Skip to main content
Erschienen in: Rheologica Acta 2/2012

Open Access 01.02.2012 | Original Contribution

Suspension-like hardening behavior of HDPE and time-hardening superposition

verfasst von: Peter C. Roozemond, Vincent Janssens, Peter Van Puyvelde, Gerrit W. M. Peters

Erschienen in: Rheologica Acta | Ausgabe 2/2012

Aktivieren Sie unsere intelligente Suche, um passende Fachinhalte oder Patente zu finden.

search-config
loading …

Abstract

The rheology of solidifying high-density polyethylene (HDPE) is investigated. Experiments on an HDPE were performed with a novel RheoDSC device. Results agree quantitatively with simulations for a suspension of elastic spheres in a viscoelastic matrix except for very low values of space filling (<5%), indicating that the rheological behavior of the crystallizing melt in the frequency range investigated is purely suspension like. The hardening behavior of the material is characterized in two different ways; a normalized rheological function and a time-hardening superposition (THS) master curve of rheological properties. An improvement is proposed to the procedure for performing THS that was previously used in the literature. Based on this procedure, a novel method for predicting the rheological properties of crystallizing melts is presented.

Introduction

Detailed knowledge of the change in rheology of semi-crystalline polymers during solidification is of vital importance for the modeling of processes such as injection molding, fiber spinning, electrospinning, etc. because the rheological properties, flow field, and crystallization kinetics are all intimately coupled (Custódio et al. 2009; Steenbakkers and Peters 2010). A vast amount of experiments investigating the change of rheological properties can be found in the literature, cf. Acierno and Grizzuti (2003), Boutahar et al. (1998), Han and Wang (1997), Lamberti et al. (2007), Pantani et al. (2001), Pogodina and Winter (1998), and Titomanlio et al. (1997). To quantify the hardening behavior of the material, a normalized rheological function (NRF) is commonly used.
$$ \Gamma(\xi)=\frac{|\eta^*(\xi)|}{|\eta^*(\xi=0)|} $$
(1)
where Γ is the hardening function, ξ space filling or relative crystallinity, and η * the complex viscosity. Among the experimental data gathered, little agreement is found between studies, as shown by Lamberti et al. (2007). This is largely because the experimental procedure commonly used to obtain the hardening curve is prone to experimental errors: rheological properties and space filling (both as a function of time) are separately measured in a rheometer (usually with less precise temperature control) and a differential scanning calorimetry (DSC) apparatus, respectively. Time–viscosity and time–space filling data are then combined to obtain a hardening curve. A slight difference in behavior in the two experiments might result in a large error in the hardening function. For instance, Steenbakkers and Peters (2008) found the experiments from Boutahar et al. (1996, 1998) showed suspension-like behavior but in order to get quantitative agreement between experiments and simulations, the space filling obtained from DSC measurements had to be adjusted significantly.
An improvement on this cumbersome method was proposed by Lamberti et al. (2007). They developed a technique that eliminates the need for obtaining space filling and rheological properties at exactly the same temperature. However, experiments were still carried out in two different devices. More recently, the group in Leuven developed a novel apparatus that combines DSC and rheometry: the RheoDSC device (Janssens et al. 2009, 2010; Kiewiet et al. 2008). With this device, rheological and thermal signals can be obtained simultaneously on the same sample and hence both types of experiments are conducted under identical conditions.
Lamberti et al. (2007) studied a novel method to describe the viscosity increase in solidifying polymers. They found that, similar to a time–temperature master curve in time–temperature superposition (TTS), a time-hardening master curve can be constructed. An empirical relation for the space filling shift factor was proposed. Custódio et al. (2009) implemented this method of predicting viscosity as a function of space filling in their code for the simulation of shear flow in a multi-pass rheometer.
A large number of analytical models for the hardening function of suspensions have been developed (Ball and Richmond 1980; Batchelor 1977; Einstein 1906, 1911; Frankel and Acrivos 1967; Graham 1981; Katoaka et al. 1978; Kitano et al. 1981; Krieger and Dougherty 1959; Mooney 1951). However, many of these models are not applicable to crystallizing polymer melts. Empirical models can also be found in abundance in literature (Boutahar et al. 1998; Doufas et al. 2000; Han and Wang 1997; Hieber 2002; Katayama and Yoon 1985; Khanna 1993; Shimizu et al. 1985; Tanner 2003; Titomanlio et al. 1997; Yarin 1992; Ziabicki 1988; Zuidema et al. 2001), but there is little agreement between these models. Thus, there is need for a model that is able to accurately describe rheological properties of semi-crystalline polymers during solidification.
Steenbakkers and Peters (2008) reviewed a number of suspension-based models and found that the generalized self-consistent method (GSCM) developed by Christensen (1990); Christensen and Lo (1979, 1986) describes accurately the rheology of crystallizing isotactic polypropylene (iPP). However, the GSCM is not applicable to the low frequency range, where gel-like behavior instead of suspension-like behavior is observed (Pogodina and Winter 1998; Pogodina et al. 1999a, b; Winter and Mours 1997). An extension of the GSCM, including percolation (Alberola and Mele 1996), did not solve this problem.
The goal of this study is to investigate the applicability of the generalized self-consistent method to predicting the hardening curve and space filling shift factors of a material. To this end, data from experiments in the novel RheoDSC device on high-density polyethylene (HDPE) will be used for validation purposes. The experimental results and time-hardening superposition of these results are presented in the “Experiments” section. The “Simulations” section shows the comparison between simulations with the GSCM and the experimental results.

Experiments

A crystallizing HDPE was analyzed in the prototype RheoDSC device. This instrument was constructed starting from two stand-alone commercial instruments: a Q2000 TzeroTMDSC (TA Instruments) and an AR-G2 rheometer (TA Instruments). The DSC apparatus uses the TzeroTMtechnology, which is important for the correction of the thermal resistances and capacitances of the various heat flow paths in the cell caused by the modification of the calorimetric cell environment. The AR-G2 rheometer is a sensitive instrument opening the range of ultra-low torques, which is of importance as the diameter of the measurement geometry is chosen to be of comparable dimension as the DSC sensor (5 mm). For more details the reader is referred to Kiewiet et al. (2008). Both the rheology and thermal signals of the RheoDSC device have been validated, as reported by Janssens et al. (2009, 2010) and Kiewiet et al. (2008).
Experiments were performed at 124°C at five oscillatory frequencies. Results for the complex viscosity, modulus and phase angle are presented in Fig. 1. Different symbols indicate the increasing space filling. The experiments show a large difference in rheology between the experiments with zero and just six percent relative crystallinity, much larger than what would be caused by a suspension of spherulites with this amount of space filling. The effect is accounted for by extrapolating the data for the crystallizing melt to obtain adjusted values for the rheology of the melt, as will be explained in the “Early stage of crystallization” section. The modulus for the fully crystallized material (ξ = 1) is also obtained by extrapolation. Values are shown in Table 1.
Table 1
Storage and loss moduli for the melt, \(G^*_0\), and for the fully crystallized material, \(G_1^*\)
ω [rad/s]
G 0′ [MPa]
G 1′ [MPa]
G 0′′ [MPa]
G 1′′ [MPa]
1.00
0.084
2.29
0.062
0.692
3.16
0.142
2.78
0.087
0.757
10.0
0.223
3.36
0.121
0.786
31.6
0.324
3.90
0.132
0.760
100
0.432
4.42
0.148
0.734

Hardening function

Figure 2 shows the hardening function, calculated using a NRF based on the magnitude of the complex viscosity versus space filling for the five frequencies investigated. The NRF is given by
$$ \Gamma(\xi,\omega)=\frac{|\eta^*(\xi,\omega)|}{|\eta^*(\xi=0,\omega)|}. $$
(2)
The hardening curve is similar in shape to what was observed in the literature (Janssens et al. 2010; Lamberti et al. 2007), taking larger values for lower frequencies throughout the range of space filling. It is emphasized that in accordance with Lamberti et al. (2007) and Janssens et al. (2010), little change in rheology is observed for space filling below 10%, whereas other authors have reported order of magnitude changes for ξ < 10% (Acierno and Grizzuti 2003; Titomanlio et al. 1997).

Time-hardening superposition

Lamberti et al. (2007) found that in a frequency-viscosity plot, measurements for crystallizing iPP with different space fillings can be shifted to a single master curve, analogous to a TTS master curve. This was done by shifting the relaxation times of the material with a ξ , so that the shifted frequency and viscosity become
$$ \omega_{\rm master}=\omega*a_\xi(\xi)$$
(3)
$$ \eta_{\rm master}=\eta/a_\xi(\xi) $$
(4)
Lamberti et al. also found that the shift factor a ξ was reasonably well described by the empirical relation
$$ \log_{10}a_\xi=\alpha_1\xi^{\alpha_2} $$
(5)
where α 1 and α 2 are fitting parameters.
Viscosity data from the present experiments were shifted in the same way. The resulting master curve is shown in Fig. 3a. The curve for ξ = 0 corresponds to the values given in Table 1. Shift factor versus space filling is shown in Fig. 3b. The empirical relation from Lamberti et al. is depicted with the solid line. We observe that Lamberti’s empirical relation is able to capture our data quite well with parameters α 1 = 5.19, α 2 = 1.20. Lamberti studied two grades of iPP and found α 1 = 3.875, α 2 = 1.236 for one of the grades and α 1 = 12.073, α 2 = 3.166 for the other.
Actually, the procedure that Lamberti et al. used to obtain a master curve differs from the way in which TTS is conventionally conducted. In TTS the shift factor for the relaxation times of the material a ξ is obtained by shifting the phase angle and subsequently the master curve for the viscosity would be obtained by applying the shift factor for relaxation times, and an additional vertical shift factor b ξ ;
$$ \eta_{\rm master}=b_\xi(\xi)*\eta/a_\xi(\xi) $$
(6)
If the vertical shift factor b ξ is of order 1, which is usually the case in TTS, the difference between the one-step procedure that Lamberti et al. used and the conventional two-step shifting procedure is small. However, if we apply Lamberti’s procedure, i.e., obtaining a ξ through shifting the complex viscosity, to our measurements, we observe that the phase angle curves shifted with this a ξ do not form a single master curve. This is shown in Fig. 4. This means that although this procedure could be used to calculate viscosity as a function of space filling, a ξ does not shift the relaxation times of the material. Therefore, the viscosity master curve constructed in this way (Fig. 3a), with its promising range of eight decades, has no physical meaning. Consequently, this master curve can not be used to obtain the complex viscosity of the pure melt at very high frequency by measuring the complex viscosity of the partly crystallized material at a lower frequency.
For this reason, we propose to use the two-step procedure commonly used in TTS. The result of this procedure is shown in Fig. 5. Contrary to the one-step procedure, this yields a smooth master curve for the phase angle, indicating that the shift factor a ξ obtained in this way is the correct shift factor for the relaxation times of the material. Compare Figs. 4 and 5a. For the complex viscosity we also obtain a smooth master curve. Again the measurements with ξ = 0 correspond to the values in Table 1. Unfortunately, the frequency range that the master curves span is reduced to four decades. The shift factors that were used to construct these master curves are given in Fig. 6 and Table 2. Although the procedure for constructing a master curve is different, the empirical fit by Lamberti et al. is still able to describe the data for the timescale shift factor. The values for vertical shift factor b ξ can be described by
$$ \log_{10}b_\xi=\beta_1\xi^2-\beta_2\xi. $$
(7)
We observe that the phase angle measurements for the lowest frequency (ω = 1 rad/s) slightly deviate from the master curve, see Fig. 5a. This indicates that time-hardening superposition (THS) might not give a good representation at lower oscillatory frequencies. We come back to this, and to our interpretation of the meaning of the parameters a ξ and b ξ , in “Time-hardening superposition with the GSCM” section.
Table 2
Space filling shift factors
ξ
a ξ (experimental)
a ξ (simulations)
b ξ (experiments)
b ξ (simulations)
0
1.00
1.00
1.00
1.00
0.063
1.05
1.01
0.89
0.88
0.21
1.56
1.19
0.60
0.65
0.42
3.03
2.17
0.41
0.42
0.50
3.45
2.97
0.35
0.35
0.64
6.81
6.35
0.28
0.28
0.83
19.4
15.7
0.23
0.21
0.95
53.2
32.3
0.19
0.18
0.99
194.4
75.3
0.19
0.17

Simulations

The generalized self-consistent method

The crystallizing melt is modeled as a suspension of spherical soft particles (the spherulites) in a viscoelastic matrix (the amorphous melt). The properties of such a system can be calculated using the three-dimensional (3D) GSCM, which was developed to describe the rheological properties of a suspension of spherical soft particles in an elastic medium. The basic problem of the 3D GSCM is shown in Fig. 7. The suspension is represented by a unit cell consisting of a particle with properties of the dispersed phase and a surrounding matrix shell with properties of the matrix phase. Their radii are a and b, respectively. The volume fraction of the particles is given by
$$ \xi=\left(\frac{a}{b}\right)^3 $$
(8)
The unit cell is suspended in an infinitely extending medium, which has the (unknown) effective properties of the suspension. The solution of the three phase problem yields the effective properties of the composite medium of an isotropic matrix phase into which is embedded the isotropic inclusion phase. The result for the effective medium problem of Fig. 7 is given by the solution of the quadratic equation (Christensen and Lo 1979, 1986)
$$ Af_G^2+Bf_G+C=0 $$
(9)
where f G is the relative modulus,
$$ f_G = \frac{G}{G_0} $$
(10)
with G the modulus of the suspension and G 0 the modulus of the matrix. The coefficients AB and C depend on space filling ξ, the Poisson’s ratios of the matrix and dispersed phase, ν 0 and ν 1, respectively, and the ratio between the moduli of the dispersed phase and the matrix, μ = G 1/G 0. Expressions for the coefficients AB and C are given in the Appendix. Steenbakkers and Peters (2008) applied the GSCM, derived for elastic materials, to viscoelastic materials by utilizing the correspondence principle Hashin (1965, 1970a, b). In that case, the relative modulus is complex;
$$ f_G^*(\omega)=\frac{G^*(\omega)}{G^*_0(\omega)} $$
(11)
and so is the modulus ratio \(\mu^*=G_1^*/G_0^*\). The relative modulus can be calculated from
$$ A^*{f_G^*}^2+B^*f_G^*+C^*=0 $$
(12)
where the complex coefficients A *, B * and C * follow from AB and C when μ is replaced by μ *. The absolute value of the complex relative modulus, \(\vert f_G^* \vert\), is equal to the NRF as defined in Eq. 2.
The 3D GSCM was developed for systems with purely spherical dispersed particles. Growing spherulites will impinge at some point and as a result the dispersed particles in this system will not be purely spherical at high volume fractions (above ~70 %). Previous studies do show a discrepancy between the GSCM and rheology experiments at very high volume fractions (~98%) (Housmans et al. 2009; Steenbakkers and Peters 2008). However, below this value there seems to be no noticeable effect of impingement on the hardening behavior.
In this study, we examine only quiescent crystallization, with crystalline structures that are spherical (before impingement). Flow-induced crystallization may yield oriented cylindrical crystalline structures. Steenbakkers and Peters (2008) found that these have qualitatively similar hardening behavior but were not able to quantitatively capture the behavior using a 2D suspension model. More research on this will be done in future work.

Early stage of crystallization

We observe that the change in material properties from 0% to 6% space filling is much larger than would be expected from a suspension. This is illustrated in Fig. 8, where we show measurements and simulations for the storage and loss moduli versus space filling for ω = 1 rad/s. Figure 8a shows simulations where we have taken the modulus of the pure melt \(G_0^*\) as the measured modulus of the material with no space filling. The simulations consistently underpredict the measured moduli. The discrepancy can be resolved by extrapolating \(G_0^*\) from the data for the crystallizing melt. This was done for the simulations shown in Fig. 8b. With this value for \(G_0^*\) the simulations and experiments show quantitative agreement for the full range of space filling.
In our view, this indicates that between 0% and 6% space filling a process has taken place that has altered the melt in a way that can not be described by suspension-like behavior. Something very similar is observed in the data from Carrot et al. (1993), who show the evolution of storage and loss modulus for a crystallizing HDPE melt under isothermal conditions at different temperatures. A strong and very steep increase in modulus is seen soon after the onset of crystallization (~200 seconds). After this, the modulus increases gradually until crystallization is complete (103–104 s, depending on the temperature). Carrot et al. (1993) did not comment on the phenomenon.
Perhaps a network of physical crosslinks is formed (Boutahar et al. 1996, 1998; Coppola et al. 2006; Peters et al. 2002; Pogodina et al. 1999a; Zuidema et al. 2001). Subsequently, spherulites grow in this altered melt, the effect of which is well described by suspension-like behavior. What exactly happens during the early stages of crystallization remains a lively research topic. For now, we deal with the phenomenon by adjusting the values for the moduli of the melt as described above. Results that are presented in this study are obtained with the adjusted value for the rheology of the melt.

Comparison with experiments

Figure 9 shows the experimental quantities depicted in Fig. 1 and the results for simulations with the GSCM for the same conditions. Quantitative agreements are observed for the storage and loss modulus and for the magnitude of the complex viscosity (for ξ ≠ 0, as explained above). For 0.063 < ξ < 0.64 the GSCM is less accurate in predicting the phase angle, however agreement is still qualitative.

Hardening function

Figure 10 shows the hardening curve for the material, both from experiments and simulations. We observe quantitative agreement, the hardening function of this material is very well described by suspension-like behavior.

Time-hardening superposition with the GSCM

The master curves for phase angle and viscosity obtained using the two-step procedure are shown in Fig. 11. The corresponding shift factors are given in Fig. 12 and Table 2. We observe a good agreement; the 3D GSCM model can be used to calculate a ξ as well as b ξ .
For the lowest frequency measurements and simulations slightly deviate from the phase angle master curve. Because this is observed both for measurements and simulations we cannot attribute this to gel-like behavior, as the simulations capture only suspension-like behavior. This means that for lower frequencies, THS might be less accurate.
In their review of the applicability of suspension models to polymer crystallization, Steenbakkers and Peters (2008) proposed a phenomenological model for hardening behavior. Suppose a relaxation spectrum (G i λ i ) with M modes is known for the melt so that storage and loss modulus as a function of frequency are given by
$$ G'(\omega)=\sum\limits_i^MG_{0,i}\frac{\lambda_{0,i}^2\omega^2}{1+\lambda_{0,i}^2\omega^2}, $$
(13)
$$ G''(\omega)=\sum\limits_i^MG_{0,i}\frac{\lambda_{0,i}\omega}{1+\lambda_{0,i}^2\omega^2}. $$
(14)
Now, if the number of modes is the same for both the liquid and solid phase, the moduli and relaxation times of the solid phase can be expressed in terms of the moduli and relaxation times of the matrix as
$$ G_i(\xi)=k_{G,i}(\xi)G_{0,i}, $$
(15)
$$ \lambda_i(\xi)=k_{\lambda,i}(\xi)\lambda_{0,i}. $$
(16)
with 1 ≤ k G,i  ≤ G 1,i /G 0,i and 1 ≤ λ G,i  ≤ λ 1,i /λ 0,i . Storage and loss moduli as a function of frequency and spacefilling can then be expressed as
$$ G'(\omega,\xi)=\sum\limits_i^Mk_{G,i}(\xi)G_{0,i}\frac{k_{\lambda,i}(\xi)^2\lambda_{0,i}^2\omega^2}{1+k_{\lambda,i}(\xi)^2\lambda_{0,i}^2\omega^2}, $$
(17)
$$ G''(\omega,\xi)=\sum\limits_i^Mk_{G,i}(\xi)G_{0,i}\frac{k_{\lambda,i}(\xi)\lambda_{0,i}\omega}{1+k_{\lambda,i}(\xi)^2\lambda_{0,i}^2\omega^2}. $$
(18)
Steenbakkers and Peters did not derive expressions for k G and k λ . The results obtained here, i.e., the fact that a time-hardening master curve can be constructed, imply that for both k λ and k G , a single value suffices for all modes, and that the storage and loss moduli as a function of space filling can be calculated using
$$ G'(\omega,\xi)=k_G(\xi)\sum\limits_i^MG_{0,i}\frac{k_\lambda(\xi)^2\lambda_{0,i}^2\omega^2}{1+k_\lambda(\xi)^2\lambda_{0,i}^2\omega^2} $$
(19)
$$ G''(\omega,\xi)=k_G(\xi)\sum\limits_i^MG_{0,i}\frac{k_\lambda(\xi)\lambda_{0,i}\omega}{1+k_\lambda(\xi)^2\lambda_{0,i}^2\omega^2}. $$
(20)
with k λ  = a ξ and k G  = 1/b ξ .
In this study, we have presented two ways to calculate the rheological properties of a crystallizing material as a function of space filling and oscillatory frequency. Figure 13 schematically shows both, for purpose of clarity. Time-hardening superposition is shown in Fig. 13a; the two shift factors k λ and k G are independent of frequency and shift the whole G *(ω) curve at once. Figure 13b shows calculation of the storage modulus via a relative modulus \(f_G^*\) which is calculated for each ξ and ω using the GSCM.

Conclusions

The evolution of the rheology of an HDPE during solidification was analyzed. Quantitative agreement with a suspension-based model, the 3D generalized self-consistent method, was found. This indicates that in the frequency range investigated, a suspension with spherical soft particles in a viscoelastic matrix is a good representation of an amorphous melt with growing spherulites. In the early stage of crystallization, a phenomenon is observed that can not be explained by suspension-like behavior. This was dealt with by using adjusted values for the complex modulus of the melt.
THS was performed. An improvement on the procedure developed in literature is proposed. The fact that a THS master curve can be constructed means that with a known Maxwell spectrum of the melt, the properties of a crystallizing melt can be obtained using two shift parameters that are mode and frequency independent and which can be easily measured or obtained from a suspension-based model.

Acknowledgements

This work is part of a collaboration with the group of Prof. Julia Kornfield at the California Institute of Technology. The project is supported by the Dutch Technology Foundation (STW), grant no. 08083 and the National Science Foundation, grant no. DMR-0710662.

Open Access

This article is distributed under the terms of the Creative Commons Attribution Noncommercial License which permits any noncommercial use, distribution, and reproduction in any medium, provided the original author(s) and source are credited.
Open AccessThis is an open access article distributed under the terms of the Creative Commons Attribution Noncommercial License (https://​creativecommons.​org/​licenses/​by-nc/​2.​0), which permits any noncommercial use, distribution, and reproduction in any medium, provided the original author(s) and source are credited.
Anhänge

Appendix

Coefficients in the GSCM

Expressions for the coefficients AB, and C in the 3D generalized self-consistent model derived by Christensen (1990), Christensen and Lo (1979, 1986) are shown in Table 3.
Table 3
Expressions for the coefficients in the 3D GSCM
\(\left. \begin{array}{l} A \\ B \\ C \end{array} \right\} = c_1\left(\mu-1\right)\eta_1\xi^{(10/3)} + c_2\left[63\left(\mu-1\right)\eta_2+2\eta_1\eta_3\right]\xi^{(7/3)}+c_3\left(\mu-1\right)\eta_2\xi^{(5/3)} + c_4\left(\mu-1\right)\eta_2\xi+c_5\eta_2\eta_3 \)
with...
... for A:
... for B:
... for C:
c 1
8(4 − 5ν 0)
− 4(1 − 5ν 0)
− 4(7 − 5ν 0)
c 2
− 2
4
− 2
c 3
252
− 504
252
c 4
\(-50(7-12\nu_0+8\nu_0^2)\)
150(3 − ν 0)ν 0
\(-25(7-\nu_0^2)\)
c 5
4(7 − 10ν 0)
3(7 − 15ν 0)
− (7 + 5ν 0)
η 1
\(\left(\mu-1\right)(7-10\nu_0)(7+5\nu_1)+105(\nu_1-\nu_0)\)
  
η 2
\(\left(\mu-1\right)(7+5\nu_1)+35(1-\nu_1)\)
  
η 3
ł
  

Poisson’s ratios

To solve Eq. 12, the Poisson’s ratios of the continuous and dispersed phase are required. They can be calculated using the well-known formula (Timoshenko and Goodier 1951)
$$ \nu=\frac{3K-2G}{2(3K+G)}, $$
(21)
where K is the bulk modulus and G is the shear modulus. The bulk modulus of HDPE K ≈ 1.5 GPa (Aleman 1990). With the plateau for the storage modulus at approximately 0.4 MPa for ξ = 0 and at approximately 6 MPa for ξ ≈ 1 (see Fig. 1), we find estimates of ν 0 = 0.4999, ν 1 = 0.4980.
Literatur
Zurück zum Zitat Acierno S, Grizzuti N (2003) Measurements of the rheological behavior of a crystallizing polymer by an “inverse quenching” technique. J Rheol 47(2):563–576CrossRef Acierno S, Grizzuti N (2003) Measurements of the rheological behavior of a crystallizing polymer by an “inverse quenching” technique. J Rheol 47(2):563–576CrossRef
Zurück zum Zitat Alberola N, Mele P (1996) Viscoelasticity of polymers filled by rigid or soft particles: theory and experiment. Polym Compos 17(5):751–759CrossRef Alberola N, Mele P (1996) Viscoelasticity of polymers filled by rigid or soft particles: theory and experiment. Polym Compos 17(5):751–759CrossRef
Zurück zum Zitat Aleman J (1990) Bulk and surface compression viscositeis of high and low density polyethylenes. Polym Eng Sci 30(6):326–334CrossRef Aleman J (1990) Bulk and surface compression viscositeis of high and low density polyethylenes. Polym Eng Sci 30(6):326–334CrossRef
Zurück zum Zitat Ball R, Richmond P (1980) Dynamics of colloidal dispersions. Phys Chem Liq 9(2):99–116CrossRef Ball R, Richmond P (1980) Dynamics of colloidal dispersions. Phys Chem Liq 9(2):99–116CrossRef
Zurück zum Zitat Batchelor G (1977) Effect of brownian motion on bulk stress in a suspension of spherical particles. J Fluid Mech 83:97–117CrossRef Batchelor G (1977) Effect of brownian motion on bulk stress in a suspension of spherical particles. J Fluid Mech 83:97–117CrossRef
Zurück zum Zitat Boutahar K, Carrot C, Guillet J (1996) Polypropylene during crystallization from the melt as a model for the rheology of molten-filled polymers. J Appl Polym Sci 60:103–114CrossRef Boutahar K, Carrot C, Guillet J (1996) Polypropylene during crystallization from the melt as a model for the rheology of molten-filled polymers. J Appl Polym Sci 60:103–114CrossRef
Zurück zum Zitat Boutahar K, Carrot C, Guillet J (1998) Crystallization of polyolefins from rheological measurements—relation between the transformed fraction and the dynamic moduli. Macromolecules 31:1921–1929CrossRef Boutahar K, Carrot C, Guillet J (1998) Crystallization of polyolefins from rheological measurements—relation between the transformed fraction and the dynamic moduli. Macromolecules 31:1921–1929CrossRef
Zurück zum Zitat Carrot C, Guillet J, Boutahar K (1993) Rheological behavior of a semi-crystalline polymer during isothermal crystallization. Rheol Acta 32:566–574CrossRef Carrot C, Guillet J, Boutahar K (1993) Rheological behavior of a semi-crystalline polymer during isothermal crystallization. Rheol Acta 32:566–574CrossRef
Zurück zum Zitat Christensen R (1990) A critical evaluation for a class of micro-mechanics models. J Mech Phys Solids 38:379–404CrossRef Christensen R (1990) A critical evaluation for a class of micro-mechanics models. J Mech Phys Solids 38:379–404CrossRef
Zurück zum Zitat Christensen R, Lo K (1979) Solutions for effective shear properties in three phase sphere and cylinder models. J Mech Phys Solids 27:315–330CrossRef Christensen R, Lo K (1979) Solutions for effective shear properties in three phase sphere and cylinder models. J Mech Phys Solids 27:315–330CrossRef
Zurück zum Zitat Christensen R, Lo K (1986) Erratum: solutions for effective shear properties in three phase sphere and cylinder models. J Mech Phys Solids 34:639CrossRef Christensen R, Lo K (1986) Erratum: solutions for effective shear properties in three phase sphere and cylinder models. J Mech Phys Solids 34:639CrossRef
Zurück zum Zitat Coppola S, Acierno S, Grizzuti N, Vlassopoulos D (2006) Viscoelastic behavior of semicrystalline thermoplastic polymers during the early stages of crystallization. Macromolecules 39:1507–1514CrossRef Coppola S, Acierno S, Grizzuti N, Vlassopoulos D (2006) Viscoelastic behavior of semicrystalline thermoplastic polymers during the early stages of crystallization. Macromolecules 39:1507–1514CrossRef
Zurück zum Zitat Custódio F, Steenbakkers R, Anderson P, Peters G, Meijer H (2009) Model development and validation of crystallization behavior in injection molding prototype flows. Macromol Theory Simul 18(9):469–494CrossRef Custódio F, Steenbakkers R, Anderson P, Peters G, Meijer H (2009) Model development and validation of crystallization behavior in injection molding prototype flows. Macromol Theory Simul 18(9):469–494CrossRef
Zurück zum Zitat Doufas A, McHugh A, Miller C (2000) Simulation of melt spinning including flow-induced crystallization - part I. Model development and predictions. J Non-Newton Fluid Mech 92(1):27–66CrossRef Doufas A, McHugh A, Miller C (2000) Simulation of melt spinning including flow-induced crystallization - part I. Model development and predictions. J Non-Newton Fluid Mech 92(1):27–66CrossRef
Zurück zum Zitat Einstein A (1906) A new determination of molecular dimensions. Ann Phys 19:289–306CrossRef Einstein A (1906) A new determination of molecular dimensions. Ann Phys 19:289–306CrossRef
Zurück zum Zitat Einstein A (1911) Correction to: a new determination of molecular dimensions. Ann Phys 34:591–592CrossRef Einstein A (1911) Correction to: a new determination of molecular dimensions. Ann Phys 34:591–592CrossRef
Zurück zum Zitat Frankel N, Acrivos A (1967) On viscosity of a concentrated suspension of solid spheres. Chem Eng Sci 22(6):847–853CrossRef Frankel N, Acrivos A (1967) On viscosity of a concentrated suspension of solid spheres. Chem Eng Sci 22(6):847–853CrossRef
Zurück zum Zitat Graham A (1981) On the viscosity of suspensions of solid spheres. Appl Sci Res 37(3–4):275–286CrossRef Graham A (1981) On the viscosity of suspensions of solid spheres. Appl Sci Res 37(3–4):275–286CrossRef
Zurück zum Zitat Han S, Wang K (1997) Shrinkage prediction for slowly-crystallizing thermoplastic polymers in injection molding. Int Polym Process 12(3):228–237 Han S, Wang K (1997) Shrinkage prediction for slowly-crystallizing thermoplastic polymers in injection molding. Int Polym Process 12(3):228–237
Zurück zum Zitat Hashin Z (1965) Viscoelastic behavior of heterogeneous media. J Appl Mech 32(3):630–636CrossRef Hashin Z (1965) Viscoelastic behavior of heterogeneous media. J Appl Mech 32(3):630–636CrossRef
Zurück zum Zitat Hashin Z (1970a) Complex moduli of viscoelastic composites—I. General theory and application to particulate composites. Int J Solids Struct 6:539–552CrossRef Hashin Z (1970a) Complex moduli of viscoelastic composites—I. General theory and application to particulate composites. Int J Solids Struct 6:539–552CrossRef
Zurück zum Zitat Hashin Z (1970b) Complex moduli of viscoelastic composites—II. Fiber reinforced materials. Int J Solids Struct 6:797–807CrossRef Hashin Z (1970b) Complex moduli of viscoelastic composites—II. Fiber reinforced materials. Int J Solids Struct 6:797–807CrossRef
Zurück zum Zitat Hieber C (2002) Modeling/simulating the injection molding of isotactic polypropylene. Polym Eng Sci 42(7):1387–1409CrossRef Hieber C (2002) Modeling/simulating the injection molding of isotactic polypropylene. Polym Eng Sci 42(7):1387–1409CrossRef
Zurück zum Zitat Housmans JW, Steenbakkers RJA, Roozemond PC, Peters GWM, Meijer HEH (2009) Saturation of pointlike nuclei and the transition to oriented structures in flow-induced crystallization of isotactic polypropylene. Macromolecules 42:5728–5740CrossRef Housmans JW, Steenbakkers RJA, Roozemond PC, Peters GWM, Meijer HEH (2009) Saturation of pointlike nuclei and the transition to oriented structures in flow-induced crystallization of isotactic polypropylene. Macromolecules 42:5728–5740CrossRef
Zurück zum Zitat Janssens V, Block C, van Assche G, van Mele B, van Puyvelde P (2009) RheoDSC: design and validation of a new hybrid measurement technique. J Therm Anal Calorim 98(3):675–681CrossRef Janssens V, Block C, van Assche G, van Mele B, van Puyvelde P (2009) RheoDSC: design and validation of a new hybrid measurement technique. J Therm Anal Calorim 98(3):675–681CrossRef
Zurück zum Zitat Janssens V, Block C, van Assche G, van Mele B, van Puyvelde P (2010) RheoDSC: analysis of the hardening of semi-crystalline polymers during quiescent isothermal crystallization. Int Polym Process 25(4):304–310CrossRef Janssens V, Block C, van Assche G, van Mele B, van Puyvelde P (2010) RheoDSC: analysis of the hardening of semi-crystalline polymers during quiescent isothermal crystallization. Int Polym Process 25(4):304–310CrossRef
Zurück zum Zitat Katayama K, Yoon M (1985) Chapter 8. Polymer crystallization in melt spinning: mathematical simulation. In: Ziabicki A, Kawai H (eds) High speed fibre spinning. Wiley, New York Katayama K, Yoon M (1985) Chapter 8. Polymer crystallization in melt spinning: mathematical simulation. In: Ziabicki A, Kawai H (eds) High speed fibre spinning. Wiley, New York
Zurück zum Zitat Katoaka T, Kitano T, Sasahara M, Nishijima K (1978) Viscosity of particle filled polymer melts. Rheol Acta 17(2):149–155CrossRef Katoaka T, Kitano T, Sasahara M, Nishijima K (1978) Viscosity of particle filled polymer melts. Rheol Acta 17(2):149–155CrossRef
Zurück zum Zitat Khanna Y (1993) Rheological mechanism and overview of nucleated crystallization kinetics. Macromolecules 26(14):3639–3643CrossRef Khanna Y (1993) Rheological mechanism and overview of nucleated crystallization kinetics. Macromolecules 26(14):3639–3643CrossRef
Zurück zum Zitat Kiewiet S, Janssens V, Miltner H, Assche G, van Puyvelde P, van Mele B (2008) RheoDSC: a hyphenated technique for the simultaneous measurement of calorimetric and rheological evolutions. Rev Sci Instrum 79(2 Pt 1):023905CrossRef Kiewiet S, Janssens V, Miltner H, Assche G, van Puyvelde P, van Mele B (2008) RheoDSC: a hyphenated technique for the simultaneous measurement of calorimetric and rheological evolutions. Rev Sci Instrum 79(2 Pt 1):023905CrossRef
Zurück zum Zitat Kitano T, Katoaka T, Shirota T (1981) An empirical equation of the relative viscosity of polymer melts filled with various inorganic fillers. Rheol Acta 20(2):207–209CrossRef Kitano T, Katoaka T, Shirota T (1981) An empirical equation of the relative viscosity of polymer melts filled with various inorganic fillers. Rheol Acta 20(2):207–209CrossRef
Zurück zum Zitat Krieger I, Dougherty T (1959) A mechanism for non-newtonian flow in suspensions of rigid spheres. Trans Soc Rheol 3:137–152CrossRef Krieger I, Dougherty T (1959) A mechanism for non-newtonian flow in suspensions of rigid spheres. Trans Soc Rheol 3:137–152CrossRef
Zurück zum Zitat Lamberti G, Peters GWM, Titomanlio G (2007) Crystallinity and linear rheological properties of polymers. Int Polym Process 22(3):303–310 Lamberti G, Peters GWM, Titomanlio G (2007) Crystallinity and linear rheological properties of polymers. Int Polym Process 22(3):303–310
Zurück zum Zitat Mooney M (1951) The viscosity of a concentrated suspension of spherical particles. J Colloid Sci 6(2):162–170CrossRef Mooney M (1951) The viscosity of a concentrated suspension of spherical particles. J Colloid Sci 6(2):162–170CrossRef
Zurück zum Zitat Pantani R, Speranza V, Titomanlio G (2001) Relevance of mold-induced thermal boundary conditions and cavity deformation in the simulation of injection molding. Polym Eng Sci 41(11):2022–2035CrossRef Pantani R, Speranza V, Titomanlio G (2001) Relevance of mold-induced thermal boundary conditions and cavity deformation in the simulation of injection molding. Polym Eng Sci 41(11):2022–2035CrossRef
Zurück zum Zitat Peters G, Swartjes F, Meijer H (2002) A recoverable strain-based model for flow-induced crystallization. Macromol Symp 185:277–292CrossRef Peters G, Swartjes F, Meijer H (2002) A recoverable strain-based model for flow-induced crystallization. Macromol Symp 185:277–292CrossRef
Zurück zum Zitat Pogodina N, Winter H (1998) Polypropylene crystallization as a physical gelation process. Macromolecules 31(23):8164–8172CrossRef Pogodina N, Winter H (1998) Polypropylene crystallization as a physical gelation process. Macromolecules 31(23):8164–8172CrossRef
Zurück zum Zitat Pogodina N, Siddiquee S, Egmond J, Winter H (1999a) Correlation of rheology and light scattering in isotactic polypropylene during early stages of crystallization. Macromolecules 32:1167–1174CrossRef Pogodina N, Siddiquee S, Egmond J, Winter H (1999a) Correlation of rheology and light scattering in isotactic polypropylene during early stages of crystallization. Macromolecules 32:1167–1174CrossRef
Zurück zum Zitat Pogodina N, Winter H, Srinivas S (1999b) Strain effects on physcial gelation of crystallizing isotactic polypropylene. J Polym Sci, B, Polym Phys 37:3512–3519CrossRef Pogodina N, Winter H, Srinivas S (1999b) Strain effects on physcial gelation of crystallizing isotactic polypropylene. J Polym Sci, B, Polym Phys 37:3512–3519CrossRef
Zurück zum Zitat Shimizu J, Okui N, Kikutani T (1985) Chapter 15. Fine structure and physical properties of fibers melt-spun at high speeds from various polymers. In: Ziabicki A, Kawai H (eds) High speed fibre spinning. Wiley, New York Shimizu J, Okui N, Kikutani T (1985) Chapter 15. Fine structure and physical properties of fibers melt-spun at high speeds from various polymers. In: Ziabicki A, Kawai H (eds) High speed fibre spinning. Wiley, New York
Zurück zum Zitat Steenbakkers R, Peters G (2008) Suspension-based rheological modeling of crystallizing polymer melts. Rheol Acta 47(5–6):643–665CrossRef Steenbakkers R, Peters G (2008) Suspension-based rheological modeling of crystallizing polymer melts. Rheol Acta 47(5–6):643–665CrossRef
Zurück zum Zitat Steenbakkers R, Peters G (2010) A stretch-based model for flow- enhanced nucleation of polymer melts. J Rheol 55(2):401–433CrossRef Steenbakkers R, Peters G (2010) A stretch-based model for flow- enhanced nucleation of polymer melts. J Rheol 55(2):401–433CrossRef
Zurück zum Zitat Swartjes F, Peters G, Rastogi S (2003) Stress induced crystallization in elongational flow. Int Polym Process 18:53–66 Swartjes F, Peters G, Rastogi S (2003) Stress induced crystallization in elongational flow. Int Polym Process 18:53–66
Zurück zum Zitat Tanner R (2003) On the flow of crystallizing polymers. I. Linear regime. J Non-Newton Fluid Mech 112(2–3):251–268CrossRef Tanner R (2003) On the flow of crystallizing polymers. I. Linear regime. J Non-Newton Fluid Mech 112(2–3):251–268CrossRef
Zurück zum Zitat Timoshenko S, Goodier J (1951) Theory of elasticity. McGraw-Hill Book Company, Inc Timoshenko S, Goodier J (1951) Theory of elasticity. McGraw-Hill Book Company, Inc
Zurück zum Zitat Titomanlio G, Speranza V, Brucato V (1997) On the simulation of thermoplastic injection moulding process. 2. Relevance of interaction between flow and crystallization. Int Polym Process 12(1):45–53 Titomanlio G, Speranza V, Brucato V (1997) On the simulation of thermoplastic injection moulding process. 2. Relevance of interaction between flow and crystallization. Int Polym Process 12(1):45–53
Zurück zum Zitat Winter H, Mours M (1997) Rheology of polymers near liquid-solid transitions. In: Neutron spin echo spectroscopy, viscoelasticity, rheology, vol 134, pp 165–234 Winter H, Mours M (1997) Rheology of polymers near liquid-solid transitions. In: Neutron spin echo spectroscopy, viscoelasticity, rheology, vol 134, pp 165–234
Zurück zum Zitat Yarin A (1992) Flow-induced online crystallization of rodlike molecules in fibre spinning. J Appl Polym Sci 46(5):873–878CrossRef Yarin A (1992) Flow-induced online crystallization of rodlike molecules in fibre spinning. J Appl Polym Sci 46(5):873–878CrossRef
Zurück zum Zitat Ziabicki A (1988) The mechanisms of neck-like deformation in high-speed melt spinning. 1. Rheological and dynamic factors. J Non-Newton Fluid Mech 30(2–3):141–155CrossRef Ziabicki A (1988) The mechanisms of neck-like deformation in high-speed melt spinning. 1. Rheological and dynamic factors. J Non-Newton Fluid Mech 30(2–3):141–155CrossRef
Zurück zum Zitat Zuidema H, Peters G, Meijer H (2001) Development and validation of a recoverable strain-based model for flow-induced crystallization of polymers. Macromol Theory Simul 10:447–460CrossRef Zuidema H, Peters G, Meijer H (2001) Development and validation of a recoverable strain-based model for flow-induced crystallization of polymers. Macromol Theory Simul 10:447–460CrossRef
Metadaten
Titel
Suspension-like hardening behavior of HDPE and time-hardening superposition
verfasst von
Peter C. Roozemond
Vincent Janssens
Peter Van Puyvelde
Gerrit W. M. Peters
Publikationsdatum
01.02.2012
Verlag
Springer-Verlag
Erschienen in
Rheologica Acta / Ausgabe 2/2012
Print ISSN: 0035-4511
Elektronische ISSN: 1435-1528
DOI
https://doi.org/10.1007/s00397-011-0570-1

Weitere Artikel der Ausgabe 2/2012

Rheologica Acta 2/2012 Zur Ausgabe

    Marktübersichten

    Die im Laufe eines Jahres in der „adhäsion“ veröffentlichten Marktübersichten helfen Anwendern verschiedenster Branchen, sich einen gezielten Überblick über Lieferantenangebote zu verschaffen.