Vous êtes sur la page 1sur 17

Rheol Acta (2014) 53:519–535

DOI 10.1007/s00397-014-0775-1

ORIGINAL CONTRIBUTION

Quiescent and shear-induced crystallization


of polyprophylenes
Maziar Derakhshandeh · Antonios K. Doufas ·
Savvas G. Hatzikiriakos

Received: 5 December 2013 / Revised: 8 March 2014 / Accepted: 7 April 2014 / Published online: 31 May 2014
© Springer-Verlag Berlin Heidelberg 2014

Abstract In this paper, the effect of shear on the flow- to convert raw materials (mainly polyolefins) to final plastic
induced crystallization (FIC) of several polypropylenes of products. The raw materials used in these polymer pro-
various macrostructures was studied using rheometry com- cessing methods are often subjected to mixed shear and
bined with polarized microscopy. Generally, an increase in extensional flow conditions, which cause the polymer to
strain and strain rate or decrease of temperature is found experience flow-induced crystallization (FIC). The mechan-
to decrease the thermodynamic barrier for crystal formation ical and optical properties of the final products are strongly
and thus enhancing crystallization kinetics at temperatures dependent upon the processing method and the conditions
between the melting and crystallization points. Secondly, under which crystallization takes place (Eder et al. 1990;
popular models based on suspension theory which are used McHugh and Edwards 1995; Kumaraswamy et al. 1999;
to relate the degree of crystallinity to normalized rheological Kornfield et al. 2002; Janeschitz-Kriegl 2003; Scelsi and
functions (such as viscosity) are validated experimentally. Mackley 2008).
For this purpose, the space filling of crystals in the polarized The kinetics of crystallization considerably improves
micrographs determined from image processing was plotted under flow at temperatures comparable to the melting peak
as a function of normalized viscosity under various shear temperature and is related to both strain and strain rates
rates. It is found that the constant(s) of various suspension (Eder et al. 1998; Gelfer and Winter 1999; Kitoko et al.
models should be dependent on the flow parameters in order 2003; Swartjes et al. 2003; Stadlbauer et al. 2004; Tanner
for the suspension models to describe the effect of shear on and Qi 2005, 2009; Dai et al. 2006; Hadinata et al. 2006;
FIC, particularly at higher shear rates. Chellamuthu et al. 2011; Tiang and Dealy 2012; White et al.
2012). Moreover, different types of flow deformation affect
Keywords Flow-induced crystallization · Polypropylene · FIC differently. For instance, in the case of fiber spinning
Quiescent crystallization · Halftime crystallization and film blowing, extensional flows account for the orienta-
tion and stretching of polymer molecules in the direction of
extension, thus enhancing the crystallization process signif-
Introduction icantly. Online fiber spinning data related to flow-induced
crystallization behavior of several polypropylenes (PPs)
Processes such as film blowing, film casting, fiber spinning, have been reported (Paradkar et al. 2008; Patel et al. 2008).
calendaring, and injection molding are still used extensively Shearing also enhances the crystallization kinetics; how-
ever, since it is inherently a weaker flow compared to exten-
sional flow, larger strains are required to see the effect of
M. Derakhshandeh · S. G. Hatzikiriakos () flow on crystallization (Hadinata et al. 2007; Derakhshan-
Department of Chemical and Biological Engineering, deh and Hatzikiriakos 2012). High shear rate data obtained
The University of British Columbia, Vancouver, BC, Canada using a special sliding plate rheometer was reported for
e-mail: savvas.hatzi@ubc.ca
poly-1-butene (Baert and Van Puyvelde 2006). At low shear
A. K. Doufas rates, the crystallization kinetics was enhanced mainly by
ExxonMobil Chemical Company, Baytown, TX, USA inducing a higher nucleation density. Further increasing in
520 Rheol Acta (2014) 53:519–535

the shear rate imposed oriented structures (Baert and Van polarized microscopy with some of the effects studied here
Puyvelde 2006). have also been addressed previously by others (Coppola
There are relatively few available data in the literature et al. 2004; Hadinata et al. 2005; Jay et al. 1999; Baert et al.
on the importance of different types of flows on crystalliza- 2006; Acierno et al. 2003; Duplay et al. 2000; Bourgin and
tion and even less data that address the effects of molecular Zinet 2010; Kim et al. 2005). However, in this paper, six
weight parameters on crystallization (Chellamuthu et al. different PP resins exhibiting different behavior in terms of
2011; Eder et al. 1998; Tanner and Qi 2005; Baert and Van nucleation (as observed by polarized optical microscopy)
Puyvelde 2006; Tiang and Dealy 2012; White et al. 2012). and crystal growth rate are studied. The data provided
Therefore, it is an objective of this paper to examine in this paper can be the basis for obtaining the model
the role of temperature and shear flow on the crystalliza- parameters which can be further validated by comparison
tion kinetics of several polypropylenes of various molecular with FIC behavior under extensional and/or mixed flows.
weights and polydispersities. In particular, the effects of A comprehensive model should predict crystallization
important parameters for crystallization such as strain rate, behavior under quiescent, shear, extensional, and mixed
strain, temperature, and molecular parameters are examined flow conditions at various temperatures. Moreover, as
under simple shear using rheometry coupled with polarized mentioned by Duplay et al. (2000), extra care should be paid
microscopy. when resins produced by different technologies are studied
An effective method of optimizing any industrial-scale and compared.
process and/or improving the quality of the final prod-
uct is by performing numerical simulations of the process.
To put this into perspective, in the film blowing process, Material and methods
the locked-in stresses at the frost line (location of transi-
tion from liquid to solid) dictates the tensile strength of Six different PPs with different molecular weights and poly-
the produced material (Han and Kwack 1983; Kwack and dispersities were used in this work, all provided by the
Han 1983; Kuijk et al. 1999). Numerical simulation of this ExxonMobil Chemical Company. These resins are listed
process and its further optimization heavily depend on the in Table 1 along with their molecular weight characteris-
accuracy of the models, and for the process under discus- tics (molecular weight, Mw ; polydispersity index (PDI =
sion, a reliable model for FIC is crucial (Doufas et al. Mw /Mn ), melt flow rate, mass flow rate (MFR), and tac-
1999, 2000; Doufas and McHugh 2001; Doufas 2013). In ticity). The molecular weights (weight average molecular
such models, the degree of crystallinity is often related to weight Mw and number average molecular weight Mn ) were
normalized rheology functions (NRF) by various models determined using High-Temperature Gel-Permeation Chro-
derived from the suspension theory. These models bene- matography (Polymer Laboratories PL-GPC-220) equipped
fit from being easy to implement into simulation; however, with a differential refractive index detector (DRI) (Sun et al.
they have their own limitations such as not being applicable 2001). Carbon NMR (13 C) spectroscopy was used to mea-
to higher deformation rates. Therefore, the second objec- sure the tacticity of the studied polypropylenes in terms of
tive of this paper is to evaluate suspension models which percent molar meso pentads (mmmm). 13 C NMR spectra
are typically used to relate the system zero-shear-rate vis- were acquired with a 10-mm broadband probe on a Varian
cosity with the degree of crystallization. A light microscope spectrometer having a 13 C frequency of at least 100 MHz.
equipped with a polarizer and an analyzer is used with a The samples were prepared in 1,1,2,2-tetrachloroethane-d2
parallel-plate rheometer in order to determine the degree of (TCE). Sample preparation (polymer dissolution) was per-
crystallization at various deformation rates. These data sub- formed at 140 ◦ C where 0.25 g of polymer was dissolved
sequently are used to evaluate the applicability of various in an appropriate amount of solvent to give a final polymer
suspension models in describing the degree of crystallinity solution of 3 ml.
under flow. Previous methods related NRF with relative
crystallinity obtained from independent differential scan- Differential scanning calorimetry
ning calorimetry (DSC) measurements (Acierno et al. 2003;
Boutahar et al. 1998; Lamberti et al. 2007; Pantani et al. A Shimadzu DSC-60 calorimeter was used to study the
2001; Titomanlio et al. 1997). In the present paper, the rhe- thermal behavior of PPs. Nitrogen purge was used in order
ological functions and degree of crystallization are obtained to avoid degradation of the test specimens. Melting and
simultaneously, thus avoiding possible errors from a differ- crystallization peaks of polymers were obtained by using
ent thermal history experienced by the test specimens as multiple heating and cooling protocols applied to 1–2 mg
shown below. of sample directly harvested from pellets sealed in an alu-
In summary, shear effects on the FIC of a variety of minum pan. The exact experimental protocol is shown in
PPs is studied here using shear rheometry combined with Fig. 1. Thermal and flow histories were eliminated by
Rheol Acta (2014) 53:519–535 521

Table 1 The PP resins studied


along their molecular PP resin Mw (kg/mol) PDI MFR (dg/min) % molar meso pentads (mmmm) Tc (◦ C) Tm (◦ C)
characteristics and thermal
properties PP1 182 2.7 36.0 0.950 122.0 162.8
PP2 174 1.9 24.0 0.947 108.7 148.6
PP3 230 3.1 12.0 0.954 115.0 160.5
PP4 288 3.8 5.0 0.944 124.0 164.2
PP5 366 3.3 2.0 0.956 112.0 162.1
PP6 498 3.6 0.9 0.909 113.4 161.0

heating each sample at the rate of 10 ◦ C/min from 50 to halftimes of crystallization of all resins at specific values of
200 ◦ C and soaking it at 200 ◦ C for 15 min. Eventually, T = 0.25, 0.5, and 0.75 on the same graph.
the temperature was decreased to 50 ◦ C at 10 ◦ C/min which
was followed by reheating to 200 ◦ C using the same rate in Shear rheometry
order to obtain the crystallization and melting temperature
peaks, respectively. History elimination was confirmed by An Anton Paar rotational rheometer (MCR 502) was used
obtaining a third cycle, where the crystallization and melting to study the shear rheology of samples. Polypropylene pel-
peaks were found almost identical to those obtained in the lets were melted at the temperature of 200 ◦ C and then were
second cycle. The melting and crystallization peaks Tm and pressed in a compression molding apparatus to produce
Tc , respectively, are summarized in Table 1, and the values sheets with a thickness of about 1 mm. The shear-induced
represent the averages of three different samples. crystallization behavior of polymers was examined in sim-
The halftime of isothermal crystallization was deter- ple shear flow using the parallel-plate geometry with plates
mined at different temperatures between the determined of 25 mm in diameter. These experiments are confined to
crystallization, Tc , and melting, Tm , DSC peaks. These were the temperature region between the crystallization and the
defined in DSC thermograms as the time required for a sam- melting peaks found in the DSC thermograms under the
ple to reach 50 % of its final crystallinity at the designated heating and cooling modes. This temperature range is of
crystallization temperature. The percent of crystallinity as great importance in FIC due to the tendency of molecules to
a function of time was calculated by integrating the area crystallize and reorient. Since higher temperatures impede
underneath the peak of the DSC thermogram. the crystallization kinetics excessively, FIC at higher tem-
Since the investigated PP resins exhibit different melting peratures is extremely time-consuming. The crystallization
and crystallization peaks, the isothermal crystallization for behavior of resins at shear rates larger than 2 s−1 could
the various resins was studied at fixed values  of a normal- not be investigated considering edge failure effects on the
ized temperature defined by T = (T − TC ) (Tm − TC ). measured torque. Thermal and flow histories of the sam-
This enables also a consistent and fair comparison of the ples were eliminated by heating up each test specimen to
200 ◦ C (around 40 ◦ C above the melting peak tempera-
ture) for 15 min prior to cooling at the rate of 10 ◦ C/min
DSC protocol to the desired temperature and experimental testing. Fast
200 History
and accurate cooling is crucial to reach the desired crystal-
Elimination
on
Cooling

lization temperature without any temperature undercooling


terminati

Validation
that would lead into premature crystallization. The Peltier
Elimination
Temperature, T (oC)

Peak De

150 system of the rheometer was used for this study since it
Peak De

can employ higher cooling rates more precisely compared


to those of the convection oven system. The influence of
terminati

several parameters such as temperature, deformation, and


Heating

100
History

deformation rates (shear) on the crystallization kinetics of


on

the PPs were studied at shear rates up to 2 s−1 .


In parallel-plate geometry, the shear rate varies linearly
50
from the center to the edge of the plate. In this paper, the
shear rate is reported at the radial position 0.75R since the
0 20 40 60 80 100 120 140 160
difference between the true and measured viscosity assigned
Time, t (s) to the shear rate at r = 0.75R is less than 2 % as reported
Fig. 1 The DSC protocol which is used to determine the melting and by Macosko (1994). Crystallization starts at the outer edge
crystallization peaks (Tc and Tm ), respectively of the sample due to a higher shear deformation rate at this
522 Rheol Acta (2014) 53:519–535

location. For modeling purposes, the induction times are in any image processing techniques. Therefore, the errors
usually defined at small levels of crystallinity (∼1–2 %), contributing to this technique is minimized by obtaining
and therefore the crystallization phenomenon which starts clear images at a gap size of about 0.5 mm. Since the rela-
from the outer edge has a minimal effect on the rheological tive crystallinity and the NRF are obtained using the same
behavior captured at the induction time. When the crystal- experiment, the temperature and shear deformation histories
lization further increases, the polymer shrinks, and thus the which are imposed on the samples are identical, which is an
underlying equations used for the rheological characteriza- advantage of this experimental method.
tion are not applicable without modifications regardless of
the geometry used (Godara et al. 2006).
Results and discussion
Polarized light microscopy
Linear viscoelasticity of the polypropylenes
Polarized light microscopy (Mitutoyo microscope setup
equipped with a Lumenera LU 165 color CCD camera and The master curves of the linear viscoelastic moduli of all
two polarizers) was used in conjunction with an Anton Paar PPs studied at the reference temperature of 190 ◦ C produced
MCR502 in order to study the crystalline structure evolution by the application of the time-temperature superposition
formed under different conditions. In a typical experiment, (TTS) technique is depicted in Fig. 2a–f (Ferry 1980). For
the sample (disk) is placed in the space between two glass the low molecular weight materials, the Newtonian region
plates of the rheometer (parallel-plate geometry of 43 mm (e.g., PP1, PP2) at low frequencies was reached. For the
in diameter) and the microstructure is observed with the high molecular weight materials (PP4, PP5, and PP6), stress
microscope as a function of time. The space filling of crys- relaxation experiments after imposing a sudden strain were
tals in the polarized images was calculated through image also performed to obtain data over a wide range of frequen-
analysis as a function of time and then converted to the cies. This is possible by converting the relaxation modulus
degree of crystallinity by using the densities of crystal and to dynamic data at very low frequencies in order to deter-
polymer melt (0.94 and 0.85 g/cm3 ), respectively (Natta mine the zero-shear viscosity (Schwarzl method) (Ferry
et al. 1955). The calculated crystallinity under quiescent 1980). This procedure was used just for three of the PPs
conditions at different temperatures was compared to crys- in this study, i.e., the high molecular weight resins. The
tallinity obtained using DSC so as to validate the polarized continuous lines in Fig. 2a–f represent fits of the multi-
microscopy applicability. mode Maxwell model {Gi, λi}which show that this model is
As discussed above, the microscopy videos were decom- capable of representing the data very well.
piled to frames which were analyzed by an image process- Figure 3 plots the complex viscosity of all PPs as a func-
ing software (Matlab and ImageJ) in order to calculate the tion of frequency at the reference temperature of 190 ◦ C.
amount of space filling. Two different image processing This figure is in agreement with the MFR values of the
methods were utilized to validate the obtained results. In the resins listed in Table 1. The resin with the larger viscosity
first approach, the grayscale filter was first applied to all has higher resistance to flow, and thus, it has a lower MFR.
extracted images to set each pixel value to a single number Table 2 lists the activation energy for flow, Ea , using the
that represents the brightness of the pixel. The pixel format shift factors obtained from the construction of the master
is typically the byte image which gives a range of possi- curves depicted in Figs. 2 and 3. The Arrhenius model was
ble values from 0 to 255. Frequently, zero is taken to be used:
black, while 255 are taken as white. Values in between 0 and   
Ea 1 1
255 construct different shades of gray. Each pixel value was aT = exp − − (1)
R T Tref
normalized between 0 and 1 in Matlab in order to define
a threshold. All pixels within the microscopy images were where Ea is the activation energy for flow, a measure of the
converted into black or white based on the specified thresh- sensitivity of the viscoelastic properties to temperature. It
old, which was estimated by Matlab and further tuned by the can be seen that PP5 has the highest energy of activation,
user. Eventually, the space filling was estimated as the ratio Ea .
of the number of white pixels to the total number of pixels
within the image. In the second approach, the area of crys- Thermal behavior of the polymer
tals was estimated by the area of circles with the same center
and radios as that of the crystals. In this method, any overlap Figure 4a depicts a typical thermogram of one of the PPs
between the circles is considered once. Circles were drawn (PP3 in Table 1). The melting and the crystallization peak
manually and then processed by ImageJ software. Image temperatures were observed at 160.5 and 115 ◦ C, respec-
quality plays a significant role in acquiring accurate results tively. As shown in Fig. 4b, the thermomechanical history
Rheol Acta (2014) 53:519–535 523

a b

d
c

e f

Fig. 2 a–f Master curves of the viscoelastic moduli of all PPs listed in Table 1 at the reference temperature of Tref = 190 ◦ C

causes a difference of 2.5 ◦ C between the first heating peak normalized temperature, T for all PP resins. This normal-
(163.0 ◦ C) and the melting peak temperature (160.5 ◦ C). ization of the temperature facilitates the comparison of
The difference between the second and third heating peaks halftimes of crystallization for various resins as discussed
is only 0.2 ◦ C, which is within experimental error. This above. At very low normalized temperatures (i.e., close to
further indicates that the sample soaking time of 15 min crystallization peak), the halftime of crystallization for all
at 200 ◦ C is long enough to eliminate the thermal his- polypropylenes are very small as normally expected. In this
tory (Fig. 4b). Similar comments apply to the crystalliza- region (T < 0.2), the halftimes of crystallization of differ-
tion/cooling peaks as well as to the DSC results for all other ent resins show a small difference of 5 %, which is within
samples listed in Table 1. experimental error. The crystallization kinetics varies signif-
Figure 5 plots the quiescent halftimes of crystalliza- icantly between various resins as the temperature increases.
tion obtained from DSC thermograms as a function of the The difference is potentially attributed to differences in (i)
524 Rheol Acta (2014) 53:519–535

it exhibits slightly increased crystallization kinetics due to


its chain mobility (smaller Mw ).

Flow-induced crystallization under shear flow

Figure 7a, b demonstrates the effect of cooling rate from


T = 200 ◦ C to the test temperature of 142.4 ◦ Cor
T = 0.5 on the shear stress growth coefficient η+ = τ γ̇
as a function of time for PP1 at two different shear rates.
These tests were performed to identify the optimum cool-
ing rate that would best satisfy the conditions of isothermal
flow-induced crystallization. Cooling rates smaller than a
value of about 5 ◦ C/min allow the polymer to crystallize
Fig. 3 The master curves of the complex viscosity of all PPs listed in non-isothermally as it takes significant time for the quench-
Table 1 at the reference temperature of Tref = 190 ◦ C
ing. For example, this is reflected in an earlier upturn of
the shear stress growth coefficient when the cooling rate
tacticity (see Table 1), (ii) the catalyst system, and (iii) of 1 ◦ C/min is used at both shear rates (Fig. 7a, b). This
the nucleation density, e.g., affected by additives such as earlier upturns are clearly due to premature crystallization
nucleating agents. before reaching the test temperature. Cooling rates greater
Figure 6 depicts polarized optical microscopy (POM) than about 10 ◦ C/min allow nearly isothermal crystalliza-
images obtained for various resins studied at the normalized tion by minimizing premature crystallization. The results
temperature of T = 0.5 under quiescent conditions. As can become independent of cooling rate for values greater than
be seen from Fig. 6, PP3 and PP5 exhibit similar nucleation 10 ◦ C/min.
density (the lowest), and therefore they revealed compara- To show this effect more clearly, the time required to
ble kinetics under quiescent conditions (Fig. 5). PP4 has the observe the onset of crystallization (induction time, tind ) is
highest nucleation density and thus has shown the lowest defined as the time at which η+ becomes 20 % (arbitrary
induction times in the temperature range investigated. PP1 value) greater than its steady-state value (Derakhshandeh
has a higher nucleation density compared to PP2, and thus and Hatzikiriakos 2012). The induction times as functions
it has shown enhanced crystallization kinetics. Since PP6 of cooling rates are plotted in Fig. 8 for two different shear
has a higher nucleation density compared to PP2, it should rates for PP1 at T = 0.5. High cooling rates (greater than
possess a lower growth rate in order to reveal the same 20 ◦ C/min) may also impose undercooling (undershooting
overall crystallization kinetics as that of PP2. All resins in the temperature profile due to limitations in the tem-
except PP2 were produced using a similar catalyst system, perature control of the rheometer in fast transient tests).
and thus, their growth rate could be compared. It has been Therefore, in this study, the cooling rate used is 10 ◦ C/min
reported that the Mw affects the crystal growth rate under in order to minimize the effect of undercooling and, at the
quiescent conditions. Specifically, the growth rate of crys- same time, to eliminate the effect of premature crystalliza-
tals decreases with increase of Mw due to less chain mobility tion due to long transients, except at temperatures close to
of the system (Wunderlich 1976). Therefore, since PP6 pos- the crystallization peak as discussed below.
sesses the highest Mw among all PPs studied, it should also The effect of temperature on flow-induced crystalliza-
possess the slowest growth rate. Since PP5 and PP3 have tion for PP1 is demonstrated in Fig. 9. Initially, imposition
similar nucleation density and since PP3 has a smaller Mw , of shear flow causes the shear stress growth coefficient to
increase with time, reaching its steady-state value after a
Table 2 The PP resins studied along with their activation energy for
certain time. The dashed horizontal lines in the graph show
flow, Ea the steady-state viscosity of crystal-free melt which can be
used as a basis to determine the induction time. The nor-
PP Resin Ea (kcal/mol) malized temperatures of 0.5 and 0.75, in combination with
PP1 9.55 the cooling rate (10 ◦ C/min), are high enough to minimize
PP2 7.63
premature crystallization during the cooling period before
PP3 9.63
shearing as well as during the initial ascending part of the
PP4 10.56
shear stress transient. However, crystallization occurs after a
certain time that causes a sudden increase of the shear stress
PP5 11.27
growth coefficient. At the lower normalized temperature
PP6 9.91
of 0.25, the shear stress growth coefficient, η+ , increases
Rheol Acta (2014) 53:519–535 525

Fig. 4 a The DSC thermogram 10


0
of resin PP3. The second PP3 TC=115oC
PP3
melting and crystallization DSC DSC
peaks were observed at 160.5
and 115 ◦ C, respectively. b The
5

Heat, Q (mW)
DSC thermogram of PP3 for

Heat, Q (mW)
three consecutive
heating/cooling cycles. Thermal 0
history elimination is
demonstrated by the small
differences obtained between First heating
-5 Tm=160.5oC Second and third peak
second and third melting peaks
heating peak
a b
-5
100 200 150 155 160 165 170 175
o o
Temperature,T ( C) Temperature,T ( C)

before reaching its steady state. The times for the onset significantly. In the parallel-plate geometry, the shear defor-
of crystallization and transient to reach its steady state are mation rate varies linearly with radial position, being zero
comparable as shown in the graph. In addition, nucleation at the center and attains its highest value at the edge. The
that evolved during the cooling period (due to the low tem- effect of this nonhomogeneous shearing on the nucleation
perature) also contributes to this deviation. As seen from density of PP1 is shown in Fig. 10. The nucleation density
Fig. 9, the temperature has a noticeable effect on the induc- increases significantly from the center (Fig. 10a) to the edge
tion time. A decrease of roughly 10 ◦ C (for example, from (Fig. 10c). Moreover, shear also induces crystal alignment
T = 0.5 to 0.25) causes the onset of crystallization to with the highest alignment and nuclei density close to the
occur at times of roughly 1 order of magnitude earlier. In edge of the parallel-plate geometry.
general, crystallization kinetics is significantly enhanced as Figure 11 depicts the shear stress growth coefficient as
temperature decreases. a function of the time for various imposed shear rates at
The gradual increase of viscosity as the crystallization T = 0.5 for PP1. A larger shear rate stretches and ori-
begins can be used to determine the induction time which is ents molecules more effectively after a certain strain. This
arbitrarily defined as the time at which a 20 % of increase of activates more nuclei which is also supported by more
the viscosity is obtained. These induction times correspond spherulites of smaller size in the microscopy images (Huo
to a small degree of crystallinity (typically 1–2 %) and if the et al. 2004). Therefore, a larger deformation rate decreases
true viscosity is measured in parallel plate (i.e., γ̇ reported the induction time for the onset of crystallization by enhanc-
at r = 0.75 R), the increase in torque caused by crystal- ing the nucleation density (Fig. 10) and possibly crystal
lization does not influence the induction time measurements growth rates as mentioned in previous studies (Duplay et al.
2000).
Figure 12 plots the shear stress growth coefficient data
10000
of Fig. 11, this time as function of strain. Typically,
PP1
the required shear strain for the onset of crystallization
PP2 increases with increase of the applied deformation rate. In
8000 PP3 other words, more strain is needed at higher shear rates to
Crystallization half time, thalf (s)

PP4
PP5 initiate crystallization. In particular, at the shear rates of
6000
PP6 0.01 s−1 , a strain of 70 is required to initiate crystallization,
while at the shear rate of 0.1 s−1 , crystallization occurs after
600 strain units. This is in agreement with experimental
4000
results reported by other authors (Eder et al. 1998; D’Haese
et al. 2010). The increase of the amount of strain with the
2000 increase of shear rate for the onset of crystallization (see
Fig. 12) is a counterintuitive observation.
Figure 13 shows the effect of shear rate on the induc-
0
tion time for all PPs studied. Open symbols in Fig. 13 are
0.0 0.1 0.2 0.3 0.4 0.5 0.6
obtained from rheometry as described above, while closed
Normalized temperature, T symbols are obtained by DSC. As seen, flow enhances and
Fig. 5 The quiescent (DSC) halftimes of crystallization of all PPs as speeds up the onset of crystallization mainly at rates greater
a function of the normalized temperature, T than a certain critical value. At deformation rates smaller
526 Rheol Acta (2014) 53:519–535

Fig. 6 POM images of


investigated resins at T = 0.5.
As shown, PP4 has the highest
nucleation density among all.
PP3 and PP5 exhibit comparable
nuclei density at the studied
temperature region

than this value, the flow does not disturb the molecules region (γ̇ < γ̇critical ), the crystallization behavior is essen-
from their equilibrium configurations, and thus no effect tially similar to that of quiescent condition. Larger shear
on their crystallization behavior occurs. Therefore, over this rates (γ̇ > γ̇critical ) activate more nuclei causing a dramatic

Fig. 7 The effect of cooling


Shear stress growth coefficient, + (kPa.s)
Shear stress growth coefficient, + (kPa.s)

30 30
rate on the shear stress growth a b
coefficient η+ of PP1 at the
shear rates of a 0.005 s−1 and b
0.1 s−1 at T = 142.4 ◦ C or 10 10
T = 0.5
PP1
.
PP1 = 0.1000 s-1
. T=0.5
= 0.005 s-1
T=0.5 1oC/min
o
1oC/min 5 C/min
5oC/min 10oC/min
10oC/min 20oC/min
20oC/min 30oC/min
1 1
102 103 104 102 103 104
Time, t (s) Time, t (s)
Rheol Acta (2014) 53:519–535 527

PP2; however, increasing shear rate enhances the crystal-


lization kinetics of PP6 more than that of PP2. PP2 does not
have as long-chain tail as PP6 due to its narrower molecu-
lar weight distribution (MWD) (Table 1) and thus requires
higher shear rate before significant reduction of the induc-
tion time occurs. PP6 has higher Mw and broader MWD
than that of PP2 which likely cause the induction time to
drop at a lower shear rate.
It is noted from Fig. 13 that the halftimes of crystalliza-
tion do not agree quantitatively with the induction times
from simple shear experiments at very small shear rates. The
former are greater than the latter for all resins. The half-
times of crystallization correspond to the times where 50 %
of relative crystallinity is formed. Polypropylenes studied
here have crystallinity of ∼46 %, and therefore 50 % of rel-
ative crystallinity corresponds to roughly 23 %. On the other
hand, the induction time determined from the flow experi-
Fig. 8 The effect of cooling rate on the induction time, tind , for the
onset of crystallization in startup of steady shear experiments for the ments is defined as the time at which the viscosity increases
resin PP1 at T = 142.4 ◦ C or T = 0.5 for two different shear rates by 20 % above its steady-state value for samples free
of crystallinity. Therefore, these two times are different,
reduction of the induction time. Similar observations were and therefore, the halftimes of crystallization are expected
reported for other polymers (Coppola et al. 2004; Paradkar to be somehow higher than those obtained from shear
et al. 2008). Differences in the molecular parameters of the testing at very small shear rates where the effect of flow
investigated resins such as molecular weight, molecular is minimal. To see this more clearly, we consider the sus-
weight distribution, and tacticity likely contribute to differ- pension model η/ηa = (1 − a/A)−2 , where A is a shape
ences in the induction times observed in Fig. 13. Among the constant, ηa is amorphous viscosity, and α is relative crys-
resins, PP5 and PP3 have the slowest crystallization kinetics tallinity. Setting the ratio η/ηa = 1.2 (induction time)
at small shear rates. However, flow enhances crystalliza- with A = 0.6 (frequently assumed), the crystallinity X is
tion kinetics of PP5 and causes dramatic reductions of its calculated to be about 0.024 (2.4 %), which is the time
induction time so that flow-induced crystallization is faster determined from shear experiments. This time is expected
compared to other resins. A similar induction time under a to be lower than the halftime of crystallization (α =
very low deformation rate was observed between PP6 and 0.50) which corresponds to roughly 23 % of crystallinity.
For the two times to be comparable, the induction time
for crystallization in shear should have been defined as the
time at which the ratio, η/ηa , becomes equal to η/ηa =
(1 − 0.5/0.6)−2 = 36, that is when the viscosity increases
to 36 times its steady-state value, instead of 1.2 times. In
spite of these qualitative differences, these two times agree
qualitatively.
In Fig. 13, two different regimes of crystallization can be
defined. At very low deformation rates, the crystallization
kinetics essentially resembles that of quiescent condition
where the shear rate does not change the induction time
(Coppola et al. 2004). The increase of deformation rate
causes chain orientation, lowers the thermodynamic barrier
for crystallization, and thus enhances the kinetics of crys-
tallization. In this region, shear rate significantly decreases
the induction time as mentioned by other authors (Coppola
et al. 2004). These two regions are more clearly defined for
PP2 and PP6 in Fig. 14 by means of two lines, namely a
horizontal where the induction time in independent of the
Fig. 9 The shear stress growth coefficient η+ of PP1 at the normalized rate of shear and a line having a negative slope showing
temperatures, at T of 0.25, 0.5, and 0.75 at the shear rate of 0.005 s−1 the decrease of induction time with increase of shear rate.
528 Rheol Acta (2014) 53:519–535

Fig. 10 Crystal density observed at various radial positions in the parallel-plate geometry for PP1 at T = 0.5 and shear rate of 2 s−1 a at center
of the geometry, b at r = 0.75R, and c r = R, with crystal alignment to increase with the radial position

The intersection of these two lines defines the critical shear are formed (Janeschitz-Kriegl et al. 2003). These behav-
rate required to start observing the effect of flow. These iors were also supported by various works done previously
critical shear rates for all resins studied are summarized in (D’Haese et al. 2010, 2011). PP4 has the highest nucleation
Table 3. As mentioned by Acierno and Grizzuti (2003), the density among all of the investigated resins (i.e., closer to
orientation of chains under shearing (low Weissenberg num- the saturated state), and thus a higher critical shear rate is
ber) is sufficient to enhance the crystallization phenomenon needed to enhance the quiescent crystallization kinetics as
by inducing a higher nucleation rate compared to quies- seen in Table 3. As previously mentioned, applying flow
cent conditions (Godara et al. 2006; Devaux et al. 2004; usually causes the polymer to induce a higher nuclei density.
Koscher and Fulchiron 2002). Large deformation rates (i.e., PP2 uses a different catalyst for its production and thus can-
large Weissenberg numbers) stretch chains in the direction not be compared with the other resins. The different catalyst
of flow, changing the mechanism of crystallization from system for production can introduce a different growth rate
spherulitic to rodlike nucleation which further facilitates the and kinetics since it imposes different molecular structures
kinetics of crystallization. POM images in our case sug- (Wunderlich 1976). PP5 and PP3 have very similar tacticity
gest spherulitic crystallization at low Weissenberg numbers and PDI. PP5 has a higher molecular weight in compari-
(typically of the order of 1). son with PP3 and thus shows lower critical shear rate for
It was hypothesized earlier that dormant nuclei exist flow-enhanced crystallization. Although PP6 has the high-
within polymer matrix. These precursors are activated by est molecular weight and PDI, it exhibits a higher critical
imposing undercooling and/or applying deformation rates shear rate than PP5, PP3, and PP1 due to its lower tacticity.
(Janeschitz-Kriegl and Ratajski 2005; Janeschitz-Kriegl Tacticity seems to play a major role in FIC.
et al. 2003). Higher deformation rates and higher degree of Figure 15 depicts the effect of shear rate on the
undercooling activate more dormant nuclei. It was assumed induction time at various normalized temperatures for PP1.
and supported by experimental data that at very high defor- DSC results for T = 0.75 could not be obtained since
mation, the nucleus density saturates, and no more nuclei the heat of crystallization which is released at this high
temperature is not detectable by the DSC apparatus used.
The induction time increases by 1 order of magnitude for a
change of roughly of 10 ◦ C. On the same time, an increase
of shear rate from 0 to 0.5 decreases the induction time by a
factor of 2.

Crystallization kinetics using optical microscopy

In addition to rheometry, polarized light microscopy was


used in conjunction with the Anton Paar MCR-502 parallel-
plate rheometer in order to study crystallization kinetics
of PPs under shear. The microscope is equipped with a
polarizer and an analyzer which are always crossed dur-
ing experimental testing. Therefore, only crystals which are
present within the polymer matrix can be observed due to
their birefringence (Wang and Dou 2008). Similar heat-
ing and cooling protocols as described in the experimental
Fig. 11 The shear stress growth coefficient η+ of PP1 at shear rates section were used so as to eliminate flow and thermal
ranging from 0.005 to 0.1 s−1 at 152.6 ◦ C or T = 0.75 history.
Rheol Acta (2014) 53:519–535 529

Fig. 12 The shear stress growth


coefficient η+ of PP1 at shear
rates from 0.005 to 0.1 s−1 at
152.6 ◦ C or T = 0.75

Quiescent crystallization of resins at desired tempera- the density of crystals (0.94 g/cm3 ) and that of the melt
tures was investigated by both optical microscopy and DSC (0.85 g/cm3 ) (Natta et al. 1955). The relative crystallinity
in order to experimentally validate the accuracy of this obtained from the optical microscopy was compared to that
setup. As a first step, the relative crystallinity of PP as obtained from DSC to check for accuracy of the optical
a function of time was calculated by integrating the area microscopic technique.
underneath the peak of the DSC thermogram and then nor- Table 4 shows typical microscopy images before and
malizing it by the total area. In a second step, the degree after image processing with methods 1 and 2 described in
of space filling of crystals in the polarized images was the experimental section. A satisfactory agreement between
calculated from optical microscopy by using the image pro- the DSC and image processing results was found, as can
cessing techniques described above. The degree of space
filling was converted to degree of crystallization using

Fig. 14 The effect of shear rate on the induction time for the onset
Fig. 13 The effect of shear rate on the induction time for the onset of of crystallinity for PP6 and PP2 at T = 0.5. A critical shear
crystallinity for all PPs at T = 0.5. The closed symbols correspond to rate is observed at which transition from quiescent to flow-induced
the halftime quiescent crystallization obtained from DSC crystallization occurs
530 Rheol Acta (2014) 53:519–535

Table 3 The critical shear rate above which the effect of flow on is essential in order to capture the true crystallization behav-
crystallization becomes evident at T = 0.5 ior of resins. Previous works typically measured the relative
Resin Critical shear rate (s−1 ) crystallinity from DSC and then relate it with rheological
measurements (Acierno et al. 2003; Boutahar et al. 1998;
PP1 0.02 Lamberti et al. 2007; Pantani et al. 2001; Titomanlio et al.
PP2 0.31 1997). As depicted in Fig. 17, the data at very low shear rate
PP3 0.03 (γ̇ < 0.5 s −1 ) shows comparable crystallization behavior.
PP4 0.10 Increasing the shear rates alters the crystal-melt interaction
PP5 0.02 and thus causes the normalized viscosity to show upturns at
PP6 0.06 lower degrees of space fillings (crystallizations).

Viscosity-crystallinity suspension modeling


be seen from the comparisons in Fig. 16a. The differences
are small and can be attributed to experimental errors origi- Different models based on the suspension theory were used
nating from a slightly different thermal history in DSC and to model the crystallization behavior of the PPs. These mod-
rheometer as well from errors in the resolution of the image els were studied before using the data generated from the
processing technique. Samples used in DSC are relatively coupling of the degree of crystallinity obtained by DSC and
small (1–2 mg) sealed in an aluminum pan, and thus, the normalized rheological function obtained by the rotational
crystallization heat released can be carried out of the sys- rheometer (Lamberti et al. 2007). The first model examined
tem by convection effectively. However, in the rheometer, was the one proposed by Graham (1981) for concentrated
the polymer is sandwiched in between two quartz plates of suspension of rigid interacting spheres. Graham proposed
43 mm in diameter with the upper plate having a thickness an additional term into Eq. 2 which modifies the behav-
of 6.4 mm. In the microscopy setup, the heat of crystal- ior for concentrated suspension of rigid interacting spheres
lization cannot be convected out rapidly. This can impede (Graham 1981).
crystal growth rate in the POM measurements and con- η
= 1 + 52 ∅ + 
ηa
9 
tributes to the lower degree of crystallinity as also shown 4 1


ϕ(1+0.5ϕ)(1+ϕ) 2

in Fig. 16a. Therefore, the differences observed in Fig. 16a 3 ∅  (2)
are clearly caused by the different thermal histories in each ϕ=2 1 − ∅max 3 ∅
∅max
case. Heat transfer calculations explained in Appendix A
demonstrate that the temperature rises in the POM samples ∅max is the maximum volume fraction occupied by the
in the range from 0.2 to 1.9 ◦ C at the crystallization tem- spheres. For the same system, Mooney (1951) proposed
peratures of 131.4 to 121.7 to 131.4 ◦ C. Using the Avrami
equation (see details in Appendix B) and the temperature
rise calculated from the heat transfer equations, the relative
crystallinity as a function of time was calculated for the case
of POM results, indicating excellent agreement in Fig. 16b.
This clearly explains the origin of the differences observed
in Fig. 16a.
The effect of shear on flow-induced crystallization was
also examined by the optical microscopy technique. Differ-
ent shear rates in the range of 0.0 to 2 s−1 were chosen to
investigate the effect of shear. The captured images were
undergone with image processing to obtain the degree of
space filling and relate this to degree of crystallization as
discussed above. The obtained shear stress growth coeffi-
cient (η+ ) from the rheometry was normalized by using
the value of steady-state viscosity that corresponds to a
crystal-free melt at the corresponding rate and tempera-
ture, indicated as ηa . The normalized viscosity as a function
of space filling is plotted in Fig. 17. Both space filling Fig. 15 The induction time for the onset of crystallinity as a function
of shear rate for PP1 at different normalized temperatures. The closed
images and normalized viscosity were obtained from the symbols correspond to halftime quiescent crystallization obtained from
same experimental test (identical temperature history). This DSC
Rheol Acta (2014) 53:519–535 531

Table 4 Optical microscopy images before and after image processing

Material Before image processing After image processing - method 1 After image processing - method 2

PP5
Morphology structure of PP5 Morphology structure of PP5 Morphology structure of PP5
after 685s at 138.8 ◦ C after 685s at 138.8 ◦ C after 685s at 138.8 ◦ C

Morphology structure of PP5 Morphology structure of PP5 Morphology structure of PP5


after 985s at 138.8 ◦ C after 985s at 138.8 ◦ C after 985s at 138.8 ◦ C

the following equation, while Frankel and Acrivos (1967) Richmond 1980; Kitano et al. 1981; Kataoka et al. 1978).
suggested Eq. 4: The latter is useful for the concentrated suspension of par-
  ticles of any shape with A being a shape coefficient (for
η 5 ∅
= exp ×  (3) instance, A = 0.68 for smooth spheres):
ηa 2 1−∅  − 5 ×∅max
∅max η 
= 1 − ∅ ∅max
2
(5)
 1/ ηa
∅ 3
η 9 ∅max η   −2
= × (4) = 1−∅A (6)
ηa 8  1/ ηa
1 − ∅ ∅max
3
To compare the capabilities of these equations, we have
The last two formulae which were studied in this work have set the adjustable parameters in such a way that all models
a similar form and were proposed by Ball and Richmond result in the same induction time at low shear rates. This
in 1980 (Eq. 5) and Kitano et al. in 1981 (Eq. 6) (Ball and induction time corresponds to the space filling required to

Fig. 16 The relative


crystallinity of PP5 as a function
of time obtained from DSC and
optical microscopy. a Raw data
without temperature correction
showing the effect of different
thermal histories in each test. b
The Avrami prediction of PP5
under microscopy setup
532 Rheol Acta (2014) 53:519–535

or less collapses as shown in Fig. 17. However, above a crit-


ical shear rate (e.g., 0.5 s−1 ), we likely deviate from the
Newtonian plateau, and the experimental NRF vs. degree of
filling curves deviate from the ones corresponding to the low
deformation conditions (≤0.1 s−1 ). Therefore, at the high
shear rate conditions (i.e., ≥0.5 s−1 ), the true NRF which is
affected by the shear rate and deviates from the low shear
rate curve, irrespective of the suspension model used.
The dependency of suspension model parameter on the
shear rate indicates that the viscosity of the semicrys-
talline system is a function of both crystallinity and
shear rate. Therefore, we suggest that a proper suspen-
sion/rheology model for semicrystalline systems should be a
frame-invariant constitutive equation with viscoelastic basis
and both crystallinity and deformation rate dependences
(Doufas et al. 2000; Doufas 2013; Tanner and Qi 2005,
2009).
Fig. 17 Normalized viscosity as a function of space filling (crys-
tallinity) for PP5

Conclusions
obtain normalized viscosity ratio, η/ηa of 1.2. The param-
eters which were used in the models are listed in Table 5. Flow-induced crystallization of several PP samples under a
The formulation of Frankel and Acrivos (Eq. 4) predicts an simple shear was studied in this work in order to elucidate
NRF of zero instead of one at zero degree of space filling the effects of flow deformation relative to quiescent con-
as shown in Fig. 17. Equations 3 and 5 exhibit the strongest ditions. In simple shear, there is a critical strain required
NRF dependence on the degree of space filling. Equations 2 for the onset of crystallization at given shear rates and tem-
and 6 show a similar trend and seem to predict the relation peratures. In shear experiments, the critical strain needed
between space filling and NRF better. Experimental data for a polymer to crystallize decreases with decreasing rate.
under a higher shear rate of 0.5 and 2 s−1 could not be Temperature was found to be an important variable in crys-
predicted by using the same model constants; therefore, A tallization as expected, and a change of temperature by a few
constant in Eq. 6 was reduced by a factor of 50 to mimic degrees changes the induction times for crystallization by 1
the behavior properly (see dash line curve in Fig. 17). The order of magnitude. The suspension models are applicable
suspension models are applicable at low deformation rates, only to low deformation rates, since there is only a depen-
since there is only a dependence of the viscosity on the dence of the viscosity on the degree of filling. A proper
degree of filling, but there is no explicit dependence on the model should have a viscoelastic constitutive equation basis
deformation rate (Newtonian-like models). A proper con- in order to be able to capture the crystallization behavior
stitutive equation for suspensions, however, should include well (e.g., the two-phase constitutive/microstructural model
both the degree of filling (crystallinity) and the deforma- for FIC of Doufas et al. (2000) and Doufas (2013)).
tion rate dependencies. For relatively small shear rates (e.g.,
≤0.1 s−1 in our case), the regime of viscosity of the New-
tonian plateau at the testing temperature is reached, and the Acknowledgments Financial assistance from the Natural Sciences
and Engineering Research Council (NSERC) of Canada, the scholar-
NRF vs. degree of space filling experimental curves more ship program of the University of British Columbia (4YF), and the
ExxonMobil Chemical are gratefully acknowledged.
Table 5 Models’ parameter used to predict the results in Fig. 17

Equation Parameter Value Appendix A


Eq. 6,a A 0.2
The coupled energy equations for the system shown in
Eq. 6,b A 0.004
Fig. 18 where the polymer in the space between two quartz
Eq. 5 ∅max 0.2313
plates is crystallized are developed. This is necessary to cal-
Eq. 4 ∅max 0.4233
culate the increase of temperature in the optical microscopy
Eq. 3 ∅max 0.2882
results due to the heat of crystallization which is accumu-
Eq. 2 ∅max 0.5
lated in the sample. This has an effect on the degree of
Rheol Acta (2014) 53:519–535 533

Fig. 18 The schematic of system used in this study

crystallinity which needs to be quantified in order to explain The heat transfer equations for the polymer and the quartz
the differences between the POM and DSC results with plate located on top of the polymer (Fig. 18) are as follows:
reference to Fig. 16a.     
ρpolymer Cp,polymer ∂T 1 ∂
∂t = r ∂r kpolymer r ∂T ∂
∂r + ∂z kpolymer
∂T
∂z + q̇
Crystal formation releases heat within the polymer at a  
ρquartz Cp,quartz ∂T 1 ∂ ∂T ∂ ∂T
∂t = r ∂r kquartz r ∂r + ∂z kquartz ∂z
rate of
(8)

q̇ = Hf (100% crystalline polymer) × ρ × ∅˙ (7)


The equations are solved using POLYFLOW with the
following boundary conditions. A convection boundary con-
where Hf (100% crystalline polymer) is the heat of melting if dition is imposed on the plate in contact with nitrogen gas
the polymer was 100 % crystalline, ρ is the density of crys- which is purged into the oven chamber of the rheometer as
tals, and ∅˙ is the rate of volume fraction increase of crystals. the heating/cooling gas. The heat transfer coefficient used
in these calculations was h = 100 W/(m2 •K). The quartz
plate underneath the polymer imposes an adiabatic bound-
ary condition at the lower side of the polymer. Continuity

Fig. 19 The temperature rise at r/R = 0.75 and z/h = 0.25 in


a polymer sample that crystallizes at different temperatures. The tem-
perature increases of 0.2 to 1.9 ◦ C at the crystallization temperatures
of 131.4 to 121.7 ◦ C, respectively, explain adequately the differences Fig. 20 Fit of the Arrhenius equation to the Avrami rate parameter k
in Fig. 16a using n = 2
534 Rheol Acta (2014) 53:519–535

of T and heat rate at the polymer/quartz interface completes References


the required BC.s.
After solving the system of Eq. 8, the temperature ele- Acierno S, Grizzuti N (2003) Measurements of the rheological behav-
ior of a crystallizing polymer by an inverse quenching technique.
vation within the polymer sample can be obtained as a
J Rheol 47:563
function of T . The temperature rise as a function of time at Acierno S, Palomba B, Winter HH, Grizzuti N (2003) Effect of molec-
r = 0.75R and z = 0.25 mm is shown in Fig. 19. The ular weight on the flow-induced crystallization of isotactic poly
temperature rise in the polymer matrix decreased as the tem- (1-butene). Rheol Acta 42(3):243–250
Ansari M, Hatzikiriakos SG, Sukhadia AM, Rohlfing DC (2011)
perature is increased. At higher temperatures, the kinetics Rheology of Ziegler–Natta and metallocene high-density
of crystallization slows down, and thus less heat is released polyethylenes: broad molecular weight distribution effects. Rheol
per unit of time. This essentially explains the larger differ- Acta 50(1):17–27
ences between OM and DSC results depicted in Fig. 16a at Baert J, Puyvelde PV, Langouche F (2006) Flow-induced crystalliza-
tion of PB-1: from the low shear rate region up to processing rates.
the lower temperatures. Macromolecules 39(26):9215–9222
Baert J, Van Puyvelde P (2006) Effect of molecular and processing
parameters on the flow-induced crystallization of poly-1-butene.
Appendix B Part 1: kinetics and morphology. Polymer 47(16):5871–5879
Ball RC, Richmond P (1980) Dynamics of colloidal dispersions. Phys
Chem Liq 9(2):99–116
The Avrami equation is often used to predict the isother- Bourgin P, Zinet M (2010) Thermally and flow induced crystalliza-
mal crystallization behavior under quiescent condition. This tion of polymers at low shear rate. J Non-Newton Fluid Mech
165(5):227–237
equation can be written as follows:
Boutahar K, Carrot C, Guillet J (1998) Crystallization of polyolefins
from rheological measurements relation between the transformed
 n
X Xf = 1 − e−kt
fraction and the dynamic moduli. Macromolecules 31(6):1921–
(9)
1929
Chellamuthu M, Arora D, Winter HH, Rothstein JP (2011) Extensional
flow-induced crystallization of isotactic poly-1-butene using a
where X is the degree of crystallinity, Xf is the total crys- filament stretching rheometer. J. Rheol 55:901
tallinity at the end of primary crystallization process, n is Coppola S, Balzano L, Gioffredi E, Maffettone PL, Grizzuti N (2004)
the Avrami index with a value of 2 for the resin studied, Effects of the degree of undercooling on flow induced crystalliza-
and k is the Avrami rate parameter. The Avrami parameter k tion in polymer melts. Polymer 45(10):3249–3256
Dai SC, Qi F, Tanner RI (2006) Strain and strain-rate formulation for
contains the temperature dependence of the nucleation and flow-induced crystallization. Polymer Eng Sci 46(5):659–669
crystal growth processes, while the exponent n depends on Derakhshandeh M, Hatzikiriakos SG (2012) Flow-induced crystalliza-
the geometry and dimensionality of the growth as well as on tion of high-density polyethylene: the effects of shear and uniaxial
extension. Rheol Acta 51(4):315–327
the nature of the nucleation process. The Avrami equation
Devaux N, Monasse B, Haudin JM, Moldenaers P, Vermant J (2004)
best predicts the behavior in primary growth region of crys- Rheooptical study of the early stages of flow enhanced crystalliza-
tals which corresponds to a relative crystallinity in between tion in isotactic polypropylene. Rheol Acta 43(3):210–222
30 and 70 %. The Avrami index determined experimentally D’Haese M, Goderis B, Van Puyvelde P (2011) The influence of
calcium-stearate-coated calcium carbonate and talc on the quies-
varied in the range of 1.9–2.5, and therefore, the theoretical
cent and flow-induced crystallization of isotactic poly (propylene).
value of 2 (1.9–2.7) was chosen for the investigated tem- Macromol Mater Eng 296(7):603–616
peratures (120–135 ◦ C). The Arrhenius Eq. 10 is used in D’Haese M, Van Puyvelde P, Langouche F (2010) Effect of particles
this study to represent the dependency of the Avrami rate on the flow-induced crystallization of polypropylene at processing
parameter on temperature. The Arrhenius equation is then speeds. Macromolecules 43(6):2933–2941
Doufas AK, Dairanieh IS, McHugh AJ (1999) A continuum model for
used in combination with the Avrami equation to predict the flow-induced crystallization of polymer melts. J Rheol 43:85–109
behavior of the sample in the microscopy setup. Doufas AK, McHugh AJ, Miller C (2000) Simulation of melt spinning
including flow-induced crystallization: part I. Model development
      and predictions. J Non-Newton Fluid Mech 92(1):27–66
k (T ) k (To ) = exp Ea R 1 T − 1 To (10) Doufas AK, McHugh AJ (2001) Simulation of film blowing including
flow-induced crystallization. J Rheol 45:1085
Doufas AK (2013) A microstructural flow-induced crystallization
The value of the Avrami rate parameter obtained using model for film blowing: validation with experimental data. Rheol
Acta 53:269–293
the temperature history in Fig. 19 along with the Arrhe- Duplay C, Monasse B, Haudin JM, Costa JL (2000) Shear-induced
nius equation is used in the Avrami equation (Fig. 20). The crystallization of polypropylene: influence of molecular weight. J
prediction of the Avrami model is shown in Fig. 16b and Mater Sci 35(24):6093–6103
compared with the experimental data from the microscopy Eder G, Janeschitz-Kriegl H, Liedauer S (1990) Crystallization pro-
cesses in quiescent and moving polymer melts under heat transfer
setup. The agreement was found to be excellent which conditions. Progr Polymer Sci 15(4):629–714
shows that the difference in Fig. 16a is related to different Ferry JD (1980) Viscoelastic properties of polymers, vol 3. Wiley, New
thermal histories during crystallization. York, p 641
Rheol Acta (2014) 53:519–535 535

Frankel NA, Acrivos A (1967) On the viscosity of a concentrated between in situ rheo-optics and ex situ structure determination.
suspension of solid spheres. Chem Eng Sci 22(6):847–853 Macromolecules 32(22):7537–7547
Eder G, Janeschitz-Kriegl H, Meijer HEH (eds) (1998) Materials Kwack TH, Han CD (1983) Rheology-processing-property relation-
science and technology, vol 18, chap 5. Wiley, Weinheim ships in tubular blown film extrusion. II. Low-pressure low-
Gelfer Y, Winter HH (1999) Effect of branch distribution on rheology density polyethylene. J Appl Polymer Sci 28(11):3419–3433
of LLDPE during early stages of crystallization. Macromolecules Lamberti G, Peters GWM, Titomanlio G (2007) Crystallinity and
32(26):8974–8981 linear rheological properties of polymers. Int Polymer Process
Godara A, Raabe D, Van Puyvelde P, Moldenaers P (2006) Influ- 22(3):303–310
ence of flow on the global crystallization kinetics of iso-tactic Macosko CW (1994) Rheology: principles, measurements, and appli-
polypropylene. Polymer Test 25(4):460–469 cations, 1st edn. Wiley, New York
Graham AL (1981) On the viscosity of suspensions of solid spheres. McHugh AJ, Edwards BJ (1995) Flow-induced structure formation
Appl Sci Res 37(3–4):275–286 in polymer solutions. In: LJ Jorgensen, K Sondergaard (eds)
Hadinata C, Gabriel C, Ruellman M, Laun HM (2005) Comparison Rheo-Physics of Multiphase Polymeric Systems, vol 5. Technomic
of shear-induced crystallization behavior of PB-1 samples with Publishing Company, Lancaster, pp 227–267
different molecular weight distribution. J Rheol 49(1):327–349 Mooney M (1951) The viscosity of a concentrated suspension of
Hadinata C, Gabriel C, Ruellmann M, Kao N, Laun HM (2006) Shear- spherical particles. J Colloid Sci 6(2):162–170
induced crystallization of PB-1 up to processing-relevant shear Natta G, Pino P, Corradini P, Danusso F, Mantica E, Mazzanti G,
rates. Rheola Acta 45(5):539–546 Moraglio G (1955) Crystalline high polymers of α-olefins. J Am
Hadinata C, Boos D, Gabriel C, Wassner E, Rüllmann M, Kao N, Laun Chem Soc 77(6):1708–1710
HM (2007) Elongation-induced crystallization of a high molecu- Pantani R, Speranza V, Titomanlio G (2001) Relevance of crystallisa-
lar weight isotactic polybutene-1 melt compared to shear-induced tion kinetics in the simulation of the injection molding process. Int
crystallization. J Rheol 51(2):195 Polymer Process 16(1):61–71
Han CD, Kwack TH (1983) Rheology-processing-property relation- Paradkar RP, Patel RM, Knickerbocker E, Doufas AK (2008) Raman
ships in tubular blown film extrusion. I. High-pressure low-density spectroscopy for spinline crystallinity measurements. I. Experi-
polyethylene. J Appl Polymer Sci 28(11):3399–3418 mental studies. J Appl Polymer Sci 109(5):3413–3420
Huo H, Jiang S, An L, Feng J (2004) Influence of shear on crystal- Patel RM, Doufas AK, Paradkar RP (2008) Raman spectroscopy for
lization behavior of the β phase in isotactic polypropylene with spinline crystallinity measurements. II. Validation of fundamental
β-nucleating agent. Macromolecules 37(7):2478–2483 fiber-spinning models. J Appl Polymer Sci 109(5):3398–3412
Jay F, Haudin JM, Monasse B (1999) Shear-induced crystalliza- Scelsi L, Mackley MR (2008) Rheo-optic flow-induced crystallisa-
tion of polypropylenes: effect of molecular weight. J Mater Sci tion of polypropylene and polyethylene within confined entry–exit
34(9):2089–2102 flow geometries. Rheol Acta 47(8):895–908
Janeschitz-Kriegl H (2003) How to understand nucleation in crystal- Stadlbauer M, Janeschitz-Kriegl H, Eder G, Ratajski E (2004) New
lizing polymer melts under real processing conditions. Colloid extensional rheometer for creep flow at high tensile stress. Part
Polymer Sci 281(12):1157–1171 II. Flow induced nucleation for the crystallization of iPP. J Rheol
Janeschitz-Kriegl H, Ratajski E (2005) Kinetics of polymer crystal- 48:631–641
lization under processing conditions: transformation of dormant Sun T, Brant P, Chance RR, Graessley WW (2001) Effect of short
nuclei by the action of flow. Polymer 46(11):3856–3870 chain branching on the coil dimensions of polyolefins in dilute
Janeschitz-Kriegl H, Ratajski E, Stadlbauer M (2003) Flow as an solution. Macromolecules 34:6812–6820
effective promotor of nucleation in polymer melts: a quantitative Swartjes FHM, Peters GW, Rastogi S, Meijer HE (2001) Stress
evaluation. Rheologica Acta 42(4):355–364 induced crystallization in elongational flow. Technische Univer-
Kataoka T, Kitano T, Sasahara M, Nishijima K (1978) Viscosity of siteit Eindhoven
particle filled polymer melts. Rheol Acta 17(2):149–155 Tanner RI, Qi F (2005) A comparison of some models for describing
Kim KH, Isayev AI, Kwon K (2005) Flow-induced crystallization in polymer crystallization at low deformation rates. J Non-Newton
the injection molding of polymers: a thermodynamic approach. J Fluid Mech 127(2):131–141
Appl Polymer Sci 95(3):502–523 Tanner RI, Qi F (2009) Stretching, shearing and solidification. Chem
Kitano T, Kataoka T, Shirota T (1981) An empirical equation of the Eng Sci 64(22):4576–4579
relative viscosity of polymer melts filled with various inorganic Tiang JS, Dealy JM (2012) Shear induced crystallization of iso-
fillers. Rheol Acta 20(2):207–209 tactic polypropylene studied by simultaneous light intensity and
Kitoko V, Keentok M, Tanner RI (2003) Study of shear and elonga- rheological measurements. Polymer Eng Sci 52(4):835–848
tional flow of solidifying polypropylene melt for low deformation Titomanlio G, Speranza V, Brucato V (1997) On the simulation of
rates. Korea Aust Rheol J 15:63–73 thermoplastic injection moulding process. 2. Relevance of inter-
Koscher E, Fulchiron R (2002) Influence of shear on polypropylene action between flow and crystallization. Int Polymer Process
crystallization: morphology development and kinetics. Polymer 12(1):45–53
43(25):6931–6942 Wang J, Dou Q (2008) Crystallization behaviors and optical properties
Kornfield JA, Kumaraswamy G, Issaian AM (2002) Recent advances of isotactic polypropylene: comparative study of a trisamide and
in understanding flow effects on polymer crystallization. Ind Eng a rosin-type nucleating agent. Colloid Polymer Sci 286(6-7):699–
Chem Res 41(25):6383–6392 705
Kuijk EW, Tas PP, Neuteboom P (1999) A rheological model for the White EEB, Winter HH, Rothstein JP (2012) Extensional-flow-
prediction of polyethylene blown film properties. J Reinforc Plast induced crystallization of isotactic polypropylene. Rheol Acta
Compos 18(6):508–517 51(4):303–314
Kumaraswamy G, Issaian AM, Kornfield JA (1999) Shear-enhanced Wunderlich B (1976) Macromolecular physics, crystal nucleation,
crystallization in isotactic polypropylene. 1. Correspondence growth, annealing, V2. Academic, New York

Vous aimerez peut-être aussi