Skip to main content

A study of the controlled degradation of polypropylene containing pro-oxidant agents

Abstract

Intentional degradation by pro-oxidant agents, many of which are metal-based, can result in uncertainty as to the time of biodegradation. Polyacetal (POM) is a thermoplastic polymer commercially classified as an engineering polymer and contains carbon, hydrogen and oxygen. The depolymerization of POM during processing can enhance thermal decomposition. The aim of this study was to investigate the controlled degradation of polypropylene induced by the degradation of POM or d2w®. Mixtures of polypropylene containing different concentrations of POM or d2w® were prepared by extrusion. The properties of the mixtures (blends) were evaluated based on the melt index (MFI), tensile properties, Fourier transform infrared spectroscopy (FTIR), Time inductive oxidation (OIT) and Thermogravimetric analysis (TGA). The two additives (POM and d2w®) enhanced the oxidative thermal degradation of polypropylene and the degradation of the polypropylene/POM mixture could be controlled by altering the POM concentration.

Introduction

Packaging waste accounted for 78.81 million tons or 31.6% of municipal solid waste (MSW) in the United States in 2003, 56.3 million tons or 25% of MSW in Europe in 2005, and 3.3 million tons or 10% of MSW in Australia in 2004. In the US, the predominant method of waste disposal is currently landfill packaging, followed by recycling, composting and incineration (Kale et al. 2007). Commodity polymers (polyethylene PE, polypropylene PP, polystyrene PS, polyvinyl chloride PVC and polyethylene terephtalate PET) prevail in packaging applications (PlasticsEurope 2011) and polyolefins are increasingly being used in new applications (Gahleitner 2011). An excellent way of producing degradable polyethylene is to mix this polymer with pro-oxidant additives that can effectively improve the degradability of these materials (Roy et al. 2007). Intentional degradation by pro-oxidant agents, many of which are metal-based (Roy et al. 2007), has generated uncertainties in the evaluation of biodegradation (European Bioplastics 2012) and several surveys it is claimed that polyolefins (PE, PP) is an inert polymer with good resistance to microorganisms (Albertsson 1978 2003). The controlled degradation of polypropylene has been used in rheological control by distributing and reducing the molar mass of organic peroxides in reactive extrusion (Rocha et al. 1994; Kim 1996). Polyacetal (POM) is a thermoplastic polymer that is susceptible to thermal decomposition (depolymerization) (Cottin et al. 2000). The objective of this study was to investigate the controlled degradation of polypropylene induced by the degradation of an organic oxidizing agent (POM) in extrusion. The additive d2w®, a commercial metal-based pro-oxidant, was used for comparison.

Materials and methods

Materials

Isotactic polypropylene (iPP) H603 (density: 0.905 g/cm3; MFI: 1.5 g/10 min) was used in granulated form, as supplied by Braskem (Triunfo, RS, Brazil). The polyacetal copolymer (density: 1.42 g/cm3; MFI: 14.0 g/10 min) was used in powder form as supplied by Ticona (São Paulo, SP, Brazil). The commercial pro-oxidant additive d2w® was supplied by RES Brazil (São Paulo, SP, Brazil).

Methods

Preparation of the mixtures

The additive d2w® was incorporated in granular form and POM in powder form. The incorporation of POM or d2w® into polypropylene (PP) initially involved homogenization in Drais with a load capacity of ~100 g of material. The mixing time was ~30 s. Blends of PP with POM or d2w® were prepared as shown in Table 1 using a single screw extruder fitted with a 25 mm diameter screw, a heating cylinder with an L/D ratio of 25:1 and four wire screens in series (60, 150, 150 and 100 mesh) to maximize the homogeneity. The extrusion conditions were 220°C, 250°C and 250°C for the first, second and third zones, respectively.

Table 1 The polymer/additive mixtures used in this study

Analysis

Melt flow index (MFI)

The MFI was determined in a plastometer (model 7023.000, CEAST, Ohio, USA) according to ASTM D-1238 (ASTM 2004). The test conditions were set at a load of 2,160 kg and a temperature of 230°C for all mixtures.

Mechanical tests

Type IV specimens (ASTM D-638-10) (ASTM 2010) were injected into a model PIC-BOY 22 machine (Petersen & Cia Ltda, São Paulo, SP, Brazil) with an injection capacity of 22 g of polystyrene. The total cycle time was 30 s and the temperatures of zones 1 (injection nozzle), 2 and 3 were 220°C, 220°C and 180°C, respectively. The tensile test was done in a universal testing machine (model 5569, Instron), according to ASTM D638-10, at a test speed of 25 mm/min and cell load of 50 kN. The tensile strength at break and elastic modulus were determined.

Fourier transform infrared spectroscopy (FTIR)

Films 30 ± 2 μm thick were prepared at 190°C with a compression pressure of 2000 psi and compression time of 80 s. FTIR measurements were obtained using a Varian 660-IR FT-IR spectrometer operated in transmittance mode. Thirty-two scans were obtained in triplicate from 4000 cm-1 to 400 cm-1 at a resolution of 4 cm-1. The influence of POM and d2w® on polypropylene oxidation was determined from the spectra by calculating the carbonyl (CI) and hydroxyl (HI) indices based on the relationships CI = A1725/A2722 and HI = A3500/A2722, respectively.

Differential scanning calorimetry (DSC)

DSC was done in a TA Instruments calorimeter at a nitrogen flow of 50 ml/min. Approximately 10 mg of each sample was heated, cooled and heated again over a temperature range of 25–250°C at a heating and cooling rate of 10°C/min. The melting temperature (Tm), crystallization temperature (Tc) and degree of crystallinity were calculated using the enthalpy of fusion values of 209 J.g-1 and 306 J.g-1 for 100% crystalline polypropylene and polyacetal, respectively (Canevarolo 2003; Kumar et al. 1995).

Oxidation induction time (OIT)

The OIT was determined by exposing ~10 mg of each mixture to a nitrogen flow of 50 ml/min and a heating rate of 20°C/min. An oxygen flow of 50 ml/min was used after melting at 200°C.

Thermogravimetric analysis (TGA)

TGA was done in equipment from TA Instruments. Approximately 10 mg of each mixture was placed in a nitrogen atmosphere and heated at a rate of 10°C/min over a temperature range of 25–550°C. The nitrogen flow over the measurement cell was 50 ml/min. The activation energy of degradation (Ea) was determined according to ASTM E1641 (ASTM 2007).

Results and discussion

Melt flow index (MFI)

The MFI is inversely related to sample viscosity and can be used to estimate the interaction between the phases in polymer mixtures (Huang et al. 2003). Figure 1 shows that pure PP had the lowest MFI of all samples, indicating that it had the highest viscosity under the test conditions. The MFI increased as the content of POM in the mixtures increased (PP2, PP4, PP6 and PP8) and probably reflected the immiscibility between the phases (PP/POM) since the variation in the measurements was proportional to the mass of POM in the PP/POM mixture (blend) (Huang et al. 2003). In polypropylene mixtures PP3, PP5 and PP7 the flow behavior was reversed, i.e., the MFI decreased as the concentration of d2w® increased. Partial miscibility between the vehicle solution for the additive d2w® and the polypropylene matrix could explain this result, although the possibility of degradation of the matrix and dispersed phase to generate cross-linked additive carrier material (dispersed phase) and polypropylene (matrix) should also be considered (Huang et al. 2003; Waldman and De Paoli 1998; Bouhelal et al. 2010).

Figure 1
figure 1

MFI for pure materials (PP, POM, d 2 w®) and blends (PP/POM, PP/d 2 w®). The points are the mean ± SD of 6 determinations. b) Average results and their respective estimated standard deviation of the melt index of the compositions and pure materials.

Tensile testing

Figure 2 shows the tensile modulus of elasticity and flow of polypropylene and blends with d2w® or POM. Mixtures containing d2w® showed larger standard deviations than those containing POM. This difference probably reflected the degree of dispersion of the additive in the polypropylene matrix since d2w® was incorporated into polypropylene in granular form while POM was incorporated in powder form. This conclusion suggests interaction between the degradative processes associated with polypropylene and POM during sample preparation. The differences in the tensile strength of the two compositions (Figure 2a) most likely reflect variations in the inherent mechanical strength of the carrier material and the additive d2w®. Likewise, differences in the modulus of elasticity (Figure 2b) reflect variations in the stiffness of the additive d2w® and POM (Huang et al. 2003). The standard deviations for the modulus of elasticity of pure polypropylene and the PP8 mixture were also high. The presence of tiny, randomly distributed bubbles in these two preparations could have contributed to this high standard deviation; differences in the composition of the samples are unlikely to be a cause of this variation.

Figure 2
figure 2

a) Tensile strength and b) elasticity modulus of pure polypropylene (PP 1 ) and blends with POM and d 2 w®. The points are the mean ± SD of 6 determinations.

Infrared spectroscopy (FTIR)

Absorbance in the region of 1725 cm-1 and 3500 cm-1 indicated the presence of carbonyl and hydroxyl groups (De Paoli 2008; Cáceres and Canevarolo 2009), respectively, and the absorbance peak at 2722 cm-1 was related to angular molecular vibrations in CH and axial molecular vibrations in CH3, as suggested elsewhere (Cáceres and Canevarolo 2009; Babetto and Canevarolo 2002; Rabello and White 1997; Garton et al. 1978). Comparison of these absorbances can be used to normalize IR spectra since these peaks are insensitive to the oxidative degradation of pure polypropylene (Wang et al. 2011). Figure 3 and Figure 4 show variation in the carbonyl (CI) and hydroxyl (HI) indices, respectively. The CI index showed increased formation of carbonyls in mixture PP8, whereas PP3 and other mixtures containing d2w® (PP5 and PP7) were quite stable in this parameter. The HI index showed similar behavior to the CI index shown in Figure 3, but mixture PP8 had high hydroxyl formation, probably indicating greater susceptibility to the formation of nOH (acetic acid) (Duan et al. 2006) during processing, with absent or low oxygen supply during extrusion of the mixtures and injection of the samples. Together, these results indicate that the addition of d2w® did not enhance the degradation of polypropylene, whereas the presence of POM resulted in greater thermal degradation of polypropylene that was proportional to the content of POM and greater than the concentration of POM in mixture PP6 (3% by weight).

Figure 3
figure 3

Carbonyl indices of pure material (PP 1 ) and PP/POM and PP/d 2 w® blends.

Figure 4
figure 4

Hydroxyl indices of pure material (PP 1 ) and PP/POM and PP/d 2 w® blends before and after thermal aging.

Differential scanning calorimetry (DSC)

Table 2 shows the thermal properties of the pure materials (PP1, POM and d2w®) and mixtures (blends) (PP/d2w® and PP/POM). Figure 5 and Figure 6 show the crystalline melting temperatures (Tm) and crystallization temperatures (Tc), respectively. There was a trend towards a higher Tm for all mixtures containing POM and d2w®. In mixtures containing d2w® (PP3, PP5 and PP7) there was a reduction in Tc that was proportional to the d2w® concentration, such that the Tc was lower than for pure polypropylene. In PP/POM blends (PP2, PP4, PP6 and PP8) the Tc of PP2, PP4 and PP6 stabilized at ~5°C above the Tc of pure polypropylene. This is a desirable result in processing by injection because it allows a reduction in the cooling time. The Tc of the PP8 blend was lower than for the other mixtures and also in relation to pure POM. This finding suggests that morphological changes in POM are probably the result of its thermal degradation.

Table 2 Thermal properties of pure materials (PP 1 , POM and d 2 w®) and blends of PP/POM and PP/d 2
Figure 5
figure 5

Crystalline melting temperature (Tm) of pure materials (PP, POM and d 2 w®) and their blends (PP/POM and PP/d 2 w®).

Figure 6
figure 6

Crystallization temperature (Tc) of the pure materials (PP, POM and d 2 w®) and their blends (PP/POM and PP/d 2 w®).

Figure 7 and Figure 8 show the heat flux derived from the crystallization of PP/POM (PP2, PP4, PP6 and PP8) and PP/d2w® (PP3, PP5 and PP7). Mixtures containing POM showed additional peaks between the peaks of pure POM and polypropylene. Degradation reactions usually appear as endothermic processes and exothermic behavior is generally a response to depolymerization (Canevarolo 2003; De Paoli 2008). Indeed, an endothermic response was observed in the curve of the derivative of heat flow during the crystallization of pure POM (Figure 7). The coexistence of these two processes in PP/POM mixtures suggests a change in the Tc and the formation of volatiles that are not detected by thermogravimetric analysis at temperatures below 200°C. During the crystallization of mixtures of PP/d2w® there is no heat flow between the peaks of the pure materials (PP and d2w®).

Figure 7
figure 7

Derivative of heat flow in the crystallization of pure materials (PP 1 and POM) and PP/POM blends (PP 2 , PP 4 , PP 6 and PP 8 ) without thermal aging.

Figure 8
figure 8

Derivative of heat flow in the crystallization of pure materials (PP 1 and d 2 w®) and blends of PP/d 2 w® (PP 3 , PP 5 and PP 7 ) without thermal aging.

Oxidation induction time (OIT)

The oxidation induction time (OIT) is an accelerated aging test that allows comparison of the relative resistance of materials to thermal oxidation. Table 3 and Figure 9 provide the OIT values for the pure materials and blends studied. The additive d2w® accelerated the oxidation of polypropylene in the presence of oxygen. However, there was little change in the OIT values of blends containing different amounts of d2w® (PP3, PP5 and PP7), i.e., the OIT values essentially reflected the amount of pure additive (d2w®) present in the mixtures. This finding indicates that there is little margin for controlling the thermo-oxidation of polypropylene during processing since all samples were processed under the same condition. In addition, changes in morphology arising from processing interfere with the diffusion of volatile degradation products.

Table 3 Oxidation induction time for pure materials and blends
Figure 9
figure 9

Oxidation induction time for pure materials and blends.

In PP/POM blends, there was a marked decrease in the OIT values from blend PP6 onwards. In contrast, there was an increase in the OIT values of blends PP2 and PP4, i.e., a stabilizing (antioxidant) effect. A similar delay in the kinetics of degradation was observed in the absence of oxygen in the TGA of these two blends, i.e., the Ti of the blends was greater than that of pure polypropylene (PP1). The OIT and TGA results indicated that d2w® concentrations ≥2% increased the thermal stability of the blends. In the case of POM, there was a decrease in the stabilizing synergistic effect at concentrations up to 3%; at higher concentrations, POM had an oxidizing effect on polypropylene.

The oxidation of a polymer involves a complex chain of reactions that involves many steps such that the overall Ea is the sum of the energies of activation of individual stages. In this chain of reactions there may be temperature ranges in which deviations from Arrhenius’ law can be neglected, e.g., with blends PP2 and PP4. The oxidation of polypropylene (in powder form) has been referred to as non-homogeneous (heterogeneous) kinetics that is characterized by chemiluminescence (Celina and George 1995). This oxidation is based on a model containing small numbers of localized zones (amorphous regions) in which oxidation occurs at a high rate and from where it spreads to other regions. The presence of stabilizers retards the diffusion of volatile degradation products for a short period of time known as the induction period. Even using sensitive techniques involving photon emission, such as chemiluminescence, the investigation of this phenomenon over such a short timescale is a difficult task, even though the Ea is higher in this period (Celina and George 1993). Several studies (Wang et al. 2011; Bouhelal et al. 2010; Groening and Hakkarainen 2002; Albertsson and Hakkarainen 2008) have shown that the decomposition of hydroperoxides in polypropylene leads to the formation of volatile products and that water is a major product of degradation but does not interfere with the spread of oxidation. During this period, generally only a decrease in polymer molecular mass is observed, along with the formation of volatile, low molecular mass products. Eriksson (Eriksson et al. 2001) suggested that following the formation of peracids by the oxidation of formaldehyde, the spreading of oxidation is favored by the gas phase and that the relatively low reactivity of formaldehyde allows greater diffusion to more distant regions.

Thermogravimetric analysis (TGA)

Figure 10 shows the temperature for the onset of degradation (Ti) of the pure materials and blends. The variation in the Ti of blends compared to pure polypropylene may reflect the formation of volatile components and their diffusion in the polymeric matrix. The pure materials (POM and d2w®) had a lower Ti than polypropylene, with POM showing the lowest value. The Ti of the blends were not lower than those of POM and d2w®, except for blend PP8. The values of Ti varied with the concentration of d2w® and POM in polypropylene. In the case of d2w®, increasing the concentration of the additive led to an increase in Ti and therefore had a stabilizing effect. This enhanced response can be attributed to the greater thermal stability of the vehicle for d2w® relative to polypropylene. The MFI corroborated this stabilizing effect. Increasing the concentration of POM led to an increase in Ti up to the value seen with blend PP4, after which the Ti stabilized up to PP6 and then decreased in blend PP8.

Figure 10
figure 10

Onset degradation temperature (Ti) determined at a rate of 10°C/min for pure materials and blends.

Together, these results indicate that the addition of d2w® to polypropylene increased the thermal stability in a manner dependent on the concentration of additive. The low production of volatile components and the reduced mobility (diffusion capacity) of the polypropylene matrix were probably important factors in this thermal stability. The addition of POM to polypropylene resulted in synergistic and antagonistic effects (stabilization and degradation), with the extent of stabilization and degradation depending on the amount of volatile components produced.

Figure 11 shows the change in mass as a function of temperature for blend PP8 and for the pure materials (PP1 and POM); a theoretical curve calculated curve calculated based on the weighted average of the experimental curves of the pure materials (PP and POM) is also shown. The region of the curve for blend PP8 close to the experimental curve of pure polypropylene (PP1) may reflect a greater stabilizing effect of polypropylene. The stabilizing effect decreases as the loss of mass increases and the theoretical curve approximates the observed curve, probably because of competition with degradation reactions. The region beyond the point where the theoretical and observed curves cross corresponds to additional degradation by shifting the equilibrium of the degradation reactions of the matrix phase (PP) induced by the dispersed phase (POM).

Figure 11
figure 11

Experimental curves for changes in mass versus time for blend PP8 and pure materials (PP and POM). The theoretical (calculated) for blend PP8 is also shown.

Table 4 and Figure 12 show the energy of activation (Ea) of the pure materials and blends. Mixtures containing d2w® and POM showed opposite changes in Ea as a function of the additive concentration in the polypropylene matrix. Ea may be regarded as the energy needed to cause the diffusion of one mole of atoms such that high Ea results in a relatively small diffusion coefficient (Callister 2006). In the blends studied here, the low mobility of pro-oxidant derivatives of d2w® attached to a polymeric carrier material may have increased the concentration of additive thereby enhancing the Ea. For blends of PP/POM, the formation of volatile, low molecular mass products such as formaldehyde derived from POM could explain the decrease in Ea with increasing POM concentration. The propagation of oxidation at high temperatures in stabilized polypropylene involves the gas phase (Eriksson et al. 2002; Eriksson et al. 2001; Celina et al. 2006). Among the products generated by the degradation of PP, e.g., water, ethylene, isobutylene and acetic acid, formaldehyde is the one that most likely contributes to the propagation of oxidation via the gas phase in stabilized polypropylene (Eriksson et al. 2002). This conclusion agrees with the calculated curve for blend PP8 shown in Figure 10 that suggests a synergistic effect of degradation. The low Ea value for blend PP8 (lower than pure POM) is also suggestive of synergism.

Table 4 Activation energy (Ea) and correlation coefficients for pure materials and blends
Figure 12
figure 12

Energy of activation (Ea) for pure materials (PP, POM and d 2 w®) and blends of PP/POM and PP/d 2 w®.

Conclusion

The addition of POM or d2w® promoted the oxidative thermal degradation of polypropylene (PP), with the extent of degradation being regulated by the POM concentration in PP/POM blends. At concentrations <3% (w/w), POM enhanced the thermal stabilization of polypropylene under the conditions investigated, whereas at concentrations >3% POM stimulated the oxidation of polypropylene. These results suggest that the POM with a concentration >3% (w/w), may act as a pro-oxidant agent of the PP, and the synergistic effect of degradation can be maximized by increasing the miscibility at the interface of the blend PP/POM.

References

  • Albertsson AC: Biodegradation of synthetic polymers. II. A limited microbial conversion of 14C in polyethylene to 14CO2 by some soil fungi. J Appl Polym Sci 1978, 22: 3419-3433. 10.1002/app.1978.070221207

    Article  Google Scholar 

  • Albertsson I: Evaluation of degradability of biodegradable polyethylene (PE). Polym Degrad Stabil 2003, 80: 39-43. 10.1016/S0141-3910(02)00380-4

    Article  Google Scholar 

  • Gröning M, Hakkarainen G, Albertsson A-C: Quantitative Determination of Volatiles in Polymers and Quality Control of Recycled Materials by Static Headspace Techniques. In Chromatography for Sustainable Polymeric Materials Renewable, Degradable and Recyclable, Vol (Eds). Edited by: Albertsson A-C, Hakkarainen M. Berlin Heidelberg: Springer; 2008:51-84.

    Chapter  Google Scholar 

  • ASTM D-1238-04: Standard Test Method for Melt Flow Rates of Thermoplastics by Extrusion Plastometer, Annual Book of ASTM. Philadelphia, PA: American Society for Testing and Materials; 2004.

    Google Scholar 

  • ASTM D638-10: Standard test method for tensile properties of plastics. Philadelphia, PA: American Society for Testing and Materials; 2010.

    Google Scholar 

  • ASTM E1641 – 07: Standard Test Method for Decomposition Kinetics by Thermogravimetry. Philadelphia, PA: American Society for Testing and Materials; 2007.

    Google Scholar 

  • Babetto AC, Canevarolo SV: Effect of the screw element from the degradation of polypropylene during multiple extrusions. Adv Polym Tech 2002, 21: 243-249. 10.1002/adv.10028

    Article  Google Scholar 

  • Bouhelal SM, Cagiao ME, Bartolotta A, Marco GD, Garrido L, Benachour D, Calleja FJB: On Polyethylene Chain Generation Through Chemical Crosslinking of Isotactic Polypropylene. J Appl Polym Sci 2010, 116: 394-403.

    Google Scholar 

  • Cáceres CA, Canevarolo SV: Degradation of polypropylene during the extrusion and generation of volatile organic compounds. Polímeros: Ciência e Tecnologia 2009, 19: 79-84.

    Article  Google Scholar 

  • Canevarolo SVJ (Ed): Techniques for polymer characterization. Ed Artliber, Brasil: São Paulo; 2003.

    Google Scholar 

  • Callister WD Jr: Materials Science and Engineering: An Introduction. 2nd edition. Rio de Janeiro: LTC; 2006:125.19.

    Google Scholar 

  • Celina M, George GA: A heterogeneous model for the thermal oxidation of solid polypropylene from chemiluminescence analysis. Polym Degrad Stabil 1993, 40: 323-335. 10.1016/0141-3910(93)90138-9

    Article  Google Scholar 

  • Celina M, George GA: Heterogeneous and homogeneous kinetic analyses of the thermal oxidation of polypropylene. Polym Degrad Stabil 1995, 50: 89-99. 10.1016/0141-3910(95)00136-A

    Article  Google Scholar 

  • Celina M, Clough RL, Jones GD: Initiation of polymer degradation via transfer of infectious species. Polym Degrad Stabil 2006, 91: 1036-1044. 10.1016/j.polymdegradstab.2005.07.022

    Article  Google Scholar 

  • Cottin H, Gazeau MC, Doussin JF, Raulin F: An experimental study of the photodegradation of polyoxymethylene at 122, 147 and 193 nm. J Photoch Photobio A 2000, 135: 53-64. 10.1016/S1010-6030(00)00274-4

    Article  Google Scholar 

  • De Paoli MA: Degradation and Stabilization of Polymers. Brasil: Chemkeys; 2008.

    Google Scholar 

  • Duan Y, Li H, Ye L, Liu X: Study on the Thermal Degradation of Polyoxymethylene Thermogravimetry-Fourier transform infrared spectroscopy (TG-FTIR). J Appl Polym Sci 2006, 99: 3085-3092. 10.1002/app.22913

    Article  Google Scholar 

  • Eriksson P, Reitberger T, Ahlblad G, Stenberg B: Oxidation fronts in the polypropylene studied by imaging chemiluminescence. Polym Degrad Stabil 2001, 1: 73:177.

    Google Scholar 

  • Eriksson P, Reitbergerb T, Stenberga B: Gas-phase Contribution to the spreading of oxidation in the polypropylene studied by imaging chemiluminescence. Polym Degrad Stabil 2002, 78: 183-189. 10.1016/S0141-3910(02)00132-5

    Article  Google Scholar 

  • European Bioplastics position paper: comments on the study. “A Life Cycle Assessment of Oxo-biodegradable, Compostable Bags and Conventional” (May 2012, Intertek). 2012. ,d.eWU&cad=rja. Accessed 15 Jan 2013 http://www.google.com.br/url?sa=t&rct=j&q=&esrc=s&source=web&cd=1&ved=0CCsQFjAA&url=http%3A%2F%2Fen.european-bioplastics.org%2Fwp-content%2Fuploads%2F2012%2F07%2FPositionspaper_LCA_160712.pdf&ei=ZANYUrK9DIWK9QSmnoG4Cw&usg=AFQjCNFGgdl3B6DDlGeKTUerJXsZW8SlUw&bvm=bv.53899372 ,d.eWU&cad=rja. Accessed 15 Jan 2013

    Google Scholar 

  • Gahleitner M: Editorial corner - a personal view. Polyolefins for the 21st century. Express. Polymer Letters 2011, 5: 936.

    Article  Google Scholar 

  • Garton A, Carlsson DJ, Wiles DM: Role of polymer morphology in the oxidation of polypropylene. J Polym Sci Pol Chem 1978, 16: 33-40. 10.1002/pol.1978.170160103

    Article  Google Scholar 

  • Groening M, Hakkarainen M: Headspace solid-phase microextraction with gas chromatography/mass spectrometry reveals a correlation between the degradation product pattern and changes in the mechanical properties during the thermooxidation of in-plant recycled polyamide 6.6. J Appl Polym Sci 2002, 86: 3396-3407. 10.1002/app.11345

    Article  Google Scholar 

  • Huang JM, Cheng HJ, WU JS, Chang FC: Blends of poly(propylene) and polyacetal compatibilized by ethylene vinyl alcohol copolymers. J Appl Polym Sci 2003, 89: 1471-1477. 10.1002/app.12350

    Article  Google Scholar 

  • Kale G, Kijchavengkul T, Auras R, Rubino M, Selke SE, Singh SP: Compostability of bioplastic packaging materials: an overview. Macromol Biosci 2007, 7: 255-277. 10.1002/mabi.200600168

    Article  Google Scholar 

  • Kim BK: Reactive Extrusion of Polyolefins and their Blends. Korea Polym J 1996, 4: 215-226.

    Google Scholar 

  • Kumar G, Neelakantan NR, Subramanian N: Polyacetal and toughened thermoplastic polyurethane polyacetal elastomer: crystallinity and fracture mechanics. J Mater Sci 1995, 30: 1480-1486. 10.1007/BF00375252

    Article  Google Scholar 

  • PlasticsEurope: An Analysis of European Plastics Production, demand and recovery for 2010. 2011. , accessed 19/03/2012 http://www.plasticseurope.org/Document/plastics---the-facts-2011.aspx

    Google Scholar 

  • Rabello MS, White JR: The role of physical structure and morphology in the photo degradation behaviour of polypropylene. Polym Degrad Stabil 1997, 56: 55-73. 10.1016/S0141-3910(96)00202-9

    Article  Google Scholar 

  • Rocha MCG, Coutinho FMB, Balke S: Controlled degradation of polypropylene. Polímeros: Ciência e Tecnologia 1994, 4: 16-22.

    Google Scholar 

  • Roy PK, Surekha P, Rajagopal C, Choudhary V: Thermal degradation studies of LDPE containing cobalt. Express Polymer Letters 2007, 1: 208-216. 10.3144/expresspolymlett.2007.32

    Article  Google Scholar 

  • Wang X, YU W, Nie Q, Guo Y, DU J: A real-time study on the evolution of the degradation of polypropylene during mixing process. J Appl Polym Sci 2011, 121: 1220-1243. 10.1002/app.33795

    Article  Google Scholar 

  • Waldman WR, De Paoli MA: Thermo-mechanical degradation of polypropylene, low-density polyethylene and their 1:1 blend. Polym Degrad Stabil 1998, 60: 301-308. 10.1016/S0141-3910(97)00083-9

    Article  Google Scholar 

Download references

Acknowledgements

The authors thank UFABC and CAPES for financial support and scholarships.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Celso Luis de Carvalho.

Additional information

Competing interests

The authors declare that they have no competing interests.

Authors’ contributions

CLC carried preparation of the mixtures, analysis and drafted the manuscript. AFS carried melt flow index, mechanical properties and discussion of results. DSR has been involved in the analysis and interpretation of data and supervised the project. All authors read and approved the final manuscript.

Authors’ original submitted files for images

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 2.0 International License (https://creativecommons.org/licenses/by/2.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

Reprints and permissions

About this article

Cite this article

de Carvalho, C.L., Silveira, A.F. & Rosa, D.d.S. A study of the controlled degradation of polypropylene containing pro-oxidant agents. SpringerPlus 2, 623 (2013). https://doi.org/10.1186/2193-1801-2-623

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/2193-1801-2-623

Keywords