Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Tetrahedron Letters 104 (2022) 154020

Contents lists available at ScienceDirect

Tetrahedron Letters
journal homepage: www.elsevier.com/locate/tetlet

An efficient, surfactant mediated Biginelli condensation for the one pot


synthesis of dihydropyrimidine derivatives
Deboshikha Bhattacharjee, Bekington Myrboh ⇑
Centre for Advanced Studies in Chemistry, Department of Chemistry, North-Eastern Hill University, Mawlai, Shillong 793022, India

a r t i c l e i n f o a b s t r a c t

Article history: Sodium dodecyl sulfate (SDS) has been demonstrated as an efficient surfactant mediated catalyst for the
Received 2 June 2022 synthesis of dihydropyrimidinone derivatives by the Biginelli condensation of aldehydes, b-dicarbonyls,
Revised 6 July 2022 and urea/thiourea. The one-pot three component reaction was accomplished in aqueous micellar media
Accepted 11 July 2022
affording excellent yield of the products. The reaction being facile and simple is also supplemented with
Available online 14 July 2022
merits like eco-friendly catalyst, aqueous reaction media and non-chromatographic methods of purifica-
tion. The versatility of this method was examined with various substituted aldehydes and has proved to
be proficient with satisfactory results.
Ó 2022 Elsevier Ltd. All rights reserved.

Introduction proposed but unfortunately, the synthesis of 3,4-dihydropyrim-


idin-2(1H)-ones often provides generally low to moderate yields.
Multi-component reactions have emerged as an important pro- Therefore, the development more efficient protocols for the syn-
tocol in recent times for the synthesis of biologically important thesis the pharmacologically important DHMPs still remains rele-
scaffolds. One of the important scaffolds known is 3,4-dihydropy- vant amongst researchers.
rimidin-2(1H)-ones (DHPM), the derivatives of which is well A number of catalysts have been employed for the Biginelli
known to exhibit significant therapeutic and pharmacological reaction which includes both homogenous and heterogeneous cat-
properties such as antiviral, antitumour, antibacterial, and antiin- alysts like p-toluene sulfonic acid, [6] SiO2-polyphosphoric acid
flammatory. [1] In addition to this, some of DHMPS have been (SiO2-PPA), [7] boric acid, [8] HCl/FeCl3, [9] polyvinyl sulfonic acid,
employed as calcium channel blockers, a-1a-antagonists and neu- [10] graphite, [11] L-proline, [12] silica sulfuric acid, [13] zeolite,
ropeptide Y(NPY) antagonists. [2] To name a few important [14] sulphated tungstate, [15] xanthan sulfuric acid, [16] montmo-
DHMPS, Monastrol exhibits important biological property thereby rillonite KSF, [17] sulphated silica tungstic acid, [18] PPF–SO3H,
blocking mitosis by specifically inhibiting the motor activity of the [19] Al(HSO4)3, [20] vanadium(III) chloride, [21] bio glycerol-based
mitotic kinesin Eg5 and is a pilot molecule for the advancement of sulfonic acid functionalized carbon catalyst. [22] Generally classi-
novel anticancer drugs, (R)-SQ 32926, SQ 32547, SWO2 also known cal methods like heating, microwave and ultrasound irradiations
to be antihypertensive agent with potent oral activity and Mon-97 have been used for the synthesis 3,4-Dihydropyrimidin-2(1H)-
shows promising anticancer activity (Fig.1). [3] Several alkaloids ones/thiones under different catalytic condition. [23] Other impro-
obtained from marine sources have been also found to possess vised catalysts include ionic liquids, [24] heteropolyacids, [25]
the dihydropyrimidine core unit. Most noteworthy among these nanomaterials [26–28], sulphated polyborate [29] which have
are the batzelladine alkaloids, which were found to be potent been also utilised for the synthesis of these important heterocycles.
HIVgp-120-CD4 inhibitors. [4]. Although a number of improved methods have been developed,
The earliest attempt to synthesize the 3,4-dihydropyrimidin-2 most of them however suffer from drawbacks such as cumbersome
(1H)-ones was reported by Pietro Biginelli in 1893, [5] where he synthesis of catalysts, use of corrosive/toxic metal-based catalysts,
performed the three-component cyclocondensation reaction of volatile organic solvents, unsatisfactory yields, high cost, longer
ethyl acetoacetate, benzaldehyde and urea under Bronsted acid reaction times and harsh reaction conditions. Thus, there is still
catalysis. However, this reaction has drawbacks like harsh condi- scope for the development of new and more efficient methods
tions, high reaction time and low yields. Since the early days of for the synthesis of these important heterocycles keeping in mind
the discovery of this reaction several reaction routes have been the green principle of chemistry.
The thrust for clean, green synthetic processes requires atom
⇑ Corresponding author.
economy, minimal waste generation, environmentally benign
reagents & solvents as the prime ingredients and water serves as
E-mail address: bmyrboh@nehu.ac.in (B. Myrboh).

https://doi.org/10.1016/j.tetlet.2022.154020
0040-4039/Ó 2022 Elsevier Ltd. All rights reserved.
D. Bhattacharjee and B. Myrboh Tetrahedron Letters 104 (2022) 154020

Fig. 1. Pharmacologically important 3,4-dihydropyrimidin-2(1H)-ones/thiones.

the most preferred solvent whenever possible. [30] The low solu- (6a). To investigate the effect of catalyst, the model reaction was
bility of organic reagents and the instability of the majority of cat- carried out in absence of any catalyst which resulted in no product
alysts in water mediated reaction is a concern. But in recent time, formation even after 18 h. The reaction was further heated for 24 h
many surfactants have been used as phase transfer catalysts in sev- which resulted in only trace amount of the product. In the next
eral organic reactions having unique capabilities to dissolve both attempt we employed SDS a surfactant as catalyst in the model
organic and aqueous solutions to enhance the rate of the reaction. reaction at room temperature in water which resulted in very less
[31] The anionic surfactant sodium dodecyl sulphate (SDS), shows product formation even after stirring for 12 h. The reaction was
the catalytic activity almost equal with Brønsted acids, such as then heated to 60 °C which furnished the desired product in a spec-
sodium hydrosulphate or p-toluene sulfonic acid. Moreover, its cat- ulated time of 4 h. Encouraged by this result, we carried out the
alytic activity is also known be greater than that of most Lewis reaction of 4 methyl benzaldehyde (1 h) (2.0 mmol), thiourea (5)
acids [32] and therefore, SDS has been employed as catalyst in a (3.0 mmol) and ethyl acetoacetate (2) (2.1 mmol) under identical
variety of reaction schemes. [32–33] In the reaction medium SDS conditions. Upon stirring the reaction at 60 °C excellent yield
forms micelles in water and can both solubilise the organic com- (92 %) of the product (6 h) was obtained in little longer time
pounds and catalyse the reaction thus increasing the energy (5 h). The efficacy of this method was further enhanced when
efficiency. Monastrol (6 k), a biologically active molecule has been success-
Therefore, in our continuous quest [34] for more efficient green fully synthesized. The structure of the compounds formed were
reaction protocols for the synthesis of the biologically important confirmed by 1H & 13C NMR, Mass and IR analyses.
3,4-dihydropyrimidin-2(1H)-ones/thiones, the employment of In order to evaluate the catalyst loading, the model reaction was
SDS an eco-friendly and cheap catalytic system for the one-pot carried out at different catalyst loading, viz. no catalyst, 05 mol%,
three-component synthesis of 3,4-dihydropyrimidin-2(1H)-ones/ 10 mol%, 15 mol%, 20 mol%, and 25 mol% of SDS in water at
thiones derivatives in aqueous medium was explored. 60 °C. Except in case of no catalyst condition the reaction worked
in all cases. In case of 5 mol% to 15 mol% catalyst loadings the
Results and discussion expected product formation was only up to 55 % upon stirring
for 4 h. To our delight, it was observed that when catalyst loadings
As a part of our effort towards the development of environmen- was increased to 20 mol% and 25 mol% the reaction yielded 85 %
tally benign synthetic methods for multi-component reactions and 91 % yield respectively in similar time of 4 h. Interestingly,
(MCRs) to synthesize various biologically important heterocyclic enhancement of catalyst loading to 30 mol% did not result in
compounds, we have developed an efficient synthesis of 3,4-dihy- reduction of the reaction time appreciably and gave almost similar
dropyrimidin-2(1H)-ones/thiones derivatives (6/7) using SDS in yields to that of optimum catalyst loading (25 mol %) in 4 h (Fig. 2).
water at 60 °C via the condensation of aldehyde (1), ethyl acetoac- The efficiency of nonionic surfactant PEG-600, Tween-80 and
etate (2)/methyl acetoacetate (3), and urea/thiourea (4/5) cationic surfactant TBAB, CTAB were also screened in order to eval-
(Scheme 1). uate and compare the catalytic activity for the three-component
In our initial attempt to synthesize the 3,4-dihydropyrimidin-2 model reaction in water. It is evident from Fig. 3 TBAB and CTAB
(1H)-ones/thiones derivatives, we selected the model reaction of was also efficient enough to catalyze the reaction to yield 69 %
4 Chloro benzaldehyde (1a) (2.0 mmol), urea (4) (3.0 mmol) and and 75 % respectively whereas PEG-600 and Tween-80 gave poor
ethyl acetoacetate (3) (2.1 mmol) to yield ethyl 4-(4-chlorophe- result under similar reaction condition yielding only 42 % and
nyl)-6-methyl-2-oxo-1,2,3,4-tetrahydropyrimidine-5-carboxylate
2
D. Bhattacharjee and B. Myrboh Tetrahedron Letters 104 (2022) 154020

Scheme 1. SDS catalysed synthesis of 3,4-dihydropyrimidin-2(1H)-ones/thiones derivatives.

Fig. 4. Effect of solvent on the yield of the product.

Fig. 2. Effect of Catalyst loading and temperature on the yield of the product.

Fig. 5. Recyclability of the catalyst.


Fig. 3. Effect of surfactant on the yield of the product.

simple workup where the catalyst was retained in the aqueous


35 % respectively proving SDS as the best surfactant for the men- layer leaving behind the product in organic layer. The regenerated
tioned reaction yielding 91 % of the desired product. catalyst in aqueous solution was then reused for the same reaction
The effect of solvent was then investigated by carrying out the and it was observed the efficiency of catalyst did not change appre-
model reaction in different solvents like DCM, THF, chloroform, ciably even after using for 4 consecutive times for the model
DMF, H2O and in EtOH:H2O (1:1) ratio at fixed catalyst loading reaction.
(25 mol%). It was observed that the reaction was efficient in protic After the optimization of the reaction parameters such as cata-
solvents like EtOH, H2O, EtOH:H2O, whereas in other solvent the lyst, catalyst loading and solvent, the efficiency of the optimised
reaction proceeded sluggishly (Fig. 4), thus confirming water as protocol was tested for a number of aromatic aldehydes having
the best suited solvent for the micelle formation. Micelles can be varying substituent. In each case the reaction proceeded smoothly
used to promote multi-component reactions by taking advantage at optimum reaction condition to give the desired product in mod-
of the confinement effect as well as of the presence of certain func- erate to excellent yields. The presence of electron withdrawing or
tional groups on the surface of the aggregates to improve the yield releasing substituent in the ortho-, meta- and para-positions do
and reaction time. not appear to have any effect on the overall reaction yields
The catalyst recyclability was investigated for the model reac- (Table 1). While in case of thiourea derivatives product formation
tion (Fig. 5). Catalyst was recovered from the reaction mixture by took longer time as compared to that of urea derivatives.

3
D. Bhattacharjee and B. Myrboh Tetrahedron Letters 104 (2022) 154020

Table 1
SDS catalysed synthesis of dihydropyrimidinones derivatives.

Entry X Aldehyde Producta Time(hrs) Yieldb


1 O 6a 4 91

2 O 6b 4 89

3 O 6c 4 87

4 O 6d 4 90

5 O 6e 4 92

6 S 6f 5 88

7 S 6g 5 85

8 S 6h 5 91

9 S 6i 5 89

10 S 6j 5 86

11 S 6k 5 85

Reaction Condition: aldehyde (2.0 mmol), ethyl acetoacetate (2.1 mmol), urea/thiourea (3.0 mmol), and SDS (25 mol %) at 60 °C.
a
Isolated yields.
b
Products were characterized by 1H & 13C NMR, Mass and IR analyses.

Likewise, almost identical results were obtained when methyl hyde, urea/thiourea, 1,3-diones condensed to produce dihydropy-
acetoacetate (3) was replaced in place of ethyl acetoacetate (2) rimidinones derivatives which are hydrophobic in nature.
and the three-component reaction with urea (4) and 4 Chloro ben- Surfactant forms micelle at a concentration above its CMC, and
zaldehyde (1a) was carried out in presence of 25 mol% of SDS in the presence of additives decreases the CMC of ionic surfactants
water under similar condition. To study the scope of the reaction to a large extent. This decrease in the CMC is attributed to the
thiourea (5) and varying substituted aldehydes (1) were employed. counter-ion binding, which neutralize the charged headgroups of
The electrophilicity of the aldehydes resulting from the +I or I the ionic micelles. It has also been reported that aromatic com-
effect of substituent on the phenyl ring did not much effected pounds also reduce the CMC of ionic surfactants by residing at
the overall reaction yield in all the cases (Table 2). In case of the micellar interface. [35] The CMC of SDS in water is around
thiourea derivatives also gave very good yields under similar reac- 8 mM [36] and in the current reaction medium, wherein aromatic
tion conditions, albeit taking longer times than their urea counter- aldehydes are present in addition to the other additives, the CMC of
part. The products were successfully isolated and purified by SDS will be lowered as well. [37] Thus, in the micellar solution of
recrystallisation from ethanol without the use of chromatographic SDS the hydrophobic moieties escape away from water molecules
method. thereby encircling the micelle core of SDS. Moreover, the reactant
The catalytic property of SDS to accelerate the rate of reaction molecules are pushed by the water surrounding the micelle to
can be explained as follows. In the above discussed reaction alde- enter inside the hydrophobic core of the micelle and resulting

4
D. Bhattacharjee and B. Myrboh Tetrahedron Letters 104 (2022) 154020

Table 2
SDS catalysed synthesis of dihydropyrimidinones derivatives.

Entry X Aldehyde Producta Time(hrs) Yieldb


1 O 7a 4 93

2 O 7b 4 89

3 O 7c 4 85

4 O 7d 4 90

5 O 7e 4 91

6 S 7f 5 84

7 S 7g 5 92

8 S 7h 5 83

9 S 7i 5 93

10 S 7j 5 86

Reaction Condition: aldehyde (2.0 mmol), methyl acetoacetate (2.1 mmol), urea/thiourea (3.0 mmol), and SDS (25 mol %) at 60 °C.
a
Isolated yields.
b
Products were characterized by 1H & 13C NMR, Mass and IR analyses.

Fig. 6. Schematic representation of role of micellar SDS droplets on the reaction in aqueous media.

5
D. Bhattacharjee and B. Myrboh Tetrahedron Letters 104 (2022) 154020

Scheme 2. Plausible mechanism for the formation of dihydropyrimidinones/thione derivatives.

more efficient reactant collisions takes place to thus facilitating the (2012) 2872–2876;
(d) A. de Fátima, T.C. Braga, S. Neto Lda, B.S. Terra, G.F. Oliveira, D.L. da Silva, L.
reaction [38] which can be schematically represented as Fig.6.
V. Modolo, J. Adv. Res. 6 (2014) 363–373.
The plausible mechanism is depicted in Scheme 2. The reaction [2] K.S. Atwal, G.C. Rovnyak, S.D. Kimball, D.M. Floyd, S. Moreland, B.N. Swanson, J.
presumably proceeds via the reaction between aldehyde (A) with Z. Gougoutas, J. Schwartz, K.M. Smillie, M.F. Malley, J. Med. Chem. 33 (1990)
urea/thiourea (B) to form the acylimine intermediate d which is 2629–2635.
[3] (a) S. Gore, S. Baskaran, B. Koenig, Green Chem. 13 (2011) 1009–1013;
the key step, which subsequently reacts with the 1,3 diketone (b) C.O. Kappe, O.V. Shishkin, G. Uray, P. Verdino, Tetrahedron 56 (2000)
derivatives (E) followed by intramolecular cyclisation and dehy- 1859–1862;
dration afforded the corresponding 3,4-dihydropyrimidinones/ (c) C.O. Kappe, Eur. J. Med. Chem. 35 (2000) 1043–1052.
[4] (a) A.D. Patil, N.V. Kumar, W.C. Kokke, M.F. Bean, A.J. Freyer, C.D. Brosse, S. Mai,
thiones derivatives (G)(6/7). The mechanism is in good agreement A. Truneh, D.J. Faulkner, B. Carte, A.L. Breen, R.P. Hertzberg, R.K. Johnson, J.W.
to that reported in the literature [39]. Westley, B.C.M. Potts, J. Eur. J. Org. Chem. 60 (1995) 1182–1188;
(b) B.B. Snider, J. Chen, A.D. Patil, A.J. Freyer, Tetrahedron Lett. 37 (1996)
6977–6980;
Conclusion (c) L. Heys, C.G. Moore, P. Murphy, J. Chem. Soc. Rev. 29 (2000) 57–67.
[5] P. Biginelli, P. Gazz, Chim Ital. 23 (1893) 360–416.
[6] T. Jin, S. Zhang, T. Li, Synth. Commun. 32 (2002) 1847–1851.
In conclusion, an efficient, facile, one-pot three-component [7] M. Zeinali-Dastmalbaf, A. Davoodnia, M.M. Heravi, N. Tavakoli-Hoseini, A.
method for the synthesis of biologically important dihydropyrim- Khojastehnezhad, H.A. Zamani, Bull. Korean Chem. Soc. 32 (2011) 656–658.
idinones/thione derivatives via SDS a surfactant, a green and mild [8] S. Tu, F. Fang, C. Miao, H. Jiang, Y. Feng, D. Shi, X. Wang, Tetrahedron Lett. 44
(2003) 6153–6156.
catalyst has been developed. The methodology offers high conver-
[9] J. Lu, MaH Synlett 2000 (2000) 63–64.
gence, cost effectiveness and high throughput in short reaction [10] A. Rahmatpour, Catal. Lett. 142 (2012) 1505–1511.
time. In addition to this, use of environmentally safer reagents, sol- [11] K.L. Dhumaskar, S.N. Meena, S.C. Ghadi, S.G. Tilve, Bioorg. Med. Chem. Lett. 24
(2014) 2897–2899.
vent, recyclable catalysts, non-chromatographic method of purifi-
[12] J.S. Yadav, P.S. Kumar, G. Kondaji, S.R. Rao, K. Nagaiah, Chem. Lett. 33 (2004)
cations are the add-on of this protocol which supersedes many 1168–1169.
pre-existing methodologies. [13] P. Salehi, M. Dabiri, M.A. Zolfigol, M.A.B. Fard, Tetrahedron Lett. 44 (2003)
2889–2891.
[14] V.R. Rani, N. Srinias, M.R. Kishan, S.J. Kulkarni, K.V. Raghavan, Green Chem. 3
Data availability (2001) 305–306.
[15] S.D. Salim, K.G. Akamanchi, Catal. Commun. 12 (2011) 1153–1156.
[16] B.S. Kuarm, J.V. Madhav, S.V. Laxmi, B. Rajitha, Synth. Commun. 42 (2012)
Data will be made available on request. 1211–1217.
[17] F. Bigi, S. Carloni, B. Frullanti, R. Maggi, G. Sartori, Tetrahedron Lett. 40 (1999)
Declaration of Competing Interest 3465–4368.
[18] N. Ahmed, Z.N. Siddiqui, J. Mol. Catal. A: Chem. 387 (2014) 45–56.
[19] X.L. Shi, H. Yang, M. Tao, W. Zhang, RSC Adv. 3 (2013) 3939–3945.
The authors declare that they have no known competing finan- [20] H. Shaterian, A. Hosseinian, M. Ghashang, F. Khorami, N. Karimpoor,
cial interests or personal relationships that could have appeared Phosphorus, Sulfur Silicon Relat. Elem. 184 (2009) 2333–2338.
[21] G. Sabitha, G.S.K.K. Reddy, K.B. Reddy, J.S. Yadav, Tetrahedron Lett. 44 (2003)
to influence the work reported in this paper. 6497–6499.
[22] K. Konkala, N.M. Sabbavarapu, R. Katla, N.Y.V. Durga, V.K.T. Reddy, L.A. Bethala,
P. Devi, R.B.N. Prasad, Tetrahedron Lett. 53 (2012) 1968–1973.
Appendix A. Supplementary data [23] (a) J. Safari, S. Gandomi-Ravandi, New J. Chem. 38 (2014) 3514–3521;
(b) J. Lal, M. Sharma, S. Gupta, P. Parashar, P. Sahu, D.D. Agarwal, J. Mol. Catal.
Supplementary data to this article can be found online at A: Chem. 352 (2012) 31–37;
(c) J.S. Yadav, S.B.V. Reddy, J.E. Reddy, T.J. Ramalingam, J. Chem. Res., Synop.
https://doi.org/10.1016/j.tetlet.2022.154020.
2000 (2000) 354–355;
(d) I. Saxena, D.C. Borah, J.C. Sarma, Tetrahedron Lett. 46 (2005) 1159–1160;
References (e) B. Vijayakumar, G.R. Rao, J. Porous Mater. 19 (2012) 491–497;
(f) P.G. Mandhane, R.S. Joshi, D.R. Nagargoje, C.H. Gill, Tetrahedron Lett. 51
(2010) 3138–3140.
[1] (a) K.S. Atwal, B.N. Swanson, S.E. Unger, D.M. Floyd, S. Moreland, A. Hedberg, B.
[24] (a) R. Zheng, X. Wang, H. Xu, J. Du, Synth. Commun. 36 (2006) 1503–1513;
C. O’Reilly, J. Med. Chem. 34 (1991) 806–811;
(b) J.C. Legeay, J.J.V. Eynde, J.P. Bazureau, Tetrahedron 61 (2005) 12386–
(b) G.C. Rovnyak, S.D. Kimball, B. Beyere, G. Cucinotta, J.D. DiMarco, J.
12397;
Gougoutas, A. Hedberg, M. Malley, J.P. McCarthy, R. Zhang, S. Moreland, J.
(c) A.R. Gholap, K. Venkatesan, T. Daniel, R.J. Lahoti, K.V. Srinivasan, Green
Med. Chem. 38 (1995) 119–129;
Chem. 6 (2004) 147–150.
(c) J. Lal, S.K. Gupta, D. Thavaselvam, D.D. Agarwal, Bioorg. Med. Chem. Lett. 22

6
D. Bhattacharjee and B. Myrboh Tetrahedron Letters 104 (2022) 154020

[25] (a) R. Fazaeli, S. Tangestaninejad, H. Aliyan, M. Moghadam, Appl. Catal. A: Gen [30] N. Parikh, S.R. Roy, K. Seth, A. Kumar, A.K. Chakraborti, Synthesis 48 (2016)
309 (2006) 44–51; 547–556.
(b) M.M. Heravi, F. Derikvand, F.F. Bamoharram, J. Mol. Catal. A: Chem. 242 [31] (a) D.O. Jang, Tetrahedron Lett. 37 (1996) 5367–5368;
(2005) 173–175; (b) H. Yorimitsu, H. Shinokubo, K. Oshima, Chem. Lett. 29 (2000) 104–105.
(c) S.P. Maradur, G.S. Gokavi, Catal. Commun. 8 (2007) 279–284; [32] J. Qian, J. Xu, J. Zhang, Pet. Sci. Technol. 29 (2011) 462–467.
(d) F.L. Zumpe, M. Flub, K. Schmitz, A. Lender, Tetrahedron Lett. 48 (2007) [33] (a) R. Sharma, A.K. Pandey, P.M.S. Chauhan, Synlett 23 (2012) 2209–2214;
1421–1423; (b) H. Mehrabi, H. Abusaidi, J. Iran. Chem. Soc. 7 (2010) 890–894;
(e) G.P. Romanelli, A.G. Sathicq, J.C. Autino, G. Baronetti, H.J. Thomas, Synth. (c) D.R.M. Arenas, C.A.M. Bonilla, V.V. Kouznetsov, Org. Biomol. Chem. 11
Commun. 37 (2007) 3907–3916. (2013) 3655–3663.
[26] (a) J. Safari, Z. Zarnegar, New J. Chem. 38 (2014) 358–365; [34] (a) D. Bhattacharjee, B. Kshair, B. Myrboh, RSC Adv. 6 (2016) 95944–95950;
(b) J. Safari, S. Gandomi-Ravandi, S. Ashiri, New J. Chem. 40 (2016) 512–520; (b) D. Bhattacharjee, D. Sutradhar, A.K. Chandra, B. Myrboh, Tetrahedron 73
(c) A. Hassanpour, J. Abolhasani, R.H. Khanmiri, J. Korean Chem. Soc. 58 (2014) (2017) 3497–3504;
445–449; (c) D. Bhattacharjee, S.K. Sheet, S. Khatua, K. Biswas, S.R. Joshi, B. Myrboh,
(d) J. Javidi, M. Esmaeilpour, F.N. Dodeji, RSC Adv. 5 (2015) 308–315; Bioorg. Med. Chem. 26 (2018) 5018–5028.
(e) C. Karami, H. Mohammadi, K. Ghodrati, H. Ahmadian, F. Jamshidi, M. Nouri, [35] I.M. Umlong, I. Kochi, J. Surface Sci. Technol. 22 (2006) 101–117.
N. Haghnazarie, Synth. React. Inorg. Met. Org. Chem. 45 (2015) 271–276. [36] (a) A. Ali, N.A. Malika, S. Uzaira, M. Alia, Mol. Phys. 112 (2014) 2681–2693;
[27] A. Khazaei, M.A. Zolfigol, S. Alaie, S. Baghery, B. Kaboudin, Y. Bayat, A. Asgari, (b) J. Dey, K. Ismail, J. Colloid Interface Sci. 378 (2012) 144–151.
RSC Adv. 6 (2016) 10114–10125. [37] B. Kim, S. Im, S. Oh, Langmuir 17 (2001) 565–566.
[28] J. Safari, Z. Zarnegar, New J. Chem. 38 (2014) 358–365. [38] H. Firouzabadi, N. Iranpoor, M. Abbasi, Adv. Synth. Catal. 351 (2009) 755–766.
[29] C.K. Khatri, D.S. Rekunge, G.U. Chaturbhuj, New J. Chem. 40 (2016) 10412– [39] (a) C.O. Kappe, J. Org. Chem. 62 (1997) 7201–7204;
10417. (b) A.N. Dadhania, V.K. Patel, D.K. Raval, J. Chem. Sci. 124 (2012) 921–926.

You might also like