0% found this document useful (0 votes)
11 views

PETG

Uploaded by

catxinha93
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
11 views

PETG

Uploaded by

catxinha93
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 34

View Article Online

View Journal

RSC Advances
This article can be cited before page numbers have been issued, to do this please use: T. Chen, G. Jiang,
G. Li, Z. Wu and J. Zhang, RSC Adv., 2015, DOI: 10.1039/C5RA09252C.

This is an Accepted Manuscript, which has been through the


Royal Society of Chemistry peer review process and has been
accepted for publication.

Accepted Manuscripts are published online shortly after


acceptance, before technical editing, formatting and proof reading.
Using this free service, authors can make their results available
to the community, in citable form, before we publish the edited
article. This Accepted Manuscript will be replaced by the edited,
formatted and paginated article as soon as this is available.

You can find more information about Accepted Manuscripts in the


Information for Authors.

Please note that technical editing may introduce minor changes


to the text and/or graphics, which may alter content. The journal’s
standard Terms & Conditions and the Ethical guidelines still
apply. In no event shall the Royal Society of Chemistry be held
responsible for any errors or omissions in this Accepted Manuscript
or any consequences arising from the use of any information it
contains.

www.rsc.org/advances
Page 1 of 33 RSC Advances

View Article Online

RSC Advances Accepted Manuscript


DOI: 10.1039/C5RA09252C
Poly(ethylene glycol-co-1,4-cyclohexanedimethanol terephthalate)

random copolymers: effect of copolymer composition and

microstructure on thermal properties and crystallization behavior

Tingting Chen, Guodong Jiang, Guoyu Li, Zhipeng Wu, Jun Zhang

Abstract
Published on 08 July 2015. Downloaded by Carleton University on 10/07/2015 15:31:24.

A series of poly(ethylene glycol-co-1,4-cyclohexanedimethanol terephthalate)

(PETG) random copolymers were synthesized and characterized using 1H and 13


C

nuclear magnetic resonance, infrared spectroscopy and viscometry. Differential

scanning calorimetry, wide-angle X-ray diffraction, and thermogravimetric analysis

were used to probe the effects of copolymer composition and microstructure on the

thermal properties, crystallization behavior, and thermal stability of the PETG

copolymers, respectively. The mechanical properties were evaluated by tensile testing

and dynamic mechanical measurements. The partial replacement in the

polymerization feed of ethylene glycol by 1,4-cyclohexanedimethanol led to

statistically random PETG copolymers with adjustable compositions and suitable

molecular weights, which were thermally stable above 380 oC. The number-average

sequence length of ethylene glycol terephthalate decreased with the increasing

1,4-cyclohexanedimethanol terephthalate (CT) content. In contrast, the

number-average sequence length of CT increased gradually with the increasing CT

content. The crystalline structure of the PETG copolymers changed from PET-type

College of Materials Science and Engineering, Nanjing Tech University, Nanjing 210009, China.
 Corresponding author. E-mail: [email protected]; Fax: +86-25-83240205; Tel.: +86-25-83587264.

1
RSC Advances Page 2 of 33

View Article Online

RSC Advances Accepted Manuscript


lattice to PCT-type lattice at a lower CT content. The crystallinity decreasedDOI:
at 10.1039/C5RA09252C
first,

and then increased remarkably with the increasing CT content. It was interesting to

notice that the rigid structure of CT unit controlled the crystallization. The

incorporation of CT in the PET chain, significantly altered the thermal transitions of

the polyester. The glass transition temperature increased linearly with the increasing
Published on 08 July 2015. Downloaded by Carleton University on 10/07/2015 15:31:24.

CT content. The melting temperature of the segments in crystalline domains strongly

depended on the corresponding average sequence length. An increase in the average

sequence length resulted in a higher melting temperature and an increase in the

melting enthalpy.

1 Introduction

Aliphatic aromatic polyesters are a class of thermoplastic polyesters with unique

properties, such as good mechanical properties, good chemical resistance, excellent

surface appearance, and stable electrical insulation properties.1 As one of the most

famous aliphatic aromatic polyesters, semicrystalline poly(ethylene terephthalate)

(PET) has broad applications and usually used in the form of fibers or films.2-6 The

properties of PET are subjected to the degree of crystallization.7 However, at the glass

transition temperature (Tg) (~70 oC) the mobility of the polymer chains in amorphous

region increases considerably, resulting in the decrease of stiffness.8 Hence, PET is

not suitable for applications in high temperature environment.

Increasing the Tg of polyesters, PET in particular, is currently an open front in the

injection-moulding area. PET with higher Tg displays greater dimensional stability in

2
Page 3 of 33 RSC Advances

View Article Online

RSC Advances Accepted Manuscript


DOI: 10.1039/C5RA09252C
thermal processing and better barrier properties in packing.9 Streamline hotfilling

bottle process requiring materials with a Tg above filling temperature is a good

example, which drives the development of the innovative PET derivatives.10

Copolymerization has been and continues to be the most appealing approach

addressed to modify the properties of aromatic polyesters.11 Cyclic monomers, which

tend to restrict the chain mobility, are able to render polyesters with enhanced Tg.12
Published on 08 July 2015. Downloaded by Carleton University on 10/07/2015 15:31:24.

Cycloaliphatic diols, such as 1,4-cyclohexanedimethanol (CHDM) can impart rigidity

to the polyester chain. Poly(ethylene glycol-co-1,4-cyclohexanedimethanol

terephthalate) (PETG), a modified PET, is prepared by partially replacing the ethylene

glycol groups of PET with CHDM groups. The crystallization rate and crystallinity of

the prepared PETG copolymers are decreased, which can be attributed to the reduced

molecular regularity by the introduction of CHDM into PET molecular chains.13

In this study, PET and a series of PETG random copolymers with different

compositions were synthesized via bulk polycondensation. The copolymer

composition and microstructure of the PETG copolymers were determined by 1H and

13
C nuclear magnetic resonance (NMR) spectroscopy. Attenuated total reflection

Fourier transform infrared spectroscopy (ATR-FTIR) was used to analyze the

chemical structure of the PETG copolymers. Their thermal properties, crystallization

behavior, and thermal stability were investigated using differential scanning

calorimetry (DSC), wide-angle X-ray diffraction (WAXD), and thermogravimetric

analysis (TGA), respectively. The tensile testing and dynamical mechanical thermal

analysis (DMTA) were used to evaluate their mechanical properties. The aim of this

3
RSC Advances Page 4 of 33

View Article Online

RSC Advances Accepted Manuscript


study was to investigate the effects of the copolymer composition DOI:
and10.1039/C5RA09252C
the

microstructure on the thermal properties, crystallization behavior, and thermal

stability of the PETG copolymers.

2 Experimental
Published on 08 July 2015. Downloaded by Carleton University on 10/07/2015 15:31:24.

2.1 Materials

Terephthalic acid (TPA) (99%), ethylene glycol (EG) (99%), and

1,4-cyclohexanedimethanol (CHDM) (99%) were purchased from Yangzi

Petrochemical Engineering Co., Ltd., Shanghai Lingfeng Chemical Reagent Co., Ltd.,

and SK Chemicals Co., Ltd., respectively. The solvents used for purification and

characterization, such as methanol, phenol, and 1,1,2,2-tetrachloroethane, were

analytical grade and used as received. The catalysts (antimony acetate, cobalt acetate,

and organic tin) were friendly supplied by Jiangsu Jinghong New Material

Technology Co., Ltd. The antioxidants (Irganox 1010 and Irgafos 168) were

purchased from Ciba Specialty Chemicals Co., Ltd.

2.2 Synthesis

The PETG copolymers and PET used in this study were prepared from TPA, EG,

and CHDM using organic tin, antimony acetate and cobalt acetate as catalysts. The

two-step polymerization was performed on a laboratory-scale polymerization reactor

in the molten state. The first step was the esterification reaction of TPA, EG, and

CHDM with catalytic organic tin. The second step was the polycondensation reaction
4
Page 5 of 33 RSC Advances

View Article Online

RSC Advances Accepted Manuscript


DOI: 10.1039/C5RA09252C
using antimony acetate and cobalt acetate as catalysts. The monomers, catalysts, and

antioxidants were premixed in an autoclave reactor, and the mixture was gradually

heated under constant stirring. The eaterification reaction was carried out around 230

o
C under nitrogen atmosphere, and the reaction was controlled by the amount of the

distilled water. The polycondensation of bis(hydroxyethyl)terephthalate (BHET) and

bis(hydroxymethylcyclohexane)terephthalate (BHCT) was carried out around 280 oC


Published on 08 July 2015. Downloaded by Carleton University on 10/07/2015 15:31:24.

under high-vacuum condition. At the end of the reaction, the products in the melt were

quenched into a water bath and followed by drying in a vacuum oven at room

temperature for 48 hrs. The monomer composition in the feed was varied from 0, 15,

30, 50, to 70 mol% of CHDM. The polymers synthesized were purified by methanol.

The reactions involved in these syntheses are formulated in Scheme 1.

Scheme 1 Synthesis of the PETG copolymers.

2.3 Characterization

The intrinsic viscosity [η] measurements were carried out in a mixed solvent of
5
RSC Advances Page 6 of 33

View Article Online

RSC Advances Accepted Manuscript


DOI: 10.1039/C5RA09252C
phenol/1,1,2,2-tetrachloroethane (1/1, w/w) at 25 ± 0.1 oC using an Ubbelohde

viscometer.14 1
H NMR spectroscopy was used for determining the copolymer

13
composition. C NMR spectroscopy was used for determining the dyad sequence

distribution in the copolymers. 1H and 13


C NMR spectra were recorded on a Bruker

DRX 500 operating at 500 and 125 MHz, respectively. About 12 and 100 mg of

sample dissolved in 0.5 mL deuterated trifluoroacetic acid (TFA-d1) for 1H and 13


Published on 08 July 2015. Downloaded by Carleton University on 10/07/2015 15:31:24.

NMR spectra, respectively. On determination of the dyad sequence distribution, the

relative peak intensities for EE, EC, CE and CC dyads are deconvoluted, and their

peak areas are considered to be corresponding dyad quantities.

ATR-FTIR spectroscopy was performed on a FTIR spectrometer (Thermo Fisher

Nicolet IS5, America) in the attenuated total reflectance mode. The spectra were

scanned from 4000 to 600 cm-1 with a resolution of 4 cm-1.

2.4 WAXD measurements

The WAXD experiments were performed using a WAXD diffractometer

(Shimadzu 6000, Japan). The operated voltage and current were 40 kV and 30 mA,

respectively. The samples were annealed at 150 oC for 1 h prior to the measurement.

The samples were scanned from 2θ=10o to 40o at a speed of 5o min-1. The lamellar

thickness (L) was calculated using the Scherrer’s formula:15

L=Kλ/βcosθ (1)

where K is a structure factor, taken as 1.0; λ is the wavelength of the monochromatic

X-ray beam [nm] (λ=0.154 nm for CuKα radiation15); β is the full width at half

6
Page 7 of 33 RSC Advances

View Article Online

RSC Advances Accepted Manuscript


maximum for the diffraction peak [rad]; and θ is the Bragg angle [o]. DOI: 10.1039/C5RA09252C

2.5 DSC measurements

A differential scanning calorimeter (TA Q20, USA) was applied to investigate

the thermal properties of the PETG copolymers and PET. Temperature and heat flow

were calibrated using a high purity indium standard (156.6 oC and 28.45 J g-1). All
Published on 08 July 2015. Downloaded by Carleton University on 10/07/2015 15:31:24.

samples were dried in a vacuum oven at 60 oC for 12 h prior to the test. The

measurement was performed using about 10 mg sample sealed in an aluminum pan

under the nitrogen atmosphere with a purge flow of 50 mL min-1. The sample was

heated to 300 oC at 40 oC min-1 and kept isothermally for 1 min to erase the previous

thermal history. Then, the sample was quenched to the amorphous state. Thereafter,

the sample was heated to 300 oC at 10 oC min-1 to clearly detect the glass transition

temperature (Tg). Finally, a second cool scan and a third heat scan were performed at

10 oC min-1 to obtain the thermal properties.

The crystallinity (Xc) of sample was calculated as follows:16

X c  H m / H m* 100% (2)

where H m is the melting enthalpy, and H m* is the melting enthalpy of 100%

crystalline PET, which is equal to 130 J g-1.17

2.6 TGA measurements

The thermal stability of the PETG copolymers and PET were evaluated by TGA.

The thermal degradation process was recorded using a TGA thermal analyzer (TA

7
RSC Advances Page 8 of 33

View Article Online

RSC Advances Accepted Manuscript


Q50, USA). For each experiment, about 20 mg of the sample was used and a DOI: 10.1039/C5RA09252C
nitrogen

flow rate of 50 mL min-1 was adopted. The samples, under a nitrogen protective

atmosphere, were heated from 40 to 800 oC at a rate of 20 oC min-1.

2.7 Tensile and shrinkage properties measurements

Samples for tensile testing were prepared with a thickness of 0.5 mm by


Published on 08 July 2015. Downloaded by Carleton University on 10/07/2015 15:31:24.

compression molding. Subsequently, substantially amorphous samples were obtained

by minimizing crystallization, using iced water bath, and followed by drying in a

vacuum oven at room temperature for 48 hrs. The samples were cut into

dumbbell shape with a width of 4 mm while the distance between the testing marks

was 20 mm. The Young’s modulus, yield stress, and elongation at break were

measured at a stretching rate of 5 mm min-1 on a universal testing machine (CMT

5254, SANS Co. China) at room temperature.

The monoaxial shrinkage of the PETG copolymers and PET was performed on

rectangular samples (50×10 mm2) cut from amorphous, isotropic films with a

thickness of about 0.5 mm. The tests were conducted on a universal testing machine

(CMT 4254, SANS Co. China) equipped with a temperature-controlled thermal

cabinet. The sample was stretched to a stretch ratio of 3×1 for 5 min at 90 oC. Then,

the sample was fixed under this strain and cooled to 25 oC for 10 min to measure the

original length (L0). Finally, the sample was placed into the thermal cabinet again at

90 oC for 5 min without any stress and the final length (Lf) was recorded. The

shrinkage at 90 oC was calculated by the following equation:18

8
Page 9 of 33 RSC Advances

View Article Online


Shrinkage  ( L0  L f ) / L0 100%

RSC Advances Accepted Manuscript


DOI: 10.1039/C5RA09252C
(3)

2.8 DMTA measurements

Samples for dynamic mechanical measurements were obtained by compression

molding. Subsequently, substantially amorphous samples were obtained by

minimizing crystallization, using iced water bath, and followed by drying in a vacuum
Published on 08 July 2015. Downloaded by Carleton University on 10/07/2015 15:31:24.

oven at room temperature for 48 hrs. Samples were cut into a 25×6×1 mm3

rectangular shape. Dynamic mechanical measurements were performed with a

dynamical mechanical thermal analyzer (Anton Paar MCR 302, Austria), at a

frequency of 1 Hz and a heating rate of 3 oC min-1, over a temperature range from 30

to 160 oC.

3 Results and discussion

3.1 NMR analysis

The 1H NMR spectra of the PETG copolymers and PET and the assignment of

each peak are shown in Fig. 1. The equation for determining the copolymer

composition from respective peak area is given as follow:

5(a  b  d  e) 5c 2( X  Y )
  (4)
f f Y

where a, b, c, d, e, and f represent the areas of corresponding peaks in Fig. 1, and X

and Y denote the mole fractions of ethylene glycol terephthalate (ET) and

1,4-cyclohexanedimethanol terephthalate (CT) units, respectively. When the

9
RSC Advances Page 10 of 33

View Article Online

RSC Advances Accepted Manuscript


DOI: 10.1039/C5RA09252C
copolymer compositions as calculated by Eq. (4) are compared with the feed

composition, it is revealed that the copolymer compositions are nearly equal to the

feed compositions (results are shown in Table 1).


Published on 08 July 2015. Downloaded by Carleton University on 10/07/2015 15:31:24.

1H
Fig. 1 NMR spectra of the PETG copolymers and PET and the assignment of each peak.

In 1H NMR spectra, since peak b (δ = 4.85 ppm) is slightly overlapped with peak

d (δ = 5.00 ppm), peak a (δ = 4.18 ppm) was used as a chain end proton for polymer

structure estimation.19 The total repeating units of ET (NET) and CT (NCT) in

copolymers can be calculated by the integration ratio of chain end to corresponding

repeating units by 1H NMR spectra, using the following equations:

Ia 4
 (5)
I d 4 N ET  8

Ia 4
 (6)
I e 4 N CT

where Ia, Id, and Ie are the integration values for the corresponding peak at positions a,

d, and e, and NET and NCT are the total repeating units of ET and CT, respectively. The
10
Page 11 of 33 RSC Advances

View Article Online

RSC Advances Accepted Manuscript


DOI: 10.1039/C5RA09252C
total number-average molecular weights of the whole polymer ( M n ) can thus be

calculated by M n /Dalton = 192 × NET + 274 × NCT (where 192 and 274 are the

molecular weights of the repeating units of ET and CT), with the results listed in

Table 1. The estimated number-average molecular weights are in the range of

1.2×104-1.6×104 g mol-1.
Published on 08 July 2015. Downloaded by Carleton University on 10/07/2015 15:31:24.

For the copolymers synthesized, viscosity average molecular weights ( M v ) were

estimated from the measured intrinsic viscosities ([η]) using the Mark-Houwink

equation with constants α and k which were determined previously by Ward20 for PET

homopolymer:
0.77
[ ]  2.75 104 M v (7)

The results are listed in Table 1. The estimated viscosity average molecular

weights are in the range of 2.5×104-3.1×104 g mol-1, indicating that the resulting

polymers have relatively high molecular weight, which are able to form films. In fact,

Eq. (7) was found for PET homopolymer rather than PETG copolymers. The

Mark-Houwink constants are not known for other copolymer compositions, therefore

values of M v are used for relative comparison.


Table 1 Compositions, number-average molecular weights, intrinsic viscosities, and viscosity
average molecular weights of the PETG copolymers and PETa
Feed Copolymer
CHDMb
composition compositionb Mnb [η] c Mvd
Copolymer (mol %) NET NCT
(mol %) (mol %) (g mol-1) (dL g-1) (g mol-1)
ET/CT ET/CT trans/cis
PET(100/0) 100/0 100/0 0/0 74.5 0 1.4×104 0.69 2.6×104
PETG(85/15) 85/15 84.4/15.6 66.7/33.3 64.5 13.3 1.6×104 0.79 3.1×104
PETG(70/30) 70/30 70.9/29.1 70.4/29.6 45.1 21.3 1.4×104 0.70 2.6×104
PETG(50/50) 50/50 51.6/48.4 69.9/30.1 30.8 27.3 1.3×104 0.67 2.5×104
PETG(30/70) 30/70 31.6/68.4 68.0/32.0 16.9 30.8 1.2×104 0.75 2.9×104
aN
ET, NCT: total repeating units of ET and CT, respectively.
11
RSC Advances Page 12 of 33

b View Article Online


Measured by 1H NMR.

RSC Advances Accepted Manuscript


DOI: 10.1039/C5RA09252C
c Measured in a mixed solvent of phenol and 1,1,2,2-tetrachoroethane (1/1, w/w) using an

Ubbelhode viscometer at 25 oC.


d Viscosity average molecular weights were estimated from the [η] using the Mark-Houwink

equation.
13
The microstructure of the PETG copolymers is analyzed using C NMR

spectroscopy. The quaternary aromatic carbon resonances appear to be sensitive to

dyad sequence effects.21 The signals due to these carbons appear to split giving a total
Published on 08 July 2015. Downloaded by Carleton University on 10/07/2015 15:31:24.

of four peaks since four types of dyads (EE, EC, CE and CC) are possible. The four

different dyads possible in the copolymer repeat units are depicted in Fig. 2a. The

132.5-134.5 ppm region containing the resonance signals produced by such carbons is

shown in Fig. 2b. The peaks arising from EE, EC, CE and CC dyads appear well

resolved as to be quantified by integration with reliability. The relative concentrations

of the four dyads are determined from deconvoluted areas of the four signals.

12
Page 13 of 33 RSC Advances

View Article Online

RSC Advances Accepted Manuscript


DOI: 10.1039/C5RA09252C
Published on 08 July 2015. Downloaded by Carleton University on 10/07/2015 15:31:24.

Fig. 2 13C NMR spectra: (a) the four possible dyads occurring in PETG copolymers, (b) the
non-protonated aromatic carbon region in the 13C NMR spectra of the PETG copolymers and PET.

In Fig. 3, the dyad contents are plotted against the copolymer compositions. The

Dashed lines represent the theoretically calculated contents for statistically random

copolymers, in which the dyad sequence distribution obeys the Bernoullian statistics

model:22

NEE  X ET 2 (8)

NEC  NCE  X ET X CT (9)

NCC  X CT 2 (10)

where Nij and Xi denote the mole fractions of the ij dyad sequence and i component in

the copolymer chains, respectively. Since the fractions of the dyad sequence

determined experimentally are in good agreement with the dyad sequence

distributions calculated by the Bernoullian statistics. It is suggested that the PETG

samples synthesized in this study are statistically random copolymers.

13
RSC Advances Page 14 of 33

View Article Online

RSC Advances Accepted Manuscript


DOI: 10.1039/C5RA09252C
Published on 08 July 2015. Downloaded by Carleton University on 10/07/2015 15:31:24.

Fig. 3 Plot of dyad contents vs. copolymer compositions. The dashed lines represent the
calculated distribution based on the Bernouillian statistics.

According to Yamadera and Murano,23 if four kinds of signals due to homolinks

and heterolinks are observed in the NMR spectroscopy of the copolymer, then the

average sequence length and the degree of randomness of the copolymer can be

further determined. The number-average sequence lengths (n) of the ET and CT

homogeneous sequences, as well as the degree of randomness (R), are estimated for

each PETG copolymers by using the following equations:24

nET  ( NEE  1/ 2( NEC  NCE )) 1/ 2( NEC  NCE ) (11)

nCT  ( NCC  1/ 2( NEC  NCE )) 1/ 2( NEC  NCE ) (12)

R  (1/ nET )  (1/ nCT ) (13)

For random copolymers, R is unity. The value of R equal to zero indicates a

mixture of homopolymers, while a value of 2 indicates an alternating distribution.23

Results from these calculations are summarized in Table 2, indicating that the

sequence distribution in PETG copolymers is essentially random in the entire range of

the compositions as the values of R are very close to unity. The number-average

sequence length nET decreases with the increasing CT unit content. In contrast, nCT

14
Page 15 of 33 RSC Advances

View Article Online

RSC Advances Accepted Manuscript


increases gradually with the increasing CT unit content. The concentrationDOI:of10.1039/C5RA09252C
the

esterification product BHCT increases with the increasing content of monomer

CHDM, which leads to an increase of the probability that BHCT reacts with BHCT

during the polycondensation reaction. This will increase the number-average sequence

length of the CT homogeneous sequences. On the contrary, it will leads to the


Published on 08 July 2015. Downloaded by Carleton University on 10/07/2015 15:31:24.

decreasing nET.
Table 2 Composition and microstructure of the PETG copolymersa
Dyad contentb (%) Sequence lengths
Copolymer R
NEE NCE NCC nET nCT
PETG(85/15) 74.67 23.18 2.15 7.44 1.19 0.97
PETG(70/30) 53.83 39.66 6.51 3.71 1.33 1.02
PETG(50/50) 29.30 50.60 20.10 2.16 1.79 1.02
PETG(30/70) 10.62 45.50 43.88 1.47 2.93 1.02
a
n: Number-average sequence lengths. R: Degree of randomness.
b Measured by 13C NMR.

3.2 ATR-FTIR analysis

The chemical structure of the PETG copolymers was analyzed by ATR-FTIR

spectroscopy taking PET as the reference in Fig. 4. The spectra were baseline

corrected and normalized at 1712 cm-1, the major absorbance peak representing the

C=O of ester groups.1 The peaks at 2968 and 2907 cm-1 are attributed to the

asymmetric and symmetric aliphatic C-H stretching vibrations in PET.25,26 The C-H

symmetrical and asymmetrical stretching due to existence of methylene groups in the

PETG copolymers can be observed between 2924 and 2852 cm-1.27 The C(=O)-O

stretching peak, CH2 bending peak, and C-H stretching peak of cyclohexylene ring are

observed at 1260, 1451 and 958 cm-1, respectively, in the PETG copolymers with

various CT content.28-31 The C-H stretching peak of cyclohexylene ring in the PETG
15
RSC Advances Page 16 of 33

View Article Online

RSC Advances Accepted Manuscript


DOI: 10.1039/C5RA09252C
copolymers increases with the increasing CT content. This result is accordance with

the copolymer composition measured by 1H NMR.


Published on 08 July 2015. Downloaded by Carleton University on 10/07/2015 15:31:24.

Fig. 4 ATR-FTIR spectra of the PETG copolymers and PET.

3.3 WAXD analysis

The crystallization behavior of the PETG copolymers and PET was probed by

WAXD. The diffraction patterns of the as-obtained samples exhibit a typical

semicrystalline polymer behavior. The diffraction patterns of the samples with the

composition close to the 50:50 is not well resolved. Hence, all the samples were

annealed at 150 oC for 1 h prior to the further WAXD measurement. After annealing,

the diffraction peaks were well resolved. Fakirov and co-workers investigated this

phenomena and suggested that the emergence of the well resolved diffraction peaks

might be ascribed to the crystallization-induced sequential reordering in the

copolymers during the annealing process.32 Fig. 5 shows the WAXD patterns of the

PETG copolymers and PET annealed at 150 oC for 1 h. The corresponding WAXD

data are represented in Table 3. A large amorphous hump appears in PETG(70/30),

implying that PETG(70/30) is mainly amorphous. However, the other four samples

16
Page 17 of 33 RSC Advances

View Article Online

RSC Advances Accepted Manuscript


DOI: 10.1039/C5RA09252C
present a series of sharp diffraction peaks related to their crystalline structures. The

crystalline structures of PET and poly(1,4-cyclohexylene dimethylene terephthalate)

(PCT) are reported to be triclinic.33-35 The three main peaks observed are assigned to

(010), ( 110 ), and (100) planes of PET, at 2θ values of about 17.68o, 22.78o, and

26.24o. PETG(85/15) shows a similar pattern to PET, indicating that PETG(85/15)

crystallized in the PET lattice. PCT shows three main diffraction peaks at 2θ = 16.63o,
Published on 08 July 2015. Downloaded by Carleton University on 10/07/2015 15:31:24.

19.10o, and 23.41o, indexing as (010), ( 1 10 ), and (100) planes.21 The diffraction

peaks shown in the patterns of PETG(50/50) and PETG(30/70) are at the same

position as that of PCT, revealing that PETG(50/50) and PETG(30/70) crystallized in

the PCT lattice. The WAXD patterns (Fig. 5) can be divided into two groups

according to the CT content in the copolymer: the PET-type crystal and the PCT-type

crystal. Under 15 mol% CT content, the PET-type crystalline structure is in

dominance, whereas above 50 mol% CT the PCT-type crystalline structure develops.

This result indicates that the crystalline structure of the PETG copolymers changes

from PET-type lattice to PCT-type lattice at a lower CT content. These findings

suggest that the rigid structure of CT unit controls the crystallization.

17
RSC Advances Page 18 of 33

View Article Online

RSC Advances Accepted Manuscript


DOI: 10.1039/C5RA09252C
Published on 08 July 2015. Downloaded by Carleton University on 10/07/2015 15:31:24.

Fig. 5 WAXD patterns of the PETG copolymers and PET annealed at 150 oC for 1 h.

As shown in Table 3, the lamellar thicknesses of the crystal planes of PET are

larger than that of PETG(85/15) at the same crystallization condition, which implies

that the crystallization rate of PET is superior to that of PETG(85/15). As for the

PETG copolymers crystallized in the PCT lattice, the lamellar thicknesses of

PETG(30/70) in the crystal planes are larger than that of PETG(50/50), indicating that

PETG(30/70) has a faster crystallization rate than that of PETG(50/50). These results

reveal that the crystallization rate of the PETG copolymers decreases at first and then

increases remarkably with the increasing CT content.


Table 3 WAXD data of the PETG copolymers and PET annealed at 150 oC for 1 ha

Samples Crystal plane 2θ (o) d (Å) β (o) L (Å)

0 11 16.56 5.35 1.10 81.06


010 17.68 5.01 1.36 65.66
PET(100/0) 111 21.80 4.07 1.32 68.07
110 22.78 3.90 — —
100 26.24 3.39 2.29 39.56
0 11 16.44 5.39 1.16 76.85
PETG(85/15) 010 17.66 5.02 — —
111 21.60 4.11 — —
18
Page 19 of 33 RSC Advances

View Article Online


110 22.68 3.92 — —

RSC Advances Accepted Manuscript


DOI: 10.1039/C5RA09252C
100 26.08 3.41 2.60 34.84
011 15.12 5.85 1.22 72.96
010 17.06 5.19 1.67 53.43
PETG (50/50) 1 10 19.42 4.57 — —
100 23.72 3.75 1.98 45.54
111 26.10 3.41 1.66 54.56
011 15.18 5.83 1.15 77.40
010 16.91 5.24 1.07 83.37
PETG (30/70) 1 10 19.10 4.64 — —
100 23.40 3.80 1.75 51.49
Published on 08 July 2015. Downloaded by Carleton University on 10/07/2015 15:31:24.

111 25.82 3.45 1.32 68.58


a θ: Bragg angle, d: interplanar distance, L: lamellar thickness. The WAXD data is only presented
to the samples which have measurable diffraction peaks.

3.4 DSC analysis

The DSC cooling traces and heating traces of the PETG copolymers and PET at

10 oC min-1 are depicted in Fig. 6. The detailed data, crystallization temperature (Tc),

cold crystallization temperature (Tcc), melting temperature (Tm), and the

corresponding enthalpies (ΔHc, ΔHcc, and ΔHm), are listed in Table 4. As shown in Fig

6a, a crystallization peak is observed in both PET(100/0) and PETG(30/70). The

crystallization peak occurs due to the melt crystallization of the samples. However,

the crystallization peak is not found in PETG(85/15), PETG(70/30), and PETG(50/50),

which suggests that these three copolymers could not melt crystallization under this

condition. These results indicate that the crystallization rate of PET(100/0) and

PETG(30/70) is faster than these three copolymers. As regards the DSC heating traces

(Fig. 6b), a cold crystallization peak and a melting peak are observed in both

PETG(85/15) and PETG(50/50). The cold crystallization peak occurs due to the cold

crystallization of the samples during the heating process. The melting peak is

19
RSC Advances Page 20 of 33

View Article Online

RSC Advances Accepted Manuscript


attributed to the crystals formed from the cold crystallization. Moreover, DOI:
just10.1039/C5RA09252C
one

melting peak is observed in PET(100/0) and PETG(30/70), indicating that these two

copolymers have completed the crystallization during the cooling process. However,

neither the cold crystallization peak nor the melting peak is found in PETG(70/30),

which suggests that PETG(70/30) is completely amorphous.


Published on 08 July 2015. Downloaded by Carleton University on 10/07/2015 15:31:24.

Fig. 6 DSC traces of the PETG copolymers and PET at 10 oC min-1: (a) cooling traces, (b)
heating traces.

The incorporation of CT in the PET chain, significantly altered the thermal

transitions of the polyester. The CT unit is a stiff and bulky structure, which destroys

the regularity of the PET chain. As shown in Table 4, although PET is semicrystalline
20
Page 21 of 33 RSC Advances

View Article Online

RSC Advances Accepted Manuscript


DOI: 10.1039/C5RA09252C
with a melting temperature of around 242.8 oC, PETG(70/30) appears to be essentially

amorphous, which is in agreement with the WAXD analysis. In fact, obvious

endothermic peaks characteristic of melting are observed in PETG(85/15),

PETG(50/50), and PETG(30/70). Also, their melting temperatures are lower than that

of PET, and meanwhile involved with much smaller associated enthalpies. The
Published on 08 July 2015. Downloaded by Carleton University on 10/07/2015 15:31:24.

crystallinity of the PETG copolymers decreases at first, and then increases remarkably

with the increasing CT content. These results clearly indicate that there is a strong

effect exerted by the CT units on the crystallinity. According to the molecular features

of CT, both chain packing and crystallization rates appeared to be hampered by

copolymerization, which will be discussed in the following part.

Fig. 7 DSC traces of the PETG copolymers and PET recorded at heating from quenched samples
for Tg observation.

The Tg of the PETG copolymers could be clearly detected in the heating DSC

traces of samples previously quenched from the melt (Fig. 7). A cold crystallization

peak is observed in PET(100/0), PETG(85/15), and PETG(30/70). The cold

crystallization peak occurs due to the cold crystallization of the quenched samples

during the heating process. However, the cold crystallization peak is not found in

21
RSC Advances Page 22 of 33

View Article Online

RSC Advances Accepted Manuscript


DOI: 10.1039/C5RA09252C
PETG(70/30) and PETG(50/50), which suggests that these two copolymers could not

crystallization under this condition. These results reveal that the crystallization ability

of PET(100/0), PETG(85/15), and PETG(30/70) is stronger than these two

copolymers.
Published on 08 July 2015. Downloaded by Carleton University on 10/07/2015 15:31:24.

Fig. 8 Dependence of glass transition and melting temperatures on the copolymer composition.

The variation of melting and glass transition temperatures as the function of

composition is plotted in Fig. 8. The Tg of the PETG copolymers increases linearly

with the increasing CT content, which also confirmed the fact that the copolymers

synthesized in this study have a random distribution of ET and CT units. Furthermore,

the CT structure is able to render stiffness and viscosity to the polymer chain, and

therefore, it will reduce the free volume as well. It is interesting to note that Tg plays

an important role in controlling the crystallization. Moreover, the copolymer

composition of around 30 mol% CT showing the minimum melting temperature

corresponds to a critical composition where the crystal transition between PET-type

and PCT-type crystal lattices occurs. Therefore, it is reasonable to believe that the

PET-type crystal develops when CT content is not above than 30 mol%, while the

PCT-type crystal develops when CT content is above 30 mol%. This result is in


22
Page 23 of 33 RSC Advances

View Article Online

RSC Advances Accepted Manuscript


DOI: 10.1039/C5RA09252C
accordance with the WAXD analysis.

Research indicated that the melting temperature of the segments in crystal region

strongly depends on the corresponding average sequence length.36 According to the

WAXD analysis, PET(100/0) and PETG(85/15) crystallize in the PET lattice, while

13
PETG(50/50) and PETG(30/70) crystallize in the PCT lattice. The C NMR results
Published on 08 July 2015. Downloaded by Carleton University on 10/07/2015 15:31:24.

reveal that the number-average sequence length nET decreases with the increasing CT

unit content. On the contrary, nCT increases gradually with the increasing CT unit

content. As shown in Table 4, the melting temperature and the melting enthalpy of the

copolymers crystallized in PET lattice, such as PET(100/0) and PETG(85/15),

decrease significantly with the decreasing average sequence length nET. As for the

copolymers crystallized in PCT lattice, such as PETG(50/50) and PETG(30/70), an

increase in the average sequence length nCT results in the higher melting temperature

and the increase of the melting enthalpy.


Table 4 Thermal properties of the PETG copolymers and PET obtained from DSCa
Sequence lengthsb Cooling Heating
Sample Codes Tc ΔHc Tcc ΔHcc Tm ΔHm Xc
nET nCT
(oC) (J g-1) (oC) (J g-1) (oC) (J g-1) (%)
PET(100/0) 185.4 43.8 — — 242.5 45.1 34.7
PETG(85/15) 7.44 1.19 — — 157.1 22.9 210.1 22.9 17.6
PETG(70/30) 3.71 1.33 — — — — — — —
PETG(50/50) 2.16 1.79 — — 156.5 5.1 197.1 9.6 7.4
PETG(30/70) 1.47 2.93 174.7 27.4 — — 243.6 25.5 19.6
a Crystallization (Tc), cold crystallization (Tcc) and melting (Tm) temperatures, as well as their
enthalpies (ΔHc, ΔHcc, and ΔHm) measured by DSC with the cooling/heating rate of 10 oC min-1.
b Calculated using the data from 13C NMR

3.5 TGA analysis

In order to investigate the thermal stability of the PETG copolymers, TGA was

23
RSC Advances Page 24 of 33

View Article Online

RSC Advances Accepted Manuscript


DOI: 10.1039/C5RA09252C
conducted under a nitrogen atmosphere using PET as the reference. Fig. 9 shows the

TGA thermograms of the PETG copolymers and PET. The detailed data, including the

initial degradation temperature at 5 wt% loss of the original weight (T5%), the

temperature of the maximum degradation rate (Td,max),37 and the weight percentage of

the residue at 800 oC, are summarized in Table 5.


Published on 08 July 2015. Downloaded by Carleton University on 10/07/2015 15:31:24.

Fig. 9 TGA thermograms of the PETG copolymers and PET at a heating rate of 20 oC min-1 in
nitrogen.

The PETG samples appear as materials with high thermal stability. Under

nitrogen flow the degradation process begins above 380 oC and occurs in a single step

up to a complete weight loss, as shown in Fig. 9. The temperatures of the initial

degradation (T5%), reported in Table 5, indicate that all samples have a very similar

behavior. Moreover, the PETG copolymers show a decreased Td,max from 449.1 oC of

PETG(85/15) to 435.7 oC of PETG(30/70). A close inspection of data reveals that

although the thermal stability of these copolymers decreased slightly with the

increasing CT content, it remains close to that of PET. The residue of PET is 12.7

wt% and the residue of the PETG copolymers ranges from 9.4 to 2.6 wt%. PETG with

the more cyclohexylene ring contents exhibit less residue at high temperature, which

might result from the reduced carbonization due to the presence of cyclohexylene
24
Page 25 of 33 RSC Advances

View Article Online

RSC Advances Accepted Manuscript


DOI: 10.1039/C5RA09252C
ring.

Table 5 TGA data of the PETG copolymers and PET in nitrogena


T5% Td,max Residue
Samples
(oC) (oC) (wt%)
PET(100/0) 407.5 451.5 12.7
PETG(85/15) 405.7 449.1 9.4
PETG(70/30) 404.6 442.2 6.5
PETG(50/50) 401.5 435.9 4.3
PETG(30/70) 401.2 435.7 2.6
Published on 08 July 2015. Downloaded by Carleton University on 10/07/2015 15:31:24.

a
T5%: the initial decomposition temperature at 5 wt% loss of the original weight. Td,max: the
temperature at the maximum degradation rate.

3.6 Tensile and shrinkage properties

A preliminary evaluation of the mechanical properties of the PETG copolymers

has been carried out by tensile testing. The stress-strain curves of the PETG

copolymers and PET at room temperature are shown in Fig. 10. The detailed data,

including Young’s modulus, yield stress, and elongation at break, are presented in

Table 6. At room temperature, the films undergo yielding followed by plastic

deformation region and ultimate failure of the films. The Young’s modulus and yield

stress of the PETG copolymers decreases with the increasing CT content. The

elongation at break of PETG(85/15) and PETG(50/50) is relatively low, whereas a

high deformation is reached for PETG(70/30) and PETG(30/70).


Table 6 Tensile properties of the PETG copolymers and PET at room temperature

Young’s modulus Elongation at break


Samples Yield stress (MPa)
(MPa) (%)

PET(100/0) 1085.5±21.8 51.3±1.0 55±10.2


PETG(85/15) 927.8±39.7 37.3±1.7 27±2.5
PETG(70/30) 839.3±45.7 36.4±2.0 151±8.6
PETG(50/50) 875.2±10.0 34.8±2.1 21±7.9
PETG(30/70) 712.2±32.3 33.1±1.2 190±34.1

25
RSC Advances Page 26 of 33

View Article Online

RSC Advances Accepted Manuscript


DOI: 10.1039/C5RA09252C

Fig. 10 Stress-strain curves of the PETG copolymers and PET at room temperature.
Published on 08 July 2015. Downloaded by Carleton University on 10/07/2015 15:31:24.

The shrinkages of the PETG copolymers with different CHDM content (0%,

15%, 30%, 50%, and 70%), stretched into 3×1 at 90 oC, tested at 90 oC, are 58%, 62%,

63%, 62%, and 48%, respectively. The testing temperature is higher than the Tg values

detected by DSC. PETG(70/30) has a higher shrinkage than other copolymers.

Presumably, this occurs due to the lowest crystallization ability of PETG(70/30)

among these copolymers and it is essentially amorphous according to the DSC

analysis. The shrinkages of the PETG copolymers and PET are more than 40% when

the testing temperature is higher than Tg. This result indicates that both the PETG

copolymers and PET could be potentially used as heat-shrinkable materials with the

operation temperature at 90 oC.

3.7 DMTA analysis

Fig. 11 shows the DMTA spectra of the PETG copolymers and PET. In the

temperature range from 60-100 oC the samples exhibit one peak of tan δ, donated as α

in order of increasing temperature. As regards the assignment of the observed

relaxation, the peak is confidently assigned to the glass-to-rubber transition. Indeed,

there is a good correspondence between the Tg values detected by DSC and the Tα
26
Page 27 of 33 RSC Advances

View Article Online

RSC Advances Accepted Manuscript


DOI: 10.1039/C5RA09252C
values determined by DMTA. The value of Tα increases with the increasing CT

content, which indicates that the copolymer with higher CT content is suitable for

applications in higher temperature environment. Moreover, from Fig. 11a, it is evident

that the intensity of the α peak of the PETG copolymers is higher than that of PET.

This might be ascribed to the cyclic structure (cyclohexylene ring) of CHDM, which

can absorb energy through the interconversion of chair and boat conformations.38
Published on 08 July 2015. Downloaded by Carleton University on 10/07/2015 15:31:24.

Fig. 11 DMTA spectra of the PETG copolymers and PET: (a) tan δ, (b) storage modulus.

As regards the storage modulus, at relatively low temperature it remains

basically unchanged with increasing temperature, whereas, in correspondence with

the glass relaxation, it shows a steep decrease of about two orders of magnitude. From

the DMTA spectra of PET(100/0), PETG(85/15), and PETG(30/70), an activated


27
RSC Advances Page 28 of 33

View Article Online

RSC Advances Accepted Manuscript


DOI: 10.1039/C5RA09252C
dispersion region corresponding to the glass relaxation can be observed, followed by a

slight pre-crystallization stiffening. The latter is presumably related to the density

modulation preceding the formation of a crystalline phase, which was already

observed by Imai et al.39 The rather steep increase of the elastic modulus observed in

the range 105-130 oC for PET(100/0), 120-145 oC for PETG(85/15), and 125-145 oC
Published on 08 July 2015. Downloaded by Carleton University on 10/07/2015 15:31:24.

for PETG(30/70) must be ascribed to the growth of crystalline domains. It is evident

that PET(100/0), PETG(85/15), and PETG(30/70) cold crystallize in the DMTA

measurement process. Moreover, a slight pre-crystallization stiffening is not observed

in the spectra of PETG(70/30) and PETG(50/50), indicating that these two

copolymers could not crystallize under the DMTA measurement condition. These

results suggest that the crystallization ability of PET(100/0), PETG(85/15), and

PETG(30/70) is stronger than that of PETG(70/30) and PETG(50/50). These results

are in accordance with the DSC analysis (Fig. 7).

4 Conclusions

A series of statistically random PETG copolymers with adjustable compositions

and suitable molecular weights were synthesized in this study. Their copolymer

composition and microstructure were determined by 1H NMR and 13


C NMR

spectroscopy. The effects of the copolymer composition and the microstructure on the

thermal properties, crystallization behavior, and thermal stability of the PETG

copolymers were systematically investigated. Their copolymer compositions were

nearly equal to the feed compositions. The number-average sequence length nET

28
Page 29 of 33 RSC Advances

View Article Online

RSC Advances Accepted Manuscript


DOI: 10.1039/C5RA09252C
decreased with the increasing CT content. On the contrary, nCT increased gradually

with increasing CT content. The crystalline structure of the PETG copolymers

changed from PET-type lattice to PCT-type lattice at a lower CT content. The

crystallinity of the PETG copolymers decreased at first, and then increased

remarkably with the increasing CT content. It was interesting to notice that the rigid
Published on 08 July 2015. Downloaded by Carleton University on 10/07/2015 15:31:24.

structure of CT controlled the crystallization. The incorporation of CT in the PET

chain, significantly altered the thermal transitions of the polyester. The Tg of the

PETG copolymers increased linearly with the increasing CT content, which also

confirmed the fact that the copolymers synthesized in this study had a random

distribution of ET and CT units. The melting temperature of the segments in crystal

region strongly depended on the corresponding average sequence length. An increase

in the average sequence length resulted in a higher melting temperature and an

increase in the melting enthalpy.

Acknowledgements

This work was supported by the Innovation Foundation for Graduate Students of

Jiangsu Province (KYLX_0744) and the Priority Academic Program Development of

Jiangsu Higher Education Institutions (PAPD). The authors would like also to express

their appreciation to Mr. Hao Wang and Mr. Mingfu Chen for their help in the

preparation of the manuscript.

29
RSC Advances Page 30 of 33

View Article Online

RSC Advances Accepted Manuscript


DOI: 10.1039/C5RA09252C
References

1 I. N. Strain, Q. Wu, A. M. Pourrahimi, M. S. Hedenqvist, R. T. Olsson and R.

L. Andersson, J. Mater. Chem. A, 2015, 3, 1632-1640.

2 G. Xu, S. Qin, J. Yu, Y. Huang, M. Zhang and W. Ruan, RSC Adv., 2015, 5,

29924-29930.
Published on 08 July 2015. Downloaded by Carleton University on 10/07/2015 15:31:24.

3 Y. Rao, J. Greener, C. A. Avila-Orta, B. S. Hsiao and T. N. Blanton, Polymer,

2008, 49, 2507-2514.

4 M. R. Patel, J. V. Patel and V. K. Sinha, Polym. Degrad. Stab., 2005, 90,

111-115.

5 X. K. Jing, X. S. Wang, D. M. Guo, Y. Zhang, F. Y. Zhai, X. L. Wang, L. Chen

and Y. Z. Wang, J. Mater. Chem. A, 2013, 1, 9264-9272.

6 Y. Jahani, M. Ghetmiri and M. R. Vaseghi, RSC Adv., 2015, 5, 21620-21628.

7 S. Fakirov, E. W. Fischer, R. Hoffmann and G. F. Schmidt, Polymer, 1977, 18,

1121-1129.

8 J. M. Huang, P. P. Chu and F. C. Chang, Polymer, 2000, 41, 1741-1748.

9 T. Higashioji, T. Tsunekawa and B. Bhushan, Tribol. Int., 2003, 36, 437-445.

10 US pat., 5 250 333, 1993.

11 D. PR. Kint and S. Muñoz-Guerra, Polym. Int., 2003, 52, 321-336.

12 C. Japu, A. Martínez de Ilarduya, A. Alla and S. Muñoz-Guerra, Polymer,

2014, 55, 2294-2304.

13 Y. Tsai, C. H. Fan, C. Y. Hung and F. J. Tsai, J. Appl. Polym. Sci., 2008, 109,

2598-2604.
30
Page 31 of 33 RSC Advances

View Article Online

RSC Advances Accepted Manuscript


DOI: 10.1039/C5RA09252C
14 J. Yang, W. Li, A. Yu, P. Xi, X. A. Huang and S. Li, J. Appl. Polym. Sci., 2009,

111, 2751-2760.

15 A. Bigi, E. Boanini, C. Capuccini, M. Fini, I. N. Mihailescu, C. Ristoscu, F.

Sima and P. Torricelli, Biomaterials, 2009, 30, 6168-6177.

16 N. Vasanthan, N. J. Manne and A. Krishnama, Ind. Eng. Chem. Res., 2013, 52,
Published on 08 July 2015. Downloaded by Carleton University on 10/07/2015 15:31:24.

17920-17926.

17 R. J. Müller, H. Schrader, J. Profe, K. Dresler and W. D. Deckwer, Macromol.

Rapid Comm., 2005, 26, 1400-1405.

18. W. K. Shih, Polym. Eng. Sci., 1994, 34, 1121-1128.

19. W. Huang, Y. Wan, J. Chen, Q. Xu, X. Li, X. Yang, Y. Li and Y. Tu, Polym.

Chem., 2014, 5, 945-954.

20. I. Ward, Nature, 1957, 180, 141-142.

21 T. E. Sandhya, C. Ramesh and S. Sivaram, Macromolecules, 2007, 40,

6906-6915.

22 Y. G. Jeong, W. H. Jo and S. C. Lee, Macromolecules, 2000, 33, 9705-9711.

23 R. Yamadera and M. Murano, J. Polym. Sci., Part A: Polym. Chem., 1967, 5,

2259-2268.

24 C. Japu, A. Martinez de Ilarduya, A. Alla, M. G. Garcia-Martin, J. A. Galbis

and S. Muñoz-Guerra, Polym. Chem., 2013, 4, 3524-3536.

25 R. H. Guo, S. Q. Jiang, C. W. M. Yuen and M. C. F. Ng, J. Mater. Sci.: Mater.

Elect, 2009, 20, 735-740.

26 O. Prasad, L. Sinha, N. Misra, V. Narayan, N. Kumar and J. Pathak, J. Mol.

31
RSC Advances Page 32 of 33

View Article Online

RSC Advances Accepted Manuscript


DOI: 10.1039/C5RA09252C
Struct.: Theochem., 2010, 940, 82-86.

27 R. Jayakumar, M. Rajkumar, R. Nagendran and S. Nanjundan, J. Appl. Polym.

Sci., 2002, 85, 1194-1206.

28 A. Bessadok, S. Roudesli, S. Marais, N. Follain and L. Lebrun, Compos. Part

A: Appl. S., 2009, 40, 184-195.


Published on 08 July 2015. Downloaded by Carleton University on 10/07/2015 15:31:24.

29 S. S. Umare, A. S. Chandure and R. A. Pandey, Polym. Degrad. Stab., 2007,

92, 464-479.

30 H. Nakamura, H. Sakai, S. Aoshima and M. ABE, J. Oleo Sci., 2002, 51,

781-787.

31 S. W. Lee, W. Huh, Y. S. Hong and K. M. Lee, Korea Polym. J., 2000, 8,

261-267.

32 Z. Denchev, M. Sarkissova, H. J. Radusch, T. Luepke and S. Fakirov,

Macromol. Chem. Phys., 1998, 199, 215-221.

33 Y. Kitano, Y. Kinoshita and T. Ashida, Polymer, 1995, 36, 1947-1955.

34 R. P. Daubeny and C. Bunn, Proc. R. Soc. Lond. A, 1954, 226, 531-542.

35 C. Boye, J. Polym. Sci., 1961, 55, 275-284.

36 A. A. Deschamps, D. W. Grijpma and J. Feijen, Polymer, 2001, 42,

9335-9345.

37 Y. Zhang, L. Chen, J. J. Zhao, H. B. Chen, M. X. He, Y. P. Ni, J. Q. Zhai, X. L.

Wang and Y. Z. Wang, Polym. Chem., 2014, 5, 1982-1991.

38. H. Ni, J. L. Daum, P. R. Thiltgen, M. D. Soucek, W. J. Simonsick, W. Zhong

and A. D. Skaja, Prog. Org. Coat., 2002, 45, 49-58.

32
Page 33 of 33 RSC Advances

View Article Online

RSC Advances Accepted Manuscript


DOI: 10.1039/C5RA09252C
39. M. Imai, K. Kaji and T. Kanaya, Macromolecules, 1994, 27, 7103-7108.
Published on 08 July 2015. Downloaded by Carleton University on 10/07/2015 15:31:24.

33

You might also like