Sie sind auf Seite 1von 11

DOI: 10.1002/slct.

201900695 Full Papers

1
2 z Materials Science inc. Nanomaterials & Polymers
3
4
5
Synthesis and Catalytic Application of Poly(2-acrylamido-2-
6 methyl-1-propanesulfonic acid-co-acrylamide) Grafted on
7
8 Graphene Oxide
9
10 Kaveh Parvanak Boroujeni,*[a] Zeinab Tohidiyan,[b] Abdulhamid Fadavi,[c]
11 Mohammad Mehdi Eskandari,[d] and Hekmat Allah Shahsanaei[a]
12
13
14 Graphene oxide was grafted with poly(2-acrylamido-2-methyl- acidic site loading of 0.85 mmol g 1, the maximum degradation
15 1-propanesulfonic acid-co-acrylamide) (GO–poly(AMPS–co– temperatures (Tmax) of 212, 292, and 406 °C, the char yield of
16 AM)) by free radical copolymerization of the corresponding 30% at 800 °C, the limiting oxygen index (LOI) of 29.5, an
17 monomers namely acrylamide (AM) and 2-acrylamido-2-meth- estimated band gap of 4.7 eV, and the fluorescence emission
18 yl-1-propanesulfonic acid (AMPS). The benzoyl peroxide (BPO) bands of 722, 800, and 880 nm at the excitation wavelengths
19 was used as initiator. The GO–poly(AMPS–co–AM) was charac- (λmax) of 360, 400, and 440 nm, respectively. The GO–poly
20 terized by Fourier transform infrared spectroscopy (FT–IR), (AMPS–co–AM) was used as a reusable catalyst for the
21 Raman spectroscopy, X–ray diffraction (XRD), thermal gravimet- preparation of 1,8-dioxo-octahydroxanthenes through the con-
22 ric analysis (TGA), differential thermal analysis (DTA), scanning densation of an aldehyde with 5,5-dimethyl-1,3-cyclohexane-
23 electron microscopy (SEM), energy-dispersive X–ray spectro- dione (dimedone) in H2O/EtOH (1:1) as green solvents at room
24 scopy (EDX), UV–Vis and fluorescence spectroscopy, and temperature.
25 elemental analysis. The GO–poly(AMPS–co–AM) exhibited the
26
27 Introduction
optical properties, electrical conductivity).[9–11] There are two
28
Graphene, which has been considered a “rising star” material, approaches for the covalent grafting of polymers to GO sheets,
29
has been used as filler in the preparation of polymer nano- including “grafting to” and “grafting from” methods. In
30
composites owing to its unique properties.[1–4] However, the “grafting to” method the pre-synthesized polymer or copoly-
31
major challenge for the fabrication of high performance mer are attached to GO sheets. The “grafting from” method
32
graphene/polymer nanocomposites is the tendency of gra- involves direct polymerization and copolymerization of the
33
phene sheets to agglomerate, due to strong π-π stacking and corresponding monomers on the surface of GO. The “grafting
34
van der Waals interaction,[5–7] and therefore their dispersion in from” method is a better approach for graphene modification
35
polymers is very difficult. Alternatively, graphene oxide (GO), a than “grafting to” method because it overcomes the low
36
product of graphene oxidation, bears a number of functional grafting density and slow reactivity problem associated with
37
groups such as OH, COOH and epoxy C–O–C which permit the the “grafting to” approach.[11]
38
attachment of organic molecules or polymers to the GO Xanthene derivatives are very important heterocyclic com-
39
surface.[8] The covalent grafting of polymers containing a pounds and occupy a prominent position in medicinal
40
suitable functional group onto GO sheets imparts an enhanced chemistry. For example, antimicrobial,[12] antidiabetic,[13]
41
dispersity and consequently easy fabrication (by strong polar- anticancer,[14] and antitubercular[15] activities of a series of
42
polar interactions or by their bulky size) as well as introduces xanthene derivatives have been investigated (Scheme 1).
43
some additional functional properties (e. g., thermal stability, Furthermore, these compounds can be used in laser
44
technologies[16] and as pH sensitive fluorescent materials for
45
visualization of biomolecules.[17] Among different kinds of
46 [a] Prof. K. P. Boroujeni, H. A. Shahsanaei
Department of Chemistry, Shahrekord University, Shahrekord, 115, Iran xanthenes, 1,8-dioxo-octahydroxanthenes have attracted sig-
47
E-mail: parvanak-ka@sku.ac.ir nificant attention. For preparation of 1,8-dioxo-octahydroxan-
48 [b] Dr. Z. Tohidiyan thenes a mixture of an aldehyde (1 equiv) and dimedone
49 Department of Chemistry, Shahrekord Branch, Islamic Azad University,
(2 equiv) along with a catalyst was heated. p-Dodecylbenezene-
50 Shahrekord, Iran
[c] Dr. A. Fadavi sulfonic acid (DBSA),[18] Amberlyst-15,[19] SmCl3,[20] poly(2-acryl-
51
Department of Chemistry, Marvdasht Branch, Islamic Azad University, amido-2-methyl-1-propanesulfonic acid-co-acrylic acid) (poly
52 Marvdasht, Iran (AMPS–co–AA)),[21] graphene oxide–SO3H (GO–SO3H),[22] [Et3NH]
53 [d] Dr. M. M. Eskandari
[HSO4],[23] cerric ammonium nitrate (CAN) supported HY-
54 Nanotechnology Research Center, Research Institute of Petroleum
Industry, Tehran, Iran zeolite,[24] Mg–Al hydrotalcite,[25] Mg(BF4)2 in ionic liquid 1-butyl-
55
Supporting information for this article is available on the WWW under 3-methyl imidazolium tetrafluoroborate ([bmim]BF4),[26] sulfated
56
https://doi.org/10.1002/slct.201900695 zirconia,[27] L-pyrrolidine-2-carboxylic acid sulfate (LPCAS),[28]
57

ChemistrySelect 2019, 4, 7734 – 7744 7734 © 2019 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Full Papers
In continuation of our recent works on the synthesis and
1
application of nanomaterials[39,40] in this study, we report the
2
grafting of sulfonated polyacrylamide to graphene oxide by
3
radical polymerization of the corresponding monomers and
4
investigate its thermal stability, morphology, optical properties,
5
and catalytic activity. To the best of our knowledge, GO grafted
6
with polymer have not yet been evaluated as a catalyst in
7
organic reactions.
8
9
10
Results and Discussion
11
12 Synthesis and characterization of GO–poly(AMPS–co–AM)
13
We have applied the direct-synthesis approach to the prepara-
14
tion of GO–poly(AMPS–co–AM) starting from its starting
15
materials, namely AM and AMPS monomers and BPO initiator
16
(Scheme 2). The amount of the acidic site loading in GO–poly
17
18
19
20
21
22
23
24
25
26
27 Scheme 2. Preparation of GO–poly(AMPS–co–AM).
28
29
30
(AMPS–co–AM) obtained by means of acid-base titration was
31
found to be 0.85 mmol g 1.[41] The mol % percentage of AM and
32
AMPS grafted onto the surface of GO determined by the
33
gravimetric and titration methods were about 70 and 30%,
34
respectively. This ratio was not altered by adjusting the feed
35
ratio of monomers. The higher loading of AM than AMPS is
36
related to both steric resistance effects and electrostatic
37
repulsion of AMPS monomers against the approach of them to
38
each other. Also, in a complementary experiment, we prepared
39
poly(AMPS-co-AM)[42] and then it was exposed to GO in the
40
presence of BPO initiator for 5 h in ethanol under reflux
41
condition. The reaction does not take place and this observa-
42 Scheme 1. Biological activities of a series of xanthene derivatives.
tion indicated that poly(AMPS-co-AM) was grafted on GO
43
surface through “grafting from” method not “grafting to”
44
method.
45
and furfural-imine-functionalized graphene oxide-copper oxide The FT–IR spectra of the pristine GO and GO–poly(AMPS–
46
((Cu(II)-Fur-APTES/GO))[29] have been used as catalysts for this co–AM) are shown in Figure 1. In the FT–IR spectrum of GO–
47
reaction. However many of the reported procedures need to poly(AMPS–co–AM) (Figure 1b), absorption peaks at 3347 and
48
relatively high temperature conditions[18–29] or an additional 3210 cm 1 (the stretching vibrations of amide groups), 2931
49
energy (ultrasound or microwave)[30–33] and are performed in and 2867 cm 1 (the C–H asymmetric and symmetric tensile
50
toxic organic solvents. The synthesis of 1,8-dioxo-octahydrox- vibrations, respectively),[43] 1662 cm 1 (the stretching vibrations
51
anthenes from the reaction of aldehydes with dimedone at of carbonyl groups), 1546 cm 1 (the secondary amide N–H
52
room temperature are relatively rare.[34,35] Usually, in the deformation peaks of AMPS units), 1449 and 1297 cm 1 (the
53
absence of catalyst or at ambient temperature the uncyclized bending vibrations of –CH2– and –CH–)„[43] 1210, 1113,
54
adduct 2,2’-aryl-methylene bis(3-hydroxy-2-cyclohexene-1-one) 1038 cm 1 (the stretching vibrations of S=O),[40] and 625 cm 1
55
is formed.[36–38] (the stretching vibrations of S–O)[44] implied the grafting
56
reaction. The FT–IR spectrum of GO (Figure 1b) did not show
57

ChemistrySelect 2019, 4, 7734 – 7744 7735 © 2019 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Full Papers

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21 Figure 1. FT–IR spectra of GO (a) and GO–poly(AMPS–co–AM) (b).
22
23
24
any the above mentioned characteristic peaks of GO–poly
25
(AMPS–co–AM).
26
The synthesized GO–poly(AMPS–co–AM) was also surveyed
27
by Raman spectroscopy. Figures S1,S2 show the Raman spectra
28
of the pristine GO and the GO–poly(AMPS–co–AM), respec-
29
tively. Figure S1 exhibits two characteristic bands at 1356 and
30
1603 cm 1, due to the D band (sp3 carbon atoms of defects and
31
disorders) and G band (sp2 carbon atoms in sheets of graphite)
32
of GO structure, respectively.[45] The corresponding bands of
33
Raman spectrum of GO–poly(AMPS–co–AM) (Figure S2) ap-
34
peared at 1367 and 1598 cm 1, respectively. Clearly, the
35
position of the characteristic D and G bands in spectrum of
36
GO–poly(AMPS–co–AM) have shifted with respect to those of
37
the pristine GO which can be attributed to the powerful
38
interaction between the GO sheets and the polymer chains.
39
The TGA data also support the immobilization of poly
40
(AMPS-co-AM) on GO. The corresponding TGA/DTA curves for
41
the pristine GO and the GO–poly(AMPS–co–AM) are shown in
42
Figures 2a,b. In both curves the first weight loss (about 3%)
43
centers around 100 °C is referred to the elimination of the
44
surface-adsorbed water and interlayer water molecules. The
45
second large weight loss in TGA curve of GO (starting at 110 °C
46
and continuing to 811 °C) is related to removal of function-
47 Figure 2. TGA/DTA curves for GO (a) and GO–poly(AMPS–co–AM) (b).
alities on GO and its own decomposition (Figure 2a). In TGA
48
curve of GO–poly(AMPS–co–AM) (Figure 2b and Figure 3), the
49
second weight loss (15%), which start at 100 °C and continue to
50
250 °C, is ascribed to the decomposition of a low molecular polymer backbone and the decomposition of GO structure.
51
weight polymer and the functionalities on GO. The third weight From the third weight loss of the GO–poly(AMPS–co–AM), it
52
loss by 27 wt% from 250 to 350 °C is mainly related to the can be concluded that the loading level of the grafted polymer
53
removal of the pendant groups of the grafted polymer (amide in GO–poly(AMPS–co–AM) is about 0.89 mmol g 1, which has
54
and sulfonic acid groups) and some remaining functionalities good agreement with the results obtained by means of acid-
55
on GO. The last weight loss (25%) which takes place at base titration and gravimetric, methods.
56
temperatures above 350 °C is due to the degradation of the
57

ChemistrySelect 2019, 4, 7734 – 7744 7736 © 2019 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Full Papers
respectively. On the basis of LOI values, the GO–poly(AMPS–co–
1
AM) can be classified as fire retardant materials (with LOI values
2
higher than 21).[47]
3
Figure 4 exhibits XRD patterns of GO and GO–poly(AMPS–
4
co–AM). The pattern of GO (Figure 4a) manifested a main peak
5
at 10.79° duo to an interlayer spacing (d-spacing) of about
6
0.80 nm by Debye–Scherrer equation.[48] This peak atributed to
7
presence of oxygen species on the sheet of GO.[49,50] The XRD
8
pattern of GO–poly(AMPS–co–AM) (Figure 4b) exhabits a
9
diffraction peak at 9.43°, related to GO sheet (d-spacing of
10
about 1.52 nm). This peak shows a shift at position and value of
11
d-spacing in comparison with its corresponding peak in GO.
12
The shift was attributed to the presence of the supported
13
species in this compound. The SO3H groups in GO–poly(AMPS–
14
co–AM) can lead to an increase in the distance between sheets
15
of GO significantly. Furthermore, the broad peak at 2θ = 10–
16
20° was assigned to amorphous structure of the grafted
17
Figure 3. TGA curves for GO (2, red line) and GO–poly(AMPS–co–AM) (1, blue polymer.[51]
18 line). To more characterization of GO–poly(AMPS–co–AM), energy
19
dispersive X-ray spectroscopy (EDX) analysis and the corre-
20
sponding EDX elemental mappings were provided in Figure 5.
21
Based on the DTA curves of GO and the GO–poly(AMPS– Clearly, Figure 5a demonstrates the presence of C, O, N, and S
22
co–AM), the maximum degradation temperatures (Tmax) of GO elements in the GO–poly(AMPS–co–AM) without any impurities.
23
appeared at 193 and 214 °C (Figure 2a). However, from Fig- The weight and atomic percentage of the elements have been
24
ure 2b, it is clear that the maximum degradation temperatures presented in a Table inserted in Figure 5a. Additionally,
25
for GO–poly(AMPS–co–AM) are at higher temperatures, 212, Figures 5b–e and Figures 5f show the elemental distribution
26
292, and 406 °C. Also, the char yields of GO and the GO–poly map of C, O, N, and S and the corresponding mapping
27
(AMPS–co–AM), which is percentage weight of material left distribution of all elementals for GO–poly(AMPS–co–AM),
28
undecomposed after TGA analysis at maximum temperature respectively. We can see that all of the above elements are
29
800 °C in a nitrogen atmosphere, were found to be 20 and 30%, highly dispersed on GO, although the concentration of carbon
30
respectively (Figure 3). These observations indicate that GO– atoms is higher than the others (green areas).
31
poly(AMPS–co–AM) exhibits slower weight loss and thermal SEM images of pristine GO and GO–poly(AMPS–co–AM) are
32
decomposition rate as well as higher thermal stability than shown in Figure 6. It can be seen that the surface morphology
33
those of the pristine GO. of these two samples is different. The observed morphology in
34
The limiting oxygen index (LOI) is the minimum concen- Figure 6a shows that GO consists of crumpled and flake-like
35
tration of oxygen which will support combustion of a material. sheets. In contrast, GO–poly(AMPS–co–AM) has porous mor-
36
LOI can be calculated from van Krevelen and Hoftyzer phology (Figure 6b), which can be due to the entry of grafted
37
equation,[46] LOI = 17.5 + 0.4 CR where CR is the char yield. polymer during the functionalization of GO surface. In fact, flat
38
Using this equation, the calculated LOI values of the GO and and smooth morphology of graphene sheets was disturbed
39
the GO–poly(AMPS–co–AM) were found to be 25.5 and 29.5, during the functionalization processes. The porous micro-
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56 Figure 4. XRD patterns of GO (a) and GO–poly(AMPS–co–AM) (b).
57

ChemistrySelect 2019, 4, 7734 – 7744 7737 © 2019 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Full Papers

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
Figure 5. EDX spectrum (a), O element mapping (b), C element mapping (c), N element mapping (d), S element mapping (e), and the corresponding mapping
45 distribution of all elementals for GO–poly(AMPS–co–AM).
46
47
48
structure of GO–poly(AMPS–co–AM) certainly affects its superi- from π-π* transitions of C=C of GO sheets and a shoulder at
49
or water uptake and swelling properties. The SEM image of the around 270 nm can be assigned to the n-π* transition of C=O
50
recovered GO–poly(AMPS–co–AM) catalyst during repeated bonds.[52] The above mentioned bands in GO–poly(AMPS–co–
51
runs in the synthesis of 1,8-dioxo-octahydroxanthenes (in the AM) in comparison with those of GO have shifted due to the
52
following discussion) did not reveal any significant change in presence of functional groups. The band gap (Eg) of the GO and
53
the morphology of the reused catalyst, which proved its GO–poly(AMPS–co–AM) were estimated by the equation:
54
robustness (Figure 6c). (Ahυ)2 = B(hυ-Eg) where A is the absorption coefficient, hυ is
55
The UV–Vis spectra of the GO and GO–poly(AMPS–co–AM) the energy of photon, and B is a constant relevant to the type
56
were show in Figure 7a. The bands at around 200 nm arise of materials. Figure 7b shows the (Ahυ)2 versus hυ curve for GO
57

ChemistrySelect 2019, 4, 7734 – 7744 7738 © 2019 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Full Papers

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15 Figure 6. SEM images of GO (a) and GO–poly(AMPS–co–AM) (fresh (b) and five cycles (c).
16
17
18
tively, by extrapolation of the corresponding curves. Clearly,
19
the synthesized GO–poly(AMPS–co–AM) with a narrow band
20
gap (toward GO) can be better semiconductor than the pristine
21
GO and may have an application in optical fields and/or can be
22
also used as a photocatalyst for degradation of different
23
organic dye polluters.[53]
24
Origin of the fluorescence band of the GO arises from
25
electronic transitions between the antibonding and bonding
26
molecular orbitals (δ* ! n, π* ! n, and π* ! π). These
27
emission transitions are generally duo to functional groups,
28
such as –COH, –COOH, C=O, C–O–C, and C=C of GO sheets.[54]
29
Usually, an emission broad peak is observed when these peaks
30
overlap with each other.[55] The fluorescence emission bands of
31
the GO and GO–poly(AMPS–co–AM) were investigated upon
32
being excited at different wavelengths (Figure 8). The
33
fluorescence emission bands of GO–poly(AMPS–co–AM) are
34
located at 722, 800, and 880 nm at excitation wavelengths of
35
360, 400, and 440 nm, respectively (Figures 8a-c). As it can be
36
seen, in the GO–poly(AMPS–co–AM) the intensity fluorescence
37
peak is enhanced compared to that of GO. This is due to the
38
fact that after GO is grafted with poly(AMPS–co–AM), it
39
becomes enriched with C=O and SO3H groups. Actually, the
40
transition energy can change with the presence of functional
41
groups, lateral size, dopants, localized domains, and strain in
42
GO which may cause changes in the width, shape, and position
43
of fluorescence peak.[56] It is well known that the compounds
44
consisting of powerful fluorescence emission bands were used
45
in various fields such as in fluorescent biosensors[56] and in vivo
46
fluorescence imaging and bimodal photodynamic and photo-
47
thermal destruction of tumors.[57] Thus, according to our results,
48
the GO–poly(AMPS–co–AM) prepared in the present investiga-
49
tion is predicted to be an eco-friendly applicable industrial
50
material for the above mentioned fields.
51 Figure 7. Electronic spectra of the GO (1, blue line) and GO–poly(AMPS–co–
52 AM) (2, red line) (a) and (Ahυ)2–hυ curves (b).
53 Investigation of catalytic activity of GO–poly(AMPS–co–AM)
54
The catalytic ability of the synthesized GO–poly(AMPS–co–AM)
55
and GO–poly(AMPS–co–AM). The band gap of the GO and GO– was investigated in the synthesis of 1,8-dioxo-octahydroxan-
56
poly(AMPS–co–AM) were estimated of 5.9 and 4.7 eV, respec- thenes through the reaction of aldehydes with dimedone. At
57

ChemistrySelect 2019, 4, 7734 – 7744 7739 © 2019 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Full Papers

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23 Figure 9. The screening of the solvent nature (a) and catalyst amount (b) on
the reaction of 2,4-dichlorobenzaldehyde (1 equivalent) with dimedone (2
24
equivalents) after 14 min. catalyzed by GO–poly(AMPS–co–AM).
25
26
27
28
the best yields were obtained in EtOH/H2O (1:1) at room
29
temperature in the presence of 4 mol% (0.05 g) of catalyst.
30
With the optimal reaction condition in hand, the efficiency of
31
this approach was explored for the synthesis of a wide variety
32
of 1,8-dioxo-octahydroxanthenes. All of the reactions delivered
33
high to excellent product yields and accommodated a wide
34
range of aromatic aldehydes bearing both electron-donating
35
and electron-withdrawing substituents (entries 1–8,11,12). Also,
36
heteroaromatic aldehydes smoothly reacted with dimedone in
37
the presence of GO–poly(AMPS–co–AM) to afford their corre-
38
sponding 1,8-dioxo-octahydroxanthenes (entries 9,10). It is
39
worth to note that the reaction of 2,4-dichlorobenzaldehyde
40
(1 mmol) with dimedone (2 mmol) in the presence of poly
41
(AMPS-co-AM) (0.05 g)[42] as a catalyst in EtOH/H2O (1:1) at room
42
temperature gave only 30% of the corresponding 1,8-dioxo-
43
octahydroxanthene product after 2 h. A mixture of EtOH and
44
H2O (1:1) is believed to elevate the solubility of reactants as
45
well as the dispersity of the GO–poly(AMPS–co–AM), bearing
46
hydrophilic or hydrophobic groups. Also, when the above
47
mentioned reaction was carried out in the presence of GO
48
(0.05 g) as a catalyst the desired 1,8-dioxo-octahydroxanthene
49
Figure 8. Fluorescence emission spectra of GO (1, blue line) and GO–poly was obtained in low yield (about 20%) after 2 h. Therefore, the
50
(AMPS–co–AM) (2, red line) measured at excitation wavelengths of 360 (a), combination of GO and poly(AMPS-co-AM) is reasonably
51 400 (b), and 440 nm (c).
believed to possess much better catalytic performance than
52
both GO or poly(AMPS-co-AM) alone. In fact, in GO–poly
53
(AMPS–co–AM) catalyst, GO with high surface area affects the
54
first, to optimize the reaction conditions, a series of reactions of catalytic activity, the full accessibility of the reactants to the
55
2,4-dichlorobenzaldehyde (1 equivalent) with dimedone (2 active sites of the catalyst. All reactions were carried out under
56
equivalents) were studied. As it can be seen from Figures 9a,b, mild reaction conditions and no by-product was observed.
57

ChemistrySelect 2019, 4, 7734 – 7744 7740 © 2019 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Full Papers

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
Figure 10. The 1H NMR spectrum of the 9-(2,4-dichlorophenyl)-3,3,6,6-tetramethyl-2,3,4,5,6,7,8-octahydro-1H-1,8-xanthenedione.
31
32
33
34
Figure 10 shows the 1H NMR spectrum of the 9-(2,4-dichloro-
35
phenyl)-3,3,6,6-tetramethyl-2,3,4,5,6,7,8-octahydro-1H-1,8- xan-
36
thenedione.
37
Notably, when the GO–poly(AMPS–co–AM) was stirred in
38
EtOH/H2O (1:1) at room temperature for 2 h and then the
39
catalyst was filtered off, the filtrate showed no catalytic activity
40
for the synthesis of 1,8-dioxo-octahydroxanthenes. Also, the
41
filtrate was analyzed for its acid content, which showed a
42
negligible release of the acidic sites. These observations
43
confirm the true heterogeneity and stability of the catalytic
44
system as well as prove that the polymer precursors are
45
attached on the surface of the GO through covalent bonding
46
and not by physical adsorption.
47
We proposed a mechanism for the condensation reaction
48
of aromatic aldehydes with dimedone (Scheme 3). First, a
49
molecule of enolized dimedon can easily undergo a condensa-
50
tion reaction with aromatic aldehyde activated by GO–poly
51 Scheme 3. The proposed mechanism for the synthesis of 1,8-dioxo-octahy-
(AMPS–co–AM) giving intermediate 1, which undergoes dehy-
52 droxanthenes using GO–poly(AMPS–co–AM) catalyst.
dration to give the intermediate 2. Then, intermediate 2 reacts
53
further with a second molecule of dimedon to form uncyclized
54
adduct 2,2’-aryl-methylene bis(3-hydroxy-2-cyclohexene-1-one)
55
3, which affords the 1,8-dioxo-octahydroxanthenes after intra- arylaldehydes bearing the electron withdrawing groups (Ta-
56
molecular cyclodehydration. According to this mechanism, ble 1, entries 2–4) react with dimedone faster than those
57

ChemistrySelect 2019, 4, 7734 – 7744 7741 © 2019 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Full Papers

1 Table 1. Preparation of 1,8-dioxo-octahydroxanthenes catalyzed by GO–


poly(AMPS–co–AM)[a]
2
3
4
5
6
7 No. Ar Time [min] Yield [%] mp [°C] [Lit.][ref.]
8
1 17 95 198–203 (205–206)[23] Scheme 4. The synthesis of bis(indolyl)methanes using GO–poly(AMPS–co–
9
AM) catalyst.
10
11 2 10 98 223–225 (226–228)[23]
12
13 3 11 98 163–169 (170–171)[23]
14
15
4 14 93 215–219 (217–221)[35]
16
17
5 20 94 239–244 (244–246)[20]
18
19
6 15 96 231–234 (237–238)[23]
20
21
7 15 96 231–235 (240–241)[23]
22
23
24 8 14 94 251–254 (247–248)[23]
Figure 11. Recyclability of GO–poly(AMPS–co–AM) (4 mol%) in the reaction
25
of 2,4-dichlorobenzaldehyde (1 mmol) with dimedon (2 mmol) in EtOH/H2O
26 (1:1) at room temperature after 14 min.
9 17 94 64–68 (62–63)[25]
27
28
29 10 17 96 165–167 (160–165)[35]
For comparative purpose, a collection of recently reported
30
11 23 94 236–240 (242–245) [20] catalysts used for the preparation 1,8-dioxo-octahydroxan-
31
thenes was presented in Table 2. As it can be seen, GO–poly
32
12 19 93 213–217 (212–214)[20] (AMPS–co–AM) has a good efficiency compared to many of
33
those reported catalysts in view of time, yield, reusability, and
34
[a] All products are known compounds and were identified by comparison mild reaction condition.
35
of their physical and spectral data with those of the authentic samples.
36
37 Conclusions
38
In conclusion, GO–poly(AMPS–co–AM) was prepared by free
39
containing electron releasing groups (entries 11,12) because in radical graft copolymerization. FT–IR, XRD, EDX, TGA/DTA, and
40
the presence of the electron withdrawing groups the inter- Raman spectroscopy indicated that AM and AMPS monomers
41
mediate 2 is more reactive to the nucleophilic addition of were grafted onto the GO successfully, forming GO–poly
42
dimedone. (AMPS–co–AM). TG/DTA suggested that grafted poly(AMPS-co-
43
The above results encouraged us to examine the catalytic AM) onto GO improved the thermal stability of the GO and
44
activity of GO–poly(AMPS–co–AM) catalyst in the condensation enhanced LOI to good level of 29.5, which can make GO–poly
45
of aldehydes with indole. As shown in Scheme 4, GO–poly (AMPS–co–AM) as a fire retardant material. SEM showed that
46
(AMPS–co–AM) was effective in catalyzing the reaction, and bis GO–poly(AMPS–co–AM) has porous morphology, which can
47
(indolyl)methanes (BIMs) were produced in EtOH/H2O with affect its superior water uptake and swelling properties. The
48
excellent yields without any side reactions, which normally are band gap of 4.7 eV indicates that the synthesized GO–poly
49
observed under the influence of strong acids. (AMPS–co–AM) could be considered as a semiconductor which
50
From the viewpoint of economy and green chemistry, the may have an application in optical fields. According to our
51
reusability of catalysts for chemical transformations is especially results from the fluorescence study, the GO–poly(AMPS–co–
52
important. The reusability of the GO–poly(AMPS–co–AM) cata- AM) with powerful fluorescence emission bands is predicted to
53
lyst was investigated using a recycling experiment involving be useful for the optical applications. Finally, the obtained GO–
54
five consecutive runs of the condensation reaction of 2,4- poly(AMPS–co–AM) showed high catalytic activity for the
55
dichlorobenzaldehyde with dimedone. As shown in Figure 11, synthesis of 1,8-dioxo-octahydroxanthenes and bis(indolyl)
56
the recovered catalyst exhibited consistent catalytic activity. methanes in EtOH/H2O as green solvents under mild reaction
57

ChemistrySelect 2019, 4, 7734 – 7744 7742 © 2019 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Full Papers
[3] D. Li, W. Zhang, X. Yu, Z. Wang, Z. Su, G. Wei, Nanoscale 2016, 8, 19491–
1 Table 2. Comparison of the efficiencies of a number of different reported
19509.
catalysts with that of the GO–poly(AMPS–co–AM) in the reaction of
2 [4] X. Zhao, P. Zhang, Y. Chen, Z. Su, G. Wei, Nanoscale, 2015, 7, 5080–5093.
benzaldehyde with dimedone
3 [5] C. Shan, H. Yang, D. Han, Q. Zhang, A. Ivaska, L. Niu, Langmuir 2009, 25,
No. Reaction conditions Time Yield 12030–12033.
4
[min] [%][a] [6] X. Yu, W. Zhang, P. Zhang, Z. Su , Biosens. Bioelectron. 2017, 89, 72–84.
5 [7] M. Zhang, Y. Li, Z. Su, G. Wei, Polym. Chem. 2015, 6, 6107–6124.
6 1 DBSA, H2O, reflux 360 89[18] [8] D. R. Dreyer, S. Park, C. W. Bielawski, R. S. Ruoff, Chem. Soc. Rev. 2010, 39,
2 Amberlyst-15, CH3CN, reflux 300 92[19] 228–240.
7
3 SmCl3, solvent-free, 120 °C 540 98[20] [9] X. Chen, L. Yuan, P. Yang, J. Hu, D. Yang, J. Polym. Sci. 2011, 49, 4977–
8 4 Poly(AMPS-co-AA), solvent-free, 110 °C 25 85[21] 4986.
9 5 GO–SO3H, H2O, 80 °C 180 97[22] [10] D. Vuluga, J. M. Thomassin, I. Molenberg, I. Huynen, B. Gilbert, C. Jérôme,
10 6 [Et3NH][HSO4], solvent-free, 100 °C 20 94[23] M. Alexandre, C. Detrembleur, Chem. Commun. 2011, 47, 2544–2546.
7 CAN supported HY-zeolite, solvent-free, 90 88[24] [11] R. K. Layek, A. K. Nandi, Polymer 2013, 54, 5087–5103.
11
80 °C [12] M. Kaya, E. Demir, H. Bekci, J. Enzyme. Inhib. Med. Chem. 2013, 28, 885–
12 8 Mg–Al hydrotalcite, H2O, reflux 180 85[25] 893.
13 9 Mg(BF4)2, [bmim]BF4, 80 °C 30 97[26] [13] Y. Kwon, P. Song, J. H. Yoon, J. Ghim, D. Kim, B. Kang, T. G. Lee, J. A. Kim,
14 10 Sulfated zirconia, EtOH, 70 °C 480 95[27] J. A. Kim, J. K. Choi, I. K. Youn, H. K. Lee, S. H. Ryu, PLoS One 2014, 9,
11 LPCAS, solvent free, 100 °C 5 95[28] e108771–e108779.
15
12 Cu(II)-Fur-APTES/GO, EtOH/H2O, 50 °C 30 95[29] [14] D. Kumar, P. Sharma, H. Singh, K. Nepali, G. K. Gupta, S. K. Jain, F. Ntie-
16 13 [cmmim][BF4],[b] US,[c] r.t. 50 87[30] Kang, RSC Adv. 2017, 7, 36977–36999.
17 14 CAN, 2-propanol, US,[c], 50 °C 35 98[31] [15] T. S. Chitre, V. M. Khedkar, K. D. Asgaonkar, A. S. Dube, K. Shurpali, D.
18 15 [H-NMP] + [HSO4]–,[d] US,[c] H2O, r.t. 50 86[32] Sarkar, M. Kathiravan, Glob. Drugs Therap. 2017, 2, 1–5.
16 [Hbim]BF4,[e] MeOH, US,[c] r.t. 45 85[33] [16] M. Ahmad, T. A. King, D. K. Ko, B. H. Cha, J. Lee, J. Phys. D: Appl. Phys.
19
17 Carbon nanotube-BuSO3H, EtOH, r.t. 30 95[34] 2002, 35, 1473–1476.
20 18 Cellulose/Al2O3-[MeIm]Cl-XAlCl3,[f] EtOH, r.t. 25 91[35] [17] C. G. Knight, T. Stephens, Biochem. J. 1989, 258, 683–687.
21 19 GO–poly(AMPS–co–AM), EtOH/H2O, r.t. 17 95 [18] T. S. Jin, J. S. Zhang, J. C. Xiao, A. Q. Wang, T. S. Li, Clean, Synlett 2004, 5,
22 866–870.
[a] Isolated yield. [b] 1-Carboxymethyl-3-methylimidazolium tetrafluoro-
23 [19] B. Das, P. Thirupathi, I. Mahender, V. S. Reddy, Y. K. Rao, J. Mol. Catal. A:
borate. [c] Ultrasound irradiation. [d] 2-Pyrodidonium hydrogen sulfate. [e]
Chem. 2006, 247, 233–239.
24 1-n-butylimidazolium tetrafluoroborate. [f] Cellulose aluminum oxide
[20] A. Ilangovan, S. Malayappasamy, S. Muralidharan, S. Maruthamuthu,
25 composite supported imidazolium chloroaluminate ionic liquid.
Chem. Cent. J. 2011, 5, 81–87.
26 [21] B. Maleki, S. Barzegar, Z. Sepehr, M. Kermanian, R. Tayebee, J. Iran. Chem.
27 Soc. 2012, 9, 757–765.
[22] A. Shaabani, M. Mahyari, F. Hajishaabanha, Res. Chem. Intermed. 2014,
28
conditions. The solid acid catalyst is easily separated from 40, 2799–2810.
29 [23] Z. Zhou, X. Deng, J. Mol. Catal. A: Chem. 2013, 367, 99–102.
products by simple filtration. The high yields, short reaction
30 [24] P. Sivaguru, A. Lalitha, Chin. Chem. Lett. 2014, 25, 321–323.
times, and tolerance to a wide variety of reactants are other [25] R. Gupta, S. Ladage, L. Ravishankar, Chem. J. 2015, 1, 1–4.
31
obvious advantages of the present method. [26] B. Banerjee, ChemistrySelect 2017, 2, 8362–8376.
32
[27] S. S. Kahandal, A. S. Burange, S. R. Kale, P. Prinsen, R. Luque, R. V.
33 Jayaram, Catal. Commun. 2017, 97, 138–145.
34 Supporting Information summary [28] V. W. Godse, S. S. Rindhe, L. Kótai, S. R. Bembalkar, R. D. Ingle, R. P.
35 Pawar, Eur. Chem. Bull. 2017, 6, 1–4.
Detailed information about the experimental procedures, [29] Subodh, N. K. Mogha, K. Chaudhary, G. Kumar, D. T. Masram, ACS
36
Raman spectra of GO and GO–poly(AMPS–co–AM) and NMR Omega. 2018, 3, 16377–16385.
37 [30] A. N. Dadhania, V. K. Patel, D. K. Raval, C. R. Chim. 2012, 15, 378–383.
(13C and 1H) and IR spectra of some products are provided in
38 [31] N. Mulakayala, G. P. Kumar, D. Rambabu, M. Aeluri, M. B. Rao, M. Pal,
the supporting information. Tetrahedron Lett. 2012, 53, 6923–6926.
39
[32] H. Naeimi, Z. J. Sadat Nazifi, J. Ind. Eng. Chem. 2014, 20, 1043–1049.
40
[33] K. Venkatesan, S. S. Pujari, R. J. Lahoti, K. V. Srinivasan, Ultrason.
41 Acknowledgements Sonochem. 2008, 15, 548–553.
42 [34] K. Parvanak Boroujeni, Z. Heidari, R. Khalifeh, Acta Chim. Slov. 2016, 63,
The authors are thankful to the Research Council of Shahrekord
43 602–608.
University, Marvdasht Branch Islamic Azad University, Shahrekord [35] K. Parvanak Boroujeni, P. Tahani, Inorg. Nano-Met. Chem. 2017, 47, 1150–
44
Branch Islamic Azad University, and Research Institute of 1156.
45 [36] J. J. Yu, L. M. Wang, J. Q. Liu, F. L. Guo, Y. Liu, N. Jiao, Green Chem. 2010,
Petroleum Industry for their partial support of this work.
46 12, 216–219.
47 [37] S. Kumari, A. Shekhar, D. D. Pathak, Chem. Sci. Trans. 2014, 3, 652–663.
[38] R. M. Naidu Kalla, R. S. Karunakaran, M. Balaji, I. Kim, ChemistrySelect
48 Conflict of Interest 2019, 4, 644–649.
49 [39] Z. Tohidiyan, S. Hashemi, K. Parvanak Boroujeni, IET Nanobiotechnol.
The authors declare no conflict of interest.
50 2019, 13, 101–106.
51 [40] K. Parvanak Boroujeni, E. Rezazadeh Shirazi, Sulfur Silicon Relat. Elem.
52 Keywords: 1,8-dioxo-octahydroxanthenes · graphene · 2016, 191, 683–688.
[41] K. Parvanak Boroujeni, P. Ghasemi, Catal. Commun. 2013, 37, 50–54.
53 heterogeneous catalysis · poly(2-acrylamido-2-methyl-1-
[42] K. Parvanak Boroujeni, S. Hadizadeh, S. Hasani, A. Fadavi, M. Shahrokh,
54 propanesulfonic acid-co-acrylamide) · polymerizations Acta Chim. Slov. 2017, 64, 692–700.
55 [43] L. Zhang, H. Gao, Y. Liao, React. Funct. Polym. 2016, 104, 53–61.
56 [1] X. Zhao, Q. Zhang, D. Chen, Macromolecules 2010, 43, 2357–2363. [44] J. Yu, H. Zhang, Y. Li, Q. Lu, Q. Wang, W. Yang, Colloid Polym. Sci. 2016,
[2] H. Song, X. Zhang, Y. Liu, Z. Su , Chem. Rec. 2019, 19, 534–549. 294, 257–270.
57

ChemistrySelect 2019, 4, 7734 – 7744 7743 © 2019 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Full Papers
[45] Y. Sun, D. Shao, C. Chen, S. Yang, X. Wang, Environ. Sci. Technol. 2013, [53] Y. Yu, Q. Yang, X. Yu, Q. Lu, X. Hong, ChemistrySelect 2017, 2, 5578–5586.
1 47, 9904–9910. [54] S. J. Zhu, J. H. Zhang, C. Y. Qiao, S. J. Tang, Y. F. Li, W. J. Yuan, B. Li, L.
2 [46] Z. Rafiee, S. Mallakpour, R. Hatamvand, High Perform. Polym. 2013, 18, Tian, F. Liu, R. Hu, H. N. Gao, H. T. Wei, H. Zhang, H. C. Sun, B. Yang,
3 373–380. Chem. Commun. 2011, 47, 6858–6860.
[47] A. D. La Rosa, A. Recca, J. T. Carter, P. T. McGrail, Polymer 1999, 40, 4093– [55] M. Li, S. K. Cushing, X. J. Zhou, S. W. Guo, N. Q. Wu, J. Mater. Chem. 2012,
4
4098. 22, 23374–23379.
5 [48] H. P. Klug, L. E. Alexander, X-ray diffraction procedures for polycrystalline [56] P. Zheng, N. Wu, Chem. Asian J. 2017, 12, 2343–2353.
6 and amorphous materials, 2nd ed. John Wiley and Sons, New York, 1974. [57] P. Kalluru, R. Vankayala, C. S. Chiang, K. C. Hwang, Biomaterials 2016, 95,
7 [49] Y. J. Choi, S. Gurunathan, J. H. Kim, Int. J. Mol. Sci. 2018, 19, 710–733. 1–10.
[50] M. S. Jia, S. Y. Chang, X. L. Hua, L. S. Dong, H. Te, T. H. Bo, Chin. Phys. Lett.
8
2013, 30, 096101–096103.
9 [51] H. S. Moon, J. K. Park, Synth. Met. 1998, 92, 223–228.
10 [52] Z. Luo, Y. Lu, L. A. Somers, A. C. Johnson, J. Am. Chem. Soc. 2009, 131, Submitted: February 22, 2019
11 898–899. Accepted: June 26, 2019
12
13
14
15
16
17
18
19
20
21
22
23
24
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 2019, 4, 7734 – 7744 7744 © 2019 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Das könnte Ihnen auch gefallen