Вы находитесь на странице: 1из 16

Determination of method-invariant activation energies

of long-chain branched low-density polyethylenes


Ute Kener, Joachim Kaschta, and Helmut Mnstedta)
Institute of Polymer Materials, Friedrich-Alexander-University
Erlangen-Nrnberg, Martensstrasse 7, D-91058 Erlangen, Germany
(Received 29 January 2009; final revision received 6 April 2009

Synopsis
The idea to use the temperature dependence of rheological properties, especially the flow
activation energy, as a tool to investigate branching structures is well-known from literature.
However, there is no common method to obtain activation energies, which are independent of the
measuring quantity chosen, particularly, in the case of slightly thermorheologically complex
polymers like low-density polyethylene LDPE. Hence, differing activation energies result, which
cannot unequivocally be correlated with the branching structure. This paper describes a method for
the determination of method-independent activation energies for thermorheologically complex
polymers like LDPE. From a generalized approach to the time-temperature superposition principle,
a vertical shift factor is introduced, which is related to the temperature dependence of the linear
steady-state compliance. In the case of the complex LDPE, a decrease in the linear steady-state
compliance with temperature is found. Taking this experimentally determined shift factor into
account leads to activation energies independent of the rheological quantity chosen. These values
can be taken to analyze differences of the branching architecture.

2009 The Society of Rheology. DOI: 10.1122/1.3124682

I. MOTIVATION
The time-temperature superposition and following from that the construction of master
curves has been known from literature since 70 years. Initially this method developed
empirically, but in 1950 Ferry established the general theoretical background for the
superposition of linear viscoelastic properties of polymers Ferry 1950; 1980.
This principle is based on the fact that, in the case of thermorheologically simple
polymers, all relaxation times involved exhibit the same temperature dependence, i.e.,

T = T0aTT,T0,

with T being the actual and T0 being the reference temperature.


The shift factor aT is a function of temperature. For polymer melts at temperatures not
too far away from the glass transition temperature Tg, the so-called WLF-equation holds
a

Author to whom correspondence should be addressed; electronic mail: helmut.muenstedt@ww.uni-erlangen.de

2009 by The Society of Rheology, Inc.


J. Rheol. 534, 1001-1016 July/August 2009

0148-6055/2009/534/1001/16/$27.00

1001

Redistribution subject to SOR license or copyright; see http://scitation.aip.org/content/sor/journal/jor2/info/about. Downloaded to IP:


147.27.58.93 On: Tue, 21 Apr 2015 17:43:40

1002

KENER, KASCHTA, AND MNSTEDT

log aTT,T0 =

d1T T0
,
d2 + T T0

d1 and d2 are constants specific for a material.


For temperatures significantly higher than Tg, an Arrhenius-equation of the type

aT = exp

Ea 1 1

R T T0

is valid. Ea is the activation energy and R is the universal gas constant.


Deduced from the Rouse theory Ferry 1950; Rouse 1953 an additional vertical
modulus shift factor bT is introduced according to Eq. 4,
GT =

1
GT0.
bT

This additional contribution is required in order to correctly superimpose relaxation


moduli in accordance with the theory. bT considers the change of the density with the
temperature T and is defined as
bT =

T 0 0
.
T

According to Eq. 4 the relaxation modulus G requires a vertical shift with bT. However,
this vertical shift is very small, if the measuring temperatures differ less than 100 C see
Sec. IV. In many cases this shift lies within the limits of the uncertainty of the measurement and hence is often neglected.
The time-temperature superposition principle is widely used in practice in order to
extend the time or frequency range of rheological measurements. This method reduces
the experimental effort to cover the large time window often required for a complete
description of the rheological behavior of polymer melts.
The flow activation energy varies with the molecular structure and branching topography. For example, branched polymers exhibit higher activation energies than linear
ones Bersted 1985. Therefore, analyzing the temperature dependence of rheological
properties may become interesting for getting an insight into the branching structure of
polymers.
Several studies in literature already deal with the relationship between molecular
structure and rheological properties of polyethylenes. Some of them try to interpret temperature dependencies and flow activation energies in relation to branching e.g., Raju et
al. 1979; Carella et al. 1986; Wasserman and Graessley 1996; Vega and Santamara
1998; Shroff and Mavridis 1999; Vega et al. 1999; Villar et al. 2001; Wood-Adams
and Costeux 2001; Starck et al. 2002; Bonchev et al. 2003; Ye et al. 2005.
However, there is no convincing method for the determination of activation energies,
which are independent of the rheological quantities taken and, therefore, suitable for
analytical purposes. Mostly, the zero-shear viscosities are used, but the data obtained are
significantly different in many cases from the results derived from the storage, loss, or
complex modulus. Sometimes even the relaxation spectra are used to determine the
activation energy. The results differ widely depending on the evaluation method making
the activation energy a somewhat dubious tool for the characterization of the molecular
structure of polymers. This situation becomes evident from the literature discussed in the
following.

Redistribution subject to SOR license or copyright; see http://scitation.aip.org/content/sor/journal/jor2/info/about. Downloaded to IP:


147.27.58.93 On: Tue, 21 Apr 2015 17:43:40

ACTIVATION ENERGIES OF LDPE

1003

Linear, high density polyethylene HDPE is known as thermorheologically simple. In


the literature, an activation energy of about 27 kJ/mol independent of molar mass and
molar mass distribution is reported e.g., Raju et al. 1979; Wasserman and Graessley
1996; Stadler et al. 2007. In contrast, in the case of commercial low-density polyethylene LDPE both thermorheological simplicity and complexity are reported in the
literature. For example, Jacovic et al. 1979, Meiner 1987, and Laun 1987 investigated the activation energy of LDPE as a function of the shear stress far into the nonlinear viscoelastic regime with capillary rheometry. Thermorheological complexity resulting
in a stress-dependent activation energy was observed. Moreover, activation energies derived from different rheological properties provide different activation energies in the
case of the same LDPE. For example, Wood-Adams and Costeux 2001 published activation energies of thermorheologically complex polyethylenes long-chain branched HDPEs resulting from the storage and loss modulus, the zero-shear viscosity, as well as their
spectra, which do not match each other.
Others determined the activation energy simply by means of zero-shear viscosities.
The resulting shift factors are often checked by constructing master curves of G and
G in double-logarithmic plots. The data may superimpose in a coarse approximation,
concealing a slight but physically relevant thermorheological complexity if regarded
more precisely. Such a distinct value of the activation energy is often found for LDPE in
the literature. In general, the insertion of long-chain branches is believed to be responsible for the thermorheological complexity of polyethylenes. A sensitive but qualitative
way to check the thermorheological behavior offers the plot of the phase angle versus
the absolute value of the complex modulus. Data of thermorheologically simple materials
measured at different temperatures lead to a good superposition, while data of thermorheologically complex materials split up with temperature van Gurp and Palmen
1998.
Mavridis and Shroff 1992 published a new approach toward the evaluation of activation energies of thermorheologically complex polymers like LDPE, which show a
split-up in G. They found that there exists a vertical modulus shift larger than the
value predicted by Eqs. 4 and 5, which allows the construction of a master curve. This
shift is empirically determined and applied toward the data in such a way that a thermorheologically simple behavior is obtained. By that way a constant, stress-independent
activation energy is determined. This study is often cited and the vertical shift is accordingly performed. However, there exists no explanation so far, why the vertical shift
should be larger than the prediction from the theory of rubber elasticity and what its
physical meaning is. Thus, the question remains how to gain reliable physically based
activation energies.
Additional contradictions in literature result from partly incomplete molecular characterizations making it difficult to relate activation energies to a definite architecture of the
polymer.
The aim of this paper is to establish a method for the undisputable determination of
activation energies. Consequently, the careful analysis of the temperature dependence of
rheological properties combined with a reliable molecular characterization will open up a
new approach toward the use of activation energies for an analytical characterization of
branching. This paper lays the experimental foundations for such a method.
II. MATERIALS
The molecular structure was characterized by means of a high-temperature size exclusion chromatograph PL 220, Varian Inc. equipped with Shodex columns UT 807 1x

Redistribution subject to SOR license or copyright; see http://scitation.aip.org/content/sor/journal/jor2/info/about. Downloaded to IP:


147.27.58.93 On: Tue, 21 Apr 2015 17:43:40

1004

KENER, KASCHTA, AND MNSTEDT


TABLE I. Molecular characteristics.

LLDPE L63
LDPE 2
LDPE 3

Tm
C

Mw
kg/mol

Mz/Mw

Mw/Mn

123.5
110.0
106.7

113
230
150

1.7
17.2
18.3

3.2
13.7
12.0

and UT 806M 3x coupled with a multiangle laser light scattering MALLS apparatus
Wyatt Dawn EOS, Wyatt Corp. at 140 C. 1,2,4-trichlorobenzene TCB was used as
the solvent.
The molecular characteristics are given in Table I. LDPE 2 and LDPE 3 are commercial tubular grades, which differ in molecular weight. The LLDPE L63 is a commercial
linear low density polyethylene with 1.2 mol.-% hexene as comonomer from NMR
measurements synthesized by a metallocene catalyst. Figure 1 shows the determined
radii of gyration r2g0.5 as a function of the molar mass MLS. r2g0.5 smaller than 20 nm
cannot be detected. In the case of the LLDPE L63, the linear relationship known from
literature is valid indicating the absence of long-chain branches Beer et al. 1999. The
long-chain branches in LDPE result in a contraction of the molecules and thus the radii of
gyration deflect from the linear relationship between r2g0.5 and MLS. The deviation of the
two LDPE investigated from the linear reference is similar in the range of equal molar
masses. LDPE 2 possesses higher molar masses than LDPE 3, which are significantly
branched.
The radii of gyration of the two LDPE have been truncated at molar masses below
200 000 g/mol as indications for non-ideal separation in the GPC became visible. In this
molar mass regime, highly branched molecules of very high molar mass Podzimek et
al.2001 and/or high molar mass stars Frater et al. 1997 may elute together with low
molar mass species making an interpretation with respect to structure impossible.

FIG. 1. Radii of gyration as a function of molar mass.

Redistribution subject to SOR license or copyright; see http://scitation.aip.org/content/sor/journal/jor2/info/about. Downloaded to IP:


147.27.58.93 On: Tue, 21 Apr 2015 17:43:40

ACTIVATION ENERGIES OF LDPE

1005

FIG. 2. G-plot of the three polyethylenes. LDPE 3 is shifted by the factor of 2 along the modulus axis for
the matter of a better distinction.

III. RHEOLOGY
The polymer pellets were stabilized with the antioxidants Irganox 1010 and Irgafos
168 from Ciba 0.5 wt % each. Disks with a diameter of 25 mm and a height of 2 mm
were hot-pressed under vacuum.
Dynamic-mechanical as well as creep and creep-recovery experiments were conducted
with a Bohlin Gemini C-VOR rheometer using a plate-plate geometry. The measurements
were performed under a nitrogen atmosphere granting a sufficient thermal stability. Data
for each material were obtained for at least four different temperatures in the range from
130 to 210 C. A new sample was used for each measurement. Reproducibility was
verified.
Stress-controlled dynamic-mechanical experiments were performed in the linear regime for frequencies between 100 and 0.01 s1. Stationarity and linearity with regard to
creep and creep-recovery experiments were ensured according to the method described in
detail in the literature e.g., Gabriel et al. 1998.
IV. RESULTS AND DISCUSSION
The thermorheological behavior in the linear viscoelastic regime can qualitatively be
distinguished by plotting the phase angle as a function of the magnitude of the complex
modulus at different temperatures. Figure 2 shows this plot for the three polyethylenes
investigated. In the case of the LLDPE L63, all curves superimpose pointing at a
thermorheological simplicity. In contrast, the curves of the two branched polyethylenes
do not superimpose. There is a slight split-up between the data at different temperatures.
A clear trend with temperature can be observed: the higher the temperature the larger the
phase angle at the same modulus. This behavior is a result of thermorheological complexity, which is therewith affirmed. G is more sensitive toward changes in the
temperature dependency than double-logarithmic plots of rheological data like the storage
and loss moduli as a function of the angular frequency.
The slight thermorheological complexity found for the two LDPE samples is in contrast to the observation of Stadler et al. 2008b that the LDPE investigated there, is
thermorheologically simple. The reason for this discrepancy of the very similar products

Redistribution subject to SOR license or copyright; see http://scitation.aip.org/content/sor/journal/jor2/info/about. Downloaded to IP:


147.27.58.93 On: Tue, 21 Apr 2015 17:43:40

1006

KENER, KASCHTA, AND MNSTEDT

FIG. 3. a Arrhenius plot of the shift factors from different rheological properties in the terminal zone for
L63. b Arrhenius plot of the shift factors from different rheological properties in the terminal zone for LDPE
3.

is due to the use of different rheometers and various degrees of experimental diligence. It
demonstrates the care which has to be taken to get reliable information on the thermorheological behavior of polymer melts.
In the following, the thermorheological behavior of the polyethylenes is analyzed in
more detail. For the linear polyethylene and LDPE 3, the flow activation energy can be
determined from rheological properties in the terminal zone like the storage, loss, or
complex modulus or the zero-shear viscosity. For this purpose, the shift factors aT were
determined and plotted according to the Arrhenius relationship given by Eq. 3. The flow
activation energy is then derived from the slope of linear fits of the experimental values
obtained. Figures 3a and 3b show the Arrhenius plots of the differently determined
shift factors for the LLDPE L63 and LDPE 3. The thermorheological simplicity of the
LLDPE L63 is again confirmed. All rheological quantities analyzed lead to the same
flow activation energy within the experimental error. The mean value of about 29 kJ/mol
is in accordance with the literature e.g., Stadler et al. 2007. Such a behavior cannot be
found for the LDPE 3. The flow activation energy derived from the phase angle is the
highest. The value from the storage modulus G is about 4.6 kJ/mol lower, that one from
the loss modulus G even differs by 8.3 kJ/mol. However, there is an agreement between
the results from G and the zero-shear viscosity 0 cf. Fig. 3b. These results are
confirmed by several studies in the literature. For example, Wood-Adams and Costeux
2001 simulated the spectra of a thermorheologically complex material, which led also
to differing activation energies in dependence on the rheological quantity chosen.
A detailed analysis of the activation energies in dependence on various rheological
quantities outside the terminal zone is shown for LDPE 2 and LDPE 3 in Figs. 4 and 5.
The evaluation of the phase angle results in approximately constant values, while the
analysis of the moduli leads to lower moduli-dependent activation energies, which again
reflect the thermorheological complexity of LDPE. They are found to decrease with
growing moduli. These results demonstrate the difficulties to determine a materialspecific activation energy. For the determination of the activation energies, a vertical shift
see Eq. 4 has been neglected.
Some more light on these findings is thrown by a more generalized approach to the
time-temperature superposition principle, which includes a modulus shift and a time or
frequency shift. Such a procedure was proposed by Verser and Maxwell 1970 in the
following way:

Redistribution subject to SOR license or copyright; see http://scitation.aip.org/content/sor/journal/jor2/info/about. Downloaded to IP:


147.27.58.93 On: Tue, 21 Apr 2015 17:43:40

ACTIVATION ENERGIES OF LDPE

1007

FIG. 4. Activation energies of LDPE 2 derived from different rheological properties outside the terminal zone.

G,T0 = bTGaT,T,

G,T0 = bTGaT,T,

G,T0 = bTGaT,T,

tan ,T0 = tan aT,T.

From these definitions follows:

According to these relationships, G should have the same shift factors as G and G,
but tan and, consequently, the phase angle is independent of the vertical shift factor
b T.

FIG. 5. Activation energies of LDPE 3 derived from different rheological properties outside the terminal zone.

Redistribution subject to SOR license or copyright; see http://scitation.aip.org/content/sor/journal/jor2/info/about. Downloaded to IP:


147.27.58.93 On: Tue, 21 Apr 2015 17:43:40

1008

KENER, KASCHTA, AND MNSTEDT

As can be concluded from Figs. 4 and 5, the activation energies of LDPE 2 and LDPE
3 determined from the temperature shift of the phase angle are independent of within
the accuracy of the measurements indicating a thermorheologically simple behavior of
these two materials. However, the activation energies calculated from the shift factors of
G, G, and G, respectively, are distinct functions of the moduli. This finding can be
interpreted as the result of a thermorheologically complex behavior.
The fact that a constant activation energy is found from the phase angle in comparison to the contradictory results from G, G, and G supports the supposition that the
choice of the vertical shift factor bT plays a decisive role for the determination of the
activation energy, particularly, as according to Eq. 9 tan is not affected by a vertical
shift of the moduli. Therefore, it can be assumed that this value may be the right quantity
characterizing a material and that by choosing adequate vertical shift factors an adaption
of the activation energies from the other methods can be obtained. The question remaining is how a physically reasonable shift factor bT could be determined. One way to
achieve this goal is sketched in the following.
As for small , i.e. in the terminal range,
G,T = 0T2Je0T2

and
10

G,T = 0T
are valid, it follows from Eq. 6 that

0T02Je0T02 = bT0T2

1
a T2

Je0T2

or

0T02 Je0T0 bT
=
0T2 Je0T aT2

11

and from Eq. 7 that

0T0 = bT0T

aT

or

0T0 bT
= .
0T aT

12

Replacing the ratio of the zero shear-rate viscosities in Eq. 11 by Eq. 12 results in
bT2 Je0T0 bT
=

aT2 Je0T aT2

or

bT =

Je0T
Je0T0

13

This generalized approach by Verser and Maxwell 1970 leads to a vertical shift factor,
which is given by the temperature dependence of the linear steady-state compliance Je0.
If the linear steady-state compliance is not temperature dependent, then bT = 1, i.e., a
vertical shift does not have to be applied, if it is dependent on the temperature, however,
a vertical shift bT given by Eq. 13 has to be considered.
Following from this general derivation of the vertical shift factor, the question arises
which effect the application of the vertical shift factor would have on the thermorheological complexity presented in Figs. 4 and 5. For this purpose, the temperature dependence of Je0 has to be known.

Redistribution subject to SOR license or copyright; see http://scitation.aip.org/content/sor/journal/jor2/info/about. Downloaded to IP:


147.27.58.93 On: Tue, 21 Apr 2015 17:43:40

ACTIVATION ENERGIES OF LDPE

1009

FIG. 6. Exemplary results of creep and creep-recovery experiments on LDPE 2. tc is the creep time; tr is the
time for recovery.

There are several studies concerning molecular influences like molar mass and molar
mass distribution on the elastic properties of polyolefins e.g., Carella et al. 1986;
Carella and Graessley 1984; Takigawa et al. 2006, but almost none regarding their
temperature dependencies. In the literature, linear steady-state compliances are mostly
determined from the storage and loss modulus in the terminal regime of dynamicmechanical experiments. In many cases, this method leads to inaccurate results, particularly if the moduli are low or the terminal regime is not reached at frequencies typically
chosen not smaller than 0.01 s1. Due to the experimental limitations, the terminal regime for long-chain branched polyethylenes normally is not accessible. Expanding the
frequency range to lower frequencies results in such long measuring times that the thermal stability of the material may cause problems.
A convenient experimental way to determine linear steady-state compliances more
reliably are creep and creep-recovery experiments. These measurements are time consuming and demanding with respect to equipment and know-how. For example, linearity
and stationarity are crucial and the thermal stability of the material investigated has to be
taken into account. Details about creep and creep-recovery experiments and their performance can be found in the literature, e.g., Gabriel et al. 1998, Gabriel and Mnstedt
1999, and Stadler and Mnstedt 2008a.
Linear steady-state compliances at different temperatures were measured in creeprecovery experiments, in order to determine the vertical shift factor bT defined by Eq.
13. In Fig. 6 the results of creep and creep-recovery experiments for LDPE 2 at 170 C
are presented as an example for typical investigations performed. Three experiments
differing in the preceding creep time tc or the applied creep stress are shown. Each set of
parameters leads to the same steady-state compliance of 6.7 104 Pa1, i.e., the linear
steady-state value of J0e is obtained.
Figure 7 compares creep-recovery experiments at different temperatures in the range
between 130 and 210 C for LDPE 2. The preceding creep times tc to reach the steadystate had to be chosen the longer, the lower the temperature. The corresponding linear
steady-state compliances distinctly depend on temperature. They are listed in Table II.
Higher temperatures lead to distinctively lower values of the linear steady-state compliance.
Table II gives the linear steady-state compliances Je0 for all the three polyethylenes.

Redistribution subject to SOR license or copyright; see http://scitation.aip.org/content/sor/journal/jor2/info/about. Downloaded to IP:


147.27.58.93 On: Tue, 21 Apr 2015 17:43:40

1010

KENER, KASCHTA, AND MNSTEDT

FIG. 7. Creep-recovery experiments at different temperatures for LDPE 2.

The values result from at least three different measurements at each temperature. The
LLDPE L63 exhibits a temperature-invariant linear steady-state compliance of about
3.8 105 Pa1, which is in the order of typical linear polyethylenes but significantly
lower than the values for LDPE 2 e.g., Pedersen and Ram 1987; Gabriel and Mnstedt
1999; Pen et al. 2003; Stadler and Mnstedt 2008a. The temperature independece
of Je0 results in bT = 1 cf. Eq. 13 and is in good agreement with the thermorheological
simplicity of LLDPE L6-3 demonstrated in Figs. 2 and 3a.
A temperature dependence of Je0 is found for LDPE 3, too. The values are significantly
lower, however, than those of LDPE 2.
The question arises now whether master curves can be obtained by applying the
vertical shift factor bT experimentally determined from the temperature dependence of
the linear steady-state compliances.
Figure 8 shows the G-plot of the two long-chain branched polyethylenes, applying the vertical modulus shift. The curves measured at different temperatures superimpose well indicating that accounting for bT by using the temperature-dependent steadystate compliances yields thermorheological simplicity. Using the temperature-dependent
density function of Fig. 9 for a vertical shift does not result in a master curve of a
comparable perfection.
The logarithms of the linear steady-state compliances of the LDPE investigated are
plotted in Fig. 9 as functions of the inverse absolute temperature. In the temperature
range given, they follow an Arrhenius relationship. There is a slight difference between
TABLE II. Linear steady-state compliances at different temperatures for the materials investigated.
J e0
10 Pa1
5

T
C
LLDPE L63
LDPE 2
LDPE 3

130

140

150

170

190

210

n.d.
81.3 0.4
44.8 0.8

3.8 0.4
n.d.
n.d.

3.7 0.3
73.5 0.6
41.2 0.8

3.8 0.2
66.6 0.9
38.1 0.7

3.9 0.3
59.9 0.8
34.8 0.3

3.9 0.8
55.0 0.5
32.4 0.5

Redistribution subject to SOR license or copyright; see http://scitation.aip.org/content/sor/journal/jor2/info/about. Downloaded to IP:


147.27.58.93 On: Tue, 21 Apr 2015 17:43:40

ACTIVATION ENERGIES OF LDPE

1011

FIG. 8. G-plot of the two long-chain branched polyethylenes considering the vertical shift factor bT.
LDPE 3 is shifted along the modulus axis by the factor of 2 for the matter of better distinction.

the temperature dependencies of the two branched polyethylenes, but they are significantly stronger than that of the density of LDPE 2 plotted in Fig. 9, for comparison.
For a quantitative discussion, the measured functions GaT , T and GaT , T were
corrected by bT and then the activation energies determined. The results are given in Fig.
10. Within the accuracy of the measurements the activation energies obtained from the
corrected shifts are the same as that from the phase angle . An average activation energy
of 67 kJ/mol is found for the LDPE 2.
The same method was applied to LDPE 3. The results are plotted in Fig. 11. Again,
identical activation energies can be found independently of the rheological quantity
evaluated, if the vertical shift factor bT is considered. The activation energy follows as 56
kJ/mol. The activation energies obtained for the two LDPE are significantly different
indicating the possibility of using them for the characterization of the molecular structure.
Determining the activation energies directly from the temperature dependence of the
zero-shear viscosities as done frequently in the literature naturally gives values smaller

FIG. 9. Temperature dependence of the linear steady-state compliance branched polyethylenes in comparison
to the density correction.

Redistribution subject to SOR license or copyright; see http://scitation.aip.org/content/sor/journal/jor2/info/about. Downloaded to IP:


147.27.58.93 On: Tue, 21 Apr 2015 17:43:40

1012

KENER, KASCHTA, AND MNSTEDT

FIG. 10. Activation energies for LDPE 2 obtained from different vertically corrected rheological properties.

than those from the other methods after applying a vertical shift factor. Considering the
vertical shift factor according to Eq. 13, however, leads to distinctively higher values,
which are consistent with the analysis from other rheological properties. The correction
of the zero-shear viscosities according to Eq. 12 leads also to a flow activation energy
around 67 kJ/mol.
V. CONCLUSIONS
The long-chain branched polyethylenes investigated are thermorheologically complex.
Their activation energies are dependent on the determination method chosen and, except

FIG. 11. Activation energies for LDPE 3 obtained from different vertically corrected rheological properties.

Redistribution subject to SOR license or copyright; see http://scitation.aip.org/content/sor/journal/jor2/info/about. Downloaded to IP:


147.27.58.93 On: Tue, 21 Apr 2015 17:43:40

ACTIVATION ENERGIES OF LDPE

1013

for the phase angle evaluation, not constant. For example, G and G cannot be
shifted to give a master curve and, therefore, the activation energy is not a constant
quantity.
It was shown that the introduction of a vertical shift factor bT as defined in Eq. 13
leads to satisfactory master curves and following from that to constant activation energies. This shift factor does not take the temperature dependence of the density into
account as it is usually done but is based on the temperature dependence of the linear
steady-state compliance.
However, it is not obvious, why the linear steady-state compliance should change with
temperature. Generally, the linear recoverable compliances Jrt can be related to their
retardation spectra in the following way:
Jrt = Ji1 et/i.

14

i are the retardation times and Ji the retardation strengths. For a thermorheologically
simple polymer the temperature dependence of the compliance is given by
Jrt,T = Ji1 eaTTt/i,

15

indicating that all retardation times i are shifted by the same shift factor aTT if the
temperature is changed. From these simple relationships, it becomes obvious that a temperature change leaves the shape of Jrt , T unaffected and that the linear steady-state
compliance
Je0T = lim JrT = Ji
t

16

is temperature independent.
Such a behavior is found for the LLDPE L63 cf. Table II, which is known to be
thermorheologically simple.
In the case of the thermorheologically complex LDPE, the temperature dependence of
Je0 and that of the retardation times i have to be considered. Applying these shifts a
master curve can be obtained as Figs. 12a and 12b demonstrate. This experimental
result can be described by introducing temperature-dependent retardation strengths JkT,
i.e.,
Jrt,T = JiT1 eaTTt/i

17

JiT = bTT,T0JiT0.

18

with

The thermorheological complexity of the two LDPE investigated, therefore, can formally
be interpreted by a temperature dependence of the retardation strengths in addition to the
usual temperature shift of the retardation times found for thermorheologically simple
materials.
Recent investigations point out, however, that the separability of the temperature dependence as found for the LDPE of this paper is not generally valid for long-chain
branched polyethylenes.
The question remains, however, why the linear steady-state compliance should change
with temperature. The decrease in the steady-state compliances with increasing temperature points to a temperature-dependent variation in the entanglement structure. A similar

Redistribution subject to SOR license or copyright; see http://scitation.aip.org/content/sor/journal/jor2/info/about. Downloaded to IP:


147.27.58.93 On: Tue, 21 Apr 2015 17:43:40

1014

KENER, KASCHTA, AND MNSTEDT

FIG. 12. a Vertically shifted creep-recovery experiments of LDPE 2. b Vertically and horizontally shifted
creep-recovery experiments of LDPE 2.

but mechanically induced change of the entanglement structure is postulated to explain


the findings of the influence of various deformation histories on rheological properties of
LDPE Mnstedt 1981; Rokudai and Fujiki 1981. A quantitative explanation of the
peculiar behavior typical of LDPE is still missing, however.
Nevertheless, the significant difference of the activation energies of LDPE 2 and
LDPE 3 offers the possibility to use the thermorheological behavior as an analytical tool.
There is a considerable difference of about 10 kJ/mol between the activation energies of
LDPE 2 and LDPE 3 if the vertical shift factor is considered.
The molecular characterization cf. Fig. 1 revealed a significant difference of the
long-chain branching structure of the two LDPE: LDPE 2 is long-chain branched at
higher molar masses than LDPE 3. But measurements of such kind are not easy and need
a sophisticated equipment. The determination of the activation energy is very sensitive to
discriminate between these structural differences and not difficult to perform from the
experimental point of view. The long-chain branches of the high molar masses of LDPE
2 could be related to the higher activation energy. The contribution of long-chain
branches at low molar masses to the activation energy, as found in LDPE 3, seems to be
weaker. This finding raises the question, which influence the chain microstructure has on
the thermorheological behavior and how relationships could be used for analytical purposes.
VI. SUMMARY
Polyethylenes can be thermorheologically simple or complex. LLDPE is consistently
classified as thermorheologically simple in the literature. However, in the case of LDPE
thermorheological simplicity and complexity are reported. A sensitive qualitative proof of
the thermorheological behavior is the van Gurp and Palmen plot, which uncovered a
slight thermorheological complexity of the two branched polyethylenes of this study. The
determination of flow activation energies from different rheological properties led to
diverging values. Moreover, a temperature dependence of the linear steady-state compliances investigated with creep and creep-recovery experiments was observed for the two
LDPE, but not for the LLDPE L63. Their temperature dependence is distinctly stronger
than that of the density often used in the literature as a vertical shift factor.
For a more generalized approach of the time-temperature superposition principle, a
vertical shift factor bT was introduced for the shift of G and G. A simple consideration
within the terminal zone leads to an interpretation of the vertical shift factor by the ratio

Redistribution subject to SOR license or copyright; see http://scitation.aip.org/content/sor/journal/jor2/info/about. Downloaded to IP:


147.27.58.93 On: Tue, 21 Apr 2015 17:43:40

ACTIVATION ENERGIES OF LDPE

1015

of the temperature-dependent linear steady-state compliances. Thermorheological complexity could be avoided by applying the factor bT to the moduli before the timetemperature superposition. In the case of the phase angle, the factor bT is not involved.
Therefore, the phase angle yields the true activation energy without applying the newly
proposed correction factor. The evaluation of the activation energy from the zero-shear
viscosities also needs the vertical shift factor bT. A constant activation energy independent of the rheological quantity chosen could be obtained by applying bT. Consequently,
this correction offers the possibility to determine definite activation energies for various
polyethylenes as it was demonstrated for the two LDPE. These findings open up a promising way to use activation energies for the characterization of long-chain branched
polyethylenes. The question how these activation energies are related to various molecular structures is being addressed in more detail in further investigations.
ACKNOWLEDGMENTS
The authors express their gratitude to Dr. F. Stadler for discussions and to Ir. X.
Drooghag Universit catholique de Louvain for the NMR analysis of the LLDPE L63.

References
Beer, F., G. Capaccio, and L. Rose, High molecular weight tail and long-chain branching in SRM 1476
polyethylene, J. Appl. Polym. Sci. 73, 28072812 1999.
Bersted, B. H., On the effects of very low levels of long chain branching on rheological behavior in polyethylene, J. Appl. Polym. Sci. 30, 37513765 1985.
Bonchev, D., A. H. Dekmezian, E. Markel, and A. Faldi, Topology-rheology regression models for monodisperse linear and branched polyethylenes, J. Appl. Polym. Sci. 90, 26482656 2003.
Carella, J. M., and W. W. Graessley, Effects of chain microstructure on the viscoelastic properties of linear
polymer melts: Polybutadienes and hydrogenated polybutadienes, Macromolecules 17, 27752786 1984.
Carella, J. M., J. T. Gotro, and W. W. Graessley, Thermorheological effects of long-chain branching in
entangled polymer melts, Macromolecules 19, 659667 1986.
Ferry, J. D., Mechanical properties of substances of high molecular weight. VI. Dispersion in concentrated
polymer solutions and its dependence on temperature and concentration, J. Am. Chem. Soc. 72, 3746
3752 1950.
Ferry, J. D., Viscoelastic Properties of Polymers, 3rd ed. Wiley, New York, 1980.
Frater, D. J., J. W. Mays, and C. Jackson, Synthesis and dilute solution properties of divinylbenzene-linked
polystyrene stars with mixed arms lengths: Evidence for coupled stars, J. Polym. Sci., Part B: Polym. Phys.
35, 141151 1997.
Gabriel, C., J. Kaschta, and H. Mnstedt, Influence of molecular structure on rheological properties of polyethylenes. I. Creep recovery measurements in shear, Rheol. Acta 37, 720 1998.
Gabriel, C., and H. Mnstedt, Creep recovery behavior of metallocene linear low-density polyethylenes,
Rheol. Acta 38, 393403 1999.
Jacovic, M. S., D. Pollock, and R. S. Porter, A rheological study of long branching in polyethylene by
blending, J. Appl. Polym. Sci. 23, 517527 1979.
Laun, H. M., Orientation of macromolecules and elastic deformations in polymer melts. Influence of molecular
structure on the reptation of molecules, Prog. Colloid Polym. Sci. 75, 111139 1987.
Mavridis, H., and R. N. Shroff, Temperature dependence of polyolefin melt rheology, Polym. Eng. Sci. 32,
17781791 1992.
Meiner, J., Proceedings of the 4th International Congress on Rheology Interscience, New York, 1987, Part 3,
p. 437.
Mnstedt, H., The influence of various deformation histories on elongational properties of low density poly-

Redistribution subject to SOR license or copyright; see http://scitation.aip.org/content/sor/journal/jor2/info/about. Downloaded to IP:


147.27.58.93 On: Tue, 21 Apr 2015 17:43:40

1016

KENER, KASCHTA, AND MNSTEDT

ethylene, Colloid Polym. Sci. 259, 966972 1981.


Pedersen, S., and A. Ram, Prediction of rheological properties of well-characterized branched polyethylenes
from the distribution of molecular weight and long chain branches, Polym. Eng. Sci. 18, 990995 1978.
Pen, J., C. Domnguez, J. F. Vega, M. Aroca, and J. Martnez-Salazar, Viscoelastic behaviour of metallocenecatalysed polyethylene and low density polyethylene blends: Use of the double reptation and Palierne
viscoelastic models, J. Mater. Sci. 38, 47574764 2003.
Podzimek, P., T. Vlcek, and C. Johann, Characterization of branched polymers by size exclusion chromatography coupled with multiangle light scattering detector. I. Size exclusion chromatography elution behavior
of branched Polymers, J. Appl. Polym. Sci. 81, 15881594 2001.
Raju, V. R., G. G. Smith, G. Marin, J. R. Knox, and W. W. Graessley, Properties of amorphous and crystallizable hydrocarbon polymers. I. Melt rheology of fractions of linear polyethylene, J. Polym. Sci. 17,
1831195 1979.
Rokudai, M., and T. Fujiki, Influence of shearing history on the rheological properties and processability of
branched polymers. IV. Capillary flow and die swell of low-density polyethylene, J. Appl. Polym. Sci. 26,
13431350 1981.
Rouse, P. E., A theory of the linear viscoelastic properties of dilute solutions of coiling polymers, J. Chem.
Phys. 21, 12721280 1953.
Stadler, F. J., C. Gabriel, and H. Mnstedt, Influence of short-chain branching of polyethylenes on the temperature dependence of rheological properties in shear, Macromol. Chem. Phys. 208, 24492454 2007.
Stadler, F. J., and H. Mnstedt, Terminal viscous and elastic properties of linear ethene/-olefin copolymers,
J. Rheol. 52, 697712 2008a.
Stadler, F. J., J. Kaschta, and H. Mnstedt, Thermorheological behavior of various long-chain branched
polyethylenes, Macromolecules 41, 13281333 2008b.
Shroff, R. N., and H. Mavridis, Long-chain-branching index for essentially linear polyethylenes, Macromolecules 32, 84548464 1999.
Starck, P., A. Malmberg, and B. Lfgren, Thermal and rheological studies on the molecular composition and
structure of metallocene- and ZieglerNatta-catalyzed ethylene-olefin copolymers, J. Appl. Polym. Sci.
83, 11401156 2002.
Takigawa, T., H. Kadoya, T. Miki, T. Yamamoto, and T. Masuda, Dependence of zero-shear viscosity and
steady-state compliance on molecular weight between entanglements for ethylenecycloolefin copolymers,
Polymer 47, 48114815 2006.
van Gurp, M., and J. Palmen, Time-temperature superposition for polymeric blends, Rheol. Bull. 67, 58
1998.
Vega, J. F., and A. Santamara, Small-amplitude oscillatory shear flow measurements as a tool to detect very
low amounts of long chain branching in polyethylenes, Macromolecules 31, 36393647 1998.
Vega, J. F., M. Fernndez, A. Santamara, A. Muoz-Escalona, and P. Lafuente, Rheological criteria to characterize metallocene catalyzed polyethylenes, Macromol. Chem. Phys. 200, 22572268 1999.
Verser, D. W., and B. Maxwell, Temperature dependence of the properties of low-density polyethylene,
Polym. Eng. Sci. 10, 122130 1970.
Villar, M. A., M. D. Failla, R. Quijada, R. S. Mauler, E. M. Valls, G. B. Galland, and L. M. Qunizani,
Rheological charaterisation of molten ethylene--olefin copolymers synthesized with EtInd2ZrCl2/MAO
catalyst, Polymer 42, 92699279 2001.
Wasserman, S. H., and W. W. Graessley, Prediction of linear viscoelastic response for entangled polyolefin
melts from molecular weight distribution, Polym. Eng. Sci. 36, 852861 1996.
Wood-Adams, P., and S. Costeux, Thermorheological behavior of polyethylene: Effects of microstructure and
long chain branching, Macromolecules 34, 62816290 2001.
Ye, Z., F. AlObaidi, S. Zhu, and R. Subramanian, Long-chain branching and rheological properties of ethylene1-hexene copolymers synthesized from ethylene stock by concurrent tandem catalysis, Macromol. Chem.
Phys. 206, 20962105 2005.

Redistribution subject to SOR license or copyright; see http://scitation.aip.org/content/sor/journal/jor2/info/about. Downloaded to IP:


147.27.58.93 On: Tue, 21 Apr 2015 17:43:40

Вам также может понравиться