Aditivos Polipropileno

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Polymer 181 (2019) 121749

Contents lists available at ScienceDirect

Polymer
journal homepage: www.elsevier.com/locate/polymer

Effect of thermo-oxidation on loss of plasticizers, on crystalline features and T


on properties in a metallocene isotactic polypropylene
Enrique Blázquez-Blázquez, Rosa Barranco-García, María L. Cerrada∗, Ernesto Pérez
Instituto de Ciencia y Tecnología de Polímeros (ICTP-CSIC), Juan de la Cierva 3, 28006, Madrid, Spain

H I GH L IG H T S

• Plasticizers loss is assessed under severe thermo-oxidative conditions in a metallocene iPP.


• Degradation time of remaining content depends strongly on the plasticizer incorporated.
• Alterations along the loss of plasticizer and antioxidant are observed in the iPP microstructure.
• Overall crystallization is considerably delayed and formation of orthorhombic crystallites is hindered.
• Frame of voids and crazes is developed on degradation and an increase in rigidity is observed.

A R T I C LE I N FO A B S T R A C T

Keywords: Incorporation of a low content of several plasticizers into a metallocene isotactic polypropylene (iPP) has been
Polypropylene performed in order to learn about their stability and performance by the action of a thermo-oxidative de-
Plasticizer gradation treatment at 95 °C for different times. Comparison with the behavior exhibited by the material with di
Thermo-oxidation (2-ethylhexyl) phthalate (DEHP) is established for a feasible replacement of its use for contributing to en-
α and γ polymorphs
vironmental and human health preservation. Dependence of these aggressive degrading conditions on plasticizer
Crystallization delay
loss is analyzed as well as elucidation of degradation influence on the iPP crystalline characteristics. The results
show that chemical changes associated with this degradation process hinder formation of the orthorhombic γ
crystalline lattice and postpone the overall crystallization capability. Accordingly, the final features in the film
surface and the mechanical response are significantly affected by degradation time. The TOTM plasticizer can be
considered an excellent candidate for DEHP substitution.

1. Introduction potential carcinogens [3,8]. Based on these findings, Directive 2005/


84/EC established restrictions on the use and marketing of certain
Plasticizers are indispensable in many applications. They are em- dangerous substances and on preparation of formulations (phthalates in
ployed as softeners for rigid materials since they lead to reduction of toys and childcare articles). Therefore, some phthalates have been
viscosity as well as increase of flexibility. They are not chemically classified since 2008 as reprotoxic substances of class 1B by the Eur-
bound to the plastic materials and, therefore, they can leach, migrate or opean Union due to their toxicity. Specifically, the use of DEHP in
evaporate into indoor air and atmosphere, foodstuff, liquids, or other medical devices for administration/removal of drugs or biological li-
materials. Consequently, they are present in the environment and can quids has been restricted by the European Directive 2007/47/CE. For
be incorporated into humans and animals through ingestion, inhalation, that reason, they have been proposed to be substituted by non-phtha-
and dermal contact [1–5] so that they are classified as environmental late compounds like acetyl tributyl citrate (ATBC), diisononyl 1,2-cy-
pollutants with a strong influence on the human health. clohexanedicarboxylic acid (DINCH), di(2-ethylhexyl) terephthalate
Di-butyl phthalate (DBP) and Di(2-ethylhexyl) phthalate (DEHP) (DEHT) or trioctyl trimellitate (TOTM) in many applications. These
have been the phthalic acid esters with the largest production volume non-phthalate substances have shown advantageous toxicological pro-
worldwide. In the last decades, the possible human health effects of files and the toxic effects on human health during their release from
phthalate plasticizers have been intensely discussed and some of those polymers to the environment are strongly reduced. Due to the ongoing
compounds have been classified as endocrine disruptors [3,6,7] and substitution process and considering new regulatory requirements,


Corresponding author.
E-mail address: [email protected] (M.L. Cerrada).

https://doi.org/10.1016/j.polymer.2019.121749
Received 2 July 2019; Received in revised form 13 August 2019; Accepted 26 August 2019
Available online 27 August 2019
0032-3861/ © 2019 Elsevier Ltd. All rights reserved.
E. Blázquez-Blázquez, et al. Polymer 181 (2019) 121749

Scheme 1. Chemical structures, boiling temperature (Tb) and molecular weight (MW) of the different plasticizers.

plasticizer substitutes have subsequently been added to the list of characteristics, the features on surface of the film and the mechanical
substances of interest. Several studies have been conducted to evaluate response. Distinct plasticizers are investigated to analyze which of them
the migration behavior in poly(vinyl chloride) (PVC) medical devices could be the best candidate for the replacement of phthalate com-
that include these plasticizers of new generation [9] and to analyze pounds and features of the resultant materials based on PP are ex-
their presence in different environments and in humans [10–14] in amined to learn about the feasible substitution of PVC by PP. The
order to replace the use of DEHP. The ATBC incorporation has been also thermo-oxidation at 95 °C has been chosen because the final materials
examined in other polymeric systems, like poly(lactic acid) and poly are frequently sterilized at high temperatures and during their shelf life
(hydroxybutyrate), among others [15–17]. can be heated by action of microwaves. Moreover, it provides an ac-
An additional concern is related to the use of PVC in applications as celerated approach to learn how these substances are lost and the
containers either for food or for medical products. Severe health risks consequences of such consumption. For that, maintenance of the ad-
can exist, mainly in the latest devices with prolonged interaction body ditive after the thermal treatment in the samples is estimated from
fluids (like disposable blood bags) or tissues, since PVC can slowly extraction by Soxhlet and analysis using gas chromatography-mass
liberate free chlorine, chlorine species and dioxins along its usage. spectroscopy (GC-MS); chemical changes involved in the iPP during the
Thus, its substitution would minimize exposure to potential toxicities. process are detected by Fourier transform infrared spectroscopy cou-
Accordingly, investigations have been promoted in manufacturing of pled with an attenuated total reflectance device (ATR-FTIR); variations
single-use packaging materials that not release toxic substances during in the crystalline phase transitions are assessed by differential scanning
their medical and food handling. Polypropylene (PP) is proposed as an calorimetry (DSC) and real-time variable-temperature Wide Angle X-
alternative to PVC, among other polymers. Suitable mechanical prop- ray Diffraction (WAXD) synchrotron experiments; features of the
erties, including flexibility, are an important issue in these applications roughness at the film surface are studied by optical profilometry; and
and incorporation of plasticizers is required. Evaluation of the effect of the mechanical characteristics are evaluated by microhardness mea-
DBP or DEHP on PP has been previously described in literature [7]. surements.
Nevertheless, unlike other polymers [10–17], there is, to our knowl-
edge, no exhaustive study of the addition into PP of these safer plasti-
cizers to simultaneously substitute the phthalate plasticizers and the 2. Experimental part
PVC in this contact application with food or in that medical-clinical
one. Therefore, the aim of this research is to proceed to the in- 2.1. Materials
corporation of a small amount of different less toxic plasticizers into an
isotactic metallocene polypropylene (iPP) to evaluate its loss as a A commercially available metallocene-catalyzed isotactic poly-
function of time by the action of thermo-oxidative treatment at 95 °C, propylene (Metocene HM562P: melt flow index of 15 g/10 min, ISO
and to comprehensively understand how degradation and plasticizer 1133, kindly supplied as pellets by Lyondell Basell) was selected in this
consumption affect the chemical details in the iPP, its crystalline work as polymeric matrix.
Different plasticizers of non-phthalate nature were used: ATBC

2
E. Blázquez-Blázquez, et al. Polymer 181 (2019) 121749

(acetyl tributyl citrate), CAS 77-90-7; DEHT (di(2-ethylhexyl) ter- 2.6. Fourier transform infrared spectroscopy (FTIR)
ephthalate, CAS 6422-86-2; DINCH (di-isononyl-cyclohexane-1,2-di-
carboxylate), CAS 166412-78-8; and, TOTM (trioctyl trimellitate), CAS The chemical changes induced within each polypropylene-plasti-
3319-31-1. Moreover, DEHP (diethylhexyl phthalate), CAS 117-81-7, cizer material during its degradation at 95 °C at the different times were
was also evaluated as a plasticizer of phthalic chemical origin for studied through Fourier transform infrared spectroscopy using a total
comparative reasons. Their chemical structures are detailed together attenuated reflectance device (FTIR-ATR). Spectra were recorded on a
with boiling temperature (Tb) and molecular weight (MW) in the PerkinElmer Spectrum Two spectrophotometer with a resolution of
Scheme 1. The 5CB ([1,1′-biphenyl]-4-carbonitrile, 4′-pentyl-), CAS 4 cm−1.
40817-08-1, was employed as internal reference in the GC-MS mea-
surements. Standards with purity equal or greater than 98.0% were
2.7. Size exclusion chromatography
purchased from Sigma–Aldrich (Spain), except for DINCH that was
acquired from Oxchem (USA).
The molecular weights of selected samples were evaluated by size
exclusion chromatography (SEC) at 145 °C in a Waters GPC/V 2000
2.2. Incorporation of additives and film preparation
equipment with both refractive index and viscosimetry detectors. A set
of three columns of the PL Gel type was used with 1,2,4-tri-
The different selected plasticizers were added to the homopolymer
chlorobenzene as solvent. The equipment was calibrated with poly-
in a percentage of 2% by weight. These blends (homopolymer plus
styrene standards of narrow molecular mass distributions. The average
additive) were obtained by melt extrusion in a co-rotating twin-screw
molecular weights and polydispersity index, PI, obtained are com-
microextruder Rondol with a length-to-diameter ratio 20:1. A screw
mented in the discussion.
temperature profile of 115, 170, 180, 185 and 190 °C was used from the
hopper to the die. Then, films with a thickness around 200 μm were
processed by compression molding at 190 °C and at 25 bar for 3 min in a 2.8. Differential scanning calorimetry (DSC)
hot-plate Collin press (200x200 model). A relatively fast cooling, at a
rate of around 80 °C/min, was applied from the melt to room tem- Calorimetric analyses were performed in a TA Instruments Q100
perature between plates under pressure (25 bar). This thermal treat- calorimeter connected to a cooling system and calibrated with different
ment is similar to those applied at industrial scale. Samples were la- standards. The sample weights ranged from 6 to 8 mg. A temperature
beled as: PP for the neat iPP; PP-ATBC, PP-DEHT, PP-DINCH, PP-TOTM interval from −65 to 200 °C was studied at a scanning rate of 20 °C/
and PP-DEHP for the materials based on iPP with ATBC, DEPT, DINCH, min. The first melting and the crystallization processes were evaluated
TOTM and DEHP plasticizer, respectively. in detail. For the determination of the crystallinity, a value of 160 J/g
[18,19] was used as the enthalpy of fusion of a perfectly crystalline
2.3. Thermogravimetric analysis material.

Dynamic thermogravimetric experiments (TGA) were performed 2.9. X-ray experiments with synchrotron radiation
from 40 °C up to 800 °C in a Q500 equipment of TA Instruments under
nitrogen atmosphere at a heating rate of 10 °C/min. Samples of ap- Real-time variable-temperature WAXD experiments were carried
proximately 10 mg were used. These measurements were carried out for out with synchrotron radiation in beamline BL11-NCD at ALBA
assessment of possible losses of plasticizers during processing, and also (Cerdanyola del Valles, Barcelona, Spain) at a fixed wavelength of
for analysis of their potential effect on the PP degradation. 0.1 nm. A Rayonix detector has been used at a distance about 19 cm
from sample and a tilt angle of around 30°. A Linkam Unit, connected to
2.4. Accelerated degradation treatment a cooling system of liquid nitrogen, was employed for the temperature
control. The calibration of spacings was obtained by means of silver
Thermo-oxidative experiments were performed in a convection behenate and Cr2O3 standards. The initial 2D X-ray images were con-
oven at 95 °C during different time periods: 0, 2, 4, 8, 10 and 16 days, in verted into 1D diffractograms, as function of the inverse scattering
films containing the distinct additives. For ATBC, two additional trials vector, s = 1/d = 2 sin θ/λ. Film samples of around 5x5x0.2 mm were
at 4 and 10 h were conducted to obtain more information in the initial used in the synchrotron analysis.
stages.

2.10. Nuclear magnetic resonance


2.5. Analysis of additives

Tacticity was determined by carbon nuclear magnetic resonance


Samples were extracted in a Soxhlet with dichloromethane for 8 h.
analysis, 13C NMR, from a polymeric solution in 1,1,2,2-tetra-
The extraction solution was concentrated in a rotary evaporator. The
chloroethane-d4 (70 mg/1 mL) at 100 °C, using a Bruker Avance III/500
obtained residue was transferred to a chromatographic vial and dried
spectrometer operating at 125.76 MHz. A minimum of 4000 scans were
with a nitrogen flow. Then, it was re-dissolved in a specific volume of
recorded with broad band proton decoupling, using an acquisition time
chloroform containing [1,1′-biphenyl]-4-carbonitrile, 4′-pentyl-, CAS
of 1.3 s and a pulse delay of 5 s.
40817-08-1, as internal standard in a known concentration.
Analytical determination was carried out using a Hewlett Packard
6890 HRGC gas chromatograph equipped with an Agilent Technologies 2.11. Optical profilometer
mass spectrometry detector model 5973. The separation of the com-
pounds was performed on a DB5-HT capillary column (15 m × 250 μm Surface characteristics of the films, in terms of changes in roughness
and 0.1 μm). The carrier gas used was helium with a flow rate of 1 ml/ associated with the loss of plasticizers, were measured in a Z-20 True
min. The electronic impact (70 eV) was the selected type of ionization color 3D Optical Profiler (Zeta Instruments).
for the mass spectrometer. This equipment allows measurements of features in height or sur-
The chosen chromatographic method lasted 37.5 min. The program face roughness. Average Surface Roughness (Sa) was measured, fol-
started at 80 °C, temperature was increased at a constant rate of 8 °C/ lowing equation (1), as the arithmetic average of the absolute values of
min up to 340 °C, and was maintained for 5 min at that high tempera- profile height deviations recorded within the evaluation area and
ture. measured from the mean surface area, which is a horizontal plane [20]:

3
E. Blázquez-Blázquez, et al. Polymer 181 (2019) 121749

1 Ly Lx
Sa =
Ae
∫0 ∫0 z (x , y ) dxdy
(1)

2.12. Microhardness measurements

A Vickers indentor attached to a Leitz microhardness tester was used


to perform microindentation measurements at 23°C. A contact load of
0.98 N for a time of 25 s was employed. Microhardness, MH, values (in
MPa) were calculated according to the relationship [21]:
P
MH = 2 sin 68° ⎛ 2 ⎞
⎝d ⎠ (2)

where P (in N) is the contact load and d (in mm) is the diagonal length
of the projected indentation area. Diagonals were measured in the re-
flected light mode within 30 s of load removal, using a digital eyepiece
equipped with a Leitz computer-counter-printer (RZA-DO).

3. Results and discussion

Soxhlet extraction was initially performed on films processed from


Fig. 2. Dependence with degradation time at 95 °C of the reduction in the
the pristine iPP homopolymer. The GC of this extract points out, as plasticizer content for the different iPP based materials analyzed.
depicted in the upper plot of Fig. 1, the presence of Irgafos 168 as
additive (in a proportion of 290 ppm) and its oxidized form (Irg168 ox)
in relation to the extract from iPP because dilution for the chromato-
at slightly higher retention time. The Irgafos 168 is a common anti-
grams containing plasticizers is superior to that used for the iPP in order
oxidant agent used in the iPP formulations. Other regularly added an-
to avoid saturation.
tioxidant substances, such as Irganox 1330 or Irganox 1010, are not
Fig. 2 represents the dependence with degradation time of re-
present in this commercial polypropylene. An HPLC test was performed
maining content for each plasticizer in the distinct iPP based films
in this neat iPP and absence of other antioxidant compounds was
during their stay at 95 °C. Important differences are observed between
confirmed. The iPP selected corresponds to a commercial grade de-
them. The PP-ATBC material is the one that shows the fastest plasticizer
signed to present optimal transparency and clarity during its useful life.
loss rate. In fact, the ATBC amount is drastically reduced in only 10 h.
Consequently, phenolic antioxidants were avoided due to their un-
For this reason, additional degradation times of 4 and 10 h were also
desirable effect of yellowing on the final material.
used in this case. Detection content below 1% is seen in the first 48 h
Fig. 1 also shows in the lower plot the chromatograms for the ex-
and it is completely lost before 8 testing days.
tracts achieved from the distinct iPP-plasticizer materials. The internal
The material that incorporates the plasticizer with phthalate che-
reference used (I.S.) is clearly noticeable in all of them together with
mical nature, the PP-DEHP, whose replacement by other less toxic ad-
the extracting compounds characteristic for each individual plasticizer.
ditive is desired, is the one showing the next high leakage behavior. Its
Their retention times are rather different between them. Other re-
content is reduced above 50% in only 4 days with respect to the initial
markable aspect deduced from Fig. 1 is the multiple-peak pattern ob-
one.
served in the PP-DINCH extract. This is because DINCH plasticizer is
In contrast to the PP-DEHP, the PP-DEHT material, which adds the
composed by a mixture of isomers [22].
plasticizer with the non-phthalic isomeric structure, shows a consider-
Irgafos 168 and its oxidized form are also seen in these chromato-
grams (obviously with a much smaller intensity than the plasticizers, able greater capability of remaining over time at 95 °C within the PP-
DEHT films. This feature constitutes a good starting point because of its
due to their lower concentration). They appear significantly minimized
much better toxicological profile.
An intermediate behavior between that shown by the PP-DEHP and
PP-DEHT is exhibited in the PP-DINCH material, as seen in Fig. 2. After
8 days of permanence at 95 °C, the PP-DINCH film displays a loss of
about 55% in the DINCH content.
Fig. 2 also clearly displays that PP-TOTM material is the one where
plasticizer remains in a very significant content at this high temperature
for 16 days. Difference with the PP-DEHT, which is the next in pre-
servation of plasticizer, is rather important: a content of around 88% of
TOTM is kept versus approximately 22% of DEHT. Thus, loss of plas-
ticizer in the PP-TOTM occurs at a rate very low. Then, similar behavior
is expected in terms of structure and properties in this PP-TOTM ma-
terial compared to that shown by the pristine non-plasticized iPP
homopolymer. If this assumption is met, the high presence of TOTM
under this exposure under demanding conditions seems to indicate that
it is the best plasticizer option to be incorporated into the iPP.
These results are consistent with those described by Bernard et al.
[9] in an earlier study in which migration was analyzed in PVC speci-
mens with DEHP, DINCH, DEHT and TOTM plasticizers immersed at
Fig. 1. Comparative chromatograms for the different extracts obtained from the 40 °C in a simulant, consisting in an ethanol/water (50/50 v/v) mix-
initial materials under analysis. From top to bottom: pristine PP, PP-ATBC, PP- ture. A similar migration capacity for DEHP and DINCH was observed,
DEHP, PP-DEHT, PP-DINCH and PP-TOTM. while the TOTM migrated 20 times less. They proposed DEHT or TOTM

4
E. Blázquez-Blázquez, et al. Polymer 181 (2019) 121749

materials (upper, amplification of Fig. 3). It can be observed in the


upper plot that the amount of lost plasticizer is close to 2%, which is the
original value added to the compounded iPP samples, except in the
cases of PP-ATBC and PP-DEHP, where that value is somewhat smaller
(close to 1.6%). Importantly, these weight losses profiles match rather
perfectly those of the pure plasticizers, as shown by the guide lines in
Fig. 4. On the other hand, TOTM, which presents the chemical structure
with higher molecular weight, exhibits a degradation pattern at sig-
nificantly higher temperature than the other plasticizers.
These TGA degradation profiles can explain, at least partially, the
reduction in the plasticizer content for the different iPP based materials
observed in Fig. 2, which appears to be mostly due to its volatility.
Evidently, possible specific interactions between iPP and the plasticizer
molecules could have an additional effect.
In order to keep unchanging as much as possible the structure and
properties of iPP along the thermo-oxidation process the preservation of
the content of antioxidant additive must be of capital importance. Its
loss under the experimental conditions might boost the beginning of the
iPP degradation, involving a rapid reduction in properties. Therefore,
variation of the content of Irgafos 168 within the annealed films has
Fig. 3. TGA curves for the different samples. The temperature interval from 100
to 400 °C has been amplified in Fig. 4.
been also quantified. Fig. 5 shows that there is not Irgafos 168 left after
day 8 in the distinct films of the different iPP based materials except in
the PP-DEHP where a 10% is still observed. Simultaneously, the content
as promising plasticizers to avoid the use of DEHP in those medical in oxidized form of Irgafos 168 is increased with time, as the resultant
devices. A similar conclusion is reached here in the case of the iPP. gas chromatograms revealed. It is reported [25] that this phosphite
Kambia et al. [23] discussed the leaching ability of DEHP and TOTM stabilizer uptakes oxygen from the environment and is oxidized stoi-
in haemodialysis tubing when they included DEHP as unique plasticizer chiometrically to the corresponding phosphate. Accordingly, it turns
or in combination with TOTM. Results showed a lower migration when out rather interesting to learn the chemical changes that occur within
TOTM was added into the formulation. These data seemed to indicate the iPP chains during the loss of plasticizer and the Irgafos 168 con-
that the most resistant plasticizer to be lost was the TOTM. A recent sumption at 95 °C with increasing degradation time.
work [24] also demonstrated that when the TOTM was added to PVC, Fig. 6 shows ATR-FTIR spectra for all of the samples evaluated.
the resultant material was thermally more stable than the ones in- Development of oxidant species is clearly noticed in the spectra for
corporating other plasticizers. These investigations induce to think that thermo-oxidation times of 10 days or longer. Chemical changes related
TOTM could become the safest plasticizer candidate to replace the use to the beginning of iPP degradation have been extensively studied by
of the toxic DEHP. infrared spectroscopy [26–30]. The appearance of bands in the region
In order to assess possible losses of plasticizer during processing, between 1700 and 1800 cm−1, ascribed to the formation of carboxyl
and also the potential effect of plasticizers on the iPP degradation, TGA species, such as acids, esters, ketones or lactones, starts to be observed
measurements have been performed on the original samples. The re- as degradation is initiated [26–30].
sults are shown in Fig. 3. It can be observed that there is practically no Presence of the distinct plasticizers is also identified in the initial
effect on the main PP degradation, since the plasticizer is lost first (as films before the thermal treatment (results for day 0, d0) through the
clearly depicted in Fig. 4). Anyway, a small but appreciable stability FTIR spectra from observation of the specific bands at each additive
increase is visible in the case of PP-TOTM.
Fig. 4 shows the comparison of the TGA curves of the pristine
plasticizers (bottom) with those attained for the iPP-plasticizer

Fig. 4. Comparison of the TGA curves of the pure plasticizers (bottom) with
those obtained for the iPP-plasticized materials (upper, amplification of Fig. 3). Fig. 5. Consumption of Irgafos 168 on thermo-oxidation time at 95 °C.

5
E. Blázquez-Blázquez, et al. Polymer 181 (2019) 121749

of diverse carbonyl-carboxyl groups. No evidence of this band is seen


while antioxidant is present, even when the amount is very little. At
longer times, when Irgafos 168 consumption is complete, the multiple
bands clearly appear. The global area of all carbonyl-carboxyl changes
with degradation time, increasing as it is raised. These results are
comparable with those reported in other accelerated study under dif-
ferent weathering conditions [31], although there are significant dif-
ferences by effect of the different plasticizers.
Loss of the plasticizer and consumption of Irgafos168 at 95 °C at the
distinct thermo-oxidation times have involved changes in the chemical
composition of iPP that indicate formation of oxidative species because
of beginning of degradation. Their effect on crystalline structure and its
phase transitions is important to be understood now.
The first parameter that can be altered by the thermal treatment
applied is the molecular weight, since in polymers its variation is very
sensitive with the beginning of decomposition processes [29]. Two of
these materials have been analyzed for the day 0 (d0) and 16 (d16): the
pristine iPP and the PP-DINCH. The average values of molecular weight
derived from SEC experiments for the pure iPP are 175,000 and
136,000 for d0 and d16, respectively, while they are 166,000 and
153,000 in the PP-DINCH for d0 and d16, respectively. It seems that
incorporation of the plasticizer allows maintaining rather unchanged
the molecular weight during thermal treatment. Consequently, pristine
iPP (where only presence of antioxidant was detected) shows a more
significant decrease of this parameter. Moreover, a variation in poly-
dispersity has been also seen in the d16 samples either for the neat iPP
or PP-DINCH after treatment at 95 °C for 16 days. It becomes broader
for these d16 specimens in these two materials, mainly because of the
formation of a significant fraction of a low molecular weight tail. The
Fig. 6. ATR-FTIR spectra for the different samples in the carbonyl region (ex- decay on the average molecular weights points out the existence of
panded down to the reference band at 1460 cm−1) and for the various thermo- chain scission whereas the growth of molecular weight distribution
oxidation times at 95 °C. seems to prove that thermo-oxidation does not involve random chain
scission. As reported previously in electron beam irradiated metallo-
incorporated. These bands do not overlap with any characteristic one cene iPP [29], polydispersity should decrease slightly with increasing
from the pure iPP, although some of them appear in the carbonyl region degradation time if simple scission would take place.
of degraded iPP. Thus, ATBC shows a unique wide band at 1748 cm−1 How do these chemical changes and chain length variations affect
related to its ester groups in day 0. The DEHP is characterized at that the crystalline structure at the end of the thermo-oxidation process? In
time (day 0) by its corresponding ester functionality at 1733 cm−1 to- general, isotactic polypropylene can crystallize into different poly-
gether with other small signals at 1072 and 741 cm−1. The specific morphs by changing microstructural features, crystallization conditions
ester band of DINCH appears at 1736 cm−1 and for DEHT bands are and other factors like incorporation of specific nucleants [32–35]. Thus,
seen at 1726, 1266 and 731 cm−1. The TOTM, which is the plasticizer three different polymorphic modifications, α, β and γ, all of them
with higher molecular weight, exhibits its ester band at 1731 cm−1 sharing a three-fold conformation, have been described together with a
(and another two at 1233 and 752 cm−1). phase of intermediate or mesomorphic order obtained by fast
Fig. 6 clearly shows that the wavenumber interval corresponding to quenching [32,34–38]. In addition to these four modifications, a tri-
the ester group present in the distinct plasticizers is superimposed to the gonal form has been more recently reported in the case of isotactic
one related to formation of the different oxidative species in the iPP. A copolymers of propylene with high contents of 1-hexene [39–41] or 1-
common feature is observed in every of PP-plasticizer materials: all pentene [42,43], in propylene terpolymers with 1-pentene and 1-
show at day 0 an ester band, which is characteristic of each plasticizer hexene [44,45] as comonomers, and in propylene terpolymers with 1-
as just commented, while no band is observed for neat iPP. As the de- pentene and 1-heptene [46], synthesized all by using metallocene cat-
gradation progresses, the plasticizer and the antioxidant additive are alysts.
lost and those small bands decrease in intensity, even disappearing in Fig. 7 shows the WAXD profiles at room temperature for pristine iPP
some materials. Simultaneously, the appearance of a complex and in- and the distinct plasticized materials before beginning the thermal
creasingly intense band arising from coexistence of different oxidant treatment, i.e., for the d0 specimens. All of these patterns exhibit the
groups is observed as mentioned above for the pristine iPP. The cases of main reflections characteristic of the monoclinic lattice [32,33]. No
PP-DEHT and PP-TOTM are significantly different. Thus, when obser- evidence is seen for the distinctive (117) diffraction ascribed to the
ving the spectra at day 16, it seems that the degradation of iPP is higher orthorhombic cell. This γ polymorph has been studied in detail for iPP
in those two materials. However, it has to be considered that these two homopolymers and a close correlation was found between the con-
plasticizers are kept in very significant proportions even after 16 days centration of defects and the maximum content of the orthorhombic
(around 88% of TOTM and 22% of DEHT). Consequently, the ester polymorph that the chains could develop [47]. Its presence is then more
bands of the remaining plasticizer are added to the ones arising from common in iPP synthesized with metallocenes than with Ziegler-Natta
degradation of iPP. It should be commented that in order to have catalysts [48–50]. The absence of this crystalline lattice in the present
comparable results, the area of the peaks obtained in the carbonyl re- iPP can be related to the rapid cooling applied from the melt during
gion has been normalized taking as reference the band at 1460 cm−1. processing of the diverse films.
Fig. 6 proves how important is the presence of Irgafos 168 in the Fig. 7 also depicts that significant changes are not observed in these
observation of the intense and composite band ascribed to the existence WAXD profiles, independently of plasticizer presence and its type. Ac-
cordingly, the degree of crystallinity developed does not considerably

6
E. Blázquez-Blázquez, et al. Polymer 181 (2019) 121749

plasticizer and takes place at approximately 113.5 °C, for the pristine
iPP and the several plasticized materials. As the thermo-oxidation
progresses, crystallization in the neat iPP is first very slightly moved to
higher temperatures for the lowest times, while this exothermic process
is clearly split into two peaks (at about 112.5 and 105 °C) for the longest
stay at 95 °C. This feature is also observed in all the distinct plasticized
materials, although the degradation time at which these two exo-
thermic events take place depends on the plasticizer used. Incidentally,
the effect on the crystallinity is not significant, with values slightly
higher than those mentioned above for the first melting of d0 samples.
The PP-ATBC, PP-DEHP and PP-DINCH materials show the ex-
istence of that double peak behavior already at d10, which is a time
shorter than for the neat iPP. Moreover, a unique crystallization process
is seen in those three samples for the longest d16 time and its location
corresponds to the process that occurs at the lower temperature, i.e., at
approximately 106 °C.
On the contrary, the PP-DEHT and PP-TOTM materials behave
Fig. 7. X-ray profiles at room temperature for the pristine iPP and the different analogously to pristine iPP, so that a single crystallization process oc-
original plasticized materials (specimens d0). curs at short times and the splitting is observed for the treatment d16.
Why do these differences exist? The results exhibited up to now have
indicated similar characteristics except for the effect of degradation in
vary in these initial films. Similarly, the corresponding melting tem-
the capability of maintenance or loss of the plasticizers incorporated. In
peratures, estimated from the first melting process in the DSC experi-
fact, those exhibiting only the low-temperature exotherm at d16 are the
ments, also remain rather unchanged, at around 144.5 °C, for all the
PP-ATBC, PP-DEHP and PP-DINCH materials, i.e., those that lose faster
different materials analyzed. Moreover, the melting enthalpies are also
the plasticizer. This trend seems to be delayed in the neat iPP, which
rather similar, with a value of 99 ± 5 J/g, meaning that the DSC
does not incorporate any plasticizer, and in the PP-DEHT and PP-
crystallinity takes an average value of 0.62 ± 0.03.
TOTM, both keeping around 22% and 88% of their respective plasti-
Fig. 8 displays, however, that once the materials become completely
cizer amounts after 16 days of thermal treatment. Fig. 9 represents the
molten, the subsequent crystallization is strongly dependent on either
relative area of the crystallization event located at the lower tempera-
degradation time or the large/small loss of plasticizer (and antioxidant)
ture for the highest thermo-oxidation times tested, showing values
in the iPP matrix. Before thermo-oxidation, i.e., in the samples labeled
significantly smaller in these three materials: iPP, PP-DEHT and PP-
as d0, crystallization is not practically affected by incorporation of
TOTM, being these latest the plasticized materials with high plasticizer
amounts present in the d16 samples (approximately 22% for DEHT and
88% for TOTM).
Which is the reason behind this behavior? Presence of a double
crystallization process primarily might indicate the development of two

Fig. 8. Crystallization curves (at 20 °C/min) from the melt for the distinct Fig. 9. Relative area of the crystallization process appearing at the lowest
materials analyzed after the different degradation times at 95 °C. temperature for the different materials (d16 specimens).

7
E. Blázquez-Blázquez, et al. Polymer 181 (2019) 121749

Fig. 11. Relative content in the γ orthorhombic polymorph estimated by WAXD


for the different materials after cooling from the melt at 20 °C/min: original
Fig. 10. Synchrotron WAXD 1D diffractograms at room temperature after ones (d0, left) and after 16 days degradation (d16, right).
cooling from the melt at 20 °C/min of the different materials analyzed: upper
plot, before thermal treatment, d0 specimens; and, lower plot, after degradation that occur with the degradation time allow rising the length of the
at 95 °C for 16 days, d16 samples.
regular sequences and, consequently, a higher amount of monoclinic
crystallites are formed at the expense of the orthorhombic ones.
populations of crystals with rather different crystallite sizes or forma- To understand the cause behind the splitting of crystallization into
tion of two distinct polymorphs. Initial films of the different materials two exothermic events, real-time variable-temperature WAXD experi-
exhibit exclusively the monoclinic crystal lattice with analogous char- ments with synchrotron radiation have been performed on cooling
acteristics, including the degree of crystallinity, as depicted in Fig. 7. under identical conditions than DSC measurements. Results are de-
However, all the samples develop a considerable amount of orthor- picted in Fig. 12. Degradation affects not only the formation of or-
hombic γ modification after crystallization from the melt at 20 °C/min, thorhombic lattice but the crystallization process is considerably ham-
as observed in Fig. 10, where the (117) diffraction of that modification pered, as deduced from comparison of the profiles exhibited by PP-d0
is clearly observed. But more importantly, there are very significant and PP-d16. This hindrance is much more noticeable in the PP-DEHP
differences in the amount of γ form between d0 and d16 specimens, material. Thus, beginning of crystallization is shifted to significantly
being considerably lower in the degraded samples. lower temperatures when comparing the patterns observed in PPDEHP-
Thus, the upper plot in Fig. 10 (d0 samples) shows coexistence of d0 with those exhibited by PPDEHP-d16. Importantly, the ratio be-
two polymorphs, the α monoclinic and γ orthorhombic one. Once the tween monoclinic and orthorhombic modifications keeps approxi-
initial monoclinic crystallites (Fig. 7) developed along the films pro- mately constant during crystallization for every sample. Microstructural
cessing are molten, those two crystalline lattices can be generated, first changes derived from degradation leads to a double variation in the
of all because the metallocene origin of the iPP and, secondly, because crystalline structure: on one hand, formation of orthorhombic crystal-
the rate applied in these crystallization experiments (20 °C/min) is lites is importantly hindered; and, on the other hand, crystallization is
smaller than that used during film processing. Content in the γ or- considerably delayed. Both aspects could be associated with the de-
thorhombic modification can be tailored by changing either crystal- velopment of a small content of branches during the non-random scis-
lization temperature in isothermal tests or rate in the dynamic experi- sion deduced from the reduction of molecular weight and broadening of
ments, turning out favorable for its formation the crystallization at low its distribution, both commented previously. Existence of these lateral
rates [51]. chains slows down crystallization but could contribute to lengthen
The relative amounts of both crystalline modifications have been isotactic sequences fact that favors preferential formation of monoclinic
assessed from the area ratio of their characteristic reflections: the (130) crystallites.
one located at s of 2.01 nm−1 for monoclinic crystals; and the (117) Fig. 13 shows the excellent agreement attained in location and
diffraction at 2.25 nm−1 associated with the γ lattice [29,52]. Left plot shape between the derivative of the WAXD crystalline area and the DSC
in Fig. 11 shows that the content in the orthorhombic polymorph for the results for PP (left plots) and PP-DEHP (right plots). In order to de-
different iPP based materials before degradation is rather significant termine that WAXD crystalline area, the amorphous halo subtraction
(around 50%) despite the fact that crystallization rate used (20 °C/min) has been performed to the X-ray experimental profile at each and every
has not been very low. single temperature. Result of that subtraction involves exclusively the
The lower plot in Fig. 10 together with the right representation in crystalline contribution and, thus, its area can be estimated. The ob-
Fig. 11 shows the important effect that degradation for 16 days exerts in tainment of the amorphous halo at each temperature has been de-
the crystallization capability for all the different materials. Consump- scribed previously in detail [19].
tion of the antioxidant, formation of oxidative species, changes in the These two complementary measurements, derivative of crystalline
microstructure and/or loss of plasticizer significantly hinder the de- area and DSC cooling curve, allow deducing crystallization hindrance in
velopment of the orthorhombic modification in more than a half. addition to difficulty for formation of γ crystals.
Carbon nuclear magnetic resonance experiments have been carried out It is expected that these important changes in crystalline details
in order to achieve supplementary microstructural information on the affect the macroscopic response of the different materials. The char-
influence of degradation in the iPP macrochains. The results indicate acteristics at the surface through profilometry and the mechanical re-
that degradation involves an increase in the isotacticity of the poly- sponse with the thermo-oxidation time have been checked in some of
meric chains. Consequently, the lower number of errors in the materials the materials. Fig. 14 shows the results achieved for those materials
d16 leads to a smaller formation of the γ polymorph [47]. The changes losing the plasticizer at the fastest and slowest rate, i.e., PP-ATBC and

8
E. Blázquez-Blázquez, et al. Polymer 181 (2019) 121749

Fig. 14. Dependence of normalized average surface roughness (top) and mi-
crohardness (bottom) upon degradation time for the PP-ATBC and PP-TOTM
materials.

turned out very insightful respect to the degree of plasticizer loss, al-
though it is a superficial characteristic and these voids and crazes are
located all through the bulk. Consequently, for a given material,
Fig. 12. Real-time variable-temperature WAXD diffractograms on cooling from
roughness changes are observed when the incorporated plasticizer in-
the melt (at 20 °C/min) performed with synchrotron radiation for iPP (upper)
itiates its migration mechanism. Moreover, these variations are also
and PP-DEHP (lower) at d0 (left) and d16 (right). For clarity, only the region
between 120 and 100 °C is depicted. dependent on amount of the plasticizer left in the different materials at
a specific degradation time since the material that has lost a higher
content becomes rougher.
Microhardness (MH) experiments have been selected for estimation
of the mechanical behavior of these materials because they are also
related to the response at surface. Values obtained at the different
thermo-oxidation times are listed in Table 1. The hardness of a material
can be defined as a measure of the resistance to a permanent de-
formation or damage. The deformation of a semicrystalline polymer
under the action of the indenter is basically ruled by several effects: an
elastic deformation that yields an instant elastic recovery on unloading;
a permanent plastic deformation determined by arrangement and
structure of crystallites and their connection by tie molecules and en-
tanglements; and a viscoelastic contribution, which is time-dependent
during loading with a long delayed recovery after load removal.
Therefore, although MH measurements are ascribed to the superficial
mechanical response they also involve a complex combination of other
bulk mechanical properties (elastic modulus, yield strength, strain
Fig. 13. Top plots: Derivatives of the variation with temperature of area of the
crystalline diffractions for the PP (left representation) and PP-DEHP (right re- Table 1
presentation) at d0 and d16 degradation times. Bottom plots: DSC cooling Microhardness (MH) values for the different plasticized materials at the distinct
curves for the d0 and d16 specimens for the PP (left representation) and PP- thermo-oxidation times.
DEHP (right representation).
samples MH (MPa)

Thermo-oxidation time (days)


PP-TOTM. The upper plot clearly shows the increase in roughness at the
surface of films with the degradation time. In addition, it is un- 0 2 4 8 16
doubtedly noticed that roughness is in PP-ATBC considerably higher
than in PP-TOTM. Loss of the plasticizer leads to the formation within PP-ATBC 80 ± 4 87 ± 5 93 ± 5 97 ± 5 116 ± 6
PP-DEHP 80 ± 4 85 ± 5 92 ± 5 97 ± 5 100 ± 5
the bulk film of a frame of voids and crazes during degradation. This
PP-DINCH 80 ± 4 83 ± 5 87 ± 5 94 ± 5 98 ± 5
path is related to the run throughout the film that a specific plasticizer PP-DEHT 80 ± 4 82 ± 4 83 ± 54 89 ± 5 95 ± 5
has to perform during its leakage out. Roughness determination has PP-TOTM 80 ± 4 80 ± 4 81 ± 4 83 ± 4 91 ± 4

9
E. Blázquez-Blázquez, et al. Polymer 181 (2019) 121749

hardening, toughness) and a correlation with rigidity modulus is com- the funding received to perform these measurements. Authors would
monly found [21,53–55]. Moreover, microhardness variations with like to thank to Dr. N. García her help in the measurements with the
structural parameters can be also expected since mechanical properties optical profilometer.
are structure-dependent. Accordingly, all the variables that lead to an
increase of crystallinity and crystallite sizes [56,57] will provide higher References
microhardness values.
Fig. 14 shows a very good agreement between dependence of [1] Y. Haishima, R. Matsuda, Y. Hayashi, C. Hasegawa, T. Yagami, T. Tsuchiya, Risk
average superficial roughness and microhardness values on degradation assessment of di(2-ethylhexyl)phthalate released from PVC blood circuits during
hemodialysis and pump-oxygenation therapy, Int. J. Pharm. 274 (2004) 119–129.
times. Thus, MH increases as time does and their values are in PP-ATBC [2] D. Koniecki, R. Wang, R.P. Moody, J.P. Zhu, Phthalates in cosmetic and personal
superior than in PP-TOTM. Higher MH values mean larger rigidity, but care products: concentrations and possible dermal exposure, Environ. Res. 111
why is PP-ATBC stiffer than PP-TOTM? Mechanical magnitudes are (2011) 329–336.
[3] F. Chiellini, M. Ferri, A. Morelli, L. Dipaola, G. Latini, Perspectives on alternatives
associated, as just mentioned, with structural parameters, such as to phthalate plasticized poly(vinyl chloride) in medical devices applications, Prog.
crystallinity and/or crystal size. Nevertheless, there are not important Polym. Sci. 38 (2013) 1067–1088.
differences in both magnitudes as deduced from X-ray diffraction and [4] A. Prasannan, J.J. Jhu, C.J. Wu, S.Y. Lin, H.C. Tsai, Evaluation of the temperature
and molecular weight dependent migration of di(2-ethylhexyl) phthalate from
melting temperatures results. The higher values found in PP-ATBC isotactic polypropylene composites, React. Funct. Polym. 113 (2017) 70–76.
might have relation to the amount of void and craze paths, these being [5] K.L. Howdeshell, A.K. Hotchkiss, L.E. Gray, Cumulative effects of antiandrogenic
greater than in PP-TOTM because the ATBC plasticizer is completely chemical mixtures and their relevance to human health risk assessment, Int. J. Hyg
Environ. Health 220 (2017) 179–188.
lost while TOTM at the end of degradation test remains around 88%.
[6] A.D. LaFleur, K.A. Schug, A review of separation methods for the determination of
Those hollow channels impose residual stresses [58,59] that lead to the estrogens and plastics-derived estrogen mimics from aqueous systems, Anal. Chim.
increase of mechanical parameters ascribed to rigidity, like micro- Acta 696 (2011) 6–26.
hardness values. Accordingly, the more holes within films there are, [7] H.Q. Fang, J. Wang, R.A. Lynch, Migration of di(2-ethylhexyl)phthalate (DEHP) and
di-n-butylphthalate (DBP) from polypropylene food containers, Food Control 73
either because degradation time increases or a given plasticizer content (2017) 1298–1302.
decreases, the greater the MH value will be. MH also rises with time for [8] X.L. Cao, Phthalate esters in foods: sources, occurrence, and analytical methods,
the rest of the plasticized materials, as deduced from Table 1. Moreover, Compr. Rev. Food Sci. Food Saf. 9 (2010) 21–43.
[9] L. Bernard, R. Cueff, C. Breysse, B. Decaudin, V. Sautou, G. Armed Study,
the smaller is the amount of plasticizer remaining at a given time the Migrability of PVC plasticizers from medical devices into a simulant of infused
greater is the increase in MH. It should be also reminded that a plas- solutions, Int. J. Pharm. 485 (2015) 341–347.
ticizer is added to increase flexibility within polymers through weak- [10] H. Fromme, A. Schutze, T. Lahrz, M. Kraft, L. Fernbacher, S. Siewering, R. Burkardt,
S. Dietrich, H.M. Koch, W. Volkel, Non-phthalate plasticizers in German daycare
ening inter and intramolecular polymeric chains interactions. There- centers and human biomonitoring of DINCH metabolites in children attending the
fore, higher plasticizer contents remaining in the material will lead to centers (LUPE 3), Int. J. Hyg Environ. Health 219 (2016) 33–39.
smaller stiffness. Relationships might change in more advanced stages [11] L. Correia-Sa, A. Schutze, S. Norberto, C. Calhau, V.F. Domingues, H.M. Koch,
Exposure of Portuguese children to the novel non-phthalate plasticizer di-(iso-
of degradation, when extent in scission reactions is significantly en- nonyl)-cyclohexane-1,2-dicarboxylate (DINCH), Environ. Int. 102 (2017) 79–86.
larged and reduction in average in molecular weight is more severe. [12] P. Gimeno, S. Thomas, C. Bousquet, A.F. Maggio, C. Civade, C. Brenier, P.A. Bonnet,
Identification and quantification of 14 phthalates and 5 non-phthalate plasticizers
in PVC medical devices by GC-MS, J. Chromatogr. B Anal. Technol. Biomed. Life
4. Conclusions
Sci. 949 (2014) 99–108.
[13] D. Bourdeaux, M. Yessaad, P. Chennell, V. Larbre, T. Eljezi, L. Bernard, V. Sautou,
A small content of different plasticizers (2 wt%) has been added to a A.S. Grp, Analysis of PVC plasticizers in medical devices and infused solutions by
metallocene isotactic polypropylene to evaluate their loss rate under GC-MS, J. Pharm. Biomed. Anal. 118 (2016) 206–213.
[14] T. Eljezi, P. Pinta, D. Richard, J. Pinguet, J.M. Chezal, M.C. Chagnon, V. Sautou,
severe thermo-oxidative conditions together with the subsequent effects G. Grimandi, E. Moreau, In vitro cytotoxic effects of DEHP-alternative plasticizers
of degradation time on crystalline characteristics, on film surface and and their primary metabolites on a L929 cell line, Chemosphere 173 (2017)
on final mechanical performance. 452–459.
[15] M. Erceg, T. Kovacic, I. Klaric, Thermal degradation of poly(3-hydroxybutyrate)
The remaining content for each plasticizer displays important dif- plasticized with acetyl tributyl citrate, Polym. Degrad. Stab. 90 (2005) 313–318.
ferences with degradation time during their thermo-oxidation at 95 °C [16] M.P. Arrieta, J. Lopez, D. Lopez, J.M. Kenny, L. Peponi, Effect of chitosan and ca-
in the distinct iPP based films. Results indicate that PP-TOTM material techin addition on the structural, thermal, mechanical and disintegration properties
of plasticized electrospun PLA-PHB biocomposites, Polym. Degrad. Stab. 132
is the one where plasticizer remains in a very significant content after (2016) 145–156.
the 16 days of thermal treatment whereas the ATCB amount is drasti- [17] A. Greco, A. Maffezzoli, Cardanol derivatives as innovative bio-plasticizers for poly-
cally reduced in only 10 h. Antioxidant consumption takes place si- (lactic acid), Polym. Degrad. Stab. 132 (2016) 213–219.
[18] R. Krache, R. Benavente, J.M. López-Majada, J.M. Perena, M.L. Cerrada, E. Pérez,
multaneously along degradation and chemical changes occur within the
Competition between α, β, and γ polymorphs in a β-nucleated metallocenic isotactic
iPP chains, which involve formation of carboxyl-carbonyl species, re- polypropylene, Macromolecules 40 (2007) 6871–6878.
duction of average molecular weight and a widening of molecular [19] E. Pérez, M.L. Cerrada, R. Benavente, J.M. Gómez-Elvira, Enhancing the formation
of the new trigonal polymorph in isotactic propene-1-pentene copolymers: de-
weight distribution.
termination of the X-ray crystallinity, Macromol. Res. 19 (2011) 1179–1185.
All of these changes in the iPP microstructure promoted a delay in [20] ASME, Surface Texture (Surface Roughness, Waviness, and Lay), B46.1-2002
the overall crystallization and a hindrance in formation of the orthor- Standard, American Society of Mechanical and Engineering, New York, 2002.
hombic crystallites, mostly in the materials with higher plasticizer loss [21] F.J.B. Calleja, Microhardness relating to crystalline polymers, Adv. Polym. Sci. 66
(1985) 117–148.
rates. An increase in roughness at the film surface is simultaneously [22] E.J. Dziwinski, B.P. Pozniak, J. Lach, GC/MS and ESI/MS identification of the new
observed with degradation time together with an enlargement of ri- generation plasticizers - cis and trans isomers of some 1,2-cyclohexane dicarboxylic
gidity. acid di(n-and isononyl) esters, Polym. Test. 62 (2017) 319–324.
[23] K. Kambia, T. Dine, R. Azar, B. Gressier, M. Luyckx, C. Brunet, Comparative study of
In summary, PP could be considered a friendly matrix compared the leachability of di(2-ethylhexyl) phthalate and tri(2-ethylhexyl) trimellitate from
with PVC and TOTM seems to be a safe candidate as plasticizer to re- haemodialysis tubing, Int. J. Pharm. 229 (2001) 139–146.
place the use of the toxic DEHP. [24] Q. Wang, W. Wu, Y.F. Tang, J.J. Bian, S.W. Zhu, Thermal degradation kinetics of
plasticized poly (vinyl chloride) with six different plasticizers, J. Macromol. Sci.
Part B-Physics 56 (2017) 420–434.
Acknowledgements [25] J. Pospíšil, Chemical and photochemical behaviour of phenolic antioxidants in
polymer stabilization: a state of the art report, part II, Polym. Degrad. Stab. 39
(1993) 103–115.
The financial support from project MAT2016-79869-C2-1-P (AEI/ [26] G. Geuskens, M.S. Kabamba, Photooxidation of polymers-Part V: a new chain
FEDER, UE) is greatly acknowledged. The synchrotron experiments scission mechanism in polyolefins, Polym. Degrad. Stab. 4 (1982) 69–76.
were performed at beamline BL11-NCD at ALBA Synchrotron Light [27] P. Delprat, X. Duteurtre, J.L. Gardette, Photooxidation of unstabilized and HALS-
stabilized polyphasic ethylene-propylene polymers, Polym. Degrad. Stab. 50 (1995)
Facility with the collaboration of ALBA staff. Authors are grateful for

10
E. Blázquez-Blázquez, et al. Polymer 181 (2019) 121749

1–12. PP in random C3/C5/C6 terpolymers, Polymer 54 (2013) 1656–1662.


[28] J.L. Philippart, C. Sinturel, R. Arnaud, J.L. Gardette, Influence of the exposure [45] A. García-Peñas, J.M. Gõmez-Elvira, E. Pérez, M.L. Cerrada, Isotactic poly(propy-
parameters on the mechanism of photooxidation of polypropylene, Polym. Degrad. lene-co-1-pentene-co-1-hexene) terpolymers: synthesis, molecular characterization,
Stab. 64 (1999) 213–225. and evidence of the trigonal polymorph, J. Polym. Sci. A Polym. Chem. 51 (2013)
[29] M.L. Cerrada, E. Pérez, R. Benavente, J. Ressia, C. Sarmoria, E.M. Vallés, Gamma 3251–3259.
polymorph and branching formation as inductors of resistance to electron beam [46] A. García-Peñas, J.M. Gómez-Elvira, R. Barranco-García, E. Pérez, M.L. Cerrada,
irradiation in metallocene isotactic polypropylene, Polym. Degrad. Stab. 95 (2010) Trigonal δ form as a tool for tuning mechanical behavior in poly(propylene-co-1-
462–469. pentene-co-1-heptene) terpolymers, Polymer 99 (2016) 112–121.
[30] C. Rouillon, P.O. Bussiere, E. Desnoux, S. Collin, C. Vial, S. Therias, J.L. Gardette, Is [47] R.G. Alamo, M.H. Kim, M.J. Galante, J.R. Isasi, L. Mandelkern, Structural and ki-
carbonyl index a quantitative probe to monitor polypropylene photodegradation? netic factors governing the formation of the gamma polymorph of isotactic poly-
Polym. Degrad. Stab. 128 (2016) 200–208. propylene, Macromolecules 32 (1999) 4050–4064.
[31] A. Niemczyk, K. Dziubek, M. Grzymek, K. Czaja, Accelerated laboratory weathering [48] F. Auriemma, C. De Rosa, Crystallization of metallocene-made isotactic poly-
of polypropylene composites filled with synthetic silicon-based compounds, Polym. propylene: disordered modifications intermediate between the alpha and gamma
Degrad. Stab. 161 (2019) 30–38. forms, Macromolecules 35 (2002) 9057–9068.
[32] A. Turner-Jones, J.M. Aizlewood, D.R. Beckett, Crystalline forms of isotactic poly- [49] O. Prieto, J.M. Pereña, R. Benavente, E. Pérez, M.L. Cerrada, Viscoelastic relaxation
propylene, Makromol. Chem. 75 (1964) 134–158. mechanisms of conventional polypropylene toughened by a plastomer, J. Polym.
[33] S. Brückner, S.V. Meille, V. Petraccone, B. Pirozzi, Polymorphism in isotactic Sci. B Polym. Phys. 41 (2003) 1878–1888.
polypropylene, Prog. Polym. Sci. 16 (1991) 361–404. [50] R.G. Alamo, J.A. Blanco, P.K. Agarwal, J.C. Randall, Crystallization rates of mat-
[34] J. Varga, Supermolecular structure of isotactic polypropylene, J. Mater. Sci. 27 ched fractions of MgCl2-supported Ziegler-Natta and metallocene isotactic poly
(1992) 2557–2579. (propylene)s. 1. The role of chain microstructure, Macromolecules 36 (2003)
[35] B. Lotz, J.C. Wittmann, A.J. Lovinger, Structure and morphology of poly(propy- 1559–1571.
lenes): a molecular analysis, Polymer 37 (1996) 4979–4992. [51] C. De Rosa, F. Auriemma, M. Paolillo, L. Resconi, I. Camurati, Crystallization be-
[36] J. Grebowicz, S.F. Lau, B. Wunderlich, The thermal properties of polypropylene, J. havior and mechanical properties of regiodefective, highly stereoregular isotactic
Polym. Sci., Polym. Symp. 71 (1984) 19–37. polypropylene: effect of regiodefects versus stereodefects and influence of the
[37] M.J. Polo-Corpa, R. Benavente, T. Velilla, R. Quijada, E. Pérez, M.L. Cerrada, molecular mass, Macromolecules 38 (2005) 9143–9154.
Development of the mesomorphic phase in isotactic propene/higher a-olefin co- [52] R. Barranco-García, J.M. López-Majada, J.C. Martínez, J.M. Gómez-Elvira, E. Pérez,
polymers at intermediate comonomer content and its effect on properties, Eur. M.L. Cerrada, Confinement of iPP crystallites within mesoporous SBA-15 channels
Polym. J. 46 (2010) 1345–1354. in extruded iPP-SBA-15 nanocomposites studied by Small Angle X-ray scattering,
[38] E. Pérez, J.M. Gómez-Elvira, R. Benavente, M.L. Cerrada, Tailoring the formation Microporous Mesoporous Mater. 272 (2018) 209–216.
rate of the mesophase in random propylene-co-1-pentene copolymers, [53] M.L. Cerrada, J.L. de la Fuente, M. Fernandez-Garcia, E.L. Madruga, Viscoelastic
Macromolecules 45 (2012) 6481–6490. and mechanical properties of poly(butyl acrylate-g-styrene) copolymers, Polymer
[39] B. Poon, M. Rogunova, A. Hiltner, E. Baer, S.P. Chum, A. Galeski, E. Piorkowska, 42 (2001) 4647–4655.
Structure and properties of homogeneous copolymers of propylene and 1-hexene, [54] M.L. Cerrada, R. Benavente, E. Pérez, Crystalline structure and viscoelastic behavior
Macromolecules 38 (2005) 1232–1243. in composites of a metallocenic ethylene-1-octene copolymer and glass fiber,
[40] C. De Rosa, S. Dello Iacono, F. Auriemma, E. Ciaccia, L. Resconi, Crystal structure of Macromol. Chem. Phys. 203 (2002) 718–726.
isotactic propylene-hexene copolymers: the trigonal form of isotactic poly- [55] M.L. Cerrada, E. Pérez, J.P. Lourenço, J.M. Campos, M. Rosário Ribeiro, Hybrid
propylene, Macromolecules 39 (2006) 6098–6109. HDPE/MCM-41 nanocomposites: crystalline structure and viscoelastic behaviour,
[41] M.L. Cerrada, M.J. Polo-Corpa, R. Benavente, E. Pérez, T. Velilla, R. Quijada, Microporous Mesoporous Mater. 130 (2010) 215–223.
Formation of the new trigonal polymorph in iPP - 1-hexene copolymers. [56] C. Fonseca, J. Pereña, R. Benavente, M.L. Cerrada, A. Bello, E. Pérez, Microhardness
Competition with the mesomorphic phase, Macromolecules 42 (2009) 702–708. and thermal study of the annealing effects in vinyl alcohol-ethylene copolymers,
[42] C. De Rosa, O.R. de Ballesteros, F. Auriemma, M.R. Di Caprio, Crystal structure of Polymer 36 (1995) 1887–1892.
the trigonal form of isotactic propylene-pentene copolymers: an example of the [57] J. Arranz-Andrés, J.L. Guevara, T. Velilla, R. Quijada, R. Benavente, E. Pérez,
principle of entropy-density driven phase formation in polymers, Macromolecules M.L. Cerrada, Syndiotactic polypropylene and its copolymers with alpha-olefins.
45 (2012) 2749–2763. Effect of composition and length of comonomer, Polymer 46 (2005) 12287–12297.
[43] A. García-Peñas, J.M. Gómez-Elvira, V. Lorenzo, E. Pérez, M.L. Cerrada, [58] L.C.E. Struik, Internal Stresses, Dimensional Instabilities and Molecular
Unprecedented dependence of stiffness parameters and crystallinity on comonomer Orientations in Plastics, John Wiley & Sons Ltd, Chichester, U.K., 1990.
content in rapidly cooled propylene-co-1-pentene copolymers, Polymer 130 (2017) [59] M.L. Cerrada, J.M. Pereña, R. Benavente, E. Pérez, Viscoelastic processes in vinyl
17–25. alcohol-ethylene copolymers. Influence of composition and thermal treatment,
[44] L. Boragno, P. Stagnaro, F. Forlini, F. Azzurri, G.C. Alfonso, The trigonal form of i- Polymer 41 (2000) 6655–6661.

11

You might also like