You are on page 1of 14

Polymer 227 (2021) 123848

Contents lists available at ScienceDirect

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

Microstructural insight on strain-induced crystallization of ethylene/


propylene(/diene) random copolymers
Finizia Auriemma a, *, Miriam Scoti a, **, Fabio De Stefano a, Giovanni Talarico a,
Odda Ruiz de Ballesteros a, Rocco Di Girolamo a, Anna Malafronte a, Claudio De Rosa a,
Roberta Cipullo a, Martin Van Duin b
a
Dipartimento di Scienze Chimiche, Università di Napoli “Federico II”, Complesso Monte Sant’Angelo, Via Cintia, 80126, Napoli, Italy
b
ARLANXEO Performance Elastomers, Innovation, P.O. Box 185, 6160AD Geleen the Netherlands

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

Keywords: The tensile properties and the structural transformations occurring by effect of deformation of un-vulcanized
Ethylene/propylene(/diene) copolymers ethylene/propylene(/diene) elastomers with ethylene contents ranging from 57 to 83 mol% are analyzed and
Chain microstructure interpreted in terms of the chain microstructure. It is shown that the samples with ethylene contents larger than
Mechanical properties
or equal to 73 mol% experience a slight but significant increase of crystallinity upon stretching and partial
Strain induced crystallization
Strain hardening
melting of the new oriented crystals formed by stretching after releasing the tension. The degree of crystallinity,
the tensile properties, the incremental crystallinity achieved upon stretching and the residual crystallinity left
upon release of the tension all show well-defined correlations with the average ethylene content and the ethylene
sequence length distribution. The tensile properties and the strain hardening behavior are correlated with the
incremental crystallinity achieved upon stretching at high strain.

1. Introduction (D)M elastomers are essentially due to the fully saturated and apolar
polymer backbones, the high chain flexibilities, the random distribu­
Ethylene/propylene (EPM) and ethylene/propylene/diene (EPDM) tions of comonomer units along the chains and the capability of the long
random copolymers (EP(D)M) with ethylene contents in the range of ethylene sequences to crystallize by cooling and deformation [10–22].
45–75 wt% are elastomers of high commercial relevance. They are The industrial production of EP(D)M is based on insertion poly­
characterized by outstanding properties, such as long-term resistance to merization facilitated by coordinating catalyst systems, typically in a
oxygen, ozone, heat, UV irradiation, water and polar solvents in general, solution process though an emulsion process is also possible [23,24].
even under extreme environmental conditions. As a result, EPDM rub­ The oldest and today still very important catalyst system consists of a
bers find numerous applications, such as in the automotive and electrical Ziegler-Natta vanadium-based pre-catalyst, such as VCl4, V(acac)3 (acac
cable industries and for the production of building and construction = acetylacetonate), or VOCl3, and an organoaluminum co-catalyst, such
materials, and many others [1–9]. In EPDM copolymers, ethylene and as AlEt3, AlEt2Cl or Al2Et3Cl3 [23–26]. The addition of a halogenated
propylene are copolymerized in the presence of non-conjugated dienes ester as a so-called promoter (e.g. chloroacetate, n-butyl per­
which are characterized by two C– – C unsaturations of different re­ chlorocrotonate, and trichlorotoluene) limits the deactivation tendency
activities towards insertion polymerization, such as 5-ethylidene-2-nor­ of the catalytic center during polymerization, thus leading to products
bornene (ENB), dicyclopentadiene (DCPD), or 5-vinyl-2-norbornene (EP(D)M − V) of high molar mass, narrow molar mass distribution and
(VNB). The easy incorporation of the strained unsaturations of the cyclic narrow chemical composition distribution [27,28]. However, the
norbornene units results in fully saturated polymer backbones bearing V-based catalyst systems suffer from a number of drawbacks, including a
the less reactive unsaturations in side groups. These pendant unsatura­ relatively low productivity, a relatively low affinity for the dienes and
tions, in turn, enable subsequent sulfur vulcanization and facilitate the need of operating at low temperatures (below 60 ◦ C), thus requiring
peroxide crosslinking reactions [1–9]. The outstanding properties of EP deep-cooling of the monomer/solvent feed in order to obtain high molar

* Corresponding author.
** Corresponding author.
E-mail addresses: finizia.auriemma@unina.it (F. Auriemma), miriam.scoti@unina.it (M. Scoti).

https://doi.org/10.1016/j.polymer.2021.123848
Received 1 March 2021; Received in revised form 14 April 2021; Accepted 1 May 2021
Available online 12 May 2021
0032-3861/© 2021 Elsevier Ltd. All rights reserved.
F. Auriemma et al. Polymer 227 (2021) 123848

mass EP(D)M− V products and prevent catalyst deactivation. Finally, the ethylene contents larger than or equal to 71 wt%, stretched at low de­
V-catalyst residues need to be removed from the rubber cement to obtain formations, crystallinity starts increasing already at deformations close
high-quality EP(D)M− V rubber products [8,23–36]. to the yielding (20–40%). In contrast, for the samples with 64 wt%
The second relevant catalytic family for EP(D)M production is based ethylene, the crystallinity increase starts only at strains higher than
on (post-)metallocene or so-called advanced catalysts. Currently, all 170%. In all cases, high degrees of orientation of the amorphous phases
major EPDM producers exploit (post-)metallocene catalysts, such as bis are achieved already at low deformations. In particular, for samples with
(substituted cyclopentadienyl) zirconium dichloride and dimethylsilyl high ethylene contents (>64 wt%), small, but significant amounts of
tetramethylcyclopentadienyl titanium dichloride. ARLANXEO has crystals are initially present at room temperature already in the unde­
developed and exploits the so-called Keltan® Advanced Catalyst Elas­ formed state. These pre-existing crystals undergo fragmentation and/or
tomer (ACE™) technology [23–25,36–39]. More than 50% of the global mechanical melting at low deformations and experience oriented
EPDM production capacity consists of products from (post-)metallocene recrystallization at higher deformations. The crystallinity increases
catalysts (EPDM-M). A major advantage of (post-)metallocene catalysts upon stretching, already at low deformations, not only as the result of
over the traditional V-based catalysts for the production of EPDM is the the high degrees of orientation achieved by the chains in the amorphous
much higher catalyst activity, resulting in the use of substantially lower phase, but also because the fragments of the initial crystals act both as
catalyst levels, thereby eliminating the need for deashing of the final physical crosslinks able to efficiently transmit the stress and as nuclei
rubber product. Another advantage is the much higher able to trigger further crystallization. Because of the non-trivial role of
reaction-temperature operating window, thus requiring less energy for the pre-existing crystals, this behavior was coined “strain-induced post-
the deep-cooling of the monomer/solvent reactor feed, the recycling of crystallization”.
unreacted monomers, and the solvent stripping. For EP(D)M samples with 64 wt% ethylene the amount of initially
EP(D)M rubbers are in general amorphous or low crystalline mate­ present crystals is very small though, making the crystallization
rials. Samples with an ethylene content larger than approximately 65 wt behavior upon stretching more similar to that of classic crystallizable
% are crystalline at room temperature and may experience further amorphous elastomers, such as natural rubber [51–55]. Indeed, the
crystallization either at low temperatures or by stretching [10–22].The crystallinity increase in these samples is “merely” induced by strain
bulky diene units are expelled from the crystals and reduce the length of (classic SIC), as it starts occurring only at high deformations, once the
the crystallizable ethylene sequences. The methyl side groups derived amorphous phase has reached a high degree of orientation [51–55].
from the propylene monomers are instead partially included in the Regardless of the ethylene content and the role of the pre-existing
crystals as defects and, thus, increase the effective length of the crys­ crystals, the crystallinity increases rather steeply after the onset. It
tallizable sequences [18,19,40–45]. The distribution of the ethylene then reaches a quasi-plateau value at about 100% deformation for the
sequence lengths and the chain microstructure in general are established more crystalline samples undergoing strain-induced post-crystallization.
in the polymerization step by the choice of the catalyst system and the For the low crystalline samples experiencing classic SIC, instead, the
reaction conditions, including the temperature, the pressure, the crystallinity increase reaches a quasi-plateau value at 200–300%
monomer feed ratio, and the chemical nature of the diene. In particular, deformation, as the initial crystals do not play a role.
under given copolymerization conditions, the relative comonomer To date, the relationships between the initial degree of crystallinity,
conversion and the resultant chain microstructure depend on the reac­ the mechanical properties and the maximum degree of crystallinity
tivity ratios of the ethylene, propylene and diene comonomers towards achieved by stretching of EP(D)M elastomers with the chain micro­
the active metal centers [46–50]. The distribution of the ethylene structure have been understood only in part, considering the sole role of
sequence lengths has a great influence on the crystallization and, comonomer content. The relationships between the crystallization and
consequently, on the mechanical properties of EP(D)M and on their mechanical properties of EP(D)M with the average length of ethylene
rheological behavior during mixing and processing. Therefore, the un­ sequence and the distributions of comonomers are only known quali­
derstanding of the effects of the chain microstructure on the final per­ tatively but have not been studied quantitatively. The work described in
formance of EP(D)M products is a fundamental issue. As an example, it this paper is aimed at filling this gap. In particular, the separate effects of
has been shown that the EP(D)M-M samples often achieve the perfor­ the ethylene content and the length of ethylene sequences on the crys­
mance shown by their EP(D)M− V counterparts at lower ethylene con­ tallization and tensile properties are outlined. Samples obtained with V-
tents, because of the greater tendency of (post-)metallocene catalysts based and (post-)metallocene catalysts, characterized by a large varia­
towards the incorporation of consecutive ethylene units rather than the tion of the ethylene, propylene and diene contents, are analyzed to
formation of hetero-sequences [23–25]. As a consequence, such EP(D) unravel to which extent the properties of these products are controlled
M-M products show similar crystallinities and properties as EP(D) by the average comonomer content and/or the differences in the
M− V products at somewhat lower ethylene contents. Furthermore, EP comonomer distribution.
(D)M-M products with similar ethylene contents as their EP(D)M− V
counterparts show a higher crystallinity and, hence, also different 2. Experimental section
properties, namely higher hardness, higher tensile strengths, but also
higher permanent deformations in low-temperature compression set Ethylene/propylene(/diene) copolymer samples with an ethylene
tests [23–25]. (E) content in the range of 44–77 wt% (57–83 mol%) have been
In this paper the mechanical properties and the structural trans­ analyzed. They have been prepared at ARLANXEO, using a V-based
formations occurring by effect of deformation of EP(D)M samples with Ziegler-Natta catalyst system or Keltan ACE™ technology in a solution
ethylene contents in the range of 57–83 mol% (44–77 wt%) are inter­ polymerization process, except for the EPDM sample with 74 mol% E,
preted in terms of the chain microstructure. The analysis is focused on which has been obtained with a V-based catalyst in a slurry process.
the evaluation of the tensile properties and the maximum degree of Except for the EPM samples with ≈77 wt% ethylene content, all the
crystallinity achieved by stretching at high deformations. The aim is to analyzed samples are commercially available in the ARLANXEO
establish a relationship between these properties and suitable micro­ portfolio.
structural parameters that describe the statistics of comonomer distri­ The characteristics of the EP(D)M samples are shown in Table 1. First
bution and the sequence length as established in the synthetic step. the EPM samples and next the EPDM samples are listed in the sequence
The strain induced crystallization (SIC) behavior occurring at low/ of decreasing ethylene contents and, thus, decreasing crystallinities.
moderate deformations of some V-based EP(D)M copolymers with They are coded as EPM-Xn and EPDM-Xn-Ym, respectively. X = V or M
ethylene contents in the range of 64–78 wt% (57–83 mol%) was stands for samples prepared with vanadium- (EPM-V or EPDM-V) or
analyzed in Ref. 16. It was shown that, for EP(D)M samples with (post-)metallocene-based (EPDM-M) catalysts in solution

2
F. Auriemma et al. Polymer 227 (2021) 123848

Table 1
Concentrations of ethylene (E), propylene (P), diene (ENB or DCPD) comonomeric units, Mooney viscosity and used catalyst of EPM and EPDM samples.
Sample E (wt%)a E (mol%)a P (wt%)a P (mol%)a ENB (wt%)a,b ENB (mol%)a,b DCPD (wt%)a,b DCPD (mol%)a,b Mooney viscosity Catalyst
ML (1 + 4) 125 ◦ C
(MU)c

Ethylene/propylene copolymers
EPM-V83 77.2 83.5 22.8 16.5 35 V
EPM-V80 72.4 79.7 27.6 20.3 25 V
EPM-V58 48.4 58.5 51.6 41.5 51 V
Ethylene/propylene/diene copolymers
EPDM-V80-N1.2 70.5 80.0 24.9 18.8 4.6 1.2 55 V
EPDM-V78-N0.4 70.0 78.3 28.5 21.3 1.5 0.4 59 V
EPDM-M78-N1.2 68.1 77.9 27.3 20.8 4.6 1.2 55 M
EPDM-M78-N1.3 67.6 77.7 27.4 21.0 5.0 1.3 80 M
EPDM-Q74-N0.8 64.6 74.3 32.2 25.0 2.8 0.8 67 V, slurry
EPDM-V73-D0.3 64.0 73.2 34.8 26.5 1.2 0.3 63 V
EPDM-V62-N1.2 51.1 62.4 44.6 36.3 4.3 1.2 46 V
EPDM-M57-N2.7 44 56.8 47 40.5 9.0 2.7 60 M
a
Determined by high-resolution13C NMR solution spectroscopy (vide infra).
b
ENB = 5-ethylidene-2-norbornene; DCPD = dicyclopentadiene.
c
Mooney viscosity (in Mooney units MU) determined using a large rotor at 125 ◦ C, with 1 min preheating, using the final torque after 4 min of testing as a measure of
molar mass.

polymerization processes, respectively, whereas X = Q stands for the Wide angle X-ray diffraction (WAXS) profiles were recorded with an
sample prepared with a vanadium-based catalyst in slurry polymeriza­ Empyrean diffractometer (Malvern Panalytical) using Ni-filtered CuKα
tion process (EPDM-Q). The label “n” indicates the ethylene mole pre­ radiation (λ = 0.15418 nm). The degree of crystallinity xc(WAXS) was
cent (mol%), whereas Y = N or D stands for ENB or DCPD diene evaluated from the area subtending the WAXS profiles of the sample (As)
monomers, respectively, and “m” indicates the corresponding mol% and of the amorphous phase (Aa) as xc(WAXS) = 100 (As-Aa)/As. The
content. amorphous contribution was approximated by the diffraction profile of
The Mooney viscosities [ML (1 + 4) 125 ◦ C] of the EPM samples the melt (at 120 ◦ C) after suitable scaling and translation along the x-
range from 25 to 51 MU and of the EPDM samples between 46 and 80 axis, up to make the position of the maximum coincident with the
MU, corresponding to mass average molar masses Mw between 150 and maximum of the amorphous phase at 25 ◦ C.
175 kg/mol and between 200 and 300 kg/mol, respectively. The EPDM Bidimensional (2D) WAXS patterns (Ni-filtered CuKα radiation) were
samples contain about 1 mol% of ENB, except for the sample EPDM- recorded on fibers obtained by stretching compression-molded films at
M57-N2.7 with the lowest ethylene content (56.8 mol%), which con­ different degree of deformations, while keeping the sample in tension.
tains 2.7 mol% of ENB, and the sample EPDM-V73-D0.3 with 73.2 mol% The 2D-WAXS images were recorded in a cylindrical camera (radius
ethylene, which contains 0.3 mol% DCPD. 57.30 mm) on BAS-MS imaging plates (FUJIFILM) and digitalized with a
Solution NMR spectra were recorded with a Bruker Avance III 400 reader PerkinElmer Cyclone Plus (storage phosphor system). Radial
spectrometer equipped with a 5 mm high-temperature cryoprobe. The profiles were extracted from the 2D WAXS images with the help of the
samples (~25 mg) are dissolved at 120 ◦ C in 1,2-dichlorobenzene-d4 Fit2D and ImageJ softwares.
(0.6 mL) with 0.40 mg mL− 1 of BHT stabilizer. Operating conditions for Mechanical tests were performed at room temperature by stretching
1
H NMR spectroscopy are: 90◦ pulse; acquisition time, 2.7 s; relaxation rectangular specimens having width, length and thickness equal to 5,
delay, 10.0 s; 1 to 16 transients. Operating conditions for 13C NMR 100 and 0.5 mm, respectively, cut from the compression molded films,
spectroscopy are: 45◦ pulse; acquisition time, 2.7 s; relaxation delay, 5 s; using an Instron 5566H1543 mechanical tester, following the ASTM
2000 transients. Broad-band 1H decoupling was achieved with a modi­ D882-83 standard. The deformation rate (v) was set at 10 and 0.1 times
fied WALTZ16 sequence (BI_WALTZ16_32 by Bruker). ENB or DCPD the initial gauge length L0 (that is v = L0 10 and L0 0.1 mm min− 1,
contents have been determined by 1H NMR spectroscopy. Average respectively) for the determination of the tensile properties up to
ethylene and propylene contents, and the ethylene sequence lenghts breaking and of the Young modulus, respectively.
have been determined by 13C NMR spectroscopy using the method The recovered strain after breaking (elastic recovery) was measured,
described in Ref. 56. The 13C NMR resonances have been assigned ac­ according to the ASTM D412-87 standard, on compression molded
cording to Refs. [57, 58]. specimens stretched up to breaking (final deformation εb = 100(Lb-L0)/
Films with a uniform thickness of approximatively 0.5 mm and L0, with Lb the final length achieved at break and L0 the initial gauge
smooth surfaces were prepared by compression molding in a hydraulic length). Ten minutes after breaking, the lengths of the fully relaxed
press by heating the samples, sandwiched between flat Teflon plates, at specimens Lr were measured by moving the two broken pieces close to
temperatures of ≈30–40 ◦ C higher than the melting temperature (i.e. each other till to fit them along the fracture line. The percentage of
100-150 ◦ C) for the crystalline samples, 100 ◦ C for the amorphous recovered strain after breaking is calculated as R = 100(Lb-Lr)/(Lb-L0).
samples, while applying only a small pressure (<1 bar) in order to not The values of L0, Lb and Lr were determined from the distances between
induce any preferred chain orientation, according to the ASTM D-2292- two marks, drawn on the samples with a pencil having a smooth tip, in
85 standard. After 5 min isotherm at 100–150 ◦ C, the films were slowly the transverse direction to the gauge length. The reported stress strain
cooled to 25 ◦ C at average rate of ≈15 ◦ C/min. All subsequent analyses curves and the values of the mechanical parameters are the result of the
were performed on these films. average of a minimum of five independent reproducible experiments.
Melting thermograms were recorded with a differential scanning
calorimeter (DSC-1) by Mettler Toledo under a N2 atmosphere with a 3. Results and discussion
scanning rate of 10 ◦ C/min. The degree of crystallinity xc(DSC) was
evaluated as xc(DSC) = 100 ΔHm/ΔH0m, with ΔHm the melting enthalpy Solution 13C NMR spectroscopy. All samples of EP(D)M copolymers
of the sample and ΔH0m = 293 J/g the thermodynamic melting enthalpy have been analyzed by 13C NMR spectroscopy to evaluate the chain
of a 100% crystalline polyethylene (PE) [59]. microstructure in terms of composition, comonomers distribution, and

3
F. Auriemma et al. Polymer 227 (2021) 123848

lengths of ethylene sequences. 13CNMR spectra of representative sam­ It is worth noting that the results of solution 13C NMR analysis of
ples are reported in the Supporting Information (Fig. S1). The statistical Table 2 are in agreement with literature data, indicating that in the
analysis of comonomers distribution has been performed using a five- copolymerization of propylene with ethylene the conventional V-based
parameter model, considering all samples as terpolymers constituted catalyts [60–63] and the less-conventional (“constrained geometry”)
by three monomers (viz. ethylene, primary 1,2-propylene and secondary (post-)metallocene catalysts [64] are both non regiospecific, even
2,1-propylene units), following the procedure of Ref. 56. Furthermore, though for the copolymers synthesized with the latter systems the con­
in the statistical analysis, the contribution of the diene has been centration of regiodefects (viz. 2,1 propylene units, P*) is lower.
neglected in a first approximation, since its content is by far much lower Furthermore, both classes of catalysts systems are poorly stereospecific
than that of the other comonomers and the available fifteen independent [60–64].
experimental data are not sufficient to handle a more complex statistical For EP(D)M samples with ethylene content in the range 73–83 mol%,
model. Therefore, the statistical analysis has been performed only for the average length of ethylene sequences LM1 determined using a first
the EP(D)M copolymers with diene contents below or around 1 mol%. order Markov statistics of comonomer addition [57] is comprised in the
The mole fraction of ethylene units sequences of length n (p(E ≥ n)) range 3–6 (Table 2). The so obtained values of LM1 are slightly smaller
with respect to ethylene sequences of all lengths has been calculated than those calculated using a Bernoulli type statistics LBe as the inverse
according to the following general equation: of the sum of the contents of propylene (P) and diene (D) units (LBe =
100/(P + D)), in agreement with the tendency to alternation denoted by
p(E≥n) = E pn11 + P pn−11 1 p21 + P* pn−11 1 p31 (1)
the values of the apparent reactivity ratios product smaller than 1.0.
Using both statistics, no great differences have been found in the values
where pij are the conditional probabilities of insertion of a monomer j
of the average length of ethylene sequences for the samples EP(D)M− V
after a monomer i, (with 1 = ethylene, 2 = 1,2-propylene, 3 = 2,1-
and EP(D)M-M having similar ethylene concentration of 78–80 mol%
propylene), and E, P, P* are the mole fractions of ethylene, 1,2-propyl­
(LM1 = 4.2–4.5 and LBe = 4.5–4.6 for the samples EPDM-V78-E0.4,
ene and 2,1-propylene, respectively.
EPDM-M78-N1.2 and EPDM-M78-N1.3 in Table 2), which confirms
The conditional probabilities pij have been determined by best-fit
the fact that the differences between V-based and (post-)metallocene
calculations on sequence distribution, determined by 13C NMR
samples are more related to the effective sequence distribution of
spectra, according to the method reported in ref. 56. In all cases the
comonomeric units.
statistical analysis ended up with a χ2r close to 1, proving the goodness of
Thermal and structural analysis. The WAXS profiles of the
fit testing. The conditional probabilities values for all samples are re­
compression molded films and the corresponding DSC thermograms
ported in Table S1 of the Supplementary Data file, while the relevant
recorded starting from 25 ◦ C are shown in Fig. 1.
microstructural characterization results are summarized in Table 2.
The WAXS profiles of the EP(D)M samples of Fig. 1A are typical of
From the conditional probabilities values the apparent reactivity
such polymers. The diffraction peak at 2θ ≈ 21◦ corresponds to the
ratios product “rE rP” reported in Table 2 have been calculated by Eq. (2),
(110)o and/or (100)h reflections of the orthorhombic or pseudohex­
including in rp the presence of both 1,2 and 2,1-propylene units [56]:
agonal forms of PE, respectively (curve a in Fig. 1A) [16,40,45].1 Only
˝rE rP ˝ =p11 [NP (p22 +p23 )+NP∗ (p32 +p33 )]/[(p12 +p13 )+(NP p21 +NP∗ p31 )] for the sample EPM-V83 with 83 mol% ethylene, this reflection appears
(2) as a sharp peak at 2θ ≈ 21◦ , overlaying the amorphous halo (curve a in
Fig. 1A) [16]. With a decrease of the ethylene content, the (110)o/(100)h
In Eq. (2), NP and NP* are given by Eqs. (3) and (4) as it follows:
reflection is hardly distinguishable, as it appears as a faint shoulder (at
NP = K(p32 p12 + p32 p13 + p31 p12 ) (3) 2θ ≈ 21◦ ) of the amorphous halo in the WAXS profiles of the samples
with ethylene contents in the range 80 to 73 mol% (curves b and d to i in
NP∗ = K(p13 p23 + p13 p21 + p12 p23 ) (4) Fig. 1A). It is completely absent for the samples with ethylene contents
in the range 57–62 mol%, for which only an amorphous halo is apparent
where K is an arbitrary proportionality constant. (curves c, j and k in Fig. 1A).
It should be recalled that, since the V-based Ziegler-Natta catalysts The presence of crystals is better revealed in the DSC thermograms
are considered multi-site systems [23–29], the values of the apparent shown in Fig. 1B. The samples with ethylene contents in the range
reactivity ratios product (“rE rP”) are average values over different cat­ 73–83 mol% show broad melting peaks at temperatures in the range
alytic species producing macromolecules with different composition. 40–50 ◦ C. The peak areas tend to decrease with a decrease of the
In all cases, the “rE rP” values are lower than 1.0, indicating the ethylene content (curves a, b and d to i in Fig. 1B). The absence of
tendency to alternation (Table 2). However, the EP(D)M-M samples endothermic peaks in the DSC thermograms of the samples with
exhibit slightly higher apparent reactivity ratios product values ethylene contents in the range 57–62 mol% confirms that the samples
(0.66–0.69) compared with EP(D)M− V copolymers (0.45–0.60), prob­ are amorphous (curves c, j and k in Fig. 1B). Therefore, for the EP(D)M
ably reflecting the major propensity of the adopted (post-)metallocene samples with ethylene content in the range 83 to 77 mol%, whereas the
catalysts towards the incorporation of consecutive ethylene units rather Bragg peak at 2θ ≈ 21◦ in the WAXS profiles becomes increasingly
than the formation of hetero-sequences with respect to the V-based weaker and broader with decreasing the ethylene content, up to become
counterparts [23]. barely detectable, the corresponding melting endotherms remain better
In all EP(D)M− V samples from V-based catalyst about 20 mol% of defined. The decrease of the intensity and the increase of the width of
the total propylene content consists of 2,1-propylene units P*, whereas the Bragg reflections with decreasing ethylene content are due to the
in EPDM-M samples from (post-)metallocene catalyst the P* content is easy inclusion of the propylene methyl groups in the unit cells of the
“only” about 2–3 mol% (Table 2). This difference may have relevant orthorhombic and/or pseudo-hexagonal forms of PE on the one hand
effects on the microstructure of the chains, since the presence of P* units and to the decrease of the crystallizable ethylene sequence fractions on
is known to favor ethylene insertion (p31 > p33 and p32 for all samples in
Table S1). Contrary to (post-)metallocene catalyst systems, for V-based
systems, after an ethylene insertion, the propylene insertion can be both 1
The crystallization of the pseudo-hexagonal form in EP(D)M rubbers is
1,2 and 2,1 (p12 and p13 > 0 in Table S1) [60]. As a result, the presence of
driven by the easy inclusion of propylene units in the orthorhombic unit cell of
a high fraction of secondary 2,1 propylene units in EP(D)M− V samples
PE. In fact, as the concentration of propylene units increases, the size of the a-
produces sequences of 2,1 propylene units followed by ethylene and
axis of the orthorhombic unit cell of PE increases, while the size of the b- and c-
both primary 1,2- or secondary 2,1-propylene units [60–63], which are axes remains almost constant. The pseudo-hexagonal form corresponds to a
absent in EPDM-M. length of the a-axis approaching the value of 31 ⁄ 2 times the b-axis.

4
F. Auriemma et al. Polymer 227 (2021) 123848

Table 2
Mole percent of ethylene (E), 1,2-propylene (P), diene (D) comonomeric units and mole percent of 2,1-propylene units (P*) relative to the total propylene content,
apparent reactivity ratios product (“rE rP”), mole fractions of ethylene sequences of length n (p(E ≥ n)) relative to sequences of ethylene units of all lengths, and average
length of consecutive ethylene sequences (LM1 and LBe) for EP(D)M copolymers.a
Sample E (mol%) P (mol%) Diene (mol%) P* (%)b “rE rP” p(E ≥ 6) p(E ≥ 7) p(E ≥ 8) p(E ≥ 9) p(E ≥ 10) p(E ≥ 11) p(E ≥ 12) p(E ≥ 13) p(E ≥ 14) p(E ≥ 15) LM1c LdBe

EPM-V83 83.5 16.5 18 0.60 0.32 0.26 0.22 0.18 0.15 0.12 0.10 0.08 0.07 0.06 5.7 6.1
EPM-V80 79.7 20.3 20 0.45 0.22 0.17 0.13 0.10 0.08 0.06 0.05 0.04 0.03 0.02 4.5 4.9
EPM-V58 58.8 41.5 24 0.47 0.02 0.01 <0.01 <0.01 <0.01 <0.01 0.00 0.00 0.00 0.00 2.0 2.4
EPDM-V80-N1.2 80.0 18.8 1.2 22 0.46 0.25 0.20 0.16 0.12 0.10 0.08 0.06 0.05 0.04 0.03 4.8 5.0
EPDM-V78-N0.4 78.3 21.3 0.4 22 0.45 0.20 0.15 0.12 0.09 0.07 0.05 0.04 0.03 0.02 0.02 4.2 4.6
EPDM-M78- 77.9 20.8 1.2 3 0.66 0.22 0.17 0.13 0.10 0.08 0.06 0.05 0.04 0.03 0.02 4.5 4.5
N1.2
EPDM-M78- 77.7 21.0 1.3 2 0.69 0.22 0.17 0.13 0.10 0.08 0.06 0.05 0.04 0.03 0.02 4.4 4.5
N1.3
EPDM-Q74-N0.8 74.3 25.0 0.8 20 0.58 0.15 0.11 0.08 0.06 0.04 0.03 0.02 0.02 0.01 <0.01 3.6 3.9
EPDM-V73-D0.3 73.2 26.6 0.3 20 0.59 0.13 0.09 0.06 0.05 0.03 0.02 0.02 0.01 <0.01 <0.01 3.4 3.7
EPDM-V62-N1.2 62.4 36.3 1.2 21 0.46 0.04 0.02 0.01 <0.01 <0.01 <0.01 <0.01 <0.01 0.00 0.00 2.3 2.7
a
The calculations of “rE rP” and p(E ≥ n) values were performed only for EP(D)M copolymers with diene contents below or around 1 mol%.
b
Relative value calculated with respect to propylene units only.
c
Calculated using a first-order Markov statistics of comonomer addition.d)Calculated as LBe = 100/(P + Diene).

the other hand [13]. This leads to an increase of the structural disorder at approximately 50 ◦ C and a small endotherm at around 112 ◦ C, the
in the crystals, to a decrease of the crystal size and to the consequent loss latter due to the presence of a small fraction of PE crystals formed by
of the long-range order in the chain conformation and packing [12–15, long ethylene sequences (Table 3). It is worth noting that the melting
22]. temperature of V-based EP(D)M samples with ethylene content of 78–80
The glass transition temperatures of the EP(D)M samples (deter­ mol% of about 41 ◦ C, is lower than that of the samples EPDM-M78-N1.2
mined from independent DSC scans recorded starting from − 100 ◦ C) are and EPDM-M78-N1.3 of similar ethylene content, synthesized with
in between − 35 ◦ C for the sample EPM-V83 and -59 ◦ C for the sample (post-)metallocene catalysts that melt at about 44 ◦ C (Table 3). These
EPDM-V62-N1.2 (Table 3). The main melting peaks of the EP(D)M differences may be due to the lower tendency of the (post-)metallocene
samples occur in the temperature range from 40 to 44 ◦ C regardless of catalysts toward alternation and/or to the lower concentration of
ethylene content, except for the EPM sample with the highest ethylene regiomistakes (viz. 2,1-propylene units P*) in the resultant copolymers
content EPM-V83 (83.5 mol% ethylene) which shows a main endotherm (Table 2).

Fig. 1. WAXS profiles (A) and DSC thermograms (B) of compression-molded EP(D)M samples. In A, the (110)o and/or (100)h reflection of the orthorhombic or
pseudo-hexagonal forms of PE, at 2θ ≈ 21◦ , respectively, is indicated with arrows, whereas the narrow diffraction peaks at 2θ ≈ 6, 9, 12, 19, 25, 28 and 31◦ observed
for several samples (curves b, and d to i) and indicated by an asterisk for EPM-V80 (curve b) are due to talc added to the samples to prevent sticking. The dashed lines
in B indicate the baselines adopted for the calculation of the melting enthalpies, drawn by extrapolating the high-temperature DSC tails.

5
F. Auriemma et al. Polymer 227 (2021) 123848

It is worth noting that the effect of regio- (and stereo)irregularities of lengths that are intermediate between those of the samples EPDM-V78-
propylene insertion on the crystallization behavior of ethylene/propyl­ N0.4, with identical ethylene content, and EPDM-V80-N1.2, with 80 mol
ene copolymers has been investigated in Ref. [65]. It has been shown % ethylene content, both synthesized with V-based catalysts. The length
that the presence of both regio- and stereoirregular propylene counits weighed fractions of ethylene sequences of length higher than 7 (cor­
induces disturbance on crystallization, but regioirregularities of pro­ responding to the integrated area of the distribution function of ethyt­
pylene insertion induce higher disturbance than stereoirregularities. As lene sequence lengths for sequence lengths higher than or equal to 7) are
the analyzed EP(D)M copolymers are both stereo- and regioirregular, 51% for the sample EPDM-V80-N1.2, 43% for the sample EPDM-V78-
the differences in the crystallization behavior are small. Therefore, the N0.4 and 48–49% for the samples EPDM-M78-N1.2 and EPDM-M78-
slightly higher melting temperature of the samples EPDM-M78-N1.2 and N1.3. These small differences reflect only in part the small differences
EPDM-M78-N1.3 of similar ethylene content, synthesized with (post-) in the values of the crystallinity index xc(WAXS) equal to 10, 8, 8 and
metallocene catalysts, than the melting temperature of the V-based co­ 8%, respectively, indicating that for regio- and stereoirregular random
polymers EPDM-V80-N1.2, EPDM-V78-N0.4, with 80 and 78 mol% EPDM copolymers, the leading parameter controlling the crystallinity
ethylene content, respectively, should be ascribed not only to the index achieved by melt-crystallization is the ethylene content.
different concentration of 2,1 counits (P*), but also to the small differ­ Tensile properties. The results of the tensile testing of the EP(D)M
ences in the co-monomer sequence distribution, associated with the samples are shown in Fig. 3. The corresponding values of the Young
lower tendency of the (post-)metallocene catalysts toward propagation modulus, strain at break and stress at break are reported in Fig. 4 and
of alternating ethylene/propylene sequences. The occurrence of these Table S2.
small microstructural differences probably accounts also for the fact that All the samples show uniform deformation up to breaking. The EP(D)
the melting endotherms of the samples EPDM-M (curves f and g of M samples with ethylene contents larger than 62 mol% show diffuse
Fig. 2B) appear slightly narrower than those of the EPDM-V systems yielding and strain at break larger than ≈1000% (Fig. 3A and B). The
(curves d and e of Fig. 2B). samples with ethylene contents larger than 73 mol% show remarkable
The values of the crystallinity index extracted from WAXS data strain hardening (Fig. 3A and B). The amorphous and sticky samples
(xc(WAXS)) and DSC analysis (xc(DSC)) (Table 3) are reported in Fig. 2 with ethylene contents smaller than or equal to 62 mol% show no
as a function of the ethylene content (A) and of the mole fraction of yielding, no strain hardening and strain at break smaller than 200%
ethylene sequences of length 7 (p(E ≥ 7)) relative to the ethylene se­ (Fig. 3C).
quences of all lengths (B). The parameter p(E ≥ 7) is selected based on the The mechanical parameters extracted from Fig. 3 are reported in
consideration that the minimum length of crystallizable ethylene se­ Fig. 4 as a function of the ethylene content and the mole fraction of
quences in ethylene/α-olefin copolymers is reported to correspond to a ethylene sequences of length 7 (p(E ≥ 7)) relative to the ethylene se­
number of consecutive ethylene units in the range 4–18 [49,66–70]. It is quences of all lengths. It is apparent that the values of the Young
apparent that the values of xc(WAXS) and xc(DSC) increase with modulus (Fig. 4A and A′ ) increase and the values of strain at yield
increasing ethylene content (Fig. 2A) and the mole fraction of ethylene decrease (Fig. 4D and D’) with increasing both ethylene content and the
sequences of length 7 (p(E ≥ 7)) relative to the ethylene sequences of all fraction p(E ≥ 7), according to single trends regardless of the catalyst
lengths (Fig. 2B). However, the plot of Fig. 2B allows highlighting the system used for the preparation of the samples. This indicates that these
effect of the presence of diene units because it shows that the values of properties are not sensitive to the details of the chain microstructure for
xc(WAXS) and xc(DSC) of EPM copolymers are larger than those of the the analyzed EP(D)M systems but are only dependent on the ethylene
corresponding EPDM samples with identical p(E ≥ 7) values. This in­ content.
dicates that even a small amount of diene co-units may induce a large It is also apparent that with increasing of the ethylene content and of
disturbance in the crystallization properties of EPDMs. the p(E ≥ 7) fraction, the values of the strain at break (Fig. 4B and B′ )
To an indepth insight, the distribution function of ethytlene sequence increase for the EPM copolymers, whereas for the EPDM copolymers
lengths in the samples EPDM-V80-N1.2, EPDM-V78-N0.4, EPDM-M78- they first increase and then they decrease. The values of the stress at
N1.2 and EPDM-M78-N1.3 are compared in Fig. S2. It is apparent that break, instead (Fig. 4C and C′ ), increase monotonically with increasing
the samples EPDM-M78-N1.2 and EPDM-M78-N1.3 with 78 mol% the ethylene content and the fraction p(E ≥ 7) for both EPM and EPDM
ethylene content and synthesized with (post-)metallocene catalysts copolymers, following two distinct trends. In particular, for identical
show nearly coincident distribution functions of ethylene sequence values of ethylene content and p(E ≥ 7) values, the values of stress at

Table 3
Ethylene content (E), glass transition temperatures (Tg), melting temperature (Tm) and enthalpy (ΔHm) and crystallinity indices determined from
WAXS profiles (xc(WAXS)) and DSC thermograms (xc(DSC)) of compression molded samples of EP(D)M copolymers, corresponding increment of crystallinity achieved
by stretching at maximum deformation (Δxcs) and residual crystallinity achieved after releasing the tension (Δxcr) evaluated from X-ray fiber diffraction
patterns of EP(D)M samples (Fig. 5).
Sample E (mol%) xc(WAXS) (%) Tg (◦ C)a Tm (◦ C) ΔHm (J/g) xc(DSC)b (%) Δxcs (%) Δxcr (%)

EPM-V83 83.5 16 − 35 50.2; 112c − 43 14 13 10


EPM-V80 79.7 10 − 46 41.2 − 19 6 16 10
EPM-V58 58.5 – − 53 – – – – –
EPDM-V80-N1.2 80.0 10 − 40 41.5 − 20 7 17 6
EPDM-V78-N0.4 78.3 8 − 45 40.4 − 12 4 15 5
EPDM-M78-N1.2 77.9 8 − 40 43.9 − 13 4 16 5
EPDM-M78-N1.3 77.7 8 − 40 43.7 − 13 4 16 4
EPDM-Q74-N0.8 74.3 5 − 46 40.1 − 3 1 12 3
EPDM-V73-D0.3 73.2 3 − 50 41.6 − 2 1 10 1
EPDM-V62-N1.2d 62.4 – − 59 – – – – –
EPDM-M57-N2.7d 56.8 – − 43 – – – – –
a)
Determined from independent DSC scans recorded starting from − 100 ◦ C.
b)
Calculated as xc(DSC) = 100 ΔHm/ΔH0m, with ΔHm the melting enthalpy of the sample and ΔH0m = 293 J/g the thermodynamic melting enthalpy of a 100%
crystalline polyethylene (PE) [59].
c)
The main endothermic peak is accompanied by a small endotherm at 112 ◦ C, due to melting of very long ethylene sequences in thicker PE crystals.
d)
The samples EPM-V58, EPDM-V62-N1.2 and EPDM-M57-N2.7 are amorphous, so that there exist no crystallinity data.

6
F. Auriemma et al. Polymer 227 (2021) 123848

Fig. 2. Crystallinity indices determined from WAXS profiles (xc(WAXS)) and DSC thermograms (xc(DSC)) of EP(D)M samples as a function of the ethylene content
(A) and the mole fraction of ethylene sequences of length 7 (p(E ≥ 7)) relative to the ethylene sequences of all lengths (B).

break is somehow related to the fact that samples with ethylene content
larger than 73 mol% display remarkable strain hardening the amount of
which is different depending on the occurrence of further crystallization
during deformation (vide infra). The low values of the stresses and
strains at break of the amorphous samples are due to the absence of any
crystallinity. This indicates that in absence of crystals acting as the
physical cross-links of the amorphous network and as efficient stress
transmitters, the entanglement density is not sufficiently high to prevent
the viscous flow of the chains and/or the breaking already at low de­
formations (Fig. 4B, B′ , C and C’).
The values of the stress at yield of the EPM and EPDM samples tend
to increase according to a single trend with increasing the ethylene
content (Fig. 4E) and according to two different trends with increasing
the values of the fraction p(E ≥ 7) (Fig. 4E’). In particular, it is apparent
that for identical values of p(E ≥ 7), the values of the yield stress of the
EPM copolymers are larger than those of EPDM samples. This indicates
that the presence of about 1 mol% of diene co-units induces a small but
significant decrease of the stress at yield, even though the values of
strain at yield of the samples are identical (Fig. 4D’). The values of the
ultimate tensile strength of EPDM samples, instead, are larger than those
of EPM copolymers (Fig. 4C,C’), due to occurrence of strong strain
hardening, which is related to the possible occurrence of further crys­
tallization during deformation (vide infra).
Finally, the percentage of recovered strain after breaking R (Fig. 4F
and F’) are in all cases larger than 70%. In particular, the crystalline
EPDM copolymers with ethylene content from 75 to 80 mol% and values
of the fraction p(E ≥ 7) larger than 0.09–0.1 show the best elastic per­
formances with R values larger than 90%. The remarkable elastic per­
formances of the crystalline un-crosslinked EP(D)M copolymers suggest
that the initial crystals present in the undeformed samples play a key
role. Indeed, not only they act as efficient physical knots of the elasto­
meric network, preventing the viscous flow of the chains during defor­
mation, but they may also play a role, after fragmentation, as nuclei in
Fig. 3. Stress-strain curves of EP(D)M samples. Dots mark the approximate
triggering further crystallization of the samples during deformation
beginning of strain hardening. A: EP(D)M samples with high ethylene contents,
(vide infra) [16].
strong strain hardening and high stress at break; B: EP(D)M samples with in­
termediate ethylene contents, some strain hardening and intermediate stress at
In the next section, it is shown that the amount of the new crystals
break; C: EP(D)M samples with low ethylene contents without any strain that are formed by stretching at deformations close to the break, that is
hardening and low stress at break. up to reach the maximum increment of crystallinity, depends on the
initial level of crystallinity in the undeformed state.
break of EPM copolymers are smaller than those of EPDM terpolymers. Fiber diffraction analysis and strain-induced crystallization. The X-
For EP(D)M samples with ethylene content larger than 70 mol% ray fiber diffraction patterns and the corresponding radial profiles of
(Fig. 4C) and p(E ≥ 7) values larger than 0.09–0.1 (Fig. 4C’), the increase selected EP(D)M samples are presented in Fig. 5. For the sake of
of the stress at break with ethylene content reflects the increase of initial simplicity, and without loss of generality, only the patterns recorded for
crystallinity of the samples (Fig. 2), but the different values of stress at the undeformed samples and for specimens stretched up to a

7
F. Auriemma et al. Polymer 227 (2021) 123848

Fig. 4. Tensile properties of EP(D)M samples as a function of the ethylene content (A to F) and the mole fraction of ethylene sequences of length 7 (p(E ≥ 7)) relative to
the ethylene sequences of all lengths (A′ to F′ ). Young modulus (A and A′ ), strain (B and B′ ) and stress (C and C′ ) at break, strain (D and D′ ) and stress (E and E′ ) at
yield, and elastic recovery (F and F′ ).

deformation close to breaking, and after the release of the tension are diffraction patterns of the sample EPM-V83 with the highest ethylene
reported in Fig. 5. It is worth noting that the maximum deformation content shows the presence of a sharp peak at 2θ ≈ 21◦ , corresponding to
achieved before breaking is higher than or equal to 0.95 εb, for all the the (110)o/(100)h reflection of the orthorhombic/pseudo-hexagonal
samples. It has been checked that, at these deformations, the samples forms of PE (Fig. 5A,A’). At high deformations, the samples achieve
have well-reached the maximum increment of crystallinity by effect of high degrees of orientation, with the chain axes parallel to the stretching
strain. Indeed, the maximum increment of crystallinity is already direction, as indicated by the polarization of the reflection at 2θ ≈ 21◦ on
reached at much lower deformations, close to 0.4–0.5 times the defor­ the equator (Fig. 5B, E, H and K). As shown in Ref. [16], a high degree of
mation at break (data not shown). orientation is achieved by the chains belonging not only to the crystals,
As already visualized in the X-ray powder diffraction profiles of but also to the amorphous phase, already at low deformations.
Fig. 1A, the radial WAXS profiles of the undeformed samples show only In particular, in the radial WAXS profiles of the samples stretched at
amorphous halos centered at 2θ ≈ 19◦ (Fig. 5D’, G′ and J′ ). Only the high deformations, the (110)o/(100)h reflection appears well-

8
F. Auriemma et al. Polymer 227 (2021) 123848

Fig. 5. X-ray fiber diffraction patterns (A to L) and corresponding radial profiles (A′ to L′ ) of the samples EPM-V83 (A to C and A′ to C′ ), EPDM-V78-N0.4 (D to F and
D′ to F′ ), EPDM-M78-N1.2 (G to I and G′ to I′ ) and EPDM-V73-D0.3 (J to L and J′ to L′ ). The radial profiles are corrected for the background contributions,
approximated by straight lines.

pronounced for the sample EPM-V83 (Fig. 5B’), less intense, but as a still samples, respectively (Fig. S4).
distinguishable reflection overlaying the amorphous halo for the sam­ As mentioned above, upon releasing the tension the high degree of
ples EPDM-V78-N0.4 (Fig. 5E’) and EPDM-M78-N1.2 (Fig. 5H’), and is orientation achieved by stretching is only partially lost (Fig. 5C, F, I and
completely concealed by the amorphous halo for the sample EPDM-V73- L). Moreover, the patterns of Fig. 5C, F, I and L seem to indicate that
D0.3 (Fig. 5K’). Indeed, for the latter sample, the (110)o/(100)h reflec­ partial melting of the new oriented crystals formed by stretching also
tion appears distinguishable only in the two-dimensional WAXS image occurs after releasing the tension [16]. The residual incremental degrees
(Fig. 5K) and in the corresponding equatorial profile (Fig. S3). of crystallinity after releasing the tension, with respect to the unde­
Upon release of the tension, the samples recover their initial di­ formed samples, have been also calculated as Δxcr = 100(APr–APu)A−Pr1,
mensions almost completely (Fig. 4F and F’), while the high degree of where APr is the area subtending the radial WAXS profile of the relaxed
orientation achieved in the stretched states by the chains in the amor­ samples Pr. The so calculated values are reported in Table 3 and in Fig. 6
phous [16] and crystalline phases is lost at least in part (Fig. 5C, F, I and as a function of the ethylene content and the mole fraction of ethylene
L). The structural and textural changes observed for the samples shown sequences of length 7 (p(E ≥ 7)) relative to the ethylene sequences of all
in Fig. 5 is common to all the EP(D)M samples of Table 1 with ethylene lengths. The values of total crystallinity index, achieved by stretching at
contents larger than or equal to 73 mol%. The samples with ethylene high deformation and after releasing the tension, are also reported. This
contents smaller than or equal to 62 mol%, instead, remain amorphous parameter is calculated as the product xc(WAXS) (1+ Δxcs/100) and
even upon stretching. xc(WAXS) (1+ Δxcr/100) for the stretched and relaxed samples,
The EP(D)M samples show small tendencies to crystallize upon respectively.
stretching.16 In particular, it was shown that for samples with ethylene It is apparent that the increment of crystallinity achieved by the EP
contents larger than or equal to 78 mol% the crystallinity starts to in­ (D)M samples upon stretching increases with increasing the ethylene
crease already at low deformations by effect of stretching [16]. For less content and the p(E ≥ 7) fraction up to reach a maximum around 80 mol%
crystalline samples (ethylene content ≈ 73 mol%), the onset of further and 0.20, respectively, then decreases for the sample EPM-83 with 83.5
crystallization occurs at strains higher than 170%. In both cases, the mol% ethylene content (Fig. 6A and B). Upon release of the tension, the
crystallinity increment tends to level off at deformations of 100–300% samples show a residual incremental crystallinity, which is lower than
[16]. Using the fiber diffraction patterns of EP(D)M samples stretched at that achieved upon stretching. Furthermore, whereas the residual in­
deformations close to the break, the maximum increment of crystallinity cremental crystallinity of EPDM samples increases with the increase of
achieved upon stretching Δxcs may be extracted by comparing the radial the ethylene content and the p(E ≥ 7) fraction, it is constant for the EPM
WAXS profiles of the highly stretched samples Ps (e.g. Fig. 5B’,E′ , H′ and samples. In particular the EPM copolymers show a residual incremental
K′ ) with the corresponding radial WAXS profiles of the unstretched crystallinity higher than that of EPDM samples (Fig. 6A and B). This is in
samples Pu (e.g. Fig. 5A’,D′ , G′ and J’), as shown in Fig. S4. In practice, agreement with the smaller values of elastic recovery shown by the EPM
after subtraction for the background contribution, the values of Δxcs are samples compared with those of the EPDM samples (Fig. 4F and F’).
calculated as Δxcs = 100 (APs–APu) A−Ps1, where APs and APu are the areas It is worth recalling that from a thermodynamical standpoint, the
subtending the radial WAXS profiles of the stretched and unstretched decrease of conformational entropy of the amorphous chains induced by

9
F. Auriemma et al. Polymer 227 (2021) 123848

Fig. 6. Increment of crystallinity (A and B) and crystallinity index (C and D) obtained by stretching the EP(D)M samples at deformation close to breaking (Δxcs, full
symbols) and after the release of the tension (Δxcr, open symbols) as a function of ethylene content (A and C) and of the mole fraction of the ethylene sequences of
length 7 (p(E ≥ 7)) relative to the ethylene sequences of all lengths (B and D).

stretching induces, in turn, a decrease of the values of the melting en­ according to two separate trends with increasing the fraction p(E ≥ 7)
tropy ΔSm. As a consequence, since the crystallization takes place at (Fig. 6D). In particular, the total crystallinities achieved by the EPM
temperatures T for which TΔSm is lower than the melting enthalpy ΔHm, samples are higher than those achieved by the EPDM samples with
the strain induced crystallization occurs at higher temperatures than in identical p(E ≥ 7) values (Fig. 6D), in agreement with the higher values of
the unstretched state [52]. However, the so formed crystals are stable at the corresponding initial crystallinity xc(WAXS) (Fig. 2).
room temperature only as long as the samples are kept in tension. By The tendency of EP(D)M samples with ethylene contents larger than
releasing the tension, the high degree of orientation of the amorphous 70 mol% to undergo incremental crystallization upon stretching is evi­
chains is lost, inducing a neat increase of the ΔSm value, and a conse­ denced also by the increase of the fractional contribution to the total
quent melting of the newly formed crystals. equatorial intensities of the (110)o/(100)h reflection of the
On the other hand, the fact that the loss of crystallinity that occurs by orthorhombic/pseudo-hexagonal forms of PE (equatorial crystallinity
releasing the tension is only partial suggests that the new crystals that index) [16]. However, the values of the equatorial crystallinity indices,
form by stretching at room temperature are characterized by different as evaluated in literature [16], represent an upper limit, as they over­
degrees of stability. It may be inferred that a portion of the crystals that estimate by a factor of 2–3 the effective values of Δxcs shown in Fig. 6.
survive even in absence of the tension are those affected by a major Effect of the incremental crystallinity on the strain hardening
degree of perfection. The crystallization of these crystals does not occur behavior of EP(D)M. A longstanding, controversial question in the
spontaneously for kinetic reasons, but it is accelerated by a decrease of realm of polymer physics and rubber elasticity in particular, is the role of
kinetic energy barriers caused by the application of the tensile stress SIC on the strain hardening behavior of amorphous entangled networks.
field. A portion of the surviving crystals may consist also of crystals that The strain induced formation of new chain-extended crystallites, well
remain entrapped in the entangled elastomeric network, even if they are oriented along the stretching direction at high deformations, is triggered
affected by only a minor stabilty. The portion of crystals that melt in by the chain extension of the network strands, as a result of application
absence of the tension are instead those which are less stable, being of uniaxial forces. On the one hand, according to the theory of rubber
characterized by a small size and/or a high degree of structural disorder. elasticity, the formation of the new oriented crystals should necessarily
The values of total crystallinities of EPM and EPDM samples achieved induce a relaxation of the connected amorphous segments, resulting, as
upon stretching and successive release of the tension increase according a consequence, in a decrease of the stress [52,71]. On the other hand, the
to a single trend with increasing ethylene content (Fig. 6C) and newly formed crystals should also produce an increase of the cross-link

10
F. Auriemma et al. Polymer 227 (2021) 123848

density and a consequent reinforcement effect especially at high defor­


mation, i.e. strain hardening [51,72–74]. Even though strain hardening
may be well-explained just on the basis of the limited extensibility of the
entangled polymer network without the occurrence of SIC [75,76], a
role of the crystals should not be discarded a priori.
The EP(D)M samples of our study, that undergo incremental crys­
tallinity by stretching, are not chemically crosslinked and are not
amorphous. They show small but significant levels of crystallinity
already in the unstretched states and tendency to increment the crys­
tallinity level upon stretching (Figs. 1 and 6 and Tables 1 and 3). In
general, strain hardening may be considered as the result of the
concomitant reinforcement due to the high level of orientation achieved
by crystal fragments derived from pristine crystals in the unstrained
sample, the high orientation attained by the amorphous chain segments
and the consequent formation of new and well-oriented crystals.
Accordingly, three different types of mechanical behavior at high de­
formations may be distinguished in EP(D)M samples depending on the
ethylene content.
This is illustrated in Fig. 7, where the stress-strain curves of samples
showing different crystallinity in the undeformed state and different
increment of crystallinity upon stretching are compared. As an example,
the stress-strain curve of the most crystalline EPM-V83 sample, with
xc(WAXS) = 16%, which experiences a medium increment of crystal­
linity upon stretching, viz. Δxcs = 13%, is compared in Fig. 7A with the
stress-strain curve of the EPDM-V80-N1.2 sample, showing a smaller
degree of crystallinity in the undeformed state, xc(WAXS) = 10%, but Fig. 7. Comparison of the stress-strain curves of the samples EPM-V83 and
the highest incremental crystallinity value Δxcs = 17%, and in Fig. 7B EPDM-V80-N1.2 (A), and EPM-V83 and EPDM-V73-D0.3 (B).
with the stress-strain curve of the sample EPDM-V73-D0.3, showing the
smallest initial crystallinity degree xc(WAXS) = 3%, and an incremental crystallization by stretching is expected to be modest. As a consequence,
crystallinity value Δxcs of 10%. also the increment of crystallinity induced by strain is expected to be not
The samples EPM-V83 and EPDM-V80-N1.2 show values of initial too high. Indeed, for EP(D)M samples with ethylene contents larger than
crystallinity xc(WAXS) of 16% and 10%, respectively and similar values 80 mol%, which are initially characterized by the presence of large and
of stress at yield (1.8 vs. 1.1 MPa, Table S2) and deformation at break well-ordered crystals, such as the sample EPM-V83, the reinforcement
(≈1600 vs. 1400%, Table S2). Moreover, the two samples show large effects occurring at high deformations (strain hardening) is essentially
differences in modulus and stress at break. In particular, the Young due to the high orientations attained by the amorphous chain segments
modulus of the sample EPM-V83 (11 MPa, Table S2) is larger than that and is augmented by the high orientations achieved by the fragments
of the sample EPDM-V80-N1.2 (3.3 MPa, Table S2), whereas the value of derived from the pristine crystals in the undeformed state rather than by
stress at break of the sample EPM-V83 (≈7 MPa, Table S2) is lower than the newly formed oriented crystals. The crystal fragments originating
that of the sample EPDM-V80-N1.2 (12 MPa, Table S2). from the initial crystals, indeed, produce a high increment of the cross-
The differences in Young moduli of these two samples are due to the link density of the elastomeric network at high deformations.
presence, in the undeformed state, of stable and well-ordered crystals in A second type of strain hardening behavior can be identified for the
the sample EPM-V83 and small, weak and highly disordered crystals in EP(D)M samples with ethylene contents around 80 mol%, such as the
the sample EPDM-V80-N1.2. The reversed difference in the values of sample EPDM-V80-N1.2 (Fig. 7A). These samples are characterized by
stress at break, instead, is essentially due to the higher strain hardening unstable and highly disordered crystals in the undeformed state and
experienced by the sample EPDM-V80-N1.2 compared with the sample experience a high level of incremental crystallinity upon stretching.
EPM-V83. Interestingly, also the value of incremental crystallinity Strain hardening, in these samples, is enhanced by the high orientation
experienced by the sample EPDM-V80-N1.2 (Δxcs = 17%) is larger than achieved by the amorphous chain segments and the consequent forma­
that of the sample EPM-V83 (Δxcs = 13%). tion of new, highly oriented crystals. The contribution from the well
The stretching of the most crystalline samples EPM-V83 and EPDM- oriented fragment derived from the pristine crystals is less important.
V80-N1.2 produces, in the initial stages, lamellar breaking coupled with The formation of new and well oriented crystals by effect of stretching
mechanical melting/recrystallization phenomena and orientation of the results in a remarkable increase of the cross-link density of the elasto­
amorphous phase [16]. As already discussed above, the increment of meric network at high deformations and, hence, in values of stress at
crystallinity in these samples starts already occurring at deformations break higher than those of samples with higher ethylene content
close to the yield point and, after a steep increase up to a deformation of (Fig. 7A).
100%, gradually approaches a quasi-plateau value at deformations The third type of behavior arises for the low crystalline EPDM sam­
higher than 100–200%. The increase of crystallinity at low deformations ples with ethylene content lower than 73 mol% and can be identified by
stems from the high degree of orientation attained by the amorphous comparing the stress-strain curves of the sample EPM-V83 and the
chains and the presence of crystal fragments, which not only act as sample EPDM-V73-D0.3 in Fig. 7B, as an example. The initial degree of
physical links of the entangled network, but also as nuclei triggering crystallinity of the sample EPDM-V73-D0.3 is only 3%. The increment of
further crystallization [16]. crystallinity as a result of stretching for EP(D)M samples with ethylene
Therefore, the data of Fig. 7A suggest that a first type of strain contents smaller than 73 mol% starts only at deformations higher than a
hardening behavior may be envisaged in EP(D)M copolymers with threshold, after the attainment of a high degree of orientation of the
ethylene contents larger than 80 mol%, for which the presence of stable amorphous chains, and is similar to SIC occurring in non-crosslinked
and ordered crystals in the undeformed state is the hallmark that the natural rubber [16,46,67–70]. The EPDM-V73-D0.3 sample, in partic­
majority of regular ethylene sequences have already crystallized, so that, ular, shows a low Young modulus (1.1 MPa, Table S2), a high ductility
the fraction of crystallizable sequences available for further (deformation at break equal to 3800%, Table S2) and no remarkable

11
F. Auriemma et al. Polymer 227 (2021) 123848

strain hardening. Furthermore, the incremental crystallinity induced by due to the high levels of orientation achieved by crystal fragments
stretching is smaller (10%) than that achieved by the sample EPM-V83 derived from pristine crystals initially present in the undeformed sam­
(13%). This is due to the intrinsic features of the chain microstructure, ple, ii) the high orientation attained by the amorphous chains and iii) the
as the p(E ≥ 7) value for the EPM-V73-D0.3 sample of ≈0.09, is close to possible consequent formation of new and well-oriented crystals.
the threshold limit necessary for the occurrence of a significant crys­ Accordingly, three different types of strain hardening behavior are
tallinity increment induced by strain. Therefore, the third type of me­ distinguished, depending on the nature of the governing factors. The
chanical behavior at high deformation may be identified for the low first type of strain hardening behavior is observed for EP(D)M samples
crystalline EP(D)M samples with ethylene contents around 73 mol%, with ethylene content larger than 80 mol%, which form large and well-
such as the sample EPM-V73-D0.3. For these samples, the small degrees ordered crystals in the undeformed state. For these samples, the rein­
of crystallinity coupled with the moderate tendencies towards devel­ forcement effect occurring at high deformations is enhanced by crystal
opment of incremental crystallinity by stretching, induce only a small or fragmentation, and the high orientations achieved by these fragments
negligible increase of the cross-link density of the elastomeric network along the stretching direction, coupled with the high orientations of the
at high deformations, and hence only very little or no strain hardening at amorphous chains. The second type of strain hardening behavior occurs
high deformations, even though the amorphous chains achieve a high in EP(D)M samples with ethylene contents around 80 mol%, which are
degree of orientation with the chain axes parallel to the stretching di­ characterized by lower crystallinity and less perfect and less stable
rections [16]. crystals in the undeformed state. These samples experience high levels of
incremental crystallinity upon stretching, leading to remarkable strain
4. Conclusions hardening. In these samples, strain hardening is determined by the high
orientations achieved by the amorphous chains and the consequent
The effects of the polymer chain microstructure of EP(D)M samples formation of new highly oriented crystals. Finally, the third type of
in terms of ethylene contents and fractional contents of ethylene se­ strain hardening behavior may be identified for the EP(D)M samples
quences longer than a threshold value on the deformation behavior and with ethylene contents around 70 mol%, which are characterized by
the crystallization induced by stretching have been investigated, small initial degrees of crystallinity. For these samples, the small degrees
focusing the analysis on EP(D)M samples polymerized using conven­ of crystallinity combined with moderate tendencies towards the devel­
tional V-based catalysts and more modern metalorganic catalysts. The opment of incremental crystallinity by stretching induce only a small or
analysis of the chain microstructure reveals that, compared with V- negligible increase of the cross-link density of the elastomeric network
based counterparts, the adopted (post-)metallocene catalysts show a at high deformations and, hence, only very little or no strain hardening,
lower tendency toward the propagation of hetero-sequences of the co- even though the amorphous chains achieve high degrees of orientation
monomeric units and the incorporation of regiomistakes (viz. 2,1-pro­ with the chain axes parallel to the stretching directions.
pylene units).
EP(D)M samples with ethylene contents in the range 73–83 mol% Declaration of competing interest
show an initial degree of crystallinity increasing from 3 to 16% with
increasing ethylene content and similar melting temperatures (in the The authors declare that they have no known competing financial
range 40–44 ◦ C). The samples with ethylene contents smaller than 63 interests or personal relationships that could have appeared to influence
mol% are instead amorphous. The initial crystallinity of EPM co­ the work reported in this paper.
polymers is higher than those of EPDM terpolymers with similar frac­
tional contents of ethylene sequences of lengths equal to or longer than 7 Appendix A. Supplementary data
ethylene units relative to the ethylene sequences of all lengths p(E ≥ 7).
The small differences in the crystallization properties of samples with Supplementary data to this article can be found online at https://doi.
similar ethylene content prepared with the two different catalytic sys­ org/10.1016/j.polymer.2021.123848.
tems are accounted for the small difference in the chain microstructure.
The analysis of the mechanical properties has shown that the values References
of the Young modulus and of the strain at yield are essentially sensitive
to the ethylene content. Other tensile properties, such as the strain and [1] A.K. Bhowmick, H.L. Stephens (Eds.), Handbook of Elastomers, second ed., Marcel
Dekker Inc., New York, 2001.
stress at break, the stress at yield and the recovered strain values, follow
[2] M. Morton (Ed.), Rubber Technology, Springer Science & Business Media, 2013.
two different trends as a function of p(E ≥ 7) for EPM and EPDM samples. [3] R.F. Ohm (Ed.), The Vanderbilt Rubber Handbook, thirteenth ed., R.T. Vanderbilt
In particular, for identical values of the fraction p(E ≥ 7), the values of Company Inc., Norwalk, 1990, p. 123.
yield stress of EPM copolymers are larger and the values of stress at [4] E. Albizzati, U. Giannini, G. Collina, L. Noristi, L. Resconi, in: E.P. Moore Jr. (Ed.),
Polypropylene Handbook, Hanser Publishers, New York, 1996, p. 419.
break and elastic recovery are smaller than those of EPDM terpolymers. [5] W. Hofmann, Rubber Technology Handbook, Hanser Publishers, Munich, 1989 ch.
This indicates that the presence of about 1 mol% diene co-units in EPDM 3.3.8.
copolymers induce a small but significant decrease of the stress at yield [6] J.A. Brydson, Rubbery Materials and Their Compounds, Elsevier, London, 1988 ch.
7.
whereas the ultimate tensile strength is larger than that of EPM co­ [7] E.K. Easterbrook, R.D. Allen, in: M. Morton (Ed.), Ethylene-Propylene Rubber in
polymers due to occurrence of strong strain hardening. Strain hard­ Rubber Technology, Springer, 1999.
ening, in turn, is related to the occurrence of further crystallization upon [8] J.W.M. Noordermeer, Ethylene–Propylene Elastomers in Encyclopedia of Polymer
Science and Technology, Wiley, 2002.
stretching, and its extent is dependent on the initial degree of crystal­ [9] R.C. Klingender (Ed.), Handbook of Specialty Elastomers, CRC Press Book, 2008.
linity in the samples. [10] B.J.R. Scholtens, The effect of variation in chain connectivity on strain-induced
The EP(D)M samples with ethylene contents larger than 73 mol% crystallization in unvulcanized EPDM elastomers, Rubber Chem. Tech. 57 (1984)
703–724, https://doi.org/10.5254/1.3536027.
experience small but significant increments of crystallinity by stretch­ [11] B.J.R. Scholtens, E. Riande, J.E. Mark, Crystallization in stretched and unstretched
ing. The incremental crystallinity has been evaluated for samples EPDM elastomers, J. Polym. Sci., Polym. Phys. Ed. 22 (1984) 1223–1238, https://
stretched at deformations close to rupture and is around 10% for EP(D) doi.org/10.1002/pol.1984.180220707.
[12] V.B. F Mathot, R. L Scherrenberg, M.F.J. Pijpers, W. Bras, D.S.C. Dynamic, SAXS
M samples with ethylene contents of 73 mol%, increases to 15 and 17%
and WAXS on homogeneous ethylene-propylene and ethylene-octene copolymers
for samples with ethylene contents of 78 and 80 mol% and then drops to with high comonomer contents, J. Therm. Anal. 46 (1996) 681, https://doi.org/
13% for the most crystalline EPM sample with an ethylene content of 83 10.1007/bf01983597.
mol%. The incremental crystallinity induced by stretching influences [13] O. Ruiz de Ballesteros, F. Auriemma, G. Guerra, P. Corradini, Molecular
organization in the pseudo-hexagonal crystalline phase of ethylene-propylene
the strain hardening behavior. Strain hardening generally originates copolymers, Macromolecules 29 (1996) 7141–7148, https://doi.org/10.1021/
from the contributions of at least three factors, i.e. i) the reinforcement ma960511z.

12
F. Auriemma et al. Polymer 227 (2021) 123848

[14] L.-Z. Liu, B.S. Hsiao, B.X. Fu, S. Ran, S. Toki, B. Chu, A.H. Tsou, P.K. Agarwal, copolymers, Polym. Chem. 9 (2018) 48–59, https://doi.org/10.1039/
Structure changes during uniaxial deformation of ethylene-based eemicrystalline C7PY01807J.
ethylene-propylene copolymer. 1. SAXS study, Macromolecules 36 (2003) [40] P.R. Swann, Polyethylene unit cell variations with branching, J. Polym. Sci. 56
1920–1929, https://doi.org/10.1021/ma020771i. (1962) 409–416, https://doi.org/10.1002/pol.1962.1205616411.
[15] L.-Z. Liu, B.S. Hsiao, R. Ran, B.X. Fu, S. Toki, F. Zuo, A.H. Tsou, B. Chu, In situ [41] E.R. Walter, F.P. Reding, Variations in unit cell dimensions in polyethylene,
WAXD study of structure changes during uniaxial deformation of ethylene-based J. Polym. Sci. 21 (1956) 561–562, https://doi.org/10.1002/pol.1956.120219925.
semicrystalline ethylene–propylene copolymer, Polymer 47 (2006) 2884–2893, [42] E.A. Cole, D.R. Holmes, Crystal lattice parameters and the thermal expansion of
https://doi.org/10.1016/j.polymer.2006.01.090. linear paraffin hydrocarbons, including polyethylenes, J. Polym. Sci. 46 (1960)
[16] F. Auriemma, M. Scoti, R. Di Girolamo, A. Malafronte, C. De Rosa, M. van Duin, 245–256, https://doi.org/10.1002/pol.1960.1204614722.
Effect of stretching on the crystallization of un-crosslinked ethylene/propylene [43] R.M. Eichorn, Unit-cell expansion in polyethylene, J. Polym. Sci. 31 (1958)
(/diene) random copolymers, Polymer 199 (2020) 122540, https://doi.org/ 197–198, https://doi.org/10.1002/pol.1958.1203112228.
10.1016/j.polymer.2020.122540. [44] J.E. Preedy, A study of branching in ethylene copolymers by X-ray diffraction, Br.
[17] R.B. Richards, Polyethylene-structure, crystallinity and properties, J. Appl. Chem. Polym. J. 5 (1973) 13–19, https://doi.org/10.1002/pi.4980050103.
1 (1951) 370–376, https://doi.org/10.1002/jctb.5010010812. [45] I.W. Bassi, P. Corradini, G. Fagherazzi, A. Valvassori, Crystallization of high
[18] B. Wunderlich, D. Poland, Thermodynamics of crystalline linear high polymers. II. ethylene EPDM terpolymers in the stretched state, Eur. Polym. J. 6 (1970)
The influence of copolymer units on the thermodynamic properties of 709–718, https://doi.org/10.1016/0014-3057(70)90020-0.
polyethylene, J. Polym. Sci., Part A 1 (1963) 357–372, https://doi.org/10.1002/ [46] J.C. Randall, C.J. Ruff, New look at the “run number” concept in copolymer
pol.1963.100010132. characterization, Macromolecules 21 (1988) 3446–3454, https://doi.org/
[19] C.H. Baker, L. Mandelkern, The crystallization and melting of copolymers 10.1021/ma00190a017.
II—variation in unit-cell dimensions in polymethylene copolymers, Polymer 7 [47] T.M. Krigas, J.M. Carella, M.J. Struglinski, B. Crist, W.W. Graessley, Model
(1965) 71–83, https://doi.org/10.1016/S0032-3861(66)80002-2. copolymers of ethylene with butene-1 made by hydrogenation of polybutadiene:
[20] G. Ver Strate, Z.W. Wilchinsky, Ethylene-propylene copolymers: degree of chemical composition and selected physical properties, J. Polym. Sci. Polym. Phys.
crystallinity and composition, J. Polym. Sci., Part A-2 9 (1971) 127–142, https:// Ed 23 (1985) 509–520, https://doi.org/10.1002/pol.1985.180230308.
doi.org/10.1002/pol.1971.160090109. [48] T. Usami, Y. Gotoh, S. Takayama, Generation mechanism of short-chain branching
[21] F.P. Baldwin, G. Ver Strate, Rubber polyolefin elastomers based on ethylene and distribution in linear low-density polyethylenes, Macromolecules 19 (1986)
propylene, Chem. Technol. 45 (1972) 709–881, https://doi.org/10.5254/ 2722–2726, https://doi.org/10.1021/ma00165a010.
1.3544730. [49] G. Natta, G. Mazzanti, A. Valvassori, G. Sartori, D. Morero, Copolimerizzazione
[22] G. Guerra, O. Ruiz de Ballesteros, V. Venditto, M. Galimberti, F. Sartori, dell’etilene con le alfa-olefine alifatiche, Chim. Ind. (Milan) 125 (1960) 42–52.
R. Pucciarello, Pseudohexagonal crystallinity and thermal and tensile properties of [50] H.G. Killian, Kristallisation von Copolymeren des Äthylens, Kolloid Z. 189 (1963)
ethene–propene copolymers, J. Polym. Sci., Part B: Polym. Phys. 37 (1999) 23–36, https://doi.org/10.1007/BF01500284.
1095–1103, https://doi.org/10.1002/(SICI)1099-0488(19990601)37:11<1095:: [51] M. Tosaka, D. Kawakami, K. Senoo, S. Kohjiya, Crystallization and stress relaxation
AID-POLB5>3.0.CO;2-J. in highly stretched samples of natural rubber and its synthetic analogue,
[23] M. van Duin, G. van Doremaele, N. van der Aar, Defining EPDM for the past and the Macromolecules 39 (2006), https://doi.org/10.1021/ma060407+, 5100-5005.
next 50 years, Kautsch. Gummi Kunstst. 14 (2017) 11. [52] P.J. Flory, Thermodynamics of crystallization in high polymers. I. Crystallization
[24] M.A. Grima, M. van Boggelen, M. Dees, G.H.J. van Doremaele, M. van Duin, induced by stretching, J. Phys. Chem. 15 (1947) 397–408, https://doi.org/
P. Henricks-Knape, High performance Keltan ACE EPDM polymers, Rubber World 10.1063/1.1746537.
250 (2014) 37–43. [53] S. Trabelsi, P.-A. Albouy, J. Rault, Crystallization and melting processes in
[25] G. Van Doremaele, M. van Duin, M. Valla, A. Berthoud, On the development of vulcanized stretched natural rubber, Macromolecules 36 (2003) 7624–7639,
titanium κ1-amidinate complexes, commercialized as Keltan ACETM technology, https://doi.org/10.1021/ma030224c.
enabling the production of an unprecedented large variety of EPDM polymer [54] S. Trabelsi, P.-A. Albouy, J. Rault, Effective local deformation in stretched filled
structures, J. Polym. Sci., Part A: Polym. Chem. 55 (2017) 2877, https://doi.org/ rubber, Macromolecules 36 (2003) 9093–9099, https://doi.org/10.1021/
10.1002/pola.28634. ma0303566.
[26] R.P. Quirk, R.E. Hoff, Transition Metal Catalyzed Polymerizations: Ziegler-Natta [55] P.-A. Albouy, A. Vieyres, R. Pérez-Aparicio, O. Sanséau, P. Sotta, The impact of
and Metathesis Polymerizations, Illustrated Ed., University of Akron. Institute of strain-induced crystallization on strain during mechanical cycling of cross-linked
Polymer Science Edison Polymer Innovation Corporation Cambridge University natural rubber, Polymer 55 (2014) 4022, https://doi.org/10.1016/j.
Press, 1988. polymer.2014.06.034.
[27] L. D’Agnillo, J.B.P. Soares, G.H.J. van Doremaele, Steady-state model for olefin [56] C.J. Carman, R.A. Harrington, C.E. Wilkes, Monomer sequence distribution in
polymerization with a two-site Vanadium catalyst in a continuous stirred-tank ethylene-propylene rubber measured by 13C NMR. 3. Use of reaction probability
reactor, Macromol. Mater. Eng. 290 (2005) 256–271, https://doi.org/10.1002/ model, Macromolecules 10 (1977) 536, https://doi.org/10.1021/ma60057a008.
mame.200400322. [57] J.C. Randall, A review of high resolution liquid 13Carbon Nuclear Mmagnetic
[28] G.G. Evens, E.M. J Pijpers, R.H.M. Seevens, Process for the Preparation of resonance characterizations of ethylene-based polymers, Macromol. Chem. Phys.
Copolymers of Ethylene with at Least One Other 1-alkene, EP0044119B1 to DSM C29 (1989) 201–317, https://doi.org/10.1080/07366578908055172.
Netherlands B.V, 1985. [58] V. Busico, R. Cipullo, A.L. Segre, Advances in the 13C NMR characterization of
[29] Y. Ma, D. Reardon, S. Gambarotta, G. Yap, Vanadium-catalyzed ethene/propene copolymers, 1 - C2-symmetric ansa-metallocene catalysts,
ethylene− propylene copolymerization: the question of the metal oxidation state in Macromol. Chem. Phys. 203 (2002) 1403–1412, https://doi.org/10.1002/1521-
Ziegler− Natta polymerization promoted by (β-diketonate)3V, Organometallics 18 3935(200207)203:10/11<1403::AID-MACP1403>3.0.CO;2-I.
(1999) 2773–2781, https://doi.org/10.1021/om9808763. [59] J. Brandrup, E.H. Immergut, Polymer Handbook, third ed., John Wiley & Sons,
[30] O. Kramer, W.R. Good, Correlating Mooney viscosity to average molecular weight, New York, 1989.
J. Appl. Polym. Sci. 16 (1972) 2677–2684, https://doi.org/10.1002/ [60] F. A Bovey, M.C. Sacchi, A. Zambelli, Polymerization of propylene to syndiotactic
app.1972.070161020. polymer. IX. Ethylene perturbation of syndiotactic propylene polymerization,
[31] S. Gambarotta, Vanadium-based Ziegler/Natta: challenges, promises, problems, Macromolecules 7 (1974) 752, https://doi.org/10.1021/ma60042a010.
Coord. Chem. Rev. 237 (2003) 229–243, https://doi.org/10.1016/S0010-8545(02) [61] G. Natta, I. Pasquon, A. Zambelli, Stereospecific catalysts for the head-to-tail
00298-9. polymerization of propylene to a crystalline syndiotactic polymer, J. Am. Chem.
[32] S.C. Davis, W. von Hellens, H.A. Zahalka, K.-P. Richter, in: J.C. Salamone (Ed.), Soc. 84 (1962) 1488, https://doi.org/10.1021/ja00867a029.
Polymer Material Encyclopedia, vol. 3, CRC Press Inc., Boca Raton, 1996, p. 2264. [62] A. Zambelli, P. Locatelli, G. Bajo, F.A. Bovey, Model compounds and 13C NMR
[33] R. Karpeles, A.V. Grossi, in: A.K. Bhowmick, H.L. Stephens (Eds.), Handbook of observation of stereosequences of polypropylene, Macromolecules 8 (1975) 687,
Elastomers, second ed., Marcel Dekker, New York, 2001, p. 845. https://doi.org/10.1021/ma60047a024.
[34] G. Ver Strate, Encyclopedia of Polymer Science and Engineering, vol. 6, Wiley, [63] A. Zambelli, I. Sessa, F. Grisi, R. Fusco, P. Accommazzi, Syndiotactic
New York, 1986, p. 522. polymerization of propylene: single-site vanadium catalysts in comparison with
[35] K. Nomura, S. Zhang, Design of Vanadium complex catalysts for precise olefin zirconium and nickel, Macromol. Chem. Rapid Commun. 22 (2001) 297, https://
polymerization, Chem. Rev. 111 (2011) 2342–2362, https://doi.org/10.1021/ doi.org/10.1002/1521-3927(20010301)22:5<297::AID-MARC297>3.0.CO;2-P.
cr100207h. [64] M. Galimberti, N. Mascellari, F. Piemontesi, I. Camurati, Random ethene/propene
[36] N.M. Bravaya, E.E. Faingol’d, E.R. Badamshina, E.A. Sanginov, Advances in copolymerization from a catalyst system based on a “constrained geometry” half-
synthesis of ethylene–propylene–diene elastomers by ion-coordination sandwich complex, Macromol. Rapid Commun. 20 (1999) 214, https://doi.org/
polymerization on single-site catalytic systems of new generation, Polym. Sci., 10.1002/(SICI)1521-3927(19990401)20:4<214::AID-MARC214>3.0.CO;2-A.
Series C. 62 (2020) 1–16, https://doi.org/10.1134/S1811238220010014. [65] G. Guerra, M. Galimberti, F. Piemontesi, O. Ruiz de Ballesteros, Influence of regio-
[37] A. Berthoud, G.H.J. van Doremaele, V. Quiroga Norambuena, R.T. W Scott, and stereoregularity of propene insertion on crystallization behavior and elasticity
Catalyst system and a process for the preparation of a polymer using the same, of ethene-propene copolymers, J. Am. Chem. Soc. 124 (2002) 1566, https://doi.
WO2015052184A1 to Arlanxeo Netherlands B.V. (2015). org/10.1021/ja017199f.
[38] R.A. Collins, A.F. Russell, R.T.W. Scott, R. Bernardo, G.H.J. van Doremaele, [66] P.J. Flory, Theory of crystallization in copolymers, Trans. Faraday Soc. 51 (1955)
A. Berthoud, P. Mountford, P. Mountford, Monometallic and bimetallic Titanium 848–857, https://doi.org/10.1039/TF9555100848.
κ1-amidinate complexes as olefin polymerization catalysts, Organometallics 36 [67] J.F. Jackson, Crystallinity in ethylene—propylene copolymers, J. Polym. Sci., Part
(2017) 2167–2181, https://doi.org/10.1021/acs.organomet.7b00225. A 1 (1963) 2119–2126, https://doi.org/10.1002/pol.1963.100010630.
[39] Z.-Q. Zhang, J.-T. Qu, S. Zhang, Q.-P. Miao, Y.-X. Wu, Ethylene/propylene [68] H. Luo, Q. Chen, G. Yang, Studies on the minimum crystallizable sequence length
copolymerization catalyzed by half-titanocenes containing monodentate anionic of semicrystalline copolymers, Polymer 42 (2001) 8285–8288, https://doi.org/
nitrogen ligands: effect of ligands on catalytic behaviour and structure of 10.1016/S0032-3861(01)00327-5.

13
F. Auriemma et al. Polymer 227 (2021) 123848

[69] D.R. Burfield, Correlation between crystallinity and ethylene content in LLDPE and [73] S. Toki, I. Sics, L. Liu, B.S. Hsiao, S. Murakami, M. Tosaka, S. Poompradub,
related ethylene copolymers. Demonstration of the applicability of a simple S. Kohjiya, Y. Ikeda, Probing the nature of strain-induced crystallization in
empirical relationship, Macromolecules 20 (1987) 3020–3023, https://doi.org/ polyisoprene rubber by combined thermo-mechanical and in situ X-ray diffraction
10.1021/ma00178a013. techniques, Macromolecules 38 (2005) 7064–7073, https://doi.org/10.1021/
[70] D.R. Burfield, N. Kashiwa, DSC studies of linear low density polyethylene. Insights ma050465f.
into the disrupting effect of different comonomers and the minimum fold chain [74] L.-R.G. Treloar, The Physics of Rubber Elasticity, third ed., Oxford University Press,
length of the polyethylene lamellae, Makromol. Chem. 186 (1985) 2657–2662, Oxford, 1975.
https://doi.org/10.1002/macp.1985.021861226. [75] S.F. Edwards, T.A. Vilgis, The tube model theory of rubber elasticity, Rep. Prog.
[71] Y. Miyamoto, H. Yamao, K. Sekimoto, Crystallization and melting of polyisoprene Phys. 51 (1988) 243–297, https://doi.org/10.1088/0034-4885/51/2/003.
rubber under uniaxial deformation, Macromolecules 36 (2003) 6462–6471, [76] J.D. Davidson, N.C. Goulbourne, A nonaffine network model for elastomers
https://doi.org/10.1021/ma0342877. undergoing finite deformations, J. Mech. Phys. Solid. 61 (2013) 1784–1797,
[72] J.C. Mitchell, D.J. Meier, Rapid stress-induced crystallization in natural rubber, https://doi.org/10.1016/j.jmps.2013.03.009.
J. Polym. Sci., Part A-2 6 (1968) 1689–1703, https://doi.org/10.1002/
pol.1968.160061001.

14

You might also like