PP Crystallization

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 134

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 LATEX 2 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 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 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
1 Introduction
1.1 Background . . . . . . . . . .
1.2 Structure-properties relation .
1.3 Processing-structures relation
1.4 Scope of the thesis . . . . . .
References . . . . . . . . . . .

ix
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

1
1
2
4
5
6

2 Using rheometry to determine nucleation density in a colored system


containing a nucleating agent
2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2 Experimental . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.1 Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.2 Rheological characterization . . . . . . . . . . . . . . . . . . . .
2.2.3 X-ray characterization . . . . . . . . . . . . . . . . . . . . . . .
2.3 Data analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.4 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.4.1 Determination of the nucleation density for NA-iPP . . . . . . .
2.4.2 Reproducibility . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.4.3 Effect of mild flow . . . . . . . . . . . . . . . . . . . . . . . . .
2.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

9
10
11
11
11
12
12
14
14
16
17
20
20

3 Pressure quench of flow-induced crystallization precursors


3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.2 Experimental . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.2.1 Materials . . . . . . . . . . . . . . . . . . . . . . . .
3.2.2 Protocol . . . . . . . . . . . . . . . . . . . . . . . . .
3.2.3 X-ray characterization . . . . . . . . . . . . . . . . .
3.3 Results and Discussion . . . . . . . . . . . . . . . . . . . . . .
3.3.1 Reference experiment: no pressure quench after flow .
3.3.2 Pressure quench after flow . . . . . . . . . . . . . . .

23
24
26
26
27
29
30
30
31

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

Contents

vi

3.3.3 Pressure quench after annealing . .


3.3.4 Inverse quench by depressurization
3.4 Conclusions . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . .
Appendices
. . . . . . . . . . . . . . . . . . . . . . . . .
4 Short-term flow induced crystallization in
how short is short?
4.1 Introduction . . . . . . . . . . . . . . . . . .
4.2 Experimental . . . . . . . . . . . . . . . . .
4.2.1 Material . . . . . . . . . . . . . . .
4.2.2 Methods . . . . . . . . . . . . . . .
4.2.3 Optical microscopy . . . . . . . . .
4.3 Results . . . . . . . . . . . . . . . . . . . . .
4.3.1 Rheological evolution during flow .
4.3.2 Structural evolution . . . . . . . .
4.4 Discussion . . . . . . . . . . . . . . . . . . .
4.4.1 SAXS delay time . . . . . . . . . .
4.4.2 Implications of the viscosity rise . .
4.4.3 Conditions of the viscosity rise . . .
4.5 Conclusions . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . .

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

34
38
39
39
42

isotactic polypropylene:
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

5 The influence of flow induced precursors and nuclei on


of isotactic polypropylene
5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . .
5.2 Experimental . . . . . . . . . . . . . . . . . . . . . . . .
5.2.1 Material . . . . . . . . . . . . . . . . . . . . . .
5.2.2 Flow device . . . . . . . . . . . . . . . . . . . .
5.2.3 X-ray characterization . . . . . . . . . . . . . .
5.2.4 Birefringence . . . . . . . . . . . . . . . . . . .
5.3 Results and Discussion . . . . . . . . . . . . . . . . . . .
5.3.1 Flow-induced distinguishable nuclei/precursors .
5.3.2 Isothermal crystallization . . . . . . . . . . . . .
5.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . .
6 High-stress shear induced crystallization in
and propylene/ethylene random copolymers
6.1 Introduction . . . . . . . . . . . . . . . . . . .
6.2 Experimental . . . . . . . . . . . . . . . . . .
6.2.1 Materials . . . . . . . . . . . . . . .
6.2.2 Flow device . . . . . . . . . . . . . .
6.2.3 X-ray characterization . . . . . . . .
6.2.4 Optical microscopy . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

45
46
48
48
48
50
50
50
52
56
57
59
60
62
62

crystallization
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

65
66
68
68
68
68
72
73
73
75
87
87

isotactic polypropylene
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

89
90
91
91
92
92
93

Contents

6.3
6.4

6.5

Depth sectioning method . . . . . . . . . . . . .


Results and Discussion . . . . . . . . . . . . . .
6.4.1 iPP homopolymer . . . . . . . . . . . .
6.4.2 Propylene/ethylene random copolymers
6.4.3 Quantification of nuclei . . . . . . . . .
Conclusions . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . .

vii

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

7 Conclusions and recommendations


7.1 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.2 Recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.

93
96
96
99
102
106
106

109
. 109
. 111
. 113

Samenvatting

115

Acknowledgements

119

Curriculum vitae

121

List of publications

123

Summary

Semi-crystalline polymers cover over two thirds of commercial applications of


polymeric materials. With most industrial processing technologies, semi-crystalline
polymers 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 shape
geometrically complex parts as with injection molding. The deformation not only
accelerates the crystallization kinetics but also can cause crystalline structures to change
from isotropic spherulites to the highly oriented cylindrical structures (e.g. shishkebab) which determine the ultimate properties. Therefore, understanding the interplay
between flow fields and the resulting crystalline structures is of great importance. It
enables the design of processing procedures and tailoring final product properties. The
objective of this thesis is to reveal how a flow field affects the crystal structures by
studying the early stages of the crystallization process which are dominated by the
creation of shear-induced precursors and nuclei.
Polymer crystallization involves two steps: nucleation and growth. Nucleation
provides quantitative (more or less) and qualitative (isotropic or oriented) nuclei and
these 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 (nucleating
agents) 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 nuclei
is the main work of this thesis.
According to the resulting morphology, nuclei can be divided into two groups: pointlike nuclei and fibrillar nuclei. The former give rise to spherulites, the latter mainly
induce oriented structures like the well-known shish-kebab; i.e. fibrils with transverse
lamellae. Oriented nuclei can be further classified into row nuclei and shish nuclei by
whether they can be directly observed with X-ray characterizations. In the first part of
this thesis, the creation of point-like nuclei is studied for an isotactic polypropylene with
a 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 size
so small that optical microscope is not suitable to count nuclei numbers. Therefore, a
suspension-based rheological model is used to quantify the nuclei density.

ix

Summary

In the second part, precursors of row nuclei in a bimodal Polyethylene (PE) are
studied. Since row nuclei cannot be observed directly by X-ray characterizations, the
crystallization step is triggered by increasing the pressure which raises the equilibrium
melting temperature and thus the under-cooling: a so-called pressure quench. It is
shown that shear-induced precursors can be generated at a temperature close to the
nominal 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 is
found that a critical shear rate exists for the formation of shish nuclei within very short
flow times (0.25 s). Shish precursors are formed during flow and these were found to
develop into shish afterwards, during or after flow depending on the flow strength.
Besides the external effects, the nuclei formation is also influenced by the molecular
structure. Therefore, in the last part, the effect of the molecular architecture on
shear-induced crystallization is studied for isotactic polypropylene (iPP) and two
propylene/ethylene random copolymers with varying ethylene monomer contents. These
three grades had very similar rheological behavior; there is only a small but important
difference in the longest relaxation time. Flow enhanced nucleation density was found
to be the lowest for the iPP homopolymer which is due to the shorter longest relaxation
time. However the reduced growth rate due to the added ethylene monomer leads
to slower overall crystallization kinetics of the random copolymers compared to the
homopolymer 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

Chapter 1

Mtonne

300

2010: 265
2009: 250

250

2002: 204

200
150

1989: 99

100
1976: 47

50
1950: 1.5

0
1950

1960

1970

1980

1990

2000

2010

year

Figure 1.1: World plastics production between 1950 and 2010. Data are obtained from
reference [3]. Only continuous growth is shown and temporary fluctuation is not
included.

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

Chapter 1

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. Semicrystalline 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 semicrystalline 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 processingrelevant conditions.

Chapter 1

70 x 70 x 1 mm

injection of
polymer

Figure 1.3: Variation in microstructure over the thickness in an injection molded HDPE
plate. Right column shows the resulting mechanical properties of three
specimens 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

(a)

(b)

(c)

Figure 1.4: Typical morphologies of: a) less and large spherulites; b) more but small
spherulites; 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-

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 undercooling, 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

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

Chapter 2

10

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

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
0.2 wt%. 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 8 mm. 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

11

Chapter 2

12

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 20041335
pixels of 44 m 44 m was placed at 178 mm. The wavelength used was 1.033
A and
the 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 fG = G /G0 can be obtained from
A fG 2 + B fG + 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 fG .
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 ,




4
3
3
(t, T ) = 1 exp N(T )G (T )t
(2.2)
3
from which the nucleation density can be obtained:
N(T ) =

3 ln (1 )
4G3 (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


U
G(T ) = G0 exp
R (T Tg + T )


G Tm2 (Tm + T )
exp
2T 2 (Tm T )

(2.4)

(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.

Chapter 2

14

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.
8

10

G' (Pa)

10

10

10

145

10

147

148

151

10

100

1000

10000

time (s)

Figure 2.2: Time evolution of the storage modulus for NA-iPP samples crystallized at
different 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

1.0

space filling

0.8

0.6

0.4

0.2

145

147

148

151

0.0
10

100

1000

10000

time (s)

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

(110)

(-131)
(040)

Intensity (a.u.)

151

148

147

145

10

(111)

(130)

11

12
2

13

14

15

( )

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


temperatures.

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 is
relatively 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

Chapter 2

16

22

10

10

20

-3

(m )

145

10

18

147
148

10

16

10

151

14

0.0

0.2

0.4

0.6

0.8

1.0

space filling

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

19

10

17

15

10

-3

(m )

10

NA+iPP
iPP

13

10

11

10

130

135

140
temperature

145
(

150

155

C)

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 1011 m3 , while we
obtain the (interpolated) value N = 6 1012 m3 . 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.

10

17

-3

(m )

10

19

15

10

st

10

13

st

nd

nd

10

11

130

135

140

145

temperature

Series NA+iPP
Series iPP
Series NA+iPP
Series iPP

150

155

C)

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

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

Chapter 2

18

to determine space filling for crystallization after shear. Moreover, the results can be
compared to those of Housmans et al. [9].

10

G' (Pa)

10

10

10

shear time (s)


10

10

100

1000

10000

time (s)

Figure 2.8: Time evolution of the storage modulus for NA-iPP under quiescent conditions
and 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, )
= sHMW(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 aT Wis as a
criterion to compare experiments, which scales as a2T , since sHMW in Eq. 2.6 scales
with aT as well. From the linear viscoelastic data and the stretch relaxation time sHMW
reported by Housmans et al. [9], we find Wis (148 C) = 13, Wis (151 C) = 12, and
aT (151 C)Wis (151 C)
= 1.1 ,
aT (138 C)Wis (138 C)

(2.7)

where Wis (138 C) = 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 1016 m3 , six times higher than
7 1016 m3 for quiescent crystallization, i.e., adding typically 1017 m3 nuclei. This

Chapter 2

19

10

18

17

-3

(m )

10

10

16

151
148

10

15

shear time (s)

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

10

17

-3

(m )

10

19

15

10

10

13

1
2
2

10

st

st

st

nd

nd

Series NA+iPP
Series iPP
Series shear
Series NA+iPP
Series iPP

11

130

135

140

145

temperature

150

155

C)

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 1013 m3 nuclei. This indicates that, in the presence of nucleating
agent, 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.

Chapter 2

20

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. Menyh
ard, M. Gahleitner, J. Varga, K. Bernreitner, P. Jaaskelainen, H. ysd, and B. Pukanszky. 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

[23]
[24]
[25]
[26]

[27]

[28]
[29]
[30]
[31]
[32]
[33]
[34]
[35]
[36]

[37]

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


M. Avrami. Journal of Chemical Physics 8:212224, 1940.
U. R. Evans. Transactions of the Faraday Society 41:365374, 1945.
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.
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.
K. Boutahar, C. Carrot, and J. Guillet. Journal of Applied Polymer Science 60(1):103114, 1996.
K. Boutahar, C. Carrot, and J. Guillet. Macromolecules 31(6):19211929, 1998.
R. I. Tanner. Journal of Non-Newtonian Fluid Mechanics 102:397408, 2002.
R. I. Tangger. Journal of Non-Newtonian Fluid Mechanics 112(23):251 268, 2003.
N. Van Ruth, J. Vega, S. Rastogi, and J. Martnez-Salazar. Journal of Materials Science 41:3899
3905.
N. V. Pogodina, H. H. Winter, and S Srinivas. Journal of polymer science Part B-polymer physics
37(24):35123519, 1999.
S. Coppola, S. Acierno, N. Grizzuti, and D. Vlassopoulos. Macromolecules 39(4):15071514,
2006.
Y. P. Khanna. Macromolecules 26(14):36393643, 1993.
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.
W. R. Hwang, G. W. M. Peters, M. A. Hulsen, and H. E. H. Meijer. Macromolecules 39(24):8389
8398, 2006.

21

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 45:4216
4224, 2012

23

Chapter 3

24

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 134 C, 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

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.

25

Chapter 3

26

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 3 wt% 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
4hRg 2 i3/2 NA

(3.1)

where hRg 2 i is mean-square radius of gyration of chain related to the molecular weight by
hRg 2 i1/2 = 0.46Mw 1/2 , [37] is the density and NA is Avogadros number. The estimated
critical concentration of UHMWPE is around 0.35 wt%, i.e., much smaller than the
3 wt% 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 = 200 mm, W = 6 mm, and H = 1.5 mm)
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 depressurization
procedures. Moving directions of the pistons are indicated by arrows. Heating
rods 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 15 mm/s for 0.8 s.

Chapter 3

shear

temperature

28

pressure

ta

time

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

The shear rate and Weissenberg number (W i) 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, W i =
Rouse , is used to estimate the
molecular stretch induced by flow. When W i > 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 4.9 102 and 4.5 105 s (see section 3.3.3 for
the 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 + 21 Psensor , 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.

29

80

60

40

UHMWPE

20

LMWPE matrix

0
0.75

0.60

0.45
0.30
0.15
distance from center (mm)

wall

0.00

0.15

0.30

0.45

Weissenberg number

shear rate (1/s)

Chapter 3

0
0.60

0.75

distance from center (mm)

center

wall

Figure 3.3: Distributions of shear rate and Weissenberg numbers for stretching HMW tail
and 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 2048
pixels of 48.8 m 48.8 m, and placed at a distance of 0.195 m, was used for wideangle 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
100%
Acrystal + Aamorphous

(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.

Chapter 3

30

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 134 C. 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, W i 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, W i = 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).

pressure difference (bar)

160
140
120
100
80
60
40
20

experiment 1
experiment 2

0
0.0

0.2

0.4

0.6

0.8

time (s)

Figure 3.4: Pressure difference during flow. Two experiments are shown to demonstrate the
reproducibility 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 W i 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.

pressure (bar)

300

200

100
pressure - top
pressure - bottom

10

20

30

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.

Chapter 3

32

(a) tc = 0s

(b) tc = 8.5s

(c) tc = 17s

(d) tc = 102s

Figure 3.6: 2D WAXD patterns under a pressure quench after flow. Flow is along the vertical
direction. 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 flowinduced 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.

15
flow + Pressure Quench

apparent crystallinity (%)

quiescent + Pressure Quench


flow

10

0
0

20

40

60

80

100

time (s)

(a)

(b)

Figure 3.7: (a) Crystallinity evolution for shear-induced crystallization with pressure quench
(), quiescent crystallization with pressure quench (), and shear-induced
crystallization without pressure quench (). (b) 2D WAXD pattern of quiescent
crystallization 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 Tm0 = 414.6 K. [46] Its increase with pressure, described
by the Clausius-Clapeyron relation, can be approximated by T (p) = Tm0 + (dTmp /dp)p,
where dTmp /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.
p
With pressurization from 50 to 300 bar, the undercooling Texp
= Tmp Texp increases

Chapter 3

34

50 bar
300 bar
from T134
9 K to T134
18 K (experimental temperature 407.15 K). Thus, a
pressure 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 labscale 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 and
a 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
3
molecules and of the low StruglinskiGraessley number Gr = ZHMW /ZLMW
= 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 = e Z 2

(3.3)

2

1.51
D = 3e Z 1
Z

(3.4)

where e is the entanglement equilibration time, around 7 109 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 4.9 102
and 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, 22
min.
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

Chapter 3

36

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).

40

10

30
20
10

110s

90 80
70 60
Az
im

uth

51s
50
al

40

an

gle

30

20

10

( )

(a)

0s

time

Intensity
(a.u.)

20

Intensity
(a.u.)

30

40

85s

90 80
70 60
Az
im

uth

time
50

al

an

40
gle

30
o

20

10

( )

0s

(b)

Figure 3.9: Change in the azimuthal distribution of (200) diffraction during pressure quench
crystallizations for (a) unannealed and (b) annealed samples.

The structures surviving the 22 min annealing step show some specific features
that can be captured by looking at the (200) reflection. Figure 3.9a shows that the
(200) diffraction of the unannealed sample occurs at low azimuthal angle and moves
toward higher value with time. After about 51 s, the majority of the (200) diffractions
tends to develop around the azimuthal angle of 50 , quite similar to the intermediate

Chapter 3

37

orientation mode. In contrast, for annealed samples, the (200) diffraction is found along
the meridional direction (see Figure 3.9b), indicating a pronounced Keller/Machin I
mode. The data of Figure 3.9 show that, during crystallization without annealing, a
higher fraction of flat lamellae formed in comparison with the annealed sample.
The different morphology must be associated with nuclei density which is larger in
the unannealed sample. This result is consistent with the finding of Keum et al. [9]
that twisted lamellae are more prominent when the shish density is smaller due to
lower shear strength (at the lower shear rate 20 s1 compared with 70 s1 for the
same flow time). Thus, the lower nuclei density in the annealed sample indicates that
some unstable shear-induced precursors relaxed during annealing. The crystallinity
developments of the annealed and unannealed samples are compared in Figure 3.10 with
quiescent crystallization as reference. The lower crystallinity in the annealed sample
confirms the decrease in the total nuclei number, consistent with 2D patterns comparison
between Figures 3.6 and 3.8.
15
un-annealed

apparent crystallinity (%)

annealed
quiescent

10

0
0

20

40

60

80

100

time (s)

Figure 3.10: Crystallinity evolution for () unannealed, () annealed, and () quiescent


crystallization under pressure quench.

Concluding, both stable and unstable precursors are generated by the flow in a
slit with varying flow strength over the thickness. The stable precursors orient and
accelerate the following crystallization while unstable ones disappear during annealing
leading to a larger space for overgrowth and higher fraction of twisted lamellae.
These precursor evolutions, i.e., survival and relaxation, can be distinguished only by
crystallization started with high enough undercooling, which, in this work, is obtained
by a pressure quench.

Chapter 3

38

3.3.4 Inverse quench by depressurization


Finally, we present the results for resetting the pressure from 300 bar back to 50
bar before reaching complete crystallization. Depressurization suddenly eliminates
the increased undercooling (the experimental temperature is kept 134 C). Figure
3.11a shows crystallinity evolutions after depressurization (prior to depressurization
the annealed and unannealed sample have been crystallized for about 100 s). For
both samples, crystallinity decreases but does not vanish completely. The residual
crystallinity in the unannealed sample is around 2%, 4 times larger than that of the
annealed sample (final level around 0.5%). Crystallinity decrease after depressurization
is ascribed to melting of the lamellae due to the reduced undercooling by depressurizing,
which acts as inverse quench. [50]
12
un-annealed

apparent crystallinity (%)

10

annealed

8
6
4
2
0
0

100

200

300

400

500

time (s)

(a)

(b)

(c)

Figure 3.11: (a) Crystallinity evolution of () unannealed and () annealed samples during
melting. t = 0 when pressure reaches 50 bar. The 2D WAXD images of (b)
unannealed sample at 527 s and (c) annealed sample at 408 s during melting.

Cho et al. [51] suggested that PE lamellae may further thicken in an isothermal
process. Thus, when the crystallization time is not long enough for all crystals to finish
thickening, as in our work, the early formed crystals have thicker lamellae compared to
those created in the late stages. The resulting lamellar distribution leads to variation
in thermal stability. Thus, after depressurization the stable (thick) crystals can survive
but unstable (thin) ones melt. As a result, the unannealed sample with more nuclei
has more stable (thicker) crystals than the annealed sample. Figure 3.10 shows that
oriented crystallization sets in before unoriented crystallization; therefore, oriented
crystals experience longer isothermal thickening, and this makes them more likely to
survive in inverse quenching. This is confirmed by the results shown in Figure 3.11c.

Chapter 3

3.4 Conclusions
The pressure quench is a relatively simple way to mimic a temperature quench.
It has a number of advantages compared to thermal quenching (reverse cooling by
depressurizing, avoiding temperature gradients, introducing complex thermal histories).
It was applied for (a) crystallizations under quiescent condition and (b) flow and with
and without subsequent annealing. For quiescent crystallization it is demonstrated that
rising pressure to 300 bar is enough to trigger crystallization. Application on sheared
samples shows that a pressure quench effectively visualizes the shear-induced precursors
which are invisible to X-ray scattering. Crystallization kinetics is faster in this case,
and oriented morphology is observed. The orientation in the outmost layer can survive
annealing for 22 min, but the average nuclei density along the whole sample does relax.
In addition, depressurization before complete crystallization leads to partially melting
of the crystals which is explained by the variation in lamellar stability.

References
[1] H. E. H. Meijer and L. E. Govaert. Progress in Polymer Science 30(8-9):915938, 2005.
[2] B. A. G. Schrauwen, R. P. M. Janssen, L. E. Govaert, and H. E. H. Meijer. Macromolecules
37(16):60696078, 2004.
[3] 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.
[4] M. Kristiansen, M. Werner, T. Tervoort, P. Smith, M. Blomenhofer, and H. Schmidt.
Macromolecules 36(14):51505156, 2003.
[5] 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.
[6] A. Keller and H. W. H. Kolnaar. Flow-induced orientation and structure formation,. In: H. E. H.
Meijer, editor, Processing of Polymers, number 18 in Materials Science and Technology: a
Comprehensive Treatment, 189268. Wiley-VCH, Weinheim, 1997.
[7] G. Kumaraswamy. Journal of Macromolecular Science, Part C: Polymer Reviews 45(4):375397,
2005.
[8] R. H. Somani, L. Yang, L. Zhu, and B. S. Hsiao. Polymer 46(20):85878623, 2005.
[9] K. K. Jong, B. Christian, Z. Feng, and S. H Benjamin. Polymer 48(15):4511 4519, 2007.
[10] L. Fernandez-Ballester, T. Gough, F. Meneau, W. Bras, F. Ania, J. C. Francisco, and J. A.
Kornfield. Journal of Synchrotron Radiation 15:185190, 2008.
[11] G. C. Alfonso and P. Scardigli. Macromolecular Symposia 118(1):323328, 1997.
[12] F. Azzurri and G. C. Alfonso. Macromolecules 38(5):17231728, 2005.
[13] F. Azzurri and G. C. Alfonso. Macromolecules 41(4):13771383, 2008.
[14] D. Cavallo, F. Azzurri, L. Balzano, S. S. Funari, and G. C. Alfonso. Macromolecules 43(22):9394
9400, 2010.

39

40

Chapter 3

[15] O. O. Mykhaylyk, P. Chambon, C. Impradice, J. P. A. Fairclough, N. J. Terrill, and J. Ryan, A.


Macromolecules 43(5):23892405, 2010.
[16] M. R. Mackley and D. G. Hassell. Journal of Non-Newtonian Fluid Mechanics 166(910):421456,
2011.
[17] L. Balzano, N. Kukalyekar, S. Rastogi, G. W. M. Peters, and J. C. Chadwick. Physical Review
Letters 100(4):048302, 2008.
[18] J. W. Housmans, L. Balzano, D. Santoro, G. W. M. Peters, and H. E. H. Meijer. International
Polymer Processing 24(2):185197, 2009.
[19] B. Hammouda, N. P. Balsara, and A. A. Lefebvre. Macromolecules 30(18):55725574, 1997.
[20] A. A. Lefebvre, J. H. Lee, H. S. Joen, N. P. Balsara, and B. Hammouda. Macromolecules
111(13):60826099, 1999.
[21] J. van Meerveld, G. W. M. Peters, and M. Hutter. Rheologica Acta 44(2):119134, 2004.
[22] M. Seki, D. W. Thurman, J. P. Oberhauser, and J. A. Kornfield. Macromolecules 35(7):2583
2594, 2002.
[23] E. L. Heeley, C. M. Fernyhough, R. S. Graham, P. D. Olmsted, N. J. Inkson, J. Embery, D. J.
Groves, T. C. B. McLeish, A. C. Morgovan, F. Meneau, W. Bras, and A. J. Ryan. Macromolecules
39(15):50585071, 2006.
[24] O. O. Mykhaylyk, P. Chambon, R. S. Graham, J. P. A. Fairclough, P. D. Olmsted, and A. J.
Ryan. Macromolecules 41(6):19011904, 2008.
[25] S. Vleeshouwers and H. E. H. Meijer. Rheologica Acta 35(5):391399, 1996.
[26] L. Yang, R. H. Somani, I. Sics, B. S. Hsiao, R. Kolb, H. Fruitwala, and C. Ong. Macromolecules
37(13):48454859, 2004.
[27] S. Acierno, B. Palomba, H. Winter, and N. Grizzuti. Rheologica Acta 42:243250, 2003.
[28] L. Balzano, S. Rastogi, and G. W. M. Peters. Macromolecules 44(8):29262933, 2011.
[29] S. Rastogi, M. Hikosaka, H. Kawabata, and A. Keller. Macromolecules 24(24):63846391, 1991.
[30] H. Hikosaka, K. Tsukijima, S. Rastogi, and A. Keller. Polymer 33(12):2502 2507, 1992.
[31] P. C. Roozemond, R. J. A. Steenbakkers, and G. W. M. Peters. Macromolecular Theory and
Simulations 20(2):93109, 2011.
[32] H. Janeschitz-Kriegl, E. Ratajski, and M. Stadlbauer. Rheologica Acta 42(4):355364, 2003.
[33] H. Janeschitz-Kriegel and E. Ratajski. Polymer 46(11):38563870, 2005.
[34] R. Huang. Immobilization and activation of early- and late-transition metal catalysts for ethylene
polymerization using MgCl2-based supports. Ph.D. thesis, Eindhoven University of Technology,
2008.
[35] P. G. de Gennes. Scaling Concepts in Polymer Physics. Cornell University Press, NY, 1979.
[36] Y. Takahashi, Y. Isono, I. Noda, and M. Nagasawa. Macromolecules 18(5):10021008, 1985.
[37] J. Schelten, D. G. H. Ballard, G. D. Wignall, G. Longman, and W. Schmatz. Polymer 17(9):751
757, 1976.
[38] L. Balzano, D. Cavallo, T. B. van Erp, Z. Ma, J. W. Housmans, L. Fernandez-Ballester, and
G. W. M. Peters. IOP Conf. Series: Materials Science and Engineering 14:01200, 2010.
[39] M. R. Mackley, R. T. J. Marshall, and J. B. A. F Smeulders. journal of rheology 39:12931309,
1995.
[40] R. J. A. Steenbakkers and G. W. M. Peters. Journal of Rheology 55(2):401433, 2011.
[41] 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.
[42] L. Scelsi and M. Mackley. Rheologica Acta 47(8):895908, 2008.

Chapter 3

[43] Both SAXS and WAXD were used to examine if any X-ray observable precursors form during
flow and develop within the following isothermal period. But, only WAXD was further used to
track the crystallizations started with pressure.
[44] A. Keller and M. J. Machin. Journal of Macromolecular Science, Part B: Physics B1(1):4191,
1967.
[45] T. Nagasawa, T. Matsumura, and S. Hoshino. Appl Polym Symp 20:295313, 1973.
[46] B. Wunderlich. Macromolecular Physics, Vol. 3: Crystal melting. Academic Press, London, 1980.
[47] J. M. Dealy and R. G. Larson. Structure and Rheology of Molten Polymers. Hanser Gardner
Pubs, Cincinnati, 2006.
[48] A. W. Phillips, A. Bhatia, P. Zhu, and G. Edward. Macromolecules 44(9):35173528, 2011.
[49] H. An, X. Li, Y. Geng, Y. Wang, X. Wang, L. Li, Z Li, and C. Yang. The Journal of Physical
Chemistry B 112(39):1225612262, 2008.
[50] S. Acierno and N. Grizzuti. Journal of Rheology 47:563576, 2003.
[51] T.Y. Cho, B. Heck, and G. Strobl. Colloid and Polymer Science 282:825832, 2004.

41

Chapter 3

42

Appendices
3A. Temperature cooling of MPR
The measured temperature profile during cooling the flow cell from 132 to 124 C
by circulating 124 C oil is shown in Figure 3.12. The cooling rate decreases when
approaching the target temperature. The temperature cooling takes about 10 min to
obtain the similar undercooling with the pressure quench from 50 to 300 bar at 134 C.

130

temperature ( C)

132

128
126
124
0

100 200 300 400 500 600 700 800


time (s)

Figure 3.12: Temperature evolution in the flow cell during cooling from 132 to 124 C by
circulating 124 C oil.

Chapter 3

43

3B. Calculation of shear rate in the slit channel


A Rheometrics ARES rheometer is used with a plate-plate geometry for smallamplitude oscillatory shear measurements. A strain sweep was done first to determine
the linear region and accordingly the applied strain is set to 10% for subsequent
rheological measurements. The characteristic rheological properties (storage modulus
G , loss modulus G , phase angle and complex viscosity ) are obtained over the
angular frequency range between 0.0025 and 100 rad/s at three different temperatures,
140, 170 and 200 C. The experiments are performed in a nitrogen environment to avoid
polymer degradation. Time-temperature superposition is applied to obtain the master
curves at the reference temperature of 140 C.
The material used is a bimodal PE system. Its master curve of dynamic viscosity
curve is shown in Figure 3.13; the shear thinning behavior doesnt follow a simple Cross
model. A double-cross model is used:
()
=

2
1
+
1 + (K1 )
(1n) 1 + (K2 )
(1n)

(31)

We use the Cox-Merz rule, i.e. the shear dependent viscosity is taken equal to the
dynamic viscosity. The result is shown in Figure 3.13 and the fitting parameters are
listed in Table 3.1.
5

10

experimental data
fitting

(Pa*s)

10

10

10

0.01

0.1

10

100

(rad/s), shear rate (1/s)

Figure 3.13: The master curve of complex viscosity at 140 C (open points) and the fitting
of double-cross model (solid line).

Chapter 3

44

Table 3.1: All fitting parameters for double-cross model.

1 (Pa s)
331131

K1 (s)
2239

2 (Pa s)
661

K2 (s)
0.005

n
0.25

The piston speed is 15 mm/s and the area of cross-section is 61.5 (mmmm), so
the volume flux is 135 mm3 /s. Known the viscosity behavior and volume flux, the shear
rate can be determined by a simple iterative numerical calculation. Some typical values
are listed in Table 3.2 and the shear rate vs. thickness curve is in Figure 3.3.

Table 3.2: The shear rate at different position over slit thickness.

distance from slit center(mm)


0.0375
0.1125
0.1875
0.2625
0.3375
0.4125
0.4875
0.5625
0.6375
0.7125
0.75

shear rate (s1 )


1.0
5.6
11.0
16.9
23.3
30.2
37.5
45.2
53.4
62.1
66.7

Chapter four

Short-term flow induced


crystallization in isotactic
polypropylene: how short is short?

Abstract
The so-called short-term flow is widely applied in experimental flow-induced
crystallization studies in order to separate the nucleation and subsequent growth
processes. The basis of short-term flow is the assumption that structure development
during flow can be minimized and the polymer rheological behavior, i.e. the viscosity,
does not change. In this chapter we explore the validity of this assumption for short
but strong flow and reveal the structure formation in the early stages of crystallization.
Viscosity and structure evolution of an isotactic polypropylene (iPP) melt at 145 C are
measured during short-term flow (0.20-0.25 s) using the combination of a slit rheometer
and fast X-ray scattering measurements. For high enough shear rates ( 240 s1 )
a viscosity rise is observed during flow, i.e. the conditions for short-term flow are
not satisfied. Such a viscosity rise indicates structure formation in the slit which is
considered as the formation of shish or their precursors. With a time delay of about
0.1 s with respect to the viscosity rise the development of shish is observed by means
of time resolved SAXS measurements in the middle of the slit. Depending on the shear
rate these shish are detected during flow (shear rates 400 s1 ) or after flow (400
s1 > shear rates 240 s1 ). For shear rates between 80 and 160 s1 , the viscosity
This chapter is based on : Zhe Ma, Luigi Balzano, Tim van Erp, Giuseppe Portale, Gerrit W. M.
Peters. to be submitted, 2012

45

Chapter 4

46

does not change significantly and, instead of shish, oriented row nuclei are generated.
These flows qualify as short-term flow conditions, but the transient and inhomogeneous
behavior, both in flow and in flow gradient direction, has to be taken into account when
characterizing the flow field in future studies.

4.1 Introduction
Semi-crystalline polymers, especially polyethylene (PE) and isotactic polypropylene
(iPP), are widely used materials because of their low cost, easy processing, good
chemical resistance, etc. These materials are often processed from the molten state and
therefore subjected to flow fields when molded into final products. It is well known that
these flow fields can not only accelerate crystallization kinetics by orders of magnitude,
but also radically change the crystalline morphology from isotropic spherulites to
highly oriented shish-kebab. Such a morphological transition is important since the
morphological building blocks determine the final (mechanical and other) properties
of products. [1, 2] Therefore, a full understanding of the relation between flow fields,
crystallization kinetics and the resulting morphology is required to design processing
procedures for optimal properties.
Initial studies on flow-induced crystallization of polymer melts focused on the
structure evolution during continuous flow fields. [37] Crystallization of polymers is
governed by nucleation and growth and both these processes are influenced by the
flow. [8, 9] When crystals nucleate/grow, the viscosity of the melt increases and this
enhances the effect of flow on crystallization, giving rise to a self-enhancing mechanism.
To simplify this kind of experiments, Janeschitz-Kriegl and co-workers proposed a
short-term shearing method [10], where the shear duration is chosen short enough
so that during flow the effects of crystallization on viscosity and structure changes are
minimized. It is assumed that only nuclei or their precursors are created during flow and
that these structures crystallize and grow after the flow ceases. In this way, the features
of flow-induced nuclei are revealed indirectly by studying the resulting crystallization
kinetics and morphology. This short-term flow has been widely used in studies on flow
induced crystallization in order to separate the nucleation and growth processes. [1019]
Based on the assumption that viscosity is not changed by the flow, the effect of the
flow can be characterized by using flow characteristics and rheological properties. For
instance, the mechanical work [20]
w =

ts

2 (t)dt

(4.1)

with the averaged viscosity , shear rate (t)

and shear time ts , is often considered


as the controlling factor in flow-enhanced nucleation [20] and formation of oriented

Chapter 4

structures [16, 18]. For the constant shear rate applied, often the steady state viscosity
is used. The key assumption to ensure this approach to be useful is that no change in
the rheological properties occurs during flow time.
In more recent work [2123] fast time-resolved experimental methods were used
to investigate structure development, also during a short flow pulse, for conditions
where the Weissenberg number related to the stretch of the high molecular weight
tail (W iRHMWtail ) is large. Kumaraswamy et al. [21] performed slit-flow experiments
with constant wall stress (0.06 MPa) and attributed the observed unusual upturn
in birefringence to the generation of long-lived oriented structures. Balzano et al.
[23] observed the appearance of X-ray diffraction peaks indicating the formation of
crystalline structures during flows shorter than 1 s. These evidences imply that the
short-term flow protocols employed were not sufficiently short to prevent structure
formation, even though the shear time is much less than the characteristic crystallization
time.
These findings raise questions concerning short-term flow: can viscosity change
during flow? If so, under what conditions and how fast and, finally, what is the relation
between a viscosity change and structure formation in these early stages?
In addition, the early stages of shear-induced crystallization, from a structural
point of view, are still under debate. It has been proposed that so called dormant
nuclei already preexist in the amorphous melt and are activated by flow to trigger
crystallization. [24, 25] The unusual birefringence upturn observed by Kumaraswamy et
al. [21, 22] points towards the formation of shear-induced oriented structures [21].
However, X-ray scattering measurements could not resolve these structures at the
relatively high temperatures applied (168 and 173 C). [26] On the other hand, density
fluctuations were observed during extrusion of iPP [27] and this was interpreted as a
process where crystallization is preceded by spinodal-assisted phase separation enhanced
by flow [28]. Balzano et al. [17] found shear-induced bundles generated by flow
and, in line with a more classical view on crystallization, they suggested that the
bundle dimensions determine the subsequent evolution of crystallization or relaxation.
Obviously, a variety of structures can be generated by flow. The current understanding
on shear-induced crystallization, especially concerning the initial stage during flow, is
not yet clear enough to answer the above questions.
The present study focuses on structure formation and viscosity changes (other than
the normal transient response of a start-up flow) during short-term shear flow (max.
0.25 s) and explores the flow strength dependency of these events. For this purpose, a
slit rheometer and fast X-ray scattering measurements are combined to achieve a time
resolution sufficient to resolve the phenomena studied.

47

Chapter 4

48

4.2 Experimental
4.2.1 Material
The material used in this work is a commercial isotactic polypropylene (iPP)
homopolymer (HD601CF) provided by Borealis GmbH, Austria. This iPP has a weight
average molecular mass, Mw 365 kg/mol and a molecular weight distribution Mw /Mn
of 5.4. Its nominal melting and crystallization temperatures are 163 and 113 C,
respectively. [18] A full characterization of the crystallization kinetics of this grade,
both for quiescent and flow enhanced, are given in refs [18, 29, 30].
For sample preparation, the material was first compression molded at 220 C to plates
with thickness of 1.5 mm. Next, strips were machined of H W Ltot = 1.5 6 200
mm3 that fit in the slit flow cell.

4.2.2 Methods
The slit flow cell is operated on a multipass rheometer [31]. The specimen is confined
between two servo-hydraulically driven rectangular pistons that fit tightly in the slit, see
Figure 4.1. When pistons move together in one direction, they impose a shear field to
the polymer melt. The top and bottom barrels are equipped with pressure transducers
(distance between the transducers L = 160 mm) to measure the pressure history in the
slit during flow. The pressure difference P is used to determine the apparent viscosity.
A pair of diamond windows placed in the middle of the flow cell allows for in-situ X-ray
characterization during and after flow.

X-ray
esynchrotron
radiation

ESRF

slit rheometer

Pilatus detector

Figure 4.1: Combined in-situ synchrotron X-ray scattering and slit rheometry.

The polymer in the slit is first heated to 220 C and annealed for 10 min in order

Chapter 4

49

to erase the sample preparation history. Next, it is cooled to 145 C and pressurized
to 50 bar by moving the pistons towards each other. During cooling, the reference
pressure of 50 bar is kept on the sample to prevent shrinkage holes. The sample is then
sheared and subsequently isothermally crystallized at 145 C. To avoid fluctuations, the
temperature of the cell is stabilized by means of an oil bath. On the other hand, the top
and bottom barrels are kept at high temperature (220 C) to ensure proper functioning
of the pressure transducers. After isothermal crystallization, the slit is cooled to room
temperature and the sample is removed for ex-situ analysis.
The flow strength was varied by choosing piston speeds, Vpiston , from 20 mm/s to 140
mm/s and the apparent wall shear rate is calculated by [32]
=

6Q
W H2

(4.2)

where Q is the volumetric flow rate, W is the slit width (6 mm) and H is the slit
thickness (1.5 mm). Apparent wall shear rates range from 80 to 560 s1 . The shear
duration is fixed at 0.25 s for shear rates from 80 to 400 s1 and shortened to 0.23 and
0.20 s for 480 and 560 s1 , respectively, due to limitations in the piston displacement.
The apparent wall shear stress and the corresponding apparent viscosity are given by:
=

HP
2(1 + H/W )L

(4.3)

H2
P

12(1 + H/W )L Vpiston

(4.4)

Both small-angle X-ray scattering (SAXS) and wide-angle X-ray diffraction (WAXD)
were employed to characterize the flow-induced structures.
Synchrotron X-ray
measurements were carried out at the Dutch-Belgian (DUBBLE) beamline BM26B of
the European Synchrotron Radiation Facility (ESRF) in Grenoble, France. [33] The
wavelength used was 1.033
A. Fast acquisitions of SAXS and WAXD were performed
with a Pilatus 1M detector and a Pilatus 300K detector, respectively. This fast step is
carried out at an acquisition rate of 30 frame/s and lasts for 1 s. Both detectors are
triggered by the start of the piston displacement. The Pilatus 1M detector (981 1043
pixels of 172 m 172 m) was placed at a distance of 7.117 m and used for SAXS, the
Pilatus 300K detector (1472 195 pixels of 172 m 172 m) was placed at a distance
of 0.240 m and used for collecting the equatorial part of WAXD images.
Figure 4.2 shows a typical SAXS pattern (so-called streak) which is the result of a
highly oriented structure. The appearance of such streaks means that the structure
formed has a electron density which is different from its surroundings. Their equatorial
distribution implies that the maximum density contrast is perpendicular to the flow
direction, i.e. these fibrillar objects are oriented along the (vertical) flow direction.

Chapter 4

50

q (1/ nm)
az (q)

flow
SAXS streak

Figure 4.2: A typical SAXS pattern with equatorial streaks. The integrating region for
determination of the equatorial intensity ISAXS is indicated. Flow direction is
vertical.

To describe the evolution of these SAXS streaks, we define a SAXS equatorial


intensity ISAXS integrated over the specific equatorial region:
ISAXS =

0.2
0.018

10

I(az, q)daz dq

(4.5)

10

where az is the azimuthal angle and q is the norm of the scattering vector. The scattering
vector q = 4sin()/ is defined as with the scattering angle 2 and the wavelength
of X-ray.

4.2.3 Optical microscopy


Optical microscopy was used to visualize the morphology over the sample thickness
direction and to determine the thickness of final shear layers at different positions in the
slit. Cross-sections of 5 m thick were prepared at a low temperature (approximately
20 C) using a microtone (Leica RM2165) equipped with a glass knife. Two crossed
polarizers are rotated to 45 with respect to the flow direction and optical micrographs
were taken with an Axioplan imaging 2 microscope combined with an AxioCam camera.
The AxioVision software was used to analyze the micrographs and determine the shear
layer thickness.

4.3 Results
4.3.1 Rheological evolution during flow
Figure 4.3a shows the evolution of pressure difference P between the transducers
during shear. This pressure difference scales directly with the wall stress and thus with
the apparent viscosity, see Figure 4.3b. The apparent viscosity is an average over the
material in the flow channel between the two pressure transducers. The kink in the

Chapter 4

51

1.5x10

900

480 s

750

400 s
320 s

600

240 s

P (bar)

160 s
400s

450

-1

-1

-1

-1

-1

-1

80 s

-1

300
240s

-1

160s
80s

150

560s

-1

-1

-1

apparent viscosity (Pa*s)

560s

560 s

-1

480s
3

10

400s
320s
240s
160s

-1

-1

-1

-1

-1

-1

80s

-1

5x10

apparent wall shear rate

0
0.0

0.1

0.2

0.3

0.4

0.01

0.1

(a)

time(s)

time(s)

(b)

Figure 4.3: (a) Pressure difference P evolution during flow for different apparent wall shear
rates. (b) The corresponding transient apparent viscosities. For shear rates of
560 and 480 s1 , some pressure data points are missing just before flow stops.
These points were extrapolated with a linear function, as shown by the dashed
lines in Figure 4.3a.

slope (dP /dtime ) at very short times is an artifact due to a small deviation of the
piston movement during startup. First of all, Figure 4.3a clearly shows that, prior
to any structure formation, the stress is not immediately constant after imposition of
flow. Considering this, we found the apparent wall shear rate to be more suitable to
characterize the strength of flow field than the shear stress which significantly evolves
with time. Secondly, the iPP melt exhibits different behavior with increasing flow
strength; two distinct responses can be distinguished for shear rates 160 s1 and
shear rates 240 s1 .
For relatively low shear rates (80 and 160 s1 ) the pressure difference P first shows
an overshoot, then decreases and, eventually, approaches a steady-state level. This
nonlinear rheological behavior is typical for polymer melts subjected to start-up shear
flow with a constant shear rate. [32] For shear rates 240 s1 , the trend of P is rather
different; after the overshoot, P increases with time rather than leveling off. With
increasing shear rate, this upturn takes place at shorter times and the time-evolution
becomes steeper. Such an unusual trend of P indicates that, during flow, viscosity
changes due to structure formation. Since crystallinity is still rather low (see next
section) we exclude a suspension-like rheological effect (part of the melt changes into
deformable solid particles, i.e. the crystalline structures) and consider the structures
formed to act as physical cross-links at a molecular level. A clear rise of the dynamic
viscosity at low levels of crystallinity was also observed in quiescent crystallization for
a HDPE by Roozemond et al. [34] and it was found that this rise at the early stage
of crystallization could not be captured by a suspension model. From now on, we will
denote this structure as the cross-linking structure.

Chapter 4

52

What kind of structure is able to change the rheology-related behavior of a polymer


melt? Two examples of deviation from the normal rheological behavior during flow
are reported. Scelsi et al. [35] found a buildup of P in high density polyethylene
during flow (the apparent wall shear rate 200 s1 and shear time = 2 s at 130 C)
similar to Figure 4.3a. They associated it to a growing crystalline layer in a slit that is
fed from a reservoir. Kumaraswamy et al. [21] observed a birefringence (proportional
to the stress) upturn during flow at constant pressure, different from the typical melt
rheology, and attributed it to the formation of a highly oriented structure that, in a
later stage, contributes to formation of a skin layer. These highly oriented structures
were examined with in-situ WAXD by Kumaraswamy et al. [12] as well, and a correlation
between these highly oriented structure and crystal formation was found for relatively
low temperatures (141 and 163 C) only. This could not be directly linked to the results
for high temperatures (168 and 173 C). Therefore, the origin of the deviations (viscosity
and birefringence) from the amorphous melt behavior remains unclear.
The absence of a viscosity rise during the short flow period at shear rates of 80 and
160 s1 does not mean that flow-induced structures are absent. Therefore, to reveal the
origin of viscosity rise during strong flows (shear rates 240 s1 ) and probe the potential
structures developing in weak flows (shear rates 160 s1 ), structural investigations
using X-ray scattering were carried out.

4.3.2 Structural evolution


Illustrative SAXS and WAXD patterns, collected during and just after flow, for a
shear rate of 400 s1 are shown in Figure 4.4. The first three patterns are taken during
flow. Interestingly, after 0.23 s from the beginning of flow, the simultaneous appearance
of SAXS equatorial streaks and WAXD (110) diffraction of monoclinic -form crystal is
observed. The SAXS equatorial streaks are associated to the formation of fibrils oriented
in the flow direction. In addition, the simultaneous appearance of SAXS streaks and
the WAXD (110) diffraction suggests that these fibrils are already (partially) crystalline
and, therefore, we consider these as crystalline shish. For this case the onset time of
shish is during flow, just before the end of the flow time (0.25 s).
To further illustrate the differences between strong flows, the SAXS results for shear
rates below and above 400 s1 are shown in detail in Figure 4.5 (we do not show all
WAXD results since the shish formed at 320 and 5601 are also crystalline. A detailed
crystallization evolution will be presented in chapter 5).
For a wall shear rate = 560 s1 the SAXS streaks appear at 0.17 s, which is still
within the flow period of 0.20 s, while for a wall shear rate = 320 s1 the equatorial
streaks appear at 0.33 s, i.e. after cessation of flow. For = 560 s1 shish start to
develop already during the brief shear pulse while for = 320 s1 shish develop after
cessation of flow.

Chapter 4

53

SAXS

time

WAXD

0.10s

0.20s
0.23s

shish
streak

(110)

0.27s

(110)

0.40s

(110)
(130) (040)

Figure 4.4: SAXS and WAXD patterns during and just after flow at = 400 s1 . Flow time
is 0.25 s. The images acquired after flow are indicated by the time in red. Flow
direction is vertical.

560s-1 for 0.2s

320s-1 for 0.25s

t = 0.13s

t = 0.26s

t = 0.17s

t = 0.33s
streak

t = 0.20s

t = 0.40s

Figure 4.5: SAXS patterns for flow rates of 560 s1 and 320 s1 . The images acquired after
flow are indicated by the time in red. Flow direction is vertical.

The formation of shish may affect the rheological behavior of the material and, thus,
contributes to the deviation of P from a melt-like behavior. The relation between the
onset of shish and P is clarified by examining the experiment with wall shear rate
of = 400 s1 . In this experiment, the P upturn time is 0.1 s, see Figure 4.6a.
As shown in Figure 4.6b, ISAXS initially fluctuates around 0.04 but then at 0.2 s,
with the onset of the streaks, starts a quick rise that continues after the cessation of flow.

Chapter 4

54

0.10

SAXS equatorial intensity (a.u.)

750

600

"upturn"

P (bar)

450

300

150

0.08

0.06

onset time

0.04
flow cessation

0.02

0.0

0.1

0.2
time(s)

(a)

0.3

0.4

0.0

0.1

0.2

0.3

0.4

time(s)

(b)

Figure 4.6: (a) P evolution during flow and (b) SAXS equatorial intensity during and just
after flow. Lines are drawn to guide the eyes. The shear rate is 400 s1 and the
flow duration is 0.25 s.

Clearly, the time required to form an amount of shish sufficient to produce a


noticeable increase of ISAXS is longer than the time characterizing the P upturn.
From this distinct difference in onset times it seems that the apparent viscosity rise
(averaged over the channel part between pressure transducers) does not result directly
from shish formation in the middle of the slit (X-ray observation window). Or precursors
for shish are the initial cause for the pressure upswing, or shish is first created at another
(upstream) location in the slit.
Figure 4.7a shows the evolution of the ISAXS for all shear conditions in the first second
(the beginning of flow is the reference time). For wall shear rates higher than the critical
value for viscosity upturn (between 160 and 240 s1 ), ISAXS increases significantly due
to formation and densification of shish. This is illustrated by the equatorial streaks
in the 2D images (Figure 4.7b). To improve the signal-to-noise ratio, the patterns of
240 and 160 s1 were averaged over ten frames (between 0.67 and 1 s). The slightly
higher intensity left and right of the beam stop for 160 s1 are due to the beam itself;
they do not correspond to an equatorial SAXS streak signal.
It is clear that, just after flow, neither a viscosity upturn nor shish appearance is
found at the low shear rates of 80 and 160 s1 . However, the absence of a viscosity
rise and SAXS streaks cannot exclude the generation of precursor structures during
flow. Therefore, for these two weakest flows (wall shear rates of 80 and 160 s1 ), we
show for prolonged isothermal crystallization the WAXD diffraction patterns at 1000 s,
see Figure 4.8. For both conditions, the (040) diffraction is mainly distributed at the
equator, implying that oriented crystals are formed and this orientation finds its origin
in oriented structures generated by flow. From this we conclude that precursors are also
generated by the weak flows and develop to some oriented nuclei, instead of shish.

Chapter 4

55

0.4

SAXS equatorial intensity (a.u.)

560 s
480 s

0.3

400 s
320 s
240 s
160 s

0.2

-1

-1

-1

-1

-1

-1

80 s

-0.20s
-0.23s
-0.25s
-0.25s
-0.25s
-0.25s

-1

-0.25s

0.1

0.0
0.0

0.2

0.4

0.6

0.8

1.0

time (s)

(a)

(b)

Figure 4.7: (a) SAXS equatorial intensity evolution during and just after flow and (b) 2D
SAXS pattern at 1 s for different shear rates. Flow direction is vertical.

(040)

(040)

(a)

(b)

Figure 4.8: 2D WAXD patterns of isothermal crystallization at 1000 s after flow pulses at
shear rate (a) 160 s1 and (b) 80 s1 . Flow direction is vertical.

All results of the P upturn time and the shish onset time are summarized in the
Figure 4.9. Both decrease with increasing shear rate. The SAXS streaks for a wall shear
rate of 240 s1 do appear after flow but it is hard to determine the precise onset time
for these streaks. Based on Figure 4.9, it is concluded that structure development can
occur during or after flow, depending on the shear strength.
Summarizing above results, for the short flow conditions applied, a critical apparent
wall shear rate is found for the deviation of the rheology from the expected melt-like
behavior. For the iPP used in this work, this critical value lies between 160 and 240
s1 at 145 C. When shear rate is beyond this threshold, the viscosity rises during flow,
indicating that these short-term flows (0.200.25 s) are not short enough from the

Chapter 4

56

0.40
SAXS streak

0.35

viscosity rise
after flow

0.30
flow cessation

time (s)

0.25
0.20
0.15

during flow

0.10
0.05
0.00
0

100

200

300

400

500

600

700

800

shear rate (1/s)

Figure 4.9: Onset times for P rise and occurrence of the SAXS streaks for different wall
shear rates. The gray region indicates the flow period for these shear rates.
Onset time for SAXS streaks at a shear rate of 240 s1 can not be determined
accurately.

rheological point of view. This viscosity rise is followed by a subsequent appearance


of shish in the middle of the slit where the SAXS is measured. This suggests that the
viscosity rise is caused by formation of shish or their precursors before actual scattering
objects are formed in the slit middle of observation window. However, the real delay time
between viscosity rise and shish formation is not exactly known since the location where
the structure formation occurs that corresponds to the viscosity rise is not necessarily
the centre of the slit where SAXS is measured. The delay time should be something
between 0 and 0.1 s. This is discussed further in the next section. Another shear rate
threshold is between 320 and 400 s1 for which shish occur during flow. In that case,
the short durations are even not sufficiently short from a structural point of view.

4.4 Discussion
For short flows of 0.200.25 s, when the wall shear rate is strong enough ( 240
s ) both, a viscosity rise and shish formation are observed. Comparison of rheological
and SAXS data shows that the viscosity rise and the shish formation are not observed
at the same time. Unexpectedly, the former is observed first. This implies that prior
to shish appearance in the middle of the slit, where the SAXS measurements are done,
some structures have already formed in the flow channel and they, behaving as physical
cross-links, increase locally the viscosity. The reason that these structures are not
observed with SAXS can be twofold: a) the contrast (or concentration) is too low, b)
structure formation is not homogeneous over the slit length and precursors may form
first upstream where the pressure is highest. Especially the latter implies that the
1

Chapter 4

57

interpretation of the delay in occurrence of shish should be considered with caution.


No strong statements can be made about case a) unless we can exclude case b). In the
following we will first show that we cannot neglect the possibility of case b). Next we
will discuss the explanation for case a) should b) not the cause of the delay.

4.4.1 SAXS delay time


The choice of a slit geometry has the advantages of allowing for very high shear rates
and pressurizing the sample (preventing shrinkage holes) but, at the same time, brings
some complications. For high shear rates, also high pressure gradients will occur. High
pressures (107 108 Pa) will not only shift the equilibrium melting temperature, but
also will change the viscosity, the (flow induced) crystallization kinetics and, moreover,
compressibility can become important. The coupling between these variables is highly
non-linear. Only with a full numerical model [3638] the interplay between all these
material functions can be investigated quantitatively. Here, we will discuss these issues
in a qualitative way only. For this purpose, the morphology of the sample subjected to
the flow with an apparent shear rate of 240 s1 is examined along the flowing direction.

Figure 4.10: Optical microscopy pictures of cross-sections of samples along the flow direction
for an apparent shear rate of 240 s1 . The sample positions are illustrated
by the top drawing and the corresponding images are viewed between two
polarizers at 45 . The scale bar is 0.2 mm.

An ex-situ micrograph of the structure distribution over the sample length is shown
in Figure 4.10. The variation in the thickness of the shear layer demonstrates that the
structure formation depends on the position in the flow direction. Position 1 is where
the polymer that is kept at a high temperature (220 C) enters the test section of the
slit, i.e. T = 145 C. The high temperature of the melt allows for molecular relaxation
and leads to a thin shear layer. On the other hand, the polymer at the positions 25

Chapter 4

58

is at the set experimental temperature of 145 C. In these positions there is indeed a


slight but clear gradient in the shear layer thickness going from high to low pressure
zone (i.e. going in the flow direction), see Figure 4.11.

thickness of shear layer (mm)

0.5

0.4
flow direction

0.3

0.2

0.1

0.0
-60

-40

-20

20

40

60

distance from X-ray window (mm)

Figure 4.11: The thickness of shear layer at various positions along the flow direction for
an apparent shear rate of 240 s1 . The reference position is the observation
window, before and after which positions are indicated by negative and positive
numbers, respectively.

How large the effect is in terms of the time difference cannot be determined from these
experiments but we can estimate the order of the differences of flow enhanced nucleation
and shish generation at different locations in the slit by using the characterization of
this grade given in van Erps work [30], where the flow enhanced nucleation rate and
the shish growth rate were determined for a range of temperatures, pressures and shear
histories. Only pressure differences are of importance until the pressure upswing occurs,
since temperature and flow history are practically the same for all locations 15. With
a pressure difference of 100 bar (estimated form the results in Figure 4.3a between the
centre and the upstream position 1 (see Figure 4.10) it can be estimated that the flow
induced nucleation rate might be slightly larger (20%), but the shish growth rate is
considerably enhanced (200%). From this we conclude that indeed structure formation
is not homogeneous over the slit length and shish will occur upstream first where the
pressure is highest.
A much stronger non-linear coupling between the viscosity rise and the flow enhanced
crystallization may happen for more severe flow conditions. The resulting morphology
distribution along the flow direction is shown for the extreme case (shear rate 560 s1
and flow time 0.20 s) in Figure 4.12. Even for the three positions close to the middle,
2, 3 and 4, the gradient is clearly visible (average coverage of the shear layers: 92%,
87% and 84%). With a pressure difference of 200 bar for this case (see Figure 4.3),

Chapter 4

59

flow induced nucleation rate and shish growth rate are increased by 40% and 900%,
respectively, i.e. the pressure effects are dramatically increased compared to the case
with a shear rate 240 s1 . Concluding, there definitely is an influence of the location
in the slit and structure formation most probably will occur first upstream. Again, this
coupling effect will not be quantitatively discussed here, since it is too complex without
simulation tools, which is out of the scope of this work.

Figure 4.12: Optical microscopy pictures of cross-sections of samples along the flow direction
for an apparent shear rate of 560 s1 . The sample positions are illustrated
by the top drawing and the corresponding images are viewed between two
polarizers at 45 . The scale bar is 0.5 mm.

In the following we will discuss our results in terms of shish and precursors for
shish assuming a time difference between rheological changes and nucleation of shish
somewhere between 0 and 0.1 s.

4.4.2 Implications of the viscosity rise


Flow gradients can have remarkable effects on generating structures since they
effectively orient and stretch polymer chains.
Due to stretch, iPP molecules
change their conformations to adapt to the increased segmental distance between
entanglements and form ordered units of 3/1 helices along the flow direction. [15, 39]
The molecular orientation and stretch increase the equilibrium melting temperature
providing additional thermodynamic driving force (under-cooling) for crystallization.
These parallel segments are similar to their arrangement state in the nuclei and crystals.
Thus the kinetic barrier of transforming the chain segments from random coils into
ordered structures is lowered. The consequence is the formation of locally ordered
aggregates, or precursors, which can be considered as the cradle for nuclei and/or
shish. It is hypothesized that due to the interactions between helices [39], the molecular

Chapter 4

60

segments are restricted inside precursors and, as a consequence, the involved chains
cannot move freely as they do in entangled melts. Therefore, these precursors can act
as physical cross-links and slow down relaxation dynamics. [36] As precursors are being
generated continuously during flow, the very early stage of shear-induced crystallization
is a cross-linking process. Precursors can crystallize; this is the nucleation step. The
time between precursor formation and nucleation is unknown but should be (much) less
than 0.1 s. Once the density of precursors (and nuclei) is high enough, the resulting
increase in viscosity is sufficient for rheological observation; see Figure 4.3. The viscosity
rise in our results indicates a high concentration of precursors/nuclei generated by strong
flows (shear rates 240 s1 ).
This cross-linking process during flow seems very similar to the physical gelation
process found in quiescent crystallization of iPP by Pogodina et al. [40, 41] However,
they found a crystallinity of around 1% at the gel point [40], whereas we did not observe
any crystallinity when viscosity starts to change. It should be noted that, in the work
of Pogodina et al. [40], low degrees of under-cooling (10-26K) were used for quiescent
crystallization for which the nucleation density is determined by temperature only. A
low under-cooling corresponds to relatively few nuclei and, therefore, the cross-linking
effect is too low to form a sufficiently dense network; gelation is not observed in the very
early stage of crystallization. However, with increasing crystallinity, more chains get
involved in the network until this is sufficient for observing a viscosity rise. So for low
under-cooling, large clusters (size of 1 m [41]) and a crystallinity of 1% [40] are
required for achieving a sufficient network. Our case with strong flows is quite different;
a precursor/nuclei density increase of orders of magnitude is possible, and viscosity can
rise as a consequence of network formation without noticeable crystal formation.
The precursors formed under shear rates 160 s1 (indicated by the appearance
of oriented crystals during isothermal crystallization) are too dilute (or weak) to rise
the viscosity during flow. Assuming that the specific cross-linking capability is the
same for different precursors, it is inferred that the precursor concentration is the major
difference. The case of low shear rates ( 160 s1 ) is then more similar to quiescent
crystallization [40] in terms of absence of significant viscosity change; i.e. a relatively
low nucleation density and further crystallization is required to form a sufficient network
for a detectable change in the viscosity.

4.4.3 Conditions of the viscosity rise


Macroscopic flow is able to generate precursors for row nuclei or for shish. Molecular
stretch of the high molecular weight tail is the prerequisite for the creation of
precursors. [42] To illustrate the flow strength dependence of the structures formed,
two characteristic Weissenberg numbers are defined [43]: W i0 =
rep which is related

Chapter 4

61

to molecular orientation, and W is =


Rouse which is related to molecular stretch. rep
and Rouse are the reptation and Rouse time related to the high molecular weight tail,
respectively. At 145 C, the relaxation times rep and Rouse for the iPP used in this
work are 48 and 0.23 s, respectively. [18] The values of Weissenberg numbers for the
different flows are summarized in Table 4.1. W i0 103 while W is > 10 for all shear
rates.
Table 4.1: Weissenberg numbers W i0 and W is for HMW tail at 145 C

Apparent shear
rate (s1 )
W i0
(orientation)
W is
(stretch)

80

160

240

320

400

480

560

3.8103

2.7103

11.5103

15.4103

19.2103

23103

26.9103

18.4

36.8

55.2

73.6

92

110

129

This means that all shear rates applied can effectively orient the contour path
along the flow direction and stretch molecular segments to deviate from the rotational
isomerization corresponding to the equilibrium Gaussian configuration. [43] However,
rheological results in Figure 4.3a show that viscosity rise only occurs for 240 s1 .
This indicates that stretch of the HMW tail is the necessary but not a sufficient factor
for viscosity rise. [16,44] The HMW tail stretch should be above a critical value to start
the growth of fibrillar nuclei. [37] Since the Deborah number > 1 (De = relaxation time
/ shear time) for ts < 0.23 s, which is nearly the same as the max flow time, the long
chain deformations can be considered as affine during flow and the molecular stretch
scales with the macroscopic strain (= shear time shear rate). The stretch values
obtained in this way are 20 and 40 for the shear rates of 80 and 160 s1 , respectively.
This would imply that the critical stretch at the wall is 40 or more. For shear rates
240 s1 all stretch values are 60 but the stretch values obtained from the onset
time of the viscosity rise for these shear rates are 36, 38, 39, 35, 35. This matches quite
well with the critical value of 40 estimated from the shear rate of 160 s1 . Fibrillar
precursors are created for shear rates > 160 while for shear rates 160 s1 this is not
the case due to the too short shear duration. Finally, the critical stretch value compares
surprisingly well with the critical value used by Custodio et al. [37] (critical stretch =
40). Although this does not mean that we should attribute too much meaning to the
exact value of the critical stretch, the results are quite consistent. If a critical stretch
is the criterion for the start of forming oriented shish precursors, it is questionable if a
work criterion as defined with Eq. 4.1 is applicable since most of the period for reaching
this value is transient.

Chapter 4

62

4.5 Conclusions
The rheological and structural evolution during and after short-term flow of
0.200.25 s at 145 C were studied for iPP. The rheological results show that viscosity
rises beyond the normal pressure overshoot for apparent shear rates 240 s1 , and
the onset time for this rise decreases with increasing shear rate. This viscosity rise
implies that these flows do not satisfy the basic requirement for a short-term flow,
i.e. that the polymer viscosity does change during flow. Therefore, for these cases, a
more advanced analysis is required. For example, these results can be combined with a
detailed model for flow induced crystallization including a fully characterized non-linear
viscoelastic model [36,37] from which the relaxation times are coupled to the structural
development.
X-ray measurements do not show simultaneous structure development with the
viscosity rise; the observation of shish is delayed, typically 0.1 s. The viscosity
rise may partially be explained by the creation of shish or precursors for shish that
act as physical cross-links. However, we can not be sure about the exact value of the
delay time; the influence of the pressure on the local values of rheological and kinetic
parameters will cause nucleation events to occur first upstream. Only a numerical model
can help to reveal this complex interaction. The development from precursors to shish
can occur during ( 400 s1 ) or after flow (400 > 240 s1 ).
Shear rates (160 and 80 s1 ) below the threshold also generate precursors which, due
to their low concentration, do not change the melt viscosity and develop into row nuclei.
These flows can be considered as short-term shear flows.

References
[1] H. E. H. Meijer and L. E. Govaert. Progress in Polymer Science 30(8-9):915938, 2005.
[2] 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.
[3] T. W. Haas and B. Maxwell. Polymer Engineering and Science 9(4):225241, 1969.
[4] A. Wereta and C. G. Gogos. Polymer Engineering and Science 11(1):1927, 1971.
[5] R. R. Lagasse and B. Maxwell. Polymer Engineering and Science 16(3):189199, 1976.
[6] C. H. Sherwood, F. P. Price, and R. S. Stein. Journal of Polymer Science: Polymer Symposia
63(1):7794, 1978.
[7] M. D. Wolkowicz. Journal of Polymer Science: Polymer Symposia 63(1):365382, 1978.
[8] F. Jay, J. M. Haudin, and B. Monasse. Journal of Materials Science 34(9):20892102, 1999.
[9] R. Pantani, I. Coccorullo, V. Volpe, and G. Titomanlio. Macromolecules 43(21):90309038, 2010.
[10] S Liedauer, G Eder, H Janeschitz-Kriegl, P Jerschow, W Geymayer, and E. Ingolic. International
Polymer Processing 8:236244, 1993.

Chapter 4

[11] S Liedauer, G Eder, and H Janeschitz-Kriegl. International Polymer Processing 10:243250,


1995.
[12] G. Kumaraswamy, R. K. Verma, J. A. Kornfield, F. Yeh, and B. S. Hsiao. Macromolecules
37(24):90059017, 2004.
[13] R. H. Somani, L. Yang, L. Zhu, and B. S. Hsiao. Polymer 46(20):85878623, 2005.
[14] E. L. Heeley, C. M. Fernyhough, R. S. Graham, P. D. Olmsted, N. J. Inkson, J. Embery, D. J.
Groves, T. C. B. McLeish, A. C. Morgovan, F. Meneau, W. Bras, and A. J. Ryan. Macromolecules
39(15):50585071, 2006.
[15] H. An, B. Zhao, Z. Ma, C. Shao, X. Wang, Y. Fang, L. Li, and Z. Li. Macromolecules 40(14):4740
4743, 2007.
[16] O. O. Mykhaylyk, P. Chambon, R. S. Graham, J. P. A. Fairclough, P. D. Olmsted, and A. J.
Ryan. Macromolecules 41(6):19011904, 2008.
[17] L. Balzano, N. Kukalyekar, S. Rastogi, G. W. M. Peters, and J. C. Chadwick. Physical Review
Letters 100(4):048302, 2008.
[18] J. W. Housmans, R. J. A. Steenbakkers, P. C. Roozemond, G. W. M. Peters, and H. E. H. Meijer.
Macromolecules 42(15):57285740, 2009.
[19] T. Yan, B. Zhao, Y. Cong, Y Fang, S. Cheng, L. Li, G. Pan, Z. Wang, X. Li, and F. Bian.
Macromolecules 43(2):602605, 2009.
[20] H. Janeschitz-Kriegl, E. Ratajski, and M. Stadlbauer. Rheologica Acta 42(4):355364, 2003.
[21] G. Kumaraswamy, A. M. Issaian, and J. A. Kornfield. Macromolecules 32(22):75377547, 1999.
[22] J. Baert, P. Van Puyvelde, and F. Langouche. Macromolecules 39(26):92159222, 2006.
[23] L. G. Balzano, S. Rastogi, and G. W. M. Peters. Macromolecules 42(6):20882092, 2009.
[24] H. Janeschitz-Kriegel and E. Ratajski. Polymer 46(11):38563870, 2005.
[25] P. C. Roozemond, R. J. A. Steenbakkers, and G. W. M. Peters. Macromolecular Theory and
Simulations 20(2):93109, 2011.
[26] G. Kumaraswamy, J. A. Kornfield, F. J. Yeh, and B. S. Hsiao. Macromolecules 35(5):17621769,
2002.
[27] N. J. Terrill, P. A. Fairclough, E. Towns-Andrews, B. U. Komanschek, R. J. Young, and A. J.
Ryan. Polymer 39(11):23812385, 1998.
[28] P. D. Olmsted, W. C. K. Poon, T. C. B. McLeish, N. J. Terrill, and A. J. Ryan. Physical Review
Letters 81(2):373376, 1998. PRL.
[29] M. van Drongelen, T. B. van Erp, and G. W. M. Peters. Polymer 53:47584769, 2012.
[30] T. B. van Erp. Structure development and mechanical performance of polypropylene. Ph.D. thesis,
Eindhoven University of Technology, Eindhoven, The Netherlands, 2012.
[31] Z. Ma, L. Balzano, and G. W. M. Peters. Macromolecules 45(10):42164224, 2012.
[32] C. W Macosko. Rheology, principles, measurements and application. Wiley-VCH, New York,
1994.
[33] 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.
[34] P. C. Roozemond, V. Janssens, P. Van Puyvelde, and G. W. M. Peters. Rheologica Acta 51(2):97
109, 2012.
[35] L. Scelsi and M. Mackley. Rheologica Acta 47(8):895908, 2008.
[36] H. Zuidema, G. W. M. Peters, and H. E. H. Meijer. Macromolecular Theory and Simulations
10(5):447460, 2001.
[37] F. J. M. F. Custodio, R. J. A. Steenbakkers, P. D. Anderson, G. W. M. Peters, and H. E. H.
Meijer. Macromolecular Theory and Simulations 18(9):469494, 2009.

63

64

Chapter 4

[38] G. W. M. Peters, F. H. M. Swartjes, and H. E. H. Meijer. Macromolecular Symposia 185(1):277


292, 2002.
[39] H. An, X. Li, Y. Geng, Y. Wang, X. Wang, L. Li, Z Li, and C. Yang. The Journal of Physical
Chemistry B 112(39):1225612262, 2008.
[40] N. V. Pogodina and H. H. Winter. Macromolecules 31(23):81648172, 1998.
[41] N. V. Pogodina, S. K. Siddiquee, J. W. van Egmond, and H. H. Winter. Macromolecules
32(4):11671174, 1999.
[42] R. J. A. Steenbakkers and G. W. M. Peters. Journal of Rheology 55(2):401433, 2011.
[43] J. van Meerveld, G. W. M. Peters, and M. Hutter. Rheologica Acta 44(2):119134, 2004.
[44] O. O. Mykhaylyk, P. Chambon, C. Impradice, J. P. A. Fairclough, N. J. Terrill, and J. Ryan, A.
Macromolecules 43(5):23892405, 2010.

Chapter five

The influence of flow induced


precursors and nuclei on
crystallization of isotactic
polypropylene

Abstract
By using ultra-fast WAXD and SAXS measurements we were able to explore shish
formation during and just after flow in a slit, within a short period of 1 s. Moreover, by
using two WAXD detectors with different acquisition rates we achieved a full track
of the entire crystallization process, from the start-up of flow up to completion of
crystallization. The flow duration is very short, max 0.25 s. The X-ray results show that
for apparent wall shear rates 400 s1 , shish appear during flow and that for apparent
wall shear rates at 320 and 240 s1 , shish precursors are generated during flow and
develop to shish after flow cessation. In contrast, no shish or shish precursors are found
for 160 and 80 s1 . Therefore, birefringence is applied for these two relatively weak
flows to probe potential shear-induced structures, where SAXS fails. The birefringence
measurements of 160 s1 show an unusual rise during flow and a partial relaxation
after flow, implying a sort of long-lived oriented structures generated by flow. Neither
SAXS nor birefringence signals of precursors are observed for 80 s1 , but the subsequent
crystallization, observed with WAXD, does show accelerated kinetics and an oriented
morphology. These WAXD results demonstrate existence of oriented precursors which
This Chapter is based on: Zhe Ma, Luigi Balzano, Giuseppe Portale, Gerrit W. M. Peters. to be
submitted, 2012

65

Chapter 5

66

are below the detection limitations of SAXS and birefringence. By combing results from
different characterization methods (SAXS, WAXD and birefringence) we can distinguish
between different shear-induced structures. They are classified into various types: shish,
shish precursors, threadlike precursors and needlelike precursors, according to whether
they can be detected by SAXS and/or birefringence.
Besides this detailed picture of the initial stages of flow induced nucleation, tracking
the entire structure evolution reveals further interesting issues on crystallization. Firstly,
daughter lamellae appear later than parent lamellae and the ratio between parent and
daughter lamellae can be varied by flow. Secondly, orientation of the initial shish is high
and lamellar orientation decreases with lateral growth. Finally, depending on the shear
strength, -phase of iPP can be induced at the experimental temperature of 145 C.

5.1 Introduction
For semi-crystalline polymers, the ultimate mechanical properties of end products
strongly depend on the crystalline structures formed, i.e. crystal modification,
degree of crystallinity, lamellar thickness, orientation, etc. According to the classical
crystallization theory, polymer crystallization is a two-step process: nucleation and
crystal growth. [1] To start polymer crystallization, stable nuclei are required.
Formation of stable nuclei follows the thermodynamic rule that the total Gibbs free
energy change from melt to nuclei, G=Gnuclei Gmelt , has to be negative. This free
energy change can be expressed as: G=Gb +Gs , where Gb is the bulk free energy
change (a negative value) and Gs surface free energy change (a positive value) [1], i.e.
the nucleation process is determined by the competition between bulk and surface free
energy changes, of which the former should overcome the latter to generate stable nuclei.
In a pure polymer melt without foreign objects, the surface free energy is associated to
the sum of contribution of all surfaces, i.e. Gs =A (with the specific surface energy
and A the corresponding surface area). In general, homogeneous nucleation requires a
large driving force, i.e. a sufficient under-cooling. [2, 3] However, in practice polymers
contain contaminates like catalyst residues. These foreign objects provide extra surface
and consequently lower the nucleation barrier (the change of free surface energy). This
heterogeneous nucleation occurs much easier (in terms of nucleation temperature and
nucleation rate) than the homogeneous case. Knowing the effective approach of utilizing
the foreign surface to lower the surface free energy change, addition of nucleating agents
is, therefore, frequently used to enhance nucleation and control optical properties (e.g.
transparency and color). Flow, on the other hand, affects crystallization in a different
way. From a thermodynamic point of view, the application of flow raises the free
energy, decreasing Gb further and, therefore, promotes nucleation. Flow also induces
anisotropy to the final morphology, for example, the typical shish-kebab structure [4]
which consists of fibrillar core oriented along the flow direction (shish) with transverse,

Chapter 5

periodically stacked lamellae (kebab). Shear-induced structures can be stable nuclei,


capable of initiating crystal growth, or precursors with a degree of ordering (positional
or/and orientational; intra- or/and inter- molecule) which can nucleate under sufficient
under-cooling or may relax in case of a relatively high temperature.
Considerable work has been devoted to the effects of various precursors/nuclei on
crystallization. Cavallo et al. [5] found that unstable precursors for nucleation
of isotactic polypropylene (iPP) disappear with time above the nominal melting
temperature. As a consequence, this relaxation of precursors for nucleation
results in reduced final orientation and slower crystallization kinetics. Similar, for
polyethylene (PE), relaxation of shear-induced precursors leads to a larger fraction of
twisted lamellae. [6] The threadlike precursors of a shear-induced structure found
by Kumaraswamy et al. [7], possess oriented molecular segments, detectable with
birefringence, which causes the appearance of a skin layer. Shear-induced bundles
that are densely packed and observable by SAXS can start crystallization in the vicinity
of the melting temperature [8, 9]. Moreover, shear is even able to determine crystal
modification, for example, shear-induced iPP -row nuclei (by means of fiber-pulling)
can initiate the growth of the -phase as well as the monoclinic -phase. [10]
To explore flow-induced precursors/nuclei and their influences on crystallization,
the ideal experimental strategy should cover both the initial generation process
of precursors/nuclei (from start of flow) and the subsequent growth process (until
completion of crystallization). However, precursors/nuclei can form quite rapidly, e.g.
shorter than 1 second [11], which puts severe demands on the experimental methods
used. The first step is really a challenge, i.e. to achieve the in-situ tracking of
nuclei/precursors formation. Most of the previous studies focus on the crystallization
after cessation of flow and infer the formation and property of nuclei through the kinetics
and morphology of the resulting crystallization. In this chapter, we investigated the
rapid formation of structure in detail using synchrotron radiation X-ray with a high
time resolution. For sufficient high shear rates, the combined results of small angle
X-ray scattering (SAXS) and wide angle X-ray diffraction (WAXD) demonstrate that
crystalline shish form during a short flow time (0.20-0.25 s). Moreover, shish precursors
can be distinguished from shish; the former develop to shish after flow. However, no
structures are detected during and shortly after flow by X-ray for some relatively weak
flows, probably because the density contrast and crystalline order are not sufficient
or concentration is too low. Therefore, the more sensitive experimental method of
birefringence is utilized to probe structure development which is below the detection
limit of X-ray. Once precursors/nuclei are characterized, the next step is to explore how
the subsequent isothermal crystallization is affected by them.
The present work aims to clarify what kind of precursors/nuclei are generated for
various flow conditions and to bridge the gap between various nuclei and the resulting
crystallization process.

67

Chapter 5

68

5.2 Experimental
5.2.1 Material
A commercial isotactic polypropylene (iPP, Borealis HD601CF) is used in this work.
Material properties are summarized in Table 5.1. The melting (Tm ) and crystallization
(Tc ) temperatures are measured using DSC at a heating/cooling rate of 10 C/min,
sample weight: 30.5 mg.
Table 5.1: Properties of the isotactic polypropylene used.

Material
iPP

Mw (kg/mol)
365

Mw /Mn
5.4

Tm ( C)
163

Tc ( C)
113

5.2.2 Flow device


The flow device used in this work is a slit flow based on a multipass rheometer which
has been described elsewhere [6]. The sample is confined between two servo hydraulically
driven pistons in a slit with a cross-section, A, of 6mm1.5mm (width, W height,
H). To impose a shear field to the molten polymer the two pistons move simultaneously
in one direction. Two diamond windows are mounted in the middle section of the flow
cell allowing the passage of X-ray and laser light for in-situ characterization of structure
evolution during and after flow.
The sample was first heated and kept at 220 C for 10 min in order to erase previous
thermo- and mechanical histories from sample preparation. Next, the sample was cooled
to 145 C. All experiments were performed at this temperature. The applied piston
speeds, Vpiston , were 20, 40, 60, 80, 100, 120 and 140 mm/s. The apparent wall shear
rates are 80, 160, 240, 320, 400, 480 and 560 s1 , determined by = W6Q
[12] with the
H2
volumetric flow rate Q = Vpiston A. The flow time is fixed at 0.25 s for piston speeds
ranging from 20 to 100 mm/s, and shortened to 0.23 and 0.20 s for 120 and 140 mm/s,
respectively, due to limitation in the maximum piston displacement.

5.2.3 X-ray characterization


Wide angle X-ray diffraction (WAXD) and small angle X-ray scattering (SAXS)
were used to characterize the entire isothermal crystallization process, including the
period of flow. All measurements were performed at the Dutch-Belgian (DUBBLE)
beamline BM26B of the European Synchrotron Radiation Facility (ESRF) in Grenoble,
France. [13] The wavelength used was 1.033
A.
Two WAXD detectors (Pilatus and Frelon, see Figure 5.1) with different acquisition

Chapter 5

69

rates and different pixel resolutions were used. The Pilatus (300K) detector has a high
acquisition rate mode of 30 frame/s, enabling the tracking of the structural evolution
during (0.20-0.25 s) and just after flow for a period of 2 s. After this period the Pilatus
detector was switched to the regular slow acquisition mode (exposure time of 4.9 s,
total acquisition time 5.7 s per frame) to follow the resulting isothermal crystallization.
This Pilatus (300K) detector has a resolution of 1472195 pixels of 172m172m.
The sample-to-detector distance was fixed at 0.240 m.
time
0.2- 0.25s

flow

SAXS

fast

slow
1s

Pilatus

fast

slow
2s

WAXD

1000s
Frelon

fast

slow
10s

Figure 5.1: Structural evolution is measured by the combination of the Pilatus and Frelon
X-ray detectors. The scheme shows how the detetors are used during different
periods.

SAXS measurements were performed with a Pilatus (1M) detector (9811043 pixels
of 172m172m) placed at a distance of 7.117 m. For SAXS, the fast step was carried
out at an acquisition rate of 30 frame/s for 1 s and the slow step had an exposure period
of 5.7 s for the following isothermal process.
Figure 5.2a shows a typical WAXD pattern obtained with the Pilatus (300K)
detector. For the chosen distance, the azimuthal range of this detector is less than
90 for iPP -phase. The diffractions in the equatorial region are related to the highly
oriented crystals including shish and parent lamellae. In order to explore other crystals
oriented along different directions, such as daughter lamellae, a WAXD azimuthal region
beyond a quarter is required. For this purpose, the Frelon detector was employed, which
has a resolution of 20482048 pixels of 48.8m48.8m. The corresponding 2D Frelon
pattern is shown in Figure 5.2b. A full diffraction pattern can be obtained by using
symmetry considerations. The measurements with the Frelon detector also include a
fast and a slow period, see Figure 5.1. Since the maximum acquisition rate of Frelon
detector is 3 frame/s, one order slower than that of Pilatus, the Frelon fast step covers

Chapter 5

70

(040)(110)

(a)
(b)

isotropic
part

155o

(c)

azimuthal
angle
192o

isotropic part
daughter
parent

(040)(110)

Figure 5.2: WAXD patterns obtained with (a) Pilatus detector and (b) Frelon detector.
(c) Azimuthal scan of (110) diffraction of the Pilatus WAXD pattern at tc = 25
s. Flow direction is vertical.

(040)(110)

(a)

(c)

(b)

overlapped

(040)(110)

Figure 5.3: WAXD patterns obtained with (a) Pilatus detector and (b) Frelon detector.
(c) Integrated intensity versus scattering angle (2) of the Frelon WAXD pattern
at tc = 900 s. Flow direction is vertical.

the first 10 s (from start of flow) with an acquisition period of 0.33 s, including an
exposure time of 0.1 s, per frame. The slow step has an exposure time of 2 s and an
acquisition period of 4.5 s. The distance from sample to the Frelon detector is 0.228 m.
The combination of the ultra-fast Pilatus detector and the Frelon detector provides
the possibility to record crystallization from start-up of flow until completion of growth
for both parent and daughter lamellae.

Chapter 5

In our results of shear-induced crystallization of iPP the monoclinic -phase is the


dominant crystal modification (see Figure 5.2a and 5.2b). It is clear that the azimuthal
locations of the (110) plane diffractions can be in either the equatorial or the meridional
region. This is due to the different c-axes orientation in parent and daughter lamellae
with respect to the flow direction. Therefore, we use the azimuthal scan of the (110)
reflection to separate parent and daughter lamellae. On the other hand, parent and
daughter lamellae possess the same b-axis, so the equatorial (040) reflection without
the isotropic contribution, as shown in Figure 5.3b, demonstrates that only oriented
crystals are grown. Considering this, when we see the non-perfectly separated (110)
diffraction between parent and daughter lamellae, as shown in Figure 5.3b, we know that
overlapping intensities come from the broad azimuthal distribution of (110) parent and
daughter diffractions, instead of from isotropic crystals. Therefore, the (110) azimuthal
scan indeed can be used to quantify oriented parent and daughter lamellae.
To extract the signal of crystalline fraction from the total intensity which includes
contribution of amorphous phases as well, two methods are introduced in the following.
In the 2D WAXD pattern, the scattered intensity I(2, az) of a specific point is
associated to its two coordinate variables, scattering angle (2), and azimuthal angle
(az). This means that two types of one-dimensional (1D) integration curve can be
obtained. In our data analysis, the 1D intensity vs. azimuthal angle curve, see Figure
5.2c, is averaged over the scattering angle range between 9.29 and 9.51 , and 1D
intensity vs. scattering angle curve, see Figure 5.3c, is averaged over a 90 azimuthal
range.
Method I: When the degree of crystal orientation is very high, usually due to a strong
flow or at the very beginning of crystallization, parent and daughter lamellae can be
easily distinguished by their distinct azimuthal (110) distributions, as shown by Figure
5.2b. In this case, the minimum intensity of (110) azimuthal scan (i.e. the intensity vs.
azimuthal angle curve), directly presents the amorphous contribution, see Figure 5.2c.
Method II: When the orientation is not high enough, the (110) diffractions might
be azimuthally so broad that 1) Pilatus detector is not able to acquire all equatorial
diffraction, see Figure 5.3a, or/and that 2) parent and daughter diffractions may overlap,
see Figure 5.3b. For either case, amorphous diffraction can not be determined from the
minimum value of (110) azimuthal scan, as done in Method I. Then, we use the intensity
vs. the scattering angle (I vs. 2) curve to estimate the amorphous contribution, since
from the only oriented (040) we know that all crystals are oriented.
Figure 5.3c shows the 1D intensity vs. scattering angle data including the crystalline
and amorphous contributions. After substracting air and detector backgrounds, the
diffraction of the amorphous fraction is scaled to fit the amorphous halo region, see
Figure 5.3c, and subtracted from the total scattering intensity to obtain the diffraction
signal of the (110) plane only. The (110) peak is fitted with a Lorentzian function to

71

Chapter 5

72

verify the construction of the baseline of the amorphous phase. As stated before, the
azimuthal (110) scan is averaged over the 2 range from 9.29 to 9.51 . Then, for this
2 region, the contribution of the amorphous fraction is equal to:
R 9.51
Iamorp
Iamorp (2) d2
= 9.29
R 9.51
Itot
Itot (2) d2

(5.1)

9.29

Subtracting the amorphous baseline from the total intensity gives the distribution of
diffraction intensity with azimuth of (110) planes.
When the curves for the intensity of the oriented crystalline fraction vs. azimuth are
obtained, Lorentzian functions are used to fit parent and daughter lamellae together.
The fitting area, Area, and the full width at half maximum, FWHM, are measures
for the amount of crystals and the degree of orientation, respectively. Note that the
equatorial (110) diffraction is associated to highly oriented crystals that include both
crystalline shish and parent lamellae.
Since the two detectors have different resolutions, fitting area values, Areaequatorial110 ,
of frames at 25 s from these two detectors are used to determine the ratio between the
absolute intensity values.

5.2.4 Birefringence
1

linear polarizer 2
aligned at 45o

polarizing beam splitter


aligned at -45o

laser

shear cell

D
Figure 5.4: Optical train used for birefringence characterization.

Figure 5.4 shows a schematic picture of the optical train used for the birefringence
characterization. We used a 2mW HeNe laser (wavelength = 633 nm) and two
photodiode detectors (Thor Labs Inc.) Dk and D that read the intensity of parallel
(Iparallel ) and crossed (Icrossed ) light intensity, respectively. The sum of parallel and
crossed light intensity is the total intensity transmitted, Itot = Iparallel + Icrossed .

Chapter 5

73

5.3 Results and Discussion


5.3.1 Flow-induced distinguishable nuclei/precursors
Table 5.2 summarizes the SAXS results for various flow conditions (scattering
patterns discussed in Chapter 4). The flow times are quite short, 0.20-0.25 s, and the
ultra-fast SAXS characterization covers 1 s in total, including the periods during and
just after flow. Clearly, structure formation depends on flow strength. Only when the
shear rate is beyond a critical shear rate (between 160 and 240 s1 ) fibrillar structures
with sufficient density contrast (shish) are observed.
Furthermore, the time at which structures are observed first (by SAXS streaks)
depends on shear rate as well. Shish appear during flow for a shear rate 400 s1 . In
the case of moderate shear rates, i.e. for 400 s1 > shear rate 240 s1 , shish can be
observed only after cessation of flow. This post-flow observation of shish suggests that
the structures generated during flow are precursors for shish which develop into shish
later on.
Table 5.2: Flow conditions applied and the corresponding shish formation characterized
using SAXS.

piston speed (mm/s)


apparent wall shear rate (s1 )
flow time (s)
appearance of SAXS streaks within 1 s
time of observing the first SAXS streaks

140
560
0.20

120
480
0.23

100
400
0.25

Yes
during shear
0.17 0.20 0.23

80
320
0.25

60
240
0.25

40
160
0.25

20
80
0.25

No
after shear
0.33 < 1

SAXS equatorial streaks cannot be detected (within 1 s) for the relatively weak
flows of shear rate 160 s1 . The absence of SAXS equatorial streaks could be
due to the fact that the shear-induced structures, if any, have no sufficient density
contrast with their surrounding melt or that the concentration is too low to be detected.
The isothermal crystallization after flow is examined in order to indirectly explore
potential shear-induced precursors. Since quiescent crystallization needs over 1000 s to
become noticeable (data not shown), substantial appearance of an oriented crystalline
morphology during the first 1000 s has a two-folded meaning for these conditions (see
also Figure 4.8 in Chapter 4 and the next section). Such accelerated kinetics shows that
indeed precursors are generated by flow, although they are below the detection limit of
SAXS. The oriented feature of precursors induced by these conditions is demonstrated
by the resulting oriented morphology.

Chapter 5

74

Therefore, birefringence was applied to characterize the orientation of shear-induced


precursors generated by the relatively weak flows (shear rate 160 s1 ). In
birefringence, molecular orientation becomes manifested in Icrossed /Itot . Figure 5.5 shows
the birefringence evolution during flow (0.25 s) and after flow. By comparing the results
for shear rates of 80 and 160 s1 it is observed that the melt behaves differently in both
periods, i.e. during and after flow. Figure 5.5a shows that Icrossed /Itot of shear at 80 s1
first increases quickly and subsequently reaches a plateau. Immediately after cessation of
this flow (lasting 0.25 s), the birefringence signal Icrossed /Itot completely relaxes to zero.
In contrast, the birefringence signal Icrossed /Itot shows an additional increase instead of
reaching plateau during shear at 160 s1 , see inset Figure 5.5b. After flow cessation,
Icrossed /Itot decays to a nonzero value and rises again.
1.2

1.2

1.0

0.9

flow time 0.25s

flow time 0.25s


0.8

1.0

0.3

0.1

0.2

0.3

0.4

/I

0.0

crossed

time(s)

0.6

0.5

0.6

0.0

0.1

0.2

0.3

0.4

time(s)

0.4

0.4

0.7

0.8

0.2

0.6

crossed

0.4

crossed

/I

tot

0.8

tot

crossed

/I

/I

tot

0.5

tot

0.6

0.2

0.2

0.0

0.0

time(s)

(a)

10

10

time(s)

(b)

Figure 5.5: Birefringence during and after short-term (0.25 s) flows. Shear rate: (a) 80 s1 ,
(b) 160 s1 at 145 C.

In the case of flow at a shear rate of 80 s1 , the transient response on shearing and
cessation of flow (increase, steady-state plateau and complete relaxation) is the typical
nonlinear rheological behavior for polymer melts. [12] So far, birefringence and SAXS
results suggest that precursors generated by the flow at shear rate of 80 s1 for 0.25 s,
possess a poor packing and orientation order. These precursors will be called needlelike
precursors in this work.
In contrast, for a shear rate of 160 s1 the birefringence shows an unusual raise
during flow and only partial relaxation after cessation of flow. This demonstrates
the generation of some long-lived structure, invisible for SAXS but clearly reflected
in birefringence. Such oriented structures are denoted as threadlike precursors [7, 14]
in this work. The combination of birefringence and SAXS distinguishes this threadlike
precursors from shish that has a density contrast with the surrounding melt and a
sufficient concentration for SAXS detection. Similar results for precursors were found

Chapter 5

by Fernandez-Ballester et al. [15] They employed simultaneous birefringence, SAXS


and WAXD, and found a sort of shear-induced precursor in iPP melt at 160 C with
a birefringence upturn signal during flow and a nonzero low birefringence after flow,
without detectable SAXS or WAXD signal. In fact, the shear-induced precursor they
found is consistent with the threadlike precursors in our results.
Therefore, different flow-induced structures can be formed: shish, precursors for
shish, threadlike precursors and needlelike precursors, depending on the flow conditions.
By using a combination of SAXS, WAXD and birefringence we can distinguish between
these structures. Next, we explore how these precursors/nuclei affect the successive
crystallization. Moreover, by studying the resulting crystallization features (kinetics and
orientation), a better understanding on the needlelike precursors should be gained,
since direct characterizations by SAXS and birefringence fail.

5.3.2 Isothermal crystallization


The detailed study of the whole crystallization evolution is realized by combining
the Pilatus detector, aiming for a high time-resolved characterization during and
just after flow, and the Frelon detector to acquire 2D WAXD patterns covering the
azimuthal range beyond 90 during further isothermal crystallization. In this way, the
crystallization process can be fully tracked, from start-up of short-duration flow until the
end of the crystallization process. Specific crystallization aspects, including growth of
parent and daughter lamellae, orientation, and polymorphism are revealed and related
to the various precursors/nuclei generated by flow.
Kinetics affected by various precursors/nuclei
Shish (shear rate 400 s1 )
Figure 5.6 shows Pilatus WAXD patterns of the early stage, i.e. during and just after
the strongest flow applied (560 s1 for 0.20 s). The first diffraction appears at 0.17 s, i.e
during flow, and belongs to (110) planes of monoclinic -phase of iPP. Other diffraction
patterns, (040) and (130) of -phase, are observed as crystallization further develops,
see Figure 5.6. When -crystals are highly oriented along flow direction, diffractions of
the (111) and (-131) planes are in the off-axis direction rather than in the equatorial
region, so these diffractions were not observed by the Pilatus detector.
Remarkably, the time of the first WAXD appearance (0.17 s) is exactly the same as
for observing the first SAXS equatorial streaks, see Table 5.2 (and Figure 4.5 and 4.9 in
Chapter 4). This excellent agreement between SAXS and WAXD in terms of formation
time demonstrates that -crystal shish are formed during flow of 560 s1 . The SAXS
results show that for shear rates 400 s1 densely packed objects appear during flow

75

Chapter 5

76

time

WAXD

0.13s

0.17s

(110)

(110)

flow 0.20s
cessation

(110)

0.27s
(130) (040)

Figure 5.6: Pilatus WAXD patterns during and after shear of 560 s1 (shear time = 0.20 s).
Flow direction is vertical.

and the present WAXD results reveal that these fibrillar structures correspond with
crystalline shish (data of 480 s1 is not shown and that of 400 s1 is shown in Figure 4.4
of Chapter 4). Shish crystals provide the ideal lattice-matching template to effectively
and efficiently initiate further crystallization (lamellar growth).
Figure 5.7 represents in a quantitative way the entire evolution of -crystals of
both parent and daughter lamellae during and after the strongest flow. The equatorial
(110) intensity quickly starts rising during flow. After cessation of flow, the crystalline
structures continue to develop. The evolutions of the equatorial (110) diffractions of the
Pilatus and Frelon detectors overlap with each other. The agreement demonstrates the
success of this approach of combining two detectors to fully characterize the whole
crystallization process. The crystallization rate is so fast that nearly half of the
intensity level reached at a long crystallization time (around 800 s) is achieved within
10 s. The equatorial diffraction may have contributions from both shish crystals and
parent lamellae, and their contributions cannot be separated. Considering the fact that
the shear-induced shish crystals directly provide a perfect crystal lattice to nucleate
subsequent crystal growth and considering the fast evolution rate of equatorial (110)
diffraction, we qualitatively associate the first appearance of WAXD diffraction with
crystalline shish and relate the main of the following significant increase to the growth
of parent lamellae.

Chapter 5

77

35000
30000

Area 110 (a.u.)

25000

10s

Equatorial (110)-Pilatus-fast
Equatorial (110)-Pilatus-slow
Equatorial (110)-Frelon-fast
Equatorial (110)-Frelon-slow
Daughter (110)-Frelon-fast
Daughter (110)-Frelon-slow

6.7s

20000

daughter
lamellae

15000
10000

1.7s

flow

5000
0
0.01

0.1

10

100

1000

time(s)
Figure 5.7: Time evolution of the area of equatorial and daughter (110) diffractions during
and after flow (560 s1 for 0.20 s). Selected Frelon patterns are shown in the
right column. Flow direction is vertical.

Daughter lamellae evolve differently from the parent lamellae; they appear later.
Figure 5.7 shows that parent lamellae have already grown significantly within the first
1.7 s, whereas daughter lamellae can not be observed yet. The first noticeable daughter
diffraction is detected at 3.3 s. Daughter lamellae are nucleated on the (010) lateral
surface of existing -crystals due to homoepitaxy. [16] This means that the number of
nucleation sites for daughter lamellae depends on the amount of (010) surface from the
parent lamella. Although the parent lamellae develop significantly within the first 1.7 s,
the time is too short for generating sufficient nucleation sites and the daughter lamellae
nucleated, if any, have limited time to grow enough for detection. As crystallization
proceeds, the growing parent lamellae provide more lateral surface for nucleation and
the daughter lamellae initially nucleated grow further. At a later stage of crystallization,
daughter lamellae diffraction becomes comparable to parent diffraction.
Two other, relatively intensive flow conditions in which crystalline shish are generated
during flow are shear rates of 480 and 400 s1 for 0.23 and 0.25 s, respectively. Figure
5.8 shows the time evolution of parent and daughter lamellae for these two flows. In
both cases the growth of parent lamellae is initiated by crystalline shish and this growth
continues after cessation of flow. Similarly, daughter lamellae appear later than parent
lamellae. Since daughter diffraction is initially very weak, each 5 frames of the Frelon
WAXD results in the fast acquisition period are added to get a better signal-noise ratio.
Considering the time-resolution of 1.7 s/frame, appearance of daughter lamellae is 3.3
s for shear rate of 480 s1 and 6.7 s for shear rate of 400 s1 .

Chapter 5

78

35000

35000
Equatorial (110)-Pilatus-fast

Equatorial (110)-Pilatus-fast

30000

Equatorial (110)-Pilatus-slow

Equatorial (110)-Pilatus-slow

30000

Equatorial (110)-Frelon-fast

Equatorial (110)-Frelon-fast
Equatorial (110)-Frelon-slow

Equatorial (110)-Frelon-slow

Daughter (110)-Frelon-slow

110

15000
10000
flow (0.23s)

15000

1000

0.1

flow (0.25s)

0
100

500

time(s)

0
10

0.25s

1000

0.01

5000

1500

10000

5000

0.1

(a.u.)

20000

110

20000

2000

Area

(a.u.)

Daughter (110)-Frelon-slow

0.01

Daughter (110)-Frelon-fast

25000

Daughter (110)-Frelon-fast

Area

Area

110

(a.u.)

25000

0.01

0.1

time(s)

(a)

10

100

1000

time(s)

(b)

Figure 5.8: Time evolution of the area of equatorial and daughter (110) diffractions during
and after flow (a) shear rate = 480 s1 and shear time = 0.23 s and (b) shear
rate = 400 s1 and shear time = 0.25 s.

These strong shear flows work similarly in generating crystalline shish. A difference
is found when comparing parent and daughter diffractions in the late crystallization
stage. As seen in Figure 5.7, for a shear rate of 560 s1 , after 800 s the (110)
diffraction of the daughter lamellae approaches a level comparable with that of
parent lamellae. Differently, Figure 5.8 shows that for 480 and 400 s1 , daughter
diffraction levels exceed that of the parent lamellae. It is found that decreasing the
shear rate from 560 to 400 s1 changes the relative diffraction intensity between
parent and daughter lamellae. Note that for parent and daughter lamellae of iPP,
the Area110 ratio estimated from diffraction pattern is not equal to the real weight
ratio of crystals. [17] Applying geometrical correction [17] gives parent/daughter ratio
at tc =800 s of around 4.4 and 2.7 for 560 s1 and 400 s1 , respectively. We will
not apply this correction further for our results, since it does not work well (for yet
unknown reasons) for the lower shear rates. Moreover, we are more interested in the
relative time scales rather than in absolute values. So the qualitative conclusion can
still be drawn that lowering the shear rate from 560 s1 to 400 s1 decreases the
parent/daughter ratio. For quiescent crystallization, the ratio between iPP parent and
daughter lamellae is a function of crystallization temperature. For the flow-induced
isothermal case, Fernandez-Ballester et al. [18] found that stronger flow can increase the
relative ratio between parent and daughter lamellae, which is consistent with our results.
Shish precursors (400 s1 > shear rate 240 s1 )
When lowering the shear rate to 360 and 240 s1 , shish precursors are generated
instead of crystalline shish. These precursors develop into densely packed structures
upon cessation of flow. The crystallization after shearing at a rate of 320 s1 for 0.25s
is considered as an example. Since no crystal diffractions appear during flow, these

Chapter 5

79

WAXD

flow cessation time


at 0.25s

0.26s

(110)

0.33s

(110)

1s

(110)

7.7s
(130)(040)

Figure 5.9: Pilatus WAXD patterns after flow of 320 s1 (shear time = 0.25 s). Flow
direction is vertical.

12000

15000

10s

10000

10000

daughter
lamellae

flow
5000

flow (0.25s)

8000

flow
direction

6000

2s

4000
2000

0
0.01

Equatorial (110)-Pilatus-slow
Equatorial (110)-Frelon-slow
Daughter (110)-Frelon-slow

7.8s

Area 110 (a.u.)

Area 110 (a.u.)

20000

Equatorial (110)-Pilatus-fast
Equatorial (110)-Pilatus-slow
Equatorial (110)-Frelon-fast
Equatorial (110)-Frelon-slow
Daughter (110)-Frelon-fast
Daughter (110)-Frelon-slow

flow (0.25s)

0
0.1

10

time(s)

(a)

100

1000

0.01

0.1

10

100

1000

time(s)

(b)

Figure 5.10: Time evolution of area of equatorial and daughter (110) diffractions for different
flows (a) shear rate = 320 s1 and (b) shear rate = 240 s1 , with the same shear
time of 0.25 s. Figure 5.10a inset image is tilted to make the weak diffraction
patterns more clear. The arrow shows flow direction. The initial diffraction for
the shear rate of 240 s1 is too weak to quantify the start of crystallization.

Chapter 5

80

frames are not shown. Figure 5.9 shows the WAXD results immediately after flow. No
crystal diffraction appears at 0.26 s (shear time is 0.25 s) and weak (110) diffraction is
observed at 0.33 s. The same holds for the onset time of SAXS streaks, see Table 5.2.
This indicates formation of crystalline shish after flow of 320 s1 .
The ability to distinguish between shish formation during and after flow is crucial
for understanding what is exactly generated by the flow. Most in-situ tracking on
crystallization is performed with acquisition periods which are in the order of seconds,
much longer than 0.033 s we achieved. In that case the exposure period covers not only
the flow duration but also some time after flow, and the structure that appears is very
often thought to be generated during flow. In fact, our results show that validity of
such interpretation depends on the flow conditions. Crystalline shish are indeed formed
during flow for shear rate between 400 and 560 s1 , while for 320 s1 crystalline shish
actually appear after flow, which, in other words, means the structures generated during
flow are just shish precursors.
Moreover, for shear rates from 320 to 560 s1 , once the densely packed structures
are observed, they are crystal, irrespective of whether they appear during or after flow.
For 240 s1 , SAXS streaks are also observed within 1 s but SAXS and WAXD signals
are both too weak to precisely determine whether they appear at the same time.
The nucleation effectiveness of the shear-generated shish precursors are expected
to have similar impact on crystallization as shear-generated crystalline shish. Figure
5.10a shows the crystallization evolution after formation of shish precursors by a flow
at 320 s1 for 0.25 s. Parent lamellae develop continuously after the occurrence of
shish. Daughter lamellae appear later, as also observed for stronger flow conditions.
Figure 5.10a shows that the amount of daughter lamellae is very little at the beginning.
However, daughter lamellae grow quite fast with time.
Threadlike precursors (shear rate = 160 s1 )
Further lowering the shear rate to 160 s1 leads to the formation of threadlike
precursors. This is confirmed by the characteristic birefringence raise during flow
and the absence of SAXS equatorial streaks. Such threadlike precursors comprise
oriented molecular segments and can survive upon cessation of flow. In addition,
Figure 5.5b shows that the residual birefringence signal increases gradually during
subsequent isothermal crystallization. Thus, at 145 C, the residual precursors will
develop into activated nuclei, called threadlike nuclei accordingly, and as a consequence,
start crystallization.
However, the WAXD results do not show any crystallization in the first 2 s (data not
shown). This is due to the low crystallization rate at the beginning, as also shown by
the little rise of birefringence within the initial 2 s in Figure 5.5b.

Chapter 5

81

Figure 5.11a shows 1D WAXD evolution of isothermal crystallization for threadlike


precursors. Results show that the first weak (110) diffraction is observed at 7.8 s and
rises as crystallization proceeds. Although threadlike precursors are not crystals that
provide exactly matched substrates, they are able to effectively trigger crystallization
at 145 C. It is imagined that the stretched and oriented segments arrange into local
ordered aggregations and subsequently promote nucleation. In addition, Figure 5.11b
shows that WAXD (110) and (040) diffractions mainly locate in the equatorial direction,
implying high degree of crystal orientation. The anisotropic growth of crystallization
confirms the orientation of these threadlike precursors. Our results indicate that a
shear rate of 160 s1 generates threadlike precursors, which develop into stable nuclei
to accelerate crystallization kinetics and orient the resulting crystalline structures.

1800

(040)

(110)

Intensity (a.u.)

1600
1400

1000s
1200
1000

(110)
(040)

800

7.8s

amorphous phase

600
8

10

11

12

2T ( )

(a)

(b)

Figure 5.11: (a) 1D WAXD evolution during isothermal crystallization after flow of 160 s1
(shear time = 0.25 s) and (b) 2D WAXD pattern at a crystallization time of
13.5 s. Flow direction is vertical.

It is interesting to find these threadlike precursors, which are birefringence visible


but below the limitation of X-ray characterizations (SAXS and WAXD). When
X-ray technique is used only, revealing threadlike precursors depends on whether
crystallization can be triggered, i.e. the degree of undercooling. When the experimental
temperature is relatively low, undercooling is sufficient to trigger crystallization and,
consequently, make the effect of these flow-induced precursors detectable, as shown
by our results at 145 C. Without crystallization, these threadlike precursors can not
be observed but might survive for long time. For example, Fernandez-Ballester et
al. [15] found that oriented precursors (the same structure indicated by threadlike
precursors in this work) can survive as long as 20 min at 160 C. In this case, cooling
can activate these surviving precursors and thus influence subsequent crystallization.

Chapter 5

82

Needlelike precursors (shear rate = 80 s1 )


Neither X-ray nor birefringence measurements show structure formation during flow
for a shear rate of 80 s1 for 0.25 s. When these direct methods fail in probing
shear-induced precursors/nuclei, the subsequent crystallization can be used to indirectly
explore the consequences of this flow.
The isothermal crystallization process at 145 C after the weakest flow applied (80
s1 for 0.25 s) is shown in Figure 5.12a. Observable (040) and (110) diffractions show
up at a crystallization time of around 200 s. Previous rheological measurements on the
same iPP find that even at a 140 C (lower than the 145 C used in this work), quiescent
crystallization takes more than 1000 s to be detected. The much faster crystallization
kinetics after flow implies the formation of extra nuclei. The nucleation density is lower
than that for other stronger flows, so it takes much more time to form a sufficient
amount of crystals for detection.
1600

Intensity (a.u.)

1400

(040)

(110)

1000s

1200
1000

200s

800

7.8s
600
8

10o

2T ( )

(a)

11

12

(b)

Figure 5.12: (a) 1D WAXD evolution during isothermal crystallization after flow of 80 s1
(shear time = 0.25 s) and (b) 2D WAXD pattern at a crystallization time of
1000 s. Flow direction is vertical.

In addition, the crystallization orientation is studied by looking at the (040)


diffraction which was shared by the oriented parent and daughter lamellae with the
same b-axis. Figure 5.12b shows that the (040) diffraction of crystallization at 1000
s is located in the equatorial direction, implying the shear-induced nuclei are indeed
oriented, although they are not reflected in birefringence. Another issue we want to point
out is the absence of isotropic (040) diffraction in Figure 5.12b. The low orientation
in the late stage is due to the low orientation of nuclei and lamellar curving and
twisting during lateral growth [17], which excludes the possibility of a mixed morphology
comprising of highly oriented lamellae and isotropic spherulites. These nuclei must be

Chapter 5

related to some precursors generated by flow.


In conclusion, one more kind of oriented precursors is found, which is neither
visible for X-ray nor for birefringence. In this work, such precursors are denoted
as needlelike precursors to differentiate them from shish (observed with X-ray)
and threadlike precursors (observed with birefringence). Actually, these needlelike
precursors which are below X-ray and birefringence but are able to ultimately nucleate
oriented crystallization, were hidden in the results of Kumaraswamys systematical
study on flow-induced iPP crystallization. [7, 14] For the material PP-300/6 used by
Kumaraswamy et al., the oriented structures indicated by the birefringence upturn
signal, similar to threadlike precursors in our results, only appear under intense flows,
i.e. for 0.06 MPa the shear time has to be larger than 4 s (for 0.03 MPa a flow time up
to 16 s is insufficient). [7] On the other hand, their X-ray results [14] show that oriented
crystallization occurs for the flow conditions of 0.06 MPa for 2 s and 0.03 MPa for 2 s,
which are actually weaker than the above mentioned threshold for forming birefringence
detectable oriented structures [7].
Crystal orientation
Shear-induced structures provide not only nucleating sites but also orientation
templates for crystal growth. The influences of the different nuclei on orientation are
explored as well. The FWHM values, based on equatorial (110) diffractions, indicating
the orientation degree of crystals affected by various flows and their time evolutions are
shown in Figure 5.13.
For shear rates of 320-560 s1 , crystallization orientation starts with a similar low
FWHM value of around 4 . At the very beginning, crystal growth strictly follows the
orientation of nuclei/precursors, so the initial small FWHM reflects the high orientation
of shish nuclei/precursors generated under these conditions, i.e. shear rates of 320-560
s1 . For a lower shear rate of 240 s1 , the equatorial (110) diffraction is too weak in the
beginning (period of 2 s) to accurately determine the FWHM evolution. However, the
equatorial (110) diffraction is quite narrow azimuthally, see the 2D pattern of Figure
5.10b inset-2s, and qualitatively demonstrates the strong orientation of generated shish
precursors.
Therefore, it is found that shish and shish precursors all possess a high orientation
with respect to the flow direction. Moreover, crystallization after a flow (160 s1 and
0.25 s) is first observed at 7.8 s, rather than during the initial 2 s. The FWHM of early
crystallization is also very small, around 6 at 7.8 s. This indicates that threadlike
precursors are also highly oriented, which is consistent with the birefringence signal
shown in Figure 5.5b.

83

Chapter 5

84

20

30

10

Intensity (a.u.)

20

10

0
50

100

150

200
o

azimuthal angle ( )
30

(c)
Intensity (a.u.)

FWHMequatorial 110 ( )

15

(b)

(a)

-1

240s (0.25s) -- Pilatus


-1
240s (0.25s)-- Frelon
-1
320s (0.25s)-- Pilatus
-1
320s (0.25s)-- Frelon
-1
400s (0.25s)-- Pilatus
-1
400s (0.25s)-- Frelon
-1
480s (0.23s)-- Pilatus
-1
480s (0.23s)-- Frelon
-1
560s (0.20s)-- Pilatus
-1
560s (0.20s)-- Frelon

20

10

50

100

150

200
o

0.1

10

100

1000

azimuthal angle ( )

time (s)

Figure 5.13: (a) FWHM evolution of equatorial (110) diffraction during crystallization after
flows at shear rates from 240 s1 to 560 s1 . The azimuthal distribution of
(110) diffraction after crystallizing for 1000 s after flows at shear rates of (b)
80 s1 and (c) 160 s1 .

Figure 5.13a also shows that the orientation decreases (indicated by increase of
FWHM) as crystallization proceeds. This is due to imperfect lamellar growth: curving
and twisting. [17] It is clear that final FWHM values for the two strongest flows are
still very low, i.e. around 6 . For shear rates of 400, 320 and 240 s1 (the same flow
time 0.25 s), the FWHM reaches various orientation levels of 9.6 , 11 , 14.5 ,
respectively. Figure 5.13b and 5.13c show the overlapping diffraction of parent and
daughter lamellae at t=1000 s for the relatively weak flows (80 and 160 s1 for 0.25 s).
Their final FWHM values are around 50 and 40 , although the latter starts with a high
orientation level. Therefore, the orientation of the ultimate structure depends on both
the orientation of initial nuclei/precursors and on the evolution during crystallization
where the contribution of curving and twisting depends on the available lateral space for
growth, which is associated to the nucleation density. For a higher nucleation density,
the average distance between neighboring nuclei is smaller leaving less space for lateral
growth and, consequently, less room for change in lamellar orientation, while in case of a
low nucleation density the orientation may decrease dramatically because of significantly
curving and twisting during the imperfect lamellar growth. Cavallo et al. [5] found
that when starting with the same oriented nuclei, the final crystallization orientation
decreases with reduction of nuclei density due to the relaxation. Notice that in our
results, the orientation is averaged over all interior layers experienced with various flow
histories, so the final orientation is determined by the corresponding average nuclei
density.

Chapter 5

85

Crystallization polymorphism: -phase


As a polymorphous polymer, sheared iPP melt may crystallize into various crystal
modifications, i.e. the -, - and -phase. In this section, the influence of flow strength
on crystallization polymorphism is discussed for this specific temperature of 145 C.
Figure 5.14 shows the WAXD results for different flows. The longer crystallization
times were chosen to present results after flow at lower shear rates to compensate for the
slower kinetics and thus lower crystallinity levels. The substantial (110) and (040) peaks
of iPP -phase appear for all flow conditions. Differently, the weak (300) diffraction
related to the -phase, can be distinguished only for shear rates from 240 to 560 s1 ,
while for shear rates of 160 and 80 s1 it was not observed, even after crystallizing for
1000 s. The formation of -crystals thus depends on the flow conditions.

Intensity (a.u.)

(110)D

-1

240s
-1
160s
-1
80s
8

10

11

(040)D

12

2T ( )

(300)E
-1

560s 480s-1
-1
400s
-1
320s-1
240s
-1
-1
80s 160s

10
o
2T ( )

11

12

Figure 5.14: 1D WAXD curves at different crystallization times after varying flow strength,
in terms of apparent shear rates: 560 s1 (at 48 s), 480 s1 (at 48 s), 400 s1 (at
500 s), 320 s1 (at 500 s), 240 s1 (at 500 s), 160 s1 (at 1000 s), 80 s1 (at 1000
s). The crystallization temperature is 145 C.

Varga et al. [10, 19] used fiber pulling to shear a polymer melt and found that the
shear-induced -row nuclei can nucleate -crystals. When the growth rate of -crystals
is larger than that of -crystals, typically in a temperature range between 105 and
140 C, the resulting crystalline structure can be polymorphous and consists of mainly
-crystals based on the -phase core. [10] Somani et al. [20] found that, at 140 C,
the amount of -crystal depends on the shear rate for a fixed strain and can go up
to 65%-70% of the total crystallinity. It was suggested that the growth of -crystals is
associated with the total surface area of oriented -crystals. Therefore, in our isothermal
crystallization at 145 C, the substantial growth of parent lamellae of -crystals is
sufficient to nucleate -crystals. The much lower intensity of (300) diffraction compared

Chapter 5

86

to those of (110) and (040) diffractions, qualitatively indicates that the total amount of
-crystals formed is very little.
The orientation of the -crystals formed is explored as well. To quantify the degree of
orientation, the FWHM values of -crystal (300) diffraction are determined and shown
in Figure 5.15. It can be seen that all FWHM values are less than 13 and slightly
decrease with increasing shear rate. Clearly, for all flow conditions applied in this study
the -crystals are oriented even after long-term crystallization for 500 s. This is different
from the observations of Somani et al. [20] and Chen et al. [21] that initially occurring
-crystals are oriented and the further spontaneous growth without any preferential
orientation leads to an oriented-to-isotropic transformation. The appearance of oriented
-crystals in our work can be explained from the growth point of view.
20
30o

FWHM(300) ( )

15
(300)

10

0
160

240

320

400

480

560

640

-1

apparent shear rate (s )


Figure 5.15: FWHM of azimuthal (300) diffraction at the crystallization time of 500 s after
different flows. Inset is the WAXD image for 480 s1 after crystallization for
500 s. Flow direction is vertical.

It is known that -crystals and -crystals both can be nucleated by the same -crystal
assemblies (e.g. -row nuclei) and these two phases are growing in competition. [10, 19]
At 145 C the growth rate of -crystal is faster than that of -crystal. [22] In this
case, continuous growth of -crystals might be suppressed by the growth of -crystals.
For example, oriented nuclei initiate dense appearance of -crystals and -crystals
and both develop outwards. Initially, - and -lamellae are growing in parallel and
perpendicular to the flow direction because of spatial confinement along the nuclei
orientation direction. However, with crystallization, the faster growing front of crystals can exceed to that of -crystals and, may grow into the potential space for their
neighboring -crystals. The initial grown -crystals are oriented and soon stopped by
-crystals. Therefore, the fraction of -crystals is little and keeps a high orientation, as
shown in our results.

Chapter 5

When the crystallization temperature is between 105 and 140 C, where growth rate
of -crystals is slower than -crystals, Varga et al. [10] found that -crystals can be the
dominant crystal. Moreover, van Erp et al. [23] found the prevailing -crystals in ZieglerNatta iPP homopolymer crystallization with shear and a mild pressure (maximum 1200
bar), instead of - or -crystals. This is due to a similar kinetic reason that high
crystallization temperature and pressure result in the faster growth rate of -crystals
than for -crystals. [24]

5.4 Conclusions
We utilized WAXD, SAXS, and birefringence to explore the development of shearinduced structures in iPP starting from the beginning of their creation. Depending on
the flow strength, oriented precursors or nuclei are generated that differ in terms of their
detection level. For strong enough flow, shish were detected during short term flow (0.200.25 s) by means of WAXD/SAXS. Birefringence is used to measure the early stages
of flow induced structures which were below the detection limitation of WAXD/SAXS.
They are also investigated indirectly by measuring the resulting crystallization (faster
kinetics and oriented morphology). The upturn in the birefringence during flow and
the non-zero relaxation after flow demonstrates that threadlike precursors are formed
during flow at 160 s1 for 0.25 s; they have sufficient orientation to be observable by
birefringence. In contrast, needlelike precursors that are generated by the flow at 80
s1 for 0.25 s possess a too low density and orientation to be detected by birefringence.
By combining two different WAXD detectors (Pilatus and Frelon) the entire
crystallization during and after flow could be tracked. It is found that parent
lamellae are directly nucleated on shish formed during flow or developed from
precursors for shish, while threadlike precursors and needlelike precursors need
time to start crystallization. Daughter lamellae show up later than parent lamellae
and the parent/daughter ratio depends on the flow strength. In the early stages
the crystallization orientation follows the high orientation of nuclei. With further
crystallization, imperfect growth (curving and twisting) of lamellae will decrease the
orientation. This effect is stronger for lower average nuclei concentration. It is also
found that both shish and precursors for shish can induce appearance of -phase of
iPP.

References
[1] B. Wunderlich. Macromolecular Physics, Vol 2: Crystal Nucleation, Gorwth, Annealing.
Academic Press, New York, 1976.
[2] M. V. Massa and K Dalnoki-Veress. Phys. Rev. Lett. 92:255509, 2004.

87

88

Chapter 5

[3] L. Kailas, C. Vasilev, J-N. Audinot, H-N. Migeon, and J. K. Hobbs. Macromolecules 40(20):7223
7230, 2007.
[4] B. S. Hsiao, L. Yang, R. H. Somani, C. Avila-Orta, and L. Zhu. Physical Review Letters
94(11):117802, 2005.
[5] D. Cavallo, F. Azzurri, L. Balzano, S. S. Funari, and G. C. Alfonso. Macromolecules 43(22):9394
9400, 2010.
[6] Z. Ma, L. Balzano, and G. W. M. Peters. Macromolecules 45(10):42164224, 2012.
[7] G. Kumaraswamy, A. M. Issaian, and J. A. Kornfield. Macromolecules 32(22):75377547, 1999.
[8] L. Balzano, S. Rastogi, and G. W. M. Peters. Macromolecules 44(8):29262933, 2011.
[9] L. Balzano, D. Cavallo, T. B. van Erp, Z. Ma, J. W. Housmans, L. Fernandez-Ballester, and
G. W. M. Peters. IOP Conf. Series: Materials Science and Engineering 14:01200, 2010.
[10] J. Varga and J. Karger-Kocsis. Journal of Polymer Science: Part B: Polymer Physics 34:657670,
1996.
[11] L. G. Balzano, S. Rastogi, and G. W. M. Peters. Macromolecules 42(6):20882092, 2009.
[12] C. W Macosko. Rheology, principles, measurements and application. Wiley-VCH, New York,
1994.
[13] 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.
[14] G. Kumaraswamy, R. K. Verma, J. A. Kornfield, F. Yeh, and B. S. Hsiao. Macromolecules
37(24):90059017, 2004.
[15] L. Fernandez-Ballester, T. Gough, F. Meneau, W. Bras, F. Ania, J. C. Francisco, and J. A.
Kornfield. Journal of Synchrotron Radiation 15:185190, 2008.
[16] B. Lotz and J. C. Wittmann. Journal of Polymer Science Part B: Polymer Physics 24(7):1541
1558, 1986.
[17] D. M. Dean, L. Rebenfeld, R. A. Register, and B. S. Hsiao. Journal of Materials Science
33(19):47974812, 1998.
[18] L. Fernandez-Ballester, D. W. Thurman, and J. A. Kornfield. Journal of Rheology 53(5):1229
1254, 2009.
[19] J. Varga and J. Karger-Kocsis. Polymer 36:48774881, 1995.
[20] R. H. Somani, B. S. Hsiao, A. Nogales, H. Fruitwala, S. Srinivas, and A. H. Tsou. Macromolecules
34(17):59025909, 2001.
[21] Y. Chen, Y. Mao, Z. Li, and B. S. Hsiao. Macromolecules 43(16):67606771, 2010.
[22] T. B. van Erp. Structure development and mechanical performance of polypropylene. Ph.D. thesis,
Eindhoven University of Technology, Eindhoven, The Netherlands, 2012.
[23] T. B. van Erp, L. Balzano, and G. W. M. Peters. ACS Macro Letters 1(5):618622, 2012.
[24] M. van Drongelen, T. B. van Erp, and G. W. M. Peters. Polymer 53:47584769, 2012.

Chapter six

High-stress shear induced


crystallization in isotactic
polypropylene and
propylene/ethylene random
copolymers

Abstract
Crystallization of an isotactic polypropylene (iPP) homopolymer and two
propylene/ethylene random copolymers (RACO), induced by high-shear stress, was
studied using in-situ synchrotron wide-angle X-ray diffraction (WAXD) at 137 C. The
depth sectioning method was applied in order to isolate the contributions of different
layers in the stress gradient direction and to relate specific structural evolution to the
corresponding local stress. This approach gives quantitative results in terms of the
specific length of fibrillar nuclei as a function of the applied stress. As expected,
crystallization becomes faster with increasing stressfrom the inner to the outer
layerfor all three materials. Stress-induced crystallization in a RACO with 7.3 mol%
ethylene content was triggered at only 1 C below its nominal melting temperature. The
comparison of iPP and RACOs with 3.4 and 7.3 mol% ethylene monomer reveals the
effect of ethylene defects on high-shear stress induced crystallization at 137 C. It is
This chapter is based on : Zhe Ma, Lucia Fernandez-Ballester, Dario Cavallo, Gerrit W. M. Peters.
to be submitted, 2012

89

Chapter 6

90

found that, for a given applied stress, the specific nuclei length that forms during flow
increases with ethylene content, which is attributed to a greater high molecular weight
tail. However, the linear growth rate is significantly reduced by the presence of ethylene
comonomers and it is found that this effect dominates the overall crystallization kinetics.
Finally, a time lag is found between development of parent lamellae and the emergence
of daughter lamellae, consistent with the concept of daughter lamellae nucleated by
homoepitaxy on the lateral faces of existing parent lamellae.

6.1 Introduction
Semi-crystalline polymers are usually processed from their molten state and subjected
to intense shear or/and elongation flows. Such flow fields not only accelerate
crystallization kinetics, which shortens the processing time, but can also change the
morphology from isotropic spherulites to highly oriented shish-kebabs [13] and, as a
consequence, determine the ultimate properties. Therefore, understanding the interplay
between strong flow fields and the resulting structures is of importance for designing
processing procedures to tailor these ultimate properties of products.
Considerable work [47] has been devoted to this topic in the past half century. Many
researchers have focused on the relation between shear flow and polymer crystallization,
because shear fields are easily created with rotational [3] or sliding [8, 9] plate-plate
devices, on rotational rheometers [10, 11] and in pressure-driven slit flows [12, 13]. They
are typically combined with time-resolved characterization techniques like mechanical
spectrometry [10, 11, 14], light scattering [15, 16], birefringence [13], X-ray scattering
[7, 17, 18] and Fourier Transform Infrared Spectroscopy [9, 19, 20]. Significant progress
has been made in understanding some of the fundamental issues in shear-induced
crystallization [47], while others remain unknown. Particularly, the knowledge of
crystallization under high shear rate and stress, close to realistic processing conditions,
is still limited. The need of such information is becoming urgent in order to improve
the latest simulation models, since the results of numerical predictions of, for example,
injection molding [21], have to be validated and further refined with experimental
evidence.
Imposing a strong shear flow at chosen high shear rate or stress under well-defined
conditions requires a specially designed flow device. Both the pressure-driven slit flow
apparatus constructed by Janeschitz-Kriegl et al. [12] and improved by Kornfield et
al. [13], and the piston-driven slit rheometer developed by Mackley et al. [22] and
modified by Peters and coworkers [23, 24] can operate in the high stress region (in the
order of 0.1 MPa) and are easily combined with time-resolved birefringence [23, 25]
or/and X-ray scattering [2629]. The drawback of these channel devices is the nonhomogeneous shear stress distribution over the thickness [30] and, consequently, any

Chapter 6

observation in the shear gradient direction represents an average over the thickness
of the sample. To solve this problem, Fernandez-Ballester et al. recently proposed
and verified the depth sectioning method [29], which takes advantage of the linear
variation of shear stress over the thickness, from a maximum and known shear stress
at the wall to zero at the center of the rectangular channel. This method separates the
contributions from specific layers by performing a series of experiments with varying
wall stress but fixed shearing time.
In this chapter, the pressure-driven flow device [13] and the depth sectioning
method [29] are combined to quantitatively study polymer crystallization induced by
high shear stress. An isotactic polypropylene (iPP) and two propylene/ethylene random
copolymers with various ethylene monomer contents are studied and compared to reveal
the effect of molecular architecture. Recent studies found that adding ethylene monomer
to the propylene chain can improve transparency, relative softness and low-temperature
impact strength [3133]. Also, it has been found that the presence of ethylene monomer
along the polypropylene chain disturbs the chain regularity and, consequently, decreases
polymer crystallization ability [3335], e.g. decreases crystallinity and linear growth rate
and, moreover, induces the formation of the orthorhombic -phase [35, 36]. However,
the effects above have mostly been studied for quiescent crystallization, or under a
rheometric flow [34] unable to impose high shear stresses similar to those typically
encountered in processing conditions. Here, we focus on the influence of the presence of
defects in the molecular architecture on shear-induced crystallization under high stress.
Moreover, we show the importance of the high molecular weight tail on the effect of
flow-induced crystallization.

6.2 Experimental
6.2.1 Materials
The materials used are an isotactic polypropylene (Borealis HD234CF) and
two propylene/ethylene random copolymers (Borealis RD204CF and RD208CF),
polymerized using Ziegler-Natta type catalysts. All three materials have very similar
weight average molecular mass Mw 310 kg/mol and a polydispersity of Mw /Mn 3.4,
but they vary in their ethylene content between 0 7.3 mol%. Their molecular and
physical properties [34] are summarized in Table 6.1. In this study, the homopolymer
is denoted as iPP while the copolymers are denoted as RACO3 and RACO7,
according to their respective ethylene content in mol%.

91

Chapter 6

92

Table 6.1: Molecular and physical properties of iPP homopolymer and ethylene/propylene
random copolymers.

Ethylene content
Xc (%) Tm ( C)(a) Tc ( C)(a)
FTIR (wt%) NMR (mol%)
iPP
HD234CF
0
0
48.3
159
110
RACO3 RD204CF
2.2
3.4
42.3
147
105
RACO7 RD208CF
4.9
7.3
34.4
138
98
(a) Measurements were performed under a heating and cooling rate of 10 C/min

Polymer

Grade

6.2.2 Flow device


Flow-induced crystallization experiments were carried out in a pressure-driven flow
cell designed by Kumaraswamy et al. [13] The flow cell, described previously [13,25,29],
has a shear slit with a rectangular cross-section of 6.35 mm (width) 0.5 mm (thickness)
and a channel length of 63.5 mm. It is equipped with two diamond windows mounted
flush on the slit channel which allow the passage of an X-ray beam through the thickness
of the sample for in-situ measurements. The experimental protocol is as follows: First,
the material in the slit is heated to 215 C and kept at this temperature for 5 min to
erase all thermal and mechanical history. Next, the relaxed melt is cooled to the desired
crystallization temperature T = 137 C. Once the sample is stabilized at 137 C, a shear
pulse is imposed on the molten polymer at a specific value of wall stress (0.11, 0.103,
0.091 and 0.079 MPa) for a fixed duration of 2 s. The sample is held at 137 C after the
shear pulse and the progress of crystallization under isothermal conditions is monitored
by acquiring X-ray diffraction patterns. Then, the depth sectioning method [29] is used
on the diffraction patterns to isolate the structural information from various layers and
relate this to the local stress (see section 6.3).

6.2.3 X-ray characterization


Time-resolved wide-angle X-ray diffraction (WAXD) characterization was carried out
at the BM26B (DUBBLE) beamline at the European Synchrotron Radiation Facility.
[37] The wavelength used was 1.22
A. Two-dimensional (2D) images were recorded with
a Frelon detector with a resolution of 1024 1024 pixels of 97.6 m 97.6 m. The
sample-to-detector distance was 157 mm. The incoming beam intensity was measured
with an ionization chamber to correct for changes in the primary beam intensity. The
data acquisition time was 15 s per image. The shear pulse lasting 2 s was applied
at the beginning of the acquisition of the first diffraction image. Therefore, this first
image combines the information of 2 s of shear and of the subsequent 13 s of isothermal
crystallization, and is noted as corresponding to a crystallization time of 13 s.

Chapter 6

93

6.2.4 Optical microscopy


The linear growth rates were measured by following quiescent melt-crystallization
using a Leica DMLP polarized optical microscope equipped with a 20 objective lens.
The microscope was coupled with a Linkam CSS450 stage to enable a careful control of
the thermal history while acquiring optical micrographs with a dedicated digital videocamera. The samples were initially loaded in the cell as a pellet, melted and compressed
into a film of approximately 20 m thick by moving the stage plates gently towards
each other. The polymer films were annealed for 5 minutes at 210 C and then cooled
to the selected crystallization temperatures at 30 C/min. Optical micrographs were
taken during the isothermal crystallization, with adequate time-resolution. Spherulitic
growth rate was determined by measuring the evolution of the spherulites diameter over
time, by means of an image analysis software. The reported values of growth rate are
the results averaged over three measurements and the reproducibility was within 3%.

6.3 Depth sectioning method


The depth sectioning method [29] uses the linear relationship between layer depth
from the wall and shear stress to separate the local structure in a specific layer, which
is a pre-requisite to reveal the relation between the shear history and the structural
evolution. For slit flow, the shear stress varies linearly along the channel thickness
direction from zero in the center to maximum at walls (see Figure 6.1). Because Xrays propagate through the whole sample along the thickness direction, i.e. the stress
gradient direction, the acquired X-ray patterns correspond to the total diffraction from
all layers. In order to apply the depth sectioning method and separate the diffraction
signal corresponding to a specific sample layer, a set of experiments is performed at
different wall shear stresses while keeping all other parameters fixed (e.g., temperature,
shear duration, crystallization time).
Consider an experiment in which a wall shear stress of max is applied and for which
max
(see Figure 6.1). For this experiment,
the corresponding scattering X-ray signal is Itot
the shear stress d at a specific depth d with respect to the nearest wall is given by
d = max (1

d
)
D

(6.1)

where D corresponds to half the channel thickness, 250 m. According to the depth
sectioning method [29], the contribution of the scattering signal arising from the interior
portion, between centerline D and boundary depth d in such experiment using wall
max
, can be determined by performing another experiment in which the
stress max , IDd
wall shear stress of d is imposed and by subsequently rescaling the obtained scattering
d
d
d
max
.
max
= Itot
) by the stress ratio, i.e. IDd
signal (Itot

Chapter 6

94
V max
2 I tot

V max

wall
V outer

dinner d outer
D

250P m

V inner

V max

I D  dinner

I DV max
 douter

V max
I tot

center
max
I dVinner
 douter

wall
X-ray
Figure 6.1: Schematic of the linear relationship between the layer depth with respect to the
wall, d and its local stress, d .

Figure 6.1 shows that each specific layer has two boundaries, douter and dinner , with
corresponding stresses, outer and inner , respectively. According to the depth sectioning
method [29], the contribution to scattering of the layer located between douter and dinner
max
(Idinner
douter ) to the intensity measured for an experiment where a wall stress max is
max
), can be determined by:
imposed (Itot
max
max
outer
max

Idinner
douter = IDdouter IDdinner = Itot

inner
outer
inner
Itot

max
max

(6.2)

inner
outer
correspond to half of the total intensity signals obtained
and Itot
in which Itot
from two separated experiments using prescribed wall stresses of outer and inner ,
respectively, calculated according to Eq. 6.1.

A series of experiments with wall stresses of 0.11, 0.103, 0.091, 0.079 MPa were carried
out to isolate four layers at depths of 0 16 m (L1), 16 43 m (L2), 43 70 m (L3)
and 70 250 m (L4) from the wall. An example of depth-sectioned patterns for iPP
after 88 s of isothermal crystallization is shown in Figure 6.2. Due to the relatively high
stress, crystallization develops fast in the outer layers L1 and L2, where L1 has a higher
orientation. The core part experiences lowest stress, so in the L4 layer the polymer is
still mainly in the amorphous state after 88 s.
In order to enable the comparison of crystallization between different layers, the
depth sectioned intensities are further normalized by the thickness of each layer d =
douter dinner . The amount of parent lamellae can be extracted by fitting the azimuthal
scan of the (110) diffraction arising from oriented crystals [29] after subtraction of the
isotropic part calculated from the (040) diffraction and after geometrical correction [38].
As an example, the results for layer L1 are given in Figure 6.3. The (110) diffraction

Chapter 6

95

L1: 0-16 m

L2: 16-43 m

L3: 43-70 m

L4: 70-250 m

0.11-0.103 MPa

0.103-0.091 MPa

0.091-0.079 MPa

0.079-0 MPa

Figure 6.2: 2D depth-sectioned diffraction patterns (top row) corresponding to the


crystallization of specific layers (bottom row) in iPP at t = 88 s after a wall
shear pulse of 0.11 MPa and 2 s. The layer depths and corresponding boundary
stresses are indicated. Flow is along the horizontal direction.

area and full width at half maximum (FWHM) from the parent and daughter lamellae
can be determined; they are relative measures of the amount of crystals and degree of
orientation for the parent and daughter lamellae, respectively.

Intensity (a.u.)

parent lamellae

FWHM
daughter lamellae

90

180

270

360

azimuthal angle ( )

Figure 6.3: Azimuthal scan of the oriented (110) diffraction of iPP at t = 88 s for the L1
layer after thickness normalization and geometrical correction. The solid line
corresponds to the Lorentzian fittings.

Chapter 6

96

6.4 Results and Discussion


The depth sectioned X-ray patterns are first used to examine the influence of shear
stress on the crystallization kinetics and orientation of each of the three materials; the
homopolymer (iPP) and two random copolymers (RACO3 and RACO7). Next, the
crystallization kinetics of these three different polymers is compared at a specific level
of shear stress to reveal the effect of the macromolecular architecture, i.e. copolymer
content, on crystallization.

6.4.1 iPP homopolymer


Prior to flow, the diffraction pattern of the iPP shows only an isotropic broad ring
(data not shown), irrespective of the layer, i.e. application of depth sectioning. This is
consistent with the undeformed amorphous melt with no crystallinity and no orientation.
Selected 2D WAXD depth sectioned patterns during shear and following isothermal
crystallization of the iPP for various layers are shown in Figure 6.4.

L1

(040)

(110)
parent
(110)
daughter

13s

28s

43s

208s

13s

28s

43s

208s

L2

(110)
parent

L3
(110)

13s

28s

(110)
daughter

43s

208s

Figure 6.4: WAXD depth-sectioned patterns of isothermal iPP crystallization at 137 C for
a wall shear stress of 0.11 MPa for 2 s. The positions and corresponding stresses
of layer L1 (top row), L2 (middle row) and L3 (bottom row) are illustrated by
Figure 6.2. The flow direction is horizontal.

Chapter 6

The results clearly show that stress has a remarkable influence on triggering
crystallization. The different layers, from L1 to L3, exhibit various crystallization
behaviors because of the decreasing local stress. Figure 6.4-L1 presents the structural
evolution in the outermost layer L1, subjected to the highest stress range 0.11-0.103
MPa. A pair of clear arc-like diffractions emerge quickly after flow at a scattering
vector of q = 10.0 nm1 in the vertical direction perpendicular to shear (Figure 6.4L1-13s). These WAXD diffraction arcs correspond to the (110) diffraction plane [39]
of monoclinic -phase in iPP, which indicates that the oriented -phase forms quickly
after shear and that the c-axis aligns along the flow direction in the so-called parent
lamellae. Differently, crystallization under lower shear stresses in the L2 and L3 layer
is more sluggish and requires a longer time, 28 s, to form detectable parent crystals
(Figure 6.4-L2 and L3). Likewise, in the two outermost layers L1 and L2, daughter
lamellae, also described as lamellar branches at an angle of 80 to a specific parent
lamellar surface [40], are already observed at 28 s as two pairs of (110) diffraction arcs
located close to the meridian (Figure 6.4-L1-28s and L2-28s). In contrast, for the L3
layer, daughter lamellae only appear at later times (43 s).
It should be noticed that only the -phase appears in the 2D diffraction patterns.
Although some studies have observed the emergence of dominant -phase crystals in
shear-induced iPP crystallization, [41, 42] the appearance of only -phase in our results
is consistent with previous studies [27, 29] that found only or predominantly -phase as
a result of shear-induced crystallization.
The results in Figure 6.4 indicate that the imposition of a shear pulse generates
nuclei which can significantly speed up crystallization kinetics and orient the crystal
morphology [12,18,43]. At 137 C, quiescent crystallization of iPP is too slow to generate
any detectable structure within 400 s (data not shown), and only isotropic crystallites
would ultimately form. In contrast, 208 s after the imposition of flow, the diffractions
patterns for the L1 layer are quite narrow in the azimuthally direction (see Figure 6.4L1208s) implying that crystal morphology in the layer that was subjected to the highest
stress range is highly oriented. For the inner layers subjected to lower levels of shear
stress, however, the orientation of structures is qualitatively lower at 208 s. Therefore,
depth-sectioned WAXD images qualitatively show that stress has a significant influence
on the start and evolution of crystallization, which is quantitatively analyzed below.
Next, a quantitative evolution of the amount of crystals and the degree of orientation
is extracted from the area and FWHM of the azimuthal (110) peak corresponding to
the parent lamellae, see Figure 6.5a and b. Irrespective of the layer considered, parent
lamellae grow rapidly in the early stage and then reach a shoulder, after which they
either halt their growth or they continue to grow at a much slower rate. Knowing
that the growth is stopped by the impingement of the growth fronts of the parent
lamellae, a shorter time to reach this shoulder must relate to less space between
neighboring nuclei. Therefore, more nuclei are generated in the outer layer by the

97

Chapter 6

98

6x105

30

L1

4x105

L3

20

2x10

L3

L2

3x105

FWHM ( )

Area

P.110

(a.u.)

5x105

L2
10

L1

1x10

0
0

200

400

600
time(s)

(a)

800

1000

1200

200

400

600

800

1000

1200

time(s)

(b)

Figure 6.5: Evolution of (a) area and (b) FWHM of parent (110) diffraction in iPP
isothermal crystallization in different layers.
Layer positions and their
corresponding stresses are illustrated by Figure 6.2.

higher stress [29]. Interestingly, the crystallization in the L1 layer not only shows the
fastest kinetics at the early stages, but it also has the highest value when it reaches
the shoulder. The larger value results from the combined effect of flow-enhanced
crystallization and parent/daughter lamellae ratio when crystallization is completed,
as found by Fernandez-Ballester et al. [29] that the relative ratio between parent and
daughter lamellae is higher in the outer layer than in the inner layers of lower stress.
Shear-induced nuclei are known to template the oriented growth of parent lamellae
in the early stage. The orientation of parent lamellae was illustrated by the FWHM
of the parent (110) diffraction, see Figure 6.5b. Lower FWHM values refer to a higher
average lamellar orientation. The initial FWHM in the L1 layer is the lowest, around
6 , but the larger ones in the L2 and L3 layers are quite similar ( 9 ) in the first short
period of tens of seconds. These low FWHM values in the initial stage of crystallization
suggest that nuclei generated at various stress levels have a high orientation. This result
is consistent with the observation of Fernandez-Ballester et al. [29], where crystallites
induced by the three strongest conditions all show very high and similar degrees of
orientation at the beginning (see Figure 9b in ref [29]).
As crystallization proceeds, the change in the orientation becomes more pronounced
for the inner layers with lower shear stress. For the highest stress range of 0.11-0.103
MPa the FWHM stays nearly constant during the observation period, whereas those for
0.103-0.091 and 0.091-0.079 MPa vary from 9 to 12 and from 9 to 19 , respectively.
The change in the orientation indicates that lamellar growth does not strictly follow
that of the nuclei or initially grown lamellae because of occurrence of lamellar curving
and twisting during lateral growth. [38] This orientation variation depends on the space
between neighboring nuclei. When nuclei density is smaller, there is more space between
nuclei for lateral growth during which the possibility to curve increases leading to the

Chapter 6

99

reduction of orientation [29, 44]. Therefore, orientation evolution shows that the L3
layer has the least nuclei, which is consistent with the results of Area evolution.

6.4.2 Propylene/ethylene random copolymers


Only three out of the four wall shear stresses for the homopolymer were imposed on
the random copolymers (0.11, 0.103, and 0.091 MPa). For the random copolymers, the
crystallization at a wall stress of 0.091 MPa was quite sluggish, so the stress was not
lowered further to 0.079 MPa. Therefore, the innermost layer of random copolymers
(named L3+L4, see Figure 6.6) should be compared to the sum of the two individual
L3 and L4 layers in iPP. Note that at the same experimental temperature Texp , iPP
and random copolymers have different undercooling T = Tm0 Texp , where Tm0 is the
equilibrium melting temperature, because the addition of ethylene monomer decreases
the equilibrium melting temperature Tm0 [34]. Accordingly, the lamellar linear growth
rate under quiescent conditions decreases with the increase of ethylene content (see
section 6.4.3).

(110)
daughter

L1
q

13s
13s

28s
28s

73s
73s

208s

13s

28s

103s

208s

13s

103s

208s

703s

L2

L3+4

Figure 6.6: WAXD depth-sectioned patterns of isothermal RACO3 crystallization at 137 C


for a wall shear stress of 0.11 MPa for 2 s. The positions and stresses of layer L1
(top row), L2 (middle row) and L3+4 (bottom row) are illustrated by Figure 6.2.
The flow direction is horizontal. Images are tilted to make the weak diffraction
patterns more clear.

Chapter 6

100

Figure 6.6 shows a representative series of depth-sectioned WAXD patterns for


RACO3. The influence of stress is also found to be significant for RACO3: in the
outermost L1 layer, some oriented crystallites can already be observed within the first
13 s after imposing the shear pulse, while in the inner L2 and L3+4 layers, crystallites
can only be detected after 28 s and 103 s, respectively. The faster kinetics in the outer
layer indicates that, as for iPP, increasing applied shear stress induces more nuclei also
for RACO3. Compared with iPP, the crystallization of RACO3 in the L1 and L2 layer
starts at the same time (13 s and 28 s, respectively), but in the L3+4 layer it is much
slower than for iPP in the L3 layer (see Figure 6.4L3-28s and Figure 6.6L3+4-103s).
Interestingly, all layers of RACO3 show a time lag between the development of
parent and daughter lamellae. For instance, in the L1 layer, RACO3 parent lamellae
development is pronounced from 13 to 28 s, while no daughter lamellae are observed at
all at 28 s. Similarly, for low stresses (shown by Figure 6.6L2-28s and Figure 6.6L3+4103s) the first crystals that develop after flow belong to parent lamellae only. This
growth lag between different lamellae is not specific for RACO3; it is also observed
for iPP in the L3 layer (Figure 6.4L3-28s) and for RACO7 in the L1 layer (data not
shown). In fact, this time lag is consistent with the mechanism of initiation of parent
and daughter lamellae: Parent lamellae are templated from shear-induced nuclei while
daughter lamellae are nucleated by the homoepitaxy on the lateral (010) faces of existing
parent lamellae with monoclinic -modification [40]. In other words, daughter lamellae
need parent lamellae to initiate the second-generation growth.
30

6x10

1.5x10

L3+4

1.0x10

5x10

L1

0.0
10

100

1000

4x10

FWHM ( )

Area

P.110

(a.u.)

5.0x10
5

3x10

L2

2x10

20

L2

10

1x10

L3+4

0
0

200

400

600
time(s)

(a)

800

1000

1200

L1
0

200

400

600

800

1000

1200

time(s)

(b)

Figure 6.7: Evolution of (a) area and (b) FWHM of parent (110) diffraction in RACO3
during shear and isothermal crystallization.
Layer positions and their
corresponding stresses are illustrated by Figure 6.2.

RACO3 orientation evolution is shown in Figure 6.7b. Layers L1 and L2 have a


quite similar orientation at the beginning but develop differently with time. This is
qualitatively consistent with the difference in nuclei density in the different layers; again,

Chapter 6

101

the larger space between neighboring nuclei allows for curving and twisting resulting in
a lower orientation (a larger FWHM value).
For an ethylene content of 7.3 mol% in RACO7, the nominal melting temperature
decreases to 138 C. Quiescent crystallization will not be detected at the experimental
temperature of 137 C, since this is just 1 C lower than its nominal melting temperature
and, as a consequence, the quiescent nucleation density and the linear growth rate are
both very small. However, crystallization of RACO7 in L1 layer proceeds immediately
after flow and for the L2 layer, some crystallites become observable at 200 s (Figure
6.8a), providing a clear example of the effect of shear stress on crystallization even in
the vicinity of the nominal melting temperature. The linear growth rate is the same
as under quiescent conditions, so the accelerated rate of oriented crystallization results
from the abundant oriented nuclei generated by the high shear stresses applied.
6x105

30
L3+4 (1003s)

(040)

4x105
(040)

3x105

20
flow

FWHM ( )

Area

P.110

(a.u.)

5x105

L1

2x10

1x105

L1

10

L2

0
0

200

400

600
time(s)

(a)

800

1000

1200

200

400

600

800

1000

1200

time(s)

(b)

Figure 6.8: Evolution of (a) area and (b) FWHM of parent (110) diffraction in RACO7
during shear and isothermal crystallization. Inset is the WAXD image for
RACO7 after crystallization for 1003 s. Layer positions and their corresponding
stresses are illustrated by Figure 6.2.

Comparing with iPP and RACO3, the time at which crystallization can first
be detected for RACO7 is similar in L1 but much slower in L2 and L3. During
crystallization, FWHM varies from 5 to 8 , see Figure 6.8b. Interestingly, the
slow RACO7 crystallization in L1 layer is comparable to that of iPP in L3 layer, but
the orientation in RACO7 is much higher.
Based on above results, a qualitative conclusion can be drawn that for each of the
three materials: the number of nuclei formed increases with applied shear stress, i.e.
from the inner to the outer layers. For a given material, the comparison is simple
because the quiescent growth rate of lamellae is fixed. However, polymers with various
ethylene contents have different quiescent growth rates which affect the crystallization

Chapter 6

102

kinetics. Therefore, to quantitatively study the effects of stress and ethylene content
on polymer crystallization, the kinetic model [4548] is used to estimate the amount of
oriented nuclei formed by shear in the different materials below.

6.4.3 Quantification of nuclei


In the KolmogorovAvramiEvans model [4548], the progress of space filling in
time,(t), can be described by the expression:
(t) = 1 exp(kGm tn )

(6.3)

where k is the factor involving the nuclei density, G the linear growth rate, m the
exponent indicating the growth dimension (1, 2 or 3-directional) and n the nucleation
mechanism (sporadic, n = m + 1 or predetermined, n = m). For shear-induced
crystallization, the number of nuclei is fixed prior to growth and does not increase with
space filling, so the exponent number n equals the growth dimension m. In the present
work, oriented nuclei are dominant and space is mainly filled by the lamellar growth that
develops perpendicular to these nuclei. Structural perfection (e.g. lamellar perfection,
branching and thickening) behind the growth front is not taken into account for space
filling. Therefore, we assume that the space filling (t) is directly proportional to the
development of the parent lamellae diffraction A(t). With this assumption, space filling
can be obtained by:
(t) = (A(t) A0 )/(A A0 )

(6.4)

where A0 is the (110) diffraction area at t = 0 that is caused by flow, and A the
shoulder (110) diffraction area when space filling is completed. Since the first data point
for the L1 layer is obtained after flow and nonzero, it contains information concerning
both the 2s of shear and the 13s of isothermal crystallization, A0 cannot be determined
directly for all L1 layer cases. On the other hand, for most of the results in the L1 layer,
after crystallizing for 13 s, A0 is still very low with respective to the shoulder value, so
the contribution of A0 to space filling is negligible and will be assumed to be 0 in the
calculation of space filling. Therefore, space filling can be assessed by (t) = A(t)
.
A
Assuming that the linear growth rate G is constant in time, the crystallization kinetics
is examined by plotting the rewritten form of Eq. 6.3 (see Figure 6.9):
ln{ln[1 (t)]} = nln(t) + ln(kGm )

(6.5)

The fitted exponent n are all in the range 1.62 (n = 2 is for ideal 2D growth with
predetermined nuclei) while all initial slopes are nearly 2. Therefore, the theoretical
integer exponent n = 2 will be used for the assessment of nucleation density. The

Chapter 6

103

Ln(-Ln(1- ))

-2
iPP_L1
iPP_L2
iPP_L3
RACO3_L1

-4

RACO3_L2
RACO7_L1

Ln(t)

Figure 6.9: Avrami plots of space filling evolution for different materials in various layers.

description of 2-dimensional growth reads [49]:


(t) = 1 exp(lNG2 t2 )

(6.6)

where l is the long period of stacked lamellae, N is the number of lamellae and total
length of nuclei per volume, L, can be easily derived from the time for filling half space,
t1/2 :
lN =L=

ln2
(G t1/2 )2

(6.7)

Note that for the random copolymers the addition of ethylene leads to defects in the
regular polypropylene chain and, consequently, decreases the crystallization ability and
the linear growth rate G.
3.0
iPP

2.5

RACO3

Log(G/(nm/s))

RACO7

2.0

1.5

1.0

0.5

0.0
120

130

140

150

160

temperature ( C)

Figure 6.10: Open points are measured growth rates and solid lines represent linear fittings.

Chapter 6

104

Table 6.2: Fitting parameters and the estimated growth rates at 137 C.

y=a+bT
iPP
RACO3
RACO7

a
12.17
12.24
11.15

b
0.079
0.085
0.079

growth rate at 137 C (nm/s)


24.5
4.4
2.1

The quiescent growth rates for the three materials at different temperatures are
plotted in Figure 6.10. Because the measured temperature range is limited, a linear
function [50] (Log(G) vs. T ) is used to estimate growth rates at 137C, obtaining 24.5,
4.4 and 2.1 nm/s for iPP, RACO3 and RACO7, respectively. All fitting parameters are
listed in Table 6.2. The growth rate of RACO7 at 137C was estimated by extrapolating
experimental results, since 137 C is too high to measure the quiescent growth rate.
Using these growth rate values, we calculated the estimated lengths of the oriented
nuclei per volume given in Table 6.3.

Table 6.3: Total length per volume (L) of oriented nuclei calculated for iPP, RACO3 and
RACO7 for different layers.

Total length per volume


of oriented nuclei (m/m3 )
L1
L1
L3

iPP (1011 )

RACO3 (1011 )

RACO7 (1011 )

5.3
2.8
0.26

34
8.9

69

For each material, the estimated nuclei length per volume of nuclei increases with
increasing stress, i.e. from the inner to the outer layers, consistent with the trend of
faster overall crystallization in the outer layers. In iPP, the nuclei length per volume
generated by the highest levels of stress is in the order of 1011 m/m3 (Table 6.3).
Knowing that the normal long period of iPP is typically tens of nanometers [51], the
number of lamellae growing directly on the oriented nuclei should be in the order of
1019 1/m3 . One could think of each lamella to be nucleated from a single nucleus site
(stress generated) which would imply a much larger number than the shear-induced
nuclei density of 1016 1/m3 in previous studies [52]. Such high nuclei density is
usually approached by adding an efficient nucleating agent [53].
For random copolymers, the high stresses of 0.11 and 0.103 MPa are able to trigger
significant crystallization for a small degree of undercooling, particularly for RACO7,
since the experimental crystallization temperature is just 1 C below its nominal melting
temperature. This effect is comparable with that observed for iPP when stresses between
0.080.19 MPa are imposed at 165 C, 2 C above its nominal melting temperature. [28]

Chapter 6

105

Therefore, for high enough stress (around 0.1 MPa for RACO7), polymer crystallization
can be initiated even in the vicinity of the nominal melting temperature.
It is surprising to see that, for identical flow conditions, the higher regularity of the
chains of iPP, i.e. higher crystallization ability, does not imply a higher total length
per volume of oriented nuclei L (Table 6.3). Notice that the long period of stacked
lamellae (independent of shear) and nuclei density N (dependent on shear) together
determine the total length of oriented nuclei per volume L = l N. One could suggest
that the difference in the long period for the three materials leads to the varying total
length of oriented nuclei. However, Hosier et al. [51] used AFM and found that the
long period at 110 C will decrease with addition of ethylene defects. Based on their
data, the estimated long periods of our materials at 110 C change from 20 nm for
iPP to 15 and 10 nm for random copolymer with 3.4 mol% and 7.3 mol% ethylene
monomer, respectively. Even though the long period at 137 C might be different from
the above numbers for 110 C, it still can be concluded that the lower total nuclei length
for iPP is not caused by its larger long period, because the long period, l, decreases
with increase of ethylene content and results in a larger difference in nucleation number
density, N = L/l.
The influence of the stress history on crystallization is determined by the response
of a polymer at the molecular level, i.e. the molecular stretch [21]. As described
by the nucleation and growth model [21], the total length of nuclei is determined by
both continuous generation of new nuclei N and the growth L of these nuclei during
shear. The nucleation rate depends on the stretch HMW of high molecular tail and
the longitudinal growth rate on the average molecular stretch AVG . Therefore, the
high molecular tail and average stretch are playing different roles in increasing the
amount of oriented nuclei. For a pressure-driven flow device used in the present work,
imposing the same shear stress ensures that the average stretch is the same, irrespective
of ethylene content in polymer. However, a difference in stretch history of the high
molecular tail may cause a significant change in nuclei quantity. In fact, the longest
relaxation time of the RACOs, determined from dynamic rheological measurements,
is larger than for the iPP homopolymer; 1.46, 2.07 and 3.14 s for iPP, RACO3 and
RACO7 at 220 C, respectively. [34] Since the temperature dependence of relaxation
time follows the Arrhenius equation, the longest relaxation times at 137 C are known
1
T
according to ref
= exp( E
( 1 Tref
)) with the activation energy E (43.0, 42.04 and
R T
45.19 kJ/mol for iPP, RACO3 and RACO7, respectively) [34] and universal gas constant
R. It is found that the longest relaxation times of RACO3 and RACO7 are 16.5 and
29.2 s at 137 C, respectively, which are 1.4 and 2.4 times larger than that of 12.2 s in
iPP. Thus, for the same imposed stress, the high molecular weight tail becomes more
oriented with increasing ethylene content, during flow. Moreover, the stretch history of
high molecular weight tail lasts longer in random copolymers, after cessation of flow.
These two aspects result in an enhanced effect when subjected to the same stress. In the

Chapter 6

106

same layer, an identical stress causes a larger influence on crystallization with increasing
ethylene content, which is consistent with our estimation of oriented nuclei shown in
Table 6.3. Under such high stress of around 0.1 MPa, flow-induced nucleation in this
set of Ziegler-Natta type homopolymer and random copolymers is dominated by their
high molecular weight tails.
Although the quantity of nuclei is higher for the random copolymers, the growth rate
of random copolymers is much lower than that of iPP. For 2D growth, crystallization
kinetics is determined by the total length of nuclei and the square of growth rate,
(t) = 1 exp(LG2 t2 ), so the overall crystallization kinetics is still dominated by
the growth rate and decreases with increasing ethylene content.

6.5 Conclusions
Using a pressure-driven slit flow device and the depth sectioning method, the
crystallization of an iPP homopolymer and two random copolymers with 3.4 and 7.3
mol% ethylene were studied. For the same material, the crystallization rate becomes
faster going from the inner to outer layers because of the increase in stress. With
crystallization, the emergence of daughter lamellae is found to occur later than the
development of parent lamellae. The lager number of oriented nuclei in outer layer
leads to a smaller orientation change during crystallization because of less space between
neighboring nuclei, i.e. the absence of space for lamellar curving and twisting. The high
stress can generate up to 1011 m/m3 of oriented nuclei and start the crystallization of
RACO7 in vicinity of its melting temperature. High-stress induced nuclei are quantified
using kinetic analysis and the results show that the total length of nuclei in iPP is less
than that in random copolymers. The increase of nuclei length with ethylene is explained
by the longer largest relaxation time for random copolymers, which determines the
molecular stretch of the longest molecules and thus the nucleation rate (as given by the
nucleation and growth model). However, since the growth rate is reduced significantly
by adding ethylene monomer, the crystallization kinetics, dominated by growth rate, is
still faster for the homopolymer than for the random copolymers, even with less nuclei.

References
[1] A. Pennings and A. Kiel. Colloid and Polymer Science 205(2):160162, 1965.
[2] A. Keller. Philosophical Magazine 2(21):11711175, 1957.
[3] B. S. Hsiao, L. Yang, R. H. Somani, C. Avila-Orta, and L. Zhu. Physical Review Letters
94(11):117802, 2005.

Chapter 6

107

[4] 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.
[5] A. Keller and H. W. H. Kolnaar. Flow-induced orientation and structure formation,. In: H. E. H.
Meijer, editor, Processing of Polymers, number 18 in Materials Science and Technology: a
Comprehensive Treatment, 189268. Wiley-VCH, Weinheim, 1997.
[6] G. Kumaraswamy. Journal of Macromolecular Science, Part C: Polymer Reviews 45(4):375397,
2005.
[7] R. H. Somani, L. Yang, L. Zhu, and B. S. Hsiao. Polymer 46(20):85878623, 2005.
[8] F. Langouche. Macromolecules 39(7):25682573, 2006.
[9] H. An, B. Zhao, Z. Ma, C. Shao, X. Wang, Y. Fang, L. Li, and Z. Li. Macromolecules 40(14):4740
4743, 2007.
[10] S. Vleeshouwers and H. E. H. Meijer. Rheologica Acta 35(5):391399, 1996.
[11] N. V. Pogodina and H. H. Winter. Macromolecules 31(23):81648172, 1998.
[12] S Liedauer, G Eder, H Janeschitz-Kriegl, P Jerschow, W Geymayer, and E. Ingolic. International
Polymer Processing 8:236244, 1993.
[13] G. Kumaraswamy, R. K. Verma, and J. A. Kornfield. Review of Scientific Instruments 70(4):2097
2104, 1999.
[14] J. W. Housmans, R. J. A. Steenbakkers, P. C. Roozemond, G. W. M. Peters, and H. E. H. Meijer.
Macromolecules 42(15):57285740, 2009.
[15] N. V. Pogodina, S. K. Siddiquee, J. W. van Egmond, and H. H. Winter. Macromolecules
32(4):11671174, 1999.
[16] J. Baert and P. Van Puyvelde. Macromolecular Materials and Engineering 293(4):255273, 2008.
[17] L. Balzano, N. Kukalyekar, S. Rastogi, G. W. M. Peters, and J. C. Chadwick. Physical Review
Letters 100(4):048302, 2008.
[18] L. Balzano, S. Rastogi, and G. W. M. Peters. Macromolecules 44(8):29262933, 2011.
[19] H. An, X. Li, Y. Geng, Y. Wang, X. Wang, L. Li, Z Li, and C. Yang. The Journal of Physical
Chemistry B 112(39):1225612262, 2008.
[20] Y. Geng, G. Wang, Y. Cong, L. Bai, L. Li, and C. Yang. Macromolecules 42(13):47514757,
2009.
[21] F. J. M. F. Custodio, R. J. A. Steenbakkers, P. D. Anderson, G. W. M. Peters, and H. E. H.
Meijer. Macromolecular Theory and Simulations 18(9):469494, 2009.
[22] M. R. Mackley and D. G. Hassell. Journal of Non-Newtonian Fluid Mechanics 166(910):421456,
2011.
[23] J. W. Housmans, L. Balzano, D. Santoro, G. W. M. Peters, and H. E. H. Meijer. International
Polymer Processing 24(2):185197, 2009.
[24] Z. Ma, L. Balzano, and G. W. M. Peters. Macromolecules 45(10):42164224, 2012.
[25] G. Kumaraswamy, A. M. Issaian, and J. A. Kornfield. Macromolecules 32(22):75377547, 1999.
[26] L. Fernandez-Ballester, T. Gough, F. Meneau, W. Bras, F. Ania, J. C. Francisco, and J. A.
Kornfield. Journal of Synchrotron Radiation 15:185190, 2008.
[27] G. Kumaraswamy, R. K. Verma, J. A. Kornfield, F. Yeh, and B. S. Hsiao. Macromolecules
37(24):90059017, 2004.
[28] L. Balzano, D. Cavallo, T. B. van Erp, Z. Ma, J. W. Housmans, L. Fernandez-Ballester, and
G. W. M. Peters. IOP Conf. Series: Materials Science and Engineering 14:01200, 2010.
[29] L. Fernandez-Ballester, D. W. Thurman, and J. A. Kornfield. Journal of Rheology 53(5):1229
1254, 2009.

108

Chapter 6

[30] C. W Macosko. Rheology, principles, measurements and application. Wiley-VCH, New York,
1994.
[31] M. Gahleitner, P. Jskelinen, E. Ratajski, C. Paulik, J. Reussner, J. Wolfschwenger, and W. Neil.
Journal of Applied Polymer Science 95(5):10731081, 2005.
[32] M. Ll. Maspoch, J. Gamez-Perez, E. Gimenez, O. O. Santana, and A. Gordillo. Journal of Applied
Polymer Science 93(6):28662878, 2004.
[33] J. Gamez-Perez, P. Muoz, O. O. Santana, A. Gordillo, and M. Ll. Maspoch. Journal of Applied
Polymer Science 101(5):27142724, 2006.
[34] J. W. Housmans, G. W. M. Peters, and H. E. H. Meijer. Journal of Thermal Analysis and
Calorimetry 98(3):693705, 2009.
[35] K. Jeon, Y. L. Chiari, and R. G. Alamo. Macromolecules 41(1):95108, 2008.
[36] I. L. Hosier, R. G. Alamo, P. Esteso, J. R. Isasi, and L. Mandelkern. Macromolecules 36(15):5623
5636, 2003.
[37] 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.
[38] D. M. Dean, L. Rebenfeld, R. A. Register, and B. S. Hsiao. Journal of Materials Science
33(19):47974812, 1998.
[39] J. W. Housmans, M. Gahleitner, G. W. M. Peters, and H. E. H. Meijer. Polymer 50(10):2304
2319, 2009.
[40] B. Lotz and J. C. Wittmann. Journal of Polymer Science Part B: Polymer Physics 24(7):1541
1558, 1986.
[41] R. H. Somani, B. S. Hsiao, A. Nogales, H. Fruitwala, S. Srinivas, and A. H. Tsou. Macromolecules
34(17):59025909, 2001.
[42] Y. Chen, Y. Mao, Z. Li, and B. S. Hsiao. Macromolecules 43(16):67606771, 2010.
[43] G. Kumaraswamy, R. K. Verma, A. M. Issaian, P. Wang, J. A. Kornfield, F. Yeh, B. S. Hsiao,
and R. H. Olley. Polymer 41(25):89318940, 2000.
[44] D. Cavallo, F. Azzurri, L. Balzano, S. S. Funari, and G. C. Alfonso. Macromolecules 43(22):9394
9400, 2010.
[45] A. N. Kolmogorov. Bull. Acad. Sci. USSR. Ser. Math 3:355360, 1937.
[46] M. Avrami. Journal of Chemical Physics 7(12):11031112, 1939.
[47] M. Avrami. Journal of Chemical Physics 8(2):212224, 1940.
[48] U. R. Evans. Transactions of the Faraday Society 41:365374, 1945.
[49] B. Wunderlich. Macromolecular Physics, Vol 2: Crystal Nucleation, Gorwth, Annealing.
Academic Press, New York, 1976.
[50] F. L. Binsbergen and B. G. M. de Lange. Polymer 11(6):309332, 1970.
[51] I. L. Hosier, R. G. Alamo, and J. S. Lin. Polymer 45(10):34413455, 2004.
[52] H. Janeschitz-Kriegl, E. Ratajski, and M. Stadlbauer. Rheologica Acta 42(4):355364, 2003.
[53] Z. Ma, R. J. A. Steenbakkers, J. Giboz, and G. W. M. Peters. Rheologica Acta 50(11-12):909915,
2011.

Chapter seven

Conclusions and recommendations

7.1 Conclusions
This study is devoted to gain understanding on how flow introduces extra nucleation
and morphology changes in polymers by using advanced experimental methods such
as high speed X-ray scattering. High density polyethylene (HDPE) and isotactic
polypropylene (iPP) were chosen as representative materials. The goal was to design
experiments at conditions comparable to real processing conditions and use these to
measure time resolved structure development in terms of quantities that are predicted
by state of the art models [1] for flow induced crystallization of polymers, i.e. number
and dimensions of crystalline structures such as spherulites and shish-kebab and the
orientation of these.
Shear significantly accelerates crystallization kinetics by increasing the amount of
nuclei and generates an anisotropic morphology by inducing orientation. These effective
nuclei can be directly generated during flow or developed from precursors which
appear within flow. This thesis focuses on formation of shear-induced precursors/nuclei
and ultimately attempts to reveal the resulting effects on crystallization. Therefore,
several novel methods are developed and applied to achieve qualitative and quantitative
knowledge on shear-induced substances.
Rheological measurements provide a convenient method to determine nucleation
densities. Using a recently proposed suspension-based rheological model [2], the
number of point-like nuclei in quiescent and mildly sheared pure and nucleated
(U-Phthalocyanine) iPPs, is assessed, see Chapter 2. Results show that UPhthalocyanine is quite efficient for nucleating iPP; it raises the nucleation density
by six decades for quiescent crystallization. Moreover, it was found that the effect
109

110

Chapter 7

of shear is enhanced by the presence of the nucleating agent. The rheological


method is easy to apply since rheometers are readily available in most academic
and industrial labs.
When shear-induced oriented precursors cannot be detected directly; crystallization has to be triggered, to reveal the nature of these precursors. This can
be reflected in the crystallization features such as kinetics and orientation. In
Chapter 3, a pressure quench method is described, i.e. a way to mimic a
temperature quench, which provides a convenient and flexible way to obtain
sufficient under-cooling to start crystallization. This method effectively lightens
up the formation of precursors, which leads to accelerated kinetics and oriented
morphology. With the pressure quench, temperature gradients can be avoided.
Moreover, it allows for reverse cooling by depressurizing and, by varying pressure
histories, for applying complex thermal histories. It was found that the precursors
with different stability can be formed by shear and the least stable ones relax
back to the melt, resulting in higher fraction of twisted lamellae in the ultimate
crystalline structure of PE. Depressurization before crystallization completion
leads to partially melting of the crystals, which is explained by the variation
in lamellar stability.
For X-ray observable crystalline shish nuclei, the combination of a slit flow
with high speed synchrotron X-ray measurements, provides a method for in-situ
structural characterization during flow, see Chapter 4. The rheological and
structural evolution during and after short-term were studied for iPP. It was
found that, depending on the shear strength, both, a rise of the apparent viscosity
and, with some delay ( 0.1 s), the formation of crystalline shish can occur
during flow. The formation conditions and appearance time of crystalline shish
are determined. These rheological and structural changes demonstrate that
these flows do not satisfy the basic requirement for what is called a short-term
flow [3], a flow where no rheological changes occur during flow. The viscosity
rise may be explained by the creation of shish or precursors for shish that act as
physical cross-links and that become detectable after the observed delay time.
However, we cannot be sure about the exact value of the delay time; i.e. if the
viscosity rise and the observation of shish are directly related. The influence of
the pressure on the local values of rheological and kinetic parameters will cause
nucleation events to occur first upstream which starts the apparent viscosity
rise while the X-ray measurements are done half way the slit. Only a numerical
model can help to reveal this complex interaction. For that, the results should be
combined with a detailed model for flow induced crystallization including a fully
characterized non-linear viscoelastic model [1, 4] from which the relaxation times
are coupled to the structural development.

Chapter 7

111

In Chapter 5 it is shown that combing the advanced experimental methods (insitu X-ray and birefringence measurements) distinguishes the various flow-induced
structures. By tracking the isothermal crystallization, a direct relation between
nuclei and crystallization behavior, including kinetics, orientation and -crystals,
are determined for iPP.
The influence of molecular structure on flow induced crystallization was
investigated (see Chapter 6). Crystallizations in an iPP homopolymer and two
rheologically similar ethylene/propylene random copolymers are compared after
imposing the same stress. By using a pressure-driven slit flow [5] and application
of the depth sectioning method [6] crystallization kinetics can be related to
different stress levels, i.e. the amount of fibrillar nuclei length per volume can
be related to the stress history. Stress induced nuclei are quantified using kinetic
analysis [710] and the results show that the total length of fibrillar nuclei is less
in iPP than in random copolymers. The increase of nuclei length with ethylene
is explained by the longer largest relaxation time for random copolymers, which
determines the molecular stretch of the longest molecules and thus the nucleation
rate. However, since the growth rate is reduced significantly by adding ethylene
monomer, the crystallization kinetics, dominated by growth rate, is still faster for
the homopolymer than for the random copolymers, even with less nuclei.

7.2 Recommendations
In this thesis, a variety of methods is developed and applied to probe and quantify
shear-induced precursors/nuclei. These methods involve rheometry combined with
a suspension model for a crystallizing polymer, advanced experimental setups and
corresponding experimental protocols. In particular, the application of a multi-pass
rheometer (MPR) provides a strong tool to explore the crystallization under processingrelevant conditions, i.e., intensive flow and mutual influence between flow and pressure.
Some challenges are emphasized for future work.
The main recommendations of this work are:
Using a suspension-based rheological model developed by Steenbakkers et al. [2],
we were able to derive the space filling of growing spherulites from rheological
measurements. This model is applied to a iPP system with a colored nucleating
agent subjected to quiescent and mild flow conditions. The results in Chapter 2
demonstrate the success of this rheological model in assessing the amount of pointlike nuclei for such a complex system. It is known that a crystalline morphology
can change from isotropic spherulites to an oriented structure when the flow is
strong enough. Therefore, it would be very useful to derive a rheological model
that is capable of capturing oriented crystallization as well.

112

Chapter 7

The pressure quench method presented in Chapter 3 provides an alternative


way of achieving sufficient under-cooling to start crystallization of shear-induced
precursors. A pressure quench imposed immediately after flow effectively triggers
crystallization and consequently reveals the generation of precursors. Moreover,
the crystallization after an extra annealing step between cessation of flow and
application of a pressure quench shows partial relaxation, i.e., the variation in
stability, of precursors generated. In Chapter 3, the annealing process is performed
at only a single set of experimental parameters (apparent wall shear rate = 67
s1 , temperature = 134 C, pressure = 50 bar). A more systematic study should
be carried out to explore the dependency of precursor stability on flow strength,
temperature and, more interestingly, on pressure.
In Chapter 4, the early stage of structure evolution during flow is explored in
depth using a slit rheometer and ultra-fast X-ray characterizations. Both pressure
rise and start of crystallization are observed during flow if the flow is strong
enough, and a time delay ( 0.1 s) is found between pressure rise (averaged over
the whole channel) and X-ray signal (specific to the observation window in the
middle of the slit). Our microscopy images (see Figure 4.10 and 4.12) show the
inhomogeneous morphology along the flow direction due to the pressure gradient,
i.e. it demonstrates the large effect of pressure on crystallization kinetics. Such
complexity suggests that above time delay may (or partially) result from the
earlier crystallization upstream than in the X-ray observation window. It is really
a challenge to experimentally determine the precise time delay and to capture this
delay using advanced modeling [11] of flow and pressure enhanced crystallization.
To avoid pressure gradients, an extensional flow rheometer [12] can be used to
explore the relation between change of rheological property and crystallization
under atmospheric pressure. To further explore pressure effect, a dilatometry flow
device is suggested. For instance, extended dilatometry (Pirouette PVT) [1315]
can achieve both homogeneous shear and desired pressure. Further combination
of PVT and X-ray would be very helpful.
The entire crystallization evolution was monitored for a slit flow; from start
of flow until complete crystallization. The results in Chapter 5 show that
crystal orientation is quite high initially because of the highly oriented nuclei
and decreases during crystallization due to the lamellar curving and twisting [16].
The reduction of orientation implies that lamellar growth is not perfect twodimensional, i.e. only perpendicular to the nuclei orientation. These observations
provide experimental data for modification of shear-induced crystallization
models.
In Chapter 6, the influence of molecular structure on crystallization is discussed
by comparing crystallization of iPP homopolymer and two propylene/ethylene

Chapter 7

113

random copolymers after the same shear. This set of materials is synthesized with
Ziegler-Natta type catalysts and has the heterogeneous distribution of ethylene
monomers along polymer chain. Knowing this, the question how ethylene defects
that are homogeneously distributed will influence polymer crystallization emerges.

References
[1] F. J. M. F. Custodio, R. J. A. Steenbakkers, P. D. Anderson, G. W. M. Peters, and H. E. H.
Meijer. Macromolecular Theory and Simulations 18(9):469494, 2009.
[2] R. J. A. Steenbakkers and G. W. M. Peters. Rheologica Acta 47(5-6):643665, 2008.
[3] S Liedauer, G Eder, H Janeschitz-Kriegl, P Jerschow, W Geymayer, and E. Ingolic. International
Polymer Processing 8:236244, 1993.
[4] H. Zuidema, G. W. M. Peters, and H. E. H. Meijer. Macromolecular Theory and Simulations
10(5):447460, 2001.
[5] G. Kumaraswamy, R. K. Verma, and J. A. Kornfield. Review of Scientific Instruments 70(4):2097
2104, 1999.
[6] L. Fernandez-Ballester, D. W. Thurman, and J. A. Kornfield. Journal of Rheology 53(5):1229
1254, 2009.
[7] A. N. Kolmogorov. Bull. Acad. Sci. USSR. Ser. Math 3:355360, 1937.
[8] M. Avrami. Journal of Chemical Physics 7(12):11031112, 1939.
[9] M. Avrami. Journal of Chemical Physics 8(2):212224, 1940.
[10] U. R. Evans. Transactions of the Faraday Society 41:365374, 1945.
[11] T. B. van Erp. Structure development and mechanical performance of polypropylene. Ph.D. thesis,
Eindhoven University of Technology, Eindhoven, The Netherlands, 2012.
[12] Y. Liu, W. Zhou, K. Cui, N. Tian, X. Wang, L. Liu, L. Li, and Y. Zhou. Review of Scientific
Instruments 82(4):045104, 2012.
[13] M. H. E. Van der Beek, G. W. M. Peters, and H. E. H. Meijer. Macromolecules 39(26):92789284,
2006.
[14] M. H. E. Van der Beek, G. W. M. Peters, and H. E. H. Meijer. Macromolecules 39(5):18051814,
2006.
[15] T. B. van Erp, L. Balzano, and G. W. M. Peters. ACS Macro Letters 1(5):618622, 2012.
[16] D. M. Dean, L. Rebenfeld, R. A. Register, and B. S. Hsiao. Journal of Materials Science
33(19):47974812, 1998.

Samenvatting

Bij het vervaardigen van polymeren producten wordt het materiaal in gesmolten
toestand gebracht en vervolgens gevormd. Specifieke eigenschappen kunnen worden
gecreeerd, zoals een hoge stijfheid in een richting middels verstrekken of het vormen
van geometrische complexe producten via spuitgieten. In ongeveer tweederde van
toepassingen worden semi-kristallijne polymeren gebruikt. Tijdens het stromen in het
vormgevingsproces wordt niet alleen het kristallisatieproces versneld, maar kunnen ook,
in plaats van gemiddeld isotrope spherulitische kristallijne structuren, sterk georinteerde
(cilindervormige) structuren worden gevormd. Deze zogenaamde shish-kebabs vertonen
een overeenkomstige vorm met dit van oorsprong Persische voedsel en bepalen in
hoge mate de uiteindelijke producteigenschappen. Het is daarom van belang om de
wisselwerking tussen stroming en de resulterende structuren te begrijpen, waardoor het
mogelijk wordt de gewenste procedures en procescondities te ontwikkelen om benodigde
eindeigenschappen te verkrijgen.
Het hoofddoel van dit onderzoek is daarom het ophelderen van de relatie tussen
specifieke stromingscondities en de resulterende kristallijne structuren. De gebruikte
methode is het bestuderen van de eerste stadia in het kristallisatie proces, waarin de
vorming van precursors en kiemen de overhand heeft. Een precursor is een geordende
locale structuur die niet groeit maar wel kan fluctueren in grootte, en die dus ook kan
verdwijnen. Hij kan echter ook uitgroeien tot kiem.
Dit is van belang, immers kristallisatie van polymeren is een twee-staps proces:
kiemvorming en groei. De eerste stap bepaalt in kwantitatieve (aantal) en in
kwalitatieve (isotroop of georinteerd) zin wat voor kiemen er worden gevormd en
daarmee in grote mate hoe verdere groei van kristallijne structuren zal plaatsvinden.
Kiemvorming kan tevens sterk worden beinvloed door het toevoegen van kiemvormers
en/of door de invloed van stroming, meer in het bijzonder door de invloed van
stromingsgradienten. Vorming en specifieke eigenschappen van kiemen zijn daarom
sleutelfactoren die de uiteindelijke kristallijne structuren bepalen. Het meten en
kwantificeren van precursors en kiemen omvat deshalve het grootste deel van dit
115

116

Samenvatting

onderzoek.
Overeenkomstig de resulterende morfologie kunnen kiemen worden opgedeeld in
twee groepen: puntvormige en georienteerde kiemen. Uit de eerste groep vormen
zich later spherulieten, uit de tweede voornamelijk de shish-kebabs: een dunne
cylinder met transversale lamellen. Afhankelijk van of ze waarneembaar zijn met
rontgendiffractiemetingen (de belangrijkste experimentele methode die in dit onderzoek
is gebruikt)worden georienteerde kiemen ook wel onderscheiden in zogenaamde shishkiemen en rij-kiemen. De laatsten worden nog onderscheiden in draad- of naaldvormige
kiemen, als alternatieve technieken als dubbele breking worden gebruikt.
In het eerste deel van dit proefschrift wordt de vorming van extra puntvormige
kiemen, in een isotactisch polypropyleen, door middel van toevoeging van een
kiemvormer (U-Phthalocyanine) beschreven. Hierbij wordt zowel de situatie met als
zonder een afschuifstroming beschouwd en in beide situaties verhoogt de kiemvorming
dramatisch. De kiemdichtheid is zo hoog, en daarmee worden de resulterende kristallijne
structuren zo klein, dat ze niet meer zichtbaar kunnen worden gemaakt met behulp van
normale optische microscopie. Daarom is gebruik gemaakt van een reologisch suspensie
model dat de fractie kristallijn material relateert aan het mechanische gedrag. Wanneer
de groeisnelheid van de kristallijne structuren bekend is, kan via dit model het aantal
kiemen worden afgeschat.
In het tweede deel van dit onderzoek worden de resultaten omtrent precursors en
rij-kiemen in een bi-modale polyetheen beschreven. Omdat rij-kiemen niet direct
detecteerbaar zijn met behulp van rontgendiffractie, wordt gebruik gemaakt van
plotse drukverhoging om de smelttemperatuur, en daarmee de onderkoeling, snel te
verhogen (een zogenaamde pressure quench). Het blijkt dat afschuivings-genduceerde
precursors kunnen worden gevormd bij temperaturen die nagenoeg gelijk zijn aan de
evenwichtssmelt-temperatuur, en die bovendien bij die temperatuur slechts langzaam
relaxeren.
Vervolgens is de vorming van naaldvormige kiemen tijdens stroming bestudeerd.
Hiervoor is gebruik gemaakt van de combinatie van een snelle rontgendiffractie methode
(30 beelden/s) en reologische metingen. Er blijkt een kritische waarde voor de
afschuifsnelheid te bestaan voor het al dan niet ontstaan van deze naaldvormige kiemen
tijdens een (zeer korte) stroming van 0.25 s. Wanneer georienteerde precursors ontstaan
tijdens de stroming ontwikkelen deze zich tot naaldvormige kiemen na het stoppen
van de stroming. Bij deze experimenten is aangetoond dat reologie gevoeliger is dan
rontgendiffractie, gegeven de gevoeligheid van de huidige detectoren.
Behalve door externe stromingseffecten wordt de vorming van kiemen ook beinvloed
door de interne moleculaire structuur van het onderhavige polymeer. Daarom is
in het laatste deel van het onderzoek de invloed van moleculaire eigenschappen op
afschuivingsgenduceerde kristallisatie onderzocht. Dit is gedaan aan de hand van

Samenvatting

117

een isotactisch polypropyleen en twee propyleen/etheen random copolymeren met


varieerende hoeveelheid etheen. Deze drie materialen hebben een nagenoeg gelijk
reologisch gedrag; er is alleen een klein maar belangrijk verschil in de grootste
relaxatietijd. Omdat voor het homo-polymeer deze relaxatietijd het laagst is, is ook
de door stroming verhoogde kiemdichtheid het laagst voor dit materiaal. Echter,
de verlaagde groeisnelheid ten gevolge van de toegevoegde etheen comonomeer leidt
uiteindelijk toch tot een lagere kristallisatiesnelheid voor de twee random copolymeren
bij vergelijkbare thermische condities.

Acknowledgements

During the past four years, many people have helped me a lot, which lead to the
completion of this thesis and my enjoyable life in The Netherlands. It is my pleasure to
use this last part of my thesis to express my sincere gratitude and thanks to all of you.
First, I would like to thank Prof. Han Meijer and Prof. Gerrit Peters for the
opportunity to study in Eindhoven. Han, I really appreciate such rare chance of pursuing
my PhD in m@te group. During my stay, your innovative thinking continuously inspires
me and your encouragement also brought me a lot of confidence.
I would like to particularly thank Prof. Gerrit Peters, my daily supervisor. This thesis
could not have been finished without your professional guidance and continuous support.
Thank you, Gerrit, for your countless work from the initial proposal, constructive
discussions, to the tireless corrections of the thesis, etc. I learned a lot from your
direct and effective way of working and thinking, especially how to clearly understand
the physical picture behind a chaotic presentation and how to make life simple. I also
truly appreciate your extra inputs on my personal development and your patience of
bearing my boring questions.
My Eindhoven dream came true also thanks to the introduction and recommendation
from Mr. Jacques Joosten (DPI) and Prof. Liangbin Li (USTC). I would like to thank
both of them.
Concerning the abundance of X-ray experiments in the thesis, all members of our
beam-time team, Luigi, Tim, Dario, Peter, Martin, Lucia deserve my appreciations.
Guys, thank you all for the help and support during the stressed moments (e.g. oil
leakage) and also the funs we have experienced together (e.g. cleaning the oil?). In
particular, I am deeply indebted to Luigi for his constant and valuable helps with
experiments, data analysis and discussions, etc. Luigi, owing to your professional xray eyes, I really enjoyed working with you on that reciprocal space. In addition, I am
grateful to Dr. Giuseppe Portale and Wim Bras for the strong support at DUBBLE
beamline BM26 (ESRF).

119

120

Acknowledgements

One reason I enjoy working in Eindhoven is that we have a very friendly and open
environment here. First of all, I want to express my gratitude to Marleen and Yvon for
all their help to make my stay easier and smooth. Next, I would like to thank my room
mates: Michiel, Isa, Lambert, Sebastiaan, Sam, Nick, Iaroslav for the pleasant office
atmosphere. Also, I am thankful to other m@te members: (former) Tom, Rudi, JanWillem, Amin, Joris, Young Joon, Frederico, Arash; (present) Leon, Patrick, Markus,
Lambert, Martien, Peter, Leo, Dirk, Marc, Danqing, Yang, Ye, Carina, Zahra, Amin,
Oleksandr, Daniel, Fabio, Jang Min, Panayiotis, and others not mentioned here, for all
the happy time we experienced together during the past four years. Although we are
colleagues for a very short time, I believe that we can be friends for ever.
My life in The Netherlands is always enjoyable thanks to my friends who I
met mainly in Eindhoven and Amsterdam. Here I would like to acknowledge, (in
GEM) Zhipeng&Yanru, Miao, Lei, Shuiquan; (in Helix) Piming&Xiaoxia, Donglin,
Chunxia, Weizhen, WuJing, DaiMian, Yulan, Jiaqi, Tamara, Camille, Yogesh,
Maurizio; (in Amsterdam) Shoumin&LiDi, Anbang&HuoChao, QuanWei, ZhuHao,
Fangyong&Hairong, Yifan, LiChao, Longyuan&ZhangZhen, SongYang, Shangsong,
ZhaoJing, ZhouJing, Zhongyu, YangQiang, FuJian; and my old friends in delft: Haining
and Guanglin. Thanks a lot for your kindness and hospitality.
Last but not least, I want to express my gratitude to my family for their unconditional
love. Papa and mama, thank you for the endless cares and constant supports, which
are always accompanying me under any circumstance. The special love goes to my wife.
Dear Liyuan, your accompanying and constant love make everyday a great time. Your
tolerating, support and encouragement help me go through all the different moments.
Language is not enough to express my appreciation; I will use the rest of my life to love
you.

Zhe
Eindhoven, October 2012

Curriculum vitae

The author, Zhe Ma was born in Hejian, Hebei Province, China, on December 13th
1983. After finishing his bachelor study in 2005 at the Zhejiang University of Technology,
he studied Material Processing Engineering in the group of Prof. Zhongming Li at
Sichuan University in Chengdu. During 2006-2008, he completed his master thesis
Study on the relationship between crystal and mesophase in isotactic polypropylene
under the supervision of Prof. Liangbin Li at University of Science and Technology of
China in Hefei.
In 2008, he started his PhD study in the Polymer Technology group of Prof.
Han Meijer at Eindhoven University of Technology under the supervision of Prof.
Gerrit.W.M. Peters. During his PhD study, the author completed Polymer Physics,
Polymer Properties and Rheology and Polymer Processing modes of the course
Registered Polymer Scientist (RPK) organized by the National Dutch Research
School of Polymer Science and Technology (PTN).

121

List of publications

This thesis has resulted in the following publications:


Z. Ma, R.J.A. Steenbakkers, J. Giboz, G.W.M. Peters. Using rheometry to
determine nucleation density in colored system containing a nucleation agent.
Rheologica Acta 50:909915, 2011.
Z. Ma, L. Balzano, G.W.M. Peters.
Pressure Quench of flow-induced
crystallization precursors. Macromolecules 45:42164224, 2012.
Z. Ma, L. Balzano, T. van Erp, G. Portable, G.W.M. Peters. Short-term flow
induced crystallization in isotactic polypropylene: how short is short? to be
submitted, 2012.
Z. Ma, L. Balzano, G. Portable, G.W.M. Peters. The influence of flow induced
precursors and nuclei on crystallization of isotactic polypropylene. to be submitted,
2012.
Z.Ma, L. Fernandez-Ballester, D. Cavallo, G.W.M. Peters. High-stress shear
induced crystallization in isotactic polypropylene and propylene/ethylene random
copolymers. to be submitted, 2012.
L. Balzano, D. Cavallo, T.B. van Erp, Z. Ma, J.W. Housmans, L. FernandezBallester, G.W.M. Peters. Dynamics of fibrillar precursors of shishes as a function
of stress. IOP Conf. Ser.: Mater. Sci. Eng. 14:012005, 2010.
L. Balzano, Z. Ma, D. Cavallo, T.B. van Erp, L. Fernandez-Ballester, G.W.M.
Peters. Molecular aspects of the transformation of oblong density fluctuations
into shish-kebabs. to be submitted, 2012.
The author contributed to several publications outside the scope of this thesis:
Z. Ma, C. Shao, X. Wang, B. Zhao, X. Li, H. An, T. Yan, Z. Li, L. Li.
Critical stress for drawing-induced crystal-mesophase transition in isotactic
polypropylene. Polymer 50:27062715, 2009.
123

124

List of publications

C. Shao, Z. Ma, R. Zhuo, R. Zhang, C. Shen. Inhomogeneous deformation of


crystalline skeleton of syndiotactic polypropylene under uniaxial stretching. J.
Mater Sci 47:33343343, 2012.
P. Ma, Z. Ma, W. Dong, Y. Zhang, P.J. Lemstra. Structure-property relationships
of partially crosslinked poly(butylene succinate). Macromol. Mater. Eng. DOI:
10.1002/mame.201200209, 2012.
Y. Liu, K. Cui, N. Tian, W. Zhou, L. Meng, L. Li, Z. Ma, X. Wang. StretchInduced Crystal-Crystal Transition of Polybutene-1: An in Situ Synchrotron
Radiation Wide-Angle X-ray Scattering Study. Macromolecules 45:27642772,
2012.
B. Zhao, X. Li, Y. Huang, Y. Cong, Z. Ma, C. Shao, H. An, T. Yan, L. Li.
Inducing Crystallization of Polymer through Stretched Network. Macromolecules
42:14281432, 2009.
X. Li, J. Sun, Y. Geng, X. Wang, Z. Ma, C. Shao, X. Zhang, C. Yang, L.
Li. Inducing New Crystal Structures through Random Copolymerization of
Biodegradable Aliphatic Polyester. Macromolecules 41:31623168, 2008.

You might also like