole products (Table 2, 3a-c). The dipolar 1). Formation of 75% isolated yield of the desired 1,4-disubstituted
cycloaddition with acrylonitrile afforded a comparative lower yield triazole product after one hour using only 1 mol % of the Cu-
of the product (Table 2, 3d). The presence of electron-withdrawing complex inspired us to study the reaction further. To increase the
Tetraamminecopper(II) Sulfate Monohydrate in Oxidative Azide-olefin Cycloaddition Current Organic Synthesis, 2020, Vol. 17, No. 1 69

Table 3. Optimization of reaction conditionsa.

NaN3 N
OH Cu-source, solvent Ph N N
B +
OH Base, rt, air Ph
4a 5a 6a

Entry Catalyst Base Solvent Yield (%)b

1 Cu-complex - DMSO:H2O (1:10) 75

2 Cu-complex K2CO3 DMSO:H2O (1:10) 92

3 Cu-complex Cs2 CO3 DMSO:H2O (1:10) 86

4 Cu-complex CaCO3 DMSO:H2O (1:10) 86

5 Cu-complex K3PO4 DMSO:H2O (1:10) 96

6 Cu-complex DIPEA DMSO:H2O (1:10) 75

7 Cu-complex DBU DMSO:H2O (1:10) 93

8 Cu-complex Guanidinehydrochloride DMSO:H2O (1:10) 51

9 Cu-complex K3PO4 DMSO:H2O (1:5) 80

10 Cu-complex K3PO4 DMSO:H2O (1:1) 71

11 Cu-complex K3PO4 DMSO 52

12 Cu-complex K3PO4 H2O 96

13 Cu-complex K3PO4 CH3OH 43

14 Cu-complex K3PO4 DMF 25

15 Cu(NO3)2 K3PO4 DMSO:H2O (1:10) 82

16 CuBr K3PO4 DMSO:H2O (1:10) 55

17 CuCN K3PO4 DMSO:H2O (1:10) 67

18 CuI K3PO4 DMSO:H2O (1:10) 69

19 Cu-complex K3PO4 DMSO:H2O (1:10) 73c, 0d


[a]
Reaction conditions: 4a (1.0 equiv., 0.369 mmol), 5a (1.0 equiv., 0.369 mmol), NaN3 (3.0 equiv., 1.107 mmol), catalyst (1 mol %), and base (2.0 equiv.) were allowed to react at
rt. [b] Product yields. [c] Yield of 6a employing 0.5 mol % of the Cu-complex. [d] No Cu-source.

efficiency of the catalyst, different bases were added to the reaction (both aliphatic and aromatic) and sodium azide leading to 81-96%
mixture and analyzed the reaction after one hour (Table 3, entries 2- isolated yields of products under the optimized conditions.
8). The highest yield (96%) was observed by employing potassium
As displayed in Table 4, the reactions proceeded swiftly in the
phosphate (K3PO4) as an additive (Table 3, entry 5). To standardize
presence of a series of structurally diverse phenyl boronic acids by
the condition of the reaction a sequence of experiments were
furnishing very high yields of products. The aliphatic terminal
carried out by varying the copper source, catalyst loading, and
alkynes were treated successfully with aryl boronic acids (Table 4,
solvent system. Screening of different solvent systems showed that
entries 6b, 6c, and 6e). The internal alkyne, DMAD provided 93%
the reactants did not react well in the presence of DMSO, CH3OH,
yield of the desired fully substituted triazole at room temperature
and DMF (Table 3, entries 11, 13, and 14). 96% yield of the
(Table 4, entry 6d). Satisfactory yields of products were observed
product was observed by employing the polar, easily available,
in the presence of both electron-donating and electron-withdrawing
harmless, and economical solvent water under the same reaction
functional groups on the aromatic ring of phenyl boronic acid and
conditions (Table 3, entry 12). However, from solubility perception,
no difference in yields was observed regarding the substitution
the DMSO-H2O mixtures were selected as the best solvent system
pattern and electronic nature of the aryl moiety. Finally, testing
for this CuAAC reaction. As increasing the DMSO content in the
heteroaryl boronic acids as a coupling partner with aromatic alkyne
mixture showed a decrease in the product yield, further experiments
under the above-optimized click cyclization reaction gave the
were carried out in 1:10 mixture of this solvent. Analysis of a few
desirable outcome (Table 4, entries 6m, 6n).
more Cu-sources, such as Cu(NO3)2, CuBr, CuCN, and CuI
provided the respective triazoles with 82%, 45%, 67%, and 69%
yields, respectively (Table 3, entry 15-18). Again, decreasing the CONCLUSION
amount of catalyst from 1 mol % to 0.5 mol % resulted in a In conclusion, the efficiency of the homogeneous, robust, and
decrease in reaction rate (Table 3, entry 19). air-stable copper amine complex has been explored in the field of
triazole synthesis. Relying on this Cu-complex, two adequate
The screened conditions were then applied to the one-pot three- catalytic systems have been developed for dipolar cycloaddition of
component strategy by varying the substrates (boronic acid and electron-deficient olefins with organic azides and alkynes with in
alkyne). Different phenyl boronic acids (both substituted and situ generated azides. The Cu-catalyzed azide-olefin cycloaddition
unsubstituted) were successfully treated with functionalized alkynes reaction has not been studied so far and our method can be a
70 Current Organic Synthesis, 2020, Vol. 17, No. 1 Sultana and Sarma

Table 4. Substrate scope of three-component reactiona .

OH Cu-complex (1 mol %) N N
B + R' N
R OH K3PO4, DMSO:H2O (1:10) R'
R
NaN3
4 5 6

N N N
N N N
O
N N N O
Ph

6a, 96% 6b, 95% 6c, 85%


N O OEt
N N N N N
N N N
O O
O
O H3C H3CO
6d, 93% 6e, 81% 6f, 95%

N N N N N N
H3C N N N

NC F
6g, 90% 6h, 88% 6i, 87%

N N N N N N
N Br N Cl N

Cl
6j, 92% 6k, 96% 6l, 91%

N N N N
S O N
N -

6m, 81 % 6n, 86 %
[a]
Reaction conditions: 4 (1.0 equiv., 0.369 mmol), 5 (1.0 equiv., 0.369 mmol), NaN3 (3.0 equiv., 1.107 mmol), Cu-complex (1 mol %), and K3PO4 (2.0 equiv.) were allowed to react
in DMSO/H2O (1:10) at rt. [b] Isolated yields.

potential substitute for the existing methods because of low reaction CONFLICT OF INTEREST
time and green solvent media. The need for no external oxidant for The authors declare no conflict of interest, financial or
oxidative dehydrogenation step and formation of “click-friendly” otherwise.
products make this strategy green. To the best of our knowledge, it
is the first ligated homogeneous copper complex-mediated system
for the OAOC approach. The novel one-pot cascade synthetic ACKNOWLEDGEMENTS
strategy developed by employing low catalyst loading was also Declared none.
efficiently applied for the regio-controlled synthesis of a variety of
triazole products. We hope that the versatility and green credentials
SUPPLEMENTARY MATERIAL
of these protocols will make them more attractive to organic
synthesis. Supplementary material is available on the publisher’s website
along with the published article.
CONSENT FOR PUBLICATION
Not applicable. REFERENCES
[1] (a) Kharb, R.; Sharma, P.C.; Yar, M.S. Pharmacological significance of
triazole scaffold. J. Enzyme Inhib. Med. Chem., 2011, 26(1), 1-21.
AVAILABILITY OF DATA AND MATERIALS Available at http://dx.doi.org/10.3109/14756360903524304.
[PMID: 20583859]
The authors confirm that the data supporting the findings of this (b) Bohacek, R.S.; McMartin, C.; Guida, W.C. The art and practice of
study are available within the article. structure-based drug design: A molecular modeling perspective. Med. Res.
Rev., 1996, 16(1), 3-50.
Available at http://dx.doi.org/10.1002/(SICI)1098-1128(199601)16:1<3::
FUNDING AID-MED1>3.0.CO;2-6.
[PMID: 8788213]
D.S. is thankful to DST, New Delhi, India for a research grant (c) Meldal, M.; Tornøe, C.W. Cu-catalyzed azide-alkyne cycloaddition.
[No. EMR/2016/002345]. The authors also acknowledge the Chem. Rev., 2008, 108(8), 2952-3015.
Department of Science and Technology for financial assistance Available at http://dx.doi.org/10.1021/cr0783479.
[PMID: 18698735]
under DST-FIST programme and UGC, New Delhi for Special (d) Angell, Y.L.; Burgess, K. Peptidomimetics via copper-catalyzed azide-
Assistance Programme (UGC-SAP) to the Department of alkyne cycloadditions. Chem. Soc. Rev., 2007, 36(10), 1674-1689.
Chemistry, Dibrugarh University. Available at http://dx.doi.org/10.1039/b701444a.
Tetraamminecopper(II) Sulfate Monohydrate in Oxidative Azide-olefin Cycloaddition Current Organic Synthesis, 2020, Vol. 17, No. 1 71

[PMID: 17721589] (c) Majireck, M.M.; Weinreb, S.M.J. A study of the scope and
[2] Kolb, H.C.; Finnand, M.G.; Sharpless, K.B. Click chemistry: Diverse regioselectivity of the ruthenium-catalyzed [3 + 2]-cycloaddition of azides
chemical function from a few good reactions. Angew. Chem. Int. Ed., 2001, with internal alkynes. J. Org. Chem., 2006, 71(22), 8680-8683.
40, 2004-2021. Available at http://dx.doi.org/10.1021/jo061688m. [PMID: 17064059]
Available at http://dx.doi.org/10.1002/1521-3773(20010601)40:11<2004:: (d) Drez-Gonzalez, S.; Stevens, D.D.; Nolan, S.P. Chem. Commun. (Camb.),
AID-ANIE2004>3.0.CO;2-5 2008, 4747.
[3] (a) Kolb, H.C.; Sharpless, K.B. The growing impact of click chemistry on Available at http://dx.doi.org/10.1039/b806806b.
drug discovery. Drug Discov. Today, 2003, 8(24), 1128-1137. (e) Zhang, H.; Tanimoto, H.; Morimoto, T.; Nishiyama, Y.; Kakiuchi, K.
Available at http://dx.doi.org/10.1016/S1359-6446(03)02933-7. Regioselective rapid synthesis of fully substituted 1,2,3-triazoles mediated by
[PMID: 14678739] propargyl cations. Org. Lett., 2013, 15(20), 5222-5225.
(b) Whiting, M.; Tripp, J.C.; Lin, Y-C.; Lindstrom, W.; Olson, A.J.; Elder, Available at http://dx.doi.org/10.1021/ol402387w. [PMID: 24087927]
J.H.; Sharpless, K.B.; Fokin, V.V. Rapid discovery and structure-activity [11] (a) Ramachary, D.B.; Ramakumar, K.; Narayana, V.V. Amino acid-catalyzed
profiling of novel inhibitors of human immunodeficiency virus type 1 cascade [3+2]-cycloaddition/hydrolysis reactions based on the push-pull
protease enabled by the copper(I)-catalyzed synthesis of 1,2,3-triazoles and dienamine platform: synthesis of highly functionalized NH-1,2,3-triazoles.
their further functionalization. J. Med. Chem., 2006, 49(26), 7697-7710. Chemistry, 2008, 14(30), 9143-9147.
Available at http://dx.doi.org/10.1021/jm060754. Available at http://dx.doi.org/10.1002/chem.200801325. [PMID: 18767077]
[PMID: 17181152] (b) Danence, L.J.T.; Gao, Y.; Li, M.; Huang, Y.; Wang, J. Organocatalytic
[4] Hawker, C.J.; Fokin, V.V.; Finn, M.G.; Sharpless, K.B. Bringing efficiency enamide-azide cycloaddition reactions: Regiospecific synthesis of 1,4,5-
to materials synthesis: The philosophy of click chemistry. Aust. J. Chem., trisubstituted-1,2,3-triazoles. Chemistry, 2011, 17(13), 3584-3587.
2007, 60, 381. Available at http://dx.doi.org/10.1002/chem.201002775. [PMID: 21341323]
Available at http://dx.doi.org/10.1071/CH07107. (c) Wang, L.; Peng, S.; Danence, L.J.T.; Gao, Y.; Wang, J. Amine-catalyzed
[5] (a) Link, A.J.; Tirrell, D.A. Cell surface labeling of Escherichia coli via [3+2] Huisgen cycloaddition strategy for the efficient assembly of highly
copper(I)-catalyzed [3+2] cycloaddition. J. Am. Chem. Soc., 2003, 125(37), substituted 1,2,3-triazoles. Chemistry, 2012, 18(19), 6088-6093.
11164-11165. Available at http://dx.doi.org/10.1002/chem.201103393. [PMID: 22461307]
Available at http://dx.doi.org/10.1021/ja036765z. (d) Seus, N.; Goncalves, L.C.; Deobald, A.M.; Savegnago, L.; Alves, D.;
[PMID: 16220915] Paixao, M.W. Synthesis of arylselanyl-1H-1,2,3-triazole-4-carboxylates by
(b) Wang, Q.; Chan, T.R.; Hilgraf, R.; Fokin, V.V.; Sharpless, K.B.; Finn, organocatalytic cycloaddition of azidophenyl arylselenides with β-keto-
M.G. Bioconjugation by copper(I)-catalyzed azide-alkyne [3 + 2] esters. Tetrahedron, 2012, 68, 10456.
cycloaddition. J. Am. Chem. Soc. 2003, 125, 3192. Available at http://dx.doi.org/10.1016/j.tet.2012.10.007.
c) Lutz, J.-F.; Zarafshani, Z. Efficient construction of therapeutics, (e) Li, Z.L.; Xie, Y.; Zhou, W. Professor Wang Ju-yi’s experience on clinical
bioconjugates, biomaterials and bioactive surfaces using azide-alkyne "click" application of Siguan points. Zhongguo Zhenjiu, 2013, 33(3), 255-257.
chemistry. Adv. Drug Deliv. Rev., 2008, 60, 958. [PMID: 23713316]
[6] (a) Evans, R.A. The rise of azide–alkyne 1,3-dipolar ‘click’ cycloaddition (f) Ramachary, D.B.; Shashank, A.B. Organocatalytic triazole formation,
and its application to polymer science and surface modification. Aust. J. followed by oxidative aromatization: Regioselective metal-free synthesis of
Chem., 2007, 60, 384. benzotriazoles. Chemistry, 2013, 19(39), 13175-13181.
Available at http://dx.doi.org/10.1071/CH06457. Available at http://dx.doi.org/10.1002/chem.201301412. [PMID: 24038664]
(b) Johnson, J.A.; Koberstein, J.T.; Finn, M.G.; Turro, N.J. Construction of
(g) Ramachary, D.B.; Shashank, A.B.; Karthik, S. An organocatalytic azide–
linear polymers, dendrimers, networks, and other polymeric architectures by
aldehyde [3+2] cycloaddition: Highyielding regioselective synthesis of
coppercatalyzed azidealkyne cycloaddition “click” chemistry. Macromol.
1,4disubstituted 1,2,3triazoles. Angew. Chem. Int. Ed., 2014, 53, 10420.
Rapid Commun., 2008, 29, 1052.
Available at http://dx.doi.org/10.1002/anie.20140672.
Available at http://dx.doi.org/10.1002/marc.200800208.
(h) Cheng, G.; Zeng, X.; Shen, J.; Wang, X.; Cui, X. A metalfree
[7] (a) Huisgen, R. Angew. 1,3dipolar cycloadditions. past and future. Chem.
multicomponent cascade reaction for the regiospecific synthesis of
Int. Ed. Engl., 1963, 2, 565-598.
1,5disubstituted 1,2,3triazoles. Angew. Chem. Int. Ed., 2013, 52, 13265.
Available at http://dx.doi.org/10.1002/anie.196305651.
Available at http://dx.doi.org/10.1002/anie.201307499.
(b) Huisgen, R. Kinetics and mechanism of 1,3dipolar cycloadditions.
Angew. Chem. Int. Ed. Engl., 1963, 2, 633-645. (i) Tian, L.; Hu, X-Q.; Li, Y.H.; Xu, P-F. Organocatalytic asymmetric
multicomponent cascade reaction via 1,3-proton shift and [3+2]
Available at http://dx.doi.org/10.1002/anie.196306331.
cycloaddition: an efficient strategy for the synthesis of oxindole derivatives.
(c) Huisgen, R. 1.3 dipolar cycloadditions review and outlook. Angew.
Chem. Commun. (Camb.), 2013, 49(65), 7213-7215.
Chem., 1963, 75, 604-637.
Available at http://dx.doi.org/10.1002/ange.19630751304. Available at http://dx.doi.org/10.1039/c3cc43755h. PMID: 23838686
[12] Amantini, D.; Fringuelli, F.; Piermatti, O.; Pizzo, F.; Zunino, E.; Vaccaro, L.
(d) Huisgen, R. Kinetics and mechanism of 1.3  dipolar cycloadditions.
Angew. Chem., 1963, 75, 742-754. Synthesis of 4-aryl-1H-1,2,3-triazoles through TBAF-catalyzed [3 + 2]
cycloaddition of 2-aryl-1-nitroethenes with TMSN3 under solvent-free
Available at http://dx.doi.org/10.1002/ange.19630751603.
[8] (a) Tornøe, C.W.; Christensen, C.; Meldal, M. Peptidotriazoles on solid conditions. J. Org. Chem., 2005, 70(16), 6526-6529.
phase: [1,2,3]-triazoles by regiospecific copper(i)-catalyzed 1,3-dipolar Available at http://dx.doi.org/10.1021/jo0507845. [PMID: 16050724]
cycloadditions of terminal alkynes to azides. J. Org. Chem., 2002, 67(9), [13] (a) Li, W.; Wang, J. Lewis base catalyzed aerobic oxidative intermolecular
3057-3064. azide–zwitterion cycloaddition. Angew. Chem. Int. Ed., 2014, 53, 14186-
Available at http://dx.doi.org/10.1021/jo011148j. 14190.
[PMID: 11975567] Available at http://dx.doi.org/10.1002/anie.201408265.
(b) Rostovtsev, V.V.; Green, L.G.; Fokin, V.V.; Sharpless, K.B. A stepwise (b) Li, W.; Du, Z.; Zhang, K.; Wang, J. Green Chem., 2015, 17, 781-784.
huisgen cycloaddition process: Copper(i)catalyzed regioselective “ligation” Available at http://dx.doi.org/10.1039/C4GC01929F.
of azides and terminal alkynes. Angew. Chem. Int. Ed., 2002, 41, 2596-2599. [14] Janreddy, D.; Kavala, V.; Kuo, C.W.; Chen, W.C.; Ramesh, C.; Kotipalli, T.;
Available at http://dx.doi.org/10.1002/1521-3773(20020715)41:14<2596:: Kuo, T.S.; Chen, M.L.; He, C.H.; Yao, C.F. Organocatalytic 1,3-dipolar
AID-ANIE2596>3.0.CO;2-4. cycloaddition reaction of α,β-unsaturated ketones with azides through
[9] (a) Zhang, L.; Chen, X.; Xue, P.; Sun, H.H.Y.; Williams, I.D.; Sharpless, iminium catalysis. Adv. Synth. Catal., 2013, 355, 2918-2927.
K.B.; Fokin, V.V.; Jia, G. Ruthenium-catalyzed cycloaddition of alkynes and Available at http://dx.doi.org/10.1002/adsc.20130034.
organic azides. J. Am. Chem. Soc., 2005, 127(46), 15998-15999. [15] (a) Huisgen, R.; Szeimies, G.; Mobius, L. 1.3Dipolare Cycloadditionen,
Available at http://dx.doi.org/10.1021/ja054114s. [PMID: 16287266] XXIV. Triazoline aus organischen Aziden und α.βungesättigten
(b) Rasmussen, L.K.; Boren, B.C.; Fokin, V.V. Ruthenium-catalyzed Carbonylverbindungen oder Nitrilen. Chem. Ber., 1966, 99, 475-490.
cycloaddition of aryl azides and alkynes. Org. Lett., 2007, 9(26), 5337-5339. Available at http://dx.doi.org/10.1002/cber.19660990216.
Available at http://dx.doi.org/10.1021/ol701912s. PMID: 18052070 (b) Broeckx, W.; Overbergh, N.; Samyn, C.; Smets, G.; L’abbe, G.
(c) Boren, B.C.; Narayan, S.; Rasmussen, L.K.; Zhang, L.; Zhao, H.; Lin, Z.; Cycloaddition reactions of azides with electron-poor olefins : Isomerization
Jia, G.; Fokin, V.V. Ruthenium-catalyzed azide-alkyne cycloaddition: scope and thermolysis of the resulting Δ2-triazolines. Tetrahedron, 1971, 27, 3527-
and mechanism. J. Am. Chem. Soc., 2008, 130(28), 8923-8930. 3534.
Available at http://dx.doi.org/10.1021/ja0749993. [PMID: 18570425] Available at http://dx.doi.org/10.1016/S0040-4020(01)97763-4.
[10] (a) Krasiński, A.; Fokin, V.V.; Sharpless, K.B. Direct synthesis of 1,5- [16] (a) Husinec, S.; Porter, A.E.A. Roberts, J.S.; Strachan, C.H. Some
disubstituted-4-magnesio-1,2,3-triazoles, revisited. Org. Lett., 2004, 6(8), approaches to the synthesis of kainic acid. J. Chem. Soc., Perkin Trans.,
1237-1240. 1984, 1, 2517-2522.
Available at http://dx.doi.org/10.1021/ol0499203. [PMID: 15070306] Available at http://dx.doi.org/10.1039/p19840002517.
(b) Wu, W-M.; Deng, J.; Li, Y.; Chen, Q-Y. Regiospecific synthesis of 1,4,5- (b) Anderson, G.T.; Henry, J.R.; Weinreb, S.M. High-pressure induced 1,3-
trisubstituted-1,2,3-triazole via one-pot reaction promoted by copper(i) salt. dipolar cycloadditions of azides with electron-deficient olefins. J. Org.
Synthesis, 2005, 8, 1314. Chem., 1991, 56, 6946-6948.
Available at http://dx.doi.org/10.1055/s-2005-861860. Available at http://dx.doi.org/10.1021/jo00024a047.
72 Current Organic Synthesis, 2020, Vol. 17, No. 1 Sultana and Sarma

(c) Prager, R.H.; Razzino, P. Heterocyclic synthesis with azides. iii. reactions [25] Prez, J.M.; Crosbie, P.; Lal, S.; Gonzalez, S.D. Lowtemperature preferential
of triazolines made from arylmethylidenemalonates. Aust. J. Chem., 1994, oxidation of carbon monoxide on pt3ni alloy nanoparticle catalyst with
47, 1375-1385. engineered surface (chemcatchem 1/2016). Chem.Cat.Chem., 2016, 8, 1-6.
Available at http://dx.doi.org/10.1071/CH9941375. Available at http://dx.doi.org/10.1002/cctc.201501371.
(d) Yang, C-H.; Lee, L-T. ang, J.-H. Spiropyrazolines from tandem reaction [26] Zhang, J.; Jin, G.; Xiao, S.; Wu, J.; Cao, S. Novel synthesis of 1,4,5-
of azides and alkyl vinyl ketones. Tetrahedron, 1994, 50, 12133-12142. trisubstituted 1,2,3-triazoles via a one-pot three-component reaction of
Available at http://dx.doi.org/10.1016/S0040-4020(01)89566-. boronic acids, azide, and active methylene ketones. Tetrahedron, 2013, 69,
[17] Janreddy, D.; Kavala, V.; Kuo, C-W.; Chen, W-C.; Ramesh, C.; Kotipalli, 2352-2356.
T.; Kuo, T-S.; Chen, M-L.; He, C-H.; Yao, C-F. Copper(i)catalyzed aerobic Available at http://dx.doi.org/10.1016/j.tet.2012.12.086.
oxidative azide–alkene cyclo addition: An efficient synthesis of substituted [27] Kaboudin, B.; Abedi, Y.; Yokomatsu, T. One-pot synthesis of 1,2,3-triazoles
1,2,3triazoles. Adv. Synth. Catal., 2013, 355, 2918-2927. from boronic acids in water using Cu(II)-β-cyclodextrin complex as a
Available at http://dx.doi.org/10.1002/adsc.201300344. nanocatalyst. Org. Biomol. Chem., 2012, 10(23), 4543-4548.
[18] Zhang, Y.; Li, X.; Li, J.; Chen, J.; Meng, X.; Zhao, M.; Chen, B. CuO- http://dx.doi.org/10.1039/c2ob25061f PMID: 22576790
promoted construction of N-2-aryl-substituted-1,2,3-triazoles via azide- [28] Lua, J.; Maa, E.; Liub, Y.; Lia, Y.; Moa, L.; Zhang, Z. Cobalt(ii)-catalyzed
chalcone oxidative cycloaddition and post-triazole arylation. Org. Lett., remote C5-selective C–H sulfonylation of quinolines via insertion of sulfur
2012, 14(1), 26-29. dioxide. RSC Adv., 2017, 7, 51313-51317.
[29] Kaboudin, B.; Mostafalua, R.; Yokomatsu, T. Fe3O4nanoparticle-supported
Available at http://dx.doi.org/10.1021/ol202718d. [PMID: 22133007]
Cu(ii)-β-cyclodextrin complex as a magnetically recoverable and reusable
[19] Chen, Y.; Nie, G.; Zhang, Q.; Ma, S.; Li, H.; Hu, Q. Copper-catalyzed [3 +
catalyst for the synthesis of symmetrical biaryls and 1,2,3-triazoles from aryl
2] cycloaddition/oxidation reactions between nitro-olefins and organic
boronic acids. Green Chem., 2013, 15, 2266-2274.
azides: highly regioselective synthesis of NO2-substituted 1,2,3-triazoles.
Available at http://dx.doi.org/10.1039/c3gc40753e.
Org. Lett., 2015, 17(5), 1118-1121.
[30] Garg, A.; Ali, A.A.; Damarla, K.; Kumar, A.; Sarma, D. Aqueous bile salt
Available at http://dx.doi.org/10.1021/ol503687w. [PMID: 25695309]
accelerated cascade synthesis of 1,2,3-triazoles from arylboronic acids.
[20] Rohilla, S.; Patel, S.S.; Jain, N. Eur. J. Org. Chem., 2016, 847-854. Tetrahedron Lett., 2018, 59, 3975-4045.
Available at http://dx.doi.org/10.1002/ejoc.201501301. Available at http://dx.doi.org/10.1016/j.tetlet.2018.09.064.
[21] (a) Barral, K.; Moorhouse, A.D.; Moses, J.E. Efficient conversion of [31] (a) Clareen, S.S.; Marshall, S.R.; Price, K.E.; Royall, M.B.; Yoder, C.H.;
aromatic amines into azides: A one-pot synthesis of triazole linkages. Org. Schaeffer, R.W. J. Chem. Educ., 2000, 77.
Lett., 2007, 9(9), 1809-1811. (b) Glemser, O.; Sauer, H. Tetraamminecopper (II) SulfateHandbook of
Available at http://dx.doi.org/10.1021/ol070527h. [PMID: 17391043] Preparative Inorganic Chemistry. 2nd Ed., Academic Press: Cambridge,
(b) Guo, S.; Lim, M.H.; Huynh, H.V. Copper(i) heteroleptic bis(nhc) and 1963, p. 11021.
mixed nhc/phosphine complexes: Syntheses and catalytic activities in the [32] Morosin, B. The crystal structures of copper tetraammine complexes. A. Acta
one-pot sequential cuaac reaction of aromatic amines. Organometallics, Crystallogr., 1969, B25, 19-30.
2013, 32, 7225-7233. Available at http://dx.doi.org/10.1107/S0567740869001725.
Available at http://dx.doi.org/10.1021/om400911u. [33] Clareen, S.S.; Marshall, S.R.; Price, K.E.; Royall, M.B.; Yoder, C.H.;
[22] (a) Zhu, W.; Ma, D. Synthesis of aryl azides and vinyl azides via proline- Schaeffer, R.W. J. Chem. Educ., 2002, 17, 904.
promoted CuI-catalyzed coupling reactions. Chem. Commun. (Camb.), 2004, [34] (a) McNulty, J.; Keskar, K.; Vemula, R. The first well-defined silver(I)-
(7), 888-889. complex-catalyzed cycloaddition of azides onto terminal alkynes at room
Available at http://dx.doi.org/10.1039/b400878b. [PMID: 15045114] temperature. Chemistry, 2011, 17(52), 14727-14730.
(b) Feldman, A.K.; Colasson, B.; Fokin, V.V. One-pot synthesis of 1,4- Available at http://dx.doi.org/10.1002/chem.201103244. [PMID: 22125272]
disubstituted 1,2,3-triazoles from in situ generated azides. Org. Lett., 2004, (b) McNulty, J.; Keskar, K. Discovery of a robust and efficient homogeneous
6(22), 3897-3899. silver(i) catalyst for the cycloaddition of azides onto terminal alkynes. Eur.
Available at http://dx.doi.org/10.1021/ol048859z. [PMID: 15496058] J. Org. Chem., 2012, 5462-5470.
(c) Andersen, J.; Bolving, S.; Liang, X. Synlett, 2005, 2941-2947. Available at http://dx.doi.org/10.1002/ejoc.201200930.
[23] Mukherjee, N.; Ahammed, S.; Bhadra, S.; Ranu, B.C. Solvent-free one-pot [35] (a) Speers, A.E.; Cravatt, B.F. Profiling enzyme activities in vivo using click
synthesis of 1,2,3-triazole derivatives by the ‘Click’ reaction of alkyl halides chemistry methods. Chem. Biol., 2004, 11(4), 535-546.
or aryl boronic acids, sodium azide and terminal alkynes over a Cu/Al2O3 http://dx.doi.org/10.1016/j.chembiol.2004.03.012 PMID: 15123248
surface under ball-milling. Green Chem., 2013, 15, 389. (b) Salic, A.; Mitchison, T.J. A chemical method for fast and sensitive
Available at http://dx.doi.org/10.1039/C2GC36521A. detection of DNA synthesis in vivo. Proc. Natl. Acad. Sci. USA, 2008,
[24] Anil Kumar, B.S.P.; Reddy, K.H.V.; Karnakar, K.; Satish, G.; Nageswar, 105(7), 2415-2420.
http://dx.doi.org/10.1073/pnas.0712168105 PMID: 18272492
Y.V.D. Copper on chitosan: An efficient and easily recoverable
[36] Abel, G.R., Jr; Calabrese, Z.A.; Ayco, J.; Hein, J.E.; Ye, T. Measuring and
heterogeneous catalyst for one pot synthesis of 1,2,3-triazoles from aryl
suppressing the oxidative damage to DNA during cu(i)-catalyzed azide-
boronic acids in water at room temperature. Tetrahedron Lett., 2015, 56,
alkyne cycloaddition. Bioconjug. Chem., 2016, 27(3), 698-704.
1968.
Available at http://dx.doi.org/10.1021/acs.bioconjchem.5b00665. [PMID:
Available at http://dx.doi.org/10.1016/j.tetlet.2015.02.107.
26829457]

You might also like