Vous êtes sur la page 1sur 10

Rheol Acta (2015) 54:159168

DOI 10.1007/s00397-014-0822-y

ORIGINAL CONTRIBUTION

The effect of molecular structure on rheological behavior


of tubular LDPEs
Masood Khabazian Esfahani &
Nadereh Golshan Ebrahimi & Ehsan Khoshbakhti

Received: 2 June 2014 / Revised: 26 October 2014 / Accepted: 10 November 2014 / Published online: 26 November 2014
# Springer-Verlag Berlin Heidelberg 2014

Abstract Molecular characterization of low-density polyethylene has long been faced with many challenges. Due to the
lack of control over the radical polymerization under very
high pressures, the low-density polyethylene (LDPE) molecular structure is very complex. This paper uses the attenuated
total reflectance (ATR)Fourier transform infrared characterization method to introduce a measuring tool to calculate the
average number of branches in LDPE molecular chain.
Among 28 gel permeation chromatography (GPC)-analyzed
LDPEs, two GPC-identical tubular LDPE were chosen. The
branching structures of the LDPEs are investigated by the vanGurp Palmen plot. To clarify the difference between rheological behavior of LDPEs, GPC, temperature rising elution
fractionation, ATR, dynamic oscillatory shear, and
elongational stress growth coefficient tests were performed.
It was found that high molecular weight portion of LDPE 2
has higher number of branches than LDPE 1, which is responsible for the different rheological behavior.
Keywords Rheology . Polyethylene . FTIR . Structure .
Gel permeation chromatography

Introduction
Nowadays, many polymerization methods are employed to
produce different kinds of polymers or even the same polymer
M. Khabazian Esfahani : N. G. Ebrahimi (*) : E. Khoshbakhti
Polymer Engineering Department, Chemical Engineering Faculty,
Tarbiat Modares University, P.O. Box 14115-114, Tehran, Iran
e-mail: ebrahimn@modares.ac.ir
M. Khabazian Esfahani
e-mail: Masood.khabazian@modares.ac.ir
E. Khoshbakhti
e-mail: Ehsan.khoshbakhti@modares.ac.ir

with different molecular structures, each of which is engineered


to meet special processing needs. Characterization of molecular
structure of polymers, on one hand, determines the molecular
structure of existing polymers and, on the other hand, helps
engineers to feedback control the polymer molecular structure
by changing the polymerization control factors. Low-density
polyethylene (LDPE) as a highly consumed commodity polymer is produced radically under high pressures. In such a
condition, polyethylene with many side branches is produced.
Radical polymerization of LDPE under high pressures leads to
the formation of two types of branches, short- and long-chain
branches. The former is created by intramolecular hydrogen
transfer via transient ring formation, whereas the latter is due
to the intermolecular hydrogen transfer (Roedel 1953).
Understanding the topology of branches is very crucial for
processing reasons. In this respect, many characterization
methods have found their way into polymer molecular structure
determination. Gel permeation chromatography (GPC) with
refractive index as detector is a very powerful tool to determine
the molecular characterization of linear polymers (Moore 1964);
however, for branched polymers, this technique faces some
problems. Podzimek (1994) used multiangle laser light scattering (MALLS) coupled with intrinsic viscometer (IV) detector to
determine the side-chain branches. In another research, Yau and
Gillespie (2001) used temperature rising elution fractionation
(TREF) in addition to MALLS and IV detectors to characterize
polyolefins. Tribe et al. (2006) used Fourier transform infrared
(FTIR) as GPC detector to determine short-chain branches in
polyethylene. Surez and Coto (2013) compared GPC-IV with
GPC-MALLS results and showed that GPC-IV is more sensitive
towards lower values of molecular weight whereas GPCMALLS is more sensitive to higher values of molecular weight.
End group analysis by FTIR is also used to characterize short-chain branches (Bryant and Voter 1953) and
long-chain branches in polyethylene (Rugg et al. 1953).
The bending mode of CH3 happens at 1378 cm1.

160

Rheol Acta (2015) 54:159168


Table 1 Peak intensities
of LDPE/HDPE blends
and the pure materials,
comparison of measured
peak intensities and
mixing rule

Fig. 1 Relationship between the number of CH3 groups and a number


average molecular of weight of polymer and b average number of
branches per molecule

This peak represents chain end groups together with


side branches (Krimm et al. 1956). Attenuated total
reflectance (ATR), as a variant to FTIR, is also used
to investigate the chemical bonds. This technique scans
the surface of the material with the penetration depth of
1 m (Schuttlefield and Grassian 2008). For ATR-FTIR,
as the surface of the scanned material and the penetration of evanescent wave are constant, by applying the
same pressure between the sample and the ATR crystal,
one can be sure that the same amount of sample is
scanned by laser beam, which consequently means that
the peaks from one sample can quantitatively be compared with the other samples.
ATR is a light characterization method. Unlike its
counterpart, FTIR, ATR only scans the surface of the
material. The evanescent wave with a penetration depth
Fig. 2 Focused absorbance ATRFTIR peaks

Material

ATR peak
intensities

Mixing
rule

HDPE
HD/LD (75/25)
HD/LD (50/50)
HD/LD (25/75)
LDPE 1

0.32
0.46
0.61
0.80
1

0.49
0.66
0.83

of 1 m travels through the sample. The intensity of


ATR peak depends on the area of scanning surface and
on the applied pressure between the sample and the
ATR crystal. One difficulty of FTIR is that quantitative
comparison of sample is relative; however, having a
constant scan volume in ATR makes this method an
absolute measure of characterization. In other words,
with a constant scan volume and pressure between sample and ATR crystal, the intensity of bending peak of
methyl at 1378 cm1 for the two samples is a comparative scale of CH3 (branches) presence in the samples;
the higher the peak intensity, the higher the number of
branches in the sample. In Fig. 1, circles represent the
ATR crystal; the number of CH3 groups in a constant
volume depends on the number of average molecular
weight of the polymer and the average number of
branches per molecule. For a linear sample, the number
of methyl end groups decreases by increasing the number average molecular weight (Fig. 1a). For a branched
sample, the number of methyl groups increases with
increasing the average number of branches per molecule
(Fig. 1b). To evaluate the sensitivity of ATR-FTIR towards the presence of methyl groups in polyethylene
chemical structure, LDPE 1 is blended with a linear

Rheol Acta (2015) 54:159168


Table 2

161

Fraction temperatures

Fraction

Fr-1

Fr-2

Fr-3

Fr-4

Fr-5

Fr-6

Fr-7

Temperature (C)

40

50

60

70

80

90

100

high-density polyethylene (HDPE) at weight ratios of


25, 50, and 75 %. Figure 2 shows the focused absorbance ATR-FTIR peaks, normalized with respect to
LDPE 1. As concentration of LDPE decreases, the peak
intensity of methyl group at 1378 cm1 also decreases.
Table 1 represents the normalized values of methyl peak
intensities for the blends and the pure materials with the
values from the mixing rule. There is a very good
correlation between the peak intensities and the mixing
rule values, which means that the peak intensity at
1378 cm1 has a linear relation with the concentration
of methyl groups.
Equation (1) describes the relation between the average
numbers of methyl group in a LDPE molecule and the ATR
peak intensity at 1378 cm1.
LDPEnCH3 k

PI1378
LDPEM n

where k is a constant and PI1378 is the intensity of methyl


bending at 1378 cm1. For a specific ATR crystal type and
constant pressure between the sample and the crystal and by
having a linear polyethylene (there are only two methyl
groups existing for each polymer chain), the equation constant
can be determined (k=8.14106).
TREF is also used to investigate the branching structure of LDPE (Kulin et al. 1988). This separation technique is based on the ability of semicrystalline polymers
to crystallize at different temperatures in accordance to
the chain topologies. It is assumed that long-chain

Fig. 3 Gel permeation


chromatography graphs of
LDPEs

branches do not affect the elution temperatures (John


and Ronald 2006).
Parallel to the aforementioned characterization techniques, rheology as an easy to access and powerful
method has made its way into polymer characterization
field. Janzen and Colby (1999) determined low level of
long-chain branches in high-density polyethylene by
melt Newtonian viscosity, which could not be determined by GPC or NMR. Later on, Shroff and
Mavridis used intrinsic and zero shear viscosities to
define long-chain branching index of scarcely branched
polyethylenes (Shroff and Mavridis 1999). Stadler
(2012) utilized rheology as a tool to detect very low
concentration of long-chain branches in metallocene catalyzed polyethylene. It is very common for linear viscoelastic data to plot one property against the other.
Trinkle et al. (2002) used the correlation of loss angle
versus the logarithm of the absolute value of complex
modulus (van-Gurp Palmen plot) to determine the topology of branches in polymer. The van-Gurp Palmen plot
showed sensitivity towards length and amount of
branches for long-chain branched polyethylenes (Lohse
et al. 2002). For nonlinear viscoelastic, on the other
hand, the transient uniaxial/biaxial elongational viscosity
is a very important property. Most of the polymer
processing operations such as fiber spinning, film blowing, and blown molding impose such deformation on
polymer chains (Mnstedt et al. 2006). Some polymers
show the rise in uniaxial elongational viscosity above
trouton ratio. It is demonstrated that the reason for the
well-known strain hardening of polyethylenes is the
presence of long-chain branches (Wood-Adams and
Dealy 2000). Many attempts have been made to model this
behavior; most of these theories are based on Doi and
Edwards (DE) tube model (Doi and Edwards 1978a, b, c).
DE model assumes that, during the deformation, the tube

162
Table 3

Rheol Acta (2015) 54:159168


Molecular characterization of LDPEs

Material

Mn

Mw

PDI

LDPE 1
LDPE 2

13,921
14,162

123,517
128,962

8.87248
9.10562

diameter is constant, and as a result, no chain stretching is


taking place (Doi and Edwards 1987). Consequently, the model predictions for elongational viscosity of long-chain-branched
polymers fall well below the experiments. Molecular stress
function (MSF) is a modification to the DE tube model. This
model can predict the uniaxial elongational viscosity of lowdensity polyethylene with only two fitting parameters (Wagner
et al. 2003). Many attempts have been made to reduce the
nonlinearity of these parameters (Abbasi et al. 2012;
Roln-Garrido and Wagner 2007). The MSF model assumes upon deformation; the backbone segment
stretches, and side-chain branches are compressed. This
assumption introduces the parameter, f, into the DE
extra stress tensor. This parameter is defined as the
relative tension of a tube segment at time t (3kT/a(t))
to that of the initial time (3kT/a0) (Wagner et al. 1998).
 0
f t; t

a0
at; t 0

This paper uses Curie approximation to calculate the


measure strain tensor (Pattamaprom et al. 2000).
For long-chain-branched polymers, the molecular stress
function f is the solution of Eq. (4).
df 2

dt

 2  r!
f 1
f 2
1
 S 11 S 22
S S 
 1 11 22 f 2max 1
2
1
4
f
4

where is Hencky strain rate, is the number of segments in a


tube section form, which one of them is the backbone segment
that stretches and 1 are side chain segments that compress
under the deformation. At very long time, the tube diameter
reaches its minimum diameter; in such a case, the f reaches to
its maximum value, which is expressed in Eq. (4) as fmax. This
indicates the steady-state condition. For some polymers, fmax
is not a constant value; it may change as the value of strain rate
decreases (Abbasi et al. 2012).
Many articles used rheology as a complimentary characterization tool for conventional methods (Roln-Garrido and
Wagner 2014; Rolon-Garrido et al. 2013).

2
Experimental

In Eq. (2), a0 and a(t,t) are the tube segment diameters at


time zero and time t, respectively.
The extra stress tensor, (t), of the MSF model is described
by Eq. (3) (Wagner 2006).

Zt
t

 0  0
 0 0
m tt f 2 t; t S IA
DE t; t dt

In Eq. (3), m(tt) is the memory function, f is the


molecular stress function, and SIA
DE(t,t) is the measure
strain tensor with independent alignment assumption.
Fig. 4 Superimposed storage
modulus of LDPEs

LDPE 1 and LDPE 2 were purchased from Ariya Sasol


Petrochemical Co. and Laleh Petrochemical Co., respectively.
Both LDPEs are radically polymerized in tubular reactors
under high pressure. HDPE is supplied by Bandar Imam
Petrochemical Company.
Dynamic oscillatory shear was performed with Anton paar
MCR 301 at 130, 160, and 190 C under nitrogen atmosphere
and superimposed at 160 C by timetemperature superposition. The amplitude of oscillation is chosen by dynamic amplitude sweep test as to fall within the linear viscoelastic
regime.

Rheol Acta (2015) 54:159168

163

Fig. 5 Superimposed loss


modulus for LDPEs

Table 4 The relaxation spectrum


of LDPEs

LDPE 1

LDPE 2

130 C
Gi

160 C
i

5.74E+01
9.70E+01
3.79E+02
2.64E+01
1.36E+03
7.19E+00
3.37E+03
1.96E+00
7.14E+03
5.32E01
1.29E+04
1.45E01
2.16E+04
3.94E02
3.44E+04
1.07E02
4.70E+04
2.92E03
1.34E+05
7.95E04
0 =39,100 (Pa s)

Gi

7.08E02
2.54E+02
8.61E+00
5.53E+01
3.35E+02
1.20E+01
1.30E+03
2.62E+00
4.55E+03
5.70E01
1.04E+04
1.24E01
1.98E+04
2.70E02
4.04E+04
5.87E03
4.98E+04
1.28E03
1.54E+05
2.78E04
0 =12,700 (Pa s)

All uniaxial elongational viscosity measurements were


performed with a strain control rheometer ARES G2 with
EVF fixture.
Gel permeation chromatography was performed by
Polymer Laboratory PL 210 with IV and RI as detectors.

Fig. 6 Dynamic viscosity of


LDPEs

130 C
Gi

160 C
i

7.94E+02
5.21E+01
4.41E+03
6.09E+00
1.40E+04
7.12E01
3.29E+04
8.33E02
6.69E+04
9.74E03
6.61E+04
1.14E03
2.73E+05
1.33E04
7.22E+05
1.56E05
2.70E+07
1.82E-06
2.36E+05
2.13E07
0 =81,800 (Pa s)

Gi

7.12E01
2.76E+02
4.83E+01
6.22E+01
7.12E+02
1.40E+01
2.09E+03
3.16E+00
5.97E+03
7.11E01
1.20E+04
1.60E01
2.19E+04
3.61E02
3.84E+04
8.13E03
5.72E+04
1.83E03
1.35E+05
4.13E04
0 =27,200 (Pa s)

LDPE is dissolved in 1,2,4-Trichlorobenzene (TCB) at 140 C


with 0.04 w/w. The volumetric flow rate was set to 0.6 ml/min.
The ATR-FTIR spectrum of the samples was collected with
Bruker-Vertex 80. The instrument resolution was set to
1 cm1.

164

Rheol Acta (2015) 54:159168

Fig. 7 van-Gurp Palmen plots of


LDPEs together with a linear
polyethylene

TREF was performed with a home-made apparatus.


The 1 % solution of each sample was prepared in
stabilized xylene (0.1 % Irganox 1010) at 120 C.
After complete dissolution, the solution was pumped
through the column, and subsequently, the temperature
was decreased to 50 C with the rate of 1 C/h. After
30 min, the xylene was pumped through the column
with a flow rate of 20 ml/min and precipitated in
methanol. The pumping was continued until no precipitation is taking place. Afterwards, the temperature was
increased to the second fraction, and this procedure was
repeated until seven fractions are collected (Table 2).
To investigate the thermal stability of the LDPE 1 and
HDPE, time sweep tests were performed with Anton paar
MCR 301 for 10 min under the atmosphere of nitrogen at
150 C. Afterwards, the blending was done in a Brabender
internal mixer at 150 C for 5 min under nitrogen atmosphere.

Fig. 8 Uniaxial elongational viscosity of LDPE 1 at 160 C

Results and discussion


Figure 3 depicts the GPC graphs of LDPEs. Both LDPEs are
nearly identical in GPC viewpoint. Table 3 represents the
molecular characterization of LDPEs, which are derived from
GPC curves. As it is conspicuous, the molecular specifications
of both LDPEs are nearly the same except that LDPE 1 has
higher molecular weight fraction at higher values of molecular
weight.
Dynamic frequency sweep test is performed for each
of the LDPEs at 130, 160, and 190 C and superimposed
at the reference temperature of 160 C. Figures 4 and 5
show the superimposed storage and loss moduli for
LDPEs, respectively. Having fallen on the same trend,
the timetemperature superposition holds for both
LDPEs. Although having the same molecular characterization for GPC, the storage and loss moduli of LDPEs

Rheol Acta (2015) 54:159168

165

Fig. 9 Uniaxial elongational viscosity of LDPE 2 at 160 C

differ considerably specially at low values of frequency.


This shows that GPC data cannot differentiate between
these two rheologically different LDPEs.
Dynamic moduli are fitted with a 10-element Maxwell
model. The relaxation spectrum is reported in Table 4. Later
on, these values are fed to the MSF model to plot the
prediction.
Dynamic viscosity of LDPEs is plotted in Fig. 6. The zero
shear viscosity of LDPE 2 gets a higher value than that of

Fig. 10 Uniaxial elongational viscosity of LDPE 1 at 130 C

LDPE 1. The transition to power law region for LDPE 2


happens sooner than that for LDPE 1. This shows that
LDPE 2 has more branches than LDPE 1.
As it has already been mentioned, the van-Gurp Palmen
plot is very sensitive towards molecular architecture. Figure 7
illustrates the van-Gurp Palmen plots of LDPEs together with
a linear polyethylene. As it can be seen, both plots show the
same trend, which means that both LDPEs have nearly the
same molecular structure. This is an expected result since both

166

Rheol Acta (2015) 54:159168

Fig. 11 Uniaxial elongational viscosity of LDPE 2 at 130 C

LDPEs are produced in tubular reactors. Another feature of


the van-Gurp Palmen plot is the verification of timetemperature superposition (Gurp and Palmen 1998). When the vanGurp Palmen plot of the same polymer at various temperatures

Fig. 12 Temperature rising elution fractions of LDPEs

Fig. 13 Attenuated total


reflectance spectra of LDPEs

holds to the same line, the timetemperature superposition


principle holds. Figure 7 shows that the timetemperature
superposition holds for both LDPEs.
One manifestation of the presence of long-chain branches in
polymer molecular structure is the rise in stress growth coefficient above trouton ratio in startup of extensional flow, the socalled strain hardening. Figure 8 shows the uniaxial elongational
viscosity of LDPE 1 at different strain rates at 160 C. The MSF
model predictions are also depicted in the graph. As it is conspicuous, at high strain rates, the MSF model predictions well
predict the extensional viscosity; however, as the rate of strain
declines, the predictions of the MSF model get higher values
than the experiments. Contrary to LDPE 1, the MSF model
predictions for LDPE 2 holds well even at low values of strain
rates (Fig. 9). Since the zero shear viscosity of LDPE 1 is lower
than that of LDPE 2, discrepancies between the MSF model

Rheol Acta (2015) 54:159168

167

Table 5 Parameters used to calculate the average number of branches


per LDPE molecule
Material

ATR peak intensity


at 1378 cm1

Mn
(kg/mol)

Average number
of branches per
molecule

LDPE 1
LDPE 2

0.020371
0.024926

13,921
14,162

10
12

predictions and the experimental values for LDPE 1 at low strain


rates may be due to differences between molecular structures or
enhanced sagging of LDPE 1 at long measuring times. To clarify
the cause of these discrepancies, uniaxial elongational viscosities
of both LDPEs are also measured at a lower temperature of
130 C. Figures 10 and 11 show the uniaxial elongational
viscosities of LDPE 1 and LDPE 2 at 130 C with the MSF
predictions, respectively. As seen, the MSF model again overpredicts the viscosity of LDPE 1 and under-predicts the viscosity
of LDPE 2 at low strain rates. Therefore, the discrepancies
between the MSF predictions and experimental data are due to
the differences between the molecular structures of LDPEs.
To investigate the branching structure of LDPEs, TREF test
is performed. Figure 12 shows the TREF fractions of LDPEs.
The weight percentage of LDPE 2, at low temperature fractions, is higher than that of LDPE 1; on the other hand, the
weight percentage of LDPE 1 is noticeably higher than that of
the LDPE 2 at high temperature fractions. This shows that the
molecular structure of LDPE 1 is more linear than that of
LDPE 2. This is in accordance with the uniaxial elongational
viscosity results. Since the linear portion of LDPE 1 is higher
than that of LDPE 2, in contrast to LDPE 2, LDPE 1 has not
shown the stain hardening behavior at low strain rates. At low
values of strain rate, time is enough for the relaxation of linear
chains via reptation mechanism. The presence of a high portion of linear molecules in LDPE 1 facilitates the reptation

Fig. 14 Calculated radius of


gyration versus molecular weight
of LDPEs

process. Because of this, the uniaxial elongational viscosity of


LDPE 1 falls well below the MSF model predictions.
To have a quantitative measure of the number of branches
per molecule, this paper uses ATR-FTIR spectroscopy.
Figure 13 shows the ATR spectra of LDPEs. The peak (at
1378 cm1) is related to the concentration of methyl groups in
a constant volume of LDPEs. Equation (1) gives the average
number of methyl groups per molecule. Having two methyl
groups at the end of each polymer chain, the average number
of branches per molecule is equal to the average number of
methyl groups per molecule minus 2. Table 5 gives the values
of peak intensities, number average molecular weight, and
average number of branches per molecule of LDPEs.
Needless to mention, these branches include both short- and
long-chain branches. To understand the branching distribution, Fig. 14 draws the calculated radius of gyration versus
molecular weight of LDPEs together with linear reference. As
seen, both LDPEs have fallen on the linear reference line at
intermediate molecular weights. However, at higher molecular
weights, the radius of gyration of LDPE 2 falls below the
radius of gyration of LDPE 1. This shows that, at high molecular weight portions, the hydrodynamic volume of LDPE 2
is less than that of the LDPE 1, which means that at high
molecular weight interval the LDPE 2 molecule has higher
number of branches per molecule than LDPE 1.

Conclusions
Although both LDPEs have almost the same GPC plots, their
rheological behavior differs considerably in Newtonian and
non-Newtonian regions. It is concluded that the branching
structure of polymer is responsible for the different rheological behaviors. However, van-Gurp Palmen shows that the
topology of branches is the same for both LDPEs. It has been
seen that the predictions of MSF model over-predict the

168

uniaxial elongational viscosity of LDPE 1, which means that


the parameter fmax decreases as the strain rate decreases. This
behavior is temperature independent, which means that the
lower viscosity of LDPE 1 at low strain rates is not due to
sagging at long measuring times and springs from its molecular structure. Together with the TREF results, it is concluded
that the large linear chain portion of LDPE 1 is responsible for
the decrease in uniaxial elongational viscosity predictions.
ATR-FTIR is used to calculate the average number of
branches per LDPE molecule. It is shown that the average
number of branches per molecule is higher for LDPE 2 than
for LDPE 1. In comparison to the radius of gyration of
LDPEs, it has been concluded that LDPE 2 has higher number
of branches per molecule at high molecular weight portions.
Therefore, having higher number of branches especially at
high molecular weight portions is responsible for higher
values of zero shear viscosity and uniaxial elongational viscosity at low strain rates for LDPE 2 than for LDPE 1.

References
Abbasi M, Ebrahimi N, Nadali M, Khabazian Esfahani M (2012)
Elongational viscosity of LDPE with various structures: employing
a new evolution equation in MSF theory. Rheol Acta 51:163177.
doi:10.1007/s00397-011-0572-z
Bryant WMD, Voter RC (1953) The molecular structure of polyethylene.
II. Determination of short chain branching1. J Am Chem Soc 75:
61136118. doi:10.1021/ja01120a006
Doi M, Edwards SF (1978a) Dynamics of concentrated polymer systems.
Part 1Brownian motion in the equilibrium state. J Chem Soc
Faraday Trans 2: Mol Chem Phys 74:17891801
Doi M, Edwards SF (1978b) Dynamics of concentrated polymer systems.
Part 2molecular motion under flow. J Chem Soc Faraday Tran 2:
Mol Chem Phys 74:18021817
Doi M, Edwards SF (1978c) Dynamics of concentrated polymer systems.
Part 3the constitutive equation. J Chem Soc Faraday Trans 2: Mol
Chem Phys 74:18181832
Doi M, Edwards SF (1987) The theory of polymer dynamics. Clarendon
Press/Oxford University Press, Oxford/New York
Gurp M, Palmen J (1998) Timetemperature superposition for polymeric
blends. Rheol Bull 67:58
Janzen J, Colby RH (1999) Diagnosing long-chain branching in polyethylenes. J Mol Struct 485486:569584
John D, Ronald L (2006) Structure and rheology of molten polymers
from structure to flow behavior and back again. Hanser Publishers/
Hanser Gardner Publications, Munich/Cincinnati
Krimm S, Liang CY, Sutherland GBBM (1956) Infrared spectra of high
polymers. II: Polyethylene. J Chem Phys 25:549
Kulin LI, Meijerink NJ, Starck P (1988) Long and short chain branching
frequency in low density polyethylene (LDPE). Pure Appl Chem 60:
14031415. doi:10.1351/pac198860091403
Lohse DJ, Milner ST, Fetters LJ, Xenidou M, Hadjichristidis N,
Mendelson RA, Garca-Franco CA, Lyon MK (2002) Well-defined,
model long chain branched polyethylene. 2. Melt rheological behavior. Macromolecules 35:30663075. doi:10.1021/ma0117559

Rheol Acta (2015) 54:159168


Moore JC (1964) Gel permeation chromatography. I. A new method for
molecular weight distribution of high polymers. J Polym Sci Part A:
Gen Pap 2:835843. doi:10.1002/pol.1964.100020220
Mnstedt H, Kurzbeck S, Stange J (2006) Importance of elongational
properties of polymer melts for film blowing and thermoforming.
Polym Eng Sci 46:11901195. doi:10.1002/pen.20588
Pattamaprom C, Driscroll JJ, Larson RG (2000) Nonlinear viscoelastic
predictions of uniaxialextensional viscosities of entangled polymers.
Macromol Symp 158:114
Podzimek S (1994) The use of GPC coupled with a multiangle laser light
scattering photometer for the characterization of polymers. On the
determination of molecular weight, size and branching. J Appl
Polym Sci 54:91103
Roedel MJ (1953) The molecular structure of polyethylene. I. Chain
branching in polyethylene during polymerization1. J Am Chem
Soc 75:61106112. doi:10.1021/ja01120a005
Roln-Garrido V, Wagner M (2007) The MSF model: relation of nonlinear parameters to molecular structure of long-chain branched polymer melts. Rheol Acta 46:583593
Roln-Garrido VH, Wagner MH (2014) Linear and non-linear rheological characterization of photo-oxidative degraded LDPE. Polym
Degrad Stab 99:136145. doi:10.1016/j.polymdegradstab.2013.11.
014
Rolon-Garrido VH, Zatloukal M, Wagner MH (2013) Increase of longchain branching by thermo-oxidative treatment of LDPE: chromatographic, spectroscopic, and rheological evidence. J Rheol 57:105
129
Rugg FM, Smith JJ, Wartman LH (1953) Infrared spectrophotometric
studies on polyethylene. I. Structure. J Polym Sci 11:120. doi:10.
1002/pol.1953.120110101
Schuttlefield JD, Grassian VH (2008) ATRFTIR spectroscopy in the
undergraduate chemistry laboratory. Part I: fundamentals and examples. J Chem Educ 85:279. doi:10.1021/ed085p279
Shroff RN, Mavridis H (1999) Long-chain-branching index for essentially linear polyethylenes. Macromolecules 32:84548464
Stadler F (2012) Detecting very low levels of long-chain branching in
metallocene-catalyzed polyethylenes. Rheol Acta 51:821840. doi:
10.1007/s00397-012-0642-x
Surez I, Coto B (2013) Determination of long chain branching in PE
samples by GPC-MALS and GPC-VIS: comparison and uncertainties. Eur Polym J 49:492498. doi:10.1016/j.eurpolymj.2012.
11.015
Tribe K, Saunders G, Meiner R (2006) Characterization of branched
polyolefins by high temperature GPC utilizing function specific
detectors. Macromol Symp 236:228234
Trinkle S, Walter P, Friedrich C (2002) Van GurpPalmen plot II
classification of long chain branched polymers by their topology.
Rheol Acta 41:103113
Wagner MH (2006) The rheology of linear and long-chain branched
polymer melts. Macromol Symp 236:219227
Wagner MH, Bastian H, Ehrecke P, Kraft M, Hachmann P, Meissner J
(1998) Nonlinear viscoelastic characterization of a linear polyethylene (HDPE) melt in rotational and irrotational flows. J NonNewtonian Fluid Mech 79:283296
Wagner MH, Yamaguchi M, Takahashi M (2003) Quantitative
assessment of strain hardening of low-density polyethylene
melts by the molecular stress function model. J Rheol 47:
779793
Wood-Adams PM, Dealy JM (2000) Using rheological data to determine
the branching level in metallocene polyethylenes. Macromolecules
33:74817488
Yau WW, Gillespie D (2001) New approaches using MW-sensitive
detectors in GPC TREF for polyolefin characterization. Polymer
42:89478958

Vous aimerez peut-être aussi