• No results found

Flow induced crystallization of isotactic polypropylenes

N/A
N/A
Protected

Academic year: 2021

Share "Flow induced crystallization of isotactic polypropylenes"

Copied!
178
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

Flow induced crystallization of isotactic polypropylenes

Citation for published version (APA):

Housmans, J. W. (2008). Flow induced crystallization of isotactic polypropylenes. Technische Universiteit Eindhoven. https://doi.org/10.6100/IR639130

DOI:

10.6100/IR639130

Document status and date: Published: 01/01/2008 Document Version:

Publisher’s PDF, also known as Version of Record (includes final page, issue and volume numbers) Please check the document version of this publication:

• A submitted manuscript is the version of the article upon submission and before peer-review. There can be important differences between the submitted version and the official published version of record. People interested in the research are advised to contact the author for the final version of the publication, or visit the DOI to the publisher's website.

• The final author version and the galley proof are versions of the publication after peer review.

• The final published version features the final layout of the paper including the volume, issue and page numbers.

Link to publication

General rights

Copyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright owners and it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights. • Users may download and print one copy of any publication from the public portal for the purpose of private study or research. • You may not further distribute the material or use it for any profit-making activity or commercial gain

• You may freely distribute the URL identifying the publication in the public portal.

If the publication is distributed under the terms of Article 25fa of the Dutch Copyright Act, indicated by the “Taverne” license above, please follow below link for the End User Agreement:

www.tue.nl/taverne

Take down policy

If you believe that this document breaches copyright please contact us at:

openaccess@tue.nl

providing details and we will investigate your claim.

(2)

Flow induced crystallization

of isotactic polypropylenes

(3)

CIP-DATA LIBRARY TECHNISCHE UNIVERSITEIT EINDHOVEN Housmans, Johannes W.

Flow induced crystallization of isotactic polypropylenes / by Johannes W. Housmans. Eindhoven : Technische Universiteit Eindhoven, 2008.

Proefschrift. - ISBN 978-90-386-1466-3 NUR 971

Subject headings: isotactic polypropylene / semi-crystalline polymers / flow induced crystallization / polymer morphology / nucleating agents / copolymerization / mechanical properties

Reproduction: University Press Facilities, Eindhoven, The Netherlands. Cover design: Sjoerd Cloos (studioCHAOS).

Cover picture: Jeroen Verreijt (Foto Verreijt).

This research forms part of the research programme of the Dutch Polymer Institute (DPI), Technology Area Polyolefins, DPI project #454.

(4)

Flow induced crystallization

of isotactic polypropylenes

PROEFSCHRIFT

ter verkrijging van de graad van doctor aan de Technische Universiteit Eindhoven, op gezag van de Rector Magnificus, prof.dr.ir. C.J. van Duijn, voor een

commissie aangewezen door het College voor Promoties in het openbaar te verdedigen op woensdag 10 december 2008 om 16.00 uur

door

Johannes Wilhelmus Housmans

(5)

Dit proefschrift is goedgekeurd door de promotor: prof.dr.ir. H.E.H. Meijer

Copromotor:

(6)
(7)
(8)

Contents

Summary xi

1 Introduction 1

1.1 Background . . . 2

1.2 Aim of the thesis . . . 4

1.3 Survey of the thesis . . . 4

References . . . 5

2 Flow induced crystallization of propylene/ethylene random copolymers 7 2.1 Introduction . . . 7

2.2 Experimental . . . 9

Materials . . . 9

Differential Scanning Calorimetry (DSC) . . . 9

Rheological properties in the melt state . . . 10

Flow induced crystallization experiments . . . 10

Equilibrium melting point, Tm0 . . . 15

2.3 Results and discussion . . . 17

DSC . . . 17

Rheological properties . . . 18

Flow induced crystallization . . . 20

2.4 Conclusions . . . 26

References . . . 27

3 Saturation of pointlike nuclei and the transition to oriented structures in flow induced crystallization of isotactic polypropylene 31 3.1 Introduction . . . 31

3.2 Experimental . . . 34

Materials . . . 34

Differential Scanning Calorimetry (DSC) . . . 34

Rheological properties in the melt state . . . 34

(9)

viii CONTENTS

Flow induced crystallization experiments . . . 35

3.3 Results and discussion . . . 36

DSC . . . 36

Rheological properties . . . 36

Flow induced crystallization experiments . . . 39

Critical work . . . 44

Determination of number of quiescent pointlike nuclei . . . 45

Number of flow induced nuclei from rheometry . . . 48

3.4 Conclusions . . . 51

References . . . 52

4 Volumetric rheology of polymers: The influence of shear flow, cooling rate and pressure on the specific volume of iPP and P/E random copolymers 55 4.1 Introduction . . . 56

4.2 Experimental . . . 56

Materials . . . 56

Experimental techniques . . . 57

4.3 Results and discussion . . . 59

Influence of cooling rate . . . 59

Influence of shear . . . 64

4.4 Conclusions . . . 68

References . . . 69

5 Dilatometry: A tool to measure the influence of cooling rate and pressure on the phase behavior of nucleated polypropylene 71 5.1 Introduction . . . 72

5.2 Experimental . . . 73

Materials . . . 73

Sample preparation . . . 74

Differential Scanning Calorimetry (DSC) . . . 74

Dilatometry . . . 74

Wide-angle X-ray diffraction . . . 76

5.3 Results and discussion . . . 77

DSC . . . 77

Specific volume . . . 78

WAXD . . . 84

Crystal structure and specific volume . . . 86

5.4 Conclusions . . . 90

(10)

CONTENTS ix

6 A design to study flow induced crystallization in a multipass rheometer 95

6.1 Introduction . . . 95

6.2 Design and instrumentation . . . 100

The multipass rheometer . . . 100

Slit flow geometry . . . 102

6.3 Experimental . . . 104

Materials . . . 104

Rheological properties in the melt state . . . 104

Birefringence . . . 105

Optical microscopy . . . 106

Fourier transform Infrared (FTIR) spectrometry . . . 106

6.4 Results and discussion . . . 107

Rheological properties . . . 107

Thermal performance . . . 107

Birefringence . . . 109

Optical microscopy . . . 110

FTIR . . . 110

Analysis of the experimental results . . . 113

6.5 Conclusions . . . 115

References . . . 116

7 Structure-property relations in molded, nucleated isotactic polypropylene 119 7.1 Introduction . . . 119 7.2 Experimental . . . 121 Materials . . . 121 Sample preparation . . . 122 Capillary rheometer . . . 122 Optical microscopy . . . 123 X-ray diffraction . . . 123 Structural analysis . . . 124 Mechanical properties . . . 127

7.3 Results and discussion . . . 127

Optical microscopy . . . 127

Structural properties from WAXD/SAXS . . . 134

Mechanical properties . . . 145

7.4 Conclusions . . . 146

References . . . 147

8 Conclusions and recommendations 151 8.1 Main conclusions . . . 151

(11)

x CONTENTS

P/E random copolymers . . . 152

Nucleated iPP . . . 153 8.2 Recommendations . . . 154 References . . . 155 Samenvatting 157 Dankwoord 161 Curriculum Vitae 163

(12)

Summary

Semi-crystalline polymers are used in many applications ranging from commodity to high-tech products. The properties of these products strongly depend on the fi-nal morphology, which itself depends on both the molecular properties used and the processing conditions applied during fabrication. In industry, emphasis is placed on materials tailor-made for each individual application. Material modifications are implemented on the molecular or the microscopic level. In the first case, the poly-mer chain itself is changed, e.g. via long chain branching or copolypoly-merization, while in the second case, e.g. particles, or a phase separated second polymer, are added. These modifications change the crystallization kinetics of the polymer, the final mor-phology and thus the resulting properties.

In this thesis the relationship between the molecular structure, material composition and processing conditions, and the morphology, determining the final properties, is investigated. Processing conditions in production processes like injection molding, film blowing and fiber spinning are often extreme; the material is subjected to high pressures, high deformation rates and high cooling rates. In this thesis the complex-ity of the processing conditions is systematically increased from isothermal homoge-neous short-term shear flow in chapters 2 and 3, via non-isothermal homogehomoge-neous shear flow including pressure dependence (chapter 4 and 5) and (non-)isothermal duct flow (chapter 6) to, finally, non-isothermal transient duct flow, which is a model process for injection molding, in chapter 7.

In chapter 2 the influence of co-monomer content and processing conditions on the crystallization kinetics of propylene/ethylene (P/E) random copolymers is studied using DSC and rheometry. The presence of ethylene increases quiescent crystalliza-tion rate at equal nominal undercooling, because both the crystal growth rate, G, and number of nuclei, N, increase. On the other hand, the effect of flow on the kinetics of crystallization decreases with ethylene content. Still, different regimes of flow-induced crystallization are observed, but their size and the position of the transitions between them depend on the ethylene content.

(13)

xii SUMMARY

In chapter 3 the influence of molecular weight and processing conditions on the crys-tallization kinetics of isotactic polypropylene is studied using rheometry. Depending on the molecular weight, a saturation of pointlike nuclei is observed with increasing shear time. The number of pointlike nuclei is derived from the rheometry experi-ments by modeling the system as a suspension. In most cases, the process acceler-ates after sufficient flow time and this change in kinetics is due to the occurrence of fibrillar nucleation resulting in the formation of row structures and/or shishes. The transition from pointlike to threadlike nuclei is marked by a critical amount of work, which decreases with increasing molecular weight.

In chapters 4 and 5 a novel dilatometer has been used to measure the evolution of specific volume at different cooling rates and at elevated pressures under quiescent conditions and under shear for a series of commercial iPP homopolymers, P/E ran-dom copolymers and an iPP nucleated with DMDBS. The transition temperature in-creases with (i) decreasing cooling rate, (ii) increasing pressure, (iii) the application of shear related to molecular weight, polydispersity and the temperature at which shear is applied, (iv) decreasing ethylene content and (v) the addition of a nucleat-ing agent. It is shown that, for the iPP-DMDBS system, the crystallization line in the phase diagram shifts to higher temperatures with increasing pressure and decreas-ing cooldecreas-ing rate, while the shift is independent of the concentration for all conditions applied. For the nucleated iPP the influence on the evolution of specific volume of cooling rate and pressure is related to the final morphology determined from X-ray diffraction. Crystallinity increases with the addition of DMDBS and the increase is independent of pressure.

The dilatometer allows for conditions near realistic processing conditions, but it is only possible to study the morphology ex-situ. In chapter 6 the design and perfor-mance of a flow geometry for the multipass rheometer (MPR) is described, creating an experimental setup to study in-situ and ex-situ structure and morphology de-velopment with a proper control over the processing conditions. Preliminary results demonstrate the possibilities of this device with the focus on the in-situ birefringence measurements of crystallization and the relation with the final morphology.

In chapter 7 the materials used in chapters 2 to 6 are all injection molded using a capillary rheometer as melting device and melt pump to have better control over the thermo-mechanical history. The morphology distribution in and properties of the rectangular strips are investigated using optical microscopy, X-ray diffraction and mechanical testing. Molecular weight (MW), molecular weight distribution (MWD) and addition of ethylene via copolymerization all influence the thickness of the ori-ented shear layer, the crystallinity, the type and amount of crystal phases, and the lamellar thickness. The addition of a nucleating agent (DMDBS), dictates the crystal-lization process and the resulting morphology, and samples with an oriented

(14)

mor-SUMMARY xiii

phology over the full thickness are created without changing other morphological features, by applying a thermal treatment to the melt prior to injection, which is based on the specific phase behavior of the iPP-DMDBS system. The thermally treated samples show a considerable improvement in mechanical properties.

Finally, in chapter 8 the main conclusions of this thesis are outlined together with recommendations for future research.

(15)
(16)

C

HAPTER ONE

Introduction

In the year 2006, 245 million metric tons of plastics were produced worldwide [1] (compared to 1240 million metric tons of steel [2]) and used for a wide range of ap-plications, see Figure 1.1, ranging from commodity products like flexible packag-ing and molded parts to high-tech applications like flexible displays, high stiffness fibers and joint replacements. Half of the plastics produced and consumed is con-stituted by polyolefins, the class of polymers to which polypropylene, polyethylene and polybutene-1 belong, and roughly 1/3 of the polyolefins produced is polypropy-lene [1, 3, 4]. Household 3.5% Furniture 6% Agriculture 2.5% Medical 1% Others 14% Packaging 33% Building and Construction 23.5% Automotive 9% Electrical/ Electronics 7.5%

Figure 1.1:Application segments of plastics in Germany, relative amounts of used materials in 2003 [4].

The properties of products of semi-crystalline polymers like polypropylene strongly depend on the final morphology, which itself depends on the complete history of the material, as depicted in Figure 1.2, which includes the synthesis determining

(17)

2 1 INTRODUCTION

the molecular characteristics and the thermo-mechanical history experienced during processing. Understanding every step from synthesis via processing to the resulting product properties will enable to tune the materials and the manufacturing processes to the product properties required.

chemical composition chain structure molecular weight and distribution reactor catalyst material formulation additives thermal history history processing mechanical crystallization properties

Figure 1.2:Flow chart illustrating the process-properties relationship for semi-crystalline polymeric materials.

This thesis studies part of this process, starting with different samples of the poly-olefin class, mainly different polypropylenes, and focuses on the influence of pro-cessing on morphology and morphology on resulting properties.

1.1 Background

Polyethylene and polypropylene were produced for the first time in the 1930s and 1950s, respectively. While in the 70s it was still assumed that the new ‘high perfor-mance polymers’ would gain an increasing share of the total polymer market, this was never realized, mainly because especially polypropylene was improved contin-uously causing that the dominant market position of standard thermoplastics was consolidated and expanded [4, 5]. In different application areas, polypropylene re-placed even the so-called high performance polymers often on price-performance. This is the reason why the relation between crystallization behavior of polypropy-lene, and in particular of isotactic polypropylene (iPP), has been studied intensively. To control processing parameters, the short-time shearing protocol as introduced by Janeschitz-Kriegl and co-workers [6] is frequently applied in different experimen-tal set-ups. Devices and techniques used include Differential Scanning Calorime-try (DSC) [7–9], rotational plate-plate devices (rheometers [9–14] and the Linkam

(18)

1.1 BACKGROUND 3

shear cell [9, 11, 15, 16]), parallel plate devices [17–21], fiber pull-out [22–26] and die extrusion set-ups [6, 27]. Only from DSC and rheometry the phase change is mea-sured directly in terms of heat flow and torque, respectively. All other set-ups need an extra measuring technique to follow the structure development in the material. Techniques applied for that part are Optical Microscopy (OM), turbidity and bire-fringence measurements, Small Angle Light Scattering (SALS), Small Angle X-ray Scattering (SAXS) and Wide Angle X-ray Diffraction (WAXD).

Different structures are formed during processing depending on the conditions ap-plied and often the final morphology of a product consists of a combination of these structures, see Figure 1.3. Crystalline structures start to grow from small objects called nuclei, which have a certain degree of order. In quiescent conditions crystals grow in all directions forming spherical shaped crystallites called spherulites. When flow is applied during processing the number of nuclei increases and, for strong enough flows, the nuclei exhibit an anisotropic morphology inducing the forma-tion of ‘shish-kebabs’, crystallites consisting of a fibrillar core (shish) overgrown by a stack of lamellae (kebabs). Different morphologies exhibit different properties and, depending on the application, a certain morphology (or combination) is favored.

Figure 1.3:Cross-section of an injection molded iPP sample (left). The enlargements illustrate the structures present in the oriented shear layer, called shish-kebabs (right, top, reproduced with permission from Figure 3 of [28]) and in the isotropic core, called spherulites (right, bottom [29]).

In industry, emphasis is placed on materials tailor-made for each individual appli-cation. Material modifications are implemented on the molecular or the microscopic level. In the first case, the polymer chain itself is changed, e.g. via long chain branch-ing or copolymerization, while in the second case, e.g. particles, or a phase separated second polymer, are added. These modifications change the crystallization kinetics of the polymer, the final morphology and thus the resulting properties.

(19)

4 1 INTRODUCTION

1.2 Aim of the thesis

The aim of this thesis is to investigate the relationship between the molecular struc-ture, material composition and processing conditions, and the morphology, deter-mining the final properties. For that purpose different experimental techniques will be used to realize different flow conditions, to monitor crystallization kinetics and to characterize the final morphology and mechanical properties. The experimen-tal techniques used comprise rheometry, DSC, dilatometry, the multipass and capil-lary rheometer, birefringence, optical microscopy, FTIR, X-ray diffraction and tensile testing. Since processing conditions in production processes like injection molding, film blowing and fiber spinning are often extreme, and the conditions applied in most of the experimental techniques in earlier studies are still far from realistic pro-cessing conditions, another objective is to develop an experimental setup with the ability to study in-situ crystallization of semi-crystalline polymers and the resulting structure and morphology (in-situ and ex-situ) with a good control over the thermo-mechanical history applied, which is in the range of processing conditions as found in injection molding or extrusion.

1.3 Survey of the thesis

In chapters 2 and 3, the influence of co-monomer content, molecular weight and processing conditions on the crystallization kinetics of isotactic polyprolylenes and propylene/ethylene (P/E) random copolymers is studied using DSC and rheometry. Different regimes of flow-induced crystallization and the position of the transitions between them are related to the rheology of the materials, the ethylene content and flow characteristics. In chapter 3, existing knowledge on FIC is applied to derive morphological information from the results. Chapters 4 and 5 discuss in depth the influence of cooling rate, pressure and shear flow on the evolution of specific volume for a series of commercial iPP homopolymers, P/E random copolymers and an iPP nucleated with DMDBS. The morphology is analyzed using optical light microscopy and X-ray diffraction. In chapter 6, the design and first testing of a flow geometry for the multipass rheometer is tested, which enables to study in-situ flow induced crys-tallization. Preliminary results demonstrate the possibilities of this device. Chapter 7 deals with the relation between processing, morphology distribution and mechan-ical properties in injection molded samples. Smart processing, based on the phase behavior of the iPP-DMDBS system, is applied in an attempt to create fully oriented samples and to improve properties. Finally, chapter 8 summarizes the most impor-tant conclusions of this thesis and gives recommendations for future research.

(20)

REFERENCES 5

References

[1] Plastics Europe, association of plastics manufacturers (2007). The Compelling Facts about Plastics - An analysis of plastics production, demand and recovery for 2006 in Europe. www.plasticseurope.org.

[2] World Steel Association (2008). 2008 Sustainability report of the world steel industry. www.worldsteel.org.

[3] Younkin, T.R., Connor, E.F., Henderson, J.I., Friedrich, S.K., Grubbs, R.H., Bansleben, B.A. (2000). Science, 287, 460-462.

[4] Gahleitner, M., Paulik, C. (2007). Polyolefin Basics, Borealis GmbH, RPOD Linz, Austria.

[5] Warzelhan, V., Brandstetter, F. (2003). Macromolecular Symposia, 201, 291-300.

[6] Liedauer, S., Eder, G., Janeschitz-Kriegl, H., Jerschow, P., Geymayer W., In-golic E. (1993). International Polymer Processing VIII, 236-244.

[7] Lamberti, G. (2004). Polymer Bulletin, 52, 443-449.

[8] Lamberti, G., Natteo, C. (2006). Polymer Bulletin, 56, 591-598.

[9] Koscher E., Fulchiron, R. (2002). Polymer, 43, 6931-6942.

[10] Vleeshouwers, S., Meijer H.E.H. (1996). Rheologica Acta, 35, 391-399.

[11] Vega, J.F., Hristova, D.G., Peters, G.W.M. (2008). submitted to Journal of Thermal Analysis and Calorimetry.

[12] Khanna, Y.P. (1993). Macromolecules, 3639-3643.

[13] Pogodina, N.V., Winter, H.H. (1998). Macromolecules, 31, 8164-8172.

[14] Boutahar, K., Carrot, C., Guillet, J. (1996). Journal of Applied Polymer Science, 60, 103-114.

[15] Somani, R.H., Hsiao, B.S., Nogales, A., Srinivas, S., Tsou, A.H., Sics, I., Balta-Calleja, F.J., Ezquerra, T.A. (2000). Macromolecules, 33, 9385-9394.

[16] Baert, J., Van Puyvelde, P. (2006). Polymer, 47, 5871-5879.

[17] Monasse, B. (1995). Journal of Materials Science, 30, 5002-5012.

[18] Tribout, C., Monasse, B., Haudin, J.M. (1996). Colloid & Polymer Science, 274, 197-208.

[19] Janeschitz-Kriegl, H., Ratajski, E., Stadlbauer, M. (2003). Rheologica Acta, 42, 355-364.

[20] Langouche, F. (2006). Macromolecules, 39, 2568-2573.

[21] Baert J., Van Puyvelde, P., Langouche, F. (2006). Macromolecules, 39, 9215-9222.

[22] Thomason, J.L., van Rooyen A.A. (1992). Journal of Materials Science, 1992, 27, 897-907.

[23] Varga, J., Karger-Kocsis, J. (1993). Polymer Bulletin, 30, 105-110.

[24] Varga, J., Karger-Kocsis, J. (1996). Journal of Polymer Science: Part B: Polymer Physics, 34, 657-670.

[25] Monasse, B. (1992). Journal of Materials Science, 27, 6047-6052.

[26] Jay, F., Haudin, J.M., Monasse, B. (1999). Journal of Materials Science, 34, 2089-2102.

(21)

6 1 INTRODUCTION

[27] Kumuraswamy, G., Verma, R.K., Kornfield, J.A. (1999). Review of Scientific In-struments, 70, 2097-2104.

[28] Hsiao, B.S. Yang, L. Somani, R.H. Avila-Orta, C.A., Zhu, L. (2005). Physical Re-view Letters, 94, 117802.

[29] Swartjes, F.H.M. (2001). PhD thesis, Eindhoven University of Technology, The Netherlands.

(22)

C

HAPTER TWO

Flow induced crystallization of

propylene/ethylene random

copolymers

1

The influence of co-monomer content and processing conditions on the crystalliza-tion kinetics of propylene/ethylene (P/E) random copolymers is studied using DSC and rheometry. The presence of ethylene lowers the melting and crystallization tem-perature compared to pure polypropylene, and the quiescent crystallization rate, ˙χ, increases at equal nominal undercooling, because both the crystal growth rate, G, and number of nuclei, N, increase. The effect of flow on the kinetics of crystallization decreases with ethylene content. Still, different regimes of flow induced crystalliza-tion are observed, but their size and the posicrystalliza-tion of the transicrystalliza-tions between them depend on the ethylene content, and can be expressed in terms of the level of molec-ular orientation, molecmolec-ular stretch and crystallization capacity of the system.

2.1 Introduction

Properties of polymer products depend on the morphology distribution within the product, which itself depends on both the molecular properties of the polymer used and the processing conditions applied during fabrication. Not only mechanical prop-erties depend on crystal structure [1–3], but also propprop-erties, like dimensional stabil-ity [4] and transparency [5]. In production processes like injection molding, film blowing and fiber spinning the polymer is subjected to high pressures and speeds,

1Reproduced from: J.W. Housmans, G.W.M. Peters and H.E.H. Meijer, Flow induced

crystalliza-tion of propylene/ethylene random copolymers, Journal of Thermal Analysis and Calorimetry, submitted, (2008)

(23)

8 2 FLOW INDUCED CRYSTALLIZATION OF P/ERANDOM COPOLYMERS

and it is cooled from temperatures well above its melting point to room temperature within seconds. Due to the complex geometries of the molds and extruder dies, the flow history has a complex nature; the material experiences both shear and elonga-tional flow. The effect of shear deformation on the crystallization kinetics and struc-ture development of semi-crystalline polymers has gained a lot of interest in the last decade, see for example [6–14] and references therein. It is now well known that the flow field applied accelerates the crystallization process and alters the morphology; the effect of shear rate is more pronounced than that of the shear time. The im-portance of the high molecular weight tail on the formation of oriented, anisotropic structures was stressed [6, 11, 12].

Van Meerveld et al. [15] proposed a rheological classification of different flow regimes, i.e. regimes leading to different types of crystalline morphologies, derived from molecular based rheology and rubber elasticity theory. The transitions between the different flow regimes, and the associated physical processes governing the flow induced crystallization (FIC) processes, are defined by critical values of the Deborah numbers (De) related to molecular orientation (Derep) and molecular stretch (Des). An

extensive evaluation of experiments reported in the literature illustrates that the tran-sition, from just an enhanced nucleation rate of spherulites towards the development of fibrillar structures, correlates with the transition from chain segment orientation to chain stretch applied at least to the high molecular weight chains in the melt. Polymers are often modified to improve their performance. The changes are imple-mented on (a) a molecular level and/or (b) a microscopic level. In the first case, the polymer chain itself is modified or another material is added and dissolved, while in the latter situation, particles, or a phase separated second polymer, are added. These alterations to the material have a strong influence on the polymer’s FIC behavior. Several researchers studied case (b), investigating experimentally the particle-morphology-properties relationship [2, 16–20]. The presence of particles lead to strong molecular orientation throughout the complete sample caused by a ‘shear amplification effect’ between the particles and sometimes an improved impact re-sistance. Numerical work on particle filled viscoelastic systems showed that the presence of particles results in the occurrence of regions of high molecular stretch between layers of particles and regions shielded by particles inside the layers [21] giving rise to anisotropic flow induced structures. Another additive is the family of nucleating agents and they are often blended with polymers to decrease conversion cycle times [22]. Crystallization on heterogeneous nuclei leads to the formation of smaller crystal structures which improves the optical and mechanical properties of the materials. An example of such an additive is based on sorbitol (DMDBS) which already in relatively small amounts (∼ 0.2 wt%) enhances the clarity and give rise to an increase in yield stress of isotactic polypropylene (iPP) [5, 23]. The effect of flow on crystallization of iPP/DMDBS blends was studied by Balzano et al. [24, 25]. Recently, even much more effective and selective nucleating agents based on 1,3,5-benzene tricarboxylic acid were developed with which the same effects are obtained for concentrations as low as 0.02 wt% [26, 27]. This family of organic compounds

(24)

2.2 EXPERIMENTAL 9

is thermally stable and some members form the iPPβ-polymorph which exhibits a higher impact toughness than the most commonαform [28].

Examples of chain modifications (case (a)) are long chain branching (J.F. Vega, per-sonal communication) and copolymerization with other monomers creating e.g. ran-dom or block-copolymers.

In this experimental study the focus is on the effect of random copolymerization of PP with ethylene on FIC. Copolymerization of polypropylene with a small amount of ethylene is an appropriate way to improve both the optical and certain mechanical properties (low-temperature brittleness) [29–31]. First, the four grades used are char-acterized using differential scanning calorimetry (DSC) and rheometry. Next short term shear experiments are performed at equal nominal undercooling and the flow effects are quantified with a characteristic time scale. Finally, the effect of copoly-merization on the rheological classification is investigated. Different regimes are ob-served and the transitions between regimes are specific for every polymer.

2.2 Experimental

Materials

Materials used are an isotactic polypropylene (iPP, HD234CF, Borealis) and three propylene/ethylene (P/E) random copolymers (P/E RACO, RD204CF, RD226CF and RD208CF, Borealis) with different ethylene contents. All four grades have a weight average molecular weight Mw310 kg/mol and a polydispersity Mw/Mn

3.4 [29]. Molecular and other physical properties are listed in Table 2.1. RACO2 dif-fers from the iPP and the other two RACO’s in that it contains some additives not present in the others like a slip agent to reduce the coefficient of friction (for post-processing steps in film application) and a synthetic silica anti-blocking aid to re-duce the adhesion between film layers [32, 33]. The anti-blocking aid consists of both spherical and plate-like particles ranging in size between 100 nm up to 10µm. We kept this material in our series of RACO’s because it shows that slight modifications of the composition of a polymer (case (b), 2.1), with no influence on the rheological behavior and expected behavior for quiescent crystallization, can have a large effect on the FIC behavior.

Differential Scanning Calorimetry (DSC)

The melting temperature (Tm), melting enthalpy (∆Hm) and crystallization

tempera-ture (Tc) were determined on a Mettler Toledo DSC (DSC823e) using a heating and

cooling rate of 10◦C min−1 on samples with a weight of 5 ±0.5 mg. A second

heat-ing cycle is performed to have a homogeneous sample distribution in the pan from which Tmand ∆Hmare obtained. Crystallinity is determined using Xc =∆Hm/∆Hm0

(25)

10 2 FLOW INDUCED CRYSTALLIZATION OF P/ERANDOM COPOLYMERS

Table 2.1: Basic characteristics of the material grades from literature [29]. Tm, Xcand

Tcare determined on the provided materials. The Tmvalues between

brack-ets are explained in section 2.3.

Name Grade Ethylene content Tm[29] Xc[29] Tm Xc Tc

FTIR [wt%] NMR [mol%] [◦C] [%] [◦C] [%] [◦C]

PP HD234CF 0 0 164 49.5 159 (163) 48.3 110

RACO1 RD204CF 2.2 3.4 153 41.3 147 (153) 42.3 105

RACO2 RD226CF 3.4 5.2 145 36.3 140 (146) 36.8 99

RACO3 RD208CF 4.9 7.3 139 33.9 138 (145) 34.4 98

Rheological properties in the melt state

A Rheometrics ARES rheometer was used with a plate-plate geometry, diameter 25 mm, for small amplitude oscillatory shear measurements. The characteristic rheolog-ical properties (storage and loss modulus, G’ and G”, and loss angle,δ) were obtained over a broad range of temperatures (from 145◦C to 250C) and angular frequencies,

ω(from 0.01 to 100 rad s−1). The lower limit ofωfor the lowest temperature was 0.1

rad s−1to avoid changes due to crystallization. To determine the rheological

proper-ties in the linear viscoelastic regime, the strain applied was determined from strain sweeps and set to 5 % for all measurements. The experiments were performed in a nitrogen environment to avoid degradation of the material. Time-temperature su-perposition has been applied to obtain mastercurves at a reference temperature of 220◦C.

Flow induced crystallization experiments

The crystallization kinetics was followed using rheological experiments. A relatively small plate-plate geometry of 8 mm was used to avoid transducer instabilities caused by stiffening of the material. Isothermal temperatures were chosen such that the undercooling, which is the driving force for crystallization, was equal for all four materials. Undercooling is normally defined as ∆T = Tm0Texp, with Tm0 the

equi-librium melting temperature and Texp the isothermal crystallization temperature at

which the experiment is performed. For the polymer grades used, Tm0is not known

and the determination of Tm0 is a lengthy and precise task, as shortly discussed in

the next section. Therefore, we define undercooling as ∆T = TmTexp, where Tm

is the nominal melting temperature determined from DSC, see Table 2.1. For the homopolymer Texp was set to 138◦C and for the P/E RACO the temperature was

determined accordingly, i.e. to get the same ∆T. The experimental procedure is as follows:

• Samples were molten at a temperature of 230◦C for 10 minutes to remove any history before the start of the test.

(26)

2.2 EXPERIMENTAL 11

• Subsequently, the samples were cooled to the desired crystallization tempera-ture with a controlled cooling rate of 15◦C min−1to avoid any undershoot. The

gap was adjusted continuously to compensate for thermal expansion (shrink-age) of the samples.

• Oscillatory tests using an angular frequency of 5 rad s−1 and a strain of 0.5 %

were performed to monitor the evolution of G, G” andδin time.

For the short-term FIC experiments, a shear flow for a certain shear time, ts was

applied prior to the third step. The shear rate values, ˙γ, varied between 3 and 60 s−1

and the total strain,γ, was kept constant at 60 strain units.

Figure 2.1 shows flow induced crystallization experiments on an iPP grade (Borealis HD120MO, J.F. Vega, personal communication, used in [35]) that follow the above-mentioned procedure (Texp = 135◦C). The figure is illustrative for the results

ex-pected. The storage modulus, G′(φ), evolves with the growing crystallites in the melt, since it is a function of degree of space filling,φ. When flow is applied, the two main observations are:

1. Curve 2: An acceleration of the crystallization process, the modulus curve shifts to lower times. The shape is similar to that of the quiescent experiment.

2. Curve 3: For longer flow times, the modulus rise shifts to lower times, but now its shape changes and the slope decreases.

The kinetics of degree of space filling is usually described by the Avrami equation [36, 37]:

φ =1−exp[−Ctn] (2.1)

where C is the overall rate of crystallization and n the Avrami exponent, which re-flects the crystal dimensionality and the type of nucleation (pre-determined or spo-radic) [38]. When nucleation is heterogeneous and the crystal growth rate, G, is constant, the n values of 1, 2, and 3 correspond to a crystal geometry of rods (1-dimensional, 1D), disks (2-(1-dimensional, 2D) and spheres (3-(1-dimensional, 3D), re-spectively. For sporadic nucleation, the n values of 2, 3 and 4 correspond to the crystal geometry of rods, disks and spheres [37, 39]. For the change of the modulus it can be derived, that:

dGdt = dG′ dφ dφ dt (2.2)

in which dφ/dt is given by (using eq. 2.1):

dt =Cnt

n−1 exp[−Ctn] (2.3)

From eqs. 2.1 and 2.3 it can be derived, that, to achieve a shift over the time axis to lower times, without changing the shape of the curve (curve 2), the Avrami exponent,

(27)

12 2 FLOW INDUCED CRYSTALLIZATION OF P/ERANDOM COPOLYMERS 100 10 10 10 104 104 105 106 107 108 time [s] G’ [Pa] 1 1 2 2 3 3 (3) (2) (1)

Figure 2.1:Evolution of the storage modulus during crystallization of iPP HD120MO at Texp = 135◦C, measured under quiescent conditions (◦, 1) and after

shearing at ˙γ =60s−1for ts =1s (, 2) and ts= 6s (△, 3). Optical

micro-graphs indicate the characteristic morphology for the three crystallization experiments [40].

n, has to be the same, and only the overall rate of crystallization, C, increases. Hence, for a shift and a change in slope (curve 3), C increases and n decreases. Under quies-cent conditions in the temperature range applied, no sporadic nucleation occurs and the number of nuclei, N, is constant. Crystal lamellae grow with a constant growth rate, G, in all directions, thus a 3D growth, from these nuclei forming spherulites (Figure 2.1, micrograph 1 [40]). In that case, n =3 and C =4/3πNG3[37, 38, 41, 42].

When flow, applied for a short time, only leads to a shift in modulus (curve 2), only C increases and n stays the same and we expect no sporadic nucleation. This implies, that the same type of morphology is formed as in the quiescent case (spherulites), but with a higher number density, N, since C increases and the growth rate G is

(28)

un-2.2 EXPERIMENTAL 13

altered [42] (Figure 2.1, micrograph 2 [40]). For longer flow times, where a change in the slope is observed (curve 3), n decreases. Oriented, shish-kebabs are formed by crystalline lamellae growing in two dimensions (2D) off fibrillar nuclei (Figure 2.1, micrograph 3 [40]), for which it can be derived, that n =2 and C = πNLG2, with L,

the total length of the shishes [37, 41]. Also in this situation, no sporadic nucleation is expected. Of course, not only fibrillar structures are formed, spherulites will grow as well and thus 2 < n < 3. For the experiments shown in Figure 2.1 the result-ing morphologies were confirmed usresult-ing optical microscopy (OM) and wide-angle X-ray diffraction (WAXD) (D.G. Hristova, personal communication, used in [35]). To summarize the above, three morphology types are distinguished:

1. Quiescent crystallization: Spherulites are formed by crystalline lamellae grow-ing in three dimensions off point-like nuclei, whose number density depends on the temperature.

2. Shift of modulus curve: The number density of point-like nuclei increases with the strain rate. This leads to a more fine grained but still spherulitic morphol-ogy.

3. Shift and shape change: Shish-kebabs are formed by crystalline lamellae grow-ing in two dimensions off fibrillar nuclei, whose number density and length increases with strain rate.

Four flow regimes, based on the rheology of the polymer, are defined by van Meerveld et al. [15]. The transitions between these regimes are defined by critical values of the Deborah number, De (τ/t), or Weissenberg number, Wi (=γτ˙ ), related to molecular orientation and molecular stretch. Both orientation and stretch are char-acterized by a time scale: the reptation time, τrep, and the stretch relaxation time or

Rouse time,τs, respectively. In regime I, for Derepand Des <1, the chains are at

equi-librium. The transition to regime II corresponds to orientation of the contour path (Derep >1, Des<1). The onset of chain stretching (Derep >1, Des >1) marks regime

III and in the fourth and final regime, the chains are strongly stretched and deviate from the Gaussian configuration caused by rotational isomerization. The analysis of FIC experiments reported in literature indicated that the high molecular weight (HMW) chains dictate the FIC dynamics [6, 12]. Furthermore, the number density of spherulites increases in regime II (Derep > 1, Des < 1−10) and the shish-kebab

morphology develops for Des >1−10 based onτsof the longest chains.

Character-ization of the melt provides the relaxation time spectrum, i.e.τrep, from whichτscan be estimated:

τs =

τrep

3Z (2.4)

with Z = Mw/Me, the number of entanglements per chain and Me the weight

aver-age molecular weight between entanglements, being 5200 g/mol for iPP [43].

Although the classification gives a clear indication in which flow regime experiments are performed and whether fibrillar morphologies will develop or not, the flow time

(29)

14 2 FLOW INDUCED CRYSTALLIZATION OF P/ERANDOM COPOLYMERS

is not specifically included. It is stated that the flow time, ts, has to be sufficiently

long in order to fulfill the conditionλ > λ∗(T), i.e. the molecular stretch ratioλhas

to be larger than a certain critical valueλ∗(T), marking the transition between weak and strong stretching conditions. However, it is expected that this critical value has to be maintained long enough to cause significant shish growth.

To clarify the relation between the flow regimes and the morphology types and to show the effect of shear rate ( ˙γ), shear time (ts) and shear strain (γ = γ˙ ts) on the

crystallization behavior of the polymers, a schematic picture is presented, see Fig-ure 2.2. A common way to characterize crystallization kinetics is to define a time, t, characteristic for the process, e.g. the induction time (ti) or the half time of

crystal-lization (t1/2), and to normalize this with the characteristic time of the quiescent case

(tQ) [44, 45]. According to the classification, the transitions from flow regime I to II

and flow regime II to III are marked by Derep = 1 and Des = 1, corresponding to

a shear rate value of 1/τrep and 1/τs, respectively. These transitions are included in

Figure 2.2 as the two vertical dashed lines. First, let’s consider FIC experiments with a constant level of applied strain (bold line, Figure 2.2). At low values of the shear rate, flow does not influence the crystallization behavior, i.e. the material crystallizes as in quiescent conditions, represented by the horizontal line at t/tQ = 1

(morphol-ogy type 1). For higher values of ˙γ, the crystallization process is accelerated and t/tQ

becomes smaller than 1. Shear rate is more effective than shear time in accelerating polymer crystallization and, hence, for increasing shear rates t/tQ decreases. The

shape (slope) of the modulus curve versus time is the same compared to the qui-escent experiment, which means that the crystallization kinetics are unaltered. The decrease in crystallization time is associated with an increase of the number of point-like nuclei (morphology type 2). With a further increase of ˙γ, a change in the slope of the G′-evolution curve is observed, indicative for a change in the crystallization ki-netics, and the crystallization process is even faster. It is related to the transition from isotropic spherulitic crystals to oriented shish-kebab structures (morphology type 3). Consequently, the slope of t/tQbecomes steeper. As can be seen in the schematic

pic-ture, the transitions in t/tQ fall within flow regime II and flow regime III. However,

they do not necessarily coincide with the shear rate values of 1/τrepand 1/τs, i.e. the

transitions from flow regime I to II and flow regime II to III, respectively. When the shear strain level is increased, the two transition points in the t/tQ curve move to

lower shear rates, eventually coinciding with the transitions between the different flow regimes. This shows the importance of flow time: flow has to be applied for a sufficiently long time in order to orient or stretch the polymer chains enough to observe an effect on the crystallization behavior and final morphology, which is in line with the conclusions of Van Meerveld et al. [15]. The transition that still needs explanation is from flow regime III to IV, the ’weak’ and ’strong’ stretching regime, respectively. In flow regime III the chains are stretched, but still keep their Gaussian configuration. For λ > λ∗(T) the chain configuration becomes non-Gaussian and

the amount of rotational isomerization (RI) is large (regime IV). Based on the anal-ysis of FIC experiments reported in literature it was concluded that the shish-kebab morphology develops when the HMW chains are subjected to the condition Des >1

(30)

2.2 EXPERIMENTAL 15 t/tQ γ 1/τrep 1/τs ts ˙ γ I II III IV Derep Des 1 1 1 Spherulites Fine spherulites Oriented structures (shish-kebabs) λ > λ∗(T)

Figure 2.2:Schematic representation of the flow regimes and morphology types. See the text for explanation.

for a sufficiently long time in order to fulfill the condition λ > λ∗(T) [15]. Hence,

in the schematic picture (Figure 2.2) the transition from flow regime III to IV occurs for t/tQ when the second change in slope is observed i.e. it coincides with the

on-set of morphology type 3. Here, the evolution of G′ shows a change in slope, i.e. a change in the crystallization kinetics, and the shish-kebab crystal structure forms. In the schematic picture, the direction of the line that marks the transition from III to IV is arbitrarily chosen.

Equilibrium melting point,

T

m0

Usually, two extrapolation methods are used to determine the equilibrium melting point, Tm0, the Gibbs-Thomson (GT) [46] and the Hoffman-Weeks (HW) [47] method.

As a result of a finite crystal thickness, l, Tm is depressed below that of an infinite size

crystal which is expressed in the GT equation: Tm =Tm0  1− 2σe l∆Hf  (2.5) whereσe is the surface free energy of the crystal plane normal to the thickness di-rection and ∆Hf the heat of fusion. From Eq. (2.5) it can be seen that Tm0 can be

(31)

16 2 FLOW INDUCED CRYSTALLIZATION OF P/ERANDOM COPOLYMERS

estimated by extrapolating Tm versus 1/l data to infinite thickness. This method

re-quires X-ray scattering techniques to determine the thickness of the crystals and a method (e.g. DSC) to measure the melting point of these lamellae.

The HW procedure correlates directly the temperature of crystallization Tc to the

melting temperature Tm of the structures formed during subsequent melting

(Equa-tion 2.6). Tm =Tm0  1− 1 γ  + Tc γ (2.6)

The parameterγ = l/li accounts for isothermal lamellar thickening after formation

of an initial crystal with thickness li. Thicker crystals are usually formed when a

polymer crystallizes at higher temperatures, resulting in a higher Tm. Tm0 can be

determined by linear extrapolation of Tm as a function of Tc to the equilibrium line

Tm = Tc. This method only requires a DSC device which makes it easy to obtain the

necessary data.

In both methods lamellar thickening plays an important role. When the lamel-lar thickness for GT is determined at room temperature, an increase of the crystal height can occur during melting. A correct extrapolation is only possible for the HW method when the thickening coefficient is independent of the crystallization tem-perature, i.e. only samples crystallized at different temperatures, but with the same value forγ, can be used to correctly determine Tm0[48].

Both methods were applied by Mezghani et al. [49] to evaluate the discrepancies be-tween two groups of Tm0 values reported for iPP, one group around 187◦C and the

other around 210◦C. They found for the GT method an equilibrium melting

tem-perature of around 186◦C. The result determined with the HW method was how-ever greatly influenced by the crystallization time. Where impinged, big spherulites (∼ long crystallization times) gave a value of 210◦C, T

m0 determined from small

spherulites, crystallized for a short time, was around 188◦C. According to Mezghani et al. [49] the most accurate result with the HW procedure would be obtained when Tmof very small, newly initiated spherulites could be measured, because the polymer

continues to crystallize during heating.

In a subsequent paper on P/E RACO’s with similar molecular weights, a significant decrease of Tm0was observed for an increase in ethylene content (∼5◦C/mol%) using

the HW method [50]. In a crystallization study on fractionated P/E RACO’s it was found that the equilibrium melting point was lowered by approximately 2.3◦C/mol% [51]. These observations imply for RACO3 with an ethylene content of 7.3 wt% a significant difference in Tm0, i.e. 150◦C [50] and 170C [51], adapting Tm0 = 187◦C

for the iPP homopolymer. The preference would maybe go to the former, because commercially available polymer grades are used with a comparable molar mass to our materials, where the latter used fractionated samples. The choice for either one would, however, be equally right (or wrong).

(32)

pro-2.3 RESULTS AND DISCUSSION 17

cedure in which errors can be made easily. It is outside the scope of this study to get exact values. The melting temperatures mentioned in Table 2.1 show a depen-dence on ethylene content of 3−3.5◦C/mol% which is of the same order as reported in [50, 51]. Therefore, we will use the nominal undercooling as mentioned in sec-tion 2.2 to determine Tc, exp.

2.3 Results and discussion

DSC

The melting temperature, Tm, the amount of crystallinity, Xc, determined from the

melting enthalpy, and the crystallization temperature,Tc, are given in Table 2.1. The

melting thermograms of all four materials show bimodal melting, see Figure 2.3(a). The Tm values in Table 2.1 are the values of the low temperature peak which is the

most pronounced. The high temperature peak/shoulder values are given in brackets.

50 100 150 200 −1.2 −1 −0.8 −0.6 −0.4 −0.2 0 0.2 Temperature [°C] Heat flow [W/g] 0 2 4 6 8 10 80 100 120 140 160 180

Ethylene content [mol%]

Temperature [

°C]

(a) (b)

Figure 2.3:(a) Melting thermograms of PP (◦), RACO1 (), RACO2 () and RACO3

(♦). (b) Melting () and crystallization () peak temperature as a function

of the ethylene content.

No influence of the additives in RACO2 is observed, i.e. the results nicely fit the trend that with increasing amount of ethylene, Tm, Xc and Tcdecrease. The bimodal

melting behavior for fractionated P/E RACO’s occurs due to recrystallization dur-ing heatdur-ing [51]. At high heatdur-ing rates the meltdur-ing behavior of their fractions was unimodal. A slightly lower melting temperature is observed compared to the val-ues of Gahleitner et al. [29], which is probably due to the use of a different DSC device and/or a different batch of material. The amount of crystallinity of the sam-ples compares well. Figure 2.3(b) shows the decrease of Tm and Tc with ethylene

content, which are approximately 3.0◦C/mol% and 1.8C/mol%, respectively. Laiho-nen et al. [51] reported that both temperatures went down by 3.0C/mol%. The small difference in slope for the crystallization temperature is probably due to the use of fractionated samples in [51].

(33)

18 2 FLOW INDUCED CRYSTALLIZATION OF P/ERANDOM COPOLYMERS

Rheological properties

The rheological properties at a temperature of 220◦C for pure iPP and P/E RACO

with 7.3 mol% are shown in Figure 2.4. All materials show very similar behavior in the linear viscoelastic regime. The horizontal and vertical shift factors are listed in Table 2.2. An Arrhenius temperature dependence is found for the horizontal shift factors (aT). The value of the activation energy, Ea, is (almost) the same for all four

materials. The vertical shift factor, bT, shows no strong temperature dependence,

which is normal behavior for linear polymers. The relaxation time spectrum of the polymers in the melt state (at 220◦C) can be calculated from the rheological properties measured. 10−2 10−1 100 101 102 103 10−1 100 101 102 103 104 105 ω aT [rad/s] G’ b T , G" b T [Pa] 10−2 10−1 100 101 102 103 10−1 100 101 102 103 104 105 ω aT [rad/s] G’ b T , G" b T [Pa] (a) (b)

Figure 2.4:Storage (△) and loss () modulus at Tre f = 220C for (a) the iPP

ho-mopolymer (PP) and (b) the P/E RACO with 7.3wt% ethylene (RACO3). The lines correspond to a 5 mode Maxwell fit to the moduli.

For that purpose, a discrete Maxwell relaxation time spectrum (gii) is used [52]:

G′(ω) =

i gi ω2τi2 1+ ω2τ2 i (2.7) G′′(ω) =

i gi ωτi 1+ ω2τ2 i (2.8) The set of relaxation moduli, gi, and times, τi, are given in Table 2.2 which nicely

fit the measured G’and G” (solid lines in Figure 2.4). In Table 2.3 the values ofτreplong

from the discrete relaxation times spectrum for the 4 materials at 220◦C are reported, together with the corresponding Rouse time,τslong (Eq. 2.4). Also reported here, are the temperature at which the FIC experiments are performed, Texp, the corresponding

aT, calculated from the Arrhenius relation, and the characteristic relaxation times at

the Texp. It is observed that the times at 220◦C are very similar for the 4 material

grades used, which shows that the viscoelastic behavior is not influenced by the incorporation of ethylene units.

(34)

2.3 RESULTS AND DISCUSSION 19

Table 2.2: Rheological parameters at 220◦C for the materials studied: time-temperature shift factors, aT and bT, Arrhenius activation energy, Ea, and

Maxwell relaxation spectra, gi and τi.

iPP HD234CF P/E RACO RD204CF T [C] a T [-] bT [-] T [◦C] aT [-] bT [-] 145 7.50016 0.88649 145 6.98125 0.89018 160 4.54793 0.93154 160 4.00255 0.88957 175 2.91501 0.95363 175 2.70305 0.92252 Shift factors 190 2.11167 0.97134 190 1.79036 0.91121 205 1.42567 1.00486 205 1.29904 0.93807 220 1 1 220 1 1 235 0.74755 1.05951 235 0.76843 0.98564 250 0.66550 1.23371 250 0.58716 1.05056 Arrhenius Ea [kJ/mol] 43.0 - Ea [kJ/mol] 42.04

-Tre f [◦C] 220 - Tre f [◦C] 220

-Mode gi 10−4[Pa] τi[s] Mode gi 10−4[Pa] τi[s]

1 11.08 0.0011 1 10.53 0.0014

Maxwell 2 3.38 0.0085 2 3.21 0.010

modes 3 0.870 0.0448 3 0.817 0.057

4 0.106 0.237 4 0.097 0.316

5 0.004 1.46 5 0.004 2.07

P/E RACO RD226CF P/E RACO RD208CF T [C] aT [-] bT [-] T [C] aT [-] bT [-] 145 7.11353 0.92549 145 6.76522 0.93119 160 4.25546 0.95089 160 4.09919 0.95672 175 2.59461 0.96162 175 2.73183 0.92839 Shift factors 190 1.86385 0.99016 190 1.97125 0.95734 205 1.29624 0.98906 205 1.31445 0.97546 220 1 1 220 1 1 235 0.67175 1.06811 235 0.59987 1.00720 250 0.51710 1.10467 250 0.50280 1.03202 Arrhenius Ea [kJ/mol] 44.70 - Ea [kJ/mol] 45.19

-Tre f [◦C] 220 - Tre f [◦C] 220

-Mode gi 10−4[Pa] τi[s] Mode gi 10−4[Pa] τi[s]

1 11.30 0.0013 1 10.83 0.0015

Maxwell 2 3.51 0.0097 2 3.11 0.013

modes 3 0.931 0.0528 3 0.703 0.073

4 0.116 0.287 4 0.067 0.439

(35)

20 2 FLOW INDUCED CRYSTALLIZATION OF P/ERANDOM COPOLYMERS

Table 2.3: Characteristic time scales at 220◦C for the materials studied: Maxwell

model longest relaxation time, τreplong and Rouse time, τslong (from Eq. 2.4).

Temperature at which the FIC experiments are performed, Texp, the

corre-sponding temperature shift factor, aT and the relaxation times, τreplong(Texp)

and τslong(Texp), resp.

τreplong[s] τslong[s] Texp[◦C] aT τreplong(Texp)[s] τslong(Texp)[s]

PP 1.46 8.11 10−3 138 8.11 11.84 6.58 10−2

RACO1 2.07 1.15 10−2 126 11.20 23.18 1.29 10−1

RACO2 1.76 9.78 10−3 119 16.61 29.23 1.62 10−1 RACO3 3.14 1.74 10−2 117 18.39 57.74 3.21 10−1

Flow induced crystallization

Figure 2.5 shows the evolution of Gwithout flow application for all grades at their

specific crystallization temperature. Clear differences between the different mate-rials are observed. The starting plateau is higher for increasing ethylene content. When a polymer melt is cooled down, the stiffness of the melt increases. Because the materials used show the same visco-elastic behavior and the crystallization exper-iments are performed at lower temperatures for the copolymers, the melt stiffness and thus the starting plateaus are higher. The horizontal shift factors, aT (Table 2.3),

at their respective Texp, calculated using the Arrhenius equation and normalized to aT

of the homopolymer, are 1, 1.4, 2.1 and 2.3 for the homopolymer PP and copolymers RACO1, RACO2 and RACO3, respectively. This corresponds well to the rise in start-ing plateau which increases by a factor 1.4 (RACO1), 1.9 (RACO2) and 2.1 (RACO3) compared to that of PP. The end levels of the storage moduli decrease with increas-ing ethylene content, which is correlated to the crystallinity values determined from DSC.

At equal nominal undercooling, crystallization is faster with increasing amount of ethylene monomers. The transition half time, t1/2, the time at which the half change

of the viscoelastic functions occur, is defined as a characteristic time scale of the pro-cess. To determine t1/2, the G′ data are converted to space filling,φ(ranging from 0

to 1). Khanna [54] definedφas:

φ = GG′ 0 G′ ∞−G ′ 0 (2.9) with G0and G∞′ the values of the start and end plateau respectively. Pogodina et

al. [55] normalized the storage modulus by:

φ = log(G/G′ 0) log(G′ ∞/G ′ 0) = log Glog G′ 0 log G′ ∞−log G ′ 0 (2.10) Whenφ, determined from optical microscopy, is plotted versus the log of time,

(36)

S-2.3 RESULTS AND DISCUSSION 21 100 101 102 103 104 105 103 104 105 106 107 108 Time [s] G’ [Pa]

Figure 2.5:Time build-up of the storage modulus under quiescent conditions at as-sumed equal undercooling for the homopolymer PP (◦) and the copoly-mers RACO1 (), RACO2 () and RACO3 (△).

shaped curves similar to those in Figures 2.5 and 2.6 are obtained [35]. Although Equation 2.10, is too simple to accurately estimate the degree of space filling and a more advanced suspension modeling is needed to derive the proper φ, the time scales obtained are similar. The end plateau shows a slight but steady increase, which makes a consistent determination of G∞′ problematic. Here, it is defined as

the intercept of the tangents drawn along the end plateau and the regime of maxi-mum changes (Figure 2.5). The homopolymer and RACO3 have a t1/2 of 7300s and

1100s, respectively, giving a ratio of 6.6. The often used Avrami equations describes

φ in case of isothermal crystallization, for which all nuclei, N, appear at the same time, t0, and the growth rate, G, is constant in time:

φ =1−exp[−4

NG

3(tt

0)3] (2.11)

Gahleitner et al. [29] determined N and G from thin slice experiments at tempera-tures between 80◦C and 120C. At all temperatures, a factor of 4 difference between the growth rate of PP (high G) and RACO3 (low G) was found. The growth rates of RACO1 and RACO2 are in between the two extremes. At their experimental crys-tallization temperatures, the difference in number of nuclei for PP and RACO3 is a factor 10 (∼ 1012.5 and∼ 1013.5 m−3, respectively) and the difference in growth rate

is a factor 3 (∼10−7.5and10−7 m/s, respectively). From Equation 2.11 the ratio of

t1/2 between the two materials can be derived, usingφ =0.5, and is equal to:

t1/2,PP t1/2,RACO3 = 3 s (NG3) RACO3 (NG3) PP (2.12)

(37)

22 2 FLOW INDUCED CRYSTALLIZATION OF P/ERANDOM COPOLYMERS

With the values found for N and G the ratio is 6.8, which is very similar to the mea-sured value.

Short-term shear flow experiments are performed at different shear rates applying a constant macroscopic strain (γ = 60) before monitoring crystallization. Results for the homopolymer and RACO3, the random copolymer with the highest ethylene content, together with those of crystallization under quiescent conditions, are shown in Figure 2.6. Applying a shear flow accelerates the crystallization process. For both materials, it is observed that there is no change in the initial G′. With increasing shear rate, the build-up of G’ starts at earlier times, but the slope with which Grises

remains the same, i.e. the kinetics of the process are not altered. The effectiveness of flow is more pronounced for the homopolymer than for the copolymer, RACO3, where the evolution of G′ for the lowest shear rate ( ˙γ = 3s−1) is almost equal to the

quiescent result. Apparently, a critical shear rate exists (at a constant macroscopic strain level) below which flow does not speed up the crystallization process. The effectiveness of flow can be quantified by defining a dimensionless crystallization half-time, Θ as [44]:

Θ= t1/2, ˙γ

t1/2, Q

(2.13) The FIC experiments are indicated by ˙γ, the Q indicates the half time for quiescent conditions. 100 101 102 103 104 105 103 104 105 106 107 108 Time [s] G’ [Pa] 100 101 102 103 104 105 103 104 105 106 107 108 Time [s] G’ [Pa] (a) (b)

Figure 2.6:Time build-up of the storage modulus for the homopolymer PP at 138◦C

(a) and the RACO3 at 117◦C (b) under quiescent conditions (◦) and after application of flow: 3s−1 for 20s (), 6s−1 for 10s (), 15s−1 for 4s (),

30s−1for 2s () and 60s−1for 1s (⊲).

Figure 2.7 shows the dimensionless crystallization half-time, Θ, as a function of shear rate, for the condition that the macroscopic applied shear strain is constant (γ=60). It shows that, under these conditions, there is a transition to morphology type 2 in which the flow is strong enough to accelerate the crystallization process. No change in crystallization kinetics is observed (Fig. 2.6) and, hence, only the number of point

(38)

2.3 RESULTS AND DISCUSSION 23 10−1 100 101 102 103 10−2 10−1 100 shear rate [s−1] Θ . γc RACO1 RACO3 PP RACO2

Figure 2.7:Dimensionless crystallization half-time, Θ, obtained from dynamical me-chanical experiments for PP (◦), RACO1 (), RACO2 () and RACO3 ()

as a function of shear rate, ˙γ, applied during the pre-shear condition with the macroscopic strain, γ = 60. The critical shear rate, ˙γc, for PP, is

indi-cated with the dotted line.

nuclei is increased. Morphology type 2 starts at a critical shear rate, ˙γc, above which

the step shear influences the crystallization kinetics. With increasing ethylene con-tent ˙γc shifts towards higher values. For RACO3, the lowest applied shear rate of

˙

γ = 3 s−1 is very close to ˙γ

c, which is a factor 4 higher than that of the

homopoly-mer (∼ 0.7 s−1). For all 4 grades, the onset of this regime, i.e. the critical shear rate,

falls within flow regime II, but it does not coincide with the transition between flow regime I and II, given by ˙γ =1/τreplong(Table 2.3). Furthermore, no transition to

mor-phology type 3, where fibrillar structures form, is observed. An exception in this picture is RACO2 which is influenced more by flow than the homopolymer, i.e. ˙γc

shifts to a lower value, due to the presence of the silica anti-blocking agent. When flow is applied, the presence of these particles constituting the additives locally in-creases the velocity gradients [21]. As such, fillers increase the influence of flow on crystallization and crystal orientation [19], and alter mechanical properties [16]. It has to be noticed that the quiescent experiments are not influenced by the presence of the particles, i.e. no extra nucleation effect is observed, both in the DSC mea-surements, which would show an increase in Tc (Figure 2.3(b)) and in the quiescent

rheometry measurements (Figure 2.5).

The minimum ˙γfor chain stretching for PP is calculated to be 15 s−1(De

s =1).

How-ever, for shear rates (equal and) higher than this critical value (15, 30 and 60 s−1, no

transition towards morphology type 3 is observed. The pre-shear conditions that ful-fill Des > 1, i.e. deformations strong enough to stretch the molecules and enter the

(39)

kinet-24 2 FLOW INDUCED CRYSTALLIZATION OF P/ERANDOM COPOLYMERS 100 101 102 103 104 105 103 104 105 106 107 108 Time [s] G’ [Pa] 100 101 102 103 104 105 103 104 105 106 107 108 Time [s] G’ [Pa] (a) (b) 100 101 102 103 104 105 103 104 105 106 107 108 Time [s] G’ [Pa] 10−1 100 101 102 103 10−2 10−1 100 shear rate [s−1] Θ γ (c) (d)

Figure 2.8:Time build-up of the storage modulus for the homopolymer PP at 138◦C

after application of flow at different shear rates: (a) ˙γ= 3 s−1, (b) ˙γ= 15s−1

and (c) ˙γ= 60s−1, and different strains; quiescent conditions (◦), FIC with γ= 60 (), γ = 120 (△) and γ = 240 (♦). (d) Dimensionless crystallization

half-time, Θ, versus ˙γfor the three different levels of shear strain, symbols are as in (a-c).

ics [15]. The duration of the flow also plays an important role. It is necessary to reach a characteristic strain at which a transition from one to the other regime is observed. This is demonstrated in Figure 2.8. Figure 2.8 (a) to (c) show FIC experiments with a shear rate of 3 s−1, 15 s−1 and 60 s−1, respectively, with increasing shear time such that the total shear strain applied is constant for all shear rates applied and equals to 60, 120, and 240. In all figures the quiescent case is included as well. For the low shear rates (up to 15 s−1) no change in crystallization kinetics is observed, only an

effect in the onset time of crystallization (Figure 2.8 (a, b)). Above a shear rate of 15 s−1, Figure 2.8 (c), the time evolution of the viscoelastic properties is changed. First,

an initial increase of G’ indicates the formation of some initial structure in the melt during flow and secondly, the slope of the G’-evolution curve is altered. For the three strain levels, the dimensionless crystallization half-time, Θ, is displayed in Figure 2.8 (d). It is shown that with an increasingγ, the critical shear rate at which flow starts to effect the crystallization process shifts to lower values. For the highest strain level

Referenties

GERELATEERDE DOCUMENTEN

Naast de veldstenen restanten zijn binnen deze fase van de kerk ook bakstenen, zogenaamde moefen, aangetroffen.. De moeilijkheid bestaat erin om uit te maken of het al

- Voor waardevolle archeologische vindplaatsen die bedreigd worden door de geplande ruimtelijke ontwikkeling en die niet in situ bewaard kunnen blijven:.. Wat is

proefsleuvenonderzoek uitgevoerd. De sleuven bevatten sporen uit de Bronstijd, Middeleeuwen en recentere periodes. Voor het oostelijke deel van het plangebied werd een

Op vraag van het Agentschap R-O Vlaanderen - Entiteit Onroerend Erfgoed werd in opdracht van IMWO-INVEST tussen 6 en 12 oktober en op 14 en 15 oktober 2009 een

Naar aanleiding van de herinrichting van de site Jonkhove, op de hoek van de Aartrijksestraat en de Hendrik Baekelandlaan voert Raakvlak op 25 oktober 2011 een

Publisher’s PDF, also known as Version of Record (includes final page, issue and volume numbers) Please check the document version of this publication:.. • A submitted manuscript is

Our mechanical robot needs sensorial elements in order to communicate with the outside environment and the computer unit.. Two encoder system give the position

waar: de diagonalen van een ruit staan loodrecht op elkaar, maar minstens één diagonaal van een vlieger wordt middendoor gedeeld.. waar: een parallellogram is een vierhoek met