Pp Crystallization

134
Flow induced crystallization of polymers: Precursors and Nuclei

description

crystallization and nucleation in PP

Transcript of Pp Crystallization

  • Flow induced crystallization of polymers:Precursors and Nuclei

  • Flow induced crystallization of polymers: Precursors and Nuclei / by Zhe Ma.Technische Universiteit Eindhoven, 2012.

    A catalogue record is available from the Eindhoven University of Technology Library.ISBN: 978-90-386-3299-5

    This thesis was prepared with the LATEX2 documentation system.Reproduction: University Press Facilities, Eindhoven, The Netherlands.Cover design: Zhe Ma and Paul Verspaget

    This research forms part of the research programme of the Dutch Polymer Institute(DPI), DPI project #714.

  • Flow induced crystallization of polymers:Precursors and Nuclei

    PROEFSCHRIFT

    ter verkrijging van de graad van doctor aan deTechnische Universiteit Eindhoven, op gezag van derector magnificus, prof.dr.ir. C.J. van Duijn, voor een

    commissie aangewezen door het College voorPromoties in het openbaar te verdedigen

    op maandag 3 december 2012 om 16.00 uur

    door

    Zhe Ma

    geboren te Hejian, Hebei, China

  • Dit proefschrift is goedgekeurd door de promotoren:

    prof.dr.ir. G.W.M. Peters

    en

    prof.dr.ir. H.E.H. Meijer

  • Contents

    Summary ix

    1 Introduction 11.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.2 Structure-properties relation . . . . . . . . . . . . . . . . . . . . . . . . . 21.3 Processing-structures relation . . . . . . . . . . . . . . . . . . . . . . . . 41.4 Scope of the thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

    References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6

    2 Using rheometry to determine nucleation density in a colored systemcontaining a nucleating agent 92.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102.2 Experimental . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

    2.2.1 Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112.2.2 Rheological characterization . . . . . . . . . . . . . . . . . . . . 112.2.3 X-ray characterization . . . . . . . . . . . . . . . . . . . . . . . 12

    2.3 Data analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122.4 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

    2.4.1 Determination of the nucleation density for NA-iPP . . . . . . . 142.4.2 Reproducibility . . . . . . . . . . . . . . . . . . . . . . . . . . . 162.4.3 Effect of mild flow . . . . . . . . . . . . . . . . . . . . . . . . . 17

    2.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

    3 Pressure quench of flow-induced crystallization precursors 233.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243.2 Experimental . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26

    3.2.1 Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 263.2.2 Protocol . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 273.2.3 X-ray characterization . . . . . . . . . . . . . . . . . . . . . . . 29

    3.3 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 303.3.1 Reference experiment: no pressure quench after flow . . . . . . . 303.3.2 Pressure quench after flow . . . . . . . . . . . . . . . . . . . . . 31

    v

  • vi Contents

    3.3.3 Pressure quench after annealing . . . . . . . . . . . . . . . . . . 343.3.4 Inverse quench by depressurization . . . . . . . . . . . . . . . . 38

    3.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

    Appendices. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42

    4 Short-term flow induced crystallization in isotactic polypropylene:how short is short? 454.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 464.2 Experimental . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48

    4.2.1 Material . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 484.2.2 Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 484.2.3 Optical microscopy . . . . . . . . . . . . . . . . . . . . . . . . . 50

    4.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 504.3.1 Rheological evolution during flow . . . . . . . . . . . . . . . . . 504.3.2 Structural evolution . . . . . . . . . . . . . . . . . . . . . . . . 52

    4.4 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 564.4.1 SAXS delay time . . . . . . . . . . . . . . . . . . . . . . . . . . 574.4.2 Implications of the viscosity rise . . . . . . . . . . . . . . . . . . 594.4.3 Conditions of the viscosity rise . . . . . . . . . . . . . . . . . . . 60

    4.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

    5 The influence of flow induced precursors and nuclei on crystallizationof isotactic polypropylene 655.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 665.2 Experimental . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68

    5.2.1 Material . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 685.2.2 Flow device . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 685.2.3 X-ray characterization . . . . . . . . . . . . . . . . . . . . . . . 685.2.4 Birefringence . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72

    5.3 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 735.3.1 Flow-induced distinguishable nuclei/precursors . . . . . . . . . . 735.3.2 Isothermal crystallization . . . . . . . . . . . . . . . . . . . . . . 75

    5.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87

    6 High-stress shear induced crystallization in isotactic polypropyleneand propylene/ethylene random copolymers 896.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 906.2 Experimental . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91

    6.2.1 Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 916.2.2 Flow device . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 926.2.3 X-ray characterization . . . . . . . . . . . . . . . . . . . . . . . 926.2.4 Optical microscopy . . . . . . . . . . . . . . . . . . . . . . . . . 93

  • Contents vii

    6.3 Depth sectioning method . . . . . . . . . . . . . . . . . . . . . . . . . . . 936.4 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96

    6.4.1 iPP homopolymer . . . . . . . . . . . . . . . . . . . . . . . . . . 966.4.2 Propylene/ethylene random copolymers . . . . . . . . . . . . . . 996.4.3 Quantification of nuclei . . . . . . . . . . . . . . . . . . . . . . . 102

    6.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106

    7 Conclusions and recommendations 1097.1 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1097.2 Recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111

    References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113

    Samenvatting 115

    Acknowledgements 119

    Curriculum vitae 121

    List of publications 123

  • Summary

    Semi-crystalline polymers cover over two thirds of commercial applications ofpolymeric materials. With most industrial processing technologies, semi-crystallinepolymers are brought in the molten state and then deformed for different reasons:i.e. to create specific properties like a high modulus with fiber spinning or to shapegeometrically complex parts as with injection molding. The deformation not onlyaccelerates the crystallization kinetics but also can cause crystalline structures to changefrom isotropic spherulites to the highly oriented cylindrical structures (e.g. shish-kebab) which determine the ultimate properties. Therefore, understanding the interplaybetween flow fields and the resulting crystalline structures is of great importance. Itenables the design of processing procedures and tailoring final product properties. Theobjective of this thesis is to reveal how a flow field affects the crystal structures bystudying the early stages of the crystallization process which are dominated by thecreation of shear-induced precursors and nuclei.

    Polymer crystallization involves two steps: nucleation and growth. Nucleationprovides quantitative (more or less) and qualitative (isotropic or oriented) nuclei andthese are the templates for further crystal growth that, ultimately, will fill the full space.The nucleation step is very sensitive and is often controlled by additives (nucleatingagents) or/and imposing a flow field. The formation and features of nuclei are, therefore,the key factors that determine the crystalline structures. Probing and quantifying nucleiis the main work of this thesis.

    According to the resulting morphology, nuclei can be divided into two groups: point-like nuclei and fibrillar nuclei. The former give rise to spherulites, the latter mainlyinduce oriented structures like the well-known shish-kebab; i.e. fibrils with transverselamellae. Oriented nuclei can be further classified into row nuclei and shish nuclei bywhether they can be directly observed with X-ray characterizations. In the first part ofthis thesis, the creation of point-like nuclei is studied for an isotactic polypropylene witha nucleating agent, with and without applied shear deformation. For such a system,nucleation is dramatically enhanced by both the nucleating agent (U-Phthalocyanine)and the flow. The nucleation density is that large and, therefore the crystallite sizeso small that optical microscope is not suitable to count nuclei numbers. Therefore, asuspension-based rheological model is used to quantify the nuclei density.

    ix

  • x Summary

    In the second part, precursors of row nuclei in a bimodal Polyethylene (PE) arestudied. Since row nuclei cannot be observed directly by X-ray characterizations, thecrystallization step is triggered by increasing the pressure which raises the equilibriummelting temperature and thus the under-cooling: a so-called pressure quench. It isshown that shear-induced precursors can be generated at a temperature close to thenominal melting temperature and the unstable ones can relax.

    Next, the formation of shish nuclei during flow is studied by using fast (30 frame/s),time resolved X-ray scattering in combination with rheological measurements. It isfound that a critical shear rate exists for the formation of shish nuclei within very shortflow times (0.25 s). Shish precursors are formed during flow and these were found todevelop into shish afterwards, during or after flow depending on the flow strength.

    Besides the external effects, the nuclei formation is also influenced by the molecularstructure. Therefore, in the last part, the effect of the molecular architecture onshear-induced crystallization is studied for isotactic polypropylene (iPP) and twopropylene/ethylene random copolymers with varying ethylene monomer contents. Thesethree grades had very similar rheological behavior; there is only a small but importantdifference in the longest relaxation time. Flow enhanced nucleation density was foundto be the lowest for the iPP homopolymer which is due to the shorter longest relaxationtime. However the reduced growth rate due to the added ethylene monomer leadsto slower overall crystallization kinetics of the random copolymers compared to thehomopolymer at similar thermal conditions.

  • Chapter one

    Introduction

    1.1 Background

    Human history is closely accompanied by the progress in material development.

    Advances in material are crucial for affecting peoples way of life because it can

    significantly change the way of working and improve output efficiency. The consequence

    of the material evolution is particularly obvious at the primary stages of civilization.

    For example, the innovation of iron farming tools such as plows and sickles, made the

    farming processes much faster and easier and, as a consequence, people produced enough

    food to survive and also gained extra time to exploit other activities. Considering such

    great importance, anthropologists even define historical epochs by the materials used

    at that moment, such as the Stone, Bronze and Iron ages. To meet the continuously

    increasing need of new functions, synthetic materials appeared and have been widely

    applied. Among current synthetic creations, plastic is quite prominent in our society

    due to its wide range of applicability. Just looking around, plastic products can be

    easily found at any time and any place. They may be a coffee stirrer, mobile shell,

    computer screen, bumper, greenhouse, construction and building parts, and so on.

    Actually, the development of truly synthetic plastics started only around one hundred

    years ago. The first completely synthetic one is poly(phenol-co-formaldehyde), a densely

    cross-linked thermoset polymeric material, well known as Bakelite, commercialized in

    1909 and applied for electrical appliances and phonograph records [1]. After its birth,

    plastic has developed dramatically during the past century. On one hand, more and

    more new plastics were synthesized and manufactured. Some major examples are:

    poly(vinyl chloride) (PVC) in 1920s, polyamide (Nylon) 66 in 1930s, linear high-density

    polyethylene (HDPE) and isotactic polypropylene (iPP) in 1950s, poly(p-phenylene

    terephthalamide) (known as Kevlar) in 1960s. [1, 2]

    1

  • 2 Chapter 1

    1950 1960 1970 1980 1990 2000 2010

    0

    50

    100

    150

    200

    250

    300

    2010: 265

    2009: 250

    2002: 204

    1989: 99

    1976: 47

    Mtonne

    year

    1950: 1.5

    Figure 1.1: World plastics production between 1950 and 2010. Data are obtained fromreference [3]. Only continuous growth is shown and temporary fluctuation is notincluded.

    On the other hand, the total amount of the plastic being used also increased

    tremendously, as shown in Figure 1.1. The world plastic production has risen from

    1.5 million tonnes in 1950 to 265 million tonnes in 2010, [3] equal to an averaged annual

    growth rate of 9% during sixty years.

    Polymers can be amorphous or semi-crystalline, depending on the degree of ordering

    of the molecular arrangement. Due to crystallization, semi-crystalline polymers can

    still be used as a solid material above the glass transition temperature. For example,

    although the glass transition temperature of HDPE is -80C, [1] its high crystallinity

    enables HDPE to be used as a solid-material at normal environmental temperatures.

    Therefore, semi-crystalline polymers, especially polyolefins such as HDPE and iPP, have

    become the most widely used polymeric materials, see Figure 1.2.

    1.2 Structure-properties relation

    Ultimate properties, e.g. mechanical and optical, of end-use plastic products strongly

    depend on the structure of the material [4, 5]. The manufacturing history includes two

    major processes: a) synthesis, starting with small molecules and b) processing, where

    the final product is given its shape. Therefore, the general issue structure includes

    two classes of defining features: the intrinsic chemical structures of the macromolecules

    and the physical arrangement structure of these macromolecules.

    The intrinsic chemical structure includes chemical composition (various monomers,

    homo-polymer/co-polymer, etc.), chain structure (linear/branched/star/crossed, ran-

    dom/block, etc.) and configuration, molecular weight, and so on. These molecular

  • Chapter 1 3

    Figure 1.2: European plastics demand by polymer type in 2010. [3]

    properties are controlled and dominated by the polymerization process.

    For amorphous polymers, considerable knowledge has been developed to understand

    the relationship between the basic molecular structure and mechanical behavior. Semi-

    crystalline polymers contain both amorphous and crystalline phases where an identical

    polymer chain often is involved in these different phases. The crystal phase has the

    hierarchic structures of conformation, crystal lattice, lamellar crystal, spherulitical or

    cylindrical assembling form of lamellae, etc. These crystal micro-structures cover a

    broad range of length scales from sub-nanometer up to microns. Polymer crystallization

    is determined by the molecular structures and, often in a very strong way, on the

    processing conditions. Therefore, when discussing structure-properties relation of semi-

    crystalline polymers, specific features of the crystal microstructure like crystallinity,

    lamellar thickness and orientation, should be taken into account. A typical and

    illustrative example [6] is shown in Figure 1.3, where an injection molded plate of HDPE

    is shown. From the optical micrographs it is observed that different structures appear

    in the thickness direction. These skin and core layers are due to the varying cooling

    and flow conditions as experienced by the material at different positions. Moreover,

    the thickness of these layers strongly depends on the position along flow path. Finally,

    for the same polymer material of HDPE, these distinct structures lead to completely

    different mechanical properties between specimens cut from the same plate but at

    different locations and in different directions. The failure mode varies from brittle

    to tough, the latter showing necking in one case and homogeneous deformation in the

    other.

    Next to molecular properties, full understanding on how crystal structures are formed

    during processing and how these subsequently affect the properties is a prerequisite of

    tailoring the ultimate properties of plastic products. This thesis focuses on the second

    step in the above knowledge chain, i.e. how polymer crystallizes under processing-

    relevant conditions.

  • 4 Chapter 1

    70x70x1mm

    injectionofpolymer

    A B

    C

    A

    B

    C

    Figure 1.3: Variation in microstructure over the thickness in an injection molded HDPEplate. Right column shows the resulting mechanical properties of threespecimens cut from the same plate but different positions. [6]

    1.3 Processing-structures relation

    Semi-crystalline polymers are often processed from the molten states and subjected

    to flow to transport the material and shape the product. Flow gradients affect

    crystallization by altering the amount, size and orientation of nuclei. In this way,

    processing-relevant flows are able to accelerate the crystallization kinetics by several

    orders of magnitude and induce anisotropic structures. According to the resulting

    morphology, flow-induced nuclei can roughly be divided into two groups: point-like

    nuclei and oriented fibrillar nuclei. During quiescent crystallization, only point-like

    nuclei are formed. The average crystal structure is isotropic and may grow into large

    spherulites because of the relatively low nuclei density, see Figure 1.4a. Depending on

    the competition between flow strength and molecular mobility, there are mild and strong

    flows. Mild flow forms extra point-like nuclei, i.e. increase nuclei density which results

    in more but smaller spherulites, see Figure 1.4b. In contrast, strong flow completely

    changes the isotropic morphology into oriented crystals. As an example highly oriented

    shish-kebab structures are shown in Figure 1.4c. [7]

    If the flow should effectively alter the crystallization behavior, the nuclei that are

    generated should be stable. However, in many cases flow-induced ordered structures

    are so-called precursors, instead of stable nuclei. These flow-induced precursors have

  • Chapter 1 5

    (a) (b) (c)

    Figure 1.4: Typical morphologies of: a) less and large spherulites; b) more but smallspherulites; c) oriented shish-kebab [7].

    a degree of ordering (position or/and orientation; intra- or/and inter- molecule) and

    favor further development to nuclei for sufficient under-cooling, or may relax back to

    amorphous melt in case of relatively high temperatures. When the unstable precursors

    are inclined to disappear, the slow kinetics of precursor relaxation may cause significant

    memory effect on subsequent crystallization, since cooling stabilizes and reactivates the

    residual precursors. [8]

    By stretching polymer chains and aligning segments parallel to each other, flow

    promotes nucleation in polymer. On the other hand, foreign substances, known as

    nucleating agent (NA), are added to the polymer melt and the additional surfaces

    lower the surface energy change in an efficient way and decrease the critical size for

    nucleation [9]. Moreover, specific nucleating agents can control the phase modification of

    growing crystals. For example, -NA is utilized to induce formation of iPP hexagonal -

    crystals [1012]. Understanding how the amount and orientation of nuclei are controlled

    by flow or/and NA is a key step in revealing processing-structure relations.

    1.4 Scope of the thesis

    The aim of this thesis is to gain understanding on how polymer crystallization under

    processing-relevant conditions (flow and NA) takes place by exploring the creation and

    evolution of precursors/nuclei. During processing, a continuous flow imposed during

    solidification, may influence both nucleation and crystal growth. To separate the

    influence of flow on these two aspects from rheological changes due to crystallization,

    Janeschitz-Kriegl and co-workers introduced the short-term shearing method [13].

    Flow duration is so short that structure formation and a related change in the viscosity

    can be minimized. The flow-induced precursors/nuclei can be revealed by direct

    observations, using state of the art experimental techniques, or by examining subsequent

    crystallization. In this thesis, we use both approaches to study (probe or/and quantify)

    precursors/nuclei induced by various flow strength. Moreover, the validity of short-

  • 6 Chapter 1

    term flow is verified for strong shear conditions. The polymers investigated are grades

    of a high density polyethylene (HDPE) and an isotactic polypropylene (iPP). They are

    not only the most used polymers but also the simplest macromolecules and provide

    suitable model systems for the academic research.

    First of all, generation of extra point-like nuclei is studied in Chapter 2. Addition of

    nucleating agents (U-Phthalocyanine) and imposition of mild flow are both utilized to

    enhance formation of point-like nuclei in an iPP system. A suspension-based rheological

    model [14,15] is applied to quantify nucleation density. In this way, the nucleation effects

    of this nucleating agent and mild flows applied are assessed.

    Oriented nuclei are investigated in Chapter 3-6. Oriented nuclei can be classified

    as row nuclei and shish, according to whether such nuclei can be detected by X-ray

    characterization.

    Chapter 3 focuses on the precursors of the X-ray invisible row nuclei. Shear-induced

    precursors cannot be observed directly by X-ray. Therefore, after increasing the under-

    cooling, subsequent crystallization kinetics and orientation evolution is examined to

    reflect the underlying precursors. A pressure quench method is presented, which

    provides an efficient way of achieving sufficient under-cooling in a fast way for triggering

    crystallization and effectively lightens up precursors.

    In the case of very strong shear, X-ray observable shish appear during flow. In

    Chapter 4, the combination of a slit flow with time resolved X-ray measurements

    provides the possibility to study in-situ structure formation of iPP during flow. In this

    way, shish formations during and after flow can be distinguished. Also the evolution

    of the average rheological behavior is tracked. The appearance of shish structure and

    the rise of the apparent viscosity identify the validity of the conditions for short-term

    flow, i.e. how short is short depends on shear strength.

    For the weakest flow conditions studied in Chapter 4, X-ray and rheology fail to detect

    shear-induced structures during flow. Therefore, in Chapter 5, birefringence is employed

    to probe the potential precursors in these flows. Subsequent crystallization features

    such as kinetics, orientation and -crystals are tracked in time, in order to establish a

    correlation between the precursors/nuclei and their influences on crystallization.

    The effect of flow strength in relation with molecular structure (defects present in the

    molecular chain and high molecular tail) on crystallization is explored in Chapter 6, by

    comparing the amount of oriented nuclei in an iPP and two ethylene/propylene random

    copolymers with various ethylene content. A pressure-driven slit flow device [16, 17]

    is employed to impose strong flows with a wall stress of up to 0.11 MPa. With this

    novel device, the depth sectioning method [18] can be applied to extract the polymer

    crystallization in certain layers and correlate that with the specific stress. In this way,

    the effect of molecular structures on flow induced nucleation can be revealed.

  • Chapter 1 7

    References

    [1] L.H. Sperling. Introduction ot Physical Polymer Science. Wiley, fourth edition edition, 2006.

    [2] D. I. Bower. An Introduction to Polymer Physics. Cambridge University Press, 2002.

    [3] Plastics Europe. Plastics - the facts 2011. an analysis of european plastics production, demand

    and recovery for 2010. Technical report, 2011. Http://www.plasticseurope.org.

    [4] B. A. G. Schrauwen, R. P. M. Janssen, L. E. Govaert, and H. E. H. Meijer. Macromolecules

    37(16):60696078, 2004.

    [5] B. A. G. Schrauwen, L. C. A. Von Breemen, A. B. Spoelstra, L. E. Govaert, G. W. M. Peters,

    and H. E. H. Meijer. Macromolecules 37(23):86188633, 2004.

    [6] B. A. G. Schrauwen. Deformation and Failure of Semicrystalline Polymer Systems: Influence of

    Micro and Molecular Structure. Ph.D. thesis, Eindhoven University of Technology, 2003.

    [7] B. S. Hsiao, L. Yang, R. H. Somani, C. Avila-Orta, and L. Zhu. Physical Review Letters

    94(11):117802, 2005.

    [8] D. Cavallo, F. Azzurri, L. Balzano, S. S. Funari, and G. C. Alfonso. Macromolecules 43(22):9394

    9400, 2010.

    [9] B. Wunderlich. Macromolecular Physics, Vol 2: Crystal Nucleation, Gorwth, Annealing.

    Academic Press, New York, 1976.

    [10] J. Garbarczyk and D. Paukszta. Polymer 22(4):562564, 1981.

    [11] W. Stocker, M. Schumacher, S. Graff, A. Thierry, J. C. Wittmann, and B. Lotz. Macromolecules

    31(3):807814, 1998.

    [12] F. Luo, C. Geng, K. Wang, H. Deng, F. Chen, Q. Fu, and B. Na. Macromolecules 42(23):9325

    9331, 2009.

    [13] S Liedauer, G Eder, H Janeschitz-Kriegl, P Jerschow, W Geymayer, and E. Ingolic. International

    Polymer Processing 8:236244, 1993.

    [14] R. J. A. Steenbakkers and G. W. M. Peters. Rheologica Acta 47(5-6):643665, 2008.

    [15] J. W. Housmans, R. J. A. Steenbakkers, P. C. Roozemond, G. W. M. Peters, and H. E. H. Meijer.

    Macromolecules 42(15):57285740, 2009.

    [16] G. Kumaraswamy, R. K. Verma, and J. A. Kornfield. Review of Scientific Instruments 70(4):2097

    2104, 1999.

    [17] G. Kumaraswamy, A. M. Issaian, and J. A. Kornfield. Macromolecules 32(22):75377547, 1999.

    [18] L. Fernandez-Ballester, D. W. Thurman, and J. A. Kornfield. Journal of Rheology 53(5):1229

    1254, 2009.

  • Chapter two

    Using rheometry to determine

    nucleation density in a colored

    system containing a nucleating

    agent

    Abstract

    A new suspension-based rheological method was applied to experimentally study the

    crystallization of a nucleating agent (NA) filled isotactic polypropylene. This method

    allows for determination of point nucleation densities where other methods fail. For

    example, optical microscopy can fail because nucleation densities become too high to

    be counted (materials with effective NA) or crystallites are not easily visible (colored

    materials), while differential scanning calorimetry does not allow the effect of flow to be

    studied. Both quiescent and mild-shear induced crystallization were investigated. The

    results show that the addition of a nucleating agent increases the nucleation density by

    six decades for quiescent crystallization. The effect of shear on crystallization in the

    presence of a nucleating agent was assessed and it is demonstrated that, at least for

    this system, the effect of applied shear is much smaller than the effect of the nucleating

    agent.

    This chapter is based on : Zhe Ma, Rudi J.A. Steenbakkers, Julien Giboz, Gerrit W.M. Peters.Rheologica Acta 50:909915, 2011

    9

  • 10 Chapter 2

    2.1 Introduction

    Semi-crystalline polymers cover over two thirds of the products used in our daily life.

    Adding a nucleating agent to a semi-crystalline polymer is a common way to control

    crystallization and tailor mechanical and optical properties [14]. The interface between

    the nucleating agent and the polymer melt has a high surface energy, which reduces

    the energy barrier associated with the formation of a nucleus. Introducing a large

    amount of nucleating agent particles therefore greatly increases the nucleation density

    and consequently accelerates crystallization [5]. In general, a higher nucleation density

    leads to more desirable properties. Several methods are available to determine the

    nucleation density in polymer crystallization. These include (a) microscopy to directly

    count the number of nuclei in the melt, (b) differential scanning calorimetry (DSC) [6],

    or (c) dilatometry [7]. The latter two indirectly measure the crystallization evolution in

    order to calculate the nucleation density using a kinetics equation. However, for colored

    nucleating agent systems, the nucleation density can become too large to be counted

    and also can change the optical properties dramatically [4], making quantification of

    the nucleation density difficult with optical microscopy (Figure 2.1).

    Figure 2.1: Morphology during quiescent crystallization at T = 151 C.

    Flow is another crucial factor affecting polymer crystallization, especially the

    nucleation step. The number of nuclei can increase dramatically [8, 9] and for a strong

    enough flow the formation of shish-kebab structures will occur [10]. The combined

    effect of nucleating agent and flow has hardly been studied. In this work we will restrict

    ourselves to point-like nucleation which will occur for moderate flow only. How much

    a flow will change the kinetics of nucleating agent crystallization depends on the type

    of nucleating agent [11]. Knowledge of the individual effect of nucleating agent and

    shear on nucleation density is a prerequisite for understanding the combined nucleation

    mechanism and to provide input for models to predict resulting structures.

  • Chapter 2 11

    To achieve this goal, conventional DSC and classical dilatometry are not suited since

    they do not allow one to impose a flow on the material. A new type of dilatometer,

    the Pirouette [12], developed in our group [13, 14], does allow for applying shear to

    the sample and thus for using the measured time dependent specific volume for our

    purpose; determining the nucleation density. However, in this chapter we will use a

    rheometer that first serves as flow device (with the possibility to vary the shear rate

    and the shear time) and, subsequently, as a mechanical spectrometer that measures

    the changing complex modulus due to the progressing crystallization process. Using an

    analytical relation between space filling and the modulus, it is possible to determine

    the nucleation density. This solves both problems: dealing with large numbers of nuclei

    and poor visibility of crystalline structures. The aim of this study is to explain this

    approach and to demonstrate its effectiveness for colored nucleating agent systems, for

    both quiescent and flow-induced crystallization.

    2.2 Experimental

    2.2.1 Materials

    The polymer used in this study is an isotactic polypropylene (iPP; HD601CF,

    Borealis, previously known as HD120MO). It has a weight average molecular weight

    Mw = 365 kg/mol and a polydispersity Mw/Mn = 5.4 [9]. The nominal melting

    temperature is 163 C. The polymer was compounded with an organic nucleating

    agent, U-Phthalocyanine of molecular weight 310 kg/mol [15] at a concentration of

    0.2wt%. This nucleating agent was also used by Lee Wo and Tanner [16] who found

    non-spherical crystallites that we did not observe, see section 2.4. From the neat iPP

    and the artificially nucleated material (NA-iPP), 1.1-mm-thick plates were injection

    molded and from those plates circular disks were cut with a diameter of 8mm. The

    NA-iPP samples were blue due to the coloring effect of the nucleating agent.

    2.2.2 Rheological characterization

    For the rheological measurements, a Rheometrics ARES rheometer with a plate-plate

    geometry was used. Samples were first heated to 230 C and kept on that temperature

    for 10 min to erase the thermal and mechanical history. Next, the melt was cooled to

    the desired temperature at a rate of 15 C/min and kept at this temperature for the

    isothermal crystallization. During cooling, gap adjustment was performed continuously

    to compensate for the thermal shrinkage of tools and sample. Two minutes of delay

    time was used to equilibrate the sample temperature before starting the dynamic

    measurements. Temperatures for crystallization were chosen between 133 and 140 C

  • 12 Chapter 2

    for the neat iPP and between 143 and 151 C for the NA-iPP. Experiments with shear

    were carried out at temperatures of 148 and 151 C. Steady shear at a rate of = 60 s1

    was imposed for shear times ts = 2, 4 and 6 s. Small-amplitude oscillatory shear

    measurements were employed to track the time evolution of the storage modulus (G)

    and loss modulus (G) using an angular frequency of 5 rad/s and a strain of 0.5%. All

    experiments were performed in a N2 atmosphere to prevent the material from degrading.

    2.2.3 X-ray characterization

    Polypropylene can crystallize in different phases. To check if both the neat iPP and

    the NA-iPP had crystallized in the -phase, the samples were analyzed afterwards using

    X-ray scattering. Wide angle X-ray diffraction (WAXD) measurements were carried

    out at the Dutch-Belgian (DUBBLE) beamline BM26B of the European Synchrotron

    Radiation Facility in Grenoble, France. [17] A Photonica CCD detector with 20041335pixels of 44m 44m was placed at 178mm. The wavelength used was 1.033 A andthe exposure time was 10 s. WAXD data were integrated with the software FIT2D.

    2.3 Data analysis

    A linear viscoelastic version of the three dimensional generalized self-consistent

    method of Christensen et al. [18] is used to couple the amount of space filling, caused

    by point-like nucleation and spherulitic growth, to the measured dynamic (or complex)

    modulus. See also the work of Christensen et al. [19,20]. This model has been validated

    experimentally in our previous work [21]. If spherulites are formed, the relative dynamic

    modulus f G = G/G0 can be obtained from

    Af G2 +Bf G + C

    = 0 , (2.1)

    where G and G0 are the complex dynamic modulus of the suspension and the

    amorphous phase, respectively. The complex coefficients A, B, and C depend on

    space filling , the ratio of the complex moduli of the amorphous (G0) and crystalline

    phase (G1), and the Poisson ratios of both phases, 0 and 1. Notice that all moduli

    are frequency and temperature dependent. Expressions for the coefficients are given in

    Appendix A of Steenbakkers et al. [21]. In this case, space filling is the unknown and

    is obtained by sloving Eq. (2.1) using the measured f G.

    Next the space filling has to be related to the nucleation density. For a fixed number

    density of nuclei N(T ), the Kolmogorov Avrami Evans equation [2225] describes the

  • Chapter 2 13

    progress of space filling in time during isothermal crystallization at temperature T ,

    (t, T ) = 1 exp(4pi

    3N(T )G3(T )t3

    )(2.2)

    from which the nucleation density can be obtained:

    N(T ) = 3 ln (1 )4piG3(T )t3

    (2.3)

    with growth rate G(T ) depending on the temperature only. The main requirement here

    is that the growth rate is independent of the nucleation density, even when this density

    is influenced by a nucleating agent or by flow. Flow experiments are so-called short

    term shear experiments; the flow can generate extra nuclei but the growth takes place

    after the flow has stopped. The growth rate for iPP is well known [26]. Note that there

    are different empirical and theoretical relations describing the temperature dependence

    of the growth rate, see Eq. (2.4) [8] and Eq. (2.5) [27]:

    G(T ) = Gref exp[cG(T TG,ref)2] (2.4)

    G(T ) = G0 exp

    [ U

    R (T Tg + T)]

    exp[GT

    2m (Tm + T )

    2T 2 (Tm T )]

    (2.5)

    One could think of using other approaches than Eq. (2.1) by applying more simple

    models [2832] or empirically based scaling laws between the space filling and the storage

    modulus [3335]. However, it was demonstrated by Steenbakkers and Peters [21] that

    such approaches do perform less than the suspension model as used here, and therefore,

    we will not apply them to our results. Summarizing, by measuring the complex modulus

    G(t, T, , ts) for different temperatures, shear rates , and shear times ts, we can

    determine the space filling (t, T ) using Eq. (2.1) and the nucleation density N(T, , ts)

    using Eq. (2.3), provided that only spherulitic growth from predetermined point-like

    nuclei occurs.

  • 14 Chapter 2

    2.4 Results and Discussion

    2.4.1 Determination of the nucleation density for NA-iPP

    First of all, Figure 2.1 shows that the structures formed during crystallization with

    nucleating agent do not show any unusual structure as was found by Lee Wo et al. [16] for

    a similar system (a commercial iPP with different concentrations of U-Phthalocyanine),

    so the above method can be applied to convert the time evolution of the rheological

    properties into kinetics of space filling. Figure 2.2 shows the time evolution of the

    storage modulus G of the nucleating agent system during crystallization at different

    temperatures. It should be noticed that for T = 145 C the initial value of G is not

    the same as those for the other temperatures. The method is based on the assumption

    of isothermal crystallization. If the degree of undercooling is large with respect to

    the cooling rate, crystallization already sets in during cooling. The problem is made

    even worse due to the 2-min delay time we used in order to equilibrate the sample

    temperature. This is the case for T = 145 C; crystallization has already started before

    we begin to track the rheological evolution (indicated by an increased initial value of

    G). Faster cooling rates are required when studying higher levels of undercooling.

    Nevertheless, we still want to show this result to demonstrate the limitations of the

    method.

    10 100 1000 10000

    10

    4

    10

    5

    151

    o

    C

    148

    o

    C

    147

    o

    C

    145

    o

    C

    10

    6

    10

    7

    10

    8

    G' (Pa)

    time (s)

    Figure 2.2: Time evolution of the storage modulus for NA-iPP samples crystallized atdifferent temperatures.

    The related space filling, determined by Eq. (2.1), is shown in Figure 2.3. For the high

    experimental temperature range used here, the growth rate is better captured by Eq.

    (2.4) [8, 26], so this expression was applied. Figure 2.4 shows the diffraction patterns

    of NA-iPP samples after quiescent crystallization and demonstrates that nucleating

  • Chapter 2 15

    10 100 1000 10000

    0.0

    0.2

    0.4

    0.6

    0.8

    1.0

    151

    o

    C

    148

    o

    C 147

    o

    C 145

    o

    C

    space filling

    time (s)

    Figure 2.3: Time evolution of space filling derived from the dynamic modulus for NA-iPPsamples crystallized at different temperatures.

    9 10 11 12 13 14 15

    145

    o

    C

    147

    o

    C

    148

    o

    C

    (-131)

    (111)

    (130)

    (040)

    Intensity (a.u.)

    2 (

    o

    )

    (110)

    151

    o

    C

    Figure 2.4: One-dimensional WAXD curves of NA-iPP samples crystallized at differenttemperatures.

    agent crystallization results in the same -modification as crystallization of the neat

    iPP. Consequently, the same growth kinetics, Eq. (2.4), applies. The number density

    of nuclei determined by Eq. (2.3) is plotted versus space filling in Figure 2.5. It is

    nearly constant between 0.1 and 0.9, except for the experiment at T = 145 C,where crystallization has already set in before the dynamic measurements have started.

    Much higher values are found in the early stage, but N(T ) becomes nearly constant

    when 0.4 for all temperatures. The influence of the unknown initial space filling isrelatively smaller for higher degrees of space filling, i.e., in later stages of the process.

    We have taken the average value for a space filling between = 0.5 and = 0.9 as a

    reasonable approximation. Nucleation densities from all experiments are plotted as a

    function of the experimental temperature in Figure 2.6. It can be concluded that this

    type of nucleating agent is very effective as it increases the nucleation density by up to

  • 16 Chapter 2

    0.0 0.2 0.4 0.6 0.8 1.0

    151

    o

    C

    148

    o

    C

    147

    o

    C

    145

    o

    C

    10

    14

    10

    16

    10

    18

    10

    20

    N (m

    -3

    )

    space filling

    10

    22

    Figure 2.5: Nucleation density versus space filling for NA-iPP samples crystallized atdifferent temperatures.

    130 135 140 145 150 155

    10

    13

    10

    15

    10

    17

    10

    11

    NA+iPP

    iPP

    N (m

    -3

    )

    temperature (

    o

    C)

    10

    19

    Figure 2.6: Nucleation density of iPP and NA-iPP versus temperature.

    six decades. Notice that the temperature dependency of the nucleation density is very

    similar for the neat iPP and the NA-iPP, and that the results for T = 145 C, although

    less reliable for < 0.4, are well in line with the results for higher temperatures.

    Results for relatively low temperatures should be treated with some caution.

    2.4.2 Reproducibility

    First we compare the results of Housmans et al. [9], obtained with the same method

    and for the same iPP, to the results presented here. For a temperature T = 138 C,

    they found, for quiescent conditions, a nucleation density N = 8 1011m3, while weobtain the (interpolated) value N = 6 1012m3. We ascribe this difference to the

  • Chapter 2 17

    sample preparation procedure. Our neat iPP and nucleated samples were processed by

    injection molding, which means that the sample preparation step already induced extra

    nuclei due to the applied (uncontrolled) flow. The samples used by Housmans et al. [9]

    were prepared by means of compression molding.

    130 135 140 145 150 155

    temperature (

    o

    C)

    1

    st

    Series NA+iPP

    1

    st

    Series iPP

    2

    nd

    Series NA+iPP

    2

    nd

    Series iPP

    10

    11

    10

    13

    10

    15

    10

    17

    10

    19

    N (m

    -3

    )

    Figure 2.7: Nucleation density of iPP and NA-iPP versus temperature for two series ofexperiments.

    We repeated our experiments to check for reproducibility. The time lapse between the

    two series of experiments was about 8 weeks. The results of these repeated experiments

    (2nd series) are shown in Figure 2.7, together with the previous results (1st series). We

    want to stress the importance of a good temperature control. An error of 1 C typically

    gives a factor two difference in the number of nuclei. Notice that in the second series

    of experiments we managed to get good results for temperatures as low as T = 143 C

    while in the first series we already encountered problems at T = 145 C.

    2.4.3 Effect of mild flow

    Figure 2.8 shows the evolution of G during crystallization at 148 C under quiescent

    conditions and after different flows (fixed shear rate and variable shear time). For a

    shear rate of 60 s1, a shear time of 2 s shows clearly accelerated kinetics. Further

    increase of the shear time to 4 s does not change the kinetics much. The acceleration

    of crystallization seems to become (nearly) independent of the shear time beyond 4 s,

    indicating that the shear-enhanced point-like nucleation saturates. This effect was also

    observed by Housmans et al. [9] for this and two other iPPs. The time evolution of

    the storage modulus of the sheared nucleating agent system has the same shape as in

    the quiescent nucleating agent system, meaning that the growth mechanism is the same

    and only spherulites are formed. This implies that our method can still be applied

  • 18 Chapter 2

    to determine space filling for crystallization after shear. Moreover, the results can be

    compared to those of Housmans et al. [9].

    10 100 1000 10000

    0

    2

    4

    time (s)

    10

    4

    10

    5

    10

    6

    10

    7

    10

    8

    G' (Pa)

    shear time (s)

    6

    Figure 2.8: Time evolution of the storage modulus for NA-iPP under quiescent conditionsand after short term shear ( = 60 s1, ts = 2, 4, 6 s) at T = 148

    C.

    For flow-induced crystallization experiments, an adequate measure of the flow

    strength is the Weissenberg number, based on the stretch relaxation time of the high

    molecular weight (HMW) tail of the molecular weight distribution:

    Wis(T, ) = HMWs (T ) (2.6)

    Recent modeling work [36] suggests that the creation rate of point-like nucleation

    precursors depends not only on the average stretch of the HMW molecules, but also

    explicitly on the temperature: the prefactor of the creation rate was found to be

    proportional to the time-temperature shift factor aT . Therefore we use aTWis as a

    criterion to compare experiments, which scales as a2T , since HMWs in Eq. 2.6 scales

    with aT as well. From the linear viscoelastic data and the stretch relaxation time HMWs

    reported by Housmans et al. [9], we find Wis(148C) = 13, Wis(151

    C) = 12, and

    aT (151C)Wis(151

    C)

    aT (138 C)Wis(138 C)= 1.1 , (2.7)

    where Wis(138C) = 8 for the strongest flow applied by Housmans et al. [9] to the same

    neat iPP as used here. Hence, these two data sets are reasonably comparable in terms

    of flow conditions.

    The calculated nucleation densities are plotted in Figure 2.9. The saturated

    nucleation density at 148 C after shear is around 40 1016m3, six times higher than7 1016m3 for quiescent crystallization, i.e., adding typically 1017m3 nuclei. This

  • Chapter 2 19

    0 2 4 6

    10

    17

    10

    16

    10

    15

    10

    18

    151

    o

    C

    148

    o

    C

    N (m

    -3

    )

    shear time (s)

    Figure 2.9: Nucleation density of NA-iPP under quiescent conditions and after shear atT = 148 C and T = 151 C ( = 60 s1).

    130 135 140 145 150 155

    temperature (

    o

    C)

    1

    st

    Series NA+iPP

    1

    st

    Series iPP

    1

    st

    Series shear

    2

    nd

    Series NA+iPP

    2

    nd

    Series iPP

    10

    11

    10

    13

    10

    15

    10

    17

    10

    19

    N (m

    -3

    )

    Figure 2.10: Nucleation densities of iPP and NA-iPP under quiescent and shear conditions.

    elevated number is much more than what was added by imposing shear to the neat iPP

    in the experiments of Housmans et al. [9]; mild flow raised the point-like nucleation

    density of the neat iPP by one to two decades before the shear effect saturated, i.e.

    adding typically 1013m3 nuclei. This indicates that, in the presence of nucleatingagent, shear-induced point nucleation can be much more effective. This should be

    directly related to the nucleating agent, i.e., the local amplification of flow effects due

    to the presence of the nucleation particles, see for example the work of Hwang et al. [37],

    and not to a change in the rheological properties, which we expect to be very small due

    to the, in this respect, still very low space filling of nuclei.

  • 20 Chapter 2

    2.5 Conclusions

    A new method to determine nucleation densities, based on rheometry, was used

    on quiescent and sheared samples of a neat and an artificially nucleated isotactic

    polypropylene. All calculated nucleation densities are summarized in Figure 2.10. It

    is quantitatively shown that U-Phthalocyanine is very effective for nucleating isotactic

    polypropylene. Moreover, it was found that the effect of shear is enhanced by the

    presence of the nucleating agent. This rheological method is easy to apply since it

    requires a standard rheometer, available in most academic and industrial labs.

    References

    [1] J. Kurja and N. A. Mehl. Nucleating agents for semicrystalline polymers. In: H. Zweifel, editor,

    Plastics Additives Handbook, 949972. Hanser, Munich, 5 edition, 2001.

    [2] M. Kristiansen, M. Werner, T Tervoort, P. Smith, M. Blomenhofer, and H. Schmidt.

    Macromolecules 36:51505156, 2003.

    [3] M. Kristiansen, T. Tervoort, P. Smith, and H. Goossens. Macromolecules 38:1046110465, 2005.

    [4] A. Menyhard, M. Gahleitner, J. Varga, K. Bernreitner, P. Jaaskelainen, H. ysd, and B. Pu-

    kanszky. Eur Polym J 45:31383148, 2009.

    [5] B. Wunderlich. Macromolecular Physics, Vol. 2: Crystal Nucleation, Growth, Annealing.

    Academic Press, London, 1976.

    [6] G. Lamberti. Polym Bull 52:443449, 2004.

    [7] F. L. Binsbergen and B. G. M. De Lange. Polymer 11:309332, 1970.

    [8] H. Zuidema, G. W. M. Peters, and H. E. H. Meijer. Macromol Theory Simul 10:447460, 2001.

    [9] J. W. Housmans, R. J. A. Steenbakkers, P. C. Roozemond, G. W. M. Peters, and H. E. H. Meijer.

    Macromolecules 42(15):57285740, 2009.

    [10] P. Zhu and G. Edward. Macromolecules 37:26582660, 2004.

    [11] D. Bylov, Panine P., K. Remerie, E. Biemond, G. C. Alfonso, and W. H. De Jue. Polymer

    49:30763083, 2008.

    [12] Http://www.imetechnologies.nl.

    [13] M. H. E. Van der Beek, G. W. M. Peters, and H. E. H. Meijer. Macromolecules 39:92789284,

    2006.

    [14] R. Forstner, G. W. M. Peters, and H. E. H. Meijer. International Polymer Processing 24:114121,

    2009.

    [15] C. Hadinata, 2009. Personal communication.

    [16] D. Lee Wo and R. I. Tanner. Rheologica Acta 49:7588, 2010.

    [17] W. Bras, I. P. Dolbnya, D. Detollenaere, R. van Tol, M. Malfois, G. N. Greaves, A. J. Ryan, and

    E. Heeley. Journal of Applied Crystallography 36:791794, 2003.

    [18] R. M. Christensen and K. H. Lo. J. Mech, Phys, Soild 27:315330, 1979.

    [19] R. M. Christensen and K. H. Lo. J. Mech, Phys, Soild 34:639, 1986.

    [20] R. M. Christensen and K. H. Lo. J. Mech, Phys, Soild 38:379404, 1990.

    [21] R. J. A. Steenbakkers and G. W. M Peters. Rheol Acta 47:643665, 2008.

    [22] A. N. Kolmogorov. Bull. Acad. Sci. USSR, Math. Series (in Russian) 1:355359, 1937.

  • Chapter 2 21

    [23] M. Avrami. Journal of Chemical Physics 7:11031112, 1939.

    [24] M. Avrami. Journal of Chemical Physics 8:212224, 1940.

    [25] U. R. Evans. Transactions of the Faraday Society 41:365374, 1945.

    [26] G. Eder and H. Janeschitz-Kriegl. Structure development during processing: crystallization. In:

    H. E. H. Meijer, editor, Processing of Polymers, number 18 in Materials Science and Technology:

    a Comprehensive Treatment, 269342. Wiley-VCH, Weinheim, 1997.

    [27] J. D. Hoffman, G. T. Davis, and J. I. Lauritzen. The rate of crystallization of linear polymer with

    chain folding. In: N. B. Hannay, editor, Treatise on Solid State Chemistry, volume 3, 497614.

    Plenum Press, New York, 1976.

    [28] K. Boutahar, C. Carrot, and J. Guillet. Journal of Applied Polymer Science 60(1):103114, 1996.

    [29] K. Boutahar, C. Carrot, and J. Guillet. Macromolecules 31(6):19211929, 1998.

    [30] R. I. Tanner. Journal of Non-Newtonian Fluid Mechanics 102:397408, 2002.

    [31] R. I. Tangger. Journal of Non-Newtonian Fluid Mechanics 112(23):251 268, 2003.

    [32] N. Van Ruth, J. Vega, S. Rastogi, and J. Martnez-Salazar. Journal of Materials Science 41:3899

    3905.

    [33] N. V. Pogodina, H. H. Winter, and S Srinivas. Journal of polymer science Part B-polymer physics

    37(24):35123519, 1999.

    [34] S. Coppola, S. Acierno, N. Grizzuti, and D. Vlassopoulos. Macromolecules 39(4):15071514,

    2006.

    [35] Y. P. Khanna. Macromolecules 26(14):36393643, 1993.

    [36] R. J. A. Steenbakkers. Precursors and Nuclei, the Early Stages of Flow-Induced

    Crystallization. Ph.D. thesis, Technische Universiteit Eindhoven, 2009. Available at

    http://www.mate.tue.nl/mate.

    [37] W. R. Hwang, G. W. M. Peters, M. A. Hulsen, and H. E. H. Meijer. Macromolecules 39(24):8389

    8398, 2006.

  • Chapter three

    Pressure quench of flow-induced

    crystallization precursors

    Abstract

    We developed a novel protocol to study the mutual influence of shear flow and

    pressure on crystallization of polymers. Here, we have applied this protocol, named

    pressure quench, to polyethylene with a bimodal molecular weight distribution. With

    pressure quench, the undercooling, required to initiate crystallization of flow-induced

    precursors generated at high temperature, is obtained by increasing pressure, i.e.,

    leaving the specimen isothermal. We find that pressure enhances the effect of shear. In

    particular, results show that the pressure quench effectively lightens up shear-induced

    precursors which otherwise are not observable, even with high-resolution synchrotron

    X-ray scattering. A pressure quench in combination with SAXS and WAXD gives

    insight into the early stages of crystallization. In this chapter we focus on the use of

    WAXD since it provides all the information required to demonstrate our main issues.

    We conclude that precursors with different stability can be formed during shear and,

    with annealing, the least stable ones relax back to the melt. Finally, it is demonstrated

    that when pressure is released after crystallization, an inverse quench takes place and

    crystalline structures partially melt, similar to an increase of the temperature.

    This chapter is based on : Zhe Ma, Luigi Balzano, Gerrit W.M. Peters. Macromolecules 45:42164224, 2012

    23

  • 24 Chapter 3

    3.1 Introduction

    Semicrystalline polymers represent the largest group of commercial polymeric

    materials. Their properties strongly depend on the crystalline structures that form

    during processing. [14] Apart from molecular features, the structures formed are

    controlled by processing variables such as flow, pressure, and temperature. These

    variables are often investigated individually, but there is also a remarkable mutual

    influence that is still unexplored. Laying ground for unraveling the relation between

    pressure and flow is the topic of this chapter.

    It is established that shear flow is able to generate precursors of crystallization in

    molten polymers. [58] Flow-induced precursors are metastable domains where stretched

    molecules assume a packing order in between the amorphous and the crystalline states.

    They can be considered as the cradle of flow-induced nuclei as, in suitable conditions,

    e.g. at low temperature, they nucleate and grow into crystalline structures. In this way,

    crystallization kinetics is accelerated, and the final morphology of the material is altered.

    Therefore, monitoring formation and evolution of crystallization precursors is crucial to

    rationalize the effect of flow on the final morphology of polymeric manufactures.

    Precursors are often invisible to scattering techniques such as SAXS and WAXD

    because their structure is neither crystalline nor densely packed or their concentration

    is too low to produce significant effects. In these cases, they are studied indirectly by

    interpreting features (such as kinetics and morphology) of the subsequent crystallization

    process. [915] Interestingly, precursors can be generated and survive for long times

    also when temperature is around and sometimes above the melting point. [9, 1114]

    Especially in these cases, cooling to a lower temperature is a convenient way to trigger

    crystallization. For instance, Hsiao et al. [9] found that, for polyethylene (PE), after

    flow ( = 20 s1, ts = 12 s) at 134C, no structure could be observed with WAXD.

    Nevertheless, after bringing the system at 129 C, oriented crystals started to grow.

    Practically, the cooling step should be fast and the final temperature not too low in order

    to keep the nucleation and growth separated on the time resolution of the experiment.

    This is commonly achieved by cooling with a liquid or gas medium. Small lab-scale

    devices, such as the commercially available Linkam shear cell, can reach cooling rates

    up to 30 C/min. Whereas, flow cells designed for strong flows and high pressures are

    typically restricted to lower values as they are constructed out of relatively large masses

    of metal. For example, the average cooling rate for the original design of the Multi-Pass

    Rheometer (MPR) is 1 C/min. [16]

    When performing flow around the melting point followed by cooling to a lower

    temperature, the lifetime of precursors becomes an important parameter as for long

    cooling times (compared with the lifetime); a fraction of the precursors dissolves and

    thus mitigates the apparent effect of flow on crystallization. Balzano et al. [17] suggested

  • Chapter 3 25

    that, shortly after flow unstable structures dissolve on the time scale of the reptation of

    the longest molecules of the melt. Therefore, experimental protocols are required that

    allow for studying the details (number, size, morphology, and dynamics) of precursors

    in the very early stages while having control over their relaxation behavior. The goal

    is to separate the nucleation from the growth step. In this light, cooling directly to

    room temperature is often not a valid option since the fast overgrowth of crystalline

    structures would obscure nucleation. On the other hand, performing experiments at

    high temperature, close to the nominal melting temperature, seems a better solution

    as growth is very slow in these conditions. The drawback is that structures created at

    high temperatures are very sporadic and easily fall beyond the detection limits even of

    high-resolution methods such as synchrotron X-ray scattering.

    An alternative way to perform controlled quenching is to utilize a pressure quench. In

    other words, apply pressure and, in this way, shift the melting temperature (according

    to the ClausiusClapeyron relation) and thus effectively increase the undercooling

    without actually changing the sample temperature. This methodology can be directly

    implemented in certain slit-flow devices such as the one developed in Eindhoven [18]

    (mounted on the MPR) and offers some clear advantages:

    The cooling step can be instantaneous.

    Temperature gradients and undershoots can be avoided (especially important for

    thick samples).

    The cooling step is easily reversed by depressurizing so complex undercooling

    histories can be applied simply by varying pressure.

    Pressurization has been used as an alternative way to shift the phase boundary

    in studies on phase separation. [19, 20] In particular, it was found that during phase

    separation of polymer blends the general features of nucleation are independent of

    whether the undercooling is obtained by decreasing temperature or by increasing

    pressure. [20] Therefore, pressure quench provides an effective way to obtain

    undercooling. Moreover, it also reflects the practically important, mutual influence

    between flow and pressure in polymer processing techniques.

    In this chapter, we demonstrate the use of pressure quench to investigate the early

    stages of nucleation of flow-induced precursors. We focus on a bimodal blend of PE

    containing 3 wt% high molecular weight tail to simplify the rheological classification [21]

    of the flow conditions and to enhance the flow-induced formation of crystallization

    precursors. [2228] Only a relatively mild increase of pressure is used in this work,

    since a few degrees of undercooling is already effective to accelerate the flow-induced

    crystallization growth. Moreover, a too high pressure (on the order of kbar) may

    induce the hexagonal phase [29, 30] of PE, which is beyond the scope of this work.

  • 26 Chapter 3

    Theoretical aspects on the origin of flow-induced precursors are not discussed here.

    Details can be found in a previous paper of our group. [31] For example, the issue

    if flow creates precursors directly from the melt or if precursors (so called dormant

    precursors [32, 33]) are always present and flow only changes their size to increase the

    number that can be activated by a quench, was dealt with in detail.

    3.2 Experimental

    3.2.1 Materials

    A bimodal polyethylene blend containing 3wt% of high molecular weight tail was

    used in this work. The ultrahigh molecular weight polyethylene (UHMWPE) has a

    weight-averaged molecular weight Mw = 1480 kg/mol and polydispersity Mw/Mn = 2.

    [34] The linear low molecular weight polyethylene (LMWPE) matrix, supplied by Basell

    Polyolefine GmbH (Frankfurt, Germany), has a Mw = 45 kg/mol and polydispersity

    Mw/Mn = 3. The critical overlap concentration of high molecular weight molecules can

    be calculated with [35, 36]

    c =3Mw

    4piRg23/2NA(3.1)

    where Rg2 is mean-square radius of gyration of chain related to the molecular weight byRg21/2 = 0.46Mw1/2, [37] is the density andNA is Avogadros number. The estimatedcritical concentration of UHMWPE is around 0.35wt%, i.e., much smaller than the

    3wt% in our bimodal blend, meaning that a significant number of entanglements exist

    between the UHMWPE molecules.

    The bimodal system was prepared by solution blending to achieve mixing at a

    molecular scale. The UHMWPE was first dissolved in a xylene solution at 130 C, and

    subsequently LMWPE was added to dissolve, where the concentration of total PEs is

    2.5% with an antioxidant (IRGANOX1010) added at a concentration of 2000 ppm. This

    solution was stirred for 1 h under a nitrogen atmosphere. Next, the hot xylene solution

    was poured into a large excess of stirred cold methanol. The precipitated gel was filtered

    and washed with methanol several times and then dried in vacuum at 80 C for 2 days.

    After further addition of 2000 ppm of antioxidant (IRGANOX1010), in order to avoid

    degradation during sample preparation, the bimodal blend was compression molded at

    160 C into strips to fit the test cell.

  • Chapter 3 27

    3.2.2 Protocol

    The slit flow device developed in Eindhoven [18,38] is an evolution of the Multi-Pass

    Rheometer of Eland Engineering Co Ltd. (UK) [39] (see Figure 3.1). The flow cell is

    specifically designed for online scattering and therefore equipped with diamond windows

    (opening angle of 45). The specimen (L = 200mm, W = 6mm, and H = 1.5mm)

    is confined between two servo hydraulically driven rectangular pistons. An important

    advantage of this slit-flow device is the possibility to impose and release pressure. The

    maximum pressure is 800 bar and the maximum temperature 250 C. The procedures

    for flow, pressurization, and depressurization are illustrated in Figure 3.1.

    (a) (b) (c)

    Figure 3.1: Schematic of the flow device and flow, pressurization, and depressurizationprocedures. Moving directions of the pistons are indicated by arrows. Heatingrods and the heating/cooling channels are not shown.

    Moving the two pistons in the same direction introduces a shear flow to the sample

    (Figure 3.1a), while moving the two pistons toward or away from each other will

    pressurize (Figure 3.1b) or depressurize (Figure 3.1c) the sample, respectively, without

    causing any flow at the observation point (center of the slit). The pressure difference

    during flow is recorded by means of two pressure transducers.

    The experimental protocol used in this chapter is shown in Figure 3.2. The polymer

    in the test cell is first heated up and annealed at 190 C for 10 min to erase the memory

    of previous thermal and mechanical histories and then cooled to 134 C at which shear,

    pressurization, and depressurization are performed. To avoid temperature fluctuations,

    the cell is stabilized by means of an oil bath, while the top and bottom barrels are

    always kept at high temperature (190 C) in order to have the pressure transducers

    functioning properly. Next, flow is imposed with a piston speed of 15mm/s for 0.8 s.

  • 28 Chapter 3

    time

    pre

    ssure

    shear

    tem

    pera

    ture

    ta

    Figure 3.2: Experimental protocol. The annealing time ta is between cessation of flow andpressurization.

    The shear rate and Weissenberg number (Wi) distributions over the slit thickness were

    estimated considering the rheology of the material and are shown in Figure 3.3 (see also

    the Appendix). The shear rate varies nonlinearly from 0 at the slit center to a maximum

    of 67 s1 at the wall. The Weissenberg number, Wi = Rouse, is used to estimate the

    molecular stretch induced by flow. When Wi > 1, molecules become stretched, and

    this is a requirement for the creation of precursors. [40] The longest Rouse times of

    UHMWPE and LMWPE at 134 C are 4.9102 and 4.5105 s (see section 3.3.3 forthe calculation), respectively, so the Weissenberg number for molecular stretch at the

    wall is 3.3 for HMW tail but nearly zero for LMW matrix.

    In order to obtain good filling of the slit and rule out wall slippage, a reference

    pressure of 50 bar was kept on the specimen. The average pressure during flow,

    PAverage = Plower +12Psensor, where Plower and Psensor are the lower pressure (around

    15 bar) and the pressure difference between the two pressure sensors (see Figure 3.4),

    is around 70 bar totally, so the influence of flow on the pressure level is negligible.

    A pressure of 300 bar was chosen as the elevated pressure level to increase the melt

    undercooling. Figure 3.2 also shows that pressurization can be applied right after flow

    (ta = 0) or after annealing (ta = 22 min) with the same flow. Both protocols were

    employed in this work. The former (ta = 0) is used to see if any precursors were formed

    during flow, and the latter (ta = 22 min) is to show how precursors develop with time.

  • Chapter 3 29

    centerwall wall

    0.15 0.30 0.45 0.60 0.75

    0

    1

    2

    3

    4

    LMWPE matrix

    distance from center (mm)

    UHMWPE

    We

    issen

    be

    rg n

    um

    be

    r

    0.75 0.60 0.45 0.30 0.15 0.00

    0

    20

    40

    60

    80

    distance from center (mm)

    she

    ar

    rate

    (1/s

    )

    Figure 3.3: Distributions of shear rate and Weissenberg numbers for stretching HMW tailand LMW matrix over the slit thickness direction.

    3.2.3 X-ray characterization

    X-ray measurements were carried out at the Dutch-Belgian (DUBBLE) beamline

    BM26B of the European Synchrotron Radiation Facility (ESRF) in Grenoble, France,

    with a wavelength of 0.95 A. [41] A Frelon detector with a resolution of 2048 2048pixels of 48.8m 48.8m, and placed at a distance of 0.195m, was used for wide-angle X-ray diffraction (WAXD). The measuring time of each WAXD frame, including

    exposure and readout of data, was 8.5 s.

    WAXD images were processed with the software package FIT2D to obtain intensity

    versus scattering angle (2) profiles. Crystallinity was calculated after deconvolution

    of the total intensity scattered by the crystalline (Acrystal) and amorphous (Aamorphous)

    domains:

    X =Acrystal

    Acrystal + Aamorphous 100% (3.2)

    Flow-induced crystallization depends on the strength of the local flow field. In the

    slit geometry, the shear rate changes from a minimum in the center to a maximum at

    the wall, as shown in Figure 3.3, so shear-induced crystalline layers will form mostly

    next to the two walls. X-ray going through the whole sample thickness will scatter from

    the two shear layers, and therefore, values calculated in this way do not represent the

    real crystallinity in certain specific layer but rather an apparent crystallinity averaged

    over the optical path of the X-ray beam through whole sample.

  • 30 Chapter 3

    3.3 Results and Discussion

    3.3.1 Reference experiment: no pressure quench after flow

    The formation and development of flow-induced precursors are investigated with a

    shear pulse ( = 67 s1, ts = 0.8 s) under 50 bar at 134C. First of all, the structural

    development during flow was examined through the rheological response of the polymer

    melt. Figure 3.4 shows the evolution of pressure difference during flow, which first

    increases rapidly, subsequently shows a small overshoot, and reaches a steady-state

    plateau. This behavior can be related to the usual nonlinear rheological response of

    the sheared molten polymer. Scelsi et al. [42] reported a buildup of pressure difference

    (HDPE, 130 C, Wi for stretch is estimated around 60) during flow in a channel with

    a contraction in the middle section and associated it to continuous crystal formation

    concentrated in the region of slit exit. Such a pressure difference buildup during flow

    was not observed in our experiments (134 C, Wi = 3.3). Therefore, our results imply

    that crystals do not form during flow or that the viscosity does not change even if

    precursors form (as discussed later).

    0.0 0.2 0.4 0.6 0.8

    0

    20

    40

    60

    80

    100

    120

    140

    160

    pressure difference (bar)

    time (s)

    experiment 1

    experiment 2

    Figure 3.4: Pressure difference during flow. Two experiments are shown to demonstrate thereproducibility of the flow history.

    Based on the nano- and mesoscale structures of crystalline planes and density

    differences, both SAXS and WAXD were employed to examine, at different length scales,

    structural aspects of flow-induced precursors. However, in-situ synchrotron SAXS and

    WAXD data acquired after flow did not show any observable signal, which confirms

    that no detectable crystallization takes place neither during flow nor within the next 20

    min when the sample is kept isothermal (data not shown since no detectable signal was

    found). More in detail, despite the Wi number for stretching the HMW tail is larger

  • Chapter 3 31

    than unity in the region close to the wall (see Figure 3.3); neither crystalline nor densely

    packed scatterers are observed. This can be a consequence of a low concentration of

    precursors, i.e., below the experimental detection limit. Therefore, crystallization has to

    be triggered to reflect the features of precursors. For PE, WAXD is sufficient to fulfill

    this purpose of characterizing crystallization features (i.e., orientation, crystallinity,

    twisted lamellae) which will be shown in the following results. Therefore, we use only

    WAXD [43] to track PE crystallization and show that a pressure quench is a suitable

    method to visualize precursors even at these very low concentrations.

    3.3.2 Pressure quench after flow

    In order to investigate the formation of flow-induced precursors, after the application

    of the same flow as in section 3.3.1, a pressure quench, with the pressure history shown

    in Figure 3.5, is performed. The pressure is raised linearly from the reference pressure

    of 50 to 300 bar without overshoot. The synchronous top and bottom pressures indicate

    that the pressure field is homogeneously distributed over the specimen and that there

    is no flow at the observation point (center of the slit). With pressure, undercooling is

    generated and immediately triggers crystallization, illustrated by the WAXD images in

    Figure 3.6.

    0 10 20 30

    0

    100

    200

    300

    pressure - top

    pressure - bottom

    pressure (bar)

    time (s)

    Figure 3.5: Pressure profile during a pressure quench.

    As soon as the maximum pressure is achieved, arched (110) and (200) reflections

    appear in the equator direction (Figure 3.6a), indicating the formation of the

    orthorhombic unit cell with c-axis aligned in the flow direction. According to Keller

    et al. [6], the diffraction pattern of Figure 3.6a can be produced both by the extended

    chain crystals of shish and by the folded chain crystals of untwisted lamellae.

  • 32 Chapter 3

    (a) tc = 0s (b) tc = 8.5s (c) tc = 17s (d) tc = 102s

    Figure 3.6: 2DWAXD patterns under a pressure quench after flow. Flow is along the verticaldirection. The reference time is the moment when 300 bar is reached.

    For the purpose of this work, it is not essential to make such a distinction. The most

    important observation is the presence of highly oriented crystals. In fact, as the sample

    is pressurized in absence of extra flow, these oriented crystals originate from deformed

    molecules involved in precursors induced by the flow during the shear step.

    As crystallization proceeds, the (110) reflection broadens in the azimuthal direction

    (Figure 3.6b) and, at tc = 17 s, splits into off-axis diffraction together with the clear

    appearance of off-axis (200) diffractions (Figure 3.6c). These off-axis (110) diffractions

    are the result of the lateral growth of twisted lamellae. During this process, the

    lamellar propagation direction holds perpendicular to the flow direction, but crystal

    units rotate along the b-axis. [6] Twisting leads to randomization in the orientation

    of c-axes. Keller et al. [6, 44] proposed two extremes of lamellar orientation for flow-

    induced PE crystallization: Keller/Machin I and Keller/Machin II. The former

    (KM-I) corresponds to the fully twisted lamellae producing off-axis (110) and meridional

    (200) diffractions. The latter refers to the flat, nontwisted lamellae (all c-axes parallel

    with flow direction) producing equatorial (110) and (200) diffractions. In the transition

    from one to another an intermediate state is observed showing the off-axis (110) and

    (200) patterns. [6, 45] Therefore, the azimuthal features of the (200) diffraction reflect

    the lateral growth of lamellae. [6, 9, 45]

    In a later stage, the diffraction of isotropic structures is also observed, e.g. isotropic

    (110), (200) diffractions at 102 s (see Figure 3.6d). These two crystallization processes,

    fast oriented crystallization and slow isotropic crystallization, characterized by different

    degrees of orientation and time scales can be rationalized by considering that the

    inhomogeneous flow field in the slit gives rise to inhomogeneous molecular stretch as

    shown by Figure 3.3. At the wall, where the stress reaches a maximum, molecules

    experience the largest stretch and many flow-induced precursors are generated. These

    are the domains where highly oriented crystals are formed upon pressurization. Toward

    the centerline, molecules experience little or no stretch and the probability of forming

    oriented precursors vanishes. In this part of the sample, isotropic crystallization can

    be triggered purely by pressure. The WAXD images (Figure 3.6a-d) are indicative of

  • Chapter 3 33

    structures averaged over the optical path of the X-ray beam (thickness of the samples),

    and therefore, they contain information on both these two processes.

    Next, the effect of precursors on crystallization kinetics is illustrated in Figure 3.7a

    where the apparent crystallinity is plotted as a function of time for shear-induced

    crystallization under a pressure quench and two reference experiments (shear-induced

    crystallization without pressure quench and quiescent crystallization with pressure

    quench). It is evident that when shear is applied, crystallization starts as soon as

    the 300 bar pressure is reached. On the other hand, without shear, crystallization

    starts only after 25 s from reaching 300 bar. Eventually, without applying pressure,

    crystallization of the sheared polymer does not occur in the experimental time window

    of 20 min. The results show that there is a remarkable interplay between flow and

    pressure on crystallization kinetics and morphology.

    0 20 40 60 80 100

    0

    5

    10

    15

    flow + Pressure Quench

    quiescent + Pressure Quench

    flow

    apparent crystallinity (%)

    time (s)

    (a) (b)

    Figure 3.7: (a) Crystallinity evolution for shear-induced crystallization with pressure quench(), quiescent crystallization with pressure quench (), and shear-inducedcrystallization without pressure quench (). (b) 2D WAXD pattern of quiescentcrystallization under pressure quench at tc = 34 s.

    From the above results, we conclude that oriented precursors are generated by the

    flow, but prior to pressurization, they are invisible to X-rays (see section 3.3.1). It is

    clear that a pressure quench effectively provides the additional undercooling that triggers

    crystallization by raising the equilibrium melting temperature. The equilibrium melting

    temperature of PE at 1 bar is T 0m = 414.6 K. [46] Its increase with pressure, described

    by the Clausius-Clapeyron relation, can be approximated by T (p) = T 0m+(dTpm/dp)p,

    where dT pm/dp = 35.2 K/kbar. [46] In this way, the equilibrium melting temperatures

    corresponding to 50 and 300 bar are calculated as 416.4 and 425.2 K, respectively.

    With pressurization from 50 to 300 bar, the undercooling T pexp = T pm Texp increases

  • 34 Chapter 3

    from T 50bar134 9K to T 300 bar134 18K (experimental temperature 407.15 K). Thus, apressure quench from 50 to 300 bar increases the undercooling by 9 K within 25 s. Such a

    pressure quench, around 22 K/min, is comparable to a temperature quench in small lab-

    scale device. On the other hand, it is much more efficient than lowering the temperature

    of such large metal device with a cooling medium, especially since the cooling rate

    decreases when approaching the target temperature, and thus it usually takes much

    longer time (minutes) to stabilize at the desired undercooling (see the Appendix).

    The azimuthally homogeneous (110) diffraction in Figure 3.7b shows that pressurizing

    does not cause flow at the observation point which might have influenced crystallization.

    Concluding, oriented crystals (Figure 3.6) and faster kinetics (Figure 3.7a) show

    that precursors can form in a sheared melt, where only the HMW chain is stretched.

    This demonstrates that the HMW tail determine the formation precursors, consistent

    with the findings of Mykhaylyk et al. [15] A pressure quench triggers crystallization

    and effectively lightens up these flow-induced precursors which could not be detected

    otherwise for these conditions (134 C, 50 bar).

    3.3.3 Pressure quench after annealing

    The effect of partial dissolution of flow-induced precursors can be semi-qualitatively

    estimated by introducing an annealing step (during which pressure is kept at 50 bar)

    between the flow pulse and the pressure quench. During the annealing step, whose length

    is arbitrarily fixed at 22 min, no structural changes were observed by SAXS/WAXD

    (data not shown). Figure 3.8 shows the 2D WAXD patterns after applying a pressure

    quench to the sheared and annealed specimen. The equatorial (110) diffraction in Figure

    3.8a indicates the survival of orientation in the precursors with annealing.

    (a) tc = 0s (b) tc = 8.5s (c) tc = 34s (d) tc = 93.5s

    Figure 3.8: 2D WAXD patterns of structures after flow, 22 min annealing under 50 bar anda final pressure quench. Flow is along the vertical direction.

    The difference with crystallization without annealing (section 3.3.2) is the meridional

    (200) diffractions (Figure 3.8c) that arises from a higher degree of randomization in the

  • Chapter 3 35

    c-axes of the unit cells. [6] This arises from the fact that, in the annealed sample, the

    lamellae can grow laterally over a longer distance; i.e., their nucleation points are further

    apart than in the nonannealed sample and also means that the annealed sample has a

    lower nucleation density.

    This observation suggests that the lifetime of some of the flow-induced precursors is

    longer than 22 min (at 134 C and 50 bar), whereas for others it is not.

    The question is whether the long lifetime of precursor relates to the rheological time

    scales of the material. Considering the molecular weight between entanglements Me =

    828 g/mol, [47] the numbers of entanglements per chain, Z, for HMW tail and LMW

    matrix are 1787 and 54, respectively. Because of the significant degree of overlap of long

    molecules and of the low StruglinskiGraessley number Gr = ZHMW/Z3LMW = 0.01,dynamic tube dilution can be neglected. [47] In other words, the relaxation time of

    HMW molecules is not reduced by the LMW matrix. According to the tube model,

    the Rouse time Rouse and reptation time D, responsible for stretch and orientation

    relaxations, can be calculated from

    Rouse = eZ2 (3.3)

    D = 3eZ3

    (1 1.51

    Z

    )2(3.4)

    where e is the entanglement equilibration time, around 7109 s for PE at 190 C. [47]Without considering molecular weight distribution and the effect of 50 bar pressure,

    the estimated reptation times of the HMW tail and LMW matrix at 134 C (Ea =

    21.8 kJ/mol) are 243 and 0.005 s, respectively, and the Rouse times are 4.9 102and 4.5 105 s, respectively. The striking feature is that the longest relaxation time,DHMW 4 min (predicted with Mw), is much shorter than the annealing time, 22min.

    Concerning precursor relaxation, previous experimental studies [17, 38, 48] showed

    that the relaxation of the most unstable shear-induced bundles, observable with

    SAXS, follows the reptation dynamics of the longest chains whereas stable precursors

    survive on time scales much longer than the rheological ones.

    Systematical studies on relaxing shear-induced nucleation precursors that are

    invisible to SAXS were done by Alfonso and co-workers. [14] They found that, in iPP,

    at temperatures as high as 190 C (above nominal melting temperature but not beyond

    equilibrium melting temperature), nucleation precursors can survive much longer than

    the longest rheological relaxation times, and that the characteristic survival time can

    be increased by a stronger or longer flow. The change in relaxation times implies that

    other effects are determining the dissolution of precursors. This idea is supported by the

    different relaxation behaviors of the flow-induced helices in iPP as observed by An et

  • 36 Chapter 3

    al. [49] They suggested that interactions between flow-induced helices of iPP dominate

    the dissolution rates and the helices with interactions relax slower than those without.

    Considering the long lifetime of X-ray unobservable oriented precursors, in our

    results, it is reasonable to infer that interactions between PE chain segments

    (comparable to very local crystallization events) contribute to the long lifetime beyond

    the rheological times. According to the classical nucleation theory, a precursor below the

    critical size tends to relax; the interaction decreases the relaxation kinetics. Once the

    total interaction (contributed by the increased volume as well) is sufficient, the volume

    free energy overcomes the surface energy, and the precursors can develop to nuclei that

    are stable. Long-term stable precursors were also observed by Mykhaylyk et al. [15] for

    hydrogenated polybutadiene; the precursors survived within the experimental window

    which was as long as 10 h. Interestingly, in our experiments, when these segmental

    interactions are initiated during flow, they do not cooperate significantly to change the

    viscosity of the whole melt as observed from the pressure difference in the presence of

    precursors (see Figure 3.4).

    0

    10

    20

    30

    40

    50

    60

    70

    80

    90

    10

    20

    30

    40

    A

    z

    i

    m

    u

    t

    h

    a

    l

    a

    n

    g

    l

    e

    (

    o

    )

    I

    n

    t

    e

    n

    s

    i

    t

    y

    (

    a

    .

    u

    .

    )

    time

    0s

    110s

    51s

    (a)

    0

    10

    20

    30

    40

    50

    60

    70

    80

    90

    10

    20

    30

    40

    A

    z

    i

    m

    u

    t

    h

    a

    l

    a

    n

    g

    l

    e

    (

    o

    )

    I

    n

    t

    e

    n

    s

    i

    t

    y

    (

    a

    .

    u

    .

    )

    time

    0s

    85s

    (b)

    Figure 3.9: Change in the azimuthal distribution of (200) diffraction during pressure quenchcrystallizations for (