• Sonuç bulunamadı

Synthesis of a novel polyester-ether copolymer and its derivatives as antistatic additives for thermoplastic films

N/A
N/A
Protected

Academic year: 2021

Share "Synthesis of a novel polyester-ether copolymer and its derivatives as antistatic additives for thermoplastic films"

Copied!
14
0
0

Yükleniyor.... (view fulltext now)

Tam metin

(1)

Polymer Testing 81 (2020) 106214

Available online 12 November 2019

0142-9418/© 2019 Elsevier Ltd. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).

Synthesis of a novel polyester-ether copolymer and its derivatives as

antistatic additives for thermoplastic films

Hüsnü Kemal Gürakın

a,b

, Ahmet Celalettin Turan

b

, Hüseyin Delig€oz

a,* aEngineering Faculty, Chemical Engineering Department, Istanbul University Cerrahpas¸a, Istanbul, 34320, Turkey bSetas¸ Kimya San A.S¸., Çerkezk€oy, Tekirda�g, Turkey

A R T I C L E I N F O Keywords:

Polyester-block-ether copolymer Antistatic polyester film Ionic liquid

A B S T R A C T

Development of antistatic thermoplastic elastomer composites is gaining much interest in scientific and indus-trial aspects. In this contribution, a novel type of polyester-block-ether copolymer (PEBE), PEBE with ionic liquid (PEBE-IL) and quaternized PEBE (PEBE-Q) were synthesized, characterized and compounded with a commercial bottle grade polyester (PET) resin. It was found that the surface resistivity of the PET composite films was around 109–1010 Ω/sq. The addition of the PEBE copolymers as an antistatic agent into PET matrix resulted in almost 106 times enhancement in surface resistivity compared to neat PET film. Among the PET films with an antistatic agent, it was emphasized that PEBE-IL and ionic liquid doping to PET matrix led to a relatively lower crystal-lization degree, surface resistivity, contact angle and breaking force with a translucent appearance compared to neat PET film. Furthermore, it was illustrated that antistatic PET films with fine-tune physical, mechanical and morphological properties can be achieved by choosing appropriate antistatic agent type and amount.

1. Introduction

Static charges accumulated on the surface of the polymers can cause some problems, especially in the packaging applications. Sensitive electronic devices such as hard disks and motherboards can be severely damaged by a short circuit caused by electrostatic discharge [1]. Most of the polymers used in the packaging industry are inherently insulators. In particular, the relatively less polar molecular structure of the poly-ethylene terephthalate (PET) (which is one of the most common pack-aging polymers) makes the polymer itself hydrophobic and facilitates the accumulation of static charges. Many studies in the past 30 years have shown that combining the polar comonomers with the polyester chain resulted in an enhancement of the wetting and static dissipative properties of the neat PET. For example, the addition of poly(ethylene glycol) (PEG) [1–3], polytetramethylene ether glycol (PTMEG, pTHF) [4], diacids like adipic [3], isophthalic [2,3], and sulfoisophthalic acid [4] to the PET polymer allows faster absorption of the moisture and eases the drain of the static charges while disordering the crystalline structure. The ether groups existing at the polyether diol molecules are substantially polar to draw moisture and static charges are more easily distributed thanks to the moisture layer absorbed on the surface of the polymer [5]. Furthermore, aliphatic diacids weaken the crystallinity of

the polymer backbone and give an amorphous structure which leads to faster discharge of static electricity [3].

Another approach to gain the antistatic effect is to blend static dissipative copolymers with the insulative resin via melt blending. This theory is based on the formation of interpenetrating continuous conductive polymer networks. In the literature, polyether block amides were widely used for this application but the antistatic copolyesters based on polyester-ethers have been scarcely reported. In a study by Li et al. [2], the polyester copolymers were obtained from diacids, namely isophthalic and terephthalic acid and the different portions of PEGs having different molecular weights. Final polyester fibers containing 18 wt% of PEG 1000 showed a volume resistivity of 5.8 � 1010 Ω/cm. In

another study, polyester-ether copolymers having PEG and other hy-drophilic segments were used to gain hyhy-drophilicity to the polyester bulk materials. They reported that the introduction of 12 wt% of PEG comonomer into the PET matrix increased the moisture absorptivity value of the neat polyester from 0.56% to 0.96% [3]. In a work pub-lished by George et al. [4], sulfoisophthalic acid was used as a como-nomer in the production of the antistatic copolyester. The final films having 30 wt% of antistatic copolymer in a polyethylene terephthalate glycol (PETG) showed surface resistivity values ranging from 1011 Ω/sq

to 109 Ω/sq. * Corresponding author.

E-mail address: hdeligoz@istanbul.edu.tr (H. Delig€oz).

Contents lists available at ScienceDirect

Polymer Testing

journal homepage: http://www.elsevier.com/locate/polytest

https://doi.org/10.1016/j.polymertesting.2019.106214

(2)

Ionic liquids can be defined as salts that are liquid at room temper-ature and can be mainly used as conductive fillers [6]. They are also very promising antistatic agents for different thermoplastic and thermoset polymers. Melt blending of ionic liquids led to great improvement in the antistatic properties of some polymers such as polyurethanes [7–9], polycarbonate [10], polyamide [11] and other polymers [12–15]. For example; in a study published by Guo et al. [16], 1,3-bis(2-hydrox-yethyl) imidazolium chloride grafted PET copolymer was synthesized via polycondensation method and the changes in hydrophilicity and conductivity of this polymer were investigated as a function of the ionic liquid content. The conductivity values up to 15 μS/cm were reported.

As far as we know from the literature quaternization of polymer backbone resulted in more polarity and facilitated charge transfer due to the positive charge of the ammonium cation [17]. Also, it was found that quaternized compounds became more hydrophilic and hygroscopic due to the moisture adsorption of these polar sites [18]. In the literature, many studies were reported with the cationic surface treatment tech-nique. For example, in a study published by Zaman et al. [19], cation-ically modified cellulose was used as a surface finish chemical for polyester fabrics to gain the hydrophilicity and the water absorption of the treated fabrics significantly increased compared to untreated fabrics. In this study, a novel type of thermoplastic polyester-ether block copolymer (PEBE) from PEG, PTMEG, isophthalic acid, adipic acid, sulfoisophthalic acid, and polyether 1,3 diol was successfully prepared and characterized. To the best of our knowledge, the use of polyether 1,3 diol as a comonomer in an antistatic polyester-ether copolymer composition and 2,3-epoxypropyltrimethyl ammonium chloride as a quaternization agent for this copolymer have not been reported formerly. Quaternary alkylammonium sulfonimide based ionic liquid and quaternized polyester-ether block copolymer were also used as antistatic discharge promoters in PEBEs and the effects of these anti-static agents on surface resistivity of PET composite films were comparatively investigated. As far as we know from the literature, this type of commercially available ionic liquid was not used as an antistatic agent for PET-based cast films before. Furthermore, physicochemical, thermal and morphological properties of the PEBE and PET composite films prepared using twin-screw and single-screw extruders respectively were also extensively characterized.

2. Experimental

2.1. Materials

Polyethylene glycol (PEG 4000 g/mol), mono ethylene glycol and diethylene glycol were purchased from DOW Chemicals. Isophthalic acid and adipic acid were purchased from Eastman Co. and Brenntag Turkey, respectively. Poly(tetrahydrofuran) (2000 g/mol) was pur-chased from BASF while 5-sulfoisophthalic acid was supplied from Yangtai Chemicals Co. Polyether 1,3 diol (Tegomer D3403, 1200 g/mol) was supplied by EVONIK, quaternary alkylammonium sulfonamide based ionic liquid (FC 5000) was kindly supplied by 3 M, 2,3 epox-ypropyl trimethyl ammonium chloride (GTMAC) was supplied by SKW Quab Chemicals Inc. Bottle grade 0.82 IV polyester chip was purchased from Indorama Ventures and it was dried at 160 �C for 6 h in a dryer

before use. In order to avoid hydrolysis reactions with the polyester end groups, the water content of 2,3 epoxypropyl trimethyl ammonium chloride was evaporated at 20 �C with 1 mbar vacuum at 24 h. All other

chemicals were used as received.

2.2. Methods

2.2.1. Synthesis of PEBE copolymer

A novel polyether block ester copolymer (PEBE) was synthesized in a two-stage process of transesterification and followed by poly-condensation. The reaction was carried out in a two-liter stainless steel reactor (Amar Instruments) with a condenser, vacuum pump, and

nitrogen inlet. Total glycol and total acid mole ratio was kept at 1.32 in the formulation. For synthesis of the copolymer (abbreviated as PEBE), reactor was loaded with 458 g (~0.114 mol) of PEG 4,000, 212 g (~1.276 mol) of isophthalic acid (IPA), 123 g (~1.981 mol) of mono-ethylene glycol (MEG), 15 g (~0.141 mol) of dimono-ethylene glycol (DEG), 40 g (~0.149 mol) of sulfoisophthalic acid (SIPA), 35 g (0.029 mol) of polyether 1,3 diol, 47 g (~0.322 mol) of adipic acid and 70 g (0.035 mol) of polyTHF 2000.300 ppm antimony trioxide (Sb2O3) was

used as a catalyst. The flask was purged with nitrogen and stirred under nitrogen atmosphere at 232 �C for 4 h under 3 bar. After 4 h of

esterifi-cation, the pressure was gradually reduced from 3 bar to 1 bar and 10 mm Hg (~0.0133 bar) for 30 min. At the final stage, the polymer was kept at 245 �C for 1.5 h to perform polycondensation. It was replaced

with a nitrogen atmosphere and allowed to cool for 30 min before the removal of polymer from the reactor. The synthesis steps of the PEBE are presented in Fig. 1. Furthermore, the intrinsic viscosity value of the PEBE copolymer was measured after each production by dissolving the polymer in phenol/1,2-dichlorobenzene (50/50 by weight) solution and determined to be between 0.58 and 0.61 dL/g according to ASTM D4603-18.

2.2.2. Preparation of PEBE-IL and PEBE-Q

In order to reduce the resistivity of PEBE, ionic liquid and quater-nization agent were added to PEBE copolymer synthesized according to the method described above. After the last stage of the polymerization, the quaternary alkylammonium sulfonimide based ionic liquid and 2,3 epoxypropyl trimethyl ammonium chloride were added individually to PEBE at 100 �C and mixed for an additional 60 min to obtain PEBE-IL

and PEBE-Q, respectively. PEBE-IL and PEBE-Q contain 10 wt% of ionic liquid or quaternization agent and 90 wt% of PEBE copolymer. After the synthesis of PEBE and its derivatives, they poured into a mold and the surface resistivity of these products were measured with a two- point probe method according to ASTM D 257 at 20 �C and 65% relative

humidity (RH).

2.2.3. Preparation of Copolyester–PET compounds

PEBE, PEBE-IL, and PEBE-Q were melt-mixed and blended with a commercial bottle grade PET resin in a co-rotating intermeshing (Leis-tritz, L/D: 45) double-screw extruder. The barrel temperature profile of the extruder was 285�C-280C-275C-270C-275C-260C–255 C from

the feed throat to the die, and the melt temperature of the blends was 273 �C. The employed screw speed was 350 rpm. The final product was

obtained in the form of granules from a filament die. All compounds were prepared at 15% wt of antistatic agent (PEBE, PEBE-IL, and PEBE- Q separately) with bottle grade commercial PET.

2.2.4. Preparation of PET composite films

Cast film samples were prepared by using a laboratory scale Dr. Collin single screw extruder equipped with a Maddock mixing screw. The employed screw speed was 160 rpm. The barrel temperature profile of the extruder was 250�C–255C-260C-270C-275C-285C–290 C

from the feed throat to the die, and the melt temperature of the blends was 271 �C. All cast films were produced at 100-μm thickness.

Com-pounds with 15 wt% PEBE, PEBE-IL and PEBE-Q loading which were pre-prepared with a twin-screw extruder were diluted to the final compositions of 1-2-3-4-5-6-8-10-15 wt% antistatic agent. The surface resistivity values of the films were measured by constant voltage two- point probe method at an applied voltage of 500 V according to ASTM D 257 at 20 �C and 65% RH. The films of PET and PEBE was donated as

PET/PEBE while PET/(PEBE-IL) showed the composition of PET and PEBE-IL.

2.3. Characterization

2.3.1. Chemical and thermal analysis

(3)

IR Prestige-21 FTIR spectrometer with attenuated total reflectance (ATR) unit on a resolution of 4 cm 1. The scanning range was altered

from 600 to 4000 cm 1. X-ray diffraction (XRD) measurements were

conducted by a Bruker AXS advance powder diffractometer with a Cu Kα

radiation source between 10 and 90 theta angles. X-ray Fluorescence Spectrometry (XRF) analyzes of the samples were conducted on EDX- 8000 (Shimadzu) spectrometer.

Differential scanning calorimetry (DSC) measurements of the syn-thesized antistatic copolymers, PET composite films, and neat PET were carried out in a DSC 2.0 (METTLER TOLEDO) analyzer. To eliminate the thermal history, samples were heated in the first heating run ( 70 �C–300 C), cooling (300 C–70 C) and second heating

( 70 �C–300 C) with a rate of 10 C/min. Glass transition (Tg) and

melting temperature (Tm) peaks were determined from the second heating run. Approximately 10–20 mg of sample was placed into the aluminum pans and measurements were performed under 50 ml/min of nitrogen flow. Glass transition temperatures were calculated from the peak of the first derivative of the inflection in the second heating curve. Crystallinity degree (%) values of the PET composite films were

calculated from the first heating cycle according to Equation (1). The melting enthalpy (ΔHm) of 100% crystalline PET was taken as 140 J/g

[20]. Since all antistatic PET films were semicrystalline, the cold crys-tallization enthalpy (ΔHcc) values were taken into account.

Crystallinity % : ΔHm ΔHcc

ΔHð100% Crystalline PETÞ

� 100 Equation 1

TG analysis of the copolymers, PET composite films, and neat PET were carried out using a Shimadzu DTG-60 series differential thermal- thermogravimetric analyzer at a heating rate of 10 �C/min from room

temperature to 600 �C. The onset of thermal decomposition was

deter-mined from the temperature at which 10% weight loss occurred.

2.3.2. Surface resistivity measurements

A surface resistivity meter (Norma Instruments) was used to measure the surface resistivity of the copolymers and PET composite films. The surface resistivity values of all samples were measured by a constant voltage two-point probe method at an applied voltage of 500 V ac-cording to ASTM D 257. The surface resistivity of the samples was Fig. 1. Schematic representation of PEBE synthesis (a) comonomer structures (b) repeating units of the PEBE copolymer with possible formations.

(4)

measured at 20 �C and 65% relative humidity (RH) conditions with a

standard deviation of �1% for all the samples.

2.3.3. Surface and optical measurements

Total transmittance and haziness measurements of the PET com-posite films were conducted with Hunterlab UltraScan Pro dual-beam spectrophotometer with a standard deviation of �0.5%. The surface morphology of neat PET and PET composite films were examined using a scanning electron microscope (FEI Quanta FEG 450). Before imaging, all the film samples were coated with a conductive layer of sputtered gold. Contact angle measurements of the samples were conducted by using an Attension Theta (Biolin Scientific) tensiometer. Contact angles were evaluated on both sides of the films with distilled water. The contact angle values reported here were calculated from the average of three measurements with a standard deviation of �2�.

2.3.4. Moisture absorption measurements

Moisture absorption of the antistatic PET films was also determined. For this purpose, the prepared films were firstly weighted accurately. Subsequently, the films were kept under 65% relative humidity at 20 �C

for one week. Finally, the moisture uptake was obtained from Equation (2).

Moisture ​ Uptake ​ %: ​ W W0

W0

�100 Equation 2

where W0 is the mass of the dried film before moisture treatment and W

is the mass of the film after moisture treatment.

2.3.5. Tensile properties

Elongation at break (%) and the breaking force (N/15 mm) values of the 100-μm thick PET-based composite films prepared by cast film

extruder were measured by using Zwick Roell Z2.5 with a tensile rate of 100 mm/min in the machine direction within a standard deviation of �1%. The dimensions of the films were 150 mm � 15 mm and the crosshead distance was 60 mm. Before the tests, the composite films were stored at 20 �C and 65% relative humidity for 24 h.

3. Results and discussion

3.1. Chemical and thermal properties of copolymers

Fig. 2 shows the FTIR spectra of PEBE, PEBE-Q, PEBE-IL, and neat PET. It can be clearly seen that C¼O stretching appeared in the range of 1750–1735 cm 1 for PEBE-based copolymers due to the formation of

ester bonds. It was illustrated that the ester-exchange reaction occurred

successfully for all compositions. In comparison with the neat PET, the C-H stretching bands of PEBE, PEBE-Q, and PEBE-IL copolyesters seen around 2875 cm 1 indicated the existence of long polyethylene oxide

units and confirmed PEG presence in the copolyester chain [21]. The band at 872 cm 1 observed in all copolyesters has belonged to

the C-H deformation of two adjacent coupled hydrogens on the aromatic ring [22]. Sharp bands at 1245 1 cm for PET and 1294 1 cm and

1240 cm 1 for other copolyesters were attributed to valence vibrations of the C-C bonds [23]. The C-H stretching vibration of GTMAC at 1482 cm 1 was overlapped with the intermolecular band of the PEBE

copolymer (C–H stretching at 1464 cm 1), while the stretching vibration of C-O of oxirane group at 914 cm 1 in the FTIR spectrum of GTMAC was

not observed at the PEBE-Q copolymer (Fig. 2) [24,25]. In addition to FTIR graphs of PEBE based copolyesters, carboxyl end group equivalent values of the PEBE and PEBE-Q copolymers were also measured and a significant decrease in the carboxylic end groups was determined after quaternization reaction. This value was measured as 96.8 meq/kgs for the PEBE and 4.12 meq/kgs for PEBE-Q according to ASTM D7409-15. These results demonstrated that the reaction between PEBE copolymer and GTMAC was achieved and PEBE-Q was successfully synthesized.

Concerning infrared spectra of ionic liquid, namely alkylammonium sulfonimide and 2,3 epoxypropyl trimethyl ammonium chloride (GTMAC), they are illustrated in Fig. 3. In the FTIR spectrum of GTMAC, the broad band at 3300 cm 1 was assigned to the O-H stretching vi-bration of water. Also, a strong absorption band at 1482 cm 1 appeared

in the IR spectrum of GTMAC due to C-H symmetric bending of methyl groups on the quaternary ammonium substituent while the absorption bands at around 950 cm 1 and 914 cm 1 showed the characteristics of

the epoxide group [25,26].

In the FTIR spectrum of ionic liquid, -CF3 symmetric stretching was

attributed to the bands at 1182 cm 1 and 1136 cm 1. The medium

intense band located at 1473 cm 1 in the spectrum was assigned to the

SO2 asymmetric stretching [27]. In the FTIR spectrum of PEBE-IL, the

bands at 1136 cm 1 and 1473 cm 1 from the ionic liquid were

over-lapped with the intermolecular C-H stretching bands of the PEBE copolyester but -CF3 symmetric stretching vibration band at 1182 cm 1

was still visible in PEBE-IL [27].

X-ray diffraction (XRD) curves of the PEBE, PEBE-IL and PEBE-Q copolymers are shown in Fig. 4. According to these results, PEBE and PEBE-Q copolymers showed similar diffraction patterns. Strong and sharp peaks at 19�, 22.8�, 25.8�and 44�(2θ) and relatively small peaks

at 35.6�, 39, 64and 81.3(2θ) were very similar in PEBE and PEBE-Q

copolymer. The peaks at 19�, 22.8and 25.8(2θ) were probably due to

the PEG presence in the copolymers [28]. On the other hand, the first three strong peaks observed for PEBE and PEBE-Q were shifted to the

(5)

higher degrees and the intensities of these peaks reduced for PEBE-IL copolymer. Three main peaks of the PEBE-IL were at 20.4�, 24.2�, and

27�(2θ) and no small peaks were observed, unlike PEBE and PEBE-Q.

This result demonstrated that the addition of ionic liquid clearly reduced the crystallinity and made the PEBE-IL copolymer more amor-phous compared to other derivatives.

X-ray Fluorescence Spectrometry (XRF) investigation was carried out to screen the expected elements in the PEBE, PEBE-IL and PEBE-Q co-polymers and the related graphs are depicted in Fig. 5. The ionic liquid was also analyzed to see the fluorine (F) and sulfur (S) arisen from bis (trifluoromethane sulfonyl) imide anion.

The sulfur peak of the PEBE (Fig. 5b) has corresponded to sulfoi-sophthalic acid comonomer and the polycondensation catalyst was the reason for the antimony (Sb) peak. According to Fig. 5a and c, it was seen that the PEBE-IL contained fluorine from the ionic liquid. The amount of the sulfur moiety was increased compared to PEBE because of the addition of the sulfonyl groups. In addition to the sulfur peak, the chlorine (Cl) and sodium (Na) peaks were clearly seen in the XRF spectra of PEBE-Q (Fig. 5d) due to the presence of GTMAC.

DSC thermograms of the PEBE, PEBE-IL, and PEBE-Q obtained from first heating ( 70 �C–300 �C), cooling (300 �C to 70 �C) and second

heating ( 70 �C–300 C) cycles at 10 C/min rate are shown in Fig. 6

and the different measurable quantities are shown in Table 1. In the first heating process (Fig. 6a), no cold crystallization peaks were observed for any of the three copolymers which indicated that at the beginning all polymers were in crystalline structure but in the sec-ond heating run, PEBE and PEBE-Q copolymers showed the cold crys-tallization peaks. For PEBE-Q (Fig. 6c), it was found that the cold crystallization (Tcc) peak was shifted to a lower temperature compared to PEBE and it reduced from 5.2 �C to 12.7 C. From this finding, we

can conclude that quaternization slightly accelerated the nucleation process of the copolyester similar to a nucleating agent. According to Fig. 6b, broad exothermal peaks appeared between 27 �C and 37 C,

indicating the poor crystallization ability of all copolymers during the cooling crystallization process.

In the second heating run, PEBE and PEBE-Q polymers displayed a relatively sharp melting endotherm enthalpies (ΔHf) of 52.2 J/gr at 40.5 �C and 54.1 J/gr at 42.5 C while PEBE-IL copolymer showed an

amorphous curve without any melting or crystallization peak. From this observation, we can conclude that PEBE-IL copolyester was completely amorphous and ionic liquid had a plasticizing effect on the polymer [29]. On the other hand, the cooling crystallization (Tcc) and glass

transition (Tg) temperatures of PEBE-IL calculated from the cooling run (Fig. 6b) and second heating run (after 15 min of cooling at 70 �C)

(Fig. 6d) respectively were slightly higher than those of PEBE and PEBE-Q. This can be explained by the plasticizing effect of ionic liquid improved the mobility of polyester chains and played a key role in accelerating the rate of crystallization [30,31]. An increase in Tg has been previously reported upon the addition of ionic liquids into sodium and lithium salts [32]. The increase in Tg with the addition of ionic liquid may be due to the enhancing ion-ion attractions between ionic liquid and sodium salt of sulfoisophthalic acid from PEBE polymer chain [33]. According to Fig. 6d and the data from Table 2, after 15 min of cooling at 70 �C, PEBE-IL had enough time to partially crystallize, so in

the second heating run, the glass transition, melting and crystallization peaks have appeared for PEBE-IL copolymer.

Thermal stability and onset degradation (~10% weight loss) tem-peratures of the PEBE based copolymers were determined by TGA. The influences of ionic liquid and quaternization on the thermal stability of PEBE copolyester is shown in Fig. 7. PEBE-IL and PEBE-Q exhibited lower thermal stability than that of neat PEBE. The onset degradation temperature for PEBE was 369.8 �C and only 2% of its weight was lost at

the extrusion melting temperature range (~273 �C). On the other hand,

PEBE-IL had an onset degradation temperature of 328 �C while PEBE-Q

showed the onset decomposition temperature of 292.2 �C. Thus, the

lowest thermal stability was observed for PEBE-Q among the prepared PEBE copolyesters. This is most probably due to the degradation of quaternary ammonium groups at ~285 �C [34,35]. In short, it was

concluded that all the copolyesters were thermally stable up to 270 �C

and they showed no thermal degradation during extrusion compounding and film preparation steps. Further, the addition of ionic liquid weakens the polymer chains and also the decomposition trigger temperature for the ionic liquid itself is expressed as 320 �C [36].

3.2. Chemical and thermal properties of PET composite films

FTIR analysis of some selected PET composite films containing 10 wt % PEBE, 10 wt% PEBE-IL, 10 wt% PEBE-Q and 5 wt% ionic liquid were conducted and the obtained spectra are shown in Fig. 8. For comparison, the FTIR spectrum of neat PET is also given. According to these results, the neat PET had four main peaks, the absorption peak at 1720 cm 1 was

attributed to carbonyl stretching vibration of C¼O terephthalate units while the peaks at 1250 cm 1 and 1090 cm 1 were trans and gauche

stretching of the ester C-O bonds. The peak at 729 cm 1 was

corre-sponded to out of plane C-H bending of the aromatic nuclei [23,37]. Since the composite films consisted of 90 wt% of PET, all antistatic films had shown similar peaks compared to neat PET. The composite films with PEBE, PEBE-Q, and PEBE-IL showed slight peaks at 2880 cm 1

attributed to the C-H stretching of the PEG moiety in copolymer com-positions [28].

To investigate the thermal transition temperatures of PET composite films, DSC analysis of the selected antistatic PET films were measured and the obtained results are given in Fig. 9 and Table 2 in detail. Three of the selected films comprising 10 wt% PEBE, PEBE-Q, and PEBE-IL, while one film was with 5 wt% ionic liquid. Neat PET reference film was also measured as a reference.

In the first heating and cooling cycles (Fig. 9a and b), DSC curves of the PET composite films showed that the cold crystallization and cooling crystallization temperature of all antistatic PET films decreased to lower temperatures compared to the neat PET film. For example, the cooling crystallization temperature of the neat PET film was measured as 181.1 �C while the composite films containing 10 wt% PEBE-Q and 5 wt

% IL had crystallization points of 207.3 �C and 203.6 C, respectively.

Probably, it can be explained by the improved mobility and re- arrangement of the PET chains due to the soft segments of antistatic copolymer structured agents [38]. Furthermore, from Fig. 9a and Table 2, the crystallinity degree (%) of the PET composite films and neat PET film were determined according to Equation (1) as a function of Fig. 4. XRD spectra of the PEBE, PEBE-IL and PEBE-Q copolymers.

(6)

antistatic agent type comparatively. The crystallinities of the PET films with 10 wt% PEBE copolyester and its derivatives were determined as in the range of 4.3% and 8.6% while the crystallinity of neat PET film was 13.1%. Therefore, it can be clearly said that crystallinity % of antistatic agent containing PET films was lower compared to neat PET film and the introduction of antistatic agents significantly reduced the crystallinity level. Moreover, the PET composite film with 5 wt% ionic liquid showed

the lowest crystallinity degree (3%) among the all prepared antistatic agent containing films. These results revealed that the use of copolyester based antistatic agents declined the crystallinity level and led to a more amorphous structure. Please note that the obtained crystallinity degrees can be acceptable for cast PET films although some oriented polymeric films had a higher crystallinity degree.

In addition to the influence of antistatic agent addition on the Fig. 5. XRF spectra of the antistatic additives (a) IL, (b) PEBE, (c) PEBE-IL, (d) PEBE-Q. The peaks rhodium (Rh), copper (Cu) and iron (Fe) are due to the XRF sample

holder and detector.

Fig. 6. DSC thermograms of PEBE, PEBE-IL and PEBE-Q: (a) first heating from 70 C to 300 C, (b) cooling from 300 C to 70 C (c) second heating from 70 C to

(7)

crystallization degree of PET matrix, the glass transition temperatures of the antistatic PET composite films were determined from the second heating run. The results indicated that the addition of antistatic additive caused to lower Tg values compared to the neat PET (85.1 �C). The

highest Tg value of pristine PET film can be attributed to the relatively larger crystallinity degree compared to PET composite films with anti-static agents. On the other hand, glass transition temperatures of PET/ PEBE, PET/(PEBE-Q), and PET/(PEBE-IL) were found to be almost

76.1 �C, 81 C and 75 C respectively (see in Table 2). The relatively

higher Tg value of PET/(PEBE-Q) can be explained by the formation of more crystalline and rigid structure compared to other antistatic agent containing PET composite films. By the way, it was found that the use of only ionic liquid as a conductive agent resulted in a more flexible structure due to its lower crystallinity as shown in Table 2. As it was discussed in the section of the mechanical properties, ionic liquid addition into the PET matrix caused less rigidity due to its function as a plasticizer. Hence, Tg of PET/IL was found to be the lowest one (73 �C).

This decrease in Tg is quite consistent with the results of previous studies [39]. According to Fig. 9c, all-composite films were showed almost similar melting points around ~249 �C and no cooling crystallization

peak was observed because all films were crystallized during the pre-vious cooling cycle.

Thermogravimetric investigation of the selected antistatic films was carried out to understand the effect of these additives on the thermal properties of the composite films. TGA curves of the four selected PET composite films (10 wt% PEBE, 10 wt% PEBE-IL, 10 wt% PEBE-Q, 5 wt% IL) and neat PET reference film are presented in Fig. 10. Generally speaking, the thermal degradation behavior of these selected films were quite consistent with the TGA values of the copolymers explained in the previous section. For example, the onset degradation temperature (~10% weight loss) of the 10 wt% PEBE-Q containing antistatic PET film was found to be 348 �C while the neat PET film was measured as

415 �C. This lower thermal stability of 10 wt% PEBE-Q containing PET

composite film (PET/PEBE-Q) was due to the less thermal durability of ammonium moiety [34,35]. PEBE, PEBE-IL, and IL bearing PET com-posite films had thermal decomposition values of 401 �C, 396 C, and

381 �C, respectively.

3.3. Surface characteristics of antistatic films 3.3.1. Influence of the copolymer content

In order to study the influence of quaternization and ionic liquid doping on surface resistivity (Rs) of PEBE copolymer, firstly the elec-trical resistivities of bulk form antistatic copolymers were measured. It was pointed out that the surface resistivities of the copolyesters with both ionic liquid (PEBE-IL) and quaternization agent (PEBE-Q) were lower than that of PEBE. While PEBE had an Rs value of 9 � 109 Ω/sq,

PEBE-IL and PEBE-Q exhibited the surface resistivities of 4 � 108 Ω/sq

and 1 � 109 Ω/sq, respectively.

Fig. 11 demonstrates the effect of different concentrations of PEBE, PEBE-IL and PEBE-Q copolymers on the surface resistivity of PET com-posite film at 20 �C and 65% RH. It was observed from Fig. 11 that the

surface resistivity of PET composite films significantly dropped depending on the amount of antistatic agents (PEBE and its derivatives). This lowering in surface resistivity was more obvious for PET/(PEBE-Q) and PET/(PEBE-IL) and it showed that the quaternization and ionic liquid doping had a positive effect on the increment of the surface conductivity of the PET composite films.

As one can see from Fig. 11, the surface resistance of the composite films dramatically decreased with the addition of copolymers (PEBE, PEBE-IL and, PEBE-Q) to the neat PET until the percolation threshold. Beyond this point, the decrease in surface resistivity was very slight and Table 1

DSC results for copolymers.

First Heating Cooling Second Heating Second Heating after 15 min of cooling at 70 �C Tg (�C) Tm (�C) ΔHm(j/gr) Tc (�C) Tg (C) Tcc (C) ΔHcc(j/gr) Tm (C) ΔHm(j/gr) Tg (�C) Tcc (�C) ΔHcc(j/gr) Tm (�C) ΔHm(j/gr) PEBE 25.4 47.8 66.6 35.4 40.9 5.2 51.5 40.5 52.2 39.5 7.0 50.9 43.5 53.1 PEBE-IL 36.5 44.7 47.1 27.4 37.6 – – – – 38.1 18.2 30.5 38.7 22.1 PEBE-Q 30.1 37.3 49.3 37.9 42.6 12.7 53.8 42.5 54.1 42.8 16.2 53.9 43.7 53.7 Standard deviation of the DSC results: within �0.2 �C for all the samples.

Fig. 7. Thermal stability of PEBE based copolymers and commercial neat PET.

(8)

resistivity values stayed nearly constant with further addition of PEBE and its derivatives. The surface resistivity of the PET film without any additive was found to be 7 � 1015 Ω/sq while antistatic additive

including PET films exhibited the resistivity of 1011 Ω/sq in the case of

1 wt% loading into the matrix. Furthermore, the surface resistivity of PET composite films containing 15 wt% of antistatic agent lowered to 109 Ω/sq. It meant that the addition of antistatic agents into the PET led

to lower surface resistivity, in which turn caused a higher surface con-ductivity around 104-106 times compared to neat PET film. Another

important result can be drawn from the figure that pure ionic liquid insertion to the PET matrix caused a better impact on lowering surface resistivity of the corresponding film. It can be clearly shown that 4 wt% of IL addition resulted in 1.1 � 1010 Ω/sq of surface resistivity while the

other antistatic agents yielded the surface resistivity of 7.4 � 1010 and 1.1 � 1011 Ω/sq at the same concentration. Probably, this higher

con-ductivity can correspond to the use of pure ionic liquid which can facilitate the mobility of the electric charges along the polymer matrix due to the existence of more amorphous sites. On the other hand,

percolation threshold points of the PET composite films were deter-mined from Fig. 11 and it was emphasized that the percolation threshold in respect of surface resistivity was about 4 wt% for ionic liquid and in the range of 6–10 wt% for other PEBEs.

Among all antistatic additives prepared in this study, PEBE had a lower impact on the improving surface conductivity of PET composite films compared to other antistatic agents. Nevertheless, 1 wt% of PEBE addition to the PET matrix caused roughly 104 times decay in surface

resistivity compared to the neat PET film. This can be interpreted as PEG, other alcohols and aliphatic acid segments from the PEBE chain increased the content of ether bonds and the irregularity of molecular chain arrangement thus resulted in more amorphous and hydrophilic regions in the film. Therefore, the mobility of the electrical charges in these amorphous structures of PEBE containing PET film was much easier. Comparing copolyester based antistatic additives, it can be briefly said that PEBE-IL containing film had lower surface resistivity than those of PEBE and PEBE-Q based corresponding composite films. For example; the surface resistivity of 8 wt% of PEBE-IL consisting PET Fig. 9. DSC thermograms of PET composite films with PEBE and its derivatives: (a) first heating from 70 C to 300 C, (b) cooling from 300 C to 70 C and (c)

second heating from 70 �C to 300 C (all cycles were run at a rate of 10 C/min).

Table 2

DSC values of the antistatic PET films.a

First Heating ( 70 �C to 300 C) Cooling (300 C

to 70 �C) Second Heating ( 70 �C to 300 C) Tg (�C) Tcc (C) ΔHcc(j/gr) Tm (C) ΔHm(j/gr) Crystalinity (%) Tc (C) Tg (C) Tm (C) Neat PET 75.1 137.6 23.0 252.3 41.4 13.1 181.1 85,1 249.3 PET/PEBE 73.9 121.0 31.5 257.3 41.6 7.2 195.6 76,1 249.1 PET/IL 73.0 123.2 34.2 254.4 38.3 3.0 203.6 73.0 249.2 PET/(PEBE-Q) 74.8 121.3 29.4 253.9 41.4 8.6 207.3 81.0 248.7 PET/(PEBE-IL) 73.4 122.2 34.1 253.1 40.1 4.3 187.9 75.0 249.4

aAll films except neat PET and PET/IL, contain 10 wt% of additive. PET/IL contains 5 wt% IL. The standard deviation of the DSC results: within �0.2 C for all the

(9)

film was found to be 1.1 � 1010 Ω/sq while the film comprising the same

amount of PEBE showed the surface resistivity of 2 � 1010 Ω/sq. To fully

understand the difference, the surface resistivity of two films containing 1 wt% of the ionic liquid in the final composition was compared. 1 wt% IL and 10 wt% PEBE-IL containing PET composite films showed the surface resistivity of 2 � 1011 Ω/sq and 9,7 � 109 Ω/sq respectively.

Thus 10 wt% PEBE-IL PET composite films had 20 times lower surface resistivity than that of corresponding on the basis of 1 wt% IL presence in the film composition. This result illustrated that the presence of PEBE and ionic liquid had a synergistic effect on the antistatic properties of PET composite films.

Unlike classical antistatic additives, the synthesized inherently dissipative copolymers are not migratory. The homogenous distribution of conductive copolymers into an insulating matrix (PET) to form conductive paths is the main reason for the formation of the antistatic effect. Although the dissipation mechanism on the polyether-based antistatic agents without any salt has not been clearly understood, the proposed conductivity mechanism of the prepared composite films is based on the ionic conductivity of polyoxyethylene repeated units with

the formation of conductive three-dimensional networks [40,41]. These ion conductive pathways are mainly responsible for electrostatic dissi-pative behavior while maintaining the immediate, permanent and low humidity dependent antistatic effect. Unlike classical polyether block amide copolymers, synthesized polyether block ester antistatic copol-ymer contains sulfoisophthalic acid segments that could ease and sup-port the ion transsup-port via SO3 sites on the polymer backbone. In

addition, the hygroscopic and hydrophilic character of the antistatic additives facilitates the moisture adsorption from the air and supports the antistatic behavior of the composite films [42].

In addition to polar ethylene oxide sites, the quaternization of PEBE leads to improve the ionic conductivity of copolymer by dissociation of ionic groups. As far as we know from the literature, ionic conduction can occur in two different mechanisms, namely hopping and vehicle for ionic group-containing polymers. In PEBE-Q, the ionic conduction can take place by hopping mechanism due to the absence of any vehicle (such as water, etc.) to support ion transfer [43,44].

Besides the self-assembly static dissipation of the block copolymers, the conductivity can be enhanced with the addition of ionic moieties like ionic liquids and other salts. In our case, the doping of the commercially available ionic liquid into the PEBE copolymers resulted in a significant increase in the conductivity. Even cation-anion based ionic conductivity is not fully understood yet, it is believed that the mechanism is based on the incorporation of the cations with the polar groups of the host polymer [45]. This type of ionic conductivity was also assisted with the amorphous segments, low glass transition temperature and low crys-tallinity of the polymer backbone [46]. As one can see from the thermal analysis results from the previous chapters, PEBE based copolymers and films produced thereof have lower crystallinity and the glass transition points of all PEBE derivatives are below 0 �C.

From a commercial point of view, there are many inherently dissi-pative polymers (IDP) that were available in the market. But few of them were reported in the PET film application. For example, polyether am-ides (PEBA), polyether ester amam-ides (PEEA) and polyether urethanes were sold for different applications such as antistatic injection molding and packaging. For polyether block amide in PET, it was reported that it is possible to achieve 1011–1012 Ω/sq surface resistivity with 10–20 wt% standard grade (without any salt) PEBA loading [47]. In another work, the ternary blends of LDPE/PEBA/PET with 25 wt% PEBA showed sur-face resistivity values between 1011-1012 Ω/sq [41]. For PBT and PETG

polymer matrices, it was reported that required polyether ester amide dosage to achieve 1 � 1011 Ω/sq surface resistivity was 15 wt% [48]. In

another work, with the addition of 12 wt% polyether ester amide into PTT based plaques resulted in surface resistivity of 1.7 � 1012 Ω/sq [49].

In this work, our standard grade inherently dissipative copolyester without ionic liquid and quaternization (PEBE) showed a surface re-sistivity of 1.6 � 1010 Ω/sq with 10 wt% loading. For example, the sur-face resistivity values of PET composite films containing 15 wt% PEBE-IL and PEBE-Q were measured as 6.9 � 109 Ω/sq and

9.9 � 109 Ω/sq respectively which was very sufficient for an antistatic flexible packaging application and it was also competitive with current IDP grades in terms of obtained surface resistivity.

3.3.2. Hydrophilicity

In order to investigate the hydrophilicity of the prepared PET com-posite films consisting of different antistatic additives, both contact angle and moisture uptake of the PET films were determined as a function of antistatic agent type and content. The obtained results are given in Figs. 12 and 13 respectively. Fig. 12 displayed that the neat PET film had a contact angle (theta) of 76�while this value descanted down

to 63�with the increasing PEBE and IL amount. A sharp decrease in the

contact angle of the PET composite films was observed in the case of using pure IL as an antistatic agent and theta value reached 63.7�if 6 wt

% of IL was introduced to the PET matrix. Furthermore, it was pointed out that the contact angle of PEBE-IL including PET films dropped to 62.1� with 15 wt% PEBE-IL addition. Generally speaking, the contact

Fig. 10. TGA curves PET composite films with PEBE and its derivatives.

Fig. 11. Surface resistivity changes of PET composite films as a function of the

(10)

angle of all prepared PET-based composite films was around at 73.6�and

62.1�. This result can be explained by the superior hydrophilicity of

synthesized copolyesters and the ionic liquid owing to their polarity compared to neat PET.

Besides, moisture uptakes of the antistatic additive containing PET films were measured at 20 �C and 65% RH for one week and the obtained

results are given in Fig. 13. The neat PET film absorbed 0.42% moisture after seven days of conditioning. Concerning the moisture uptake (%) of PET composite films with antistatic additives, IL and PEBE-IL containing films displayed higher moisture uptake values than those of PEBE and PEBE-Q based composite films. Moisture uptake values of all films obviously increased until 8 wt% of PEBE, PEBE-IL and PEBE-Q addition into PET matrix and reached to 1.9% for PEBE-IL containing films. Beyond 10 wt% of antistatic agent addition, moisture uptake values remained nearly unchanged regardless of the used antistatic agent type. These results also showed a similar trend with the contact angle behavior of all PET composite films regardless of the antistatic agent.

Thus, IL containing PET composite films showed the most hygroscopic character due to lower contact angle value while this film exhibited higher moisture uptake value at the same time.

In conclusion, these results demonstrated that the addition of syn-thesized PEBEs into the PET matrix increased the hydrophilicity, wettability, and conductivity of the composite films. The hydrophilicity of PEBE derivatives and the ionic liquid containing composite films are superior to neat PET owing to the incorporation of water with their polar moieties [50,51].

3.3.3. Surface morphology

To investigate the morphology of the antistatic PET composite films, SEM analysis of the selected films were performed and the photos are presented in Fig. 14. It can be observed from Fig. 14 that surface views of all films were homogeneous, dense and defect-free. Apart from neat PET, the IL and PEBE-IL containing PET composite films, some slightly straight lines were observed for the composite films with PEBE and PEBE-Q. Probably, this can be explained by the slight incompatibility between neat PET and these copolyester structures. Whereas, it was seen that the addition of IL and PEBE-IL resulted in more smooth surfaces of antistatic PET films. Therefore, it may be concluded that IL acts as a compatibility enhancer between neat PET and copolyester. This finding can also be supported by the optical properties of the antistatic films that will be explained in the following part.

3.3.4. Optical and mechanical characteristics

From the industrial point of view, adequate surface resistance, lower haziness, and sufficient mechanical properties are the major re-quirements for an antistatic packaging film. Blending the inherently dissipative copolymers into a transparent polymer matrix creates hazi-ness to the final film. For this purpose, the hazihazi-ness and the total light transmittance of the films were measured and the obtained results are given in Figs. 15 and 16. As one can see from Fig. 15a that the haziness of the PET composite films was almost linearly increased with the addition of antistatic agent and no significant transparency change was observed until 5 wt% of antistatic agent loading. Beyond this point, the haziness values of the PET composite films were gradually rose depending on the type of antistatic agent. Among PET composite films, PEBE-Q containing PET film had higher haziness values than those of PEBE and PEBE-IL containing films. For example; it was found that the haziness value of 15 wt% PEBE comprising PET film was found to be 10.5% while the film including the same amount of PEBE-Q exhibited the haziness of 11.1%. This result can be explained by the more crystalline structure of PEBE-Q compared to PEBE. Unlike PEBEs, the use of only ionic liquid had no obvious negative effect on the optical properties of the PET composite film. Haziness values of the ionic liquid doped films (PET/IL) slightly increased up to 2.3% with the increasing amount of ionic liquid. This result may be explained by its relatively amorphous structure. This finding was also confirmed by DSC and XRD analysis given in the former sections.

In addition to haziness values, total light transmission % of the PET composite films is given Fig. 15b. Similarly, it was found that the transmission values of the composite films linearly dropped with the antistatic agent addition regardless of its type and reached to 75–78% for composite films with 15 wt% of antistatic agent. The explanations given as above are also valid for the relationship between total trans-mission and additive agent’s type and content.

In addition, the appearances of the selected composite films are presented in Fig. 16. It was seen from the figure that there is no obvious change in transparency of PET and PET/IL composite films (Fig. 16b and c) however, a translucent appearance was observed with the addition of PEBE and its derivatives as an antistatic agent to the PET film structure. This can probably be explained by the incompatibility between long ether chains of PEG and the hard segments of commercial PET which mainly caused by the incompatibility of the polyether segments of the copolymer structure.

Fig. 12. Contact angle values of the PET composite films as a function of the

antistatic agent type and content (wt%).

Fig. 13. Moisture uptake of the PET composite films as a function of the

(11)

Fig. 14. SEM photos of (a) neat PET, (b) PET with 10 wt% PEBE, (c) PET with 10 wt% PEBE-Q, (d) PET with 10 wt% PEBE-IL, (e) PET with 5 wt% ionic liquid.

Fig. 15. (a) Haziness and (b) total transmission values of the PET composite films as a function of the antistatic agent type and content (wt%).

Fig. 16. Photos of the films with a black background (computer keyboard) (a) no film, (b) PET reference, (c) PET with 5 wt% IL, (d)PET with 10 wt% PEBE, (e) PET

(12)

Another important issue in the PET sheet manufacturing process is the recycling of post-consumer products. In this recycling process, the transparency of the recycled product is very critical and has to be maintained. For this reason, the PET film producers always tend to use additives with lower haziness impact to the final article.

In this study, we prefer a polyether-ester structure rather than a polyether-amide structure because the ester bonds are more compatible with the PET matrix rather than the amide structure. This makes pre-pared antistatic copolyester more similar, compatible and easier to recycle with the virgin PET polymer. Despite these low haziness values provided with PEBE and its derivatives, there is still a need to develop inherently dissipative copolymers that have a much lower impact on the visual properties of the polymer matrix to which they are incorporated. Breaking force and elongation at break of the PET composite films including copolyester ether (PEBE) and its derivatives as an antistatic agent were measured to understand the effect of these additives on the mechanical properties. Fig. 17 shows the breaking force and elongation at break of PET films in terms of the type and content of the antistatic agent. In general, increasing the contents of additives in the PET film resulted in an improvement in elongation at break and a substantial reduction of breaking force. Among PET composite films, PEBE-Q con-taining PET film had higher breaking force than those of other PEBE and derivatives containing PET films whereas the film with PEBE-IL and only ionic liquid showed the highest elongation at break. The elongation at break of the neat PET film was measured 95.5% while 10 wt% PEBE-IL had a value of 144%. In the case of using only ionic liquid to improve surface resistivity of PET films, it was found that the elongation at break became larger and larger with the amount of ionic liquid added into the polymer matrix and it reached 183% with 6 wt% of ionic liquid loading. This result showed that the introduction of ionic liquid to the polymer matrix resulted in a more amorphous and flexible structure. As we know from the literature, the ionic liquids as conductive filler in polymer composite systems generally act as plasticizers and the flexibility of polymer backbone improves with the ionic liquid content [52]. There-fore, both literature and our findings confirmed the effect of ionic liquid in the physical and mechanical properties of PET composite films. On the other hand, elongation at break of 10 wt% PEBE containing PET film was around 139% while the film including the same amount of PEBE-Q showed the elongation at break of 132%. This indicated that the qua-ternization of PEBE copolyester led to a slight increase in the rigidity of PET composite film. Similarly, our findings on the crystallinity degree of PEBE and PEBE derivatives were a very good agreement with these obtained results.

Besides, as one can see from Fig. 17b that the breaking forces of PET composite films with 10 wt% PEBE, PEBE-Q, and PEBE-IL antistatic additives were found to be 63, 66 and 59 N/15 mm respectively. This result can be explained by the more crystalline structure of PEBE-Q and

the relatively amorphous structure of PEBE-IL compared to PEBE copolymer. By the way, if we compare the breaking force of PET com-posite films with neat PET, we can conclude that the pristine PET film showed significantly higher breaking force (110 N/15 mm) compared to antistatic agent containing PET composite films. Most probably the reason is the larger crystallinity of the neat PET film than those of other composite films shown in Table 2. In contrast to the highest elongation at break value of PET/IL based composite film, the breaking force of that film showed the lowest value. This is an expectedly result considering the plasticizer effect of the ionic liquid.

4. Conclusion

In this study, three novel antistatic polyether block esters (PEBE) were successfully synthesized and characterized. Two different ap-proaches were applied to improve conductivity. Firstly, the quaterni-zation of PEBE with 2,3-epoxypropyl trimethyl ammonium chloride (PEBE-Q) and secondly doping the copolymer matrix with ionic liquid, namely quaternary alkylammonium sulfonamide type (PEBE-IL), were carried out. From FTIR measurements of the PEBE copolymers, it was observed that the transesterification, quaternization and ionic liquid doping of the copolyesters were successful. It was found that the qua-ternization of PEBE improved the crystallinity with a decrease in ther-mal decomposition temperature. Whereas it was detected that the presence of ionic liquid and other PEBE derivatives decreased the glass transition temperature and the crystallization temperature as well as the degree of crystallinity compared to neat PET film. XRF measurements of the PEBE copolymers confirmed the presence of the fluorine in the ionic liquid structure. Concerning the crystallinity of PEBE and its derivatives, XRD results indicated that the PEBE-IL had a more amorphous structure while the PEBE-Q relatively larger crystalline regions.

It was illustrated that the antistatic PET films can be achieved by choosing appropriate antistatic agent type and content. Also, it was pointed out that the controlling of the crystallization degree was a major factor to adjust electrical resistivity as well as optical properties. The antistatic performance of the ionic liquid alone is better than PEBE compositions at the same concentration but blending ionic liquid with a copolymer (PEBE-IL) had a synergist effect and improved the antistatic performance of the ionic liquid and copolyester.

SEM investigation indicates that antistatic agents showed good compatibility with the PET matrix without the addition of compatibil-izer. Among PET composite films, PEBE-Q containing PET film had higher breaking force than those of other PEBE derivatives containing PET films whereas the film with PEBE-IL and only ionic liquid showed the highest elongation at break. In conclusion, the addition of synthe-sized PEBEs into the PET matrix increased the hydrophilicity, wetta-bility, and conductivity of the composite films. The hydrophilicity of

(13)

PEBE and the ionic liquid containing composite films are superior to neat PET owing to the incorporation of water with the polar moieties of the antistatic agent [50,51]. From these observations, we conclude that PET composite films with fine-tune physical, mechanical and morpho-logical properties can be developed depending on the application area where it may need some special features.

Acknowledgement

The authors gratefully acknowledge the Setas¸ Kimya Sanayi A.S¸. and Istanbul University-Cerrahpas¸a, Chemical Engineering Department for providing technical support in characterization and extrusion trials.

References

[1] J. Pionteck, G. Wypych, Handbook of Antistatics, second ed., ChemTec Publishing, Toronto, 2016, pp. 11–13. https://www.sciencedirect.com/book/97818951 98959/handbook-of-antistatics.

[2] X. Li, R. Liu, L. Zhong, L. Gu, Antielectrostatic poly(ether ester) block copolymer of poly(ethylene terephthalate-co-isophthalate)–poly(ethylene glycol), J. Appl. Polym. Sci. 89 (2003) 1696–1701, https://doi.org/10.1002/app.12311. [3] S. Rwei, W. Lin, Synthesis and characterization of adipic acid/polyethylene glycol/

poly(ethylene terephthalate) copolyester fiber, Text. Res. J. 85 (2015) 1691–1703,

https://doi.org/10.1177/0040517515573406.

[4] S. E. George, D. S. McWilliams, R. S. Armentrout, U.S. Patent No: 20030124367 A1, (2003).

[5] G. Reese, Polyester fibers: fiber formation and end-use applications, in: J. Scheırs, T.E. Long (Eds.), Modern Polyesters: Chemistry and Technology of Polyesters and Copolyesters, John Wiley & Sons Ltd., England, 2003, pp. 426–427, https://doi. org/10.1002/0470090685.

[6] M. Armand, F. Endres, D. mac Farlane, H. Ohno, B. Scrosati, Ionic-liquid materials for the electrochemical challenges of the future, Nat. Mater. 8 (2009) 621–629,

https://doi.org/10.1038/nmat2448.

[7] T. Iwata, A. Tsurumaki, S. Tajima, H. Ohno, Bis(trifluoromethanesulfonyl)imide- type ionic liquids as excellent antistatic agents for polyurethanes, Macromol. Mater. Eng. 299 (2014) 794–798, https://doi.org/10.1002/mame.201300333. [8] A. Tsurumaki, S. Tajima, T. Iwata, B. Scrosati, H. Ohno, Antistatic effects of ionic

liquids for polyether-based polyurethanes, Electrochim. Acta 175 (2015) 13–17,

https://doi.org/10.1016/j.electacta.2014.12.128.

[9] B. Berns, H. Delig€oz, B. Tieke, F. Kremer, Conductive composites of polyurethane resins and ionic liquids, Macromol. Mater. Eng. 293 (2008) 409–418, https://doi. org/10.1002/mame.200700405.

[10] C. Xing, X. Zheng, L. Xu, J. Jia, J. Ren, Y. Li, Toward an optically transparent, antielectrostatic, and robust polymer composite: morphology and properties of polycarbonate/ionic liquid composites, Ind. Eng. Chem. Res. 53 (2014) 4304–4311, https://doi.org/10.1021/ie404096b.

[11] D. Li, Z. Pang, Q. Wang, H. Ke, Y. Cai, Y. Xu, F. Huang, Q. Wei, Fabrication and characterization of polyamide6-room temperature ionic liquid (PA6-RTIL) composite nanofibers by electrospinning, Fibers Polym. 14 (2013) 1614–1619,

https://doi.org/10.1007/s12221-013-1614-1.

[12] Y. Zhang, J. Xiang, Z. Hu, J. Lei, C. Zhou, Preparation and characterization of low environmental humidity sensitive ionic liquid polymer electrolytes, Appl. Polym. Sci. 131 (2014) 40675, https://doi.org/10.1002/app.40675.

[13] S. Ramesha, C. Liewb, E. Morris, R. Durairaj, Effect of PVC on ionic conductivity, crystallographic structural, morphological and thermal characterizations in PMMA–PVC blend-based polymer electrolytes, Thermochim. Acta 511 (2010) 140–146, https://doi.org/10.1016/j.tca.2010.08.005.

[14] Y. Seki, Enhancement of electrical conductivity of polyethylene terephthalate (pet) fabrics via ionic liquids, Polym. Plast. Technol. Eng. 8 (2018) 70–76, https://doi. org/10.1080/03602559.2018.1466163.

[15] M. D€obbelin, R. Marcilla, M. Salsamendi, C. Pozo-Gonzalo, P. Carrasco, J. Pomposo, D. Mecerreyes, Influence of ionic liquids on the electrical conductivity and morphology of PEDOT:PSS films, Chem. Mater. 19 (2007) 2147–2149, https:// doi.org/10.1021/cm070398z.

[16] N. Guo, D. Hu, H. Wang, R. Wang, Y. Xiong, Functional poly(ethylene terephthalate) materials prepared by condensation copolymerization with ionic liquids, Polym. Bull. 70 (2013) 3031–3040, https://doi.org/10.1007/s00289-013- 1005-1.

[17] A.M. Ramadan, S. Mosleh, S.M. Gawish, Grafting and quaternization of 2- (dimethylamino)ethyl methacrylate onto polyamide-6 fabric pretreated with acetone, J. Appl. Polym. Sci. 81 (2001) 2318–2323, https://doi.org/10.1002/ app.1672.

[18] L. Qirong, Preparation of cationic water-absorbent resin from PEI and PEO, J. Appl. Polym. Sci. 45 (2009) 1611–1616, https://doi.org/10.1002/app.1992.070450911. [19] M. Zaman, H. Liu, H. Xiao, F. Chibante, Y. Ni, Hydrophilic modification of

polyester fabric by applying nanocrystalline cellulose containing surface finish, Carbohydr. Polym. 91 (2013) 560–567, https://doi.org/10.1016/j.

carbpol.2012.08.070.

[20] A. Mehta, U. Gaur, B. Wunderlich, Equilibrium melting parameters of poly (ethylene terephthalate), J. Polym. Sci. Polym. Phys. Ed 16 (1978) 16–289,

https://doi.org/10.1002/pol.1978.180160209.

[21] J. Ji, B. Li, W. Zhong, Effects of soy protein on the crystallization and dielectric properties of PEG/PEG copolymers, Macromol. Chem. Phys. 213 (2012) 757–765,

https://doi.org/10.1002/macp.201100684.

[22] B.J. Holland, J.N. Hay, The thermal degradation of PET and analogous polyesters measured by thermal analysis–Fourier transform infrared spectroscopy, Polymer 43 (2002) 1835–1847, https://doi.org/10.1016/S0032-3861(01)00775-3. [23] C. Cazan, M. Cosnita, A. Duta, Effect of PET functionalization in composites of

rubber–PET–HDPE type, Arab. J. Chem. 10 (2017) 300–312, https://doi.org/ 10.1016/j.arabjc.2015.10.005.

[24] Z.S. Pour, P. Makvandi, M. Ghaemy, Performance properties and antibacterial activity of crosslinked films of quaternary ammonium modified starch and poly (vinyl alcohol), Int. J. Biol. Macromol. 80 (2015) 596–604, https://doi.org/ 10.1016/j.ijbiomac.2015.07.008.

[25] F. S¸en, M.V. Kahraman, Preparation and characterization of hybrid cationic hydroxyethyl cellulose/sodium alginate polyelectrolyte antimicrobial films, Polym. Adv. Technol. 29 (2018) 1895–1901, https://doi.org/10.1002/pat.4298. [26] M. Gonz�alez, J.C. Cabanelas, J. Baselga, Applications of FTIR on epoxy resins -

identification, monitoring the curing process, phase separation and water uptake, in: T. Theophanides (Ed.), Infrared Spectroscopy - Materials Science, Engineering and Technology, Intech Publishing, Croatia, 2012, pp. 261–281, https://doi.org/ 10.5772/36323.

[27] M.E.D. Lestard, L.A. Ramos, M.E. Tuttolomondo, S.E. Ulic, A.B. Altabef, Bis (trifluoromethyl) sulfone, CF3SO2CF3: synthesis, vibrational and conformational properties, Spectrochim. Acta 96 (2012) 332–339, https://doi.org/10.1016/j. saa.2012.05.049.

[28] G. Alva, Y. Lin, G. Fang, Synthesis and characterization of chain-extended and branched polyurethane copolymers as form stable phase change materials for solar thermal conversion storage, Sol. Energy Mater. Sol. Cells 186 (2018) 14–28,

https://doi.org/10.1016/j.solmat.2018.06.023.

[29] X. Zheng, Q. Lin, P. Jiang, Y. Li, J. Li, Ionic liquids incorporating polyamide 6: miscibility and physical properties, Polymers 10 (2018) 562, https://doi.org/ 10.3390/polym10050562.

[30] P. Xua, X. Luoa, Y. Zhoua, Y. Yang, Y. Ding, Enhanced cold crystallization and dielectric polarization of PLA composites induced by P[MPEGMA-IL] and graphene, Thermochim. Acta 657 (2017) 156–162, https://doi.org/10.1016/j. tca.2017.10.005.

[31] T. Wei, S. Pang, N. Xu, L. Pan, Z. Zhang, R. Xu, N. Ma, Q. Lin, Crystallization behavior and isothermal crystallization kinetics of PLLA blended with ionic liquid, 1-butyl-3-methylimidazolium dibutylphosphate, Appl. Polym. Sci. 132 (2015) 41308, https://doi.org/10.1002/app.41308.

[32] H. Yoon, H. Zhu, A. Hervault, M. Armand, D.R. MF, M. Forsyth, Physicochemical properties of N-propyl-N-methylpyrrolidinium bis(fluorosulfonyl)imide for sodium metal battery applications, Phys. Chem. Chem. Phys. 16 (2014) 12350–12355,

https://doi.org/10.1039/c4cp00365a.

[33] C.R. Pope, M. Kar, D.R. MacFarlane, M. Armand, M. Forsyth, L.A. O’Dell, Ion dynamics in a mixed-cation alkoxy-ammonium ionic liquid electrolyte for sodium device applications, ChemPhysChem 17 (2016) 3187–3195, https://doi.org/ 10.1002/cphc.201600692.

[34] Y. Xiong, Q. Liu, Q. Zhang, A.M. Zhu, Synthesis and characterization of cross- linked quaternized poly(vinyl alcohol)/chitosan composite anion exchange membranes for fuel cells, J. Power Sources 183 (2008) 447–453, https://doi.org/ 10.1016/j.jpowsour.2008.06.004.

[35] A.M. Sajjan, H.G. Premakshi, M.Y. Kariduraganavar, Synthesis and

characterization of GTMAC grafted chitosan membranes for the dehydration of low water content isopropanol by pervaporation, J. Ind. Eng. Chem. 25 (2015) 151–161, https://doi.org/10.1016/j.jiec.2014.10.027.

[36] 3M™ Ionic Liquid Antistat FC-5000, Technical Data Sheet, 2014 (Technical Data Sheet).

[37] Z. Zhu, M.J. Kelley, IR spectroscopic investigation of the effect of deep UV irradiation on PET films, Polymer 46 (2005) 8883–8891, https://doi.org/10.1016/ j.polymer.2005.05.135.

[38] J. Tao, X. Dong, T. Wang, Y. Cui, Crystallization behaviour of nano-ZnO/PET and nano-ZnO/polyethylene glycol (PEG)/PET composites by in situ polymerization, Key Eng. Mater. 334–335 (2007) 853–856. https://doi.org/10.4028/www. scientific.net/KEM.334-335.853.

[39] A. Tsurumaki, S. Tajima, T. Iwata, B. Scrosati, H. Ohno, Evaluation of ionic liquids as novel antistatic agents for polymethacrylates, Electrochim. Acta 248 (2017) 556–561, https://doi.org/10.1016/j.electacta.2017.07.181.

[40] R.B. Rosner, Conductive materials for ESD applications: an overview, IEEE Trans. Device Mater. Reliab. 1 (2001) 9–16, https://doi.org/10.1109/7298.946455. [41] J. Wang, A.R. Valencia, B.D. Favis, Assembling conductive PEBA copolymer at the

continuous interface in ternary polymer systems: morphology and resistivity, Macromolecules 49 (2016) 5115–5125, https://doi.org/10.1021/acs. macromol.6b00469.

[42] M. Young, J. Lin, Electrostatic dissipating properties of poly(oxyethylene)amine- modified polyamides, Ind. Eng. Chem. Res. 37 (1998) 4284–4289, https://doi.org/ 10.1021/ie980197s.

[43] J. Lina, S.R. Kumara, W. Maa, C. Shiha, L. Tenga, C. Yang, S.J. Lue, Gradiently distributed iron oxide@graphene oxide nanofillers in quaternized polyvinyl alcohol composite to enhance alkaline fuel cell power density, J. Membr. Sci. 543 (2017) 28–39, https://doi.org/10.1016/j.memsci.2017.08.045.

[44] W. Chen, X. Yan, X. Wu, S. Huang, Y. Luo, X. Gong, G. He, Tri-quaternized poly (ether sulfone) anion exchange membranes with improved hydroxide conductivity, J. Membr. Sci. 514 (2016) 613–621, https://doi.org/10.1016/j.

(14)

[45] S.B. Aziz, T.J. Woo, M.F.Z. Kadir, H.M. Ahmed, A conceptual review on polymer electrolytes and ion transport models, J. Sci. Adv. Mat. Dev. 3 (2018) 1–17,

https://doi.org/10.1016/j.jsamd.2018.01.002.

[46] T.J. Singh, S.V. Bhat, Morphology and conductivity studies of a new solid polymer electrolyte:(PEG)xLiClO4, Bull. Mater. Sci. 26 (2003) 707–714, https://doi.org/ 10.1007/BF02706768.

[47] Pebax® high performance permanent antistatic agent, Presentation, available at:

http://www.connellbros.com.cn/upload/funplastic/product/20140331033640. pdf. Accessed 2019 Oct 25.

[48] Modifying conductivity in plastics:meeting needs in electrical and electronics, Plastics, Addit. Compd. 6 (2004) 40–43, https://doi.org/10.1016/S1464-391X(04) 00104-7.

[49] T. Kobayashi, B.A. Wood, G.S. Blackman, A. Takemura, Morphology, dynamic mechanical, and electrical properties of bio-based poly(trimethylene terephthalate) blends, part 1: poly(trimethylene terephthalate)/poly(ether esteramide)/

polyethylene glycol 400 bis(2-ethylhexanoate) blends, Appl. Polym. Sci 120 (2011) 3519–3529, https://doi.org/10.1002/app.33498.

[50] H. Ajiro, Y. Takahashi, M. Akashi, T. Fujiwara, Surface control of hydrophilicity and degradability with block copolymers composed of lactide and cyclic carbonate bearing methoxyethoxyl groups, Polymer 55 (2014) 3591–3598, https://doi.org/ 10.1016/j.polymer.2014.06.011.

[51] P. Ji, J. Jin, G. Ji, C. Wang, H. Wang, Investigation of the state and distribution of water in poly(ethylene terephthalate)/polyethylene glycol copolymers with various molecular weight of polyethylene glycol, Polym. Eng. Sci. 55 (2015) 2195–2204, https://doi.org/10.1002/pen.24104.

[52] H. Delig€oz, M. Yılmazo�glu, S. Yılmaztürk, Y. S¸ahin, K. Ulutas¸, Synthesis and characterization of anhydrous conducting polyimide/ionic liquid complex membranes via a new route for high-temperature fuel cells, Polym. Adv. Met. Technol. 23 (2011) 1156–1165, https://doi.org/10.1002/pat.2016.

Referanslar

Benzer Belgeler

Yeniköy'de A fif Paşa yalısı, Faik ve Bekir Bey yalıları, Ar­ navutköy yalıları, Tarabya'da eski Fransız Okulu, İtalyan Konsolosluğu, Alman Sefare­ ti, Villa

Nükhet Demirtaşlı (Prof. Dr.) Ülkü Doğanay (Prof. Dr.) Ali Dönmez (Prof. Dr.) Çiler Dursun (Prof. Dr.) Pınar Ecevitoğlu (Yrd. Dr.) Selin Esen (Prof. Dr.) Yasemin Esen (Doç.

Depolama alanlarında tüm vücudun ergonomik olarak değerlendirilmesini amaçlayan bir çalışmada ise (Battini vd., 2014) raflardan malzeme alma, yerleştirme gibi

In this thesis, we study an extension of symmetric connections model, which is presented by Matthew O. Jackson and Asher Wolinsky in their paper entitled, “A Strategic

In conclusion, the results of this study indicate that surgical treatment of intra-articular calcaneus fractures allows more effective results compared to conservative

Inspired by Kaldor (1940), Ichimura (1954) explored the possibility of an economic system with a unique limit cycle; Chang and Smyth (1971) reexamined the model and stated the

The program is observed to be based upon three main components: (i) austerity of public expenditures subject to specific targets for non-interest fiscal surpluses; (ii) a

Fixing we group the scatterers into varying NoC and compute the a scatterer scenario, we calculate root mean square delay and rms angular and delay spreads of the corresponding