Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/274903231

Aqueous hydrogen peroxide-induced degradation of polyolefins: A greener


process for controlled-rheology polypropylene

Article  in  Polymer Degradation and Stability · April 2015


DOI: 10.1016/j.polymdegradstab.2015.04.001

CITATIONS READS

19 3,175

8 authors, including:

Graeme Moad Ian Dagley


The Commonwealth Scientific and Industrial Research Organisation CRC for Polymers
310 PUBLICATIONS   32,168 CITATIONS    43 PUBLICATIONS   342 CITATIONS   

SEE PROFILE SEE PROFILE

Jana Habsuda Chris Garvey


Visy Technische Universität München
22 PUBLICATIONS   447 CITATIONS    162 PUBLICATIONS   3,213 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Radical Polymerization View project

Nitroxide Mediated Polymerization View project

All content following this page was uploaded by Graeme Moad on 19 June 2018.

The user has requested enhancement of the downloaded file.


Polymer Degradation and Stability 117 (2015) 97e108

Contents lists available at ScienceDirect

Polymer Degradation and Stability


journal homepage: www.elsevier.com/locate/polydegstab

Aqueous hydrogen peroxide-induced degradation of polyolefins:


A greener process for controlled-rheology polypropylene
Graeme Moad a, b, *, Ian J. Dagley a, Jana Habsuda a, c, Christopher J. Garvey a, d,
Guoxin Li a, b, Lance Nichols a, b, George P. Simon a, c, Maria Rossella Nobile e
a
CRC for Polymers, 8 Redwood Drive, Notting Hill, Victoria 3168, Australia
b
CSIRO Manufacturing, Bayview Ave, Clayton, Victoria 3168, Australia
c
Monash University, Department of Materials Engineering, Clayton, Victoria 3800, Australia
d
Bragg Institute, Australian Nuclear Science and Technology Organisation, Lucas Heights, NSW 2234, Australia
e
University of Salerno, Department of Industrial Engineering, 84084 Fisciano, Salerno, Italy

a r t i c l e i n f o a b s t r a c t

Article history: In this work we demonstrate that aqueous hydrogen peroxide is an effective reagent for chain scissioning
Received 17 February 2015 or vis-breaking of polypropylene during melt-processing to produce a controlled rheology product. The
Received in revised form novel process involves the direct injection of aqueous hydrogen peroxide into the polypropylene melt
31 March 2015
under pressure. The polypropylene produced has reduced molar mass, narrowed molar mass distribu-
Accepted 5 April 2015
tion, and is indistinguishable in terms of melt flow rate, molar mass distribution, crystallinity and melt
Available online 11 April 2015
rheology from conventionally vis-broken polypropylene produced using an organic peroxide (2,5-
dimethyl-2,5-di-tert-butylperoxyhexane (DHBP)). However, the polypropylene produced in the current
Keywords:
Polypropylene
process is notably free of the initiator-derived organic volatiles that are formed as by-products in the case
Hydrogen peroxide where organic peroxides such as DHBP are used.
Scissioning © 2015 Elsevier Ltd. All rights reserved.
Rheology
Molecular weight

1. Introduction other peroxides [24]. However, the use of these organic peroxides
in this context suffers from a number of drawbacks, most notably,
Ex-reactor conventionally-formed polypropylene is character- the formation of organic volatiles, such as tert-butanol and acetone
ized by a broad molar mass distribution (MMD). So called in the case of DHBP (Scheme 1) [4]. This has provided the impetus
“controlled rheology” grades of lower molar mass and dispersity for a significant effort in developing other initiators for use in the
are produced in a subsequent in-line melt processing step in an process. These include cyclic peroxides [27] and a range of non-
extruder. These grades possess improved processability and the peroxide initiators, which include certain azoalkanes [28] and
specific rheological behaviour required for many applications alkoxyamines [29e31], and UV photoinitiators [32].
which include fibre and thin-wall injection-moulding [1e4]. The Hydrogen peroxide is sometimes mentioned in lists of suitable
process has been described fully in reviews describing reactive initiators in various patents and is mentioned in at least one book
extrusion processing of polyolefins [1e4] and usually involves [33] as an initiator used in forming controlled rheology poly-
radical-induced chain-scissioning or vis-breaking. The reagents propylene, though to our knowledge there are no actual examples
most commonly used as a source of radicals are organic peroxides of such use [1e4]. Given the inherent incompatibility of water and
[4], which include cumyl peroxide [5e11], di-tert-butyl peroxide polyolefin-melts, it seems counterintuitive to contemplate the
[12], 2,5-dimethyl-2,5-di-tert-butylperoxyhexane (DHBP; also direct use of an aqueous peroxide solution in melt-extrusion of
known under trade names such as Luperox™ 101 or Triganox™ polypropylene at high temperatures. Nonetheless, there is an
101) [10,13e25], di(tert-butylperoxyisopropyl)benzene [26] and obvious advantage to the use of hydrogen peroxide in the context
of vis-breaking, in that the only expected initiator-derived by-
product is water. Our initial research on the use of aqueous
* Corresponding author. CSIRO Manufacturing, Bayview Ave, Clayton, Victoria hydrogen peroxide in vis-breaking is described in a recent patent
3168, Australia.
application [34].
E-mail address: graeme.moad@csiro.au (G. Moad).

http://dx.doi.org/10.1016/j.polymdegradstab.2015.04.001
0141-3910/© 2015 Elsevier Ltd. All rights reserved.
98 G. Moad et al. / Polymer Degradation and Stability 117 (2015) 97e108

hydrogens, the methine hydrogens of 2,4-dimethylpentane were


nonetheless abstracted preferentially (Fig. 1). Similar observations
were made by Dokolas et al. [40] for hydrogen abstraction from 2,4-
dimethylpentane by tert-butoxy radicals at 60  C (Fig. 1). These data
appear in general agreement with reactivities estimated in EPR
studies by Camara et al. [41].
The process with hydrogen peroxide is anticipated to produce
hydroxy radicals thermally or by induced decomposition [42,43].
Hydroperoxy radicals can also be formed by hydrogen-atom
abstraction from hydrogen peroxide [42]. It is also possible to
form oxygen from hydrogen peroxide. We have found in previous
Scheme 1. Mechanism of thermal decomposition of 2,5-dimethyl-2,5-di-tert-butyl- work that oxygen will cause scissioning of polypropylene, though
peroxyhexane (DHBP). [H] is a hydrogen atom source, e.g., polypropylene. the product formed is discoloured and there is some formation of
volatile byproducts. We also found that the process is catalysed by
transition metals such as copper [44].
The mechanism of vis-breaking using organic peroxides is In this work we describe the process condition for vis-breaking
usually described as shown in Scheme 2 [35]. The process involves polypropylene with hydrogen peroxides in melt extrusion. We
formation of tertiary radicals (a to the polypropylene methyl) on characterize the polymer by molar mass, MMD, crystallinity and
the backbone that undergo b-scission to form a polypropylene dynamic rheological properties and compare properties with those
propagating radical and an unsaturated chain end. Bertin et al. for the similar polymer formed by vis-breaking with DHBP.
[36,37] have suggested a very different mechanism. They explored
polypropylene degradation in solution at 140e165  C and surpris- 2. Experimental
ingly found for experiments under inert atmosphere that there was
a small increase in molar mass. It should be noted that the condi- 2.1. Materials
tions used in the present study, and in the industrial process for
producing controlled rheology polypropylene (~280  C, polymer The polypropylene grades (HP400N, HP555G) used are manu-
melt), differ markedly from those used by Bertin et al. [37] factured by Lyondell-Basell and contain a proprietary stabilizer
(140e165  C, solution). They [37] found that efficient chain- package. Aqueous hydrogen peroxide (59.5% w/w) Interox®ST-60
scissioning required the presence of oxygen. They also proposed was obtained from Solvay Interox Pty Ltd Australia. DHBP was ob-
on the basis of density functional calculations that, contrary to tained from United Initiators Gmbh.
expectations based on relative bond strengths, hydrogen abstrac-
tion by tert-butoxy (and other tert-alkoxy radicals) radical should
2.2. Extrusion
occur preferentially from the polypropylene methyls due to a more
favourable entropy of activation. The primary radicals so formed do
All extrusion experiments were carried out with a JSW TEC30
not readily undergo b-scission and terminate mainly by combina-
twin screw extruder with L/D 47. A flat temperature profile was
tion. They [37] speculated that the efficiency of chain scission
used where zone 2 was set to 160  C. Zones C3eC14 (of 14) were all
observed in the commercial vis-breaking process was therefore
set to 280  C and the die was set to 220  C. Polypropylene was
dependent on the presence of oxygen and the formation of meth-
added to the nitrogen blanketed hopper through a Brabender
ylperoxy and hydroxy radicals, which they proposed should show
Flexiwall FW40/5-50 gravimetric feeder at 20 kg h1. Aqueous
less selectivity.
hydrogen peroxide was injected under pressure into the propylene
These proposals [37] are not supported by more recent theo-
melt stream at C4 using a Teldyne-Isco 500D syringe pump through
retical and experimental work. Gryn'ova et al. [38] found using high
a cooled injection nozzle. On exit of the die, the strand was cooled
level ab initio calculations that the reactions of peroxyl radicals
by passage through an ambient temperature water bath
with unactivated hydrogens is extremely endothermic such that
(2.4 m  300 mm  300 mm), pelletized, collected and dried at
they unfavourable even at elevated temperatures and are unlikely
ambient temperature (typically ~24  C). Full details of the screw
to be involved in oxidative degradation. Other studies have found
configuration and temperature profile are shown in the
that oxidative degradation provides a polypropylene containing
Supplementary material.
various functionality at the chain ends and along the backbone
which are observable using such techniques as FTIR spectroscopy
and CRYSTAF [39]. 2.3. Melt flow rate (MFR) measurements
Garrett et al. [11] used radical trapping to examine the regio-
selectivity of hydrogen abstraction from model substrates by Melt flow rate (MFR) measurements were performed according
cumyloxy radicals at 160  C. They found that, while steric factors to ASTM D1238 [45] (230  C/2.16 kg for polypropylene). The units
reduced the rate of abstraction or methine and methylene for MFR are g/10 min using an IDM instruments melt flow indexer.

Fig. 1. Experimentally determined relative reactivity per hydrogen atom in abstraction


Scheme 2. Proposed mechanism of chain scission. by (a) cumyloxy radicals at 160  C [11] (b) tert-butoxy radicals at 60  C [40].
G. Moad et al. / Polymer Degradation and Stability 117 (2015) 97e108 99

2.4. Size exclusion chromatography (SEC) stored in polyethylene bags at room temperature prior to the
experiment. In order to estimate the crystalline fraction, Gaussian
Size exclusion chromatography (SEC) or gel permeation chro- functions were used to describe the amorphous and crystalline
matography was performed with a GPC PL 220 from Polymer Lab- peaks. The overall crystallinity was then calculated as the ratio of
oratories (Agilent) instrument equipped with a refractive index the total of the fitted crystalline areas to fitted total crystalline plus
detector and two linear Olexis columns (each 300  8 mm) from amorphous areas. The procedure has been described in detail by
Polymer Laboratories. The mobile phase was 1,2,4-trichlorobenzene Zhu et al. [47]. A second estimate of the crystalline fraction was
which was purified by vacuum distillation and stabilized with 3,5- provided by the program Traces™ (version 6).
di-tert-4-butylhydroxytoluene (BHT, 200 mg L1). The injection
system, columns and detector were maintained at 150  C. The 2.9. Small angle and wide angle X-ray scattering (SAXS and WAXS)
sample concentration was 1 mg mL1. (synchrotron source)

2.5. Crystallization analysis fractionation (CRYSTAF) Samples prepared as above were packed in a 1.5 mm quartz
capillary. SAXS and WAXS scattering patterns were simultaneously
A CRYSTAF model 200 from Polymer Char S.A. (Valencia, Spain) acquired on the SAXS/WAXS beam line at the Australian Synchro-
was used for crystallization fractionation. The sample (20 mg) was tron. For each sample 2 single measurements of 1 s were averaged
dissolved in 1,2,4-trichlorobenzene (30 mL). The cooling rate used from different spots on the capillary. The sample measurements
was 0.1  C min1 [39]. were complemented by a background measurement of 2 by 1 s
measurements on an empty capillary in air. There was no difference
2.6. Differential scanning calorimetry (DSC) between duplicate measurements. Further details of methods and
data analysis are provided in the Supplementary material.
A Mettler Toledo DSC821e equipped with a TSO 801RO sample
robot and ATARe software version 9.00a was used. Experiments
2.10. Melt rheology
were performed under nitrogen (50.0 mL min1). The temperature
program involved heating the sample from 40.0 to 240.0  C at
The complex viscosity (h*), the storage (G0 ) and the loss (G00 )
10.0  C min1, maintaining the sample at 240.0  C for 10.0 min then
moduli were measured using a Rheometrics Scientific ARES
cooling from 240.0 to 40.0  C at 20  C min1. The program was
(Advanced Rheometric Expansion System), equipped with a
then repeated.
2KFRTN1 transducer, in oscillatory mode. A parallel plate geometry
50 mm in diameter and 1 mm gap was used.
2.7. Density
The samples were loaded as extruded pellets and annealed at
240  C for 10 min prior to commencing the experiments. The linear
Density of materials at 23.0  C was measured using an AccuPyc
viscoelastic region of each material was determined at two
1330 Pycnometer. The sample (1e1.5 g) each were placed in a cy-
different temperatures (190  C and 240  C), at 10 rad s1 frequency
lindrical container (5 mL), which was purged with helium. The
under nitrogen atmosphere after being assessed for thermal sta-
purge fill pressure was set to 19.5 psig (134.447 KPa) with an
bility at the same temperatures and frequency. The frequency scans
equilibration rate of 0.005 psig min1 (34.47 Pa min1). The den-
were performed at a minimum of six different temperatures be-
sities reported in Tables 1 and 2 are the average over 10
tween 130  C or 140  C and 240  C under nitrogen atmosphere at
measurements.
10% strain over a frequency range of 0.01e100 rad s1.
The crystalline fraction was calculated using the relationship
equation (1).
3. Results and discussion
ð1  lÞd ¼ ðð1=ra Þ  ð1=rÞÞ=ðð1=ra Þ  ð1=rc ÞÞ (1)
3.1. Processing
where r is the measured density and ra and rc are the densities of
completely amorphous and pure crystalline phase; 0.854 g cm3 Initial experiments on hydrogen-peroxide induced vis breaking
and 0.936 g cm3 respectively [46]. are described in our patent application [34]. These experiments
explored the manner of addition, the hydrogen peroxide concen-
2.8. Wide angle X-ray scattering (WAXS) (conventional X-ray tration and addition rate, the polypropylene feed rate, the screw
source) design and temperature profile, and the effects of additives and
stabilizers.
The measurements were carried out on a Phillips 1130 diffrac- The experiments described here were designed to establish
tometer using CuKa radiation (wavelength 0.1542 nm) operated at the dependence of the extent of vis-breaking of polypropylene on
40 kV and 25 mA. The scattering angle (2q) covered the range from the hydrogen peroxide addition rate. All experiments were con-
2 to 35 (q is Bragg angle) with a step of 0.02 at speed 1.0 min1. ducted on a JSW TEC-30 twin screw L/D 47 operating with a
The pellets were ground to powder under liquid nitrogen and 20 kg h1 throughput of polypropylene. Two grades of

Table 1
Properties of virgin and chain-scissioned polypropylene HP400N.

Material Fraction crystallinity from DSC Tm ( C) (1st heating) Tc ( C) Density (g cm3) Fraction crystallinity from density

1st Heating 2nd Heating

HP400N MFR 11 0.47 0.49 165.2 105.3 0.8893 0.453 ± 0.0012


H2O2 scissioned MFR 44.6 0.45 0.49 164.3 106.5 0.8892 0.452 ± 0.0009
DHBP scissioned MFR 44 0.48 0.49 164.4 106.6 0.8931 0.499 ± 0.0026
H2O2 scissioned MFR 77 0.51 0.50 165.6 106.7 0.8915 0.480 ± 0.0008
100 G. Moad et al. / Polymer Degradation and Stability 117 (2015) 97e108

Table 2
Properties of virgin and chain-scissioned polypropylene HP555G.

Material Fraction crystallinity from DSC Tm ( C) (1st heating) Tc ( C) Density (g cm3) Fraction crystallinity from density

1st Heating 2nd Heating

HP555G MFR 1.5 0.43 0.48 165.6 103.0 0.8900 0.461 ± 0.0016
H2O2 scissioned MFR 7.4 0.44 0.50 165.1 108.2 0.8930 0.498 ± 0.0011
DHBP scissioned MFR 9.8 0.45 0.49 164.8 106.3 0.8912 0.476 ± 0.0010
H2O2 scissioned MFR 24.0 0.48 0.51 164.9 108.8 0.8910 0.474 ± 0.0015

commercial polypropylene were used, namely Lyondell-Basell the DHBP was injected onto the conveyed polypropylene pellets in
HP400N (nominal MFR 11) and HP555G (nominal MFR 1.5). The a manner that emulates the conventional method for organic
aqueous hydrogen peroxide was injected under pressure through peroxide addition during the commercial vis-breaking process. It
a cooled injection nozzle into the polypropylene melt at zone C4, should be noted that when aqueous hydrogen peroxide was added
which is beyond the initial melt seal. The MFR changes observed to the polypropylene prior to the initial melt seal in this way,
are reported in Fig. 2a. Further details are reported in the hydrogen peroxide was not effective in vis-breaking. This is
Supporting information. attributed to the volatility of hydrogen peroxide and consequent
Even though the extruder was not vented, the extrudate loss of the peroxide through the hopper. Small amounts of
exhibited no foaming for addition levels of aqueous peroxide or hydrogen peroxide were detected using moistened peroxide test
water <1000 ppm (volume/volume), and no surface defects for strips in these circumstances.
water/peroxide addition rates <600 ppm (i.e. a smooth strand was
produced that was indistinguishable in appearance from a strand
produced by direct extrusion of polypropylene without addition of
water or peroxide).
Samples of the extrudate were collected for determination of
organic volatiles and for gel count measurements. As expected,
ethanol, acetone and tert-butanol, which are the normal by-
products observed in vis-breaking with DHBP, were below the
limit of detection. Additionally, there was no elevation of hydro-
carbon volatiles, which appeared typical of those observed for
polypropylene formed with DHBP using conventional process. Gel
count analysis showed that the level of gels was similar to or below
those seen with DHBP, and well within normal specifications for
controlled rheology polypropylene [34].
In order to evaluate the importance of shear-induced degrada-
tion under the process conditions, a series of experiments with
HP400N was carried out with addition of equal volume of water
rather than aqueous hydrogen peroxide and otherwise similar
conditions. No vis-breaking, as would be evidenced by an MFR shift,
was observed.
Vis-breaking of polypropylene with aqueous hydrogen peroxide
was directly compared the same process using DHBP. When
injected into the melt steam at 280  C under our processing con-
ditions, DHPB was ineffective and resulted in minimal vis-breaking.
This is attributed to the residence time of the peroxide in the in- Fig. 3. Comparison of the dependence of vis-breaking of commercial polypropylenes
on peroxide addition rate for hydrogen peroxide (corrected for concentration) and 2,5-
jection nozzle and its decomposition prior to injection. Accordingly,
dimethyl-2,5-di-tert-butylperoxyhexane (DHBP).

Fig. 2. Dependence of the extent of vis-breaking on peroxide addition rate for commercial polypropylenes (HP555G or HP400N) with (a) 60 %w/w aqueous hydrogen peroxide or (b)
2,5-dimethyl-2,5-di-tert-butylperoxyhexane (DHBP).
G. Moad et al. / Polymer Degradation and Stability 117 (2015) 97e108 101

A similar series of extrusion experiments was carried out with a significant MFR shift shows that the extent of chain scissioning
DHBP and both HP400N and HP555G. Given the above-mentioned increases with the level of peroxide used, the effect on the molar
factors, the temperature profile for the experiments with DHBP mass averages and MMD of increasing the hydrogen peroxide
was modified such that the set temperature over the injection addition rate from 0.2 mL min1 to 0.333 mL min1 appears small
zone was lower (i.e. C2 160, C3 220, C4 170  C) so as to avoid (within experimental error). The same conclusion follows from the
premature decomposition of the organic peroxide. The results of analysis of rheological data (vide infra).
these experiments are reported in Fig. 2b. After correcting for
peroxide concentration, the hydrogen peroxide is ca 20% more
3.3. Rheological properties
effective than DHBP on a weight or volume basis (Fig. 3). However,
on a “moles of peroxide” basis, the DHBP (molar mass 294 g mol1)
The complex viscosity (h*), the storage (G0 ) and the loss (G00 )
is substantially more effective than hydrogen peroxide (molar
moduli for the selected samples were measured as a function of
mass 34 g mol1).
temperature (T) and frequency (u) using a Rheometrics Scientific
The finding that the vis-breaking effectiveness with both
ARES (Advanced Rheometric Expansion System). The h* values
hydrogen peroxide and DHBP becomes increasingly more effec-
obtained at various temperatures for HP400N and HP555G and for
tive for higher addition rates can be attributed to the peroxide
the derived vis-broken samples, were shifted along the frequency
initially reacting with the stabilizer package present in the
axis to obtain the master curves shown in Fig. 6 at the reference
commercial polypropylene. This hypothesis is supported by the
temperature (T0) of 190  C. The timeetemperature superposition
observation of significantly higher reaction efficiency when
principle (TTS) [48] was applicable over the entire temperature
unstabilized reactor powder is used in the process with hydrogen
range and the derived horizontal shift factors (aT) for the two series
peroxide as initiator [34].
are reported in Fig. 7. The temperature dependence of the shift
factors can be expressed in terms of an Arrhenius relationship
3.2. Size exclusion chromatography (SEC) molar mass distributions (equation (2)):
  
MMD were obtained by size exclusion chromatography (SEC) for Ea 1 1
aT ¼ exp  (2)
HP400N and HP55G and the samples that were vis-broken for a 60% R T T0
aqueous hydrogen peroxide addition rate of 0.2 mL min1
(~600 ppm volume/weight) or 0.333 mL min1 (~1000 ppm). The where R is the universal gas constant, Ea is activation energy for
data are compared with data for a polypropylene scissioned with flow and T0 is the reference temperature. The calculated Ea for flow
0.165 mL min1 DHBP, which provided a similar MFR as that pro- at the reference temperature of 190  C for the HP400N series is
duced with 0.2 mL min1 of hydrogen peroxide. The molar mass 46.0 kJ mol1. A similar value (46.4 kJ mol1) was obtained in the
averages in polystyrene equivalents for these samples are reported case of the HP555G series. These values are consistent with acti-
in Fig. 4 and the MMD are displayed in Fig. 5. vation energies reported in literature for isotactic polypropylene
As anticipated, the distributions shift from the original position [49e53]. The low Ea values for isotactic polypropylene indicate that
towards lower molar mass for the scissioned samples. The vis- flow properties of polypropylene are not strongly temperature
breaking process preferentially cleaves chains of higher molar dependent.
mass, thereby providing a polymer with a narrower MMD. The The temperature window for the rheological measurements of
molar mass averages and the MMD for polymers of similar MFR the isotactic polypropylene melts is limited at the low end by the
formed by vis-breaking with aqueous hydrogen peroxide and DHBP onset of crystallization (melt temperature (Tm) ~ 165  C) and at the
appear similar and the latter are consistent with data in the liter- high end >240  C by degradation. This narrow temperature win-
ature for polymer vis-broken with organic peroxides [12,24]. While dow, combined with a low temperature coefficient of viscosity,


Fig. 4. Effect of vis-breaking with hydrogen peroxide ( ) or 2,5-dimethyl-2,5-di-tert-butylperoxyhexane (DHBP) (B) on molar mass averages (in polystyrene equivalents) and
dispersities for (a) polypropylene HP400N and (b) HP555G. The dotted lines are lines of best fit though the hydrogen peroxide data. Multiple data points indicate replicate
experiments.
102 G. Moad et al. / Polymer Degradation and Stability 117 (2015) 97e108

Fig. 5. Effect of vis-breaking with hydrogen peroxide or 2,5-dimethyl-2,5-di-tert-butylperoxyhexane (DHBP) on molar mass distributions (MMD) for polypropylene (a) HP400N:
virgin polymer (ee), H2O2-scissioned MFR 44.6 ( ), H2O2-scissioned MFR 77 ( ), DHBP-scissioned MFR 44 ( ) and (b) HP555G: virgin polymer (ee), H2O2-scissioned MFR
7.1 ( ), H2O2-scissioned MFR 24, ( ), DHBP-scissioned MFR 9.8 ( ).

Fig. 6. Master curves of the complex viscosity for virgin polypropylene and the vis-broken samples derived from (a) HP400N and (b) HP555G series for the reference temperature
(T0) of 190  C. The continuous lines correspond to the CarreaueYasuda model (equation (4)) fit to the complex viscosity data. The data points correspond to ◊ virgin polymer, D
scissioned with 0.2 mL min1 (~600 ppm) H2O2, B scissioned with 0.165 mL min1 2,5-dimethyl-2,5-di-tert-butylperoxyhexane (DHBP), , scissioned with 0.33 mL min1
(~1000 ppm) H2O2.

Fig. 7. Temperature dependence of the horizontal shift factors for a reference temperature T0 of 190  C for (a) the HP400N series and (b) HP555G series. The dotted lines correspond
to the Arrhenius equation with activation energies for flow Ea of 46.0 kJ mol1 and 46.4 kJ mol1 respectively. The data points correspond to ◊ virgin polymer, D scissioned with
0.2 mL min1 (~600 ppm) H2O2, B scissioned with 0.165 mL min1 2,5-dimethyl-2,5-di-tert-butylperoxyhexane (DHBP), , scissioned with 0.33 mL min1 (~1000 ppm) H2O2.

restricts the range of frequencies that can be obtained in the of the data collected at different temperatures was very good for
master curves [19,50e54]. To perform viscoelastic measurements all samples.
at low temperatures (130 or 140  C) the polypropylene samples The virgin polypropylene samples show shear thinning
were melt annealed then cooled to the test temperature. The behaviour at higher frequencies and a tendency for Newtonian
measurement time was set so as to be shorter than the induction behaviour at the lower frequencies. Consistent with their lower
time for crystallization suggested by DSC experiments. This molar mass averages, the results (Fig. 6a and b) show that the
maximizes the amount of complex viscosity data obtainable at controlled rheology grades are characterized by viscosity values
higher frequencies in the master curves (Fig. 6a and b). The overlap much lower than the virgin polypropylene samples. A shift
G. Moad et al. / Polymer Degradation and Stability 117 (2015) 97e108 103

toward more Newtonian behaviour is also clearly observed for all vis-broken polypropylene with aqueous peroxide and DHBP can be
the controlled rheology polypropylene samples because of their described by the same power law correlation. Indeed, the viscosi-
narrower MMD. Our results show that the flow curves of vis- tyemolar mass power law correlation for many uniform polymers
broken polypropylene obtained with aqueous peroxide and can be described using an exponent a of value 3.4 [48]. In the case
DHBP appear similar. The data are consistent with that reported of polypropylene values a ¼ 3.6 [49,53,59e61], a ¼ 3.5 [51], a ¼ 3.7
in the literature for polypropylene vis-broken with DHBP and [62,63] and a ¼ 3.9 [19] have been reported for the power law
other organic peroxides [8,19,54e56]. exponent for both isotactic and atactic polypropylene. A value of
The zero shear viscosity h0 at T ¼ 190  C of the investigated a ¼ 3.4 has been reported for metallocene polypropylene with a
samples was evaluated from the master curves as the limiting value relatively low dispersity (Mw/Mn ~2) [51,52,64,65]. Our results are,
of the complex viscosity (equation (3)): therefore, consistent with the literature results. The “Mw rule” for
the viscosityemolar mass power law correlation generally can be
h0 ¼ lim h* (3)
u/0 satisfactorily applied for mildly disperse polymers [66,67]. How-
In the case of the vis-broken polypropylene samples, there is a ever, the relationship begins to break down for higher dispersity
clear plateau in the viscosity at the lowest frequencies, indicating polymers. In this latter case, Zeichner and Patel [61] enclosed the
that the Newtonian region has been reached in the frequency (Mw/Mn)0.84 term in the viscosityemolar mass power law correla-
window investigated. In contrast, the limiting region was not tion, while Wasserman and Graessley [53] preferred to use a Mz/Mw
observed for the virgin polypropylene. In all the cases, the esti- term because it better reflects the breadth of the high molar mass
mate of h0 was obtained using a fit of the experimental viscosity tail of the MMD.
data to the CarreaueYasuda model (equation (4)) as shown in In general, the rheological behaviour of polymer melts
Fig. 6a and b [57]. strongly depends on their MMD. The increase in the breadth of
 n1 the MMD of polymers influences the shape of the viscosity vs
h ¼ h0 1 þ ðluÞa a (4) shearerate curve, which shows a decrease in the characteristic
shear rate for the onset of non-Newtonian behaviour. This is
where h0, l a and n are the fitting parameters. clearly evident from the results reported in Fig. 6 for the virgin
The CarreaueYasuda model zero shear viscosity values as a polypropylene, which are characterized by a broad MMD, and for
function of the SEC Mw are reported as a logelog plot in Fig. 8. the controlled rheology polypropylene melts, which have a nar-
The viscosity h0 data can be correlated with the scaling relation rower MMD. Concerning the linear viscoelastic response, the ef-
(equation (5)) fects of dispersity are reflected in the broadening of the
relaxation spectrum in the plateau and terminal regions. Recently
a
h0 ¼ K Mw (5) different blending laws have been proposed to describe the
mixing effect on the rheological behaviour of linear flexible
where K is a material constant and a is a scaling exponent. disperse polymers. Mixing rules have been applied both to the
The powerelaw correlation should hold for linear polymers that flow curve [68,69] as well as to the viscoelastic properties of
contain fractions with molar masses larger than the critical molar entangled polymer melts [70e77]. Among these, the double-
mass, Mc (~7000 g mol1 for polypropylene [58]) where entan- reptation model [72e74,76] has been successfully applied to
glements begin to have a significant effect on the chain dynamics. describe the effects of MMD in the plateau and terminal zone of
The scaling relation of the h0 data is shown in Fig. 8 as a straight line entangled polymer melts. The theory of the effect of chain-
with a slope of 3.6. The results demonstrate that the h0 values of scission of polypropylene (and whether it occurs randomly) on
molar mass has been described by Mead [78]. Simple mixing
rules for the melt flow index and the steady-state recoverable
compliance consistent with the double reptation model were also
derived by Mead [78] to predict the effect of the reactive extru-
sion vis-breaking of polypropylene on the resulting linear visco-
elastic material properties.
Viscoelastic measurements are a well-established and practical
experimental method for melt characterization. The “inverse
problem” of obtaining a MMD by analysis of linear viscoelastic
behaviour has been the subject of many papers [79e89] since, in
principle, rheological data is a powerful method for deriving the
MMD of flexible polymers that is complementary or an alternative
method to SEC and to polymers that are difficult to dissolve or
insoluble in common solvents. The process used for obtaining the
MMD and the molar mass averages from viscoelastic measure-
ments is described in the following section.

3.4. Molar mass distributions from rheological properties

The rheological behaviour from the viscoelastic measurements


of the virgin and controlled rheology samples has been analysed to
obtain the MMD and the Mn, Mw and Mz molar mass averages. An
inversion procedure for converting the linear viscoelastic data of
polymer melts in terms of the storage (G0 ) and loss (G00 ) moduli into
Fig. 8. Plot of the zero-shear viscosity (evaluated from the fit of complex viscosity data
MMD was implemented in previous works by Nobile and Cocchini
to the CarreaueYasuda model) as a function of the weight average molar mass Mw (as [84e86]. The problem was divided into two well-defined steps: the
obtained from SEC measurements) for the HP400N and the HP555G series of samples. first was to define a direct relationship, based on physical concepts,
104 G. Moad et al. / Polymer Degradation and Stability 117 (2015) 97e108

which gives the rheological behaviour as a function of the MMD;


the second consists in the use of a mathematical method to invert
the above relationship.
In the direct way, a viscoelastic model [84,86] was developed to
calculate G0 and G00 from known MMD. It was based on the double
reptation mixing rule [72e74,76] and the MMD dependent kernel
introduced by Thimm et al. [83,90] correlating the relaxation time
spectrum h(t) to the MMD function w(M) (equation (6)).
0 ∞ 1
  Z 0Þ
~ 2 wðM
h½tðMÞ ¼ hðMÞ ¼ G0N wðMÞ@ dM0 A (6)
a M0
M

where a z 3.4 is the scaling exponent for the uniform case in the
scaling relation for the relaxation time as a function of the molar
mass M and G0N is the plateau modulus.
This relation was generalized by Nobile and Cocchini [86] to a
BaumgaerteleSchausbergereWinter (BSW) form [91]; so the
direct relationship developed in Nobile and Cocchini [86] was
equation (7):

Z∞ ~
h½MðlÞ dl
n
hðtÞ ¼ nt (7)
ln l
t

with the scaling relation equation (8)

nþ1 a
l ¼ lðMÞ ¼ k M (8)
n

where n is the BSW parameter [91].


The model viscoelastic functions G0 and G00 can then be obtained
from the relaxation time spectrum through well-known relation-
ships [48] and are related to the MMD as shown in equation (9) and
equation (10).

Z∞
ðutÞ2 dt
G0 ðuÞ ¼ hðtÞ (9) Fig. 9. Master curves of (a) the storage modulus,G0 , and (b) loss modulus, G00 for the
1 þ ðutÞ2 t HP400N series. Reference temperature, T0 ¼ 190  C. The data points are ◊ virgin
0
polymer, D scissioned with 0.2 mL min1 (~600 ppm) H2O2, B scissioned with
0.165 mL min1 2,5-dimethyl-2,5-di-tert-butylperoxyhexane (DHBP).
Z∞
00 ut dt
G ðuÞ ¼ hðtÞ (10)
1 þ ðutÞ2 t
0
In this work, as in Nobile and Cocchini previous works [84e86]
Regardless of the approach used to correlate the MMD to the the w(M) in Equation (5) has been constrained to the GEX function.
rheological properties, the inversion of such relations is an ill-posed Therefore, the MMD function and the model rheological functions
00
problem, as the MMD is very sensitive to small changes in the G0GEX and GGEX are univocally identified by the three GEX parame-
rheological measurements. The approaches used in literature rely ters a, b and M0. The inversion procedure consists in fitting the
00 00
on the use of some a priori knowledge of the behaviour of the MMD model viscoelastic functions G0GEX and GGEX to the G0j and Gj (j ¼ 1, 2,
[79e89]. In particular, the three-parameter generalised exponential … N) raw data (Fig. 9; HP400N series) through a numerical mini-
function (GEX) [92,93] is able to describe a wide class of uniform mization of the objective c2 function in terms of the three degrees
MMD in terms of three free parameters, a, b and M0 (equation (11)) of freedom (equation (13)).

 aþ1 "   #

2
M b N
G0  G0
X

GEX a; b; M0 ; uj

b M
j
wGEX ða; b; M0 ; MÞ ¼   exp  (11) 2
c ða; b; M0 Þ ¼

G aþ1
M0 M0

j¼1
εG0j

b

00
2
N
G  G

00
X
j GEX a; b; M0 ; uj

This function gives rise to the meaningful triplets of the Mn, Mw, þ

(13)
Mz averages (equation (12)):
00
εGj

j¼1
     
G aþ1 G aþ2 G aþ3 where a constant relative error ε on the dynamic data is assumed.
b b b
Mn ¼ M0  ; Mw ¼ M0   and Mz ¼ M0   The MMD curves as well as the Mn, Mw and Mz averages for the
G ab aþ1 virgin polypropylene samples have been obtained from the rheo-
G G aþ2
b b logical G0 and G00 raw data, according to the described procedure.
(12) The MMD obtained from the inversion procedure applied to the G0
and G00 rheological data is shown in Fig. 10 for the HP400N together
G. Moad et al. / Polymer Degradation and Stability 117 (2015) 97e108 105

Fig. 10. Comparison of the SEC molar mass distributions (MMD) obtained directly from Fig. 12. Comparison of the SEC molar mass distributions (MMD) obtained directly from
SEC analysis (d) and from application of the rheological GEX model (e e e) for virgin SEC analysis (d) and from application of the rheological GEX model (e e e) for
HP400N polypropylene. HP555G polypropylene vis-broken with aqueous hydrogen peroxide to give MFR 7.1
(lower traces) and with 2,5-dimethyl-2,5-di-tert-butylperoxyhexane (DHBP) to give
MFR 9.8 (upper traces).

Therefore the Mn value for each of the controlled rheology poly-


propylene samples has been constrained to the corresponding Mn
value evaluated from the G0 and G00 data of the virgin sample. The
procedure is developed in detail in Nobile et al. [94]. In Figs. 11
and 12, the MMD obtained from the inversion procedure (rheo-
logical GEX model) applied to rheological data of polymers of
similar MFR formed by vis-breaking with aqueous peroxide and
DHBP appear similar and compare well with the corresponding
SEC results.
The results of the inversion procedure applied to the rheological
data of both the HP400N and HP555G series in terms of the molar
mass averages are reported in Fig. 13a and b respectively and
compared with the corresponding averages from SEC analysis. The
molar mass averages obtained from the rheological analysis for
polymers of similar MFR formed by vis-breaking with aqueous
peroxide and DHBP appear similar and are generally consistent
with the corresponding results from SEC analysis.
Importantly, in the present context, the results confirm the close
Fig. 11. Comparison of the SEC molar mass distributions (MMD) obtained directly from
SEC analysis (d) and from application of the rheological GEX model (e e e) for
similarity of polypropylenes of the same or similar MFR obtained
HP400N polypropylene vis-broken with aqueous hydrogen peroxide to give MFR 44.6 with hydrogen-peroxide-induced vis-breaking and DHBP-induced
(lower traces) and with 2,5-dimethyl-2,5-di-tert-butylperoxyhexane (DHBP) to give vis-breaking.
MFR 44.0 (upper traces).
3.5. Crystallinity (WAXS, SAXS, DSC, density)

with the corresponding SEC results. The predictions based on The crystallinity of the polypropylene as extruded was examined
rheological analysis closely match the MMD of the polypropylene by differential scanning calorimetry (DSC) (1st heating), density
obtained from SEC measurements. The molar mass averages are measurements [46] and wide and small angle X-ray diffraction
compared in Fig. 13. Similar findings were made for virgin HP555G (WAXS and SAXS). The data are presented in Table 1 (DSC, density
(Fig. 14). for HP400N series), Table 2 (DSC, density for HP555G series), Table 3
In the case of the vis-broken samples, whose MMD curves shift (WAXS and SAXS for HP400N series) and Table 4 (WAXS and SAXS
from the original position towards lower molar mass, the lack of for HP555G series) and the fraction crystallinity obtained by the
viscoelastic experimental data at high frequencies creates unac- various methods is compared in Fig. 15. The methods (other than
ceptable uncertainty in the estimate of Mn from the rheological SAXS) provide a consistent value for the fraction crystallinity of the
model. In this work, a further constraint has thus been introduced virgin polymer.
in the rheological model for the inversion of the G0 and G00 data of Analysis of the WAXS (conventional X-ray source) (Fig. 13) in-
the controlled rheology polypropylene samples in order to obtain dicates that the crystallinity of the vis-broken product is reduced
their MMD. In that the vis-breaking process preferentially cleaves from the precursor polypropylene (Table 3, Table 4). On the other
chains of higher molar mass, a dramatic decrease is expected in hand, The DSC (1st heating) and density measurements suggest
Mz and Mw while only minor changes are expected to occur for in that the crystallinity, if anything, increases on scissioning (Table 1,
Mn. This approximation is also justified by the SEC analysis. Table 2). This apparent discrepancy prompted us to obtain further
106 G. Moad et al. / Polymer Degradation and Stability 117 (2015) 97e108

Fig. 13. Comparison of molar mass averages and dispersities obtained by fitting the rheological data to the GEX MMD model with “constrained Mn” [94] for the vis-broken samples
(squares) and evaluated SEC measurements (circles, average of 2 measurements; refer Fig. 4) for (a) the HP400N series and (b) the HP555G series. Lines are lines of best fit through
the rheological data. Note that in the analysis of the rheological data, the Mn values were constrained to a fixed value (see text).

Fig. 14. Comparison of wide angle X-ray (WAXS e conventional X-ray source) for (a) the HP400N series and (b) the HP555G series. From bottom to top are shown the virgin
polymer, scissioned with 0.2 mL min1 (~600 ppm) H2O2, scissioned with 0.165 mL min1 2,5-dimethyl-2,5-di-tert-butylperoxyhexane (DHBP) and scissioned with 0.33 mL min1
(~1000 ppm) H2O2.

WAXS with much improved signal to noise using the Australian It is clear that crystallinity of the hydrogen peroxide-scissioned
Synchrotron. However, the data obtained showed that the samples polypropylene and DHBP-scissioned polypropylene are similar
had changed over 2 years but nonetheless confirmed the earlier regardless of the method by which it was determined.
findings with respect to fraction crystallinity. SAXS were obtained
simultaneously with WAXS in the synchrotron experiments. The 3.6. Crystallizability (CRYSTAF)
fraction crystallinity suggested by the SAXS measurements is very
different to that indicated by the other techniques for both virgin Crystallization analysis fractionation (CRYSTAF) in which a solu-
and scissioned polypropylene. This can be attributed to the tion of polypropylene is slowly cooled is a measure of the crystal-
different definition of fraction crystallinity inherent in SAXS anal- lizability of the polypropylene. This analysis showed no significant
ysis [95]. SAXS is sensitive to the amorphous component in be- differences between the commercial polymer and the vis-broken
tween the crystalline lamellae, while the analysis of WAXS is polymer or between hydrogen peroxide- and DHBP-scissioned
sensitive to the overall fraction crystallinity. samples. The traces all exhibited a single sharp peak accounting for
The DSC (2nd heating) suggests that the crystallinity is the ~90e91% of the sample for HP400N and 91e92% for HP555G
same for both virgin and scissioned samples, which suggests that (Table 5). This peak is attributed to the crystallization of isotactic
any differences seen for the “as produced” samples reflect the polypropylene. Six replicate runs were conducted for the two DHBP-
history and differences in the morphology or spherulite size, scissioned samples which showed little variability in the fraction
rather than any intrinsic differences in the polypropylene. The crystallized but some variation in the crystallization temperature.
results of CRYSTAF measurements (below) also support this The controlled rheology polypropylenes are, however, very
hypothesis. different from polypropylene produced in thermo-oxidative
G. Moad et al. / Polymer Degradation and Stability 117 (2015) 97e108 107

Table 3
X-ray diffraction properties of virgin and chain-scissioned polypropylene HP400N.

Material Conventional WAXSa Fraction crystallinity Synchrotron SAXS

Conventional WAXSb Synchrotron WAXSb

HP400N MFR 11 0.47 ± 0.002 0.45 0.45 0.75


H2O2 scissioned MFR 44.6 0.39 ± 0.001 0.38 0.39 0.82
DHBP scissioned MFR 44 0.40 ± 0.002 0.38 0.41 0.74
H2O2 scissioned MFR 77 0.38 ± 0.002 0.36 0.35 0.77
a
Data processed using Traces® v6 (GBC Scientific Equipment, Hampshire, IL). Error bars indicate the reproducibility of the determination.
b
Data processed with Origin v8 (OriginLab, Northampton, MA).

Table 4
X-ray diffraction properties of virgin and chain-scissioned polypropylene HP555G.

Material Conventional WAXSa Fraction crystallinity Synchrotron SAXS

Conventional WAXSb Synchrotron WAXSb

HP400N MFR 11 0.45 ± 0.001 0.43 0.43 0.81


H2O2 scissioned MFR 44.6 0.40 ± 0.002 0.39 0.37 0.80
DHBP scissioned MFR 44 0.40 ± 0.002 0.38 0.42 0.81
H2O2 scissioned MFR 77 0.35 ± 0.001 0.38 0.40 0.79
a
Data processed using Traces® v6 (GBC Scientific Equipment, Hampshire, IL). Error bars indicate the reproducibility of the determination.
b
Data processed with Origin v8 (OriginLab, Northampton, MA).

Fig. 15. Dependence of fraction crystallinity on vis-breaking.

degradation [39]. The CRYSTAF trace is generally unaffected by broken polypropylenes showed no spectral features consistent
changes in molar mass but is sensitive to changes in chemical with oxidative scissioning (e.g., carbonyl absorptions).
composition or microstructure [96].
The CRYSTAF results are therefore qualitative evidence that 4. Conclusions
there is no significant functionalization of the polypropylene
backbone during the vis-breaking process with either hydrogen Aqueous hydrogen peroxide has been shown to be an effective
peroxide or DHBP. Consistent with this, FTIR spectra of the vis- reagent for chain scissioning or vis-breaking polypropylene to
produce a controlled rheology polymer in a melt extrusion process.
The process produces a polypropylene with similar characteristics
Table 5 (molar mass, MMD, melt rheology, crystallinity) to that produced
CRYSTAF data for HP400N, HP555G and scissioned samples. with an organic peroxide (DHBP) according to the conventional
Sample Peak temperature ( C) Fraction procedure, but the hydrogen peroxide-scissioned polypropylene
HP400N MFR 11 78.4 0.909
has the distinct advantage that it contains no organic volatile
H2O2 scissioned MFR 44.6 78.2 0.901 contaminants attributable to use of a peroxide.
DHBP scissioned MFR 44 74.8±2a 0.912±1a
H2O2 scissioned MFR 77 78.6 0.898
HP555G MFR 1.5 78.3 0.921 Acknowledgements
H2O2 Scissioned MFR 7.4 68.4 0.926
DHBP Scissioned MFR 9.8 74.4±2a 0.911±1a High temperature SEC and CRYSTAF measurements were per-
H2O2 Scissioned MFR 24.0 72.5 0.921 formed by Robert Bruell, Fraunhofer for Structural Durability and
a
Average over two determinations. System Reliability, Darmstadt.
108 G. Moad et al. / Polymer Degradation and Stability 117 (2015) 97e108

Appendix A. Supplementary data [42] Gigue re PA, Liu ID. Can J Chem 1957;35(4):283e93.
[43] Troe J. Combust Flame 2011;158(4):594e601.
[44] Moad G, O'Shea MS, Peeters G, Postma A. Modification of polyolefins.
Supplementary data related to this article can be found at http:// Advanced Polymerik; 2009. WO2009012523.
dx.doi.org/10.1016/j.polymdegradstab.2015.04.001. [45] ASTM standard E313-10 standard practice for calculating yellowness and
whiteness indices from instrumentally measured color coordinates. West
Conshohocken, PA: ASTM International; 2010.
References [46] Isasi JR, Mandelkern L, Galante MJ, Alamo RG. J Polym Sci B Polym Phys
1999;37(4):323e34.
[1] Xanthos M. Process analysis from reaction fundamentals. In: Xanthos M, ed- [47] Zhu P-W, Tung J, Phillips A, Edward G. Macromolecules 2006;39(5):1821e31.
itor. Reactive extrusion. Munich: Hanser; 1992. p. 33e55. [48] Ferry JD. Viscoelastic properties of polymers. 3rd ed. New York: Wiley; 1980.
[2] Brown SB. Reactive extrusion: a survey of chemical reactions of monomers p. 264e320.
and polymers during extrusion processing. In: Xanthos M, editor. Reactive [49] Pearson DS, Fetters LJ, Younghouse LB, Mays JW. Macromolecules 1988;21(2):
extrusion. Munich: Hanser; 1992. p. 75e199. 478e84.
[3] Moad G. Prog Polym Sci 1999;24:81e142. [50] Mavridis H, Shroff RN. Polym Eng Sci 1992;32(23):1778e91.
[4] Hogt AH, Meijer J, Jenenic J. Modification of polypropylene by organic per- [51] Eckstein A, Friedrich C, Lobbrecht A, Spitz R, Mulhaupt R. Acta Polym
oxides. In: Al-Malaika S, editor. Reactive modifiers for polymers. London: 1997;48(1e2):41e6.
Chapman & Hall; 1997. p. 84e132. [52] Eckstein A, Suhm J, Friedrich C, Maier RD, Sassmannshausen J, Bochmann M,
[5] Azizi H, Ghasemi I. Polym Test 2004;23(2):137e43. et al. Macromolecules 1998;31(4):1335e40.
[6] Azizi H, Ghasemi I. Iran Polym J 2005;14(5):465e71. [53] Wasserman SH, Graessley WW. Polym Eng Sci 1996;36(6):852e61.
[7] Pong S-H, Lu C-P, Wang M-C, Chiu S-HJ. Appl Polym Sci 2005;95(2):280e9. [54] Bernreitner K, Neissl W, Gahleitner M. Polym Test 1992;11(2):89e100.
[8] Azizi H, Ghasemi I, Karrabi Q. Polym Test 2008;27(5):548e54. [55] Tzoganakis C, Vlachopoulos J, Hamielec AE, Shinozaki DM. Polym Eng Sci
[9] Echeverrigaray S, Cruz RD, Oliveira RB. Polym Bull 2012:1e14. 1989;29(6):390e6.
[10] Mun ~ oz PAR, Bettini SHP. J Appl Polym Sci 2013;127(1):87e95. [56] Barakos G, Mitsoulis E, Tzoganakis C, Kajiwara T. J Appl Polym Sci 1996;59(3):
[11] Garrett GE, Mueller E, Pratt DA, Parent JS. Macromolecules 2014;47(2): 543e56.
544e51. [57] Yasuda K, Armstrong RC, Cohen RE. Rheol Acta 1981;20(2):163e78.
[12] Tollefson NM, Roy S, Shepard TA, Long WJ. Polym Eng Sci 1996;36(1):117e25. [58] Wool RP. Macromolecules 1993;26(7):1564e9.
[13] Pabedinskas A, Cluett WR, Balke ST. Polym Eng Sci 1989;29(15):993e1003. [59] Petragli G, Coen A. Polym Eng Sci 1970;10(2):79.
[14] Tzoganakis C, Tang Y, Vlachopoulos J, Hamielec AE. Polym Plast Technol Eng [60] Plazek DL, Plazek DJ. Macromolecules 1983;16(9):1469e75.
1989;28(3):319e50. [61] Zeichner GR, Patel PD. In: Proc 2nd world congr chem eng Montreal; 1981.
[15] Ryu SH, Gogos CG, Xanthos M. Adv Polym Technol 1991;11(2):121e31. p. 6e9.
[16] Triacca VJ, Gloor PE, Zhu S, Hrymak AN, Hamielec AE. Polym Eng Sci [62] Carrot C, Revenu P, Guillet J. J Appl Polym Sci 1996;61(11):1887e97.
1993;33(8):445e54. [63] Minoshima W, White JL, Spruiell JE. Polym Eng Sci 1980;20(17):1166e76.
[17] Tzoganakis C. Can J Chem Eng 1994;72(4):749e54. [64] Aguilar M, Vega JF, Pena B, Martinez-Salazar J. Polymer 2003;44(5):1401e7.
[18] Krell MJ, Brandolin A, Valles EM. Polym React Eng 1994;2(4):389e408. [65] Fujiyama M, Inata H. J Appl Polym Sci 2002;84(12):2157e70.
[19] Berzin F, Vergnes B, Delamare L. J Appl Polym Sci 2001;80(8):1243e52. [66] Struglinski MJ, Graessley WW. Macromolecules 1985;18(12):2630e43.
[20] Machado AV, Maia JM, Canevarolo SV, Covas JA. J Appl Polym Sci 2004;91(4): [67] Wasserman SH, Graessley WW. J Rheol 1992;36(4):543e72.
2711e20. [68] Bersted BH. J Appl Polym Sci 1975;19(8):2167e77.
[21] Oliveira JA, Biscaia EC, Fadigas JC, Pinto JC. Macromol Mater Eng 2006;291(5): [69] Malkin AY, Blinova NK, Vinograd Gv, Zabugina MP, Sabsai OY, Shalgano Vc,
552e70. et al. Eur Polym J 1974;10(5):445e51.
[22] Kalantari B, Semnani Rahbar R, Mojtahedi MRM, Mousavi Shoushtari SA, [70] Graessley WW, Struglinski MJ. Macromolecules 1986;19(6):1754e60.
Khosroshahi A. J Appl Polym Sci 2007;105(4):2287e93. [71] Rubinstein M, Colby RH. J Chem Phys 1988;89(8):5291e306.
[23] Sirin K, Yavuz M, Çanli M, Avci A, Dogan F. J Polym Mater 2013;30(4):381e96. [72] Tsenoglou C. Polym Prepr Am Chem Soc Div Polym Chem 1987;28(2):185e6.
[24] Scorah MJ, Zhu S, Psarreas A, McManus NT, Dhib R, Tzoganakis C, et al. Polym [73] Descloizeaux J. Europhys Lett 1988;5(5):437e42.
Eng Sci 2009;49(9):1760e6. [74] Descloizeaux J. Macromolecules 1990;23(17):3992e4006.
[25] Tzoganakis C, Vlachopoulos J, Hamielec AE. Polym Eng Sci 1988;28(3): [75] Maier D, Eckstein A, Friedrich C, Honerkamp J. J Rheol 1998;42(5):1153e73.
170e80. [76] Tsenoglou C. Macromolecules 1991;24(8):1762e7.
[26] Iedema PD, Willems C, van Vliet G, Bunge W, Mutsers SMP, Hoefsloot HCJ. [77] Cassagnau P, Montfort JP, Marin G, Monge P. Rheol Acta 1993;32(2):156e67.
Chem Eng Sci 2001;56(12):3659e69. [78] Mead DW. J Appl Polym Sci 1995;57(2):151e73.
[27] Zummallen M. Controlled-rheology polypropylene, method of preparation [79] Tuminello WH. Polym Eng Sci 1986;26(19):1339e47.
and article. USA: Dow Global Technologies Inc.; 2010. US20100324225A1. [80] Mead DW. J Rheol 1994;38(6):1797e827.
[28] Aubert M, Roth M, Pfaendner R, Wile n C-E. Macromol Mater Eng 2007;292(6): [81] Wasserman SH. J Rheol 1995;39(3):601e25.
707e14. [82] Carrot C, Guillet J. J Rheol 1997;41(5):1203e20.
[29] Pfaendner R. C. R. Chim 2006;9(11e12):1338e44. [83] Thimm W, Friedrich C, Roths T, Honerkamp J. J Rheol 2000;44(6):1353e61.
[30] Psarreas A, Tzoganakis C, McManus N, Penlidis A. Polym Eng Sci 2007;47(12): [84] Cocchini F, Nobile MR. Rheol Acta 2003;42(3):232e42.
2118e23. [85] Nobile MR, Cocchini F. Rheol Acta 2001;40(2):111e9.
[31] Schneider A, Roth M. Chem Fibers Int 2006;56:177e9. [86] Nobile MR, Cocchini F. Rheol Acta 2008;47(5e6):509e19.
[32] He G, Tzoganakis C. Polym Eng Sci 2011;51(1):151e7. [87] Leonardi F, Allal A, Marin G. J Rheol 2002;46(1):209e24.
[33] Tripathi D. Practical guide to polypropylene. Shrewsbury, UK: iSmithers Rapra [88] Van Ruymbeke E, Keunings R, Bailly C. J Newt Fluid 2002;105(2e3):153e75.
Publishing; 2002p 11. [89] Guzman JD, Schieber JD, Pollard R. Rheol Acta 2005;44(4):342e51.
[34] Dagley IJ, Moad G, Nichols LV. Modification of propylene polymers. Polymers [90] Thimm W, Friedrich C, Marth M, Honerkamp J. J Rheol 1999;43(6):1663e72.
CRC; 2012. WO2012000022. [91] Baumgaertel M, Schausberger A, Winter HH. Rheol Acta 1990;29(5):400e8.
[35] Zhou W, Zhu S. Ind Eng Chem Res 1997;36(4):1130e5. [92] Gloor WE. J Appl Polym Sci 1978;22(5):1177e82.
[36] Bertin D, Grimaldi S, Leblanc M, Marque SRA, Siri D, Tordo P. J Mol Struct [93] Gloor WE. J Appl Polym Sci 1983;28(2):795e805.
Theochem 2007;811(1e3):255e66. [94] Nobile MR, Cocchini F, Dagley IJ, Habsuda J, Li G, Moad G, et al., 2015. To be
[37] Bertin D, Leblanc M, Marque SRA, Siri D. Polym Degr Stab 2010;95(5):782e91. submitted (http://www.sciencedirect.com/science/article/pii/S01413910150
[38] Gryn'ova G, Hodgson JL, Coote ML. Org Biomol Chem 2011;9(2):480e90. 0124X).
[39] de Goede S, Brüll R, Pasch H, Marshall N. Macromol Symp 2003;193(1):35e44. [95] Rabiej S, Wlochowicz A. Angew Makromol Chem 1990;175(1):81e97.
[40] Dokolas P, Loer SM, Solomon DH. Aust J Chem 1998;51(12):1113e20. [96] Anantawaraskul S, Soares JP, Wood-Adams P. Adv Polym Sci 2005;182:1e54.
[41] Camara S, Gilbert BC, Meier RJ, van Duin M, Whitwood AC. Org Biomol Chem
2003;1(7):1181e90.

View publication stats

You might also like