Vous êtes sur la page 1sur 6

Chemical Papers 68 (7) 983988 (2014)

DOI: 10.2478/s11696-013-0527-1

SHORT COMMUNICATION

Synthesis of melamineformaldehyde resin functionalised


with sulphonic groups and its catalytic activities
a

a Zhejiang

Lin-Jun Shao, b Gui-Ying Xing, a Chen-Ze Qi*

Key Laboratory of Alternative Technologies for Fine Chemicals Process, b School of Medicine,
Shaoxing University, Zhejiang Province 312000, China
Received 24 March 2013; Revised 27 September 2013; Accepted 3 November 2013

Spherical melamineformaldehyde resin (MFR) particles were synthesised using the condensation
reaction of melamine and formaldehyde with PEG-2000 as an additive. After thermal treatment
at 200 C and then sulphonation by chlorosulphuric acid, an MFR-supported strong acid catalyst
(SMFR) was prepared with an acidity of 3.23 mmol g1 . This new acid catalyst was evaluated
in the reactions of esterication and acetalisation, with the results indicating that this novel acid
catalyst was highly ecient in the traditional acid-catalysed reaction. Its high activity, stability and
reusability give it great potential for green chemical processes.
c 2013 Institute of Chemistry, Slovak Academy of Sciences
Keywords: solid acid, acidity, acetalisation, esterication

Acid catalysts are very important in the production of various chemicals (Anastas et al., 2002;
DeSimone, 2002). In recent decades, over 15 million
tons of sulphuric acid were consumed annually as an
unrecyclable homogeneous catalyst, leading to serious environmental pollution and increased purication costs. The heterogenisation of active catalysts in
the homogeneous phase is an eective way of increasing overall productivity and cost eectiveness. Moreover, heterogenisation can also aord important advantages in handling, separation and recycling procedures (Jothiramalingam & Wang, 2009; Clark, 2002;
Sabitha et al., 2010; Adam et al., 2012; Shaterian
& Rigi, 2012). Recently, a number of heterogeneous
acid catalysts have been developed to replace some
of the non-recyclable acids, especially liquids. Specifically, sulphonated carbonaceous materials have received much attention due to their high acidity and
catalytic activities (Xiao et al., 2010; Suganuma et
al., 2011; Shao et al., 2012a). However, to date, no
solid acid catalyst as active, stable and inexpensive as
sulphuric acid has been identied. Hence, preparation
of an ecient and inexpensive solid acid represents a
great challenge for chemists.
*Corresponding author, e-mail: qichenze@usx.edu.cn

Fig. 1. Schematic representation for synthesis of SMFR.

Melamineformaldehyde resin (MFR), which has


excellent physical and chemical stability, is widely
used in industry (Brooker & Mungin, 1983; Tutin,
1998; Baytekin, 2012). MFR contains many secondary
amine nitrogen atoms which can be used as an immobilisation site for acidic functionalities (Ji et al.,
1996; Qiu et al., 2002; Chehardoli et al., 2011; Hang
et al., 2011; Rezaei & Karami, 2011). In this study, a
novel MFR-supported acid catalyst (SMFR) was synthesised with chlorosulphuric acid as the sulphonating
reagent (Fig. 1). The catalytic activities of the SMFR

984

L. J. Shao et al./Chemical Papers 68 (7) 983988 (2014)

Fig. 2. SEM images of MFR without addition of PEG-2000 (a), MFR (b) and SMFR (c, d).

catalyst were evaluated in acetalisation and esterication.


All chemicals were commercial products of the
highest purity available and were used in the reactions
without further purication. Melamineformaldehyde
resin (MFR) was synthesised as follows: Melamine
(1,3,5-triazine-2,4,6-triamine, 32.0 g, 0.25 mol) in
19.8 g of a 30 mass % poly(ethylene glycol) (PEG
2000) aqueous solution in the presence of hexamethylenetetramine (1,3,5,7-tetraazatricyclo[3.3.1.13,7]
decane, 150 mg) as catalyst and 36 % aqueous solution of formaldehyde (60.0 g, 0.74 mol) were stirred
at 90 C for 5 hours. The reaction was quenched by
the addition of an HCl aqueous solution (10 mass %)
with a nal reaction solution pH value of 34. After
ltration, washing with deionised water and crushing,
the MFR was dried for 3 hours at 200 C and isolated
with a yield of 66.1 %. The MFR (1.0 g) was added
to a 10 mass % chlorosulphuric aciddichloromethane
solution (20 mL) and stirred gently at 10 C for 12
hours. After completion, the reaction solution was ltered and washed with dichloromethane, deionised water, ethanol and dried under reduced pressure at ambient temperature for 12 hours. The yield of SMFR
was 75.6 %.

The procedures employed in the SMFR-catalysed


acetalisation and esterication were as previously reported (Shao et al., 2012a, 2012b). Quantitative analysis was performed on a Shimadzu (GC-14B) gas
chromatograph. FT-IR/ATR spectra were recorded
on a FT-IR spectrometer (Nicolet, Nexus-470, USA)
equipped with accessories to attenuate total reection.
The morphologies of the samples were characterised
using a scanning electron microscope (SEM) (Jeol,
Jsm-6360lv, Japan). Samples for SEM were sputtered
with a 2030 layer of Au to render them conducA
tive. The diameters of the particles were determined
from the SEM images. The elemental analysis was performed on a EuroEA 3000 from Leeman, USA.
Poly(ethylene glycol) (PEG) has been widely
used as a dispersant to synthesise regular inorganic
nanoparticles (Yan et al., 2008; Gz ak et al., 2009;
o u
Choi et al., 2010). In the present study, PEG was initially used as an additive to synthesise MFR particles.
Fig. 2, shows that the addition of PEG-2000 could promote the formation of regular spherical MFR particles
with diameters of (3.97 0.64) m. The specic surface area of MFR also increased from 1.56 m2 g1 to
8.09 m2 g1 after the addition of PEG-2000.
Fig. 3 shows the FT-IR spectra of the MFR

985

L. J. Shao et al./Chemical Papers 68 (7) 983988 (2014)

Table 1. Acetalisation of various carbonyl compounds and diolsa


Conversion/%b

Yield/%b

98

98

99

99

93

93

86

86

99

99

89

88

99

99

90

88

Entry

Substrate

Product

a) Catalytic conditions: 20 mg of SMFR, 20 mmol of cyclohexanone, 24 mmol of ethylene glycol, 5 mL of cyclohexane, reux, 2
hours, with DeanStark apparatus; b) conversion and yield were determined by GC based on carbonyl compounds.

Fig. 3. FT-IR spectra of MFR and SMFR.

and SMFR. The absorption bands at 1626 cm1 ,

1334 cm1 and 786 cm1 assigned to CN, CN

and the triazine ring clearly indicated the synthesis


of MFR. The appearance of the absorption band at
1043 cm1 shows that the sulphonic acid groups were
successfully immobilised on the MFR (Liang et al.,
2010).
The elemental analysis gave the results: N: 42.8 %;
C: 29.2%; H: 4.5%; S: 3.1%, indicating that the
amount of sulphonic acid groups was 0.99 mmol g1 .
Hence, the acidity of SMFR was expected to be
0.99 mmol g1 . In eect, the acidity of the SMFR
was 3.23 mmol g1 , as determined by neutralisation

titration (Margelefsky et al., 2008). The titration was


carried out as follows: MFR or SMFR (40 mg) and
2 M aqueous NaCl (4 mL) were stirred at ambient
temperature for 24 hours. The solids were removed
by ltration and washed with water (4 2 mL). The
combined ltrate was titrated with 0.01 M NaOH using phenol red as the indicator. Accordingly, other
acidic groups besides the sulphonic groups would appear to be present in the SMFR. As the MFR contains
many secondary amine nitrogen atoms, which not only
can be used as the sites to immobilise the sulphonic
groups, but also can adsorb the H+ in reaction solution, the chlorosulphuric acid can decompose into HCl
during the sulphonation step. Therefore, the physical
adsorption of H+ on SMFR might be a signicant factor in the high acidity of SMFR.
Acetalisation is an important strategy in protecting carbonyl groups. The eects of reaction time on
the conversion and yield were investigated (Fig. 4),
indicating that the catalyst was very ecient in the
acetalisation. The yield was 90 % after reacting for
half an hour, and achieved the maximum at 2 h.
Various carbonyl compounds and diols were employed to evaluate the catalytic activities of the
SMFR. Table 1 shows that all aldehydes and ketones could be converted to the corresponding 1,3dioxolanes with moderate to good yields (Entries 1, 3,
5 and 7). As a seven-membered ring is not as stable as
a ve-membered ring, the yields are lower for certain
substrates (Entries 4, 6 and 8) using butane-1,4-diol.

986

L. J. Shao et al./Chemical Papers 68 (7) 983988 (2014)

Table 2. Esterication of acetic acid with various alcoholsa


Conversion/%b

Yield/%b

91

91

93

93

90

88

89

89

Entry

Substrate

Product

a) Catalytic conditions: 50 mg of catalyst, 20 mmol of alcohol, 24 mmol of acetic acid, reux, 3 h; b) conversion and yield were
determined by GC based on carbonyl compounds.

Fig. 4. Time-dependence of SMFR-catalysed acetalisation of


benzaldehyde with ethylene glycol, conversion ( ), yield
( ).

Fig. 5. Reuse of SMFR catalyst, conversion (light grey bar),


yield (dark grey bar).

Examination of entries 14 in Table 1 shows that the


aliphatic aldehydes were more active in acetalisation
than the aromatic aldehydes.

In addition to acetalisation, esterication was investigated using SMFR as the catalyst. It is recognised
that water as a by-product could signicantly decrease
the conversion and yield of esterication (Peters et al.,
2006; Bhorodwaj & Dutta et al., 2011). Interestingly,
using SMFR as catalyst, there was no need to remove
the water continuously from the reaction mixture using a DeanStark apparatus. Table 2 shows that the
alcohols were converted to the corresponding esters
with conversions of 89 % or greater, and with yields
of at least 88 %. These results demonstrated the eectiveness of the catalyst in the esterication reaction.
The recovery and reuse of heterogeneous catalysts
could greatly facilitate the purication of products
and increase overall yield of the product and reduce
production costs, resulting in a greener and more
sustainable chemical transformation process. The catalytic activity of the recovered SMFR catalyst was
carefully investigated through the acetalisation of cyclohexanone with ethylene glycol. Fig. 5 shows that
the catalytic activity of SMFR did not changed even
after four runs, indicating the excellent stability of the
SMFR catalyst.
The heterogeneous catalyst is usually less active
than the corresponding homogeneous one, due to the
less favourable kinetics of the biphasic catalytic system. A comparison of the catalytic activity of the
SMFR with traditional homogeneous catalysts was
performed using the acetalisation of benzaldehyde
with ethylene glycol. Table 3 shows that the catalytic
activity of SMFR is slightly lower than that of the homogeneous H2 SO4 and p-toluene sulphonic acid catalyst. However, the SMFR could conveniently be recovered and reused, which would avoid the purication
step and reduce overall costs.
In summary, a highly active and recyclable solid
acid catalyst was synthesised by sulphonation of MFR.
This new solid acid is as eective as the homogeneous
acid catalysts commonly used for acetalisation and es-

987

L. J. Shao et al./Chemical Papers 68 (7) 983988 (2014)

Table 3. Comparison of dierent catalysts for the acetalisation of benzaldehyde with ethylene glycola
Entry

Catalyst

Amount/mgb

Conversion/%c

Yield/%c

1
2
3
4

H2 SO4
Benzene sulphonic acid
Amberlyst
SMFR

3.2
10.2
13.7
20.0

99
77
99
91

99
77
99
91

a) Catalytic conditions: 20 mmol of cyclohexanone, 24 mmol of ethylene glycol, reux, 30 min, with DeanStark apparatus; b)
amounts of H+ for all catalysts were 6.46 105 mol; c) conversion and yield were determined by GC based on benzaldehyde.

terication. The remarkable stability and recyclability


of this unique environmentally-benign heterogeneous
acid catalyst render it attractive for large-scale industrial applications.
Acknowledgements. The authors wish to acknowledge the
nancial support received from the Zhejiang Science and Technology Innovation Team of Zhejiang Province Science and
Technology Hall (2012R10014-16) and Shaoxing University.

References
Adam, F., Batagarawa, M. S., Hello, K. M., & AI-Juaid, S. S.
(2012). One-step synthesis of solid sulfonic acid catalyst and
its application in the acetalization of glycerol: Crystal structure of cis-5-hydroxy-2-phenyl-1,3-dioxane trimer. Chemical
Papers, 66, 10481058. DOI: 10.2478/s11696-012-0203-x.
Anastas, P. T., & Kirchho, M. M. (2002). Origins, current
status and future challenges of green chemistry. Accounts of
Chemical Research, 35, 686694. DOI: 10.1021/ar010065m.
Baytekin, M. T. (2012). Monoliths from step-growth polymerization reactions. Ume Sweden: Ume University.
a,
a
Bhorodwaj, S. K., & Dutta, D. K. (2011). Activated clay supported heteropoly acid catalysts for esterication of acetic
acid with butanol. Applied Clay Science, 53, 347352. DOI:
10.1016/j.clay.2011.01.019.
Brooker, L. G., & Mungin, H. (1983). US Patent No. 4,405,690.
Washington, DC, USA: US Patent Oce.
Chehardoli, G., Zolgol, M. A., Azimi, S. B., & Alizadeh, E.
(2011). Melamine-(H2 SO4 )3 and PVP-(H2 SO4 )m as solid
acids: Synthesis and application in the rst mono- and dinitration of bisphenol A and other phenols. Chinese Chemical Letters, 22, 827830. DOI: 10.1016/j.cclet.2011.01.021.
Choi, S. G., Moon, Y. M., & Jung, H. K. (2010). Luminescent
properties of PEG-added nanocrystalline YVO4 :Eu3+ phosphor prepared by a hydrothermal method. Journal of Luminescence, 130, 549553. DOI: 10.1016/j.jlumin.2009.10.029.
Clark, J. H. (2002). Solid acids for green chemistry. Accounts of
Chemical Research, 35, 791797. DOI: 10.1021/ar010072a.
DeSimone, J. M. (2002) Practical approaches to green solvents.
Science, 297, 799803. DOI: 10.1126/science.1069622.
Gzak, F., Kseolu, Y., Baykal, A., & Kavas, H. (2009).
o u
o g
Synthesis and characterization of Cox Zn1x Fe2 O4 magnetic nanoparticles via a PEG-assisted route. Journal of
Magnetism and Magnetic Materials, 321, 21702177. DOI:
10.1016/j.jmmm.2009.01.008.
Hang, Z. S., Tan, L. H., Gao, X. M., Ju, F. Y., Ying, S. J., &
Xu, F. M. (2011). Preparation of melamine microbers by
reaction electrospinning. Materials Letters, 65, 10791081.
DOI: 10.1016/j.matlet.2011.01.010.
Ji, S., Crews, G. M., Pittman, C. U., Jr., Wang, Y., & Ran,
R. (1996). Ammelinemelamineformaldehyde resins. Preparation and properties. Journal of Polymer Science Part A:

Polymer Chemistry, 34, 25432561. DOI: 10.1002/(sici)10990518(19960930)34:13<2543::aid-pola1>3.0.co;2-t.


Jothiramalingam, R., & Wang, M. K. (2009). Review of recent
developments in solid acid, base and enzyme catalysts (heterogeneous) for biodiesel production via transesterication.
Industrial & Engineering Chemistry Research, 48, 6162
6172. DOI: 10.1021/ie801872t.
Liang, X. Z., Xiao, H. Q., Shen, Y. M., & Qi, C. Z. (2010). Onestep synthesis of novel sulfuric acid groups functionalized
carbon via hydrothermal carbonization. Materials Letters,
64, 953955. DOI: 10.1016/j.matlet.2010.01.070.
Margelefsky, E. L., Bendjriou, A., Zeidan, R. K., Dufaud,
V., & Davis, M. E. (2008). Nanoscale organization of thiol
and arylsulfonic acid on silica leads to a highly active and
selective bifunctional, heterogeneous catalyst. Journal of
the American Chemical Society, 130, 1344213349. DOI:
10.1021/ja804082m.
Peters, T. A., Benes, N. E., Holmen, A., & Keurentjes,
J. T. F. (2006). Comparison of commercial solid acid
catalysts for the esterication of acetic acid with butanol. Applied Catalysis A: General, 297, 182188. DOI:
10.1016/j.apcata.2005.09.006.
Qiu, X. Q., Yi, C. H., Yang, D. J., & Ouyang, X. P. (2002).
Synthesis of sulfonated melamine urea formaldehyde resins.
Modern Chemical Industry, 22, 2427.
Rezaei, R., & Karami, M. (2011). Microwave promoted rapid dehydration of aldoximes to nitriles using melamineformaldehyde resin supported sulphuric acid in dry media. Chinese
Chemical Letters, 22, 815818. DOI: 10.1016/j.cclet.2011.01.
008.
Sabitha, G., Prasad, M. N., Ramesh, M., & Yadav, J. S.
(2010). Silica sulfuric acid as a reusable heterogeneous
catalyst for the diastereoselective Mukaiyama aldol reaction of 2-(trimethylsilyloxy)furan: Facile synthesis of butenolides. Monatshefte fr Chemie, 141, 12451248. DOI:
u
10.1007/s00706-010-0388-z.
Shao, L. J., Du, Y. J., Xing, G. Y., Lv, W. X., Liang, X. Z., &
Qi, C. Z. (2012a). Polyacrylonitrile ber mat supported solid
acid catalyst for acetalization. Monatshefte fr Chemie, 143,
u
11991203. DOI: 10.1007/s00706-011-0706-0.
Shao, L. J., Xing, G. Y., He, L. Y., Chen, J., Xie, H. Q., Liang,
X. Z., & Qi, C. Z. (2012b). Sulfonic groups functionalized
preoxidated polyacrylonitrile nanobers and its catalytic applications. Applied Catalysis A: General, 443444, 133137.
DOI: 10.1016/j.apcata.2012.07.034.
Shaterian, H. R., & Rigi, F. (2012). Acetalization of carbonyl
compounds as pentaerythritol diacetals and diketals in the
presence of cellulose sulfuric acid as an ecient, biodegradable and reusable catalyst. Chinese Journal of Chemistry,
30, 695698. DOI: 10.1002/cjoc.201280002.
Suganuma, S., Nakajima, K., Kitano, M., Kato, H., Tamura, A.,
Kondo, H., Yanagawa, S., Hayashi, S., & Hara, M. (2011).
SO3 H-bearing mesoporous carbon with highly selective catalysis. Microporous and Mesoporous Materials, 143, 443450.
DOI: 10.1016/j.micromeso.2011.03.028.

988

L. J. Shao et al./Chemical Papers 68 (7) 983988 (2014)

Tutin, K. K. (1998). US Patent No. 5,710,239. Washington, DC,


USA: US Patent Oce.
Xiao, H. Q., Guo, Y. X., Liang, X. Z., & Qi, C. Z. (2010).
One-step synthesis of a novel carbon-based strong acid catalyst through hydrothermal carbonization. Monatshefte fr
u
Chemie, 141, 929932. DOI: 10.1007/s00706-010-0332-2.

Yan, A. G., Liu, X. H., Qiu, G. Z., Wu, H. Y., Yi, R., Zhang, N.,
& Xu, J. (2008). Solvothermal synthesis and characterization
of size-controlled Fe3 O4 nanoparticles. Journal of Alloys and
Compounds, 458, 487491. DOI: 10.1016/j.jallcom.2007.04.
019.

Vous aimerez peut-être aussi