Carvalho Et Al JAPS - Corrigido Vinícius Após Reviewers Reports - Color Graphics

You might also like

Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 30

In Situ Compatibilization of a Polyethylene, Polypropylene and Polystyrene

Ternary Blend through Friedel-Crafts Alkylation

Vinícius L. Carvalhoa, Camila Safieddinea, Nicole R. Demarquetteb,

Luís A. Pinheiroc*1
a
Postgraduate Program in Materials Engineering and Science, State University of Ponta

Grossa (UEPG), Ponta Grossa, Paraná, Brazil


b
Ecole Technol Super, Dept Genie Mecan, Montreal, Canada
c
Department of Materials Engineering, State University of Ponta Grossa (UEPG), Ponta

Grossa, Paraná, Brazil

ABSTRACT

The Friedel-Crafts alkylation reaction has been applied to reactively compatibilize a ternary

blend of high density polyethylene (HDPE), polypropylene (PP), and polystyrene (PS). The

reactions were carried out in an internal mixer using varying catalyst concentrations. The

resulting compatibilizer was quantified after Soxhlet extraction. In addition, p-substitution due

to the grafting of alkyl groups onto the polystyrene benzene ring was identified via nuclear

magnetic resonance spectroscopy (NMR). The size of the PS domain in the reactive

compositions decreased by 80%. Moreover, the phase in which PS droplets were dispersed

varied, i.e., in the non-reactive blends they were found in the PP phase, and in the reactive

blends they shifted towards the HDPE phase. The effect of the compatibilizing agent was to

improve the mechanical properties of the blend. Even with the lowest catalyst content, the

1
*Corresponding author: Luís A. Pinheiro, e-mail: lapinheiro@uepg.br
properties of elongation at break, tensile strength, toughness and elastic modulus showed

improvements.

Keywords: Compatibilization, Friedel-Crafts, Morphology, Reactive blend, Ternary blend.

1. INTRODUCTION

A considerable fraction of municipal solid waste is constituted by plastic, up to 15% [1].

These causing major environmental damage when improperly arranged [2]. The largest

polymer production, 80%, is made up of thermoplastics, with a consequent potential for

technological recycling [3]. In the last decade, only 18% of Brazilian cities had municipal solid

waste management, produce up on average 180 tons / day of thermoplastics (PE, PP and

PS) from the polymeric fraction [4-6]. Reuse this fraction technologically as a blend is a

challenge [7]. Studies of polymer blends are important because they underpin the

development of new materials with differentiated properties and post-consumer plastic

recycling, combining individual characteristics or developing new properties. The obstacles to

be overcome in such cases are polymer processing volume and the recovery of

physicochemical properties lost through degradation. Moreover, poor interfacial adhesion

between phases (given that most polymer blends are thermodynamically immiscible and

incompatible) increases interfacial stresses, resulting in phase segregation, delamination, and

low mechanical properties [8-10]. In a study of binary blends of polyolefins and polystyrene,

Kallel et al. [11] reported that one of the main problems in processing post-consumer

polymers was the deterioration of properties resulting from the incompatibility of components.

The immiscibility and incompatibility of polymer blends can be overcome by using

compatibilization techniques to improve the blend’s properties and decrease interfacial


stresses, thus enabling interphase adhesion, stabilizing the morphology and preventing the

coalescence of dispersed particles [12-18]. Compatibilizing agents must have high molecular

weight or be able to form components of high molecular weight in the blend [13]. These

agents are located at interfaces, and parts of their molecular chain interact in some way with

each polymer phase. Compatibilizing agents produced by chemical reactions during the blend

process (in situ) are of particular interest [13-20].

The chemical structure of high density polyethylene (HDPE), polypropylene (PP), and

polystyrene (PS) lacks functional groups that enable direct interaction to form copolymers.

Friedel-Crafts (F-C) alkylation may be an alternative route to promote their compatibilization

by para-substitution in the PS phenyl ring, coupling a polyolefin chain to it [21-23]. F-C

alkylation has been employed to compatibilize PE/PS and PP/PS polymer pairs [21-32]. A

proton of the aromatic ring may undergo electrophilic substitution with a halogen or an olefin

in the presence of strong Lewis acid [34]. Increased viscosity due to the presence of

compatibilizing agents used in Friedel-Crafts alkylation of polyolefin-PS binary blends was

demonstrated by Sheng and Li [34, 35] based on rheological properties and by Sun et al. [24]

based on the melt flow index.

Carrick [20] was the first to produce a PE-g-PS graft copolymer via F-C alkylation,

using aluminum chloride as catalyst in hexane solution at 81°C. The author demonstrated

that grafts reach maximum values as a function of the reaction time, and that polymer chain

scission and compatibilizer production are competing phenomena. Heikens and Barentsen

[21, 22] replicated the method and used the graft copolymer to compatibilize the molten

PE/PS blend. They identified a correlation between improved mechanical properties and

morphology, and also found that the copolymer activity had an emulsifying effect. Baker and

Sun [23] tested different Lewis acids and co-catalysts to develop the Friedel-Crafts route in a

PE/PS blend. They reported that grafting efficiency was improved by using anhydrous
aluminum chloride as catalyst and styrene monomer as co-catalyst. In another study, Sun et

al. [24] used the extrusion process to compatibilize PE/PS and PP/PS blends via the F-C

alkylation reaction. The competing behavior between graft copolymer formation and PP and

PS chain degradation was revealed by the increase in the mixing level inside the extruder.

Gao et al. [25] subjected LLDPE/PS and LLDPE/HIPS blends to F-C alkylation, and

determined the chemical structure of the resulting copolymer by means of RAMAN

spectroscopy. They confirmed that PE segments were grafted onto the para position of the

PS aromatic ring. Shahbai et al. [32], who were the first researchers to use HDPE with PS,

observed grafting at about 15% of the initial PS when the blend was compatibilized by the F-C

reaction. The authors confirmed the positive effects of the reactive compatibilization based on

the correlation between rheological properties and morphology.

Diaz et al. [26, 30] compatibilized a PE/PS blend using the F-C reaction and varying

the molar mass of the polyolefin, and examined the copolymer structures and their effects on

the mechanical properties. These authors showed that lower molecular segments are

preferentially grafted to the PS ring. This behavior confirms the theory proposed by Helfand

and Tagami [36], whereby species with lower molar mass tend to be located at the interfaces

due to their higher mobility. Diaz et al. [26, 30] also demonstrated that PE does not undergo

degradation when it reacts solely with catalysts, regardless of its molar mass. The

improvement in its chemical properties is significant when compared to non-reactive blends.

Diaz et al. [27, 28] also studied PP/PS blends compatibilized via F-C alkylation. The

degradation of PP was detected at the same time as the formation of copolymer. This

degradation process is catalytic, as is the degradation mechanism of PS, due to the presence

of Lewis acid. Diaz et al. [27, 30] also reported a point where the graft reaction was optimal

relative to the amount of catalyst they used. The authors examined the mechanical properties

and average particle size, and reported that optimum results were achieved in the graft
reaction with 0.4 wt% AlCl3 in the PE/PS blends, and with 0.7 wt% catalyst in the PP/PS

blend.

Strong Lewis acid is responsible for the catalytic degradation of PS [31, 32, 37],

whereas polyolefins undergo marginal thermo-oxidative degradation due to processing

conditions in the F-C compatibilizing route [27, 32]. The catalytic degradation of PS is reduced

by using a co-catalyst, which acts as a more reactive intermediary during the chemical

reaction mechanism [23] and as a reformer of degraded segments of PS [31].

The number of scientific studies that have employed Friedel-Crafts alkylation to

compatibilize binary blends justifies the efforts aimed at exploring its possible application in a

new scenario: ternary blends. This study evaluated the use of F-C alkylation to compatibilize

HDPE, PP, and PS blends, in terms of amounts of catalyst and blend compositions.

2. EXPERIMENTAL

2.1. Materials

Commercial grade HDPE (IE59U3) and PP (H301) were supplied by Braskem, and PS

came from Innova (N1941). These products presented densities of 0.96, 0.9 and 1.05 g/cm³

and melt flow rates of 5, 10 and 11 g/10min, respectively. Anhydrous AlCl 3 in powder form

with 99% purity was purchased from Vetec Química Fina (Brazil). The styrene monomer,

which was donated by Harima Chemicals, was produced by Brisco do Brasil in liquid form

with anti-polymerization stabilizers.

2.2. Blend preparation

Table 1 describes the blend compositions, which were mixed at 200°C in a HAAKE

internal mix, Rheomix QC Lab Mixer, operating at a rotor speed of 100 rpm. The internal
volume was kept constant at 83% of the mixer’s total volume of 69 cm³. The homopolymer

mass was calculated in blend fractions based on the polymer density. The same calculation

was applied to the catalyst (AlCl3) and co-catalyst (styrene monomer). The co-catalyst was

kept constant at 0.6% w/w in all the reactive blends. Blending was carried out in a normal

atmosphere. Physical blending consisted of melting together only the polymers until the

torque curve stabilized. In the reactive blends, the catalyst and co-catalyst were added to the

already melted and mixed homopolymers (3 minutes after beginning), followed by 3.5 minutes

of reactive mixing to stabilize the torque curve.

Table 1. Catalyst compositions and weight fractions in the blends

Catalyst, %
Blends 0 0.3 0.5 0.7
(HDPE/PP/PS) Non-reactive
Reactive blends
blend
SET 1 50/30/20 S1-00 S1-03 S1-05 S1-07

SET 2 40/40/20 S2-00 S2-03 S2-05 S2-07

SET 3 80/15/05 S3-00 S3-03 S3-05 -

SET 4 80/05/15 S4-00 S4-03 S3-05 -

The main compositions were SET 1 and SET 2, which reflect the mass composition of

these polyolefins in the polymer fraction of Brazilian urban waste [5]. However, two additional

sets of reactive blends, SET 3 and SET 4, were prepared to corroborate the morphological

behavior of SET 1 and 2. The HDPE mass fraction in SET 3 and SET 4 was kept constant to

form a matrix phase, while the composition of the dispersed phases was changed.

2.3. Scanning Electron Microscopy


Various samples were subjected to cryogenic fracturing after 80 minutes of immersion

in liquid nitrogen in order to determine the type of final morphology resulting from Friedel-

Crafts compatibilization. The PS phase was then extracted by constant stirring overnight in

chloroform at room temperature. The samples were dried at room temperature for 48 hours

and then gold sputter coated using a Quorum SC7620 sputter coater. They were then

examined under a TESCAN MIRA3 field emission scanning electron microscope (SEM-FEG)

equipped with a secondary electron detector, and the data were processed using ImageJ

software. Particle size was calculated considering approximately 500 particles, with two

images for SET 1 and three images for SET 2 at different magnifications.

2.4. Solvent Extraction

Samples from SET 1 and SET 2 were sliced by hand and placed in a Soxhlet extractor

for continuous extraction and refluxing for 76 hours, using boiling xylene as solvent. Eq.1 was

used to calculate the amount of PS grafted based on gravimetric analysis.

m
%grafed= ( )
m0
.100 (1)

where, m 0, is the initial mass of de blend and m is the final one, both dried.

2.5. Nuclear Magnetic Resonance – NMR

Blends were solubilized at 107°C in toluene-d8 with 1% tetramethylsilane (TMS) as

reference in a 4mm NMR tube. Sample concentration was kept at 5 wt%. The samples were

examined by nuclear magnetic resonance using a Bruker FT-NMR Ascend 400 MHz
1
spectrometer equipped with temperature control, for H nuclei. ¹H NMR spectra were

recorded in 32 scans and 1s acquisition time.

2.6. Mechanical Properties


The samples used for the test were hot pressed into thin films at 200°C using a 200 μm

spacer mold, in the dimensions specified by ASTM-D882. Tensile testing was performed at

room temperature on a Shimadzu Autograph AGS-10kN mechanical tester operating at a

crosshead speed of 50 mm/min.

3. RESULTS AND DISCUSSION

3.1. Process characteristics

Fig. 1 and Fig. 2 show the rheograms of the two sets of blends. The drop observed in

the reactive blends indicates the moment when the chamber was opened to add the catalyst

and co-catalyst to the melted polymers. The difference between drop time of the sets 1 and 2

and sets 3 and 4 was the time of stabilization of the torque curve for each composition. The

torque stabilization of samples containing 0.3 wt% catalyst was similar to that of the non-

reactive blends, suggesting that polymer chain degradation and copolymer formation reached

equilibrium (Fig. 1). Increasing the catalyst content resulted in a relative increase in torque,

followed by a decrease. This was attributed to the competitive behavior between the graft

reaction and the catalytic degradation of PP and PS chains described by Diaz [26, 27, 31]. As

can be seen in Fig. 1(a) and (b), the HDPE content kept the major phase constant and did not

undergo degradation due to the presence of AlCl 3 [27]. Additional resistance imposed on the

system was observed, resulting from species with high molar mass, which enhanced

viscosity.
Fig. 1. Torque vs. time inside the mixer: (a) SET 1, and (b) SET 2

Fig. 2. Torque vs. time inside the mixer: (a) SET 3 and (b) SET 4

In SET 3 (Fig. 2(a)), the torque of blend S3-05 initially increased after the F-C reaction

but subsequently decreased. The ungrafted PP chains tended to degrade by catalytic


reaction. In SET 4 (Fig. 2(b)), S4-05 had fewer PP chains available for synthesis and

therefore showed little degradation behavior. In both SET 3 and 4, samples containing 0.3 wt

% catalyst each maintained their torque after presenting some increase. This was particularly

visible in blend S4-03, which showed gains in torque, indicating the presence of

compatibilizing agents in the reaction medium. Shahbazi et al. [32] reported similar results for

their HDPE/PP/PS blend compatibilized in situ via F-C reaction. Torque curves indicated the

formation of compatibilizing agents, since crosslinking during synthesis was discarded [27]

and the degradation processes – thermo-oxidative and catalytic – did not significantly affect

the most representative blends (S1-03; S2-03; S3-03 and S4-03).

3.2. Analysis of compatibilizer after Soxhlet extraction

After 76 hours of extraction, the non-reactive blends showed no extraction residue,

either because they had been completely solubilized or because their minuscule quantity was

undetectable. This finding was confirmed by the physical solubilization of the ternary blend

from the Soxhlet extractor. In this situation, 100% of the non-reactive blend was solubilized in

hot xylene.

On the other hand, the reactive blends had a residue remaining at the end of the

process, indicating that they are not completely soluble. This was attributed to the formation

of copolymers, which were separated at the end of the process. Copolymer concentrations

were calculated by the gravimetric method. These concentrations were 5.1, 5.7, and 5.4% for

0.3, 0.5, and 0.7 of catalyst, respectively, in SET 1, and 3.0, 4.1, and 3.0% for 0.3, 0.5, and

0.7 of catalyst, respectively, in SET 2. The composition with the highest HDPE content (SET

1) presented the highest copolymer concentrations. As for the catalyst concentration, both

sets produced more copolymer with 0.5% of AlCl 3.


Residual material was dried for 24 hours at 40°C, and then analyzed by differential

scanning calorimetry (DSC) and Fourier transform infrared spectroscopy (FTIR). The infrared

spectra of the extraction residue did not reveal absorption bands between 800-860 cm -1,

which is where p-substitution occurs. Fig. 3 shows the FTIR spectrum of the S2-03 reactive

blend after extraction, along with the spectrum of pure polymer. PP signals were present at

1375 cm-1 (asymmetric stretching of methyl groups), 1000 cm -1 (angular deformation) and

1170 cm-1 (angular deformation in CH3 torsion). The extracted material did not show the

characteristic peak of HDPE (at 2017 cm -1) in FTIR spectrum, but their presence can be seen

in a small sign of crystallization in the DSC, Fig. 4. The absorption bands at 3000-2840 cm -1

characteristic of the alkanes remained unchanged in the extracted material. The PS aromatic

ring was weakly detectable at 710-690 cm-1.

Fig. 3. FTIR spectra of the homopolymers compared with the insoluble fraction of the reactive

blend (S2-03).
Fig. 4 also illustrates the melt peaks of the reactive blend S2-05 extracted and non-

extracted by the Soxhlet technique, as well as the homopolymers. All the other reactive

blends which were also extracted continuously showed a similar thermal behavior. The non-

extracted reactive blend S2-05 presented melt temperatures characteristic of PP and HDPE.

The extracted S2-05 showed a sharp peak temperature around 166ºC, which was attributed

to PP phase, and a minor peak at 130 ºC (seven times smaller) attributed to HDPE. Thus, the

copolymer produced here presented both PP and HDPE polymer chains, but with

predominance of the first. However, two possible explanations can be proposed for the

difference in the magnitude of the peak melt temperatures: 1) the copolymer was composed

mainly of PP chains; and/or 2) the HDPE chains in the copolymer were prevented from

crystallizing.
D SC (m W )
100 120 140 160 180

Temperature (ºC)
D S C (m W )
HDPE

PP

S2-05 before extraction

S2-05 after extraction

50 100 150 200

Temperature (ºC)

Fig. 4. DSC identification of the homopolymers in the reactive blend (S2-05) based on

thermal variables.

In this study, the FTIR results were confirmed by DSC, showing that the residual

material consisted mainly of PP chains. This finding is in agreement with the mechanism of

the Friedel-Crafts reaction, which requires the presence of a more stable intermediary species

in the reaction medium. The results lead to the conclusion that segments of PP polymer were

grafted onto the PS aromatic ring simultaneously to the HDPE ones, producing a terpolymer,

expanding the compatibilizing agents formed by the Friedel-Crafts reaction. The ease with

which polymer segments become diffused in the molten medium and the stability of each

segment determines the predominance of one or the other in the terpolymer structure.

3.3. Grafting reaction revealed by NMR analysis

Fig. 5 depicts the 1H NMR spectrum of the non-reactive blend (S2-00), clearly showing

the deuterated solvent (S) and the reference substance (TMS) at 7.0–7.1 and 0 ppm,

respectively [38]. The polyolefin signals were in agreement with the non-reactive blends
studied by Pozzi and Marini [38], who solubilized PE/PP blends with deuterated toluene and

quantified each phase based on 1H NMR spectra. Both HDPE and PP presented signals of

primary alkyl, R-CH3, at 0.8–1.0 ppm, and secondary alkyl, R-CH 2-R, at 1.2–1.4. However,

only PP had a clear tertiary alkyl signal, R 3C-H, at 1.4–1.7, where R corresponds to alkyl

groups [39]. Moreover, the HDPE presented a strong secondary alkyl signal at 1.35 ppm,

while the PP showed weak signals of this group between 1.4 and 1.5 ppm, because the peak

area was proportional to the number of protons that originated the signal. The studies by

Bevington and Huckerby [40] and Takeichi et al. [41] helped to identify the characteristic PS

at 6.45–6.75 ppm. However, part of it was superimposed by the solvent peaks.

Fig. 5. 1H NMR spectra of the non-reactive blend (S2-00). Homopolymer signals are

highlighted.
Fig. 6 reveals changes in the spectral signals of SET 2 in response to increasing

catalyst concentrations. Two details were observed during the in situ F-C compatibilization:

spectral signals of p-substitution arose at 6.94 and 7.04 ppm (Fig. 6(a)), confirming the

polymeric coupling to the PS aromatic ring due to the formation of copolymer and when the

catalyst concentration increased, the p-substitution signal of the alkyl groups became

stronger. Trace quantities of PS diminished due to its catalytic degradation, Fig. 6(a).

Another noteworthy fact is that the signal of PP secondary alkyls increased (Fig. 6(b)).

This was attributed to the disproportionation reaction signal caused by the formation of R=CH 2

ending groups, as reported by Diaz [27]. An increase in this range may also indicate linkage

by PP macrocarbocations in the aromatic ring through the hydrogen belonging to the primary

alkyl carbon. However, this hypothesis is unlikely to be confirmed, given the weak chemical

stability of a primary intermediary species in the reactive medium. Nevertheless, Mustafa et

al. [42] already observed this fact when a styrene monomer participated alone in the graft

reaction. Co-catalyst protons were not detected [43].


Fig. 6. Details of 1H NMR spectra in response to increased catalyst content: (a) Aromatic

range, (b) Secondary alkyl range. Circled peaks in (a) are attributed to p-substitution in the PS

aromatic ring.

3.4. Morphology and particle size

The morphological behavior of HDPE/PP/PS ternary blends of SET 2 compatibilized by

F-C alkylation showed polyolefin phases quite similar to those of the co-continuous

concentration, with PS phase acting as dispersed domains. SET 1 presented the same

behavior. Favis et al. [44-47] defined the ternary morphology as an equilibrium between
‘complete wetting’ and ‘partial wetting’ morphologies. In the former, the thermodynamic

driving force makes PS droplets spread out across the PP phase (spreading coefficient,

Λ PS ∕ PP ∕ PEAD >0 ), while in the latter, all the spreading coefficients are negative, being

Λ PS ∕ PEAD ∕ PP < Λ PS / PP/ PEAD <0 . In Fig. 7(a), HDPE and PP are distinguishable due to the cryogenic

preparation: the latter undergoes brittle fracture (smooth surface), whereas the former

undergoes ductile fracture (rough surface). The PS phase is visible by the holes left after

chloroform extraction.

In the non-reactive blend, PS domains are dispersed in the PP phase; however, the

use of a catalyst/co-catalyst system for the F-C reaction seems to draw PS particles to the

HDPE phase. The migration of PS droplets from PP to the HDPE phase and the emulsifying

effect of compatibilization become visible upon increasing the catalyst concentration. This

migration is greater at the HDPE-PS interface than at the PP-PS interface, causing the

displacement of PS dispersed domains. This phenomenon is attributed to the stronger

decrease in interfacial stresses between HDPE and PS than between PP and PS, resulting

from the action of the compatibilizing agents formed in situ. The lower grafting tendency of the

HDPE segments than the PP segments, as observed by FTIR and DSC, enabled the action of

the HDPE to be more effective at the interface. This behavior can be attributed to the higher

mobility of the HDPE segments in the corresponding homopolymer phase than that of the PP

segments.

The particle size decreased down to a content of 0.3 wt% aluminum chloride. The

emulsifier effect decreased until the particles reached the critical micelle concentration, but

began to increase in response to increasing catalyst content, with a tendency for the

occurrence of a morphological inversion. However, at any catalyst concentration, the particle


size was smaller than in the non-reactive blends. Diaz [27, 31] noted that F-C

compatibilization had different effects on PP/PS and PE/PS blends.

Fig. 7. Morphological behavior of physical and reactive blends: a) S2-00, b) S2-03, c) S2-05

and d) S2-07.

A detailed examination of the SET 2 interface can be observed in Fig. 8. The non-

reactive blend (S2-00) showed only a few PS particles at the interface, with most of them

remaining dispersed in the PP phase. However, in the reactive blend, where copolymer is

formed, numerous PS droplets were visible at the interface of the PP and HDPE phases,

which is attributed to the formation of a terpolymer at the interface [48], corroborating FTIR
and DSC data. Formation of the terpolymer was attributed to the close HDPE/PP interfacial

contact that occurred in response to partial wetting.

Fig. 8. Close-up image of the interface of the blends: a) S2-00 and b) S2-03 in SET 2.

By keeping the HDPE mass fraction high in ternary blends (matrix phase) and

changing the composition of dispersed phases, which was done in SET 3 and SET 4, the PP

phase was encapsulated by the HDPE matrix (Fig. 9). However, despite this difference

between HDPE and PP concentrations (80 and 15 in SET 3 and 80 and 5 in SET 4,

respectively), the PS droplets were distributed in the same manner as in the previous

compositions: inside the PP phase in the non-reactive blends, and at the interface and inside

the HDPE phase in the reactive blends. This behavior suggests that the concentration of PS

did not affect its distribution inside the physical and reactive blends.
Fig. 9. Affinity of PS phase altered by compatibilization in: a) S3-00, b) S3-03, c) S4-00 and d)

S4-05.

Fig. 10 shows a quantitative analysis of the emulsifying effect of the compatibilizing

agent in SET 2. The same behavior was found to occur in SET 1. The curve presented an

optimum point of particle size reduction, which means that small amounts of catalyst sufficed

to decrease the particle size in the dispersed phase. The aspect ratio indicates the low
deformation of the PS dispersed phase, possibly as a results of the viscosity ratio and particle

stabilization due to the copolymer at the interface.

Fig. 10. Effect of wt% catalyst on the particle size (Feret’s diameter) in SET 1 and SET 2.

3.5. Mechanical behavior

The effects of F-C compatibilization applied to the ternary blend were evaluated

qualitatively based on stress-strain curves. Elongation at break and toughness were found to

be correlated with interfacial adhesion, and provided information about resistance between

the compatibilized phases. On the other hand, tensile strength and yield stress were

correlated with the morphology and reduction of particle size in the dispersed phase domain.

Particle size reduction is expected to increase the blend’s tensile properties without changing

its composition [49].

Table 2 summarizes the mechanical properties shown in Fig. 11 of the blends. The

poor mechanical properties of the non-reactive blend were attributed to the weak

intermolecular interaction between HDPE and PP. Improvements in interfacial strength and

mechanical strength were achieved with a of 0.3 wt% (improved morphology). Elongation at
break of the blend with 0.3 wt% catalyst content increased by 50% in relation to the non-

reactive blend, followed by a decrease when the catalyst concentration was increased. The

decrease in mechanical properties in response to the increase in aluminum chloride

concentration indicates the competing effect of F-C compatibilization, i.e., copolymer

formation and catalytic degradation. Diaz [28, 29] reported similar results for a PP/PS blend

and opposite results for a PE/PS blend.

Table 2. Mechanical properties evaluated to correlate Friedel-Crafts compatibilization.


Fig. 11. Variation of mechanical properties in the tensile test as a function of catalyst content.

The samples containing 0.3 wt% of catalyst required twice as much energy to break as

the non-reactive blends, indicating improved toughness, and hence, greater compatibility.
This required energy was proportionally higher for S1-03 because of the higher HDPE content

in its composition. This finding is congruent with the presence of a compatibilizing agent. The

magnitude of the decrease of this property with the increase in catalyst content was similar to

that reported by Diaz [28]. The 50% improvement in the tensile strength of SET 1 and of 30%

in SET 2, and the 25% gain in yield stress of both sets indicate a better distribution and

reduction of the dispersed PS domain. Reactive blends are more favorable for load

distribution, given their larger surface area to support stresses. This parameter can be

analyzed based on the morphology.

The HDPE/PS and PP/PS interphases of the ternary blend underwent F-C

compatibilization. However, its effect on the mechanical properties of the former was more

marked than on the latter, which was attributed to the increase in catalytic degradation with

increasing PP concentration.

Fig. 12 compares the stress-strain curves of neat PS, HDPE, and PP with the reactive

blends S1-03 and S2-03. Each curve represents the average behavior of the 10 samples in

the tensile test performed under the same conditions. The reactive blends showed similar

mechanical behavior to that of neat PP but a higher Young’s modulus, indicating that the

effects of the compatibilization reaction outweighed the catalytic degradation. The reactive

blends showed higher yield stress than the neat homopolymers and their Young’s modulus

was similar to that of neat PP.


Fig. 12. Comparison of stress-strain curves of neat homopolymers and reactive blends (S1-

03 and S2-03).

4. CONCLUSIONS

It can be stated that in situ compatibilization of the PE/PP/PS ternary blend by Friedel-

Crafts alkylation improved interfacial adhesion, decreased interfacial stresses and the size of

PS dispersed phase, promoted adhesion and stabilized the morphology. As for the extracted

material, the DSC and FTIR analysis suggested that, during the compatibilization process, PP

segments were more readily coupled than HDPE chains. However, the latter are more

representative due to their high molecular mass. NMR evidenced the F-C alkylation reaction

through capture of p-substitution in the PS aromatic ring. The 1H NMR spectra proved to be

sufficiently sensitive to detect grafting onto PS chains. In the ternary blend, a balance was
achieved between complete and partial wetting morphology with the migration of dispersed

domains from one polyolefin phase to another as a result of F-C compatibilization. The

optimal emulsifying effect was achieved with 0.3% of Lewis acid. The dispersed PS droplets

at the HDPE/PP interface of the reactive blends exhibited two kinds of behavior: close

contact, in which there are copolymers at the interface to anchor polyolefins, and broad

contact, without any compatibilizing agent. The mechanical behavior of the blends indicated

that F-C compatibilization led to morphological dispersion and improved interfacial properties.

Acknowledgements

The authors thank the companies Braskem, Innova, and Harima for their donation of

the materials used in this work. We also acknowledge the Brazilian research funding agencies

CAPES (Federal Agency for the Support and Improvement of Higher Education), CNPq

(National Council for Scientific and Technological Development), and Fundação Araucária for

their financial support.

Data Availability

The raw/processed data required to reproduce these findings cannot be shared at this

time due to technical and time limitations.


References

1. Mondal, M.K.; Bose, B.P; Bansal, P. J Envirom Manage. 2019, 240, 119-125.

2. Sidorov, O.F.; Shishov, M.G.; Deryugin, A.A.; Sidelnikov, A.Y. Coke Chem. 2016, 59(3),

117-121.

3. Gawande, A.; Zamare, G.; Renge, V.C.; Tayde, S.; Bharsakale, G. JERS. 2012, 3(2), 01-

05.

4. IBGE, Pesquisa Nacional de Sanemaneto Básico. https://www.ibge.gov.br (accessed

May 9, 2019).

5. Fornari, M. Saneamento Ambiental. 2013, 167, 10-15.

6. Manrich, S. Polym. Degr. Stab. 2002, 77, 441-447.

7. La Mantia, F.P; Morreale, M.; Botta, L.; Mistreatta, M.C.; Ceraulo, M.; Scaffaro, R. 2017,

145, 79-92.

8. Fortelny, I.; Michalkova, D.; Krulis, D. Polym. Degr. Stab. 2004, 85, 975-979.

9. Koning, C.; Van Duin, M.; Pagnoulle, C.; Jerome, R. Progr. Polym. Sci. 1998, 23, 707.

10. Spinacé, M. A. de S.; Paoli, M. A. Quím. Nova 2005, 28, 65.

11. Kallel, T.; Massardier-Nageotte, V.; Jaziri, M.; Gérard, J.-F.; Elleuch, B. J. Appl. Polym.

Sci. 2003, 90, 2475.

12. Feldman, D. J. Macromol. Sci. A: Pure Appl. Chem. 2005, 42, 587.

13. Ajji, A.; Utracki, L. A. Polym. Eng. Sci. 1996, 36, 1574.

14. Przybysz, M.; Marć, M.; Klein, M.; Saeb, M. R.; Formela, K. Polym. Test. 2018, 67, 513.

15. Jose, S.; Thomas, S.; Parameswaranpillai, J.; Aprem, A. S.; Karger-Kocsis, J. Polym.

Test. 2015, 44, 168.

16. Karl Fink, J. In Reactive Polymers: Fundamentals and Applications; Elsevier Inc., Austria,

2018, Chapter 16, pp 497-546.

17. Tanrattanakul, V.; Sungthong, N.; Raksa, P. Polym. Test. 2008, 27, 794.
18. Brown, S. B. In Polymer Blends Handbook; Utracki, L. A., Ed.; Kluwer Academic

Publishers, London, 2002, Vol. 1, Chapter 5, pp 339-414.

19. Yang, J.; White, James L. In: Polyolefin Blends; Nwabunma, D.; Kyu, T., Eds.; Wiley-

Interscience: New Jersey, 2008, Part I, Chapter 2, pp 27-51.

20. Carrick, W. L. J. Polym. Sci. A-1: Polym. Chem. 1970, 8, 215.

21. Barentsen, W. M.; Heikens, D. Polymer. 1973, 14, 579.

22. Heikens, D.; Barentsen, W. Polymer. 1977, 18, 69.

23. Sun, Y.-J.; Baker, W. E. J. Appl. Polym. Sci. 1997, 65, 1385.

24. Sun, Y.-J.; Willemse, R. J. G.; Liu, T. M.; Baker, W. E. Polymer. 1998, 39, 2201.

25. Gao, Y.; Huang, H.; Yao, Z.; Shi, D.; Ke, Z.; Yin, J. J. Polym. Sci. B: Polym. Phys. 2003,

41, 1837.

26. Diaz, M. F.; Barbosa, S. E.; Capiati, N. J. Polymer. 2002, 43, 4851.

27. Diaz, M. F.; Barbosa, S. E.; Capiati, N. J. J. Polym. Sci. B: Polym. Phys. 2004, 42, 452.

28. Diaz, M. F.; Barbosa, S. E.; Capiati, N. J. Polymer. 2005, 46, 6096.

29. Diaz, M. F.; Barbosa, S. E.; Capiati, N. J. Polym. Eng. Sci. 2006, 46, 329.

30. Diaz, M. F.; Barbosa, S. E.; Capiati, N. J. Polymer. 2007, 48, 1058.

31. Diaz, M. F.; Barbosa, S. E.; Capiati, N. J. J. Appl. Polym. Sci. 2009, 114, 3081.

32. Shahbazi, K.; Razavi, M. K.; Abbasi, F.; Meran, M. P.; Mazidi, M. M. Polym. Bull. 2012, 69,

241.

33. Solomons, T. W. G.; Fryhle, C. B. In Organic Chemistry; LTC: São Paulo, 2000.

34. Li, J.; Ma, G.; Sheng, J. J. Polym. Sci. B: Polym. Phys. 2010, 48, 1349.

35. Li, J.; Ma, G.; Sheng, J. J. Macromol. Sci. B: Phys. 2009, 48, 2009.

36. Helfand, E.; Tagami, Y. J. Polym. Sci. B: Polym. Phys. 1996, 34, 1947.

37. Guo, Z.; Fang, Z.; Tong, L.; Xu, Z. Polym. Degrad. Stab. 2007, 92(4), 545-551.

38. Pozzi, P.; Marini, M. Polym. Test. 2008, 27, 161.


39. Cheng, H. N.; Lee, G. H. Polym. Bull. 1985, 13, 549.

40. Bevington, J. C.; Huckerby, T. N. Eur. Polym. J. 2006, 42, 1433.

41. Takeichi, T.; Thongpradith, S.; Kawauchi, T. Molecules 2015, 20, 6488.

42. Mustafa, S. J. S.; Azlan, M. R. N.; Fuad, M. Y. A.; Ishak, Z. A. M.; Ishiaku, U. S. J. Appl.

Polym. Sci. 2001, 82, 428.

43. Shapiro, B. L.; Mohrmann, L. E. J. Phys. Chem. Ref. Data. 1977, 3, 919-989.

44. Favis, B. D.; Reignier, J. Macromol. 2000, 33, 6998.

45. Sepehr, R.; Favis, B. D. Polymer. 2010, 51, 4547.

46. Le Correller, P.; Favis, B. D. Polymer. 2011, 52, 3827.

47. Virgilio, N.; Desjardins, P.; L'Espérance, G.; Favis, B. D. Polymer. 2010, 51, 1472.

48. Wang, H.; Zhang, X.; Nie, H.; Wang, R.; He, A. Compos Part A Appl Sci Manuf. 2019,

116, 197-205.

49. Xavier, S. F. "Properties and performance of polymer blends," In Polymer Blends

Handbook; Utracki, L. A., Kluwer Academic Publishers: London, 2002; Vol. 1, Chapter 12,

pp 861-942.

You might also like