You are on page 1of 8

Progress in Organic Coatings 102 (2017) 217–224

Contents lists available at ScienceDirect

Progress in Organic Coatings


journal homepage: www.elsevier.com/locate/porgcoat

Alkyd resins based on hyperbranched polyesters and PET waste for


coating applications
Nawal E. Ikladious, Jeannette N. Asaad, Hassan S. Emira, Samia H. Mansour ∗
Department of Polymers and Pigments, National Research Center, 33 El Bohouth st. (former El Tahrir st.) Dokki, P.O. 12622, Giza, Egypt

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

Article history: Medium oil alkyd resins based on the glycolyzed PET waste and different generations of aliphatic hyper-
Received 13 September 2016 branched polyesters with linseed and sunflower oil fatty acids, which are of interest to the coating
Accepted 14 October 2016 industry, were investigated. Different resin compositions were synthesized and characterized by IR and
Available online 22 November 2016 1
H NMR spectroscopy. The effect of oil fatty acid type, glycolyzed products, functionality of the poly-
ols on thermal and film performance properties was studied. The onset of decomposition of all samples
Keywords:
displayed at values above 200 ◦ C which render them suitable for the processing requirements of most
Alkyd resins
conventional coating applications. All films had good adhesion, bending, impact and ductility. The gloss
Coatings
Hyperbranched polyesters
and hardness increased with increasing the degree of branching and consequently, the molecular weight
PET waste in the resin backbone. The presence of G3 as polyol enhanced the thermal stability and film properties of
Oil fatty acids coatings.
© 2016 Elsevier B.V. All rights reserved.

1. Introduction them suitable for several applications, especially those coatings


with low volatile organic compounds (VOC) [21–24].
The non-biodegradability of polyethylene terephthalate (PET) The application of hyperbranched polyesters, the most studied
creates huge amount of waste and disposal materials which results class of hyperbranched polymers, in different coating technologies
in serious environmental problems. The increasing awareness of such as UV curable coatings, powder coatings and in blends with
environmental pollution is the most important cause for recycling epoxy resins have been presented [22,25,26].
and reprocessing of PET waste. Therefore, the necessity of finding Alkyd resins are the most widely used polymers for a broad
a sustainable and simple economic route for PET waste recycling is range of paint and coating applications [27]. Hyperbranched alkyds
an important practice in contribution to raw materials and energy [28–31] have been reported as promising resins for development of
conservation [1,2]. low VOC or nonpolluting coatings. They show significantly lower
The chemical recycling of PET waste and conversion into valu- viscosities and rapid air drying property with enhanced outdoor
able products might be of great interest. Glycolysis, hydrolysis, durability compared to the conventional high solid alkyds and can
methanolysis, aminolysis and ammonolysis are the most recycling be utilized for the production of high-gloss decorative paints with
strategies for this material [3–8]. Production of specialized products low content of organic solvent and good application properties.
such as saturated and unsaturated polyester resins, polyurethanes, Using the glycolyzed products of PET in production of alkyd
coating materials and additives, using PET waste, has been investi- resins is of great importance from the economic and environmen-
gated [9–17]. tal point of view. Long oil alkyd resins from phthalic anhydride,
Over the last few years, a large number of hyperbranched poly- depolymerization product of PET waste, glycerin, sunflower oil
mers has been reported in the literature [18–20]. Hyperbranched fatty acids and glycol were prepared by Güclü and Orbay [15]. Film
polymers with various end groups and architectures provide very properties and thermal stability of these alkyd resins were better
intrinsic properties of great potential values of application. Their than the properties of reference alkyd resin. Torlakoglu and Güclü
lower viscosities compared to the analogues linear ones render [32] reported that, PET-based alkyd-amino resins having desirable
properties can be prepared by changing alkyd/amino resin ratios
and type of amino resins. Polyester polyol derived from PET waste
for coating application on mild steel was studied by Purohit et al.
[33]. Alkyd resins derived from recycled PET waste have also been
∗ Corresponding author.
E-mail address: s.mansour26@hotmail.com (S.H. Mansour).

http://dx.doi.org/10.1016/j.porgcoat.2016.10.015
0300-9440/© 2016 Elsevier B.V. All rights reserved.
218 N.E. Ikladious et al. / Progress in Organic Coatings 102 (2017) 217–224

investigated by Atta et al. [34] as corrosion protective coating for


carbon steel.
In continuation of our study on either chemical recycling of PET
[3–5,9,14] or preparation of hyperbranched polyesters [31,35–38],
the present study deals with the combination of these two research
areas. Different alkyd resins based on glycolyzed PET together with
hyperbranched polyesters and different vegetable oil fatty acids,
that appear to have potential applications in coating industry, have
been studied with respect to methods of preparation, their proper-
ties, advantages over conventional counterparts and applications.

2. Experimental

2.1. Materials

PET (soft drink bottles) with number average molecular weight


26,103 g mol−1 and intrinsic viscosity 0.8 dl g−1 were used in
the glycolysis process. 2,2-Bis(hydroxymethyl) propionic acid
(bis-MPA), 1,3,5-tris(2-hydroxyethyl) isocyanurate (THECA), and p-
toluene sulfonic acid (PTSA) were purchased from Sigma–Aldrich,
Germany. Octa-Soligen155 was kindly provided as a gift by OMG
Borchers GmbH OM Group, Schönebeck, Germany. The oil fatty
acids were also kindly supplied as a gift by Oleon GmbH, Emmerich,
Germany. All materials were used as received without further
purification.

2.2. Methods

2.2.1. Synthesis of hyperbranched polyesters (HBPs)


Several generations (G1-G3) of hyperbranched polyesters were Scheme 1. The proposed structure of the hyperbranched polyester G1.
synthesized by the polycondensation of stoichiometric amounts
of the monomer bis-MPA, corresponding to each generation with
1 mol of THECA as a trifunctional core [36]. The monomer to core
ratio varied between 3 and 21. The polymerization reaction was
carried out in bulk at a maximum temperature 160 ◦ C. p-Toluene
sulfonic acid (PTSA), 0.5 wt% based on bis-MPA monomer, was used
as acid catalyst in all reactions. The yield of all obtained polymers
was above 97%. The synthesis procedure for the hyperbranched
polyester, for example, G1 is outlined in Scheme 1.

2.2.2. Glycolysis reactions


PET waste was depolymerized using propylene glycol (PG), and
diethylene glycol (DEG), in the presence of zinc acetate as transes-
terification catalyst. The concentration of the catalyst was 0.5 wt.%
based on the weight of PET. The weight ratio of PET waste to gly-
col was 54/46 wt.%, (corresponding to PET/glycol molar ratio of
1:1.8). The mixture of glycols (PG: DEG) was charged in the ratio of
50:50 wt.%, respectively. The glycolysis was carried out at 220 ◦ C for Fig. 1. Molecular weight distribution of the glycolyzed products.

10 h under reflux in a nitrogen atmosphere. The glycolyzed prod-


ucts were analyzed for hydroxyl value and was 526.85 g/mol. It was PTSA, 0.5 wt% based on acid moieties, was used as catalyst in all
also analyzed by GPC for molecular weight distribution. Fig. 1 shows reactions. The alkyd constant (K) ranged from 1.1 to 1.27 and the
a wide monomodal behavior with polydispersity index (PDI) 1.7. ratio of basic equivalents to acid equivalents (R) from 1.1 to 1.2.
The number and weight average molecular weight (Mn and Mw) The reaction of the third generation (G3) of the hyperbranched
obtained by GPC analysis were 255 and 435 g/mole, respectively. polyester with sunflower oil fatty acids is presented as an example:
The glycolyzed products were also identified by IR and 1 HNMR. Sunflower oil fatty acids (58.8 g, equivalent to 0.21 mol
1 H NMR (500 MHz, CDCl , ␦ppm): 1.3(CH -PG moiety), 3.1–4.7 and 55.09 wt%), phthalic anhydride (6.66 g, 0.045 mol, 6.24 wt%),
3 3
(CH2 -DEG, EG, and PG moieties), 8.0 (arom. H of terephthalic acid). glycolyzed products (12.78 g, 0.06 mol, 11.97 wt%), G3 (28.5 g,
The depolymerization process for PET waste is outlined in 0.15 mol) and a catalytic amount of PTSA (0.3273 g) were placed
Scheme 2. in a three necked flask, equipped with a condenser, thermome-
ter and argon inlet. The reaction vessel was evacuated for 10 min,
2.2.3. Synthesis of alkyd resins flushed with argon, and then immersed in a preheated oil bath
Medium oil alkyd resins were synthesized by reacting phthalic with magnetic stirrer at 180 ◦ C. The reaction mixture was further
anhydride and the glycolyzed products (0.8 mol) together with heated to 220 ± 5 ◦ C and maintained at this temperature until the
0.2 mol of the prepared hyperbranched polyesters, or glycerin or acid value decreased as desired (7–28 mg KOH/g). A stream of argon
pentaerithritol with either linseed or sunflower oil fatty acids. The was applied to continuously remove the water formed during the
reaction was carried out in bulk at maximum temperature 220 ◦ C. reaction. The acid value was monitored during the reaction and
N.E. Ikladious et al. / Progress in Organic Coatings 102 (2017) 217–224 219

Scheme 2. Depolymerization of PET waste.

Scheme 3. The proposed structure of the alkyd resin.

determined by end-group analysis of samples collected at different 2.2.4. Characterization


time intervals. Fourier transform infrared (FTIR) spectra were recorded on a
1 H NMR (500 MHz, CDCl , ␦ ppm) 0.85–1.4 (CH -fatty acid, −bis- FT/IR JASCO 6100 (Tokyo, Japan). 1 H NMR spectra were run at
3 3
MPA and PG), 1.7–2.8 (CH2 and CH-fatty acid), 4.2 (CH2 OR), 3.6 500 MHz on a JEOL ECA-500 NMR spectrometer using CDCl3 as
(CH2 N of the central core), 5.30 (CH CH), 7.5(4H of phthalic anhy- solvent. Thermogravimetric measurements (TG) were performed
dride), 8.0 (4H of terephthalic acid). using TA Instruments SDT-Q600 Simultaneous TGA/DSC (USA)
The preparation of the alkyd resins can be generally described under nitrogen atmosphere at heating rate 10 ◦ C/min. GPC was
by Scheme 3. carried out using Agilent1100 Series Refractive Index Detector,
In a similar manner, resins of different compositions of alkyd (Germany), a set of two of high resolution PL gel 5 ␮m columns
resins (Table 1) were prepared by changing either the fatty acid (103, 105 Å). Tetrahydrofuran was used as mobile phase at a
or the generation of the hyperbranched polyester. The yield of all flow rate 1 ml min−1 .The columns were calibrated by means of
obtained alkyd resins was above 80%. The resulting products were polystyrene standards covering the molecular weight range of
identified by IR, and 1 H NMR. 103 –3 × 106 g mol−1 .
220 N.E. Ikladious et al. / Progress in Organic Coatings 102 (2017) 217–224

Table 1
The compositions of the prepared alkyd resins.

Constituents (Weight%) Alkyd Resin Number

1 2 3 4 5 6 7 8 9

Linseed oil fatty acid 47.23 50.76 47.26 55.09 53.74


Sunflower oil fatty acid 53.92 46.24 50.18 55.09
Phthalic anhydride 14.86 13.42 10.77 6.24 3.58 3.56 12.70 9.56 6.24
Glycolyzed products 34.21 30.89 24.8 11.97 11.32 11.28 24.26 16.96 11.97
Glycerol 3.7
Pentaerethritol 4.93
G1 17.17 31.36 31.24 16.8
G2 23.3
G3 26.7 26.7
R 1.2 1.2 1.2 1.1 1.1 1.1 1.1 1.1 1.1
K 1.27 1.25 1.23 1.1 1.34 1.34 1.17 1.11 1.1
Acid no. 3 25 18 26 8 7 28 26 15
Oil length 50.6 51.47 51 58 56 56 49 53 58

Table 2 terephthalic acid, produced through trans-esterification reaction of


Thermal behaviour data of the prepared alkyd resins.
PET, were used together with the unreacted glycols in the synthesis
Alkyd Resin Number T10 [◦ C] T50 [◦ C] Residue at 550 ◦ C of the alkyd resins.
1 325 399 8.7 The preparation and characterization of G1–G3 of the used
2 310 400 9.5 hyperbranched polyesters is described in details elsewhere [36].
3 324 400 3.2 The number of hydroxyl end groups of G1, G2 and G3, determined
4 326 418 4.4 by 1 H NMR analysis, was 6, 11 and 17, respectively.
7 309 393 3
In order to reach a goal sought in the present investigation
8 318 399 2.1
9 348 410 1 and to prepare alkyd resins with improved film properties which
Cured film of 9 243 388 11.4 make them suitable for some eco-friendly industrial coatings, nine
medium oil alkyd resins were synthesized. Studied was the effect of
oil fatty acid type, glycolyzed products, functionality of the polyols
2.2.5. Film preparation and properties
and the resulting branched structure on physicochemical, thermal
Resin samples (80 wt.%) were mixed with 20 wt.% turpentine
and film performance properties. The different compositions of the
and toluene (1:1) and 4.5 wt.% Octa-Soligen 155 as drier. The con-
prepared alkyd resins are shown in Table 1.
tent of drier in the formulated coatings was based on the resin
content. The resulted mixtures were stirred to give homogeneous
solutions and then stored overnight prior to their application either
on glass or metal plates. The films were cast on the test panels with 3.1. Characterization
60 ␮ applicator. In order to accelerate curing, they were left 3 h in
an oven at 90 ◦ C, followed by one week at room temperature. In order to elucidate the structural units in the glycolyzed prod-
Hardness was determined using pendulum hardness tester ucts and the prepared alkyd resins all samples were characterized
according to ASTM D 4366-14. Adhesion was measured using cross- in view of their IR and 1 H NMR spectra, as shown in Figs. 2–5 . The
hatch cutter according to ASTM D 3359-2009. Cupping test was characteristic bands and signals of IR and 1 H NMR spectra of the
used to determine the ductility according to ISO 1520-2006. It con- glycolyzed products were observed in all spectra of the prepared
sists of measuring the minimum depth at which the coating cracks alkyd resins.
and/or becomes detached from the substrate. Gloss measurements IR spectra of all alkyd resins (Figs. 2 and 3) were quite similar
according to ASTM D 523-2008 were performed at incidence angle and illustrate the stretching frequencies of the hydroxyl group and
20◦ . Bending and impact resistance were determined according to the ester carbonyl group at 3500 cm−1 , and 1735 cm−1 respectively.
ASTM D 522 and ASTM D 2794-2010 respectively. Symmetrical, asymmetrical, stretching and bending frequencies
of the CH2 group appear at 2855, 2925, and 1460 cm−1 respec-
3. Results and discussion tively. The band at 3006 cm−1 , assigned to fatty acids unsaturation,
indicates alkyd resins formation. Stretching frequencies of C C at
The depolymerization products of PET waste is considered as 1580 and 1607 cm−1 and the CH out of plane bending frequency at
an inexpensive potential raw material for the production of alkyd 729 cm−1 suggest the presence of aromatic residues in the resins.
resins. In the present study, glycolysis of the PET waste was per- The bands at around 1129 and 1273 cm−1 , are assigned to stretching
formed using a mixture of propylene glycol (PG), and diethylene and bending frequencies of C O C, −COOR and CH, respectively.
glycol (DEG) in a ratio 50:50 wt.%. The hydroxyl terminated ester of Comparing the IR spectra of alkyd resin 9 and that of its cured film,

Table 3
Film properties of the prepared alkyd resins.

Properties Alkyd Resin Number

1 2 3 4 5 6 7 8 9

Hardness 12 12 17 24 16 15 15 15 21
Bending pass
Ductility >6 >6 >6 >6 >6 >6 >6 >6 >6
Adhesion 5B 5B 5B 5B 5B 5B 5B 5B 5B
Impact resistance (joule) >18 >18 >18 >18 >18 >18 >18 >18 >18
Gloss 20◦ 71.6 76 106.4 119.8 117.8 106.9 88 89 110
N.E. Ikladious et al. / Progress in Organic Coatings 102 (2017) 217–224 221

Fig. 3. The FTIR spectra of alkyd resins 7, 8, 9 and the cured film of resin 9.

depicted in Fig. 3, the absence of the band at 3006 cm−1 illustrates


complete curing of the sample.
1 H NMR spectra of the prepared alkyd resins (Figs. 4 and 5)

had similar features and showed the four aromatic protons of


terephthalic acid residue at 8.0 ppm which further confirm the
incorporation of the glycolyzed products in the resins. The two ill-
defined distorted doublets at 7.7 and 7.5 ppm indicate the presence
Fig. 2. The FTIR spectra of glycolyzed products and alkyd resins 1, 2, 3 and 4.
of an AB system for the aromatic protons of phthalic anhydride
residue. The peak at 5.30 ppm, stands for the CH CH protons of the
fatty acids, is an evidence of the formation of the alkyd resins.
The complex pattern at 4.6–3.5 ppm is due to the superposition
of resonances related to the methylene groups (CH2 -OR). The sig-
222 N.E. Ikladious et al. / Progress in Organic Coatings 102 (2017) 217–224

1
Fig. 5. H NMR spectra of alkyd resins 7, 8 and 9.

nals at 2.8–1.5 ppm are characteristic of CH and CH2 protons of the


fatty acids. Methyl protons of PG, bis-MPA and fatty acid residues
appear at 1.4–0.85 ppm.

3.2. Thermal properties

Thermal stability of the different prepared alkyd resins and the


cured film of resin 9 are represented in Figs. 6 and 7.
1
It is worth mentioning that, the thermal behavior of the differ-
Fig. 4. H NMR spectra of glycolyzed products and alkyd resins 1, 2, 3 and 4.
ent alkyd resin samples was similar irrespective of the type and
percentage of the multifunctional polyols. This indicates that all
samples have the same mechanism of degradation and molecular
structure. The onset of decomposition of all samples displayed at
values >200 ◦ C rendering them suitable for the processing require-
N.E. Ikladious et al. / Progress in Organic Coatings 102 (2017) 217–224 223

et al. [40] reported that: A stable residue of 11% was detected as a


result of thermal degradation of alkyd resin that does not contain
any inorganic filler.

3.3. Film properties

Film properties of all prepared alkyd resins are summarized in


Table 3. It is clear that all evaluated coatings showed excellent,
impact, adhesion, bending, and ductility indicating the flexibility
of the cross-linked films and good adhesive bond to substrate.
In order to fulfill their functions of protecting or decorating a
substrate, coating materials must adhere very well to the substrate
for the expected service life. It has been observed that all prepared
films have good adhesion strength and no failure occurred between
coat and substrate. Therefore all films were classified in class 5B.
The impact resistance of any coated substrate is considered as its
ability to withstand a high force or shock applied to it over a short
Fig. 6. TGA curves of alkyd resins 1, 2, 3 and 4. time. The nature of polymer backbone and the free volume available
between the chains play an important role in attaining satisfactory
impact resistance. All films had good impact resistance and retained
their integrity without cracking in impact test with 1.8 kg weight
and 1 m height. These samples had an impact resistance higher than
18 J, which is the upper limit of the potential energy that can be
detected by our instrument. This may be attributed to the flexible
nature of the films due to the long chain fatty acids resulting in
available free volume between adjacent chains. Furthermore, gloss
and hardness increased with increasing the degree of branching
and consequently, the molecular weight, in the resin backbone. It
is known that the resin molar mass, the amounts of the fatty acids
and the reactants containing aromatic structure greatly affect the
hardness of the coating. It is obvious that, these prepared alkyd
resin formulations had higher hardness values than those formula-
tions based only on hyperbranched polyesters and fatty acids [31],
in which hardness values ranging between 8 and 18 were reported.
Such mprovements can be related to the synergetic effect of the
molecular weight of the polyol and the aromatic nature of the gly-
colyzed products and PA. Although alkyd resins 5 and 6 include
the double percentage of G1 compared to alkyd 3 and 7, all these
Fig. 7. TGA curves of alkyd resins 7, 8, 9 and the cured film of resin 9.
samples have nearly similar film properties. This implies that nei-
ther the higher percentage of the hyperbranched polyester, G1, nor
ments of most conventional coating applications. In order to
the type of fatty acid affected the film properties of these samples.
compare the thermal behavior of the alkyd samples, the degrada-
As shown in Table 3 the gloss values of all prepared samples were
tion temperatures at 10 and 50% weight loss (T10 and T50 ), and
higher than 80, except that of resins 1 and 2. This high gloss values
the residues at 550 ◦ C of resins 1, 2, 3, 4, 7, 8, 9 and the cured
indicate homogeneous drying throughout the whole film and good
film of resin 9 are presented in Table 2. As can be seen, the 50%
compatibility of the reactants resulting in good film-forming and
weight loss of the alkyd resin 4 and 9 samples took place at 418 ◦ C
leveling properties.
and 410 ◦ C whereas, resins 1, 2, 3, 7, 8 and the cured resin 9
Based on the above results, it appears that it is feasible to
weight losses displayed at 399, 400, 400, 393, 399 and 388 ◦ C,
use such materials, namely, recycled PET and hyperbranched
respectively. It seems that the presence of G3 as polyol (cf resins
polyesters to synthesize new alkyd resins. The present alkyd resins
4 and 9), results in an improvement in thermal stability. This is
based on inexpensive recycled and renewable materials provide
attributed to the high degree of functionality and consequently
sustainable and environmentally friendly industrial coatings.
high molecular weight of G3. Table 2 demonstrates that among
all alkyd samples, the cured film exhibited the lowest thermal
stability. A similar behavior has been observed for other previ-
ously prepared alkyd resins [31]. This behavior was explained by 4. Conclusions
Stenberg et al. [39] as follow: The reaction rate of curing using a
dryer is influenced by both desired intermolecular and undesired • The glycolyzed products of PET waste together with aliphatic
intramolecular degradation reactions. In the desired stage building hyperbranched polyesters can be considered as interesting com-
up the polymer matrix occur whereas the undesired leads to photon bination for synthesizing new alkyd resins through reaction with
emission from the auto-oxidation process as well as volatile emis- linseed or sunflower oil fatty acids.
sion of low molecular weight species. Moreover, the higher residue • The prepared alkyds were characterized using IR and1 H NMR
amount at 550 ◦ C (11.4%) of the cured film may be attributed to spectroscopy.
both the carbonization of the aromatic groups, originated from PA • All alkyd samples showed initial decomposition temperature at
and glycolyzed products and to thermally induced crosslink for- values >200 ◦ C which renders them suitable for the processing
mations. These findings are in good agreement with what Cadena requirements of most conventional coating applications.
224 N.E. Ikladious et al. / Progress in Organic Coatings 102 (2017) 217–224

• A significant improvement in adhesion, hardness, bending, [17] J. Dulius, C. Ruecker, V. Oliveira, R. Ligabue, S. Einloft, Chemical recycling of
impact, ductility and high gloss values is achieved for all alkyd post-consumer PET: alkyd resin synthesis, Prog. Org. Coat. 57 (2006) 123–157.
[18] A. Hult, M. Johansson, E. Malmström, Hyperbranched polymers, Adv. Polym.
films. Sci. 143 (1999) 1–34.
• The new alkyd resins that based on recycled PET and hyper- [19] B. Voit, New developments in hyperbranched polymers, J. Polym. Sci. Part A:
branched polyesters appear to be promising candidates for the Polym. Chem. 38 (2000) 2505–2525.
[20] M. Jikei, M. Kakimoto, Hyperbranched polymers: a promising new class of
development of eco-friendly and inexpensive industrial coatings. materials, Prog. Polym. Sci. 26 (2001) 1233–1285.
[21] P.N. Mehta, Hyperbranched polymers: unique design tool for coatings, Surf.
Acknowledgement Coat. Int. Part B: Coat. Trans. 89 (2006) 333–342.
[22] C. Gao, D. Yan, Hyperbranched polymers: from synthesis to applications, Prog.
Polym. Sci. 29 (2004) 183–275.
The authors express their deep gratitude to the National [23] B. Voit, D. Beyerlein, K.-J. Eichhorn, K. Grundke, D. Schmaljohann, T. Loontjens,
Research Center, for the financial support of this work through Functional hyper-branched polyesters for application in blends coatings, and
thin films, Chem. Eng. Technol. 25 (2002) 704–707.
project no. P100105.
[24] K. Manczyk, P. Szewczyk, Highly branched high solids alkyd resins, Prog. Org.
Coat. 44 (2002) 99–109.
References [25] B.R. Frings, M. Wend, New hyperbranched polyesters for UV-curing, DIC Tech.
Rev. (9) (2003) 43–51.
[26] K. Johansson, T. Bergman, M. Johansson, Hyperbranched aliphatic polyesters
[1] V. Sinha, M.R. Patel, J.V. Patel, PET waste management by chemical recycling:
and reactive diluents in thermally cured coil coatings, ACS Appl. Mater.
a review, J. Polym. Environ. 18 (2010) 8–25.
Interfaces 1 (2009) 211–217.
[2] F. La Mantia, Handbook of Plastics Recycling, Rapra Technology, Shropshire
[27] A. Hofland, Alkyd resins: from down and out to alive and kicking, Prog. Org.
United Kingdom, 2002.
Coat. 73 (2012) 274–282.
[3] N.E. Ikladious, Identification of glycolysis products, J. Elastom. Plast. 32 (2000)
[28] E.A. Murillo, P.P. Vallejo, B.L. López, Synthesis and characterization of
140–151.
hyperbranched alkyd resins based on tall oil fatty acids, Prog. Org. Coat. 69
[4] S.H. Mansour, N.E. Ikladious, Depolymerisation of poly(ethylene
(2010) 235–240.
terephthalate) wastes using 1,4-butanediol and triethylene glycol, Polym.
[29] E.A. Murillo, P.P. Vallejo, B.L. López, Effect of tall oil fatty acids content on the
Test. 21 (2002) 497–505.
properties of novel hyperbranched alkyd resins, J. Appl. Polym. Sci. 120 (2011)
[5] S.H. Mansour, S.L. Abd-El-Messieh, N.E. Ikladious, Utilization of some
3151–3158.
oligomers based on poly(ethylene terephthalate) wastes as modifiers for
[30] R. Baloji Naik, D. Ratna, S.K. Singh, Synthesis and characterization of novel
polyvinyl chloride, J. Appl. Polym. Sci. 85 (2002) 2501–2509.
hyperbranched alkyd and isocyanate trimer based high solid polyurethane
[6] A.M. Atta, Epoxy resin based on poly(ethylene terephthalate) waste: synthesis
coatings, Prog. Org. Coat. 77 (2014) 369–379.
and characterization, Prog. Rubber Plast. Recycl. Technol. 19 (2003) 17–40.
[31] N.E. Ikladious, S.H. Mansour, J.N. Asaad, H.S. Emira, M. Hilt, Prog. Org. Coat. 89
[7] A.M. Atta, S.I. Elnagdy, M.E. Abdel-Raouf, Compressive properties and curing
(2015) 252–259.
behaviour of unsaturated polyester resins in the presence of vinyl ester resins
[32] A. Torlakoglu, G. Güclü, Alkyd-amino resins based on waste PET for coating
derived from recycled poly(ethylene terephthalate), J. Polym. Res. 12 (2005)
applications, Waste Manag. 29 (2009) 350–354.
373–383.
[33] J. Purohit, G. Chawada, B. Choubisa, M. Patel, B. Dholakiya, Polyester polyol
[8] A.M. Atta, M.E. Abdel-Raouf, S.M. Elsaeed, A.A. Abdel-Azim, Curable resins
derived from waste poly (ethylene terephthalate) for coating application on
based on recycled poly(ethylene terephthalate) for coating applications, Prog.
mild steel, Chem. Sci. J. 2012 (2012) CSJ–76.
Org. Coat. 55 (2006) 50–59.
[34] A.M. Atta, R.A. El-Ghazawy, A.M. El-Saeed, Corrosion protective coating based
[9] S.L. Abd-El-Messieh, S.H. Mansour, D. El-Nashar, N.E. Ikladious, Evaluation of
on alkyd resins derived from recycled poly(ethylene terephthalate) waste for
polyester resin as a new compatibilizer for SBR/PVC blends, Can. J. Chem. Eng.
carbon steel, Int. J. Electrochem. Sci. 8 (2013) 5136–5152.
82 (2004) 358–370.
[35] S.H. Mansour, N.N. Rozik, K. Dirnberger, N.E. Ikladious, Hyperbranched
[10] A.B. Raheem, L. Uyigue, The conversion of post-consumer polyethylene
polyesters based on polycondensation of 1,3,5-Tris(2-hydroxyethyl) cyanuric
terephthalate (PET) into a thermosetting polyester resin, Arch. Appl. Sci. Res.
acid and 3,5-dihydroxybenzoic acid, Part A: Polym. Chem. 43 (2005)
2 (2010) 240–254.
3278–3288.
[11] G.P. Karayannidis, D.S. Achilias, I.D. Sideridou, D.N. Bikiaris, Alkyd resins
[36] N.E. Ikladious, S.H. Mansour, N.N. Rozik, K. Dirnberger, C.D. Eisenbach, New
derived from glycolized waste poly(ethylene terephthalate), Eur. Polym. J. 41
aliphatic hyperbranched polyester polyols based on 1,3
(2005) 201–210.
5-Tris(2-hydroxyethyl) cyanuric acid as a core, J. Polym. Sci. Part A: Polym.
[12] D.E. Nikles, M.S. Farahat, New motivation for the depolymerization products
Chem. 46 (2008) 5568–5579.
derived from poly(ethylene terephthalate) (PET) waste: a review, Macromol.
[37] N.E. Ikladious, J.N. Asaad, N.N. Rozik, Modification and methacrylation ofsome
Mater. Eng. 290 (2005) 13–30.
new aliphatic hyperbranched polyester polyols based on
[13] E. Bulak, I. Acar, The use of aminolysis, aminoglycolysis, and simultaneous
1,3,5-tris(2-hydroxyethyl) cyanuric acid (THECA) as a core, Des. Monomers
aminolysis-hydrolysis products of waste PET for production of paint binder,
Polym. 12 (2009) 469–481.
Polym. Eng. Sci. 54 (2014) 2272–2281.
[38] J.N. Asaad, N.E. Ikladious, F. Awad, T. Müller, Evaluation of some
[14] J.N. Asaad, S.L. Abd-El-Messieh, N.E. Ikladious, Unsaturated polyester
newhyperbranched polyesters as binding agents for heavy metals, Can. J.
nanocomposites based on PET waste using different types of nanofillers,
Chem. Eng. 91 (2013) 257–263.
Proceedings of the Institution of Mechanical Engineers Part N: J.
[39] C. Stenberg, M. Svensson, M. Johansson, A study of the drying of linseed oils
Nanoengineering and Nanosystems 228 (2014) 174–183.
with different fatty acid patterns using RTIR-spectroscopy and
[15] G. Güclü, M. Orbay, Alkyd resins synthesized from postconsumer PET bottles,
chemiluminescence (CL), Ind. Crops Prod. 21 (2005) 263–272.
Prog. Org. Coat. 65 (2009) 362–365.
[40] F. Cadena, L. Irusta, M.J. Fernandez-Berridi, Performance evaluation of alkyd
[16] O. Tuna, A. Bal, G. Güclü, Investigation of the effect of hydrolysis products of
coatings for corrosion protection in urban and industrial environments, Prog.
postconsumer polyethylene terephthalate bottles on the properties of alkyd
Org. Coat. 76 (2013) 1273–1278.
resins, Polym. Eng. Sci. 53 (2013) 176–182.

You might also like