Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Composites Part B 218 (2021) 108930

Contents lists available at ScienceDirect

Composites Part B
journal homepage: www.elsevier.com/locate/compositesb

Interfacial localization of organoclay enhances the peelability of


polyethylene/polybutene-1/organoclay nanocomposite films
Raziyeh S. Mohammadi a, Ali M. Zolali b, 1, Seyed H. Tabatabaei c, Abdellah Ajji a, *
a
3SPack NSERC-Industry Chair, CREPEC, Department of Chemical Engineering, Polytechnique Montreal, C.P. 6079, Succ. Centre-ville, Montreal, QC, H3C 3A7,
Canada
b
Microcellular Plastics Manufacturing Laboratory, Department of Mechanical and Industrial Engineering, University of Toronto, Toronto, Ontario, M5S 3G8, Canada
c
ProAmpac, 12025 Tricon Road, Cincinnati, OH, 45209, USA

A R T I C L E I N F O A B S T R A C T

Keywords: Premature failure in polymer blends originates mainly at interfaces between immiscible polymers which is
Polymer nanocomposite deemed detrimental to mechanical properties in many applications. Here we exploit this phenomenon in
Peel fracture developing a peelable film with a very wide peelable seal-temperature window (ΔTp) of over 90 ◦ C. A melt-cast
Polybutene-1
films of low-density polyethylene (PE) containing both 5 wt% isotactic polybutene-1 (PB-1) and 1 phr organoclay
Organoclay
Polymorphism
shows peel strength well within the peelable region (150–650 N/m), whereas the PE/5 wt% PB-1 blend, and PE/
1 phr organoclay nanocomposite films are lock sealed similar to the neat PE film after the sealing process.
Furthermore, the characteristic aging of the PE/PB-1 blends is eliminated by the addition of organoclay to the
system. The results indicate a synergistic interaction between organoclay and PB-1 that plays a key role in the
formation and consistent propagation of cracks throughout the system. The localization of organoclay at the
interface of the PE/PB-1 blend significantly reduces the thickness of the fibrillar PB-1 phase more than three
times to 60 nm. This suggests an increase in the interfacial area that can facilitate interfacial cracking and failure
at the interface. The nanoconfinement induced by the PE matrix and organoclay at the interface also enhance the
formation of the thermodynamically stable form I and form III PB-1 crystals rendering a consistent peel
performance.

1. Introduction shown to be effective in toughening PP/CaCO3 nanocomposite up to 4


times more than the impact strength of the neat PP [9]. Lightweight
Most polymers are thermodynamically incompatible and phase microcellular polylactide (PLA)/polyolefin blends with significantly
separate when they are melt blended to generate materials with com­ improved mechanical properties, i.e. high strength, high modulus, and
plementary mechanical or thermal properties [1]. The interfaces of increased impact resistance as compared with the neat PLA, were re­
immiscible polymers in polymer blends and composites are regarded as ported through taking advantage of the poor interfacial compatibility
the weak mechanical points where premature failure is originated [2–4]. between PLA and polyolefins [10]. The materials were cold drawn over a
The compatibilization of polymer blends and nanocomposites to rein­ range of uniaxial strains (10–90%) at room temperature which resulted
force the polymer interfaces through different approaches such using in the formation of microvoids at the interface of PLA/polyolefin and a
copolymers or nanomaterials has drawn tremendous attention over the weight reduction of up to 34%.
past decades [5–7]. These studies focus on improving the interfacial The interfacial incompatibility along with a unique elongated phase
adhesion between components to enhance the mechanical performance, morphology of a dispersed phase in polyolefin-based binary blend films
particularly, the impact strength [7,8]. In contrast, a few studies have are being exploited in the packaging industry for producing cohesive
reported exploiting the interfacial incompatibility in polymer blends or easy-to-open packings, i.e. peelable seals. The peelable seals are
composites in favor of practical applications. The interfacial debonding increasingly employed in a vast variety of packages for merchandise,
at the interface of polypropylene (PP) and CaCO3 nanoparticles has electrical, food and medical packaging applications [11,12]. In contrast

* Corresponding author.
E-mail address: abdellah.ajji@polymtl.ca (A. Ajji).
1
Present address: Performance Materials Research, BASF Corporation, 1609 Biddle Ave., Wyandotte, MI 48192, United States.

https://doi.org/10.1016/j.compositesb.2021.108930
Received 20 August 2020; Received in revised form 8 April 2021; Accepted 18 April 2021
Available online 22 April 2021
1359-8368/© 2021 Elsevier Ltd. All rights reserved.
R.S. Mohammadi et al. Composites Part B 218 (2021) 108930

to adhesive peels that possess a relatively high risk of leakage and are Organoclays are effective in inducing peelability in the PE nano­
more sensitive to seal parameters, i.e. pressure and temperature [13], composites films through initiating cracks within the electrostatically
cohesive peels are welded to a substrate and minimize the chances of bonded clay stacks/layers when they are well dispersed and distributed
leakage [14]. Blending of polyolefins such as polyethylene (PE) and throughout the system [46]. The localization of organoclay at the
ethylene-vinyl acetate copolymer (EVA) with polybutene-1 (PB-1) is a interface of PE/ethylene-methyl acrylate copolymer (EMA) blends has
promising approach to produce cohesive peelable seals [15]. PB-1 has recently presented a promising strategy to generate systems with com­
attracted much attention due to its superior impact properties, tough­ patibilized but weak interfaces suitable for peel applications [47]. Here,
ness, elastic recovery and creep resistance [16,17]. PB-1 is immiscible we report a high performance peelable PE/PB-1/organoclay nano­
with PE and its counterpart polyolefins and possesses decent tensile and composite film with exceptional ΔTp and consistent peel strength. Only
yield strengths required for seal applications. It has been shown that at 1 wt% organoclay localized at the PE/PB-1 interface was enough to
least 10 wt% of PB-1 is required to achieve peelable seal with reasonable markedly tune the fibrillar structure of the 5 wt% dispersed PB-1 phase.
heat seal temperature window (ΔTp) [18]. ΔTp is a temperature window By studying the morphology development and thermal properties of the
in which a cohesive peelable seal is obtained and typically is very nar­ systems, we reveal the role of confinement imposed upon the PB-1 phase
row for most of conventional peel formulations. By increasing the heat on the polymorphism of PB-1 and mechanical properties of the nano­
seal temperature, the peel performance may convert to lock seal due to composite films. We also show that the typical aging problem of films
intense polymer chain inter-diffusion, disintegration of the fibrillar containing PB-1 can be successfully eliminated while retaining a
morphology, or increase in crystalline anchors at the weld zone [19,20]. consistent peel strength within the limits of easy-to-open packaging.
Depending on crystallization conditions, PB-1 exhibit different
crystalline structures including twined hexagonal form I with 3/1 helix 2. Experimental
[21], untwined hexagonal form Iʹ with 3/1 helix [22], twinned tetrag­
onal form II with 11/3 helix [23] and orthorhombic form III with 4/1 2.1. . Materials
helix conformation [24]. Crystals form Iʹ and III are usually obtained
from solution and can be converted to crystal form II when dried and Low density polyethylene (PE) with the trade name of Novapol LF-
heated up to 90 ◦ C [24]. When cooled from melt to ambient tempera­ 0219-A was obtained from Nova chemicals. Polybutene-1 (PB-1),
ture, PB-1 forms more kinetically favored crystalline form II. Then, the PB0300 M, with a density of 0.915 g cm− 1, melt flow index of 4 g/10 min
metastable crystal form II spontaneously converts to crystalline form I, (190 ◦ C/2.16 kg) and molecular weight of 374 kg/mol was kindly pro­
which is thermodynamically stable [22,25]. Because of the significantly vided by Lyondellbasell. Organomodified montmorillonite clay nano­
different melting temperatures and densities of the crystal forms II and I, particles with the trade name of Cloisite 15 was obtained from Byk
the transition of crystalline form II to form I results in profound thermal Company. The organoclay was modified with dimethyl dehydrogenated
and mechanical changes [26,27]. So far, extensive efforts have been tallow with approximately 65% C18, 30% C16, 5% C14 with cation
made to accelerate form II to form I crystal transition through various exchange capacity (CEC) of 125.
methods such as pressure [28], orientation and drawing [29–33],
incorporation of additives [34] and copolymerization [35,36]. Some 2.2. Sample preparation
studies have reported direct formation of form I or form Iʹ under peculiar
conditions such as from ultrathin films [37], through self seeding [38], PE/PB-1 blends containing 5 (B5), 10 (B10) and 20 (B20) wt% of PB-
stereodefects [39] and manipulating the melt temperature [40,41]. 1 were prepared using a Leistritz ZSE 18HP twin screw extruder (TSE)
Nevertheless, complicated polymorphism and inconsistent thermal, with an L/D ratio of 40. A masterbatch of PB-1 containing 20 wt%
physical and mechanical properties of PB-1 after melt processing re­ Cloisite 15 was prepared using the same TSE equipped with a separate
mains a challenging issue that restricts its commercial development. nanoclay feeder. The master-batch was then diluted with PE to prepare
It has been shown that small amounts of nanoparticles can effectively PE/PB-1/organoclay nanocomposites containing 5 (NB5), 10 (NB10)
change the morphology, thermal, and mechanical properties of immis­ and 20 (NB20) wt % PB-1 and 1 phr Cloisite 15. PE containing 1 phr
cible blends [42]. Nanoclay is one of widely used nanoparticles in Cloisite 15 (PE/C1) was also prepared as a reference. The screw speed of
polymer-based composites. Due to its unique two-dimensional structure, 100 rpm and a temperature profile of 150/160/170/180/180/190/190/
it significantly affects the phase morphology by refining and coarsening 190, from hopper to die, were used for all compoundings. The blends
dispersed domains, inducing/broadening cocontinuity, and compatibi­ and nanocomposites were prepared in a separate melt casting process
lizing phases. Generally, thermodynamics governs the morphology into thin films using the same TSE equipped with a slit die with a die
development in polymer blends and nanocomposites, however, kinetic opening of 500 μm. The extruded films were passed through an air knife
effects can also play a major role in the localization of nanoparticles due and drawn with calendar rolls chilled with cold water (see Fig. 1a). The
to high viscosity of polymer components. The addition of nanoclay to collection speed and draw ratio were manipulated to achieve films of 70
immiscible poly(vinylidene fluoride) (PVDF)/polyamide-6 (PA6) blends μm in thickness.
has reduced the dispersed PVDF phase size [43]. The dispersion and To examine the peeling performance of the nanocomposite films,
localization of the nanoclay were identified to control the degree of they were first welded through a sealing process. The sealing process
compatibilization induced by the nanoclay. The compatibilization is was performed using a hot tack/seal tester from LakoTool &
achieved by suppressing coalescence or/and reduction in the interfacial Manufacturing Inc, USA. Two strips of the films with 2.54 cm in width
tension when the nanoclay is placed at the interface [42]. A significant were cut to be welded on themselves. The films were sandwiched be­
reduction in the polystyrene (PS) phase size was reported in PP/PS tween two acetate films in order to prevent sticking of the films to the
blends when organically modified nanoclay was resided in the PP phase hot seal bars. A sealing pressure of 0.5 N/mm2, a dwell time of 1 s and an
and at the interface of PP/PS. The addition of organoclay improved the identical sealing temperature for upper and lower seal bars were set as
tensile strength and modulus while a decrease was noted for elongation the seal parameters (see Fig. 1b). Samples were sealed at temperatures
at break and impact strength of the nanocomposite blends at low PS ranging from 110 to 200 ◦ C with intervals of 5 ◦ C.
contents [44]. A micromechanical deformation analysis of PLA/natural T-peel tests were carried out using a tensile testing machine (Instron
rubber (NR)/organoclay nanocomposite blends has showed that orga­ E3000) per ASTM F88 to evaluate the peel performance of the sealed
noclays act as preferential sites for the craze formation at the PLA/NR films. The samples were peeled at room temperature with a peeling rate
interface [45]. It appears that at an optimum 1 wt% of organoclay of 200 mm/min (see Fig. 1c). The plateau of the force-displacement
content, organoclays both generate multiple crazes and preclude craze curves was reported as the peel force for each sample. To ensure
development and transformation into cracks during uniaxial stretching. reproducibility of the data, the average peel force of at least 5 specimens

2
R.S. Mohammadi et al. Composites Part B 218 (2021) 108930

Fig. 1. Film preparation process and T-peel test: (a) schematic of the melt casting process of plastic films. The calendaring parameters are controlled to obtain films
with an average thickness of 70 μm; (b) the sealing process at a constant temperature and pressure; (c) T-peel testing of sealed films.

of each sample was reported as the final peel force. of polymers and solid inclusions as they provide high levels of dispersion
and distribution of components by vigorous shear and extensional
2.3. Characterizations mixing. The flat cast die attached to the TSE in our work shapes the
polymer melt into a thin film with a controlled thickness. The thickness
Wide angle X-ray diffraction (WAXD) analysis was performed using a is further adjusted using a calendar by fine tuning the drums’ speed and
Phillips X’pert apparatus. The anode was copper (Cu) with a K<α cooling rate. The phase morphology of the blends and nanocomposite
wavelength of 1.54 Å and the generator voltage and tube current were blends were examined through SEM analysis. Fig. 2 and Fig. S1 show the
set at 50 kV and 40 mA, respectively, at room temperature. The scans morphology of the blends and nanocomposites, respectively. A fibrillar
were run over a 2θ range of 2–30◦ with a scan rate of 0.02◦ /s and a 20 cm phase morphology is identified in both blends and nanocomposites. The
distance between the sample and detector. In order to maximize the cross-section of the PB-1 fibers can be seen in the micrographs taken in
diffraction intensity, several layers of each sample were stacked to make the machine direction (MD) which are used to calculate the fiber size
samples with an approximately 2 mm in thickness. The d-spacing of the and interparticle distance between the fibers. The elongated features
clay layers was calculated using Bragg’s equation: λ = 2d<sinθ in which observed in the transverse direction (TD) micrographs prove that the
λ is the wavelength of the X-ray radiation, <d is the distance between the holes detected in the MD are fibers extended along the films. Due to high
clay layers and θ is the diffraction peak angle. A transmission electron extensional deformation exerted on the systems inside the flat die, the
microscopy (TEM) from JEOL JEM-2100F, Japan, operating at 200 kV dispersed PB-1 phase, which was dispersed and distributed into spher­
was used to examine the localization of organoclay in the nano­ ical domains during the mixing process in the TSE, extends into thin
composite blends. For sample preparation, an ultra-cryomicrotome, elongated PB-1 fibrils. This basically creates thin films with the PB-1
Leica Microsystem EM-UC7/FC7, operating at − 120 ◦ C and equipped fibrillar domains elongated along the films. Table 1 summarizes the
with a diamond knife was employed to cut thin 100 nm slices of nano­ main morphological characteristics of the blends and nanocomposite
composites suitable for TEM imaging. blends. The thickness of the PB-1 nanofibers increases from 100 nm to
Differential scanning calorimetry (DSC) was performed using a DSC 200 nm by the PB-1 content from 5 to 20 wt%.
instrument Q2000 from TA Instruments. About 10 mg of each sample The morphology analysis results show that the addition of organo­
was placed in an aluminum pan and then heated from 25 to 150 ◦ C to clay affect the phase morphology of PB-1. Although a similar trend to the
remove thermal history. Then, it was cooled to 0 ◦ C and finally heated blends is observed in the nanocomposite blends, a significant reduction
again to 150 ◦ C. All experiments were performed under nitrogen at­ in the PB-1 thickness to 60 nm can be seen for the nanocomposite
mosphere with a scan rate of 10 ◦ C/min. The heat of fusion (ΔHm) of the containing 5 wt% PB-1 (see Table 1). The dispersion and distribution of
samples was measured using TA Universal Analysis software to calculate organoclay as well as its localization can significantly influence the PB-1
the degree of crystallinity according to equation (1) in which Xc is de­ phase morphology [49]. A thermodynamic model based on the inter­
gree of crystallinity, x is the weight fraction of PE and ΔHm,100 is the facial energies of constituent components is widely used for predicting
enthalpy of 100% crystal of low density PE, which is reported to be 298 the localization of solid inclusions in polymer nanocomposite blends.
J/g [48]. The thermodynamic analysis shows that the organoclay would localize
/ at the PE/PB-1 interface (see Supporting Information). However, to
Xc = ΔHm xΔHm,100 (1) assess the predictions, TEM imagings were performed to provide a
qualitative understanding of the distribution and localization of the
Scanning electron microscopy (SEM) was conducted using a Field
organoclay. Fig. 3 shows the TEM micrographs of the
Emission SEM machine (JSM 7600TFE, JEOL) operated at a voltage of 2
PE/PB-1/organoclay nanocomposite films containing different PB-1
kV. Films were embedded in an epoxy mold in machine and transvers
contents in the machine direction. The darker domains marked by ar­
directions (MD and TD) and then cryo-microtomed using a glass knife at
rows are organoclay stacks which are distinct from the gray PB-1 do­
− 150 ◦ C. The cryo-microtomed samples were then treated using cyclo­
mains dispersed within the PE matrix. These results suggest that
hexane at 50 ◦ C to extract the dispersed PB-1 phase and create contrast
organoclay layers/stacks are mainly located at the interface of the PE
between phases for SEM imaging. The cryo-microtomed samples were
and PB-1 phases. This is in line with the thermodynamic predictions and
gold/palladium coated under plasma vacuum deposition before SEM
confirms that the organoclay thermodynamically tends to migrate to the
performed in MD direction.
PE/PB-1 interface despite being first incorporated into the PB-1 phase.
Almost no organoclay layers/stacks can be detected in the PE matrix
3. Results and discussion
whereas some are observed inside the PB-1 domains. This can be related
to the mixing sequence as the organoclay was first melt-blended with
Films preparation and heat sealing process. The blends and
PB-1 and then the masterbatch was diluted by PE. The kinetic barrier
nanocomposite blends were all prepared through a melt-mixing process
due to the high viscosity of the PB-1 phase can prevent the full segre­
using a twin-screw extruder as illustrated in Fig. 1a. Twin-screw ex­
gation of the organoclay at the interface, and thus some may be stuck
truders (TSEs) are considered ideal mixing equipment for melt-blending

3
R.S. Mohammadi et al. Composites Part B 218 (2021) 108930

Fig. 2. SEM images of (a, a’) B5, (b, b’) B10 and (c, c’) B20 in the machine and transverse directions, respectively (The white bars denote 1 μm).

nanocomposite blends with 10 and 20 wt% PB-1, minor phase size re­
Table 1
ductions are obtained which are within the margin of the errors (see
The morphological and thermal characteristics of the blends and nanocomposite
Table 1). By increasing the PB-1 content, the interfacial area increases
blends.
and consequently more organoclay is needed to completely cover and
Sample Fiber Aspect ratio Interfibrillar Ti ΔTp
saturate the interface. However, a constant amount of organoclay (1
thickness (L/D) distance (nm) (◦ C) (◦ C)
(nm) phr) is used in all the systems which seems not enough to cover the
interfacial area in the two systems with 10 and 20 wt% of PB-1. This
PE 110 5
explains why only a slight decrease in the PB-1 phase size is achieved at
– – –
PE/C1 – – – 110 5
B5 100 ± 20 ~40 365 ± 25 110 10 higher PB-1 contents. The TEM results support these morphological
B10 190 ± 35 >50 300 ± 17 115 20 observations. More organoclay stacks/layers are located at the interface
B20 200 ± 50 >50 290 ± 15 120 70 of PE/PB-1 in the nanocomposite blend containing 5 wt% PB-1 (see
NB5 60 ± 8 ~30 165 ±7 110 90 <
Fig. 3a). Much less organoclay stacks/layers can be seen at the interface
NB10 170 ± 15 >50 260 ± 11 115 45
NB20 190 ± 20 >50 270 ± 18 120 65–70 when the PB-1 content is increased to 10 and 20 wt% as shown in Fig. 3b
and c. Thus, much less compatibilization effect is observed in these
systems.
inside the PB-1 phase. The peel performances of the PE/PB-1 blends and PE/PB-1/clay
The morphologies observed in the nanocomposite blends are directly nanocomposites are demonstrated in Fig. 4. The nanocomposite films
affected by the localization of organoclay at the interface. The organo­ with a peel force value within the easy-to-open window (150–650 N/m)
clay layers at the interface prevent the coalescence of the PB-1 domains are considered as peelable films [51]. The seal initiation temperature
by both compatibilizing the interface and creating a physical barrier at (Ti), a minimum temperature at which two films are welded, and ΔTp are
the interface. This leads to a marked reduction in the phase size of the listed in Table 1. Ti of the neat PE film is around 110 ◦ C, very close to its
dispersed PB-1 phase in the NB5 nanocomposite blend containing 5 wt% melting temperature. By increasing the seal temperature to 115 ◦ C, the
PB-1. This is another indication that the organoclay is localized at the peel force of the PE significantly increases beyond 650 N/m showing a
interface and act as a compatibilizer to reduce the dispersed PB-1 phase lock seal behavior. Blending PE with 5 wt% of PB-1 slightly enhances its
size as reported elsewhere [43]. Whereas, the localization of organoclay peelability window ΔTp from 5 to 10 ◦ C. ΔTp is further increased to
within the dispersed phase generally increases the viscosity of the 20 ◦ C for B10 and to 70 ◦ C for B20. In addition, Ti of B10 and B20 in­
dispersed phase and results in coarsening its morphology [50]. In the crease to 115 and 120 ◦ C, respectively. The widening of Ti by increasing

Fig. 3. TEM images of (a) NB5, (b) NB10 and (c) NB20 nanocomposite films. The arrows indicate clay particles at the PE/PB-1 interface.

4
R.S. Mohammadi et al. Composites Part B 218 (2021) 108930

Fig. 4. Peel force versus seal temperature of (a) the blends and (b) the nanocomposites.

Fig. 5. Force-displacement curves of (a) B5, (b) NB5, (c) B10, (d) NB10, (e) B20 and (f) NB20 at different sealing temperatures.

5
R.S. Mohammadi et al. Composites Part B 218 (2021) 108930

the PB-1 content is attributed to the higher melting point of PB-1 crystals blends and nanocomposites, the fracture surface of the sealed films after
compared to that of PE as Ti depends directly on the crystallinity of peeling is investigated by SEM and the results are shown in Fig. 6. A
components [19,20]. Fig. 4b shows that the peelability of the blend rough and highly fibrillated peeled fracture surface indicates a high peel
systems was significantly improved upon the incorporation of 1 phr strength and resistance upon peeling. In contrast, a smooth and less
organoclay to the blends. ΔTp is markedly broadened for NB5 when it is rough peeled surface with less plastic deformations suggests lower
compared to that of NB10. Also, despite the NB5 and NB10 nano­ resistance to peeling. Fig. 6a and b shows the peeled fracture surface of
composite blends that demonstrate a significant improvement, the peel the B5 and NB5 films sealed at 130 and 200 ◦ C, respectively. The B5 film
force of the NB20 nanocomposite blend slightly increases while showing is not peelable at 130 ◦ C while the NB5 nanocomposite film is cohesively
a ΔTp similar to the neat B20 blend. peeled even when sealed at 200 ◦ C. These fracture results are in line with
To better understand these results, the force-displacement curves of the morphological observations and T-peel test results. The peeled sur­
the blend films were examined against the nanocomposite films during face of the easily peelable NB5 nanocomposite blend is uniformly frac­
the peeling experiments and the results are presented in Fig. 5. It should tured with less fibrillations whereas the B5 blend film is fractured with
be noted that the curves with a wide plateau represent a cohesive peel significant amount plastic deformations. It should be noted that the NB5
behavior. Films with a partially cohesive peel behavior show a small nanocomposite film shows similar peeled fracture surface to the one
plateau and then an increase in the force due to elongation upon further sealed at 200 ◦ C over its entire peelable region. The NB10 nano­
peeling. The curves without plateau are similar to stress-strain curves composite film is peelable up to the seal temperature of 160 ◦ C (Fig. 6c)
and represent the lock seal behavior as lock seals usually yield and and at a seal temperature beyond 160 ◦ C it is not easily peelable
deform until rupture. Fig. 5a indicates that the B5 blend displays a lock (Fig. 6d). These results also agree with the peel force experiments where
seal performance at seal temperatures beyond 120 ◦ C while the NB5 the NB10 nanocomposite film revealed a lock seal behavior at the seal
nanocomposite is cohesively peeled over all the tested temperatures temperature of 160 ◦ C. The NB20 nanocomposite film demonstrates the
from 120 to 200 ◦ C (Fig. 5a’). The B10 blend demonstrates a cohesive easy-to-open peelable character even when is sealed at 190 ◦ C (Fig. 6e).
peel performance when sealed under 130 ◦ C whereas increasing the seal A slightly higher seal temperature at 195 ◦ C changes the peeling per­
temperature to 135 ◦ C results in partial peeling followed by elongation formance to a lock-seal behavior with high plastic deformation at the
of the films (Fig. 5b). At seal temperatures above 140 ◦ C, the B10 blend peeled surface (Fig. 6f).
yields and becomes elongated before being peeled. In contrast, the NB10 The NB5 nanocomposite film containing only 5 wt% of PB-1 and 1
nanocomposite film shows a plateau force-displacement curves up to the phr of organoclay shows markedly improved peel performance
seal temperature of about 160 ◦ C and just yields beyond that seal tem­ compared to the other blend and nanocomposite films. The mechanical
perature (Fig. 5b’). The peel performance of the B20 blend and NB20 and microstructure analyses suggest that there is a synergistic effect
nanocomposite are very similar as displayed in Fig. 5c and c’ showing between organoclay and PB-1. The addition of organoclay enhances the
that the organoclay has almost no effect on the peeling of the system. yield strength of the NB5 film to a value that surpasses the interfacial
These results can be explained through understanding the mechan­ adhesion of the welded area. On the other hand, the reduction of the PB-
ical properties and microstructure of the films. The peeling of two 1 nanofibril size significantly increases the interfacial area and conse­
welded films will occur when the yield strength of the film is greater quently promotes crack bridging during peeling fracture. Also, the
than the adhesion strength of the interface inside the welded area. WAXD results show that organoclay forms intercalated/partially exfo­
However, if the yield strength of the film is smaller than the interfacial liated stacks during the melt mixing process (see Supporting Informa­
adhesion, the film will yield and become elongated prior to being peeled tion). It has been shown that dispersion and distribution of partially
at the interface. The force-displacement curves of the B10 films in exfoliated organoclay stacks in the PE matrix induces peelability in PE
Fig. 5b at 135 and 145 ◦ C are good examples of the elongation of the films [46]. A moderate peelability performance has also been reported
films before the peeling of the system. The effect of the yield strength of for polyolefin based nanocomposite blends through the localization of
the blend and nanocomposite films on the peeling performance is further the organoclay at the interface [47]. The organoclay stacks can be
discussed in the Supporting Information. In cohesive peel fracture, delaminated at the interface and create multiple crazes which promote
cracks are initiated at the interface and then propagate to the nearest the cohesive peeling of the system. By increasing the PB-1 content, the
domain to form a consistent fracture throughout the blend. Thus, the thickness of the PB-1 fibrils grows while the interparticle distance re­
morphology of the dispersed phase and particularly the distance be­ duces (Table 1). Although the addition of 1 phr organoclay slightly re­
tween neighboring domains plays an important role on the peel per­ duces the interparticle distance and thickness of the PB-1 nanofibrils in
formance of the sealed blend films. The SEM micrographs of the blend the nanocomposite blends with 10 and 20 wt% PB-1, the reduction is
and nanocomposite films in Fig. 2 and Figure S1 are quantified and the much significant in the system with 5 wt% PB-1. This can be attributed
results are presented in Table 1. The results indicate that the fibrillar PB- to the higher interfacial coverage of organoclay particles at the interface
1 domains in the B5 and B10 blend films are dispersed with an average in the latter system.
interparticle distance of 364 and 302 nm, respectively. The morphology One major drawback of PB-1 and its blends is their aging due to PB-
of the B20 blend is more sheet like rather than fibril like with an average 1’s polymorphism which changes the physical and mechanical proper­
interparticle distance of 287 nm. The peel force seems to have a direct ties [52,53]. It has been reported that upon aging of PB-1 at room
relationship with the interparticle distance in the blend films. By temperature the peel strength of its polyolefin blends reduces by aging
decreasing the interparticle distance the peel force reduces over a rela­ time [54,55]. Fig. 7 shows the peel force of the PE/PB-1 blends and
tively wide ΔTp of 70 ◦ C for the B20 blend when compared to the B5 and nanocomposite films with the aging time after the sealing process. The
B10 blends with a narrower ΔTp of 10 and 20 ◦ C, respectively. The peel force of the B5 blend sealed at 130 ◦ C slightly decreases from 1010
interparticle distance of the dispersed PB-1 phases in the nanocomposite to 936 N/m after 24 h and remains constant afterwards. The B10 blend
films decrease to 160, 260 and 170 nm for NB5, NB10 and NB20, shows a peel force of 645 N/m when sealed at 135 ◦ C while its peel force
respectively. The results suggest that the change in the peel behavior of decreases to 490 N/m after 72 h and then remains unchanged. The peel
the systems containing 5 wt% PB-1 from lock seal to peelable by the force of the B20 film sealed at 150 ◦ C decreases from 310 to 195 N/m
incorporation of 1 phr organoclay can be partly attributed to the after 7 days and then levels off. In general, the peel force decreases by
reduction in the interparticle distance. The crack propagation is facili­ time indicating an aging effect in all the PE/PB-1 blend films. The DSC
tated when the interparticle distance reduces. The nanocomposite films thermal results in Figure S3 confirm that the metastable form II crystal of
with higher PB-1 contents show similar peel force with wider peelable PB-1 forms after the sealing process. The endotherm peaks at 115 ◦ C
window when are compared with the blend films. correspond to the melting peak of crystal form II in these systems.
In order to better understand the peeling mechanism in the PE/PB-1 Although the melting peak of PE is at 110 ◦ C, which is very close to the

6
R.S. Mohammadi et al. Composites Part B 218 (2021) 108930

Fig. 6. SEM micrographs of the peel fractured surface of (a) B5 sealed at 130 ◦ C, (b) NB5 sealed at 200 ◦ C, (c) NB10 sealed at 160 ◦ C, (d) NB10 sealed at 165 ◦ C, (e)
NB20 sealed at 180 ◦ C and NB20 sealed at 195 ◦ C (The white bars denote 20 μm).

melting peak of the form II crystals, the melting peak at 115 ◦ C can be
only assigned to the form II crystal as the melting peak of the form I
crystal appears at 120 ◦ C. These results imply that the metastable form II
converts to the stable and thermodynamically favored form I upon time
aging. As the form I is denser than the form II, the dispersed PB-1
nanofibrils shrink during the form II to form I transition and conse­
quently the peel force diminishes by the aging time until it reaches a
constant value. The aging results reveal that the form II to form I tran­
sition progresses faster in the B5 blend film (24 h) compared to the B10
and B20 blends that require 72 and 168 h to complete the transition,
respectively.
In contrast to the blend films, the peel force of the nanocomposite
blends remains constant after the sealing process with no aging (Fig. 7).
This can be attributed to the absence of the form II crystal in the
nanocomposite films as the aging is a direct consequence of the from II to
form I transition. The WAXD results in Fig. 8 confirm the formation of
form I/I’ in the nanocomposite films immediately after the heat-sealing
process. The reflection at 2θ = 10◦ corresponds to (110) plane of form I
or Iʹ and is only observed in nanofibrillar blends of the PE/PB-1 blends
Fig. 7. Aging of the peel force of the blend and nanocomposite films after the and nanocomposites [56]. The PE/PB-1 blend films show reflections at
sealing process. 2θ = 11.9, 16.9 and 18.2 corresponding to (200), (220) and (213)/(311)
planes of crystal form II which are absent in the WAXD patterns of the
nanocomposite blends. It has been shown that both the nanofibrillated
morphology of the PB-1 domains and the localization of organoclay at

7
R.S. Mohammadi et al. Composites Part B 218 (2021) 108930

that the organoclay and nanoconfinement effect induced on the PB-1


phase significantly affect the crystallinity of PB-1 and its aging. The
blend films containing different amount of PB-1 show the aging of the
peel force in which the peel force decreases by time after the sealing
process. The aging effect is significant for the blends with higher PB-1
contents. The reduction of the peel force is attributed to the trans­
formation of form II to form I crystals in the blend films after sealing. It
was shown that the incorporation of organoclay to the blends
completely eliminates the aging in the nanocomposite films. This is
achieved by suppressing the formation of the metastable form II crystals
and promoting the direct formation of thermodynamically stable form I
and III crystals after the sealing process. These results indicate the
importance of both a nanofibrillar morphology and interfacial locali­
zation of organoclay particles in obtaining films with a high peeling
performance.

Declaration of competing interest

The authors declare that they have no known competing financial


interests or personal relationships that could have appeared to influence
the work reported in this paper.

Acknowledgement

The authors acknowledge the financial support from the Natural


Sciences and Engineering Research Council of Canada (NSERC) and
industrial partner ProAmpac for the 3S Pack NSERC-Industrial research
chair that supported this work.

Appendix A. Supplementary data

Fig. 8. WAXD patterns of the blend and nanocompositefilms of PE/PB-1 con­ Supplementary data to this article can be found online at https://doi.
taining 5, 10 and 20 wt% of PB-1 immediately after the heat-seal process.
org/10.1016/j.compositesb.2021.108930.

the interface alter the kinetics of the PB-1 crystallization [56]. Author statement
Furthermore, the PE matrix solidifies before the PB-1 domains after the
heat-sealing process due to its lower crystallization temperature Raziyeh S. Mohammadi: Conceptualization, Methodology, Inves­
compared to that of the PB-1 phase, and thus imposes a nanoconfine­ tigation, Validation, Formal analysis, Data curation, Writing-Original
ment restriction on the nanofibrillated PB-1 domains. The nano­ draft, Writing-Review & Editing.
confinement effect in conjunction with the hindrance effect created by Ali M. Zolali: Conceptualization, Methodology, Validation, Formal
the organoclay layers at the interface limit the movement of the PB-1 analysis, Writing-Review & Editing.
chains. This may disrupt the crystallization kinetic of the form II and Seyed H. Tabatabaei: Funding acquisition, Resources.
supress its formation as the form II crystal is kinetically favored. The Abdellah Ajji: Supervision, Funding acquisition, Project adminis­
result is the direct formation of form I and form III modification in the tration, Writing- Review & Editing.
nanocomposite films which are not observed in the blend films under the
similar sealing process.
References

4. Conclusions [1] Paul DR, Bucknall CB. Polymer blends. New York: John Wiley & Sons, Inc.; 2000.
[2] Fu S-Y, Feng X-Q, Lauke B, Mai Y-W. Effects of particle size, particle/matrix
A novel PE/PB-1/organoclay nanocomposite with a broad peelable interface adhesion and particle loading on mechanical properties of
particulate–polymer composites. Compos B Eng 2008;39:933–61.
seal-temperature range is reported for seal/peel applications. The [3] Ebadi-Dehaghani H, Khonakdar HA, Barikani M, Jafari SH. Experimental and
incorporation of either 1 phr organoclay or 5 wt% of PB-1 into PE results theoretical analyses of mechanical properties of PP/PLA/clay nanocomposites.
in films with a lock-seal performance after the sealing process similar to Compos B Eng 2015;69:133–44.
[4] Zolali AM, Favis BD. Compatibilization and toughening of co-continuous ternary
that of the neat PE. However, the PE/PB-1/organoclay containing both 1 blends via partially wet droplets at the interface. Polymer 2017;114:277–88.
phr organoclay and 5 wt% of PB-1 renders a peelable film with a very [5] Chen J, Chen JW, Chen HM, Yang JH, Chen C, Wang Y. Effect of compatibilizer and
broad ΔTp of 90 ◦ C. The TEM and thermodynamic analysis confirm that clay on morphology and fracture resistance of immiscible high density
polyethylene/polyamide 6 blend. Compos B Eng 2013;54:422–30.
organoclay is localized at the interface of the PE and nanofibrillar PB-1 [6] Si M, Araki T, Ade H, Kilcoyne ALD, Sokolov JC, Rafailovich MH, Fisher R.
phases. The localization of organoclay at the interface compatibilizes the Compatibilizing bulk polymer blends by using organoclays compatibilizing bulk
morphology and reduces the PB-1 nanofibrillar thickness which justifies polymer blends by using organoclays. Macromolecules 2006;39:4793–801.
[7] Eagan JM, Xu J, Di Girolamo R, Thurber CM, Macosko CW, LaPointe AM, Bates FS,
the observed versatile peel performance in the nanocomposite blends.
Coates GW. Combining polyethylene and polypropylene: enhanced performance
Although organoclay is found very effective in enhancing the peelability with PE/i PP multiblock polymers. Science 2017;355:814–6. 80.
of the PE/PB-1 films, increasing the PB-1 content to 20 wt% negatively [8] Zolali AM, Favis BD. Toughening of cocontinuous polylactide/polyethylene blends
via an interfacially percolated intermediate phase. Macromolecules 2018;51:
affects the peelability of the nanocomposite blends. This was attributed
3572–81.
to the shortage of organoclay particles to cover the increased interfacial [9] Lin Y, Chen H, Chan C-M, Wu J. High impact toughness polypropylene/CaCO 3
area in the systems with higher PB-1 contents. Furthermore, it was found nanocomposites and the toughening mechanism. Macromolecules 2008;41:
9204–13.

8
R.S. Mohammadi et al. Composites Part B 218 (2021) 108930

[10] Xu Y, Delgado P, Todd AD, Loi J, Saba SA, McEneany RJ, Tower T, Topolkaraev V, [36] Stolte I, Androsch R. Kinetics of the melt - form II phase transition in isotactic
Macosko CW, Hillmyer MA. Lightweight micro-cellular plastics from polylactide/ random butene-1/ethylene copolymers. Polymer 2013;54:7033–40.
polyolefin hybrids. Polymer 2016;102:73–83. [37] Zhang B, Yang D, Yan S. Direct formation of form I poly(1-butene) single crystals
[11] Yoxall A, Janson R, Bradbury SR, Langley J, Wearn J, Hayes S. Openability: from melt crystallization in ultrathin films. J Polym Sci Part B Polym Phys 2002;40:
producing design limits for consumer packaging. Packag Technol Sci 2006;19: 2641–5.
219–25. [38] Yamashita M, Ueno S. Direct melt crystal growth of isotactic polybutene-1 trigonal
[12] Liebmann A, Schreib I, Schlözer RE, Majschak J-P. Practical case studies: easy phase. Cryst Res Technol 2007;42:1222–7.
opening for consumer-friendly, peelable packaging. J Adhes Sci Technol 2012;26: [39] De Rosa C, Ruiz De Ballesteros O, Auriemma F, Di Girolamo R, Scarica C, Giusto G,
2437–48. Esposito S, Guidotti S, Camurati I. Polymorphic behavior and mechanical
[13] Nase M, Großmann L, Rennert M, Langer B, Grellmann W. Adhesive properties of properties of isotactic 1-butene-ethylene copolymers from metallocene catalysts.
heat-sealed EVAc/PE films in dependence on recipe, processing, and sealing Macromolecules 2014;47:4317–29.
parameters. J Adhes Sci Technol 2014;28:1149–66. [40] Wang Y, Lu Y, Zhao J, Jiang Z, Men Y. Direct formation of different crystalline
[14] Nase M, Langer B, Grellmann W. Fracture mechanics on polyethylene/polybutene- forms in butene-1/ethylene copolymer via manipulating melt temperature.
1 peel films. Polym Test 2008;27:1017–25. Macromolecules 2014;47:8653–62.
[15] Hwo CC. Polybutylene blends as easy open seal coats for flexible packaging and [41] Su F, Li X, Zhou W, Zhu S, Ji Y, Wang Z, Qi Z, Li L. direct formation of isotactic poly
lidding. J Plastic Film Sheeting 1987;3:245–60. (1-butene) form I crystal from memorized ordered melt. Macromolecules 2013;46:
[16] Luciani L, Seppälä J, Löfgren B. Poly-1-butene: its preparation, properties and 7399–405.
challenges. Prog Polym Sci 1988;13:37–62. [42] Taguet A, Cassagnau P, Lopez-Cuesta J-M. Structuration, selective dispersion and
[17] Schemm F, De Vliet V, Grasmeder J, Cocola F. Basell polybutene-1 : high compatibilizing effect of (nano)fillers in polymer blends. Prog Polym Sci 2014;39:
performance polyolefin for plumbing appilcations. 2000. 1526–63.
[18] Nase M, Zankel A, Langer B, Baumann HJ, Grellmann W, Poelt P. Investigation of [43] Vo LT, Giannelis EP. Compatibilizing poly(vinylidene fluoride)/nylon-6 blends
the peel behavior of polyethylene/polybutene-1 peel films using in situ peel tests with nanoclay. Macromolecules 2007;40:8271–6.
with environmental scanning electron microscopy. Polymer 2008;49:5458–66. [44] Tiwari RR, Paul DR. Effect of organoclay on the morphology, phase stability and
[19] Stehling FC, Meka P. Heat sealing of semicrystalline polymer films. II. Effect of mechanical properties of polypropylene/polystyrene blends. Polymer 2011;52:
melting distribution on heat-sealing behavior of polyolefins. J Appl Polym Sci 1141–54.
1994;51:105–19. [45] Bitinis N, Sanz A, Nogales A, Verdejo R, Lopez-Manchado Ma, Ezquerra Ta.
[20] Mueller C, Capaccio G, Hiltner A, Baer E. Heat sealing of LLDPE: relationships to Deformation mechanisms in polylactic acid/natural rubber/organoclay
melting and interdiffusion. J Appl Polym Sci 1998;70:2021–30. bionanocomposites as revealed by synchrotron X-ray scattering. Soft Matter 2012;
[21] Natta G, Corradini P, Bassi IW. Crystal structure of isotactic poly-alpha-butene. 8:8990.
Nuovo Cim 1960;15:52–67. [46] Mohammadi RS, Tabatabaei SH, Ajji A. Peelable clay/PE nanocomposite seals with
[22] Miller RL, Holland VF. On transformations in isotactic polybutene-1. J Polym Sci B ultra-wide peelable heat seal temperature window. Appl Clay Sci 2018;158:
Polym Lett 1964;2:519–21. 132–42.
[23] Jones AT. Polybutene-1 – type II crystalline form. J Polym Sci B Polym Lett 1963;1: [47] Mohammadi RS, Tabatabaei SH, Ajji A. Effect of nanoclay localization on the peel
455–6. performance of PE based blend nanocomposite sealants. Appl Clay Sci 2018;152:
[24] Holland VF, Miller RL. Isotactic polybutene-1 single crystals: Morphology. J Appl 113–23.
Phys 1964;35:3241–8. [48] Mirabella FM, Bafna A. Determination of the crystallinity of polyethylene/α-olefin
[25] Gohil RM, Miles MJ, Petermann J. On the molecular mechanism of the crystal copolymers by thermal analysis: relationship of the heat of fusion of 100%
transformation (Tetragonal-Hexagonal) in polybutene-1. J Macromol Sci Part B polyethylene crystal and the density. J Polym Sci Part B Polym Phys 2002;40:
1982;21:189–201. 1637–43.
[26] Azzurri F, Flores A, Alfonso GC, Sics I, Hsiao BS, Calleja FJB. Polymorphism of [49] Elias L, Fenouillot F, Majesté JC, Alcouffe P, Cassagnau P. Immiscible polymer
isotactic polybutene-1 as revealed by microindentation hardness. Part II: blends stabilized with nano-silica particles: rheology and effective interfacial
correlations to microstructure. Polymer 2003;44:1641–5. tension. Polymer 2008;49:4378–85.
[27] Alfonso GC, Azzurri F, Castellano M. Analysis of calorimetric curves detected [50] Wu L, Luo X, Wang XD. Influence of processing conditions on dual-phase
during the polymorphic transformation of isotactic polybutene-1. J Therm Anal continuous blend system of thermoplastic polyurethane with ethylene-propylene-
Calorim 2001;66:197–207. diene monomer elastomer. J Appl Polym Sci 2006;102:5472–82.
[28] Hong KB, Spruiell JE. The effect of certain processing variables on the form ii to [51] Manias E, Zhang J, Huh JY, Manokruang K, Songtipya P, Jimenez-Gasco MM.
form i phase transformation in polybutene-1. J Appl Polym Sci 1985;30:3163–88. Polyethylene nanocomposite heat-sealants with a versatile peelable character.
[29] Kalay G, Kalay CR. Interlocking shish-kebab morphology in polybutene-1. J Polym Macromol Rapid Commun 2009;30:17–23.
Sci Part B Polym Phys 2002;40:1828–34. [52] Schultz J, Hsiao B, Samon J. Structural development during the early stages of
[30] Choi CH, White JL. Crystal-crystal transformations in isotactic polybutene-1 polymer melt spinning by in-situ synchrotron X-ray techniques. Polymer 2000;41:
oriented filaments and in thick molded rods. Polym Eng Sci 2001;41:933–9. 8887–95.
[31] Samon JM, Schultz JM, Hsiao BS, Wu J, Khot S. Structure development during melt [53] Azzurri F, Flores A, Alfonso GC, Baltá Calleja FJ. Polymorphism of isotactic poly(1-
spinning and subsequent annealing of polybutene-1 fibers. J Polym Sci Part B butene) as revealed by microindentation hardness. 1. Kinetics of the
Polym Phys 2000;38:1872–82. transformation. Macromolecules 2002;35:9069–73.
[32] Liu Y, Cui K, Tian N, Zhou W, Meng L, Li L, Ma Z, Wang X. Stretch-induced [54] Nase M, Androsch R, Langer B, Baumann HJ, Grellmann W. Effect of polymorphism
crystal–crystal transition of polybutene-1: an in situ synchrotron radiation wide- of isotactic polybutene-1 on peel behavior of polyethylene/polybutene-1 peel
angle X-ray scattering study. Macromolecules 2012;45:2764–72. systems. J Appl Polym Sci 2008;107:3111–8.
[33] Hu J, Tashiro K. Relation between higher-order structure and crystalline phase [55] Nase M, Androsch R, Henning S, Grellmann W. Influence of the Form II/Form I
transition of oriented isotactic polybutene-1 investigated by temperature- crystal polymorphism of random copolymers of butene-1 with ethylene or
dependent time-resolved simultaneous WAXD/SAXS measurements. Polymer propylene on the peel behavior of peel films. Polym Eng Sci 2015;55:749–1757.
2016;90:165–77. [56] Mohammadi RS, Zolali AM, Tabatabaei SH, Ajji A. Nanoconfinement induced
[34] Zhang X, Zhang X, Shi G. The effect of some additives on the Form II to Form I direct formation of form I and III crystals inside in situ formed poly(butene-1)
phase transformation in polybutene-1. Thermochim Acta 1992;205:245–52. nanofibrils. Macromolecules 2020;53:1346–55.
[35] Nase M, Androsch R, Henning S, Grellmann W. Influence of the Form II/Form I
crystal polymorphism of random copolymers of butene-1 with ethylene or
propylene on the peel behavior of peel films. Polym Eng Sci 2015;55:749–1757.

You might also like