Download as pdf or txt
Download as pdf or txt
You are on page 1of 47

Review

Cite This: Chem. Rev. 2018, 118, 1840−1886 pubs.acs.org/CR

Multiscale and Multistep Ordering of Flow-Induced Nucleation of


Polymers
Kunpeng Cui,† Zhe Ma,*,‡ Nan Tian,§ Fengmei Su,† Dong Liu,∥ and Liangbin Li*,†

National Synchrotron Radiation Laboratory, Chinese Academy of Sciences Key Laboratory of Soft Matter Chemistry, and Anhui
Provincial Engineering Laboratory of Advanced Functional Polymer Film, University of Science and Technology of China, 96
Jinzhai Road, Baohe District, Hefei 230026, People’s Republic of China

Tianjin Key Laboratory of Composite and Functional Materials, School of Materials Science and Engineering, Tianjin University,
92 Weijin Road, Nankai District, Tianjin 300072, People’s Republic of China
§
Ministry of Education Key Laboratory of Space Applied Physics and Chemistry and Shanxi Key Laboratory of Macromolecular
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Science and Technology, School of Science, Northwestern Polytechnical University, 127 Youyi West Road, District Beilin, Xi’an
710072, People’s Republic of China

Key Laboratory of Neutron Physics and Institute of Nuclear Physics and Chemistry, China Academy of Engineering Physics, 64
Downloaded via TU EINDHOVEN on December 16, 2022 at 09:42:47 (UTC).

Mianshan Road, Mianyang, Sichuan 621999, People’s Republic of China

ABSTRACT: Flow-induced crystallization (FIC) is a typical nonequilibrium phase


transition and a core industry subject for the largest group of commercially useful
polymeric materials: semicrystalline polymers. A fundamental understanding of FIC can
benefit the research of nonequilibrium ordering in matter systems and help to tailor the
ultimate properties of polymeric materials. Concerning the crystallization process, flow
can accelerate the kinetics by orders of magnitude and induce the formation of oriented
crystallites like shish-kebab, which are associated with the major influences of flow on
nucleation, that is, raised nucleation density and oriented nuclei. The topic of FIC has
been studied for more than half a century. Recently, there have been many
developments in experimental approaches, such as synchrotron radiation X-ray scattering, ultrafast X-ray detectors with a
time resolution down to the order of milliseconds, and novel laboratory devices to mimic the severe flow field close to real
processing conditions. By a combination of these advanced methods, the evolution process of FIC can be revealed more
precisely (with higher time resolution and on more length scales) and quantitatively. The new findings are challenging the
classical interpretations and theories that were mostly derived from quiescent or mild-flow conditions, and they are triggering
the reconsideration of FIC foundations. This review mainly summarizes experimental results, advances in physical
understanding, and discussions on the multiscale and multistep nature of oriented nuclei induced by strong flow. The multiscale
structures include segmental conformation, packing of conformational ordering, deformation on the whole-chain scale, and
macroscopic aggregation of crystallites. The multistep process involves conformation transition, isotropic−nematic transition,
density fluctuation (or phase separation), formation of precursors, and shish-kebab crystallites, which are possible ordering
processes during nucleation. Furthermore, some theoretical progress and modeling efforts are also included.

CONTENTS 3.3.3. Ordering Parameter of Flow-Induced


Precursors 1849
1. Introduction 1841 3.4. Influences of Precursor on Crystallization 1850
2. Conformational Ordering Induced by Flow 1842 3.4.1. Nucleation 1850
2.1. Flow-Induced Conformational Ordering 1842 3.4.2. Crystallization Kinetics 1850
2.1.1. Coil−Helix Transition 1842 3.4.3. Morphologies 1852
2.1.2. Gauche−Trans Transition 1844 4. Factors Influencing the Formation of Flow-
2.2. Flow-Induced Isotropic−Nematic Transition 1845 Induced Precursors 1852
3. Precursors Induced by Flow 1845 4.1. Flow Parameters 1852
3.1. Flow-Induced Precursors 1846 4.2. Temperature 1853
3.2. Theoretical Considerations for Flow-Induced 4.3. Molecular Parameters 1855
Precursors 1846 5. Correlation between Deformation at Whole-
3.3. Nature of Flow-Induced Precursors 1847 Chain Scale and Nucleation 1855
3.3.1. Existence of Flow-Induced Precursors 1847 5.1. Shish-Kebab Crystallite 1855
3.3.2. Are Flow-Induced Precursors Crystalline
or Not? 1848
Received: August 17, 2017
Published: January 19, 2018

© 2018 American Chemical Society 1840 DOI: 10.1021/acs.chemrev.7b00500


Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

5.2. Hierarchical Structure of Shish-Kebab 1856 has become a unique aspect in the crystallization field.
5.3. Coil−Stretch Transition in Polymer Solution 1857 Obviously, a fundamental understanding of FIC is of great
5.4. Shish-Kebab Formation in Polymer Melt 1857 importance for both industrial and academic communities.
5.5. Other Mechanisms for Shish-Kebab Forma- After decades of effort, two major influences of flow on
tion 1858 crystallization have been revealed: acceleration of crystallization
5.6. Advances in Understanding of Shish-Kebab 1859 kinetics and formation of oriented crystallites, which are
5.6.1. Coil−Stretch Transition or Stretched associated with raised nucleation density and oriented
Network Model? 1859 nuclei.11−15 According to the ultimate morphology (spherulites
5.6.2. Is Long Chain the Only Contributor for or oriented texture), flow-induced nuclei can be roughly divided
Shish Formation? 1861 into pointlike nuclei and oriented nuclei. Formation of pointlike
5.6.3. How Does Polydispersity Influence the nuclei has been recently reviewed by Peters et al.16 and will not
Threshold for Shish Formation? 1861 be included in the present review. However, a more complex
5.6.4. How Does Short-Chain Matrix Influence
situation is the formation of flow-induced oriented nuclei. A full
Shish Formation? 1862
understanding of oriented nuclei has not been reached yet, even
5.6.5. In Situ Observation of Evolution of
after decades of effort.17−19 Typical unsolved questions include
Shish-Kebab 1862
5.6.6. Shish Formation under Near-Equilibrium
the ordering nature and precise evolution process of the oriented
and Far-from-Equilibrium Conditions 1863 nuclei. On one hand, for a polymer chain, the intrinsic structures
5.6.7. Is Shish Formation a Single-Stage or at various length scalessuch as segmental conformation,
Multistage Process? 1864 segmental orientation, density fluctuation, stretch of the whole
5.6.8. Is Shish a Thermodynamic Phase or a chain, and so oncan all be arranged in an orderly fashion
Kinetic State? 1866 during the crystallization process, especially under the effort of
6. Theories and Models of Flow-Induced Nucleation 1867 flow. Is there a basic ordering parameter or a combination of
6.1. Molecular Theory of Flow-Induced Nuclea- multiscale orderings to define nuclei? Do these flow-induced
tion 1867 ordering processes at various length scales evolve simultaneously
6.1.1. Crystallization Induced by Stretching of or sequentially? On the other hand, numerous experimental
Cross-Linked Network 1867 results indicate that nucleation under flow does not simply
6.1.2. Flow-Induced Nucleation in Free Melt 1869 follow the classical nucleation theory to form stable crystalline
6.1.3. Quantitative Description of Extension- nuclei directly and may undergo some intermediate states. In
Induced Nucleation 1871 this case, when and how the structural intermediates, if they
6.2. Macroscale Continuum Modeling 1873 exist, occur during flow-induced nucleation (FIN) should be
7. Final Remarks 1878 answered.
Author Information 1879 This review attempts to emphasize the possible processes of
Corresponding Authors 1879 ordering in various length scales. For this purpose, this work is
ORCID 1879 organized by mainly following a bottom-up length scale. We first
Notes 1879 focus on the formation of conformational ordering on a
Biographies 1879 segmental scale (e.g., helix) in section 2, which acts as a basic
Acknowledgments 1879 ordering unit to arrange into a three-dimensionally ordered
References 1879 crystal unit. Then we introduce flow-induced precursors (FIP),
the presence of which can be demonstrated but about which
1. INTRODUCTION detailed structural information has not been revealed, and their
structural analysis in section 3. Concerning precursors, the
Semicrystalline polymers cover more than two-thirds of the factors influencing their appearance are summarized to shed
global applications of polymeric materials. Just due to their light on the formation mechanism in section 4. Then, whether
intrinsic capability to arrange macromolecular segments into deformation on the whole-chain scale is essential for FIC is
three-dimensionally ordered crystal units, even in various discussed with the discovery of shish in section 5. At last, some
modifications, the ultimate products are endowed with rich theoretical and modeling efforts are compared and validated
and unique properties.1−3 In practice, polymer processing often
with experimental results in section 6 to explore possible
starts with molten materials and subsequently utilizes flow to
molecular origins. In the meantime, possible multistep processes
transport the polymeric melt and/or to shape into final
products.4−8 This means that flow is an unavoidable factor in for structure formation are also discussed for FIP and shish in
polymer processing, and flow-induced crystallization (FIC) is a sections 3 and 5, respectively.
general phenomenon for semicrystalline materials. Moreover, Note that the following multiscale and multistep aspects of
crystallization is a typical first-order phase transition existing in FIN are generic for polymer crystallization. Especially, the
macromolecular systems. A polymer is a unique substance with nonequilibrium ordering understood from macromolecular
anisotropic molecular properties, where atoms along the chain systems may be valid for other types of matter, like small
direction are connected by chemical bonds. Thus, a polymer that molecules. However, the concrete manifestation of these aspects
initially is in quiescent random-coil conformation can be easily depends on the polymer investigated. For example, conforma-
deformed into an oriented or stretched state, and the tional ordering arranges into zigzag conformation for poly-
thermodynamic state of the system is varied correspondingly.9,10 ethylene (PE),20 syndiotactic polystyrene (sPS),21 and poly-
The system under such disturbed molecular ordering and (ethylene terephthalate) (PET)22 but forms helices for isotactic
thermodynamic state crystallizes in a different manner with polypropylene (iPP),23 isotactic polystyrene (iPS),24 and
respect to the quiescent case or small-molecule system, and this poly(ethylene oxide) (PEO).25
1841 DOI: 10.1021/acs.chemrev.7b00500
Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

Figure 1. Sketch of coil−helix transition at strained end-to-end distance R. The chain is simplified into a diblock polymer composed of coil and helical
domains.

2. CONFORMATIONAL ORDERING INDUCED BY conformational ordering including coil−helix transition (section


FLOW 2.1.1) and gauche−trans transition (section 2.1.2) will be
discussed.
Since a long-chain polymer needs to orderly arrange its short
2.1.1. Coil−Helix Transition. On the premise of dense
stems into a crystal lattice, the short stems must adopt proper
accumulation and positioning into a crystal unit, polymer chains
conformations for ultimate regular packing. For FIN, basic
adopt conformations with the lowest energy. Helix is a preferred
structural ordering concerns the change of segmental con-
state not only in crystals of synthetic polymers, like iPP, iPS, and
formation. However, in classic nucleation theory, FIN was
PEO, but also in some biological macromolecules, such as DNA
described in a coarse-grained way. It was often thought that
and actin filaments.39,40 The coil−helix transition of biopol-
when external flow is imposed, polymer chains are deformed
ymers induced by deformation has gained much attention
into an oriented or stretched state. The resulting orientation or
because biological functions are closely associated with the
stretch of the chain reduces system entropy and consequently
specific structure of the formed helices.41 To better understand
elevates the melting temperature according to the entropy-
the coil−helix transition, we first have a look at some theoretical
reduction model (ERM),26 which introduces an extra under-
works about how flow influences conformational transition, and
cooling for crystallization. This physical picture is essentially
then we discuss the relevant experimental results.
reasonable, since it has been widely demonstrated that flow can
Theoretical understanding of coil−helix transition can be
effectively enhance crystal nucleation.27 However, molecular
dated back to the work of Zimm and Bragg,42 in which the
details like intramolecular conformational ordering, intermo-
transition was described on a coarse-grained level. In the
lecular positional ordering, and orientation ordering28,29 are
Zimm−Bragg model,43 segments along a polymer chain are
ignored in ERM. In principle, flow can not only give rise to
assumed to have only two states, random-coil and helical. The
orientation or stretch of individual chain but also change
state of a long chain can be described by a simple sequence of
conformational ordering. Moreover, recent ultrafast X-ray
segmental status (cchhcchccchh...), where c and h represent coil
scattering results revealed a constant critical strain for nucleation
and helix, respectively. A monomer within a helix gains potential
in a wide temperature range under strong flow, which invalidates
energy Δh by forming hydrogen bonds but bears a loss of
the prediction of strain−temperature equivalence in classical
entropy Δs. (Here the notation for energy or entropy is
ERM.30 In this section, we mainly discuss the change of
lowercase for monomer, but uppercase for total system.) The
molecular details induced by flow. The ultrafast X-ray scattering
free energy of each monomer in the helical state can be
results will be discussed in section 5.6.6.
estimated by Δf = Δh − TΔs. The monomer located at the
Flow can induce conformational ordering. When the
interface between helical and coil domains suffers entropy
concentration or length of conformational ordering segments
reduction but does not form hydrogen bonds, resulting in an
exceeds a certain threshold, those ordered stiff segments may
increase in free energy of Δf t = −Δh compared to the monomer
spontaneously aggregate and organize, as happens in isotropic−
in helical state. Knowing the microscopic state of the whole
nematic transition.31,32 If this is the case, nucleation seems to be
chain and the energetic contributions of each segment, the total
completed via multiple steps from the melt, which has been
free energy ΔF can be obtained by a statistical mechanics
evidenced by experimental observations, especially under flow
method. This approach was further extended to deal with the
conditions.33−35 In this section, we decouple the global
case where an external extension is imposed by Tamashiro and
nucleation induced by flow into two ordering aspects: (i) how
Pincus36 and Buhot and Halperin.44,45 In their approaches, the
flow induces conformational ordering of individual segments
chain is simplified into a diblock polymer composed of coil and
and (ii) possible aggregation and parallel permutation of the
helical domains with an interface energy of Δf t (Figure 1). This
formed stiff segments.
means that the mixing entropy from reordering of domains and
Before going into a deep discussion, it should be first clarified
the interfacial energy are disregarded. The detailed study of
that intermolecular and intramolecular orderings do not
Buhot and Halperin45 demonstrated that this simplification has
necessarily occur simultaneously but may happen individually.
a slight influence on the outcome only quantitatively but not
As an example, stretching-induced formation of protein α-helix
qualitatively. In this case, the total free energy can be expressed
in biopolymers is a conformational ordering without positional
as follows:
ordering.36 On the other hand, PE can crystallize into hexagonal
crystals under high pressure, which lacks full conformational
3(R − γaNϕ)2
ordering.37,38 ΔFch = ϕN Δf + 2Δft +
2(1 − ϕ)Na 2
2.1. Flow-Induced Conformational Ordering
Although early investigations on FIC can be dated back to the where N is the total number of monomers and ϕ is the
1960s, conformational ordering of chain segments did not get concentration of helical segments. The last term describes the
the attention it deserves. Here, the formation of intramolecular free energy of a coil domain, where R is the end-to-end distance
1842 DOI: 10.1021/acs.chemrev.7b00500
Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

Figure 2. (a) IR spectra of iPP under quiescent crystallization at 140 °C and (b) the time evolution of normalized relative intensities of different
conformational bands. Reprinted with permission from ref 50. Copyright 2007 American Chemical Society.

Figure 3. IR spectra of iPP before and after shear at 140 °C and (b) time evolution of normalized relative intensities of different conformational bands.
Reprinted with permission from ref 50. Copyright 2007 American Chemical Society.

of chain, γ is a factor to characterize the shortening of effective absorption bands in Fourier transform infrared spectra (FTIR)
polymer length due to formation of helix, and α is monomer has been well documented.50,51 For iPP, the absorption bands at
length. This model provides a quantitative correlation between 940, 1220, 1167, 1303, 1330, 841, 998, 900, 808, 1100, and 973
helix concentration and molecular deformation, which can be cm−1 indicate helical structures with different ordering degrees
used to study conformational transition in FIC, as synthetic from high to low, of which the 1220, 841, 998, and 973 cm−1
polymers and biological macromolecules share the same physics. bands correspond to helical lengths with monomer numbers 14,
For network, Courty et al.41,46 and Kutter and Terentjev47 12, 10, and 5, respectively.49,52 Besides, the concentration of
formulated a theoretical work by uniting statistical mechanisms these helices can be indicated by the absorbance intensity.
of single helix-forming chain under extension (as just described) Flow-induced conformational ordering of iPP coil−helix
with the phantom-chain network approach. It was theoretically transition has been studied by An et al.50,51 and Geng et al.53
predicted that extension promotes the formation of helical Figures 2 and 3 compare the IR spectra of quiescent
domains under appropriate conditions, which is in agreement crystallization and FIC of iPP at 140 °C. In Figure 2a, the
with experimental results.47 observation of bands at 973, 808, and 900 cm−1 demonstrates
Compared with biological macromolecules, flow-induced that the short helices already exist in amorphous iPP melt. Under
coil−helix transition was less studied in synthetic polymers, quiescent conditions, the intensities of those IR bands remain
either theoretically or experimentally. As polymer melt is a almost unchanged in the initial 4 h (see Figure 2b). However,
transient network constructed by entanglements, change of end- when flow is applied, the intensities of IR bands rise sharply just
to-end distance by flow should influence the coil−helix after flow, suggesting the increase in population of helices. As
transition, as in biological macromolecules. However, it is not crystallization proceeds further, long helices appear and grow.
easy to trace the coil−helix transition in polymers. Helix is a The saturation time for those IR bands, which under quiescent
short-range order and cannot be characterized with regular conditions is 20 h (see Figure 2b), is shortened to around 100
scattering techniques like X-ray scattering, which is sensitive to min by shear flow (see Figure 3b). The observation of long
long-range correlation. In this aspect, IR spectroscopy shows helices is often considered as a signature of crystallization.
superiority to probe conformational ordering. Middle IR bands Obviously, these results demonstrated that flow can promote
are mainly vibrational bands due to the existence of dipole formation of helices and thus crystallization in synthetic
moment. Therefore, even though the vibrational modes are the semicrystalline polymers.
same, differences in helical length and local environment can According to the Zimm−Bragg model,46 end-to-end distance
vary the wavenumber of absorption bands. This was revealed by is an essential parameter to determine the ratio between coil and
the calculations of Snyder and Schachtschneider48 and the helix. This is qualitatively confirmed in a cross-linked iPP by Su
experiments of Zhu et al.49 Length and concentration of helices et al.,54 who observed a direct correlation between helical
are two important parameters for coil−helix transition, which content and stress. As stress is determined by the end-to-end
can be characterized by the position and intensity of the distance, helical content is related to end-to-end distance in
characteristic absorption peaks, respectively. iPP with 3/1 helix cross-linked iPP. However, polymer melt is a transient network
conformation is chosen here as a representative, as the one-to- constructed by entanglements, and there must exist a
one correspondence between helix length of segments and competition between deformation and relaxation of polymer
1843 DOI: 10.1021/acs.chemrev.7b00500
Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

Figure 4. Schematic of three growth pathways for flow-induced coil−helix transition. Reprinted with permission from ref 50. Copyright 2007 American
Chemical Society.

chain. Thereby, the fraction of helix formed in total sequence is growth. The third one is the merging of short helices, that is, an
influenced not only by flow strength, like strain or strain rate, but incorporation process.
also by temperature that determines chain relaxation. An et al.51 Yamamoto55 used molecular simulation to study the coil−
employed the ratio between height values of band 998 cm−1 after helix transition by quickly drawing an amorphous iPP to about 9
and before shear flow to indicate a conformational order ratio times elongation within about 350 ps. Statistical results showed
(COR) and assessed the effect of various external parameters on that the weight-average length of helical sequence increases with
flow-induced helix quantitatively. Under a fixed flow strength elongation, while elongation has no influence on the length of
with strain rate of 13.5 s−1 and strain of 3375%, COR induced at nonhelical sequences. The detailed distribution of the
temperatures below 150 °C decreases linearly with increasing population of helical sequences as a function of their length
temperature due to the competition between deformation and indicated that short sequences dominate the initial stage of
relaxation. For temperatures above 205 °C, COR has negligible elongation and long sequences increase considerably with
change after flow, probably because relaxation dominates this further elongation. A threshold length of helix with about 13
process. On the other hand, at a constant temperature of 145 °C monomers was observed during elongation, in line with the IR
with strain rate of 6 s−1, a minimum strain is required to experimental results of An et al.50
significantly vary COR, in line with the prediction of Kutter and 2.1.2. Gauche−Trans Transition. In addition to helical
Terentjev47 that a threshold extension is required to induce conformation, planar zigzag is another conformation in
coil−helix transition. As strain increases, COR first increases and crystallites of many polymers, such as sPS,24 nylon,56 PET,57
then reaches saturation at strain larger than about 4000%. and PE.20,58 For example, sPS has a trans sequence of (TT)2 in
Furthermore, when temperature and strain are both fixed, COR crystals, with T representing trans conformation. It was reported
increases with rising strain rate. Therefore, An et al.51 pointed that within the crystallization induction period the content of all-
out that, due to relaxation effect on COR, a combination of trans sequence increases, whereas the sequence containing
intense flow (high strain rate or/and high strain) and high gauche conformation decreases.24 For nylon, all-trans sequence
temperature has a similar effect as combining weak flow (low can be induced by strong flow, as happens in electrospinning.56
strain rate or/and low strain) and low temperature. In drawing of PET film, gauche−trans transition may also take
Besides helical population, Geng et al.53 also revealed the place to form a mesomorphic phase constituted of trans
influence of flow on length of formed helices. Unlike the short conformation. 57 In the following, PE is chosen as a
helices that already exist in original amorphous melt, long helices representative to discuss the gauche−trans transition.
with more than 12 monomers can be found only under flow. A PE is the simplest semicrystalline polymer and often serves as
critical flow strength is required to generate long helices. At 150 a model material to study phase transitions such as
°C, an 841 cm−1 band (representing the helix with 12 crystallization, melting, and so on.4,59,60 PE can crystallize in
monomers) appears only for flow rates larger than 18 mm3/s, various modifications depending on crystallization conditions.
while a flow rate of 45 mm3/s is required to induce the For instance, mobile hexagonal crystals may appear under high
occurrence of a 1220 cm−1 band (corresponding to the helix pressure and then transform to orthorhombic crystals with
with 14 monomers). The flow rate thresholds for appearance of increasing thickness.61−63 Here we discuss the formation and
the 1220 and 841 cm−1 bands increase with temperature, evolution of locally ordered trans-rich structure in the early stage
suggesting that stronger flow is necessary to generate longer of crystallization studied with Raman and FTIR spectrosco-
helices at higher temperatures. Geng et al.53 also found that, at py.20,58,64,65
temperatures close to the melting point, the intensity of 841 Using time-resolved FTIR and small-angle X-ray scattering
cm−1 band or helices with 12 monomers increases monotoni- (SAXS), Sasaki et al.65 concluded that, in the early stage of
cally with time upon cessation of flow, which is accompanied by quiescent crystallization, disordered conformation of trans
a decrease in band intensity of short helices. In addition, an sequence first increases and then remains at the maximum for
induction time exists for helices with 12 monomers. On the basis a while. After that, the content of regular trans sequences, which
of those results, they proposed three pathways for the formation can be used to indicate the formation of orthorhombic crystals,
of long helices (see Figure 4). The first way is flow-induced increases simultaneously with the decrease of disordered trans
growth of the short helices that already existed in amorphous sequences. In contrast, the structural intermediate composed of
melt, without primary nucleation, known as propagation. The all-trans configurations can be induced by flow even before the
second path is to generate long helices via nucleation and onset of crystallization. Flow orientates polymer chains and
1844 DOI: 10.1021/acs.chemrev.7b00500
Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

increases the number of trans bonds, which in turn enhances the individual isolated long helix is not stable in free polymer melt, as
possibility for formation of all-trans configurations. Combining a the raised end-to-end distance of chain tends to relax back to
homemade shear device and Raman, Chai et al.58 performed a that of coil state due to entropy effects. The growth of long
quantitative study on how shear rate influences all-trans C−C helices suggests the occurrence of coupling between conforma-
bonds. Three peaks in the Raman spectrum of PE are assigned to tional and orientational orderings. This means that flow-induced
intrachain bands, where sharp Raman peaks at 1065 (C−C coil−helix transition provides the necessary stiff segments for
asym) and 1130 (C−C sym) cm−1 stem from consecutive trans isotropic−nematic transition, which further form intermolecular
vibrations, and the broad peak at 1080 cm−1 originates from ordering that in turn stabilizes the long helices. The recent works
gauche bonds in the amorphous component. When PE melt is of Kanaya and Matsuba and co-workers68−71 in iPS melt gave
sheared at 140 °C, the Raman bands at 1065 and 1130 cm−1 indirect evidence to support this mechanism. Stringlike objects
appear and their band intensities increase with increasing shear of iPS with micrometer dimensions were generated by flow
rate, suggesting the generation of all-trans sequences induced by above the melting temperature. Microbeam SAXS and wide-
flow. The all-trans sequence can survive for several hours after angle X-ray scattering (WAXS) characterization demonstrated
cessation of flow, indicating that the flow-induced all-trans that those structures are not stacks of lamellae and have no or
sequence is not only oriented but also stable. The experimental very limited crystallinity.68 The conformational information in
results of Chai et al.58 are consistent with the simulation of those intermediate structures was further examined with two-
Meier,64 who calculated vibrational frequencies and intensities dimensional (2D) FTIR. The IR band of long helix was observed
of various hexadecanes to address how long a trans sequence is in the center and edge of those structures, whereas in the
needed for PE to contribute to the all-trans Raman bands. These amorphous domain only short helical bands were detected. In
results showed that sequences with more than 10 trans bonds are addition, the works of Alfonso and co-workers72,73 suggested
responsible for the all-trans Raman bands at 1065 and 1130 that the intermediate structure consisted of bundles with
cm−1. As only disordered trans sequences exist in PE melt, the conformational ordering and oriented chain segments. Hsiao
appearance of all-trans structure can be expected under strong and co-workers74,75 inferred that the intermediate structure is
nonequilibrium conditions before the onset of crystallization, liquid-crystal phase. It seems reasonable and likely that flow
such as under flow field. In addition, trans and gauche induces the occurrence of isotropic−nematic transition in
conformations may differentiate in the response to external polymer melt.
flow. The work of Pigeon et al.20 demonstrated that the trans For a certain helical length, a threshold concentration of stiff
conformation in amorphous phase is easily oriented along the segment is also needed to induce the isotropic−nematic
flow direction, whereas flow has negligible influence on transition. Combining extensional rheometry and ultrafast
orientating gauche sequences. SAXS/WAXS measurements, Cui et al.30 found a constant
critical strain to observe iPP intermediate structure for a broad
2.2. Flow-Induced Isotropic−Nematic Transition temperature range from 130 to 170 °C. Based on flow-induced
The helix induced by flow can be considered as a rigid rod or stiff coil−helix transition47 and fluctuation−dissipation theories,76,77
segment with length L and diameter b. The increased rigidity the content of helical sequence was calculated as a function of
induced by conformational changes may cause the occurrence of helical length. For helical length exceeding 14 nm, flow not only
nematic ordering, although flexible polymers have no intrinsic facilitates compatible alignment of the helical sequences but also
mesogenic unit. The statistical mechanical theory of isotropic− increases the helical concentration for the isotropic−nematic
nematic transition for rodlike molecules was first proposed by transition. In addition, flow may vary the critical concentration
Onsager66 in colloidal particles and then developed by Flory for isotropic−nematic transition. With time-resolved birefrin-
(see ref 67) for polymers. According to this theory, the excluded gence, Lenstra et al.78 studied isotropic−nematic transition in
volume (Vexcl) of stiff segments increases with increasing helical suspensions of fd virus and found that the critical concentration
length. The excluded volume Vexcl can be calculated as Vexcl = under shear flow is much smaller than in the quiescent case.
2bL2|sin θ|, where θ is the angle between neighboring stiff The formation of conformational ordering discussed in
segments. When the helical sequences exceed a critical length, section 2.1, including coil−helix and gauche−trans transitions,
they start to orient parallel to each other to reduce the excluded can be effectively influenced by external stimuli such as strain,
volume or free energy of the system. The critical length of stiff strain rate, and temperature. Some qualitative conclusions can
segments for orientation occurrence is given as L = 4.19M0/ be drawn: (i) flow promotes segmental ordering not only in
bl0ρNA, where ρ is density, NA is Avogadro’s number, and l0 and sequence length but also in content; (ii) the combination of high
M0 are the length and molecular weight of monomer, flow intensity and high temperature has a similar effect to the
respectively. With this correlation, the critical helical length combination of low flow intensity and low temperature; (iii)
for parallel ordering of iPP calculated by Zhu et al.49 is 11 stronger flow is needed for the formation of ordering segments
monomers, which agrees well with their FTIR experimental with long sequence length than for short ones. However, the
observations. It was indicated that the segmental conformation quantitative relationship between flow intensity, length, and
is stable when the length of helical sequences is shorter than 10 content of ordering segments is still lacking and requires more
monomers, while the helix conformation extends quickly and investigation. The rigidity of ordering segments leads to
starts crystallization as soon as the helical length exceeds 12 intermolecular ordering, that is, isotropic−nematic ordering,
monomers. which is commonly considered as an intermediate state and will
Flow may induce isotropic−nematic transition in flexible be discussed in section 3.
polymers even above the melting temperature. From the rheo-
FTIR experiments of iPP conducted by Geng et al.,53 it was 3. PRECURSORS INDUCED BY FLOW
found that flow-induced helices with 12 monomers can exist at Although segmental conformations are the basic ordering for
170 °C and the intensity of the corresponding absorption band ultimate crystallites, polymer crystallization often starts with the
increases after flow. According to coil−helix transition theory, an formation of nuclei, of which the length scale is larger than the
1845 DOI: 10.1021/acs.chemrev.7b00500
Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

segmental conformation. According to the classical nucleation and thermal histories were imposed on the polymers, which
theory (CNT),79 crystalline nuclei are directly generated from influence both nucleation and growth steps. Despite all
initial random coil and no further structural intermediate is aforementioned limitations, some features of FIP are still
considered. However, numerous experimental observations captured: the formation of FIP can strongly influence the
suggest that FIN may be a multistep process and may undergo crystallization rate, change the crystallite morphology, and even
an intermediate ordered state (also commonly termed as induce the appearance of new crystal modification.
precursor, preordering, mesomorphic ordering, and so The understanding of FIP was further improved by progress
on).11,68,80,81 In our opinion, precursors are the structures in experimental methods. Recently, Lamberti12 produced a
whose existence can be solidly demonstrated by their substantial review paper on methods to investigate and quantify
influences on crystallization (i.e., kinetics, orientation, morphol- crystallization. Modeling approaches and main factors acting
ogy, and so on) but whose fundamental ordering nature is still on crystallization are also included in this review. Here, we just
unclear yet. give a brief introduction to the progress in experimental
Structural intermediates may exist even in quiescent methods. In 1993, Liedauer et al.90 designed a new flow
crystallization. A liquid−liquid phase separation, known as apparatus and proposed a short-term shear protocol, in which
spinodal decomposition, was observed in the first step of the flow duration is short enough to separate nucleation from
polymer crystallization by Kaji et al.,31 and then the relevant crystallization processes. This method allows one to study FIP
theoretical effort was made by Olmsted et al.35 Strobl33 by looking at the subsequent crystallization kinetics and
conjectured that crystallization starts with formation of a morphology. A similar flow apparatus was developed by
mesomorphic layer that then develops into a granular crystal Kumaraswamy et al.,91 which allows one to study structural
layer, followed by a merging process of crystal blocks. Flow is an evolution in real time by optical and X-ray probes with only a
effective external stimulus to reveal the intermediate state before small amount of sample (several grams). At the same time, the
crystallization, if there is one, because the formation of structural commercial Linkam shear cell with parallel-plate geometry was
intermediates with mesomorphic ordering in flexible-chain widely used by the community due to its convenience for X-ray
polymers is often attributed to the rigidity based on conforma- and optical measurements.11 In 2011, Liu et al.92 developed an
tional ordering, which could be considerably enhanced by flow extensional-type flow device, which can control strain and strain
as discussed in section 2. rate independently. More importantly, stress recording and real-
The theory developed by Kim and Pincus82 and experiment time X-ray measurement can be realized simultaneously by this
conducted by Li and de Jeu83 showed that the coupling between device. Meanwhile, various structure characterization techni-
coil−helix transition and orientational ordering could induce a ques such as optical microscopy,93 light scattering,94 birefrin-
nematic state. Recent results from a combination of FTIR, gence, 95 atomic force microscopy (AFM),96 and X-ray
SAXS, and WAXS suggested that coil−helix transition and scattering97 have been combined to reveal the formation and
orientational ordering indeed happen before crystallization in evolution of FIP. Especially with the development of
cross-linked iPP.54 For long-chain polymers, the highly synchrotron ultrafast X-ray scatting30,98 and X-ray micro-
asymmetric nature provides enormous possibility to form diffraction scanning,68,99 new understanding of FIP has been
precursors with orientational ordering. In this section, the obtained.
observations of structure intermediatesthat is, precursors 3.2. Theoretical Considerations for Flow-Induced
will be discussed for FIN to shed light on the detailed process Precursors
and formation mechanism of stable nuclei.
Flow can induce intrachain conformational ordering, but the
3.1. Flow-Induced Precursors conformational ordering alone cannot cause a phase separation
Katayama et al.84 probably were the first to report the formation in polymer melt. However, the conformational ordering may
of flow-induced precursors (FIP) in the 1960s. They studied a couple with density. The aggregation and parallel arrangement
melt spinning process by combining a special model-spinning of oriented segments with conformational ordering may work
apparatus with an in-house X-ray setup. Their results showed together to induce phase separation. As mentioned earlier, Kaji
that the structural signal in SAXS is observed much earlier than et al.31 proposed a spinodal (SD) type of phase-separation
the crystalline signal in WAXS, which indicates that density model for the induction period of polymer crystallization. In this
fluctuations on the length scale of several tens of nanometers model, the amorphous random coil first transforms into a stiff
occur before crystallization. This mesomorphic state was helical conformation. As soon as both length and concentration
deemed as FIP, bridging the transformation from melt to exceed their thresholds, the stiff segments tend to orient with
crystals. Several other groups also provided results to support each other. The increase in chain stiffness induces the parallel
the presence of FIP in PET.85−88 Bonart85 reported that nematic ordering of polymer segments due to excluded volume effect.
and smectic phases form sequentially during stretching of an According to isotropic−nematic theory67 mentioned in section
amorphous PET sample. Using synchrotron characterization, 2, the completely parallel orientation gives an excluded volume
Blundell and co-workers86−88 announced the existence of of zero, while the perpendicular orientation gives the maximum
mesophase during drawing of PET homopolymer and value. Such parallel orientation does not occur homogeneously
copolymer, which act as a precursors in strain-induced in the system, but it involves a microphase separation into
crystallization. At that time, the investigations of FIP were oriented and unoriented domains. Those oriented dense
mainly concentrated on steady melting spinning or polymers domains grow in size by reactions and diffusion of clusters.
with slow crystallization kinetics such as PET,85,86 poly(ethylene Olmsted et al.35 developed a more general phenomenological
naphthalate) (PEN),87,88 and polystyrene (PS),24,89 due to the theory to illustrate the coupling between chain conformation
limitation of time resolution of detection techniques. Besides, and density for phase separation. Figure 5 shows the
the experiments were mostly performed in complex processing temperature-density diagram calculated by use of phenomeno-
conditions like molding or spinning. This means complex flow logical free energy, which contains three regions: coexistence
1846 DOI: 10.1021/acs.chemrev.7b00500
Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

of diffraction maximum before crystal growth. The SAXS


evolution can be well fitted with the theory for spinodal
decomposition, indicating that the scattering behavior obeys the
predication of spinodal kinetics in the early stage of
crystallization.
Spinodal phase separation was mainly observed in melt
crystallization under flow. This is because flow aligns polymer
chains and enhances density, which thereby increases the
possibility of liquid−liquid separation. Recently, crystallization
studies on the extrusion process of iPP, PS, and PET were
conducted by Ryan et al.,106−108 where large-scale ordering
before growth of crystals was observed. Moreover, melt spinning
investigation of poly(vinylidene fluoride) conducted by Cakmak
Figure 5. Generic phase diagram for a polymer melt. Tm, Tc, and Ts are et al.109 and annealing study of iPP from a highly extended melt
the melting, liquid−liquid binodal and spinodal temperatures, conducted by Schultz et al.110 also showed spinodal-type phase
respectively. Dotted line represents quench path. Reprinted with separation.
permission from ref 35. Copyright 1998 American Physical Society.
3.3. Nature of Flow-Induced Precursors
region of melt and crystals, metastable region, and unstable Although the formation of FIP has been demonstrated by a large
region. In Figure 5, the horizontal axis indicates the product of amount of experimental observations, the nature of FIP is still
normalized density ρ and w, which is defined as a product of the under debate. In this section, we review the debates on the
average mass density ρ of melt and the specific volume w of existence (section 3.3.1), structure (section 3.3.2), and ordering
monomer core. This phase diagram is analogous to the binary parameter (section 3.3.3) of FIP. The concept of FIP is rather
component case but has a different ordering parameter, that is, vague and is commonly used to describe structures that are not
density. While the melt is quenched to a temperature below clearly revealed. The term FIP can include various structures; for
spinodal temperature Ts (that is, the melt is in an unstable example, Ma et al.111 sorted four types of FIP according to the
region), SD-type phase separation takes place. The melt
sensitivities of different detection techniques. Up to now, there
separates into two coexisting liquids, which differ in con-
has been no consensus yet reached on the existence and
formation distributions (see Figure 6). In the region of denser
structure of FIP. Various experimental techniques have been
employed to characterize dynamic and thermal abilities of FIP.
Since scattering methods are powerful for probing structure with
long-range ordering, most experimental results presented in this
section are based on X-ray scattering.
3.3.1. Existence of Flow-Induced Precursors. Many FIC
studies have shown the appearance of metastable FIP before
crystallization, among which most experiments were based on
the combination of SAXS and WAXS methods. SAXS can detect
density fluctuations without long-range order, and WAXS can
probe crystalline order. For crystallization with nucleation and
growth mechanism, WAXS should appear at the same time as
SAXS. However, if FIP forms, development of SAXS due to
density fluctuation may occur prior to that of WAXS due to
crystalline order. Most early evidence for the appearance of FIP
was obtained from extrusion processes. The pioneering works of
Katayama et al.,84 Cakmak et al.,109 Ryan et al.,107,108 and
Figure 6. Schematic representation of the late stage of spinodal Schultz et al.110 showed earlier observation of SAXS than WAXS
decomposition for coexisting liquid phases with different conforma- in various systems like iPP, PE, and PET. They concluded that
tions. Thin and thick lines represent disordered and helical density fluctuation with large-scale order occurs and that it plays
conformations, respectively. Reprinted with permission from ref 35. an important role in crystallization. However, Wang et al.112,113
Copyright 1998 American Physical Society. pointed out that the time lag between SAXS and WAXS may be
attributed to detector sensitivity. In other words, WAXS may not
liquid, density and conformation are both closer to the state be able to detect crystals with low volume fraction in the early
needed for crystal packing than that in the original melt, which stage of crystallization, whereas SAXS is capable of probing the
lowers the energy barrier for crystallization.100 structure as long as its density contrast with the surrounding
In fact, numerous experimental results reported the melt is sufficient. Thus, caution should be exercised when
emergence of SAXS signal before crystal diffraction in WAXS, comparing the first appearance of SAXS and WAXS. Heeley et
indicating the occurrence of density fluctuation prior to al.114 further addressed this issue by using an updated WAXS
crystallization.31,101−103 For example, Imai et al.101,104,105 detector, which has an improvement in count rate with a factor
investigated the induction period of PET in cold crystallization. of 104 over that used in previous experiments. Their results
In their experiments, the peak position in SAXS is smaller than demonstrated that the lag between SAXS and WAXS is not
the ordinary peak position of the intercrystal correlation peak. caused by detector sensitivity, and the appearance of SAXS
SAXS signal shows an increase of scattering intensity and a shift before WAXS indicates the formation of unknown structure.
1847 DOI: 10.1021/acs.chemrev.7b00500
Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

In addition to X-ray scattering, some other experimental Although their experimental approach is simple, with that
techniques also lead to the same conclusion on the existence of method, the formation of FIP can be corroborated and the
FIP. As already discussed in section 2, the rheo-Raman study by lifetime can be quantitatively assessed.
Chai et al.58 showed that flow can promote the formation of all- 3.3.2. Are Flow-Induced Precursors Crystalline or Not?
trans bands, which can survive for several hours after cessation of Concerning FIP, another debated question is whether they are
flow even above the melting temperature. The birefringence crystalline or not. In the early stage of nucleation, FIP must start
experiments reported by Seki et al.95 and Ma et al.111 suggested with a tiny amount. In addition, FIP is often studied at high
the appearance of oriented structure before crystallization. temperatures close to the nominal melting point. Those two
Small-angle light scattering (SALS) experiments conducted by reasons lead to difficulty in justifying the structure of FIP. With
Pogodina et al.115 and Zhang et al.116 indicated the formation of the development of synchrotron radiation and detection
highly elongated structures associated with crystallization by techniques, noncrystalline rodlike structures from SAXS,
flow. appearing earlier than crystalline scattering from WAXS around
The elegant works of Alfonso and co-workers72,117−119 about the nominal melting point, have been observed by several
the existence and lifetime of FIP deserve a separate discussion. groups.59,120−122 In contrast, some recent work on PE blends
They imposed shear flow by pulling a fiber inside a thin layer of with low and high molecular weight components and lightly
molten polymer and chose the crystalline morphology at the cross-linked PE showed that FIP contains crystals at temper-
fiber−melt interface as a marker to study the existence and ature above the equilibrium melting point.59,120,123 However,
memory of flow history. Melt was first sheared at a temperature even with a synchrotron radiation source, the detection
slightly above the melting point, followed by an annealing at this sensitivity of SAXS and WAXS is still a disputed issue, which
temperature for different durations, and finally cooled to a lower poses an obstacle to understanding whether FIP are crystalline
temperature for isothermal crystallization. For the sheared melt or not. Therefore, experimental techniques with higher
after being annealed for 100 s, a cylindrical morphology was detection sensitivity and delicate experimental design are
found during isothermal crystallization, while the completely required.
relaxed melt after being annealed for 300 s led to the typical Microdiffraction scanning with beam size of a few micro-
spherulitic morphology (Figure 7). The lifetime of FIP could be meters is a highly sensitive technique, which allows one to obtain
high-quality SAXS and WAXS patterns with high time and
spatial resolution. For FIP, some notable studies have been
carried out recently with this technique by Gutierrez et al.,117
Kanaya et al.,68 and Su et al.,99 just to name a few. Gutierrez et
al.117 observed that ordered FIP can be generated under intense
flow at temperatures above the nominal melting temperature in
iPS. They claimed that those ordered FIP are noncrystalline and
are made up by parallel chain bundles oriented along the flow
direction, while the lamellae developed subsequently with a
crystallinity less than 1%. Their results also suggested that the
smallest and least stable FIP may be destroyed by the relaxation
of orientation and loss of segments from the oriented clusters,
which is consistent with the finding of Balzano et al.120 that FIP
stability depends on size. Kanaya et al.68 reported that FIP has a
crystallinity of about 0.15% and has a higher melting
temperature than the nominal melting temperature of lamellar
crystals in iPS.
Su et al.99 performed systematic temperature-range research
to provide more evidence on the structure of FIP. Their
experimental protocol is similar to the fiber pulling experiment
of Gutierrez et al.117 A fiber was pulled through iPP films to
create an intense shear flow field in a temperature range covering
temperatures below and above the nominal melting point, and
Figure 7. Morphological evolution of PS sample as a function of then the sheared sample was kept at the shear temperature for 10
isothermal time at shear temperature after cessation of flow. The min, which was finally cooled to 138 °C for isothermal
isothermal times were (a) 100, (b) 300, and (c) 500 s (pulling rate is 5
crystallization; see Figure 8. For shear temperatures below 170
mm/s and pulling time is 1 s). Reprinted with permission from ref 118.
Copyright 2008 American Chemical Society. °C (slightly above the nominal melting point of iPP, 165 °C),
crystalline FIP is observed. For shear temperature at 175 °C, no
crystalline FIP is detected but transcrystalline morphology is
estimated from the disappearance of transcrystalline morphol- generated at the fiber−melt surface at 138 °C. The transcrystal-
ogy. They investigated a series of isotactic poly(1-butene)s with line morphology indicates the formation of noncrystalline FIP
varying molecular weights and found that the temperature induced by shear at 175 °C. For shear temperatures above 180
dependence of FIP lifetime follows an Arrhenius-type relation- °C, only spherulitic morphology appears at 138 °C. Some may
ship.72 This relationship is independent of the molecular weight argue that the formation of tiny crystals cannot be detected due
of polymer, but the lifetime of FIP increases with increasing to the detection limit of WAXS. This indeed cannot be
molecular weight. Later, they used the same method to study a completely avoided, even with synchrotron radiation scanning
series of very narrowly dispersed iPS and found that the microdiffraction. If the structures induced by flow at 175 °C are
exponent of this Arrhenius-type relationship is about 2.118 tiny crystallites below the detection limit, the absence of
1848 DOI: 10.1021/acs.chemrev.7b00500
Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

Figure 8. (a1−d1) Microscopic images after crystallization at 138 °C for about 10 min. (a2−d2) One-dimensional (1D) WAXS diffraction profiles
integrated from mapping diffraction patterns. (a3−d3) Distribution of crystallinity over time around the fiber surface (which is indicated in the
microscopic images) during the second mapping measurement of samples (a) S160−138, (b) S170−138, (c) S175−138, and (d) S180−138. Dashed line indicates
the position of fiber. S160−138, S165−138, S170−138, S175−138 and S180−138 represent samples sheared at 160, 165, 170, 175, and 180 °C, respectively.
Reprinted with permission from ref 99. Copyright 2014 American Chemical Society.

induction time is expected, because crystals formed could grow 148 °C, while Somani et al.125,126 provided evidence of
straightforwardly when sample was cooled to 138 °C. However, noncrystalline FIP at 165 and 175 °C. In a nice review by
an induction time exists for the sample sheared at 175 °C Hsiao and co-workers,11 it was pointed out that FIP may be
(S175−138), as shown in Figure 9. Thus, it seems that the existence amorphous, mesophase, or crystalline. In the work of Balzano et
of induction time demonstrates the formation of noncrystalline al.,127 kinetic analysis of early-stage crystallization indicates that
FIP, which further develop into stable nuclei during the unidirectional propagation of a growth front may occur, which
induction time. facilitates development of crystalline FIP. However, even at low
temperature (much lower than the melting point), recent
molecular simulation by Yamamoto55 showed the growth of
smectic mesophase before crystallization in a highly stretched
amorphous iPP. Therefore, probably noncrystalline FIP can be
generated at low temperature, but the transformation process
from noncrystalline FIP to crystallites is too fast to be detected
by current experimental techniques. If this is the case, further
development of ultrafast structural detection techniques should
shed light on this long-standing problem.
3.3.3. Ordering Parameter of Flow-Induced Precur-
sors. Experimental data from X-ray scattering,83 birefrin-
gence,128 and simulation results55 all suggest the existence of
FIP. What is the fundamental ordering of FIP? This question is
still unrevealed for the polymer community. Therefore, the
definition of FIP is obscure and has been queried, just due to the
Figure 9. Evolution of crystallinity during isothermal crystallization at lack of full structural information and precise ordering
138 °C of samples after shear at different temperatures. Reprinted with
permission from ref 99. Copyright 2014 American Chemical Society.
parameter.
Generally speaking, the order of polymer includes intrachain
and interchain orderings.28,29 Imposing a flow field first leads
As suggested by the results of Su et al.,99 whether FIP is polymer chains to be oriented or/and stretched, which is
crystalline or not depends on temperature. This dependence is accompanied by intrachain ordering. With Fourier transform
consistent with the findings from different groups.121,124−126 infrared (FTIR) spectroscopy, it had been demonstrated by Li
Polec et al.121 observed crystalline FIP for temperature below and co-workers50,51,53 that flow can indeed induce conforma-
155 °C, whereas noncrystalline FIP was found for temperatures tional order (helix) in iPP melt, where the effect is more
above 155 °C. Zhao et al.124 reported crystalline FIP of iPP at pronounced below the nominal melting temperature than
1849 DOI: 10.1021/acs.chemrev.7b00500
Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

above. Hence, intrachain ordering seems the starting point to FIP in some way. Of course, relaxation time is not enough to
study FIP, which, however, is commonly neglected. define the structure of FIP, but the work of Li and co-workers
The effects of strain and strain rate on the coil−helix transition presented a good starting point. A more precise ordering
were studied at various temperatures. It was noticed that long parameter is needed to describe and define the structure of FIP.
helices induced by flow can relax or grow and align to induce Interestingly, the crystallization begins after a relaxation
interchain ordering. Figure 10a shows an IR spectrum of the process of flow-induced helices rather than directly after
cessation of flow (Figure 10a), indicating that an arrangement
of helices is required for crystallization. This implies that the
helical aggregations induced by flow have a mesomorphic or
liquid-crystal ordering, consistent with the viewpoint of Somani
et al.74 that FIP are mesomorphic bundles of aligned chain
segments.
3.4. Influences of Precursor on Crystallization
Compared with disordered melt, precursors are easier to
transform into stable crystalline nuclei and consequently trigger
crystallization. This acceleration of transformation under flow is
due to two reasons. First, the barrier between crystal and
precursor is smaller than that between crystal and melt. Second,
the conformation and orientation of chain in precursor are closer
to the state in crystal than that of random-coil state. The former
is thermodynamically favorable while the latter is kinetically
favorable for this transformation, leading to accelerated
crystallization kinetics. In addition, with advanced character-
ization techniques, some other interesting phenomena were
observed recently about how the presence of precursors
influences the crystallization process, such as nucleation (section
3.4.1), crystallization kinetics (section 3.4.2), and final
morphologies (section 3.4.3).
3.4.1. Nucleation. The influence of precursor on nucleation
not only refers to its direct transformation into stable nuclei but
also facilitates the subsequent appearance of ultimate nuclei by
absorbing the surrounding molecules, creating new surface, and
even directly catalyzing the growth from relaxed melt.
Figure 10. (a) Intensities of 998 cm−1 band vs time after shear at (a) Janeschitz-Kriegl and co-workers130−132 proposed the presence
164 °C; (b) 172, 177, and 200 °C; and (c) 168 °C (shear rate 13.5 s−1, of dormant nuclei even in quiescent melt, which are local
shear strain 3375%). Reprinted with permission from ref 51. Copyright alignments or organized aggregates with a shape of fringe
2008 American Chemical Society. micelle. When flow strength (stress in their work) exceeds a
critical level, the dormant nuclei become active as point nuclei
helix containing 10 monomers, with conformational band (998 and align along the flow direction. With increasing stress, the
cm−1), before and after shearing at 164 °C. First, flow leads to a number and longitudinal dimension of those nuclei are both
sharp increase of IR band intensity, indicating a strong significantly increased. At this stage, those nuclei have the ability
enhancement of long helices. Second, the IR intensity shows a to induce growth of lamellae and can be considered as the
decay process after shear, which can be considered as the precursors for shish nuclei. As the intensity of the flow field
melting of helical structures. Finally, crystallization starts and further increases, adjacent nuclei merge into shish nuclei.
leads to a second increase in intensity of the IR band. The Later in 2002, Seki et al.95 put forward a new model within the
relaxation process of helices was analyzed with the model that framework of Janeschitz-Kriegl and co-workers130−132 that
was widely used for protein.129 Two different relaxation emphasized the role of chain absorbing in the formation of
processes were identified, with the nominal melting point shish nuclei. In this new model (Figure 11), flow promotes the
(165 °C) as a boundary temperature. For temperatures above formation of precursors, which interact with the adjacent chains
the nominal melting point, the IR band intensity follows a first- on the surfaces. The absorbed chains are easier to deform during
order decay process (Figure 10b), while for temperatures flow and generate more new precursors, which further adsorb
around or below the nominal melting point, second-order decay adjacent chains and ultimately lead to a trace of precursor
fitting is needed to model the relaxation process (Figure 10c). clusters, that is, shish nuclei. More recently, a ghost nucleation
Taking 168 °C as an example, two relaxation times of 315 and model has been proposed by Cui et al.97 It was speculated that
16 428 s are obtained, suggesting the formation of two different the formed precursors have movement relative to their
states of helices. It was tentatively speculated that the isolated surrounding matrix, which creates surface to form nucleation
helices have a fast relaxation dynamic and aggregations of helices sites. Such an effect can be generated by pulling fiber embedded
with interchain ordering have a slow relaxation process. The in polymer melt.133
relaxation time of helices is influenced by strain, strain rate, and 3.4.2. Crystallization Kinetics. In most cases, FIP develops
temperature, which can reflect the ordered degree of helical into an effective nucleus and accelerates crystallization kinetics,
aggregation. Thus, Li and co-workers50,51,53 proposed that which has been demonstrated by many experiments. However,
relaxation time can be considered as an ordering parameter of FIP may be destroyed and consequently influence the early stage
1850 DOI: 10.1021/acs.chemrev.7b00500
Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

nonmonotonic change in the onset time of crystallization was


observed. For strain smaller than 1.5 or larger than 3.0, only
amorphous scattering is observed immediately after cessation of
flow, while with intermediate strain of 2.0 and 2.5, an obvious
lamellar scattering peak appears. On the other hand, with strain
of 3.5, the lamellar crystals have a relatively weak orientation in
the initial stage of crystallization but a higher orientation in the
late stage (Figure 12b), whereas an opposite evolution trend
appears for the sample with smaller strain of 1.5. Cui et al.134
attributed this unusual phenomenon to the destruction of
formed precursors by further increasing flow intensity. In the
fragmenting process, the precursors may be forced to fracture,
rotate, and tilt, leading to an increased onset time of
crystallization and low orientation of initial lamellar crystals
under large strains.
In addition to the onset of crystallization and lamellar
orientation, the growth process is also affected by the presence of
precursors. Hsiao et al.59 revealed an unexpected multiple shish
Figure 11. Schematic diagram of the nature of shear-induced structure in a sheared PE blend containing 2 wt % crystalline
nucleation and subsequent growth of oriented lamellae during short-
term shearing. (a) A long chain (bold line) dispersed in a matrix of short
ultrahigh molecular weight (UHMW) component and 98 wt %
chains adsorbs to a pointlike precursor. Dangling segments of adsorbed noncrystalline matrix. From the high-resolution scanning
chains become oriented due to shear. (b) Additional chains adsorb and electron microscopic (SEM) image shown in Figure 13a,
their dangling segments form streamers upstream and downstream of multiple shish with a diameter of several nanometers, rather than
the pointlike precursor. (c) Increased local orientation of the chain single shish, are presented, different from the conventional view
segments increases the probability that long-lived ordered structures of shish structure. With the presence of multiple shish, a time-
will form. (d) More chains adsorb to these new nucleation sites, and the dependent growth rate of kebab is observed (Figure 13b), which
process propagates a string of nuclei along the flow line. (e) The nuclei is demonstrated as diffusion-controlled growth. This result is
along this thread lead to lateral lamellar growth. Reprinted with
permission from ref 95. Copyright 2002 American Chemical Society.
quite different from crystallization under quiescent conditions,
where diffusion is not a limiting step and causes a constant
of crystallization in a different way. Working with extensional growth rate. The results of Hsiao et al.59 are consistent with
rheology and in situ SAXS techniques, Cui et al.134 investigated AFM experiments conducted by Hobbs et al.,96,135,136 who
the FIC of bimodal PEO blends. As shown in Figure 12a, a observed a similar diffusion-controlled growth process for
kebab.
More recently, Roozemond et al.137 proposed a self-regulation
mechanism in flow-induced structure formation of iPP.
Combining a slit flow device and in situ ultrafast WAXS
measurement (30 frame/s), they found that the kinetics of the
crystallization process in all shear layers are identical, regardless
of flow rate or flow time. The intensity of flow field only
influences the thickness of shear layer. They attributed this
abnormal phenomenon to the disturbance of local flow field by
the formed shish. Flow first promotes the formation of pointlike
Figure 12. (a) 1D SAXS meridional integrations I(q) immediately after precursors, which transform into shish in the same way as in the
cessation of extension (t = 0 s). (b) Orientation parameter of lamellar model of Janeschitz-Kriegl and co-workers130−132 and Seki et
crystals of PEO blends evolving with time under different strains at 52 al.95 When the density of shish reaches a critical value, the
°C. Reprinted with permission from ref 134. Copyright 2014 American
modulus of the shear layer is sufficient to hinder further
Chemical Society.
deformation and stop shish growth. Since the volume flow rate is

Figure 13. (a) SEM image of toluene-extracted UHMW PE crystallites with a shish-kebab structure having multiple shish. (b) Kebab growth rate G as a
function of time. Reprinted with permission from ref 59. Copyright 2005 American Physical Society.

1851 DOI: 10.1021/acs.chemrev.7b00500


Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

Figure 14. Schematic of correlation between precursors and cylindrites at a local wall shear stress. (a) One-dimensional SAXS plots of structural melt
after annealing for 10 min at 180 °C (a-1; iPP was heated directly from room temperature to 180 °C) and supercooled melt at 180 °C (a-2; iPP was first
heated to 210 °C, kept at this temperature for 15 min to ensure thermal history, and finally cooled to 180 °C). (b) Schematic of shear stress distribution
in a slit shear cell. (c) Corresponding optical images obtained under crossed polarizers at wall shear stress of 20 kPa. The bright sections in optical
images represent β-form crystals of iPP. Reprinted with permission from ref 142. Copyright 2012 American Chemical Society.

constant, the material in the inner layer suffers the raised flow scattering in WAXS (Figure 14a). They found that, even with
rate, which facilitates the appearance of shish and formation of low shear stress, the presence of precursor assists the formation
shear layer. of β-cylindrites (Figure 14c). The density and content of β-
3.4.3. Morphologies. The influence of FIP on crystal- cylindrites increase with increasing precursor content. Their
lization mainly depends on the density and orientation of further work142 indicated that, in a specific structural melt
formed precursors. Mild flow gives rise to chain orientation and containing precursors, the number of α-cylindrites increases
promotes the formation of pointlike FIP, leading to a huge with shear stress at the expense of β-cylindrites. Moreover,
number of spherulites that are small in size, while strong flow lamellar structure also depends on the type of precursor. Four
results in chain stretch and favors the formation of shish different types of precursornamely, pointlike, scaffold-net-
precursors, resulting in shish-kebab structure. Some intermedi- work, microshish, and shishwere observed sequentially by
ate structures have also been reported, depending on the density increasing strain in a lightly cross-linked PE.144 The long period
and length of FIP. For example, oriented FIP with short length of lamellar stacks resulted from the former three types of
lead to distorted spherulites,138 while those with long length but precursor increasing in an orderly fashion, while the fourth type
low density give rise to sausagelike structures.139 Under strong of precursor decreased. This is due to the increase of
extensional flow, precursors with low and high densities can lead entanglement and cross-link densities in the former three
to fibrillar structure and to homogeneous network structure types of precursor, leading to a thick interlamellar amorphous
decorated by small highly oriented lamellar stacks, respec- layer. In contrast, the chain conformation inside the fourth type
of precursor is almost extended, allowing lamellae to grow
tively.134
without preservation.
Recently, Shen et al.140 performed an interesting work to
establish the relationship between initial precursors and final 4. FACTORS INFLUENCING THE FORMATION OF
fibrillar structure (the so-called cometlike shish-kebab morphol-
FLOW-INDUCED PRECURSORS
ogy in their definition). A combination of two-step shear
proposal and in situ optical microscopy was adopted in their FIP formation is influenced by many experimental parameters,
experiment. The aim of the first shear step is to generate such as flow strength, temperature, polydispersity of materials,
pointlike precursors. After the second shear step, they observed and so on. Thus, study of the crucial factors determining FIP
that all fibrillar structures are initiated from pointlike precursors. formation is expected to improve our understanding of the
In addition to morphologies, crystal modifications are also nature and formation mechanism of FIP.
influenced by the presence of precursor. Zhang et al.141−143 4.1. Flow Parameters
obtained iPP precursors by melting a sample at a temperature Flow can disturb chain configuration from the initial equilibrium
between the nominal melting point and the equilibrium melting random-coil state to oriented and aligned states and thus can
point. The formation of precursors can be demonstrated by the promote the formation of FIP as well as accelerate the
periodic structure scattering in SAXS but no crystalline crystallization. However, promotion and acceleration happen
1852 DOI: 10.1021/acs.chemrev.7b00500
Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

only when flow is above a certain strength. In this sense, the shish-kebab structure in the final crystallized sample. For stress
quantification of flow strength is essential to understand FIC. below the critical value, flow induces the formation of pointlike
The flow field can be assessed by several parameters such as precursors, as confirmed by the accelerated kinetics and
strain, strain rate, flow time, stress, and external work. The first increased number density of spherulites compared with the
three parameters (strain rate, flow time, and stress) are intrinsic quiescent conditions. These results of Kornfield and co-workers
flow properties, but the last parameter, mechanical work, is are well consistent with that of Balzano et al.,148 who performed
associated with not only flow strength but also molecular shear experiments at temperatures around the nominal melting
properties. point (163 °C for the iPP they used). Coincidentally, stable
Recently, Peters and co-workers98,111,145 carried out some threadlike FIP containing limited crystals form for stress larger
elegant works to study the effects of strain rate on the formation than 0.12 MPa, while short-length FIP with aspect ratios tending
of FIP even during short-term flow. By combining a Linkam to unity (pointlike FIP) are generated for stress smaller than
shear cell and fast X-ray scattering, they investigated the early 0.10 MPa and dissolve into the melt with time.
stage of FIC from melt of iPP at 145 °C with a total strain ε = 180 In contrast with the parameters describing flow field, such as
but different strain rate ε̇ under strong flow.145 For ε̇ ≥ 90 s−1, strain rate, strain, flow time, and stress, Janeschitz-Kriegl et al.149
crystalline FIP containing 2% crystallinity were already proposed a concept of specific mechanical work, which is
generated during flow, while for ε̇ ≤ 60 s−1, only metastable defined as W = ∫ t0flowη[ε̇(t)]ε̇ 2(t) dt, where η is the strain-rate-
FIP with no crystallinity were formed during flow, which dependent viscosity. The specific mechanical work contains the
transform into crystals when flow stops. Later, combining the effects of stress and strain, which may be chosen as a decisive
characterizations of fast X-ray scattering, birefringence, and final parameter to characterize the effect of flow on FIP formation.
morphology, Ma et al.111 defined several types of FIP for iPP at Janeschitz-Kriegl et al. found that the number of pointlike
145 °C. For ε̇ ≥ 400 s−1, crystalline FIP can be generated during precursor increases with the specific mechanical work applied,
flow, even when the flow duration is only 0.25 s. For ε̇ = 320 and but a sudden transformation from pointlike to threadlike FIP
240 s−1, noncrystalline FIP formed during flow and quickly happens when the specific mechanical work reaches a threshold.
transformed into crystalline shish after cessation of flow, which Recently, the validity of specific mechanical work was
are defined as precursor for shish. For ε̇ = 160 s−1, SAXS and demonstrated by Mykhaylyk et al.150 by using a well-defined
WAXS do not show characteristic structural signals, but the model of linear−linear hydrogenated polybutadiene (h-PBD).
birefringence signals, including a birefringence upturn during They found that, for an assigned sample, the specific mechanical
flow and nonzero relaxation after flow, were observed. The works for formation of threadlike FIP are almost the same and
birefringence signals results demonstrate the formation of FIP, independent of strain rate when the strain rate is larger than the
which is termed as threadlike precursor. For ε̇ = 80 s−1, the inverse Rouse time of long-chain molecules (Figure 15).
birefringence cannot capture a structural signal, but the However, this specific work decreases with increasing fraction
accelerated kinetics and orientation indirectly reveal the of long-chain molecules in blends. Later, Mykhaylyk et
generation of FIP, which is termed as needlelike precursor. al.138,151,152 generalized this conclusion to polydisperse
In addition to strain rate, strain also plays an important role in materials and pointed out that the specific mechanical work
FIP formation. For example, Liu et al.144 studied strain-induced for threadlike FIP formation depends on the chemical structure
crystallization in lightly cross-linked PE and observed and molecular weight distribution of polymer as well as
uncorrelated oriented FIP when the applied strain is smaller experimental temperature.
than 0.6; scaffold-network FIP when the applied strain is on the
4.2. Temperature
order of 1.0; microshish-type FIP when the applied strain is up
to 1.3; and shish-type FIP when the applied strain is larger than The effect of temperature is obvious in FIP formation, as in
2.0. nucleation and crystal growth. In this section, we mainly discuss
Although the strain rate and strain provide useful ways to some unusual phenomena about the influence of temperature on
assess flow strength, the flow time, that is, accumulation effect precursors with the presence of flow.
for the whole flow period, is not fully considered. Many results As already discussed in section 3.3, whether FIP are crystalline
have shown that the strain rate ε̇ and flow time tflow together or noncrystalline depends on temperature. Generally, in
influence FIP formation in a complex way. An empirical polymer melt, for temperatures around and below the nominal
relationship was constructed by Liedauer et al.90 that the melting point, FIP stay in a crystalline state; for temperatures
transition from spherulitic to fibrillar morphology (an indicator slightly above the nominal melting temperature but still well
of fibrillarlike FIP) is proportional to (ε̇)4tflow2. This conclusion below the equilibrium melting temperature, FIP are in a
is qualitatively consistent with the results of Somani et al.146 in noncrystalline state; and for temperatures above the equilibrium
the sense that a flow with large strain rate but short flow time has melting temperature, only amorphous melt is expected.
more remarkable effects on the formation of FIP than a flow with However, for PE at 142 °C, close to but above the equilibrium
small strain rate but long flow time. melting temperature Tm0 (141.2 °C), FIP containing limited
According to the work of Doi and Edwards,67 the anisotropy crystallinity have been observed in blends of LMW (low
in chain conformation can be directly related to the applied molecular weight) and HMW (high molecular weight)
stress. Thus, stress is also a key parameter to influence FIP components by Keum et al.153 and Balzano et al.27,120 Recently,
formation. Kornfield and co-workers95,139,147 carried out some Liu et al.123 reported that, in lightly cross-linked PE, crystalline
well-defined shear experiments with bimodal iPP blends by not FIP was observed under extensional flow at 184 °C, about 43 °C
only fixing the stress but also employing similar flow time and higher than the equilibrium melting temperature.
applied strain, aiming to control the same average chain Balzano et al.27 pointed out that, in polymer melt, precursors
orientation at the molecular level. They found a critical shear could form if the HMW chains overcome a critical molecular
stress of 0.12 MPa at 137 °C, above which threadlike FIP forms deformation at a time scale when those molecules cannot relax.
as evidenced by the anomalous stress upturn during flow and However, at such high temperatures, the crystalline FIP is not
1853 DOI: 10.1021/acs.chemrev.7b00500
Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

Figure 16. Crystallinity and SAXS intensity evolve with time after step
shear at 142 °C. The initial drop of SAXS intensity is fitted with Doi−
Edwards’ memory function; the fit is represented by the solid line.
Reprinted with permission from ref 120. Copyright 2008 American
Physical Society.

between the experimental dissolution process and the modified


memory function further indicated that the time scale for
dissolution of small FIP is related to the reputation time of
HMW chains, in line with the IR results of An et al.51
Intuitively, the threshold of FIP formation is expected to
increase with temperature because of the enhanced barrier. In
contrast, Kumaraswamy et al.154 found that the flow time for FIP
formation under strong shear flow decreases with increasing
temperature (Figure 17). However, after rescaling by the
rheological Williams−Landel−Ferry (WLF) time−temperature
shift factors, the values of the rescaled onset time for
birefringence upturn were approximately the same for various
temperatures ranging from 140 to 175 °C. This result implies

Figure 15. (a−f) Optical images of sheared blends of high and low
molar mass h-PBD (2 wt % 1700 kDa in a 15 kDa matrix). White dashed
circles mark a boundary between oriented and nonoriented lamellar
structures in samples. The plot superimposed on image d shows the
degree of orientation of lamellar structure along the flow direction,
measured across the diameter by SAXS. The color-coded scattering
patterns above were taken from the areas marked by the green and pink
squares. (g) Plot of specific mechanical work vs measured critical shear
rates, required for the formation of threadlike FIP. (Inset) Dependence
of shear rate required for orientation on shearing time. Reprinted with
permission from ref 150. Copyright 2008 American Chemical Society.

stable even though it could form. Balzano et al.120 studied the


structural evolution of FIP at 142 °C after cessation of flow,
using in situ SAXS and WAXS. At times before 600 s, some
interesting phenomena were observed; see Figure 16. After
cessation of flow, the SAXS intensity decreased exponentially,
while the corresponding crystallinity built up gradually. Balzano Figure 17. Temperature dependence of the time scale at quiescent
et al.120 speculated that the initially formed FIP show crystallization, tQ, and of the onset time of the abnormal upturn in
polydispersity in size. The FIP with sizes larger than the critical birefringence during flow, tu. The solid line presents WLF rheological
value transform into crystals, whereas those with sizes smaller shift factor with 190 °C as a reference temperature. Reprinted with
than the critical value dissolve into melt. The good correlation permission from ref 154. Copyright 2002 American Chemical Society.

1854 DOI: 10.1021/acs.chemrev.7b00500


Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

that the formation of FIP happens at a fixed strain. Thereby, they agreement with the results of Kornfield et al.128 However, Seki et
proposed that the formation of FIP is determined by a critical al.95 pointed out that, at low stress, long chains have no priority
level of molecular orientation, consistent with the viewpoint of in accelerating the formation of pointlike precursors. They
Balzano et al.27 More recently, Cui et al.30 presented a similar speculated that the formation of pointlike precursors is
result in iPP under strong extensional flow. In their study, the determined by the local stress or average orientation of chain
time resolution of SAXS/WAXS was down to 30 ms, and rather segments, which is independent of long chains.
short flow duration was adopted (less than 400 ms). By The simulation results of Wang et al.157 showed that
combining SAXS/WAXS and rheological information, they polydispersity, together with orientational relaxation, deter-
found that FIP already formed during flow and they discovered a mines precursor formation. They first studied athermal
constant critical strain for FIP formation for temperatures relaxation in bulk extended chains and then investigated
ranging from 130 to 170 °C. Cui et al.30 proposed a kinetic isothermal crystallization of intermediately relaxed melt by
pathway coupling flow-induced coil−helix transition and dynamic Monte Carlo simulation. Their results suggested a
isotropic−nematic transition, in which flow provides sufficient longer orientational memory of long chains than short ones. For
concentration and alignment of helices for FIP formation, isothermal crystallization, in the melt with uniform chain
exhibiting that the barrier is eliminated by flow. lengths, the crystallization rate can be increased by the
orientational memory, but no precursors formed during
4.3. Molecular Parameters
crystallization for either long- and short-chain systems.
Most semicrystalline polymers have very broad molecular However, in a bimodal blend with different chain lengths,
weight distribution (i.e., polydispersity). Short and long chains precursors were generated from the highly oriented long-chain
differ in their responses to imposed flow, which complicates the component. They argued that orientational relaxation leads to
explanation of FIP formation, especially the role of short and selection of long chains, due to their long orientational memory,
long chains. From the rheological point of view, whether and and makes them the source for precursor formation.
how much chains can be deformed by external flow depends on
the competition between molecular relaxation and flow strength. 5. CORRELATION BETWEEN DEFORMATION AT
This competition can be quantified by the Weissenberg number, WHOLE-CHAIN SCALE AND NUCLEATION
which is a product of strain rate ε̇ and relaxation time τ (Wi = ε̇τ,
which is named as Deborah number in the original reference155). Ordering in FIP is not completely clear yet, but flow-induced
Two characteristic time constants, terminal relaxation time τrep shish can be characterized much more easily due to their
for chain orientation and Rouse relaxation time τRouse for chain sufficiently large density contrast with the surrounding melt. The
stretch, classify flow field in three regions: (i) Wirep < 1, where length of shish can reach micrometers, comparable with that of
flow is weak and has negligible influence on chain conformation; extended chain, so shish have been recognized as extended-chain
(ii) Wirep > 1 and Wis < 1, where only chain orientation occurs crystals in early times. Hence, the formation of shish nuclei
without segmental stretch; and (iii) Wirep > 1 and Wis > 1, where seems correlated with the deformation of polymer at the whole-
flow is strong enough to ensure both chain orientation and chain chain scale. In this section, the formation mechanism of shish
stretch. The two specific relaxation times follow the relationship nuclei is our primary concern. For this purpose, we will first
τrep = τRouseZ, where Z is the number of entanglements per introduce the discovery and various formation mechanisms of
chain,145 meaning that Wirep is always larger than Wis. Also note shish nuclei. Then some new understandings of shish-kebab
revealed by recent experimental investigations will be delivered,
that the practical threshold of Weissenberg number may be
which are essential but still under debate: (i) Is coil−stretch
slightly larger than unity. According to the estimation of van
transition (CST) or stretched network model (SNM)
Meerveld et al.,155 the number density of spherulites can be
responsible for shish formation in polymer melt (section
increased for Wis < 1−10, suggesting an enhancement of
5.6.1)? (ii) How do polydispersity, long chain, and short chain
pointlike FIP in this flow strength region. While shish-kebab
matrix influence shish formation (sections 5.6.2−5.6.4)? (iii)
structure is generated for Wis > 1−10, implying the formation of
How does shish evolve and grow under flow and thermal
threadlike FIP. The validity of this Weissenberg number
conditions (section 5.6.5)? (iv) How does shish form under
approach has been further confirmed by Housmans et al.,156
near-equilibrium and far-from-equilibrium conditions (section
Somani et al.,146 Zhao et al.,124 and many others.
5.6.6)?(v) Is shish formation a single-stage or multistage process
According to this picture, it is obvious that molecular weight
(section 5.6.7)? (vi) Is shish a thermodynamic phase or a kinetic
and polydispersity play key roles in defining the flow regions and
state (section 5.6.8)?
in influencing FIP formation. At a given flow condition, long
chains are much more easily oriented or stretched than short 5.1. Shish-Kebab Crystallite
ones, as long chains have longer relaxation times. To assess the The chain character of polymer molecules was recognized
effect of molecular weight and polydispersity on FIP formation, around 80 years ago. Since then, many studies had been carried
bimodal blends composed of long and short chains in a matrix out on extension and alignment of polymer chains to exploit the
were commonly used as a model system. Kornfield et al.128 intrinsically anisotropic properties of materials.158−171 First of
studied the role of long chains in a bimodal blend of iPP under all, chain stretching in extensional flow benefits fiber spinning,
shear flow. They found that the long chains are preferentially where a uniaxial flow from a spinneret dies and a subsequent
involved in the formation of threadlike precursors. Chains mechanical drawing are utilized to induce chain orientation.
longer than about 4.7 times average are effective in promoting Ziabicki et al.172,173 and Peterlin174 constructed a relationship
threadlike precursor formation. Furthermore, the formation of between extensional flow and processing parameters for
threadlike precursors is greatly enhanced by long chain−long orientation phenomena in melt spinning. Early in 1963,
chain overlap, suggesting a cooperative effect of long chains. Seki Mitsuhashi175 reported the generation of fibrous stringlike
et al.95 reported a similar enhancement of threadlike precursor structures in polyethylene solution upon stirring. In the same
formation by adding long chains in a short-chain matrix, in year, Vanderheijde176 observed a similar crystal morphology in
1855 DOI: 10.1021/acs.chemrev.7b00500
Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

polyoxymethylene crystallized from stirred dilute solution. After flow component of Taylor vortices is responsible for shish-kebab
observing the stringlike structure (see Figure 18), Pennings and formation. Those essential experimental phenomena lead to the
genius idea of Frank182 about chain structure and end-use
properties of polymer products. Frank realized that the unit cell
of PE crystallites is similar to diamond in certain directions and
predicted that the Young’s modulus could be on the order of 285
GPa for the crystal consisting of fully extended chains. However,
in that era, the Young’s modulus was only on the order of 10
GPa, even for oriented PE. The potential of high-modulus PE, as
Frank conjectured, stimulated the polymer community greatly
and research on shish-kebab flourished, as shish have been
considered as extended-chain crystals.183−186
5.2. Hierarchical Structure of Shish-Kebab
Initially, it was thought that shish-kebab consists of parallel
extended long chains as shish or row nuclei, which nucleate the
epitaxial growth of folded-chain crystals as kebab.187,188 The
observations of shish-kebab at variant scales revealed that the
shish-kebab entity consists of two structural levels: micro- and
Figure 18. Electron micrograph of shish-kebab in stirred PE solution. macro-shish-kebab. The macro-shish is constructed by micro-
Reprinted with permission from ref 177. Copyright 1965 Dr. Dietrich shish-kebab, and the macro-kebab grows on the surface of
Steinkopff Verlag. macro-shish. As shown in Figure 19a, Barham and Keller189
schematically showed the difference between macro- and micro-
shish-kebabs. Moreover, Pennings et al.180 gave the details of
Kiel177 demonstrated that it may consist of extended-chain micro-shish-kebab in Figure 19b. In the macro-shish-kebab
crystals in 1965. Moreover, similar stringlike structure was also structure, the macro-kebab can be removed to reveal the central
reported for several other polymers.178−180 In these cases, a micro-shish-kebab, while the micro-kebab is intrinsically
fibrillar core with diameter of about 10 nm was always found, attached to the central micro-shish core. Figure 19c shows the
which acts as a nucleus for epitaxial growth of folded-chain structure of the central core of a micro-shish-kebab. The core
crystals. Such morphology was named shish-kebab. was considered as continuous extended-chain crystals at first,
For shish-kebab structure, Pennings and Kiel177 investigated but later shrinkage measurements and dark-field electron
the melting behavior with polarized optical microscopy (POM). microscopic results suggested that the core may be elongated
It was found that the shish backbone shows strong birefringence bundlelike crystals connected by fringelike amorphous re-
and can bear an unexpectedly high temperature up to 151 °C, gions.190 The chains originating from these regions may
much larger than the melting point of common lamellar crystals. crystallize later to form the platelets. The ends of the crystalline
They conjectured that the shish might be extended-chain region are tapered in order to satisfy the lower density of
crystals. Later, in 1970, Pennings et al.181 studied extensional- amorphous regions.189
flow-induced crystallization in PE solutions with a Couette-type Such a hierarchical structure of shish-kebab has been
apparatus. They found that shish-kebab could be observed only confirmed by monitoring the thermal and soluble behavior
with Taylor vortices, and they speculated that the extensional with different techniques like scanning or transmission electron

Figure 19. Schematic of (a) difference between macro- and micro-shish-kebabs, (b) micro-shish-kebab, and (c) central core of micro-shish-kebab.
Adapted with permission from ref 189. Copyright 1985 Chapman and Hall Ltd.

1856 DOI: 10.1021/acs.chemrev.7b00500


Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

Figure 20. (a) Schematic of flow device consisting of two opposed jets. (b) Birefringent bright line under elongational flow in a PE dilute solution
above crystallization temperature. Adapted with permission from ref 207. Copyright 1993 Dr. Dietrich Steinkopff Verlag.

microscopy (SEM or TEM), providing real-space information, extended chains, where no stable intermediate chain con-
and X-ray or neutron scattering, providing details in reciprocal formation exists.
space.71,72,95,117,122,147,190−203 Keller and co-workers190,200 used The elegant studies of Keller and Kolnaar207 and Mackley208
TEM to study shish-kebab structure before, during, and after the provided direct evidence of CST in polymer dilute solution
melting process in PE. They observed smooth fibers with under extensional flow. They designed a flow cell that consists of
diameters around 10−100 nm and small closely packed two opposed jets to create an elongational flow field, as depicted
plateletsthat is, micro-shish-kebabsand fibers with large in Figure 20a, and used birefringence to characterize chain
platelets with spacing distance about 0.4 μmthat is, macro- orientation. In a dilute solution of monodisperse PE, they found
shish-kebabs. The partial melting experiments in iPS of that a sharp bright line appeared along the central line of the two
Petermann and co-workers201 suggested that the entire shish- jet orifices, corresponding to full chain extension when the
kebab structure is composed of the following parts: extended- imposed flow rate exceeded a critical value (Figure 20b).
chain micro-shish crystals, overgrown micro-kebabs that Mackley and Keller209 later carried out experiments at lower
nucleate and connect with micro-shish intrinsically, and temperatures, where crystallization could happen, and found
macro-kebabs generated by secondary crystallization with that shish-kebab morphology was generated within the localized
micro-kebabs as nuclei. Different parts of shish-kebab structure regions only where birefringence appeared. On the basis of these
show different responses to external attack. It was found that results, they constructed a nice correlation between initial chain
macro-kebab can be washed away by fuming nitric acid from conformation and final morphology, that is, the CST model,
micro-shish-kebab.204 Moreover, micro-kebabs can be over- where extended chain is assigned to form shish and random coil
heated and are more thermally stable than macro-kebabs, as is assigned to form kebab.
micro-kebabs are firmly attached to the micro-shish backbone. Experimentally, the direct visualization of CST in individual
Recently, Kanaya et al.205 investigated blends of low polymers was reported by Smith et al.,210,211 who observed
molecular weight deuterated PE and high molecular weight molecular extension as a function of strain rate through use of
hydrogenated PE under shear flow by combining small-angle fluorescently labeled DNA molecules in an extensional flow.
neutron scattering (SANS), SAXS, and depolarized light Their results gave the transient rather than the final steady-state
scattering (DPLS). They elucidated that a large oriented chain conformations, which indeed showed fully extended chain
structure, on the micrometer scale, contains some extended- conformations and are in line with the predictions of CST. The
chain crystals with a diameter of approximately 10 nm, which simulation work conducted by Dukovski and Muthukumar212
seems in accordance with the picture of macro-shish-kebab. also demonstrated the discontinuous CST of isolated chains
Their work also gave some structural details on both nanometer under extensional flow. Their calculated free-energy landscape
and micrometer scales. showed that chains may be in stretched or coil state depending
5.3. Coil−Stretch Transition in Polymer Solution on the initial conformation under a specified flow field, even for a
monodisperse polymer. The stretched chains aggregate into
To generate extended-chain crystals, polymer chains have to be shish, while the coiled chains crystallize as kebabs subsequently.
stretched. Coil−stretch transition (CST) was used to account
for the formation of shish nuclei under flow. Historically, the 5.4. Shish-Kebab Formation in Polymer Melt
CST theory was first developed by De Gennes206 to study single- In concentrated solutions and melts of polymer, chain overlap
chain dynamics in polymer dilute solution under external flow. and entanglements exist. The physical entanglements between
According to the CST theory, only outer segments are exposed different chains can act as network junctions. With an opposed-
to the flow field when polymer is in random-coil conformation, jets flow device and birefringence, Keller and Kolnaar207
while all segments are subjected to flow when polymer is in demonstrated that, for concentrated solution under elongational
extended-chain conformation. Hence the hydrodynamic inter- flow, there are two critical strain rates ε̇c and ε̇n, corresponding to
actions are decreased by chain stretching and a critical strain rate CST (Figure 21a) and network deformation of stretched
exists, above which a solute polymer random coil will unwind polymer chains (Figure 21b).
abruptly. That is to say, the steady flow generates dual For an entangled polymer solution, it was demonstrated that
populations of chain conformations, random coil and fully the chains become disentangled and stretched out individually
1857 DOI: 10.1021/acs.chemrev.7b00500
Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

with the result of 3−4 nm obtained from X-ray radial


distribution function.214 Then the shish would be axially
punctuated with kebabs upon modest heat treatment. As
mentioned above, the correlation between CST in chain
conformation and the resulting crystal morphology of shish-
kebab has been firmly established for solution, while no
conclusive study on melt was conducted. On the basis of the
fact that fiber-platelet duality has been also widely observed in
melt, Keller et al.1,216 argued that the final morphology of shish-
kebab structure was a pointer to the preexistence of CST, and
thus they extended the application of CST model to polymer
melt, although there is no direct evidence for CST occurrence in
melt. Afterward, the CST model was widely accepted by the
community and recognized as a standard to explain shish-kebab
formation in both solution and melt.
However, it should be emphasized that the CST theory
Figure 21. Schematic of the response to an elongational flow field for proposed by De Gennes was originally deduced from polymer
entangled polymer solution. (a) Chain stretch at strain rate window dilute solution where hydrodynamic effect is an essential
ε̇c−ε̇n and (b) deformation of network of stretched polymer chains for ingredient. For polymer melt, this essential ingredient does
strain rate larger than ε̇n. Reprinted with permission from ref 207. not exist. Thus, in polymer melt the argument for CST actually
Copyright 1993 Dr. Dietrich Steinkopff Verlag. lacks the strict physical fundamentals. Therefore, the above
experiments conducted by Keller and Kolnaar207 can provide
when the strain rate reaches ε̇c. As the strain rate further convincing evidence of CST only for dilute and semidilute
increases to ε̇n, the chains deform like a mechanically cross- solutions. Polymer melt is a transient network constructed by a
linked network and the viscosity of the polymer solution large amount of entanglements, and the chains seem unlikely to
increases drastically. That means, for entangled polymer be extended individually. Experimental evidence on shish
solution, there is a strain rate window between ε̇c and ε̇n composed of both short and long chains147 and the critical
where stretch of individual chain occurs. Both ε̇c and ε̇n decrease strain for shish formation217 clearly revealed that CST is not a
with increasing polymer concentration, while ε̇n decreases much necessary condition to induce shish in polymer melt. Those
more steeply with concentration than ε̇c. The window ε̇c−ε̇n works will be discussed in detail later.
narrows with increasing concentration, but it is not clear
5.5. Other Mechanisms for Shish-Kebab Formation
whether there really exists a crossover at high concentrations,
due to the lack of well-defined materials. If there is no crossover, In addition to the dominant role of CST in explaining shish-
as shown in Figure 22a, the CST holds true even in polymer melt kebab formation, some other models and mechanisms were also
(C = 1); but if a crossover exists, as shown in Figure 22b, chains developed. In 1979, Hoffman218 proposed that the formation of
cannot be extended individually but deform as a network in melt. shish was a multiple nucleation event. In this model, long
Keller and Kolnaar207 found that the oriented crystal molecules are more evident to be elongated by flow, which
morphology observed in melt under flow displays the same provides a set of sites for the formation of bundlelike or fringed-
fiber-platelet duality as that in solution. Similar experimental micellar nuclei. The nuclei are interlinked by short amorphous
results have been widely obtained in polymer melt under ciliary bridges, and the interior of nuclei are partly of extended-
extensional flow.213−215 For example, Schultz and co-workers213 chain type crystals but with some chain-folding defects and chain
studied the structure development of PET during melt spinning ends. Nucleation on the same aligned long chain leads to a set of
in the early stage of crystallization. They observed the diameter nuclei, which subsequently crystallize and transform into a very
of shish directly by TEM, which is about 5 nm, in accordance long core fibril, shish. Then Hoffman219 combined continuum

Figure 22. Strain rate for CST and stretched network rate versus concentration plot in elongational flow. Two cases are shown: (a) CST exists up to
pure melt (C = 1); (b) no CST at high concentrations or pure melt. Adapted with permission from ref 207. Copyright 1993 Dr. Dietrich Steinkopff
Verlag.

1858 DOI: 10.1021/acs.chemrev.7b00500


Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

Figure 23. Engineering stress−strain curves of high-density polyethylene (HDPE) melt with (a) different strains at the same strain rate of 6.3 s−1 and
(b) different strain rates. Numbers at the largest stress indicate yield strain. (c) Extensional yield strain evolves with strain rate. Red dashed line
indicates the critical strain for shish formation. Reprinted with permission from ref 217. Copyright 2010 American Chemical Society.

mechanics and nucleation theory to describe this model become active as point nuclei and are aligned along the flow
theoretically. The theory predicts several parameters, such as direction. With increasing stress, both the number and
shish diameter and mean characteristic length, that agree well longitudinal dimension of those nuclei are significantly enlarged.
with experimental results. At this stage, those nuclei have the ability to induce the growth of
In 1984, Smook and Pennings220 emphasized that elastic flow lamellae and can be considered as precursors for shish. With
instabilities play a key role in shish-kebab formation during gel further increase in the intensity of flow field, the adjacent nuclei
spinning of ultrahigh molecular weight PE. In the model merge into shish nuclei.
proposed by Smook and Pennings,220 the solution of ultrahigh 5.6. Advances in Understanding of Shish-Kebab
molecular weight PE can be considered as a highly entangled
network constructed by entanglements with different lifetimes. In the past 20 years, researchers have made great efforts with
time-resolved technologies including synchrotron radiation
The lifetime of entanglements is mainly dependent on the strand
SAXS and WAXS, AFM, and birefringence to unveil the features
length protruding from an entanglement. During extrusion flow,
of shish-kebab.11,93,95−97,131,150,156,205 New observations have
the entangled network deforms inhomogeneously, since the
been obtained and are challenging the conventional models or
entanglements of short lifetime can be easily migrated and the
explanations for shish-kebab. In this section, we intend to
entanglements of long lifetime are more resistant to outline the current understanding of this subject.
deformation. In this process, parts of chains are stretched out 5.6.1. Coil−Stretch Transition or Stretched Network
from the cluster of unoriented molecules. The stretched chains Model? Keller and Kolnaar207 reasoned the CST for shish-
are more likely to crystallize even at higher temperature, due to kebab formation in polymer melt because similar morphologies
the decrease in conformational entropy and the squeezing effect were observed in polymer solution and melt. Actually, no direct
on solvent. Further increasing flow strength, the large clusters evidence is available hitherto. Recent experiments on entangled
break into oriented chain bundles and small clusters. Finally, the polymer melt and cross-linked network suggested that it is
oriented chain bundles evolve into shish and the unoriented unlikely for a long chain to disentangle and undergo CST at
cluster crystallize as kebab. This model was further supported by typical flow conditions.59,116,217,221−223 On the other hand,
SEM results, which provide a possible mechanism for shish- more and more researchers tend to accept that the primary
kebab formation in gel spinning. nucleus of shish comes from deformed network. Actually, early
Janeschitz-Kriegl and co-workers130−132 proposed an explan- in 1984, Smook and Pennings224 suggested that a stretched
ation for the nonlinear relationship between densities of nuclei network should be responsible for shish formation. Recent rheo-
and mechanical loading times in flow-induced nucleation (FIN). optical experiments conducted by Zhang et al.116 showed that
As mentioned in section 3, dormant nuclei exist even in shish can be generated at a very low shear rate, which cannot be
quiescent polymer melt at temperatures below the equilibrium explained by CST. A report on shish composed of both short and
melting point. Those nuclei are local alignments or organized long chains in iPP melt by Kimata et al.147 is not reconciled with
aggregates with a shape of fringed micelle, and they can become the CST model either. Hsiao et al.59 found multiple shish with
effective to induce crystal growth at low temperature or by the SEM from a solvent-extracted sheared PE blend, where adjacent
action of flow. When the strength of flow field (called stress in kebabs are interconnected by several short shish instead of single
their work) is larger than a critical level, the dormant nuclei shish. They argued that CST occurs in the chain sections
1859 DOI: 10.1021/acs.chemrev.7b00500
Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

between kebabs or probably in adjacent entanglement points,


leading to extended chain sections. Furthermore, the rheo-
optical experiments on iPP melt conducted by Seki et al.95
showed that the effect of long chain on shish formation becomes
remarkable after exceeding a critical concentration, which is
about half the overlap concentration of long chain. Ogino et
al.225 reported that, in PE melt, shish was generated when the
long chain is above a certain concentration, about 2.5−3 times
larger than the overlap concentration. Later, they further
pointed out that the critical concentration of long chain is
independent of temperature below 125 °C, while it increases
Figure 24. Critical extension ratio to induce shish formation as a
with temperature above 125 °C.197 One may argue that the function of chain concentration at 131 °C. The red line presents the
materials used in the above studies were broadly dispersed in fitting with the stretched network model (inset equation). Reprinted
matrix or long chain or both. However, the blends of well- with permission from ref 227. Copyright 2015 American Chemical
defined monodisperse samples of h-PBD in the work of Heeley Society.
et al.226 also gave the same conclusion, that is, the formation of
shish at a concentration of long chain around the overlap
concentration. These results imply that long chains should be
deformed as an entangled network rather than individual chain
behavior.
On the basis of results measured by in situ synchrotron
radiation X-ray scattering and homemade extensional rheom-
eter, Yan et al.217 argued that the occurrence of CST does not
need chains to undergo a disentanglement process in PE melt.
This argument can be justified by a combination of rheological
and structural information (Figure 23). On one hand, the
disentanglement can be judged by yielding strain. On the other
hand, the appearance of shish structure can be detected by streak
scatting in SAXS. As exhibited in Figure 23c, the yield strain
increases with increasing strain rate. However, the critical strain
for shish formation is constant when the strain rate is sufficiently
large to overcome the Rouse relaxation of chains. Meanwhile, Figure 25. Correlation between flow-induced nucleation morphologies
the critical strain for shish formation was found to be smaller and strain in polyethylene. Reprinted with permission from ref 144.
than the yield strain, which demonstrates that shish have already Copyright 2013 American Chemical Society.
been generated before CST occurs. These results excluded CST
for shish formation, and the stretched network is proved to be shish formation should be attributed to network stretching
the contributor in PE melt. instead of CST. Yang et al.228 further detected real chain
More recently, Cui et al.227 prepared a series of bimodal PE conformation in a similar lightly cross-linked system, composed
blends with various long-chain concentrations. In all bimodal of deuterated PE/hydrogenated PE blend, with real time SANS;
blends, only the long-chain component meets the requirement see Figure 26. Their results demonstrated an expectedly small
of chain stretch to induce shish formation. They utilized the chain deformation of about 1.3 for shish formation, which does
scaling law between entanglement molecular weight and not support CST. Yang et al.228 pointed out that the coupling
concentration of long-chain component to justify whether between chain conformation and density, rather than intrachain
CST or stretched network is responsible for shish formation. If conformation alone, is the key factor for shish formation.
the stretched network model holds true, the critical strain to While in iPP, Cui et al.97 revealed that shish is generated after
induce shish formation should follow this scaling law. As the collapse of entangled network during flow. They studied the
depicted in Figure 24, the critical extension ratio (strain) for rheological and crystallization behaviors in a wide flow
shish formation decrease with increasing long-chain concen- parameter space and found that at each strain rate a fracture
tration, which is fitted with the stretched network model well. strain exists. Unexpectedly, a sample can maintain integrity
With a unique PE system consisting of cross-linked network without fracture if a strain larger than the fracture strain is
and free chains, Liu et al.144 revealed four types of nucleation imposed on the sample directly. Before and beyond fracture
morphology in strain space, which coincides nicely with the four strain zones were defined, with the fracture strain as boundary.
regions defined by stress−strain curve. From rheological and SAXS results indicated that the before and beyond fracture strain
structural information, they established a correlation between zones correspond to weak and strong acceleration in
flow-induced conformations of chains and morphologies of crystallization kinetics, respectively, accompanying the tran-
nuclei, as shown in Figure 25. The uncorrelated oriented point sition from point nuclei to shish nuclei. They proposed a ghost
nuclei in region I result from the orientation of cross-linked nucleation model, in which whether a sample fractures or not
network and free chains, while the formation of scaffold-network depends on the interplay between the initial entangled chain
nuclei in region II is due to the disentanglement of free chain. network and the physical cross-linked network constructed by
Not only orientation but also stretch of chain segments is the ordering structure or point nuclei induced by flow. As
required to induce the micro-shish in region III, and finally, exhibited in Figure 27, the formation of point nuclei has two
nearly extended-chain segments give rise to shish nuclei in effects. On one hand, the chain segments between point nuclei
region IV. As the sample they used is permanently cross-linked, are more easily oriented or stretched. On the other hand, the
1860 DOI: 10.1021/acs.chemrev.7b00500
Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

Figure 26. Stress−draw ratio curve and in situ 2D SANS patterns of lightly cross-linked PE with different draw ratios (number in the upper left corner
of each pattern) of the extension processes at 170 °C. Reprinted with permission from ref 228. Copyright 2016 American Chemical Society.

HMW PE in noncrystalline LMW PE copolymer with rheo-


SAXS and rheo-WAXS, Yang et al.230 demonstrated that the
HMW component dominates the formation of FIP. Balzano et
al.120 studied a specially synthesized HDPE bimodal blend at
temperatures slightly above the equilibrium melting point and
concluded that long chains construct the backbone of shish. In
addition, Ogino et al.225 combined SAXS and SANS as well as
DPLS to investigate the blends of LMW deuterated PE and
HMW hydrogenated PE under shear flow. Their results implied
that shish were mainly generated from the HMW component.
Moreover, the simulation results of Wang et al.157 from a melt of
short- and long-chain blend concluded that the crystallization of
oriented long chains forms shish, which induces the subsequent
crystallization. However, a contrary conclusion was drawn by
Kimata et al.147 with deuterium labeling to distinguish different
chain lengths in iPP. Their results highlighted the presence of
LMW component in shish; that is, shish consist of both short
Figure 27. Ghost nucleation model. Path I indicates the chain segments
between nuclei are easily deformed to induce new nuclei, and path II and long chains. They suggested that the long chain may play a
indicates that the initial nuclei N0 move to Nt during flow, which catalytic role, recruiting short chains into the formation of shish.
induces daughter nuclei along their trails. Reprinted with permission Those contradictory viewpoints about shish were recently
from ref 97. Copyright 2012 American Chemical Society. unified by Zhao et al.231 Two simplified mixtures of free and end-
linked PEO were designed to mimic the blends of short and long
movement of initial nuclei provides a dynamic template for chains. Their results indicated that, without the entanglement−
surface nucleation. This model explains their experimental disentanglement transition (EDT) in end-linked network, the
results well and provides a new explanation for shish-kebab formation of shish is a synergetic effect between the stretch of
formation. Wang et al.229 studied the effect of formed ordering network and the fast diffusion of free chains. In other words,
structure on nucleation and crystallization in more detail. They short chains are involved in the formation of shish, consistent
utilized a two-step flow protocol and revealed a transition from with the report of Kimata et al.147 This situation may be the
chain to crystal network in extension-induced crystallization of general case in the polymer melt, where highly entangled chains
iPP. During the interval between two extensional operations, the are unlikely to disentangle themselves, while with the
structures induced by the first strain evolve into different stages appearance of EDT in end-linked network, the long chains are
before application of the second strain. As the interval time the main contributors for shish formation, in line with the results
increases, large-scale lamellar crystallites are formed, and during of Yang et al.,230 Balzano et al.,120 Ogino et al.,225 and Wang et
the second flow, the resultant concentration of stress on the al.157 This situation may be the general case in polymer solution.
crystal network increased, leading to an unusual increase of Actually, the noncrystalline PE in the work of Yang et al.230 and
crystallization half-time. However, if the crystal network is short chains at temperatures above the equilibrium melting
continuously perfected, the dynamic symmetry of the system point in the work of Balzano et al.120 can be considered as
can be rebuilt, and crystallization kinetics is accelerated again. solvent from a general point of view.
5.6.2. Is Long Chain the Only Contributor for Shish 5.6.3. How Does Polydispersity Influence the Thresh-
Formation? As mentioned earlier, whether in the CST old for Shish Formation? Polydispersity seems to be a
model207,209 or from the rheological point of view,155 only prerequisite for shish formation.207 Here we focus on the
chains above a critical molecular weight undergo stretch at a influence of polydispersity on the threshold for shish formation,
specific flow strength. Thereby, long chains were thought to be in which the effects of concentration and length of long chain are
mainly responsible for shish formation. Using a crystalline mainly discussed. Mykhaylyk et al.150 reported that the critical
1861 DOI: 10.1021/acs.chemrev.7b00500
Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

Figure 28. AFM amplitude images of the growth process of lamellae from the oriented backbone at various temperatures: (a) 132.5, (b) 132.3, (c) 132,
and (d) 131 °C. Scale bars represent 300 nm. Reprinted with permission from ref 136. Copyright 2001 American Chemical Society.

specific mechanical work for shish formation is inversely shish in a cooperative manner. When matrix viscosity is low, the
proportional to the long-chain concentration, which stems long chains have low orientation under flow but a fast diffusion
from the change in Rouse relaxation time of long chain due to rate. On the contrary, when matrix viscosity is high, the long
polydispersity. With a molecular weight ratio between HMW chains have high orientation but slow diffusion rate. The
and LMW components of about 5, Seki et al.95 demonstrated an interplay between those two factors influences the arrangement
almost constant critical stress of 0.12 MPa for shish formation, of shish. Recently, Okura et al.232 gave some quantitative results
which is independent of HMW concentration. while with use of for the effect of matrix on shish formation. By studying
the same LMW component but an increase of this ratio to 20, bidisperse blends of HMW h-PBD in LMW polymer matrices
the critical stress decreases significantly with HMW concen- under shear flow, they found that the critical specific mechanical
tration.139 Fernandez-Ballester et al.139 ascribed that the work for shish formation (revealed by an oriented morphology)
differences correlate to the onset of chain stretch during flow. has a power law dependence on the molecular weight of matrix,
Recently, Cui et al.227 found that the critical strain for shish in which the index is 2.59 ± 0.27. They speculated that the
formation decreases with increasing long-chain concentration, matrix influences the viscosity of blends, which opposes flow and
which agrees pretty well with the stretched network model. makes the long chain difficult to stretch. Thereby, the critical
Their quantitative analysis further demonstrated that the specific mechanical work increases with increasing molecular
formation of shish is determined by the deformation degree of weight of matrix.
long-chain entangled network, but not by a sole parameter such 5.6.5. In Situ Observation of Evolution of Shish-Kebab.
as strain or long chain concentration. With the help of in situ characterization techniques, in both real
5.6.4. How Does Short-Chain Matrix Influence Shish space (i.e., POM, AFM) and reciprocal space (i.e., SAXS and
Formation? As discussed above, the role of long chains in shish WAXS), the detailed evolution process of shish-kebab can be
formation was investigated quite extensively. However, less traced. Here we take AFM and X-ray scattering as examples to
attention has been paid to the effect of matrix, that is, short illustrate in situ studies of shish-kebab.
chains. The length and concentration of long chain have a strong AFM is now becoming a standard method to provide real-
influence on shish formation. Similarly, the matrix is also space information on the nanometer scale. Hobbs and co-
expected to have an influence. Using blends of long chains of workers96,233,234 developed a novel rapid AFM technique
crystalline PE with two different short chains of noncrystalline (video-AFM) with the capability to detect crystal growth on
PE copolymer, Yang et al.230 studied shear-induced shish the millisecond time scale and applied this technique to study
formation by in situ rheo-SAXS and rheo-WAXS. Their results FIC for the first time. Their AFM works on shish-kebab
suggested that the matrix viscosity influences the formation of formation are especially interesting, as structure evolution on the
1862 DOI: 10.1021/acs.chemrev.7b00500
Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

Figure 29. Critical strain for nucleation from (a) SAXS and (b) WAXD as a function of temperature. Reprinted with permission from ref 123.
Copyright 2014 American Chemical Society.

Figure 30. Schematic of morphologies and structures of four regions in the strain−temperature space, deduced from both WAXD and SAXS.
Reprinted with permission from ref 123. Copyright 2014 American Chemical Society.

lamellar scale was revealed in real time and in real decrease in volume fraction of shish due to internal stress and
space.96,135,136,235 Oriented PE melt was obtained by dragging relaxation of stretched chains, in line with the viewpoint of
a razor blade across the melt on glass coverslip. It should be Gutierrez et al.117 and Balzano et al.120 that defective or small-
noted that the direct observation of shish growth was not size shish are unstable and may relax into melt. Interestingly, in
realized. However, the subsequent overgrowth of kebab could be 2/98 blend, the length of shish was found to increase after 400 s,
monitored, which exhibits a number of interesting features. As which can be explained by the autocatalytic growth mechanism
shown in Figure 28a, when the melt was cooled from 133 to proposed by Petermann et al.236 For 5/95 blend, the length of
132.5 °C, a new nucleation event of one lamella occurs on the shish decreased monotonously during isothermal process,
shish (marked A). When the neighboring shish-kebabs meet consistent with the recent report of Cui et al.,227 who argued
each other, interdigitations may happen (marked B in Figure that if only defective and short shish melted during isothermal
28b). The kebabs change their directions to avoid joining, process, the average length of shish should increase rather than
implying that the advancing lamella has a strong influence on the decrease. Cui et al.227 pointed out a new relaxation process by
melt. In some case, lamellae can even join together to form a which some defective sections in a long shish may relax and give
single long crystal (marked C in Figure 28d). Hobbs and rise to several short shish, hence decreasing the average length of
Miles136 further measured the growth rate of lamella and found shish. Note that Ma et al.237 recently found that, during
different lamellar growth rates. Even for individual lamellae, the relaxation, the disappearance of SAXS signal does not mean the
growth rate was not constant with time, different from the complete relaxation of shish into ideal random coil, since the
constant growth rate of spherulites observed by POM. partially ordered intermediate structures that are below SAXS
In this AFM study, flow field is difficult to define and only detection limit may be formed, which can crystallize into shish
qualitative results were obtained. Quantitative information can again upon further cooling.
be obtained through X-ray scattering techniques. Keum et al.153 5.6.6. Shish Formation under Near-Equilibrium and
investigated the formation and stability of shish-kebab with in Far-from-Equilibrium Conditions. From the thermodynam-
situ rheo-SAXS and rheo-WAXS. In their experiment, PE blends ic point of view, entropy-reduction model (ERM) is the most
with long-chain concentrations of 2 and 5 wt % were used, which recognized mechanism for FIN, which is based on the idea of
were named 2/98 and 5/95 blends, respectively. They chose a flow-induced reduction of conformational entropy.238−240 In
high shear temperature (142 °C), to allow the sole formation of ERM, the entropic reduction or external work is directly added
shish without kebab, and performed isothermal measurements to the driving force of nucleation in classical nucleation theory,
at this temperature. In both blends, shish scattering intensity which lowers nucleation barrier and enhances nucleation rate.
decreases within the initial 400 s and afterward increases with The ERM predicts a strain−temperature equivalence; that is,
time. They reasoned the initial intensity decrease was due to the lowering temperature and increasing strain have the same effect
1863 DOI: 10.1021/acs.chemrev.7b00500
Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

for nucleation. However, it should be noted that in ERM the appearance of streak scattering in SAXS, crystalline scattering in
external work is treated as a perturbation, restricting itself to an WAXS, and stress-upturn in the stress−strain curve in Figure 31.
equilibrium thermodynamic framework, which may be not Unexpectedly, they discovered that the critical strain for shish
applicable to far-from-equilibrium conditions. Recently, Liu et formation remains almost constant in the temperature range
al.123 and Cui et al.30 designed two model experiments to study
nucleation under near-equilibrium and far-from-equilibrium
conditions. In the near-equilibrium experiment, Liu et al.123
verified the validity of ERM and found that free energy
determines the transition from point nucleus to shish, while in
the far-from-equilibrium experiment, Cui et al.30 invalidated the
prediction of strain−temperature equivalence in classical ERM
but demonstrated a constant critical strain for nucleation in a
wide temperature range, which reveals the nonequilibrium
nature of nucleation under strong flow.
5.6.6.1. Near-Equilibrium Experiments. With a series of
lightly cross-linked HDPE, step extension with a strain rate of
0.02 s−1 and a total strain of 2.8 were imposed to induce
nucleation. A slow strain rate was selected in order to approach a
near-equilibrium experimental condition. By a combination of
SAXS, wide-angle X-ray diffraction (WAXD), and stress−strain
results, the incipient strain (εi) for nucleation can be obtained by
the first appearance of shish scattering in SAXS or crystalline
scattering in WAXD. As shown in Figure 29, for each sample εi
increases with temperature, while at the same temperature εi
decreases with increasing irradiation dose or cross-link density.
S15, S30, and S50 represent the absorbed doses of 15, 30, and 50
kGy, respectively. Meanwhile, four regions can be defined in
strain−temperature space according to the formed crystal
morphology and structure (Figure 30): orthorhombic lamellar
crystal (OLC), orthorhombic shish crystal (OSC), hexagonal
shish crystal (HSC), and oriented shish precursor (OSP). Figure 31. (a) Engineering stress (σengr) and corresponding 2D (b)
This work indicates that flow not only reduces the entropy of SAXS and (c) WAXS patterns evolve during flow at 140 °C. The flow
direction is horizontal. Reprinted with permission from ref 30.
initial melt but also modifies the free energies of final states,
Copyright 2015 American Chemical Society.
which were often overlooked in classical ERM in FIN. They
incorporated the free energies of various final states in ERM, from 130 to 170 °C (Figure 32), which breaks the strain−
which agrees with their results pretty well. With this modified temperature equivalence predicted by ERM but unveils the
ERM, the estimated critical thicknesses for OLC, OSC, and nonequilibrium nature of FIN. To account for this temperature
HSC nuclei are on the order of 10, 100, and 500 nm,
respectively, implying that the critical thickness of nuclei
determines the transition from point nuclei to shish nuclei.
This work demonstrates the validity of classical nucleation
theory for FIC under near-equilibrium conditions, provided that
the entropy reduction of initial melt and the free-energy changes
of final states induced by flow are taken into consideration.
5.6.6.2. Far-from-Equilibrium Experiments. Nucleation
under strong flow is a typical kinetics-controlled process,
where the equilibrium ERM may not be applicable. Validation of
the equilibrium ERM theory under strong flow condition
requires simultaneous acquisition of nucleation and stress−
strain information during flow, which is rather challenging. As Figure 32. Critical strain for shish formation at different temperatures.
FIN is extremely fast (on the order of tens of milliseconds) Dashed red line indicates the average critical strain measured from 130
under strong flow, a structural detection technique with high to 170 °C. Reprinted with permission from ref 30. Copyright 2015
time resolution is a precondition, which restricts most studies on American Chemical Society.
FIC that can only detect crystallization after the cessation of flow
rather than during flow.98,124 To construct a strain−temperature independence of FIN, Cui et al.30 proposed a tentative kinetic
diagram for nucleation, the capability to collect stress−strain pathway of nucleation containing stretch-induced coil−helix
during flow is another precondition. and isotropic−nematic transitions, in which flow provides both
Cui et al.30 combined extensional rheological and ultrafast X- the concentration and the compatible alignment of helixes for
ray scattering measurements to reveal the nucleation in iPP melt nucleation. These two transitions are dictated by strong flow
under strong flow, with a strain rate of 12.6 s−1, flow duration less rather than thermal fluctuation, manifesting that FIC is a strong
than 400 ms, and time resolution of X-ray scattering up to 30 ms. external-field-driven nonequilibrium phase transition.
By combining SAXS/WAXS and rheological information, the 5.6.7. Is Shish Formation a Single-Stage or Multistage
critical strain for nucleation can be defined precisely by the Process? Early mechanisms for shish-kebab formation, such as
1864 DOI: 10.1021/acs.chemrev.7b00500
Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

Figure 33. Schematic illustration of shish-kebab formation during (A) spinning and (B−D) shear flow. B and D are optical and birefringent results,
respectively. Reprinted with permission from ref 222. Copyright 2011 American Chemical Society.

Figure 34. Schematic illustration of kinetic process for shish formation. (A) Stretched network, (B) type I shish without lamellar structure, (C) type II
shish with sporadic lamellae, and (D) type III shish containing periodic lamellae. The flow direction is horizontal. Reprinted with permission from ref
227. Copyright 2015 American Chemical Society.

CST or SNM, only propose the correspondence between initial measurements indicated that the crystallinity of inner structure
chain conformation and final crystal morphology. However, it is of the stringlike precursors is rather low, about 0.15%.
unclear how extended chains or stretched networks transform By studying the structure evolution in solution of UHMW PE
into shish. Massive studies on FIN imply that shish may nucleate under shear flow and along spinning line, Hashimoto and co-
via a precursor or metastable stage rather than a direct workers194,222,244 proposed a new scenario, in which flow-
transformation.72,120,124,241−243 On the basis of X-ray scattering induced phase separation, mediated by the dynamic asymmetry
results, Balzano et al.145 and Zhao et al.124 argued that precursors and the stress−diffusion coupling245−247 preceding nucleation,
first generated during flow, which may transform to shish was taken into account. As shown in Figure 33, the nucleation
crystals or dissolve gradually. Works by Kanaya and co- process can be divided into four regions. In the t1 region, PE
workers68,70,71,121 provided more details on the nature of shish entangled network is swollen homogeneously in solvent, as
precursors. Their results showed that large stringlike precursors, evidenced by no feature in the optical image. In the t2 region,
on the micrometer scale, form above the melting point, which plane-wave-type concentration fluctuations with wave vectors
are able to induce shish when cooled. Microbeam WAXS oriented along the flow direction appear. In the t3 region,
1865 DOI: 10.1021/acs.chemrev.7b00500
Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

Figure 35. (a) True stress−time curve during extension of cross-linked PE with strain rate of 3 s−1 at 172 °C, and corresponding SAXS and WAXD
images. (b) Nonequilibrium phase diagram of cross-linked PE in stress−temperature space. L, δ, H and O are melt, noncrystalline shish, H-crystal and
O-crystal, respectively. Reprinted with permission from ref 248. Copyright 2016 Springer Nature.

Figure 36. (a) Nonequilibrium phase diagram of cross-linked iPP in strain rate−temperature space. M, S, α, and β are melt, noncrystalline shish, α
crystals, and β crystals, respectively. (b) Relative content of β crystals in strain rate−temperature space. Reprinted with permission from ref 249.
Copyright 2016 Wiley−VCH.

demixed domains are generated and distributed in space However, whether the noncrystalline shish is a thermodynamic
randomly. In the t4 region, the domains are aligned into strings phase or a kinetic state remains unclear. Wang et al.248 studied
by flow to create bundles of stretched chains. The stretched- flow-induced phase behaviors of a cross-linked PE (with gel
chain bundles and the nearly isotropic demixed domains fraction 43.5%) in a wide temperature range up to 240 °C with
transform into shish and kebab, respectively. This model is ultrafast X-ray scattering. By combining SAXS, WAXD, and
further supported by the transient structures revealed by TEM extensional rheology, they could determine the onset or critical
along spinning line with a special fixation method, which is stress for various structure formations, such as shish structure
rational and may be general for dynamically asymmetric polymer (named δ-phase in their work), hexagonal crystal (H-crystal),
systems. and orthorhombic crystal (O-crystal). Figure 35a is an example
Very recently, Cui et al.227 revealed a kinetic process for shish to show structural evolution during extension at 172 °C, where
formation from initial chain conformation to final stable nuclei. melt, noncrystalline shish, and H-crystal appear sequentially. On
They first demonstrated that the stretched entangled chain the basis of critical stress for the structures formed at various
network is responsible for shish formation. Then, with a temperatures, they constructed a nonequilibrium thermody-
delicately designed thermal history and real-time SAXS method, namic phase diagram for FIC of PE in stress−temperature space.
they further observed three types of shish with different This phase diagram has four regions: melt (L), noncrystalline
stabilities by increasing strain, which can be considered as shish (δ), H-crystal (H), and O-crystal (O) (Figure 35b).
three stages in the formation of stable shish nuclei; see Figure 34. Increasing stress (σ) leads to a sharp increase of L → δ transition
When the deformation of entangled network reaches a critical temperature, a moderate increase of δ → H transition
degree, the stretched chains couple with each other to transform temperature, and a decrease of O → H transition temperature.
into fibrillar type I shish, which is similar to nematic phase. Type Wang et al.248 verified the reversibility of flow-induced phase
I shish has density contrast to surrounding matrix but no transitions by conducting a stress reduction experiment. The
embedded lamellar structure. Type I shish further transforms structure evolution follows a pathway L → δ → H with
into type II shish, containing sporadic lamellar structure, and increasing stress, while a reversed transition H → δ→ L occurs
type III shish, containing lamellar stacks with well-defined with decreasing stress. Interestingly, the critical stresses for H →
periodicity. The barriers for those transformations can be δ and δ → L transitions during stress reduction are smaller than
overcome by flow. It should be noted that the lamellar structure their reversed δ → H and L → δ transitions during stress
mentioned here is within shish, which is different from the kebab increase, indicating that δ ↔ H and L ↔ δ are two pairs of
that grows on the surface of shish. nonequilibrium phase transitions with stress hysteresis.
5.6.8. Is Shish a Thermodynamic Phase or a Kinetic Considering the facts that shish (δ) can be observed solely
State? Noncrystalline shish has been observed at temperatures without crystalline diffraction even at temperatures above 200
near or above the melting temperature of lamellar crystals. °C and that Figure 35b shows a clear direction and pathway for
1866 DOI: 10.1021/acs.chemrev.7b00500
Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

Figure 37. Strain rate−temperature flow structural (a) and morphological (b) diagrams of poly(1-butene). Reprinted with permission from ref 250.
Copyright 2017 Royal Society of Chemistry.

shish (δ) structure transition in stress−temperature spaces, view, it is necessary to understand experimental phenomena on a
Wang et al.248 claimed that shish is a thermodynamic phase and molecular level.
can be stable or metastable depending on the stress and Conformational entropy reduction is commonly considered
temperature. as the origin of FIC. Polymer crystallization is different from the
Ju et al.249 constructed a similar flow diagram of iPP in strain crystallization of small molecules because of the long-chain
rate−temperature space, which is composed of melt (M), feature. The covalent connection of polymer not only restricts
noncrystalline shish (S in their work), α crystals, and an α/β the diffusion of chain segments but also leads to a huge amount
crystal coexistence regions (Figure 36). What needs to be of conformational entropy. For a single chain with a fixed end-to-
emphasized here is that the noncrystalline shish presents alone end distance, the position of segments varies with time and a
in a small region at temperatures above the melting point of α great number of conformations can be obtained. The freedom to
crystals. This region is located at the most violent competition form a large number of possible conformations corresponds to
zone in the strain rate−temperature phase diagram, where large entropy. Evidently, when a coiled polymer chain is
dynamic competitions happen among four involved phases. Ju et deformed, the number of possible conformations is reduced. If
al.249 attributed the formation of noncrystalline shish to the equilibrium phase transition is still assumed with no change of
kinetic competitiveness of shish over α and β crystals but not to free energy, the reduction of entropy directly leads to the
thermodynamic stability. The frustration effect due to the increase of equilibrium melting temperature. Equivalently, the
violent competitions among four phases leaves the noncrystal- undercooling becomes larger, and thus the crystallization is
line shish winning out. For poly(1-butene),250 the constructed facilitated. This is the widely accepted molecular mechanism for
structural and morphological diagram (Figure 37), the same in flow-induced polymer nucleation.
strain rate−temperature space, revealed that metastable form III The quantitative description of FIC needs a specified chain
can be crystallized from the melt, which may further transform model to calculate the conformational entropy. At the very
into form II during flow. Concerning the crystallite morphology, beginning, it is worth pointing out that crystallization of polymer
a transition from flow-induced network to the shish structure mostly happens under a large undercooling, while for the sake of
was also observed upon elevating flow temperature from below simplification, the equilibrium thermodynamics of crystalliza-
to above the melting temperature for form II. tion is used to describe the phase transition. Furthermore, the
A recent work of Liu et al.251 even gave a quantitative enthalpy of crystallization is considered as a constant at different
estimation of the thermodynamic parameter of noncrystalline crystallization temperatures, and the entropy of undercooled
shish in cross-linked PE. They calculated the critical external melt is considered the same as that at equilibrium temperature.
works for melt−noncrystalline shish and noncrystalline shish− With these assumptions, the conformational entropy change is
crystal transitions based on rheology and X-ray scattering the only parameter that needs to be calculated. Various models
methods. The critical external work of melt−noncrystalline will give different expressions of this entropy change. With the
shish transition is larger than that of noncrystalline shish−crystal quantified entropy change, a predictable model for crystal-
transition at low temperature, but the situation is reversed at lization can be obtained.
high temperature. Thermodynamic parameters such as the In the following subsections, we will discuss some molecular
enthalpy and surface energy of noncrystalline shish were theories to describe the crystallization of polymer under either
analyzed by incorporating the critical external work into a strain or flow. This section dates back to the independent works
modified stretch network model.123 Their results help to focusing on cross-linked network and melt. Although the
illustrate the following two issues: (1) the energy barrier of external field is macroscopically imposed by extensional or/and
noncrystalline shish formation increases with increasing temper- shear flow, it is believed that molecular deformation, rather than
ature, and (2) the surface free energy and bulk free energy of macroscopic flow, is the substantial reason for changes in
noncrystalline shish are much smaller than those of correspond- crystallization behavior, so strain-induced crystallization in
ing crystals. cross-linked network and FIC in melt will not be distinguished.
However, crystal morphology and chain relaxation behavior
6. THEORIES AND MODELS OF FLOW-INDUCED differ in the two conditions, which make these models slightly
NUCLEATION different.
6.1.1. Crystallization Induced by Stretching of Cross-
6.1. Molecular Theory of Flow-Induced Nucleation Linked Network. A cross-linked network like vulcanized
Nucleation of polymer can be greatly influenced by flow, from natural rubber can sustain strain without significant relaxation,
kinetics to crystal morphology. From the prediction point of which provides a model system to link chain deformation with
1867 DOI: 10.1021/acs.chemrev.7b00500
Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

the resulting crystallization. Flory developed a thermodynamic where


theory to describe the strain-induced crystallization in cross-
linked natural rubber, which hardly crystallizes at ordinary β′ = β[n/(n − ξ)]1/2 (6)
temperatures. The segments between cross-link points are and z′ is the algebraic sum of the z displacement lengths of
defined as an individual chain, of which the end-to-end distance amorphous sections of the chain.
increases with strain. During stretch, chains are elongated and To calculate the conformation entropy change with respect to
the number of possible conformations gradually decreases, the fully crystallized state, two hypothetical steps are considered
which reduces the entropy. Gaussian distribution is still used to by Flory: (a) n − ξ segments per chain are first melted and the
describe the conformation distribution. It is assumed that chain ends are assumed to be free, and (b) the free chain ends are
crystallization does not happen before chains are stretched to the reassigned to the cross-link points, which leads to a trans-
final length, so the entropy change coming from deformation formation of conformation distribution from ν′(x, y, z′) to ν(x ,y
can be calculated from the strain. When crystallization begins, ,z):
the conformational entropy change is further divided into two
parts. The first part comes from the transition from oriented Sa = σ(n − ξ)sf (7)
amorphous segments to crystals, where the involved segments where sf is the entropy of fusion per segment and is obtained by
are completely frozen by crystallization. This reduction in the lattice model.
entropy is monotonic. The second part originates in the (partial) Sb is calculated by employing the Boltzmann relationship:
release of chain stretching of the residual amorphous segments.
Since crystals in natural rubber are found to orient well along the S = k ∑ ν ln W (8)
stretch direction, the crystals in this theory are assumed to align
along the stretch direction. Segments in the crystals have a more S b = k ∑ ν(x , y , z) ln W ′(x , y , z′)
extended conformation than that in the amorphous phase; thus, xyz
for the same end-to-end distance, the stretching of amorphous
segments will release partially. When crystallinity is low, the − k ∑ ν′(x , y , z′) ln W ′(x , y , z′)
second part of entropy change is positive. With increasing xyz′ (9)
crystallinity, the segment number may become so small that the
The obtained entropy change is given in eq 10:
entropy decreases again. This decrease in entropy will prevent
full crystallization of the segments between cross-link points. S = σk[(n − ξ)sf /k − (ξβl)2 n/(n − ξ)
Obviously, calculation of the entropy change of the
amorphous part under stretch is a major problem. Fully + (2αξβl /π 1/2)n/(n − ξ)]
crystallized state is chosen as the reference state, and partial − (α 2/2 + 1/α)n/(n − ξ) + 3/2 (10)
melting of crystals is considered to give the equilibrium
crystallinity. The detailed theory is briefly described as follows. The free energy of the system, with perfectly ordered and totally
The number of probable conformations of an amorphous crystalline chains as the standard state, becomes
chain is assumed to be dependent on their end-to-end distance r F = σ(n − ξ)hf − TS (11)
by Gaussian function:
If we define λ = (n − ξ)/n and θ = sf/k − hf/kT, the equilibrium
r = (x 2 + y 2 + z 2)1/2 (1) state can be obtained by eq 12 and the result is given by eq 13:

Ä ÉÑ ÄÅ ÉÑ|1/2
∂F/∂λ = 0 (12)

oÅÅÅÅ 3
l
W (x , y , z) = (β /π 1/2)3 exp[−β 2(x 2 + y 2 + z 2)]
ÑÑ ÅÅ 3 Ño
ÅÅ 2 − θ ÑÑÑÑ}
(2)

λ=m
oÅÅÇ 2 ÑÑ
n ÑÖ ÅÇ ÑÖo
~
where x, y, and z represent the coordinates of one end of the − ϕ( α )
chain with respect to the other end and 1/β equals the most (13)
probable value of chain displacement length, r. After stretching
along the z-axis, the distribution of chain coordinates becomes where
ϕ(α) = (6/π )1/2 α /n1/2 − (α 2/2 + 1/α)/n (14)
ν(x , y , z) = σ(β /π 1/2)3 exp[−β 2(αx 2 + αy 2 + z 2/α 2)]
(3) The incipient crystallization temperature Tm, at which
crystallization just happens, can be obtained by assigning λ =
where σ is the total number of chains under consideration and α
1 in eq 13:
is the stretch ratio. The freely jointed chain model with
independently oriented rigid segments is used. For such chains, 1/Tm = 1/Tm 0 − (R /hf )ϕ(α) (15)
the reciprocal most probable displacement length is given by
The force can be obtained by
i ∂F y i ∂F y
f = jjj zzz = jjj zzz
β = (3/2n)1/2 /l (4)
where l is the length of each segment and n is the number of k ∂α {e k ∂α { λ
segments per chain. The product nl is taken to equal the
maximum extension L of the actual chain. = σRT[(α − 1/α 2) − (2nβl /π 1/2)(1 − λ)] /λ (16)
When ξ of the n segments of the chain participate in
crystallization, the relative number of configurations available to When no crystal forms, λ equals 1 and f becomes
the remaining n − ξ segments becomes f = σRT(α − 1/α 2) (17)
1/2 3 2 2 2 2
W ′(x , y , z′) = (β′/π ) exp[−(β′) (x + y + z′ )] Force is an easily measurable parameter; thus, for an
(5) approximation, the integration of force with strain can be used
1868 DOI: 10.1021/acs.chemrev.7b00500
Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

Figure 38. Schematic illustration of the free energies of different morphologies and structures: (a) under quiescent conditions, (b) classical stretched
network only involves the effect of flow on the entropic reduction of melt, and (c) modified stretched network considers different crystal morphologies
and structures in the final stage of nucleation as well. G is free energy, where subscripts L and N represent melt and nuclei, respectively; ΔG* is the
nucleation barrier of critical nucleus; ΔGfN is the free-energy change of nuclei induced by flow. Reprinted with permission from ref 19. Copyright 2016
American Chemical Society.

to indicate the entropy change. Particularly, there should be no 6.1.2.1. Microrheological Model. The Doi−Edwards theory,
crystallization during stretch. Crystallization is the most- based on the tube model proposed by De Gennes, is widely used
encountered reason for deviation from Flory’s theory. Based as a rheological theory of polymer melt. This theory gives an
on this approach, several recent works have been reported.123 analytical expression of the constitutive equation in orientation
In addition to the aforementioned entropic reduction of initial or even strong chain-stretch conditions. With this theory,
melt, Liu et al.123 also considered the flow-induced free-energy Coppola et al.252 have calculated chain deformation and the
change of the final nuclei, which was quantitatively expressed as corresponding entropy change during flow. For the sake of
simplification of mathematical analysis, the independent align-
1
=
1

(
vk αi2 + α − 3
i
2
) ment approximationthat the segments are still of the same
length and number as at equilibrium and that deformation has
4σ only changed their orientation in spaceis introduced.253 Note
Tc(αi) Tc(1) (
2ΔH 1 − ΔHle * ) that this approximation excludes chain stretch, which is actually
important in crystal morphology change.
where Tc(αi) and Tc(1) are the crystallization temperatures at
As described by the Lauritzen and Hoffman theory,
elongation ratios of αi and 1, respectively, and v is the network-
ÄÅ É
K n ÑÑÑÑ
nucleation rate can be expressed as

jij Ea zyz ÅÅ
chain density. Concerning the final nuclei, melting entropy ΔH

z expÅÅÅ−Å
N = CkBT ΔG expjj−
j kBT zz ÑÑ
ÅÅÇ T (ΔG)n ÑÑÑÖ
depends on their specific crystal modification formed, while

k {
surface free energy σe and critical nucleus thickness l* both are ̇
determined by the crystallite morphology. For instance, the (18)
surface free energy σe can be varied from folded-chain to fringe
micellar structures, due to the morphological transition from where C includes energetic and geometrical constants, kB is
folded-chain lamellae to shish. In this way, the flow-induced Boltzmann’s constant, T is crystallization temperature, ΔG is
change of free energy is also considered for the final nuclei, as thermodynamic driving force (the volumetric free-energy
shown by Figure 38.19 difference between melt and crystal), and Ea is diffusion
6.1.2. Flow-Induced Nucleation in Free Melt. Different activation energy across the liquid−nucleus interface. Kn is a
from the permanent constraint on chain mobility in cross-linked constant containing energetic and geometrical factors of nuclei.
networks, the constraint from entanglements is temporary in For primary nucleation, n equals 2, and for secondary nucleation,
melt and the chain can relax via chain movement; for instance, n is 1.
the terminal time for a single chain to move out the original tube When flow is imposed, the volumetric free energy of melt is
is on the order of seconds or less. Two effects must be raised and ΔG changes into
considered due to the occurrence of relaxation in melt: (i)
ΔG = ΔGq + ΔGf
during flow, the deviation between macroscopic strain and (19)
molecular deformation is probably significant, which hinders a
direct correlation between chain deformation and crystalliza- Here ΔGq is the thermodynamic driving force under quiescent
tion; and (ii) after flow cessation, the continuous decay of chain conditions and can be written as

ij T yz
ΔGq = ΔH0jjj1 − 0 zzz
deformation may make crystallization also dependent on time
j Tm z{
k
after flow. A more comprehensive model is needed to include
strain rate and time-dependent chain deformation, in addition to (20)
the thermodynamic model of crystallization.
Rheological theory of polymer melt is a good candidate to where ΔH0 is the latent heat of fusion and Tm0 is the equilibrium
describe chain deformation during and after flow. From the melting point. ΔGf is the driving force contributed by external
rheological point of view, the stress response due to flow is field, as the step strain in this work. If ΔGf is known, nucleation
determined by chain deformation, which is commonly described rate under flow can be predicted. The Doi−Edwards theory with
by the constitutive equation. The correspondence between independent alignment approximation (IAA) was proposed to
macroscopic stress and microscopic chain deformation makes describe the effect of flow in the linear region. The general form
the theory verifiable and gives abundant information about the given by Marrucci and Grizzuti253 is
strain rate and time-dependent chain deformation. Two models t
have been proposed: the microrheological model (section
6.1.2.1) and the dumbbell model (section 6.1.2.2).
ΔGf = 3ckBT ∫−∞ μ(̇ t , t′)A[E(t , t′)] dt′ (21)

1869 DOI: 10.1021/acs.chemrev.7b00500


Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

where c is the primitive chain-segment concentration, equivalent This deduction assumes a monodisperse sample, while in
to entanglement density. It can be expressed as practice a polydisperse sample is commonly used. To
ρ demonstrate the validity of the theory, Acierno et al.254 also
c= NA use this model to describe a long- and short-chain blend of
Me (22) polybutene. The only change is the form of the memory
where ρ is the density of melt, Me is molecule weight between function, which is modified as
entanglements, and NA is Avogadro’s number.
The Doi−Edwards memory function μ(t,t′) gives the memory μDR (t , t ′, τdH , τdL)
of flow at time t when flow was imposed at time t′. It can be = [ϕHμSR (t , t ′, τdH) + ϕLμSR (t , t ′, τdL)]2
ÄÅ ÉÑ
(32)

ÅÅÅ (t − t ′)p2 ÑÑÑ


written as

μ(t , t ′) = 2 ∑ 2 expÅÅÅ− ÑÑ
where ϕH and τdH are the volume fraction and relaxation time of

ÅÅÅ ÑÑ
ÑÑÖ
8 1
Ç
long chain and ϕL and τdL are the volume fraction and relaxation
π p p τd time of short chain, respectively.
odd (23)
The effects of crystallization temperature were also inves-
where τd is the terminal relaxation time. tigated.255 Qualitatively, as temperature increases, chain
For steady-state flow, ΔGf can be rewritten as relaxation becomes faster. To quantitatively capture this change,
+∞ the time−temperature equivalence principle was used, which is a
ΔGf = 3ckBT ∫0 μ(̇ z)A[Wiz ] dz
(24)
standard approach in rheology. The Weissenberg number is
modified by a shift factor α, and other parts of the theory remain
Wi is the Weissenberg number (named as Deborah number in the same. Good agreement was obtained on the experimental
the original reference252), that is, deformation rate multiplied by and modeling induction time.
polymer relaxation time. z is a dimensionless value defined as In summary, the microrheological model developed by
time divided by relaxation time τd. Coppola et al.252 gives the free-energy change induced by flow
For uniaxial elongation deformation with an elongation ratio in steady state and correlates it with the acceleration of
λ, we have crystallization kinetics. The introduction of independent
alignment approximation intrinsically neglects chain stretch
tan−1(λ 3 − 1)1/2 and only accounts for the effect of orientation. The use of
A(λ) = ln λ + −1
(λ 3 − 1)1/2 (25) memory function clearly shows the importance of chain
relaxation in melt.
ÄÅ É
and for shear deformation with strain γ, we can write
ÅÅ 1 + γ 2x 2 + [x 4(γ 4 + 4γ 2) − 2γ 2x 2 + 1]0.5 ÑÑÑ
6.1.2.2. Dumbbell Model. The dumbbell model considers the

ln ÅÅÅÅ ÑÑ dx
ÑÑ
polymer chain as two rigid beads linked by a flexible spring. The
ÅÅÅ ÑÑÑ
1

Ç Ö
1
A(γ ) =
2
∫0 2
vector R, giving the relative position of the two beads, is the only
parameter to describe chain deformation. Evidently, this model
(26) is simpler than the Doi−Edwards theory. Several works have
In the small Wi limit, the integration in eq 24 has the form shown been reported to describe the crystallization of polymer based on
in eqs 27 and 28 for shear and extension: this model.
The configurational distribution function of the end-to-end
ΔGf π4
Wi 2 vector R can be written as

jij[κ·R]ψ − 2kBT ∂ ψ − 2 F ψ zyz


=

jj e z
3ckBT 600 (27)

z
k {
∂ψ ∂
=−
ΔGf π4 ∂t ∂R ζ ∂R ζ (33)
= Wi 2
3ckBT 200 (28)
where κ is the velocity gradient tensor, kB is Boltzmann’s
In the large Wi limit, the form for shear and extension will be constant, ζ is bead friction coefficient, and Fe is elastic force of
expressed as shown in eqs 29 and 30: the spring. For a Hookean spring, the force is
ΔGf Fe = K R (34)
= ln Wi − 2.41
3ckBT (29) where K is the spring elastic constant. For a finite extendable
2 nonlinear elastic dumbbell, the force will be
ΔGf π
= Wi − 1 KR
3ckBT 12 (30) Fe =
1 − ⟨(R /R 0)2 ⟩ (35)
Clearly, the microrheological model only describes steady-state
flow, which is not valid after crystallization happens. The authors where R0 is the maximum length of the dumbbell.
thus compared the predicted induction period with exper- The free energy of the system A, for the deformed chain under
imental value, where the crystallinity is rather low. flow, is
Nq̇ A = nkBT ln⟨ψ /ψ0⟩ (36)

ÅÄÅ É
Θ=

yzÑÑÑÑ
Nḟ
ÅÅ K ij
where ψ0 is the equilibrium distribution and n is the number

expÅÅÅ Å j
j z
− 1zzÑÑÑÑ
z
j
ÅÅ T (ΔGq )n jj (1 + ΔGf /ΔGq )n zÑÑ
1 1 density of molecules. In entangled melt, n is equivalent to the

ÅÇ k {ÑÖ
n
= entanglement density.
1 + ΔGf /ΔGq
With eqs 33−35, when flow field is assigned by the tensor κ,
(31) the distribution function ψ can be obtained by eq 33. When ψ is
1870 DOI: 10.1021/acs.chemrev.7b00500
Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

known, the free-energy change after flow can be calculated by eq When the expression for the Hamiltonian is substituted, the
36. rate of crystallization is
ij T yzz
Z ΔHujjjj1 − zz
In the work of Coppola et al.,252 the free-energy change by

Tm 0 z{
∂ϕc ρ
k
dumbbell model is also used as a comparison. The free-energy =
ÄÅ ÉÑ
nkTZ ÅÅÅÅ Ñ
change obtained by eq 36 is assigned directly as ΔGf. The ∂t Mu
ij 24Nc11 yz 2Ñ ÑÑ
Å
Å j
j z
z c Ñ ÑÑ
formula describing crystallization kinetics follows the Laur-
2(1 − ϕc)2 ÅÅÅÇ
1/2

k π { ÑÑÖ
itzen−Hoffman theory. It is found that the simplification of − tr c − + 6N ϕ − 3Nϕ
ÄÅ ÉÑ
̅ c

ÅÅ ÑÑ
chain conformation gives a larger free-energy change than that in
ÅÅ ÑÑ
nkTZ ÅÅÅ ÑÑ
ÑÑ
the Doi−Edwards theory. 1/2

ÅÅ
2 ÅÅÅ c c c + 3ϕ 2Nc c − ϕ 24Nc11̅ 1/2 c c ÑÑÑÑ
24Nc11
Bushman and McHugh256 further use the Hamiltonian to 6ϕ c
Nc c
̅ 33
22 ̅ − ( π
̅
)
c c
̅ 33
22 ̅

ÅÅ 11 ̅ ÑÑÑÖ
+
ÅÇ ̅ 22
derive the time evolution of crystallinity. When crystallization is
taken into account, the free-energy change of the remaining ̅ 33 ̅ c ̅ 33
22 ̅ c (
π )
̅ 33
22

ÅÄÅ ÑÉÑ
amorphous part by the dumbbell model becomes (41)

nkBT ÅÅÅ ijjj 3⟨R cR c⟩ yzzzÑÑÑ


ln ÅÅÅdetjj 2 zÑÑ
The conformational tensor during crystallization can be

ÅÅ j ⟨R N ⟩(1 − ϕ ) zzÑÑÑ ÄÅ ÉÑ
obtained as
ÅÇ k c {Ñ ÅÅ ÑÑ
Ö ÅÅ c11c 22 c33 − ϕ 6Nc11̅ Ñ
A=−

ÄÅ ÉÑ ̅ = 2ε ̇c + 1 ÅÅÅÅ ̅ ̅ ̅ ̅ ÑÑÑ
2
Å Ñ ÑÑ
1/2

Å Ñ λ ÅÅÅÅ c11 ÑÑ
( )
3nkBT ÅÅ Ñ
0
c c

− 1ÑÑÑÑ
ÑÑ
∂ c11 ̅ 33

ÅÅ
c π 22

ÅÅ ÑÑ
Å
Å ÑÑ ÅÇ ÑÖ
2 ̅
11

2 Å ⟨R N0 ⟩(1 − ϕc)
⟨R c ⟩ ∂t ̅ c 22
̅ c33
2
̅ + 3ϕc Nc 22̅ c33
ÅÇ ÑÖ Ä É
̅
ÅÅ 1/2 Ñ
+

i 24Nc11 y Å i 6Nc11 y ÑÑÑÑ


2

j z ÅÅ j z
− ϕcjj zz c 22 ̅ − λ(1 − ϕ ) ÅÅÅÅ c11 ̅ − ϕcjj π zz ÑÑÑÑ
(37) 1/2

k π { c Å k { ÑÖ
1
Ç
̅ ̅
Upon conversion of Rc to the end-to-end distance of whole ̅ c33
molecules, RN, the free-energy change in extension becomes

nk T ij ⟨RN ⟩(1 − ϕc) yzz


(42a)

A = B ln jjjj zz
zz
3

j
2 ∂ c 22 c 22
̅ = − 1 ε(1 1 ̅
k {
̇ + p) c 22
̅ + −

ÅÄÅ ÑÉÑ
2 3 ∂t 2 λ λ(1 − ϕc) (42b)

nkBT ÅÅÅ ij 8c11 yz ÑÑ


ln ÅÅdet c + c 22c33l − j
ÅÅ
j z lc 22c33ÑÑÑÑ
z
1/2 ∂ c33 c33
k { ÑÑÖ
̅ = − 1 ε(1 1 ̅
ÅÇ
2 ̇ − p) c33 + −
− ̅
ÄÅ ÉÑ
∂t 2 λ λ(1 − ϕc)
ÅÅ Ñ
2 π (42c)

ÅÅtr c + l 2 − jij 8c11 zyz l ÑÑÑ − 3nkBT


ÅÅ j z ÑÑ
where λ is the characteristic relaxation time of the medium,
2⟨RN ⟩(1 − ϕc) ÅÅÅÇ k π { ÑÑÑÖ
1/2
3nkBT
+ 2 which is chosen to reproduce the known result for a
2 noncrystalline system.
(38) The dumbbell model used above has two interesting points
where c is the conformation tensor of the whole molecule and l is making it different from other models, as the authors pointed
the length of crystal. out. First, it can simultaneously account for the dynamics of flow
The Hamiltonian can be expressed as follows: and the crystallization that is occurring. Second, the finite

l
extensibility of molecules is taken into consideration in the finite
o
∫Ω omooo 21ρ M·M + σem
extendable nonlinear elastic dumbbell. These two characters are

n
H= clearly different from the model of Marrucci and Grizzuti,253 in

ij T yzz 3nkT
which only induction time under orientation conditions can be
ϕcΔHujjjj1 − zz +
Tm 0 z{
ρ obtained.
k
− ln (1 − ϕc)
ÄÅ ÉÑ
Titomanlio and co-workers focus on the modeling of

nkT ÅÅÅÅ ÑÑ
Mu 2
ij 24Nc11 yz ÑÑ
crystallization during real polymer processing, for example,
ln ÅÅ c11 − ϕcjj zz c 22 ÑÑ
ÅÅ ÑÑ
1/2
film casting257,258 and injection molding.259 In their works, both
ÅÇ k π { ÑÖ
− c 22 c33 2
+ 3ϕc Nc 22 c33 ̅ c33
ÄÅ É
̅ ̅ ̅ ̅ ̅ ̅ ̅ the evolution of crystallinity and crystal morphology under
1/2 Ñ |
2

nkT ÅÅÅ Å Ñ o 3
i 24Nc11
ÅÅtr c̅ + 3ϕc 2 − ϕcjjj ̅ yzz ÑÑÑÑ − 3nkT o
nonisothermal crystallization processes are modeled. The
z ÑÑ }
odx
2(1 − ϕc) ÅÅÇ Å k π { ÑÑÖ 2 o o
dumbbell model is used in their model to describe chain
~
+
extension, which is transformed to entropy change and further to
elevation of the melting temperature of crystals. The complete
(39)
model is rather complicated, since many factors affecting the
Here M = ρv is the momentum density, Mu is the molecular crystallization have to be considered. Introduction of the details
weight of a single chain, σem is the interface energy between is beyond the aim of this review.
crystal and amorphous melt, and c̅ = 3c/Nb2 is the non- 6.1.3. Quantitative Description of Extension-Induced
dimensionalized conformation tensor. Nucleation. 6.1.3.1. Relaxation-Related Long Period Evolu-
Upon introduction of a parameter Z describing crystallization tion. It is widely accepted that, during flow, molecular
kinetics, the time evolution of crystallinity can be given by deformation depends on the competition between flow strength
and chain relaxation. However, whether chain relaxation plays a
∂ϕc δH role in FIC after flow is a question that is paid little attention.
= −νβ∇β ϕc − Z
∂t δϕc (40) Optical microscopic results show that all spherulites have the
same dimension, which indicates all nuclei form at the same
where νβ is a component of the velocity tensor and Z is a fitting time. Besides, the rate of lamellar growth is mainly determined
parameter. by the crystallization temperature. Then, it seems that chain
1871 DOI: 10.1021/acs.chemrev.7b00500
Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

relaxation does not affect the FIC after flow. Intuitively, just after
flow, the chain relaxation affects chain conformation, which
leads to a time-dependent entropy loss. According to nucleation
theory, the nucleation rate should correspondingly change with
time. It is convincing to us that the apparent discrepancy is
rooted in the length scale covered by the experiments, which is
spherulite scale for microscopic experiments but lamellar details
for X-ray methods.
It was found that at high extensional rate the long period of
PEO first increases just after flow; see Figure 39.260 The
abnormal increase upon flow cessation is different from the
increase of long period induced by lamellar thickening; the latter
takes a long time for chain segments in crystals and amorphous
phase sliding collectively. A model based on relaxation is
proposed to explain the increase in the distance between two
lamellae, as shown in Figure 40.
After cessation of the step strain at time t0, the highly oriented
molecular chains tend to gradually relax back to equilibrium Figure 40. Schematic illustration of nucleation model after step strain.
The nucleation density decreases (long period increases) with
random-coil state, during which FIN takes place. In the situation nucleation time, due to relaxation of polymer chains. Reprinted with
permission from ref 260. Copyright 2011 American Chemical Society.

happens during stretch. The free-energy change at time t after


cessation of deformation (t = 0 s) is the product of the memory
function and a constant, A0, which is the free-energy change in
the absence of relaxation:
ΔGf = 3ckBTA(t ) = 3ckBTA 0μ(t ) (45)
The nucleation rate of unit volume can be written as eq 46,

ÄÅ ÉÑ
ÅÅÅ ÑÑ
following the microrheological model:
ij Ea yz ÑÑ
Ṅ = CkBT ΔG expjjj− zz expÅÅ− Ñ
j kBT zz Å
ÅÅ T (ΔGq + ΔGf ) ÑÑÑ
K
k { ÅÇ ÑÖ
n
n

(46)
Figure 39. Abnormal increase of long period during extension-induced
crystallization of PEO. (a) Two-dimensional scattering patterns. (b) The volume is estimated by assuming that the filling rate has a
Lorentz-corrected one-dimensional intensity profiles. (c) Time linear relationship with the unoccupied volume, as eq 47 shows:
evolution of the long period. Reprinted with permission from ref 260.
Copyright 2011 American Chemical Society.
d V (t )
= B[V0 − V (t )]
dt (47)
where V(t) is the volume occupied by nuclei and V0 is the total
of high orientation, the long period is assumed to be
volume. B is a constant with units of reciprocal seconds and
proportional to the inverse of linear density of nuclei, as
reflects the speed of nucleation. B will be referred as the volume
shown in eq 43:
filling rate hereafter. Solving eq 47, we get
V (t )
L (t ) ∼ V (t ) = V0[1 − exp( −Bt )] (48)
N (t ) (43)
The total nucleation rate depends on the remaining volume;
Here the symbol ∼ means the coefficient is omitted. The density hence eq 46 has to be multiplied by exp(−Bt). Subsequently the
of nuclei decreases with time due to relaxation of the oriented number of nuclei N(t) can be obtained:
polymer network. Correspondingly, the long period increases
t
with nucleation time; namely, Lt 1< Lt 2, with t1< t2 as N (t ) ∼ ∫0 Ṅ (t ′) exp( −Bt ′) dt ′
(49)
schematically illustrated in Figure 40. The long period obtained
with SAXS is the inverse of nucleation density averaged over With the time-dependent number of nuclei and the occupied
total nucleation time, which also increases with time. volume, the long period obtained from SAXS is numerically
The original microrheological theory is extended to the step fitted. The fitted curves of the long period are presented in
strain case to interpret the evolution of long period Figure 41. The volume filling rate B, reflecting the speed of
quantitatively. Generally, when a deformation with finite time nucleation, is given in Figure 42, which shows good agreement
is considered, the free-energy change can be calculated as with the time of onset of a distinct scattering peak.
t One drawback of this model is the exclusion of chain stretch,
A(t ) = A[E(t , 0)] − ∫0 A[E(t ′, 0)] μ(̇ t − t ′) dt ′
(44)
which is rooted in the microrheological model by Coppola et
al.252 Nevertheless, chain stretch is definitely very important in
When step strain with a strain rate much higher than is τd−1 FIC of polymer. Besides the abrupt morphology change, some
considered, it is a good approximation that no relaxation interesting changes in long period evolution have been
1872 DOI: 10.1021/acs.chemrev.7b00500
Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

found.261,262 How to describe these transient changes is still an equivalent chain stretch, which originates in the surface
open question in modeling. crowding of PEO, can be obtained. The thickness of PLLA
6.1.3.2. Crystallization under Single Molecular Stretching. lamellae determines the strength of stretch imposed on PEO. By
Crystallization theories, particularly those based on rheological use of this block copolymer, a single molecular stretch was
achieved, as presented in Figure 43.
The crystallization kinetics can also be described by classical
nucleation theory (CNT) incorporating free-energy change
induced by stretch. The free-energy change by stretch can be
given through surface crowding theory, as shown in eq 50:

π 2kBTcNA
ΔGf = lPEO2
16NMb2 (50)

where N is the number of segments with Kuhn length b, M is the


molecular weight of PEO block, NA is Avogadro’s number, and

Figure 41. Comparison of calculated and observed long period.


Reprinted with permission from ref 260. Copyright 2011 American
Chemical Society.

Figure 43. Illustration of single molecular stretch and the effect of


temperature. Reprinted with permission from ref 266. Copyright 2011
American Chemical Society.

lPEO is the length of PEO block and is calculated from the density
of crystals and amorphous region.
The induction time is obtained by the decrease of SAXS
intensity, since after PEO crystallizes the density contrast begins
to decrease. The fitted induction time is presented in Figure 44.
The good agreement indicates a well-controlled single molecular
Figure 42. Comparison of 1/B and the time of onset of scattering peak. stretch.
Reprinted with permission from ref 260. Copyright 2011 American 6.2. Macroscale Continuum Modeling
Chemical Society. Modeling provides a very powerful tool to understand FIN.
Focusing on different length and time scales of polymer
models, often assume that material system studied is crystallization, there are various classes of modeling methods,
monodisperse. The assumption simplifies the derivation of including full molecular dynamics simulations,267−269 lattice
theory but impedes its comparison with experimental results. As Monte Carlo,270,271 coarse-grained kinetic Monte Carlo,272,273
the material practically used is polydisperse, sometimes the and macroscopic continuum models.274,275 The bottom-up
response to flow is dominated by the long-chain portion, which strategy is to develop detailed models by resolving the motion of
composes only a minority of the material.221,230,263 Even with a individual atoms. However, the corresponding high computa-
monodisperse sample, the inhomogeneous flow in strong flow tional cost hinders its wide application to the crystallization of
field may also disrupt the comparison between experimental and polymers on practical length and time scales, especially under
theoretical results.264,265 How to stretch a polymer chain in a the mutual influence of severe processing-relevant (flow,
controllable way and how to correlate deformation with temperature, and pressure) conditions.
subsequent crystallization are still unsolved issues. Alternatively, the top-down investigation starts with deriva-
Stretching a single chain directly is a promising way to solve tion of macroscopic continuum models from the empirical
the above problem. A poly(L-lactic acid) (PLLA)−poly- relationships obtained from abundant experimental observa-
(ethylene oxide) (PEO) diblock copolymer was designed and tions. Unlike molecular models, which are capable of presenting
synthesized.266 PLLA has a relatively higher melting temper- many details of nucleation, macroscopic models are disadvan-
ature and crystallizes first at 90−130 °C. The different variations taged by the absence of molecular details, and consequently, the
of crystallization temperature will cause different lamellar macroscale continuum model requires initial assumptions as
thicknesses. After crystallization, the contraction of PLLA input to build the quantitative relationship between flow
makes the PEO chains deviate from random-coil state. An strength and resultant nucleation features like density and
1873 DOI: 10.1021/acs.chemrev.7b00500
Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

Figure 44. (a) SAXS intensity during crystallization of PEO. The temperature indicates the crystallization temperature of PLLA. (b) Fitted induction
time. Reprinted with permission from ref 266. Copyright 2011 American Chemical Society.

Figure 45. Model for growth of nuclei. Reprinted with permission from ref 275. Copyright 2009 Wiley−VCH.

orientation. A detailed comparison between different model formation of fibrillar nuclei, the detailed longitudinal growth
methods and recent progress of coarse-grain kinetic Monte rate, and the influence of generated nuclei on rheology. Zuidema
Carlo and molecular dynamics models can be found in the et al.274 first proposed that the recoverable strain, that is, elastic
recent reviews of Graham13 and Rutledge.276 deformation, dominates FIN. Later, Steenbakkers and Peters280
In this section, macroscale continuum models will be mainly replaced the recoverable strain by molecular stretch, and
discussed. Besides the potential application to practical Custodio et al.275 applied it on injection molding processes. It
processes, the macroscopic model can directly use experimental has been demonstrated that the high molecular weight (HMW)
results to test and to improve mechanism understanding, in the tail dominates FIN; the part controlling molecular stretch
form of assumptions. Also note that only modeling of FIN is should be the HWM tail. Thus, flow-enhanced nucleation rate
introduced; other models describing the subsequent crystal- was modeled to be fourth-order-dependent on the molecular
lization process, including crystallite growth, are outside the stretch ΛHMW of the slowest relaxation mode, Ṅ f = gn(ΛHMW4 −
scope of this work, and readers are referred to ref 277. 1), with a scaling parameter gn dependent on temperature. Next,
Under severe flow strength, the formed nuclei have a fibrillar when the molecular stretch of HMW tail surpasses a threshold,
geometry, where the length is much larger than the lateral the formed nuclei change from pointlike to fibrillar, resulting in
diameter. It was suggested that formation of fibrillar nuclei may the morphology transition from spherulite to oriented
start with pointlike objects and subsequently grow in the crystallites, as shown in Figure 45.
longitudinal direction to form the ultimate fibrillar shape. After formation, the fibrillar nuclei grow longitudinally. The
The first model aspect concerns what determines the HMW tail plays a crucial role in triggering the formation of
appearance of fibrillar nuclei in a quantitative manner. The fibrillar nuclei, but Kimata et al.147 found that shish structure has
quantity of formed fibrillar nuclei is often macroscopically a comparable concentration of HMW parts as the rest of
indicated by the sharp morphology transition from fine-grained polymer melt. This implies that the development of fibrillar
layer to oriented layer observed with a polarized microscope. nucleithat is, L̇ , the growth along longitudinal direction
Janeschitz-Kriegl and co-workers90,278 found that a parameter should be participated in by all chains in the melt. Thus, the
(the product of the fourth power of shear rate and the square of longitudinal growth rate of fibrillar nuclei was determined by the
shear time) remains approximately constant at the boundary average stretch of all chains and written as L̇ = gLJ2(Be,avgd),
between fine-grained layer and oriented skin layer in a where J2(Be,avgd) is the second invariant of the deviatoric elastic
rectangular shear slit. This parameter was linked to the density Finger tensor of a mode representing whole-chain-length
of flow-induced fibrillar nuclei, and the developed model distribution of the melt, indicating the combined effect of
interpreted such dependence as meaning the creation rate of molecular stretch and orientation, and gL is a scaling factor. In
nuclei and their longitudinal growth rate are both proportional addition, the fibrillar nuclei and the caps can grow in the radial
to the square of shear rate. However, this empirical correlation is direction at a lamellar growth rate G; see schematic illustration
restricted to specific material (molecular weight and molecular in Figure 46.
weight distribution). The model needs to relate nucleation with Another important aspect of this model is inclusion of the link
deformation on the molecular level in order to be more generally between flow-induced nuclei and rheological properties of the
applicable to other materials and flow conditions, as McHugh et corresponding melt, which often was not considered by the
al.279 suggested that molecular strain due to flow may be the simulation of the length and time scales of individual molecules.
dominant process in FIN. FIP or nuclei are usually envisioned as clusters of chain
Systematic work has been done by Peters and co-workers on segments, in which the mobility of chain segments is lower than
macroscale continuum models to describe the onset of that of free chains in the melt due to dense packing effects; as a
1874 DOI: 10.1021/acs.chemrev.7b00500
Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

consequence, the flow-induced nuclei can act as physical cross- nonisothermal protocols, respectively. A finite element code
linking points. Zuidema et al.274 proposed that the relaxation was utilized to implement the model, and a decoupled approach
times of HMW chains in the flow melt depend on the number of was employed to compute the flow kinematics. From the growth
flow-induced nuclei in a linear function. rate of fibrillar nuclei, Eder rate equations278 were employed to
calculate the total undisturbed length, surface, and volume of
oriented crystallites per unit volume. Figure 47 compared the
modeling and experimental results for isothermal crystalliza-
tions. The features of flow-induced fibrillar nuclei, in terms of (i)
total length per volume, (ii) total number density, and (iii)
average length, are predicted along the slit thickness direction,
where various locations along the thickness direction experi-
enced distinct flow strengths. Experiment MPR1 serves as the
Figure 46. Growth of isotropic and oriented crystalline structures.
reference experiment, and its microscopic images show three
Reprinted with permission from ref 275. Copyright 2009 Wiley−VCH.
distinct regions of spherulitic core layer, fine-grained layer, and
oriented layer, moving from the slit center to the wall. These two
This model was then applied to two sets of well-defined transition locations along the thickness direction are used in the
injection molding experiments employing isothermal and predicted curves to determine critical points modeled for flow

Figure 47. Distribution across half the normalized slit thickness of (i) total length of fibrillar nuclei, (ii) number of flow-induced nuclei per volume, and
(iii) average length of fibrillar nuclei for flow conditions MPR1 and MPR 2. Reprinted with permission from ref 275. Copyright 2009 Wiley−VCH.

1875 DOI: 10.1021/acs.chemrev.7b00500


Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

Figure 48. Distribution across half the normalized slit thickness of (i) total length of fibrillar nuclei, (ii) number of flow-induced nuclei per volume, and
(iii) average length of fibrillar nuclei for flow conditions MPR3 and MPR4. Reprinted with permission from ref 275. Copyright 2009 Wiley−VCH.

condition MPR1 (shown by ●, which change to ○ under other interestingly shows that the values of Vori at the transition from
conditions). These critical values obtained from experimental fine-grained layer to oriented layer for MPR1 and MPR2 are
condition MPR1 are further compared with those values of both around 0.025.
transitions under other flow conditions, to test whether MPR1 The model was also applied to nonisothermal experiments,
modeled criteria are applicable to other flow conditions. Figure where the criterion based on the relative volume of oriented
47 right column (MPR2) clearly shows that the critical values crystals shows good agreement between predictions and
predicted from MPR1 (○) are almost on top of the prediction experimental observations (see Figure 22 in ref 275). The
curve for flow condition MPR2, indicating that the morpho- agreement between modeling predictions and experimental
logical criteria modeled are applicable to both MPR1 and MPR2 results shows that the molecular assumptions about criteria for
conditions. the formation of fibrillar nuclei, longitudinal growth rate, and
Further applied to conditions MPR3 and MPR4, the model correlation between nuclei and melt viscosity, are reasonable for
predicted a lower value of total length per volume of fibrillar these flow conditions. However, the authors also point out that
nuclei than the MPR1 criteria; see Figure 48. This agrees well the predicted critical average length of fibrillar nuclei is up to 1
with the fact that no oriented layer is observed under these two mm, which seems too high and needs more experiments to
flow conditions, implying that a critical total length of fibrillar calibrate the model parameters.275
nuclei is required to cause a detectable oriented layer. In Meanwhile, alternative insights are given into the formation
addition, the relative volume fraction of oriented crystals can be mechanism of FIN. Van Erp et al.281 studied the combined effect
obtained as Vori = ψ0/(ϕ0 + ψ0), as shown in Figure 49, which of shear and pressure on FIN in iPP. Among flow conditions,
1876 DOI: 10.1021/acs.chemrev.7b00500
Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

Figure 49. Distribution of relative volume of oriented material, Vori = ψ0/(ϕ0 + ψ0), across half the normalized slit thickness under flow conditions (a)
MPR1, (b) MPR2, (c) MPR3, and (d) MPR4. Reprinted with permission from ref 275. Copyright 2009 Wiley−VCH.

Figure 50. (a) Molecular and (b) cumulative molecular stretch over time (shear time ts = 1 s). Reprinted with permission from ref 281. Copyright 2013
Wiley−VCH.

two typical cases are 180 s−1 at 500 bar and 100 s−1 at 1200 bar, determine the formation of fibrillar nuclei under the combined
respectively causing distinct isotropic and oriented morpholo- influences of flow and pressure.
gies. Again the molecular stretch during flow is calculated and The experimental results used to compare with modeling
presented in Figure 50a, which shows a relatively larger ultimate results are the final crystallite morphologies ex situ examined
molecular stretch for 100 s−1 at 1200 bar than for 180 s−1 at 500 after complete solidification. Recently, Ma et al.98 utilized a
bar. According to the aforementioned model using the critical modified multipass rheometer and an advanced ultrafast
molecular stretch as a criterion, fibrillar nuclei are more likely to synchrotron X-ray method to successfully track time evolution
be generated in 100 s−1 at 1200 bar. However, the experimental of rheological properties during severe flows close to the
result displayed that fibrillar nuclei appear in 180 s−1 at 500 bar. processing strength and lasting only 0.2−0.25 s. A significant rise
in melt viscosity was found during the severe flow pluses, instead
To understand this deviation, van Erp et al.281 compared the
of the steady-state plateau. On the basis of these experimental
time integrals over the stretch history during the shear time
results, Roozemond et al.282 modeled the FIN at such high shear
∫ t0sΛHMW dt between these two conditions (see Figure 50b). rates and found that, when describing the dependence of
Interestingly, it was found that the accumulative molecular nucleation rate on molecular stretch of the HMW tail, an
stretch of 180 s−1 at 500 bar is always larger than that of 100 s−1 exponential relationship works better than the power law used
at 1200 bar, which is consistent with experimental results that before by Zuidema et al.274 and van Erp et al.;281 see the
fibrillar nuclei are observed under the former flow conditions. comparison of various models in Figure 51.
This indicates that, different from the critical value of the Actually, Pantani et al.283 also proposed a molecular stretch
transient molecular stretch, the accumulative molecular model to describe the transition from spherulitic to fibrillar
stretchthat is, the integral of transient molecular stretch morphology. By employing the elastic dumbbell model, the
over flow timeseems to be the controlling parameter to constitutive equation was written as
1877 DOI: 10.1021/acs.chemrev.7b00500
Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

Figure 51. (a) Pressure drop and (b) apparent crystallinity during and after flow with piston of 100 mm/s for 0.2 s. Symbols show experimental data;
solid lines show the model of Roozemond et al.282 (indicated as current model); and dashed lines show the model of Zuidema et al.274 Reprinted with
permission from ref 282. Copyright 2015 AIP Publishing LLC.

D 1 molecular mechanism of FIN should incorporate structural


A̿ − (∇ν ̅ )T A̿ − A̿ (∇ν ̅ ) = − A̿ + (∇ν ̅ ) + (∇ν ̅ )T
Dt λ intermediates.
(51) (2) Ordering parameter of precursors: Fast X-ray scattering
studies have demonstrated the formation of precursors during
where A̿ is the deformation tensor with respect to equilibrium flow, which have a large influence on nucleation, crystallization
state, ν̅ is the velocity field, and λ is the relaxation time. If a flow kinetics, and final crystal morphology. However, the definition
at the constant shear rate (γ̇) is applied to melt at the reference of precursors is still obscure due to the lack of a precise ordering
time and the relaxation time λ is considered as constant during parameter. Although relaxation time was proposed as an
flow, then the deformation tensor A̿ at time t can be written as ordering parameter of precursors, this parameter lacks structural

ÄÅ É
t yÑÑÑ
follows:
ÅÅ
information and depends on temperature. Finding a suitable
i t yi
A11 = 2λ 2γ 2̇ ÅÅÅ1 − expjjj− zzzjjj1 + zzzÑÑÑ
ordering parameter of precursors is essential to construct the
ÅÅÇ k λ {k λ {ÑÑÖ
molecular mechanism of FIN and to understand how precursors
(52) influence the nucleation process.
ÄÅ É
ÅÅ i t yÑÑÑ
A12 = λγÅÅÅ̇ 1 − expjjj− zzzÑÑÑ
(3) Nonequilibrium nature of FIN: Current theoretical

ÅÅÇ k λ {ÑÑÖ
considerations of FIN are based on the framework of classical
two-phase nucleation model. The effect of flow is incorporated
(53)
into nucleation driving force through a simple add-up method.
The difference between the two main eigenvalues of the However, nucleation under strong flow is a typical non-
deformation tensor can be taken as a measurement of the equilibrium phase transition, which is not a simple extension
elongation of the system. For simple shear, the system of equilibrium phase transition and cannot be described by small
modifications of the equilibrium case. On the other hand, the
elongation becomes Δ = (A112 + 4A12 2 )1/2 . When the Debor-
intrinsic nature of multiple lengths and multiple relaxation
ah number (defined as the ratio between relaxation time and
modes makes polymers a rather suitable model system to study
shear time) is much larger than 1, the molecular stretch reaches
nonequilibrium phase transition. Meanwhile, flow is an effective
its maximum value, Δ = 2Wi[(1 + Wi 2)1/2 ], where Wi = λγ̇. trigger to reveal the nonequilibrium nature of nucleation.
Then, to form fibrillar morphology, molecular stretch Δ needs to Therefore, exploring the nonequilibrium nature of flow-induced
reach a certain critical value of Δc. Thus, the relaxation time (λ) polymer nucleation not only helps to underscore the molecular
is considered as a fitting parameter to reproduce the measured mechanism of FIN but also enriches the study of nonequilibrium
boundary of shish formation in shear rate−shear time space. science.
(4) Diversity in mechanism of FIN: Concerning FIN, the
7. FINAL REMARKS general effects revealed for all polymers include raised
After decades of effort, a wealth of phenomenological nucleation density, oriented nuclei, and even altered crystallite
accumulations and substantial progress in theory and modeling modifications. However, the mechanism of FIN may vary greatly
have been achieved in FIN. However, there are still many in different polymers, or even for the identical polymer with
challenges in unveiling the multiscale and multistep process of changes in experimental conditions like temperature and flow
FIN, which are summarized as follows. strength. It seems difficult, at least presently, to build one
(1) Molecular mechanism of FIN: The long-chain nature of universal mechanism to describe FIN for all polymers.
polymers endows them with multiscale structure and relaxation Therefore, molecular structures and flow conditions should be
time. During nucleation induced by flow, the anisotropic long studied in a systematic manner, rather than under a single
chain assembles into hierarchical structures, where the multi- condition, to identify all similarities and differences to
scale ordering involves via both intra- and interchain ordering. comprehensively understand the FIN mechanism. Moreover,
As discussed in section 3, those ordering processes may occur experimental conditions should be well-defined to precisely
sequentially, leading to appearance of structural intermediates specify the mechanism under certain limited molecular and
like precursors. The occurrence of precursor challenges the experimental conditions as well as to avoid occurrence of
traditional two-phase nucleation model, and the future unintended effects like inhomogeneous flow and melt fracture.
1878 DOI: 10.1021/acs.chemrev.7b00500
Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

(5) Developing new and powerful experimental and modeling Zhe Ma received a B.S. in polymer science and technology from
approaches: Nucleation of polymer under flow is a multiscale Zhejiang University of Technology in 2005 and obtained his M.S. in
and multistep process. The spatial uniformity and fast dynamics polymer material processing from Sichuan University in 2008. Later, he
of this process require robust detection techniques with joined Eindhoven University of Technology, where he received a Ph.D.
sufficiently high space and time resolution. As the development in polymer materials in 2012 and worked as a postdoctoral researcher
of synchrotron radiation-based X-ray scattering benefits the from 2012 to 2015. He is currently working as an associate professor of
understanding of structural evolution during crystallization, polymer science and engineering at Tianjin University. His research
innovative techniques with ultrahigh time and space resolution focuses on the crystallization behaviors of polyolefins and their
can shed light on the complex nucleation process and help to correlations with molecular structures.
construct the molecular mechanism. Meanwhile, multiscale Nan Tian received a B.S. in polymer materials and engineering from the
modeling approaches are also expected to give verification and University of Science and Technology of China in 2009 and obtained a
new understanding of nucleation at the molecular level and Ph.D. in nuclear science and technology from the University of Science
bridge gaps between different detection techniques in both time and Technology of China in 2014. Since then, he has worked as an
and length scales. associate professor of polymer chemistry and physics at Northwestern
(6) FIN in real polymer processing: Different from most FIN Polytechnical University. His research focuses on the relation between
studies that were performed under relatively mild conditions in entanglement, chain dynamics, and crystallization of polymers.
the lab, real polymer processes have much stronger flow strength
and more complex flow conditions. For instance, in film blowing Dong Liu received a B.S. in chemistry from the University of Science
and casting, flow gradients are two-dimensional and exist along and Technology of China in 2009. Later, he joined Professor Liangbin
both stretching and lateral directions. More attention should be Li’s Soft Matter Group at National Synchrotron Radiation Laboratory,
paid to FIN events under complex real processing conditions. In University of Science and Technology of China, where he received a
the meantime, based on these obtained experimental results, the Ph.D. in nuclear science and technology in 2015. He is currently
mesoscale simulation and modeling need to be developed to working as a research assistant in neutron scattering, based at the China
gain understanding of mechanism and to also validate/modify Mianyang Research Reactor (CMRR) at Key Laboratory of Neutron
the current mechanism. Physics and Institute of Nuclear Physics and Chemistry, China
In summary, FIN is a very fundamental and industrially core Academy of Engineering Physics (CAEP). His research focuses on
subject. The intrinsic multiscale features of polymer chain lead structure analysis in polymer science with the help of neutron and X-ray
to multiscale ordered structures and multistep ordering process scattering.
for nucleation. The multiscale structures include segmental Fengmei Su received a B.S. (2011) in polymer material and engineering
conformation, packing of conformational ordering, deformation from Sichuan University and obtained a Ph.D. (2016) in nuclear
on the whole-chain scale, and macroscopic aggregation of science and technology from the University of Science and Technology
crystallites. The multistep process involves conformation of China. She is currently a postdoctoral fellow in the Synchrotron
transition, isotropic−nematic transition, density fluctuation Radiation Laboratory, University of Science and Technology of China.
(or phase separation), formation of precursors, and shish- Her primary research interest is flow-induced polymer crystallization,
kebab crystallites. The molecular mechanism of FIN, incorpo- such as conformation ordering and flow-induced precursors, etc.
rating both multiscale and multistep considerations, is in urgent Liangbin Li received a B.S. (1994) from Sichuan Normal University and
demand. A full understanding of nucleation under flow will M.S. (1997) from Sichuan University in physics and a Ph.D. (2000) in
ultimately benefit the precise control of polymer products and polymer material processing from Sichuan University. From 2000 to
the understanding of nonequilibrium phase transition. 2004, he was a postdoctoral fellow in FOM-Institute for Atomic and
Molecular Physics and Technology University of Delft, The Nether-
AUTHOR INFORMATION lands. From 2004 to 2006, he worked as a materials scientist at Unilever
Food and Health Research Institute. Since 2006, under the One-
Corresponding Authors
Hundred Talent Program of Chinese Academy of Science, Dr. Li joined
*E-mail zhe.ma@tju.edu.cn (Z.M.). National Synchrotron Radiation Laboratory, University of Science and
*E-mail lbli@ustc.edu.cn (L.L.). Technology of China as a full professor and started the Soft Matter
ORCID Group. His primary research interests are polymer physics relevant to
processing, such as flow-induced crystallization of polymer and stress-
Nan Tian: 0000-0003-1822-876X induced deformation and phase transitions of crystalline polymers.
Liangbin Li: 0000-0002-1887-9856
Notes ACKNOWLEDGMENTS
The authors declare no competing financial interest. This work was supported by the National Natural Science
Biographies Foundation of China (51325301, 51633009, 51573132, and
51227801) and the key research and development tasks of
Kunpeng Cui received a B.S. in material science and engineering from MOST (2016YFB0302501).
Zhengzhou University in 2010. Then he joined the research group of
Professor Liangbin Li at the University of Science and Technology of REFERENCES
China, where he investigated flow-induced polymer crystallization and
(1) Keller, A.; Kolnaar, J. W. H. Chain Extension and Orientation:
received a Ph.D. in 2015. He is currently working as a JSPS (Japan Fundamentals and Relevance to Processing and Products. In Orienta-
Society for the Promotion of Science) research fellow at Hokkaido tional Phenomena in Polymers; Steinkopff: 1993; pp 81−102; DOI:
University. His research interests include polymer crystallization, 10.1007/BFb0115440.
toughening mechanism of hydrogels, and synchrotron-based X-ray (2) Liu, X.; Dai, K.; Hao, X.; Zheng, G.; Liu, C.; Schubert, D. W.; Shen,
scattering techniques. C. Crystalline Structure of Injection Molded β-Isotactic Polypropylene:

1879 DOI: 10.1021/acs.chemrev.7b00500


Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

Analysis of the Oriented Shear Zone. Ind. Eng. Chem. Res. 2013, 52, (25) Hikichi, K.; Furuichi, J. Molecular Motions of Polymers Having
11996−12002. Helical Conformation. I. Poly(ethylene Glycol) and Polyoxymethylene.
(3) Liu, Q.; Sun, X.; Li, H.; Yan, S. Orientation-Induced J. Polym. Sci., Part A: Gen. Pap. 1965, 3, 3003−3013.
Crystallization of Isotactic Polypropylene. Polymer 2013, 54, 4404− (26) Tosaka, M. A Route for the Thermodynamic Description of
4421. Strain-Induced Crystallization in Sulfur-Cured Natural Rubber.
(4) Samon, J. M.; Schultz, J. M.; Hsiao, B. S.; Seifert, S.; Stribeck, N.; Macromolecules 2009, 42, 6166−6174.
Gurke, I.; Saw, G. C. Structure Development during the Melt Spinning (27) Balzano, L.; Rastogi, S.; Peters, G. Self-Nucleation of Polymers
of Polyethylene and Poly(vinylidene Fluoride) Fibers by in Situ with Flow: The Case of Bimodal Polyethylene. Macromolecules 2011,
Synchrotron Small- and Wide-Angle X-Ray Scattering Techniques. 44, 2926−2933.
Macromolecules 1999, 32, 8121−8132. (28) Auriemma, F.; De Rosa, C.; Corradini, P. Solid Mesophases in
(5) Doufas, A. K.; McHugh, A. J.; Miller, C. Simulation of Melt Semicrystalline Polymers: Structural Analysis by Diffraction Techni-
Spinning Including Flow-Induced Crystallization - Part I. Model ques. Adv. Polym. Sci. 2005, 181, 1−74.
Development and Predictions. J. Non-Newtonian Fluid Mech. 2000, 92, (29) Bermejo, F. J.; Criado, A.; Fayos, R.; FernandezPerea, R.; Fischer,
27−66. H. E.; Suard, E.; Guelylah, A.; Zuniga, J. Structural Correlations in
(6) Lin, Y.; Yao, Y.; Yang, X.; Shen, L.; Li, R.; Wu, D. Effect of Gas Disordered Matter: An Experimental Separation of Orientational and
Flow Rate on Crystal Structures of Electrospun and Gas-Jet/ Positional Contributions. Phys. Rev. B: Condens. Matter Mater. Phys.
Electrospun Poly(vinylidene Fluoride) Fibers. Chin. J. Polym. Sci. 1997, 56, 11536−11545.
2009, 27, 511−516. (30) Cui, K.; Liu, D.; Ji, Y.; Huang, N.; Ma, Z.; Wang, Z.; Lv, F.; Yang,
(7) Xu, Z.; Su, L.; Wang, P.; Peng, M. Effect of Oscillatory Shear on the H.; Li, L. Nonequilibrium Nature of Flow-Induced Nucleation in
Mechanical Properties and Crystalline Morphology of Linear Low Isotactic Polypropylene. Macromolecules 2015, 48, 694−699.
Density Polyethylene. Chin. J. Polym. Sci. 2015, 33, 1114−1124. (31) Kaji, K.; Nishida, K.; Kanaya, T.; Matsuba, G.; Konishi, T.; Imai,
(8) Zheng, G.-Q.; Yang, W.; Huang, L.; Yang, M.-B.; Li, W.; Liu, C.-T.; M. Spinodal Crystallization of Polymers: Crystallization from the
Shen, C.-Y. Flow-Induced Fiber Orientation in Gas-Assisted Injection Unstable Melt. Interphases Mesophases Polym. Cryst. III 2005, 191, 187−
Molded Part. Mater. Lett. 2007, 61, 3436−3439. 240.
(9) Qin, J.; Milner, S. T. Tube Diameter of Oriented and Stretched (32) Zhang, W.; Gomez, E. D.; Milner, S. T. Predicting Nematic
Polymer Melts. Macromolecules 2013, 46, 1659−1672. Phases of Semiflexible Polymers. Macromolecules 2015, 48, 1454−1462.
(10) Li, J.; Nie, Y.; Ma, Y.; Hu, W. Stress-Induced Polymer (33) Strobl, G. Crystallization and Melting of Bulk Polymers: New
Deformation in Shear Flows. Chin. J. Polym. Sci. 2013, 31, 1590−1598. Observations, Conclusions and a Thermodynamic Scheme. Prog.
(11) Somani, R. H.; Yang, L.; Zhu, L.; Hsiao, B. S. Flow-Induced Polym. Sci. 2006, 31, 398−442.
Shish-Kebab Precursor Structures in Entangled Polymer Melts. Polymer (34) Heeley, E. L.; Poh, C. K.; Li, W.; Maidens, A.; Bras, W.; Dolbnya,
2005, 46, 8587−8623.
I. P.; Gleeson, A. J.; Terrill, N. J.; Fairclough, J. P. A.; Olmsted, P. D.;
(12) Lamberti, G. Flow Induced Crystallisation of Polymers. Chem.
et al. Are Metastable, Precrystallisation, Density-Fluctuations a
Soc. Rev. 2014, 43, 2240−2252.
Universal Phenomena? Faraday Discuss. 2003, 122, 343−361.
(13) Graham, R. S. Modelling Flow-Induced Crystallisation in
(35) Olmsted, P. D.; Poon, W. C. K.; McLeish, T. C. B.; Terrill, N. J.;
Polymers. Chem. Commun. 2014, 50, 3531−3545.
Ryan, A. J. Spinodal-Assisted Crystallization in Polymer Melts. Phys.
(14) Zhou, M.; Xu, S.; Li, Y.; He, C.; Jin, T.; Wang, K.; Deng, H.;
Zhang, Q.; Chen, F.; Fu, Q. Transcrystalline Formation and Properties Rev. Lett. 1998, 81, 373−376.
(36) Tamashiro, M. N.; Pincus, P. Helix-Coil Transition in
of Polypropylene on the Surface of Ramie Fiber as Induced by Shear or
Dopamine Modification. Polymer 2014, 55, 3045−3053. Homopolypeptides under Stretching. Phys. Rev. E: Stat. Phys., Plasmas,
(15) Hamad, F. G.; Colby, R. H.; Milner, S. T. Onset of Flow-Induced Fluids, Relat. Interdiscip. Top. 2001, 63, No. 021909.
Crystallization Kinetics of Highly Isotactic Polypropylene. Macro- (37) Wunderlich, B.; Grebowicz, J. Thermotropic Mesophases and
molecules 2015, 48, 3725−3738. Mesophase Transitions of Linear, Flexible Macromolecules. Adv.
(16) Peters, G. W. M.; Balzano, L.; Steenbakkers, R. J. A. Flow- Polym. Sci. 1984, 60-61, 1−59.
Induced Crystallization. In Handbook of Polymer Crystallization; John (38) Wunderlich, B.; Möller, M.; Grebowicz, J.; Baur, H. Conforma-
Wiley & Sons Inc.: Hoboken, NJ, 2013; pp 399−431; DOI: 10.1002/ tional Motion and Disorder in Low and High Molecular Mass Crystals.
9781118541838.ch14. Adv. Polym. Sci. 1988, 87, 1−121.
(17) Hamad, F. G.; Colby, R. H.; Milner, S. T. Lifetime of Flow- (39) Chou, H. P.; Spence, C.; Scherer, A.; Quake, S. A Microfabricated
Induced Precursors in Isotactic Polypropylene. Macromolecules 2015, Device for Sizing and Sorting DNA Molecules. Proc. Natl. Acad. Sci. U.
48, 7286−7299. S. A. 1999, 96, 11−13.
(18) Mackley, M. R.; Keller, A. Flow Induced Crystallization of (40) Sheils, C. A.; Käs, J.; Travassos, W.; Allen, P. G.; Janmey, P. A;
Polyethylene Melts. Polymer 1973, 14, 16−20. Wohl, M. E.; Stossel, T. P. Actin Filaments Mediate DNA Fiber
(19) Wang, Z.; Ma, Z.; Li, L. Flow-Induced Crystallization of Formation in Chronic Inflammatory Airway Disease. Am. J. Pathol.
Polymers: Molecular and Thermodynamic Considerations. Macro- 1996, 148, 919−927.
molecules 2016, 49, 1505−1517. (41) Courty, S.; Gornall, J. L.; Terentjev, E. M. Induced Helicity in
(20) Pigeon, M.; Prud'homme, R. E.; Pezolet, M. Characterization of Biopolymer Networks under Stress. Proc. Natl. Acad. Sci. U. S. A. 2005,
Molecular-Orientation in Polyethylene by Raman-Spectroscopy. 102, 13457−13460.
Macromolecules 1991, 24, 5687−5694. (42) Zimm, B. H.; Bragg, J. K. Theory of the One-Dimensional Phase
(21) Ishihara, N.; Seimiya, T.; Kuramoto, M.; Uoi, M. Crystalline Transition in Polypeptide Chains. J. Chem. Phys. 1958, 28, 1246−2147.
Syndiotactic Polystyrene. Macromolecules 1986, 19, 2464−2465. (43) Zimm, B. H.; Bragg, J. K. Theory of the Phase Transition between
(22) Ebewele, R. O. Polymer Science and Technology; CRC Press: Boca Helix and Random Coil in Polypeptide Chains. J. Chem. Phys. 1959, 31,
Raton, FL, 2000. 526−535.
(23) Tadokoro, H.; Kobayashi, M.; Ukita, M.; Yasufuku, K.; (44) Buhot, A.; Halperin, A. Extension of Rod-Coil Multiblock
Murahashi, S.; Torii, T. Normal Vibrations of the Polymer Molecules Copolymers and the Effect of the Helix-Coil Transition. Phys. Rev. Lett.
of Helical Conformation. V. Isotactic Polypropylene and Its 2000, 84, 2160−2163.
Deuteroderivatives. J. Chem. Phys. 1965, 42, 1432−1449. (45) Buhot, A.; Halperin, A. Extension Behavior of Helicogenic
(24) Matsuba, G.; Kaji, K.; Nishida, K.; Kanaya, T.; Imai, M. Polypeptides. Macromolecules 2002, 35, 3238−3252.
Conformational Change and Orientation Fluctuations prior to the (46) Courty, S.; Gornall, J. L.; Terentjev, E. M. Mechanically Induced
Crystallization of Syndiotactic Polystyrene. Macromolecules 1999, 32, Helix-Coil Transition in Biopolymer Networks. Biophys. J. 2006, 90,
8932−8937. 1019−1027.

1880 DOI: 10.1021/acs.chemrev.7b00500


Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

(47) Kutter, S.; Terentjev, E. M. Networks of Helix-Forming Melting Temperature by Microbeam X-Ray Scattering. Macromolecules
Polymers. Eur. Phys. J. E 2002, 8, 539−547. 2013, 46, 3031−3036.
(48) Snyder, R. G.; Schachtschneider, J. H. Valence Force Calculation (69) Zhao, Y.; Matsuba, G.; Moriwaki, T.; Ikemoto, Y.; Ito, H. Shear-
of the Vibrational Spectra of Crystalline Isotactic Polypropylene and Induced Conformational Fluctuations of Polystyrene Probed by 2D
Some Deuterated Polypropylenes. Spectroehim. Acta 1964, 20, 853− Infrared Microspectroscopy. Polymer 2012, 53, 4855−4860.
869. (70) Deng, C.; Fujiwara, T.; Polec, I.; Matsuba, G.; Jin, L.; Inoue, R.;
(49) Zhu, X. Y.; Yan, D. Y.; Fang, Y. P. In Situ FTIR Spectroscopic Nishida, K.; Kanaya, T. Precursor of Shish-Kebab in Atactic
Study of the Conformational Change of Isotactic Polypropylene during Polystyrene/Isotactic Polystyrene Blend above Nominal Melting
the Crystallization Process. J. Phys. Chem. B 2001, 105, 12461−12463. Temperature. Macromolecules 2012, 45, 4630−4637.
(50) An, H. N.; Zhao, B. J.; Ma, Z.; Shao, C. G.; Wang, X.; Fang, Y. P.; (71) Hayashi, Y.; Matsuba, G.; Zhao, Y. F.; Nishida, K.; Kanaya, T.
Li, L. B.; Li, Z. M. Shear-Induced Conformational Ordering in the Melt Precursor of Shish-Kebab in Isotactic Polystyrene under Shear Flow.
of Isotactic Polypropylene. Macromolecules 2007, 40, 4740−4743. Polymer 2009, 50, 2095−2103.
(51) An, H. N.; Li, X. Y.; Geng, Y.; Wang, Y. L.; Wang, X.; Li, L. B.; Li, (72) Azzurri, F.; Alfonso, G. C. Lifetime of Shear-Induced Crystal
Z. M.; Yang, C. L. Shear-Induced Conformational Ordering, Nucleation Precursors. Macromolecules 2005, 38, 1723−1728.
Relaxation, and Crystallization of Isotactic Polypropylene. J. Phys. (73) Ziabicki, A.; Alfonso, G. C. A Simple Model of Flow-Induced
Chem. B 2008, 112, 12256−12262. Crystallization Memory. Macromol. Symp. 2002, 185, 211−231.
(52) Budevska, B. O.; Manning, C. J.; Griffiths, P. R.; Roginski, R. T. (74) Somani, R. H.; Yang, L.; Hsiao, B. S.; Fruitwala, H. Nature of
Step-Scan Fourier Transform Infrared Study on the Effect of Dynamic Shear-Induced Primary Nuclei in iPP Melt. J. Macromol. Sci., Part B:
Strain on Isotactic Polypropylene. Appl. Spectrosc. 1993, 47, 1843− Phys. 2003, 42, 515−531.
1851. (75) Ran, S. F.; Burger, C.; Fang, D. F.; Zong, X. H.; Cruz, S.; Chu, B.;
(53) Geng, Y.; Wang, G. L.; Cong, Y. H.; Bai, L. G.; Li, L. B.; Yang, C. Hsiao, B. S.; Bubeck, R. A.; Yabuki, K.; Teramoto, Y.; et al. In-Situ
L. Shear-Induced Nucleation and Growth of Long Helices in Synchrotron WAXD/SAXS Studies of Structural Development during
Supercooled Isotactic Polypropylene. Macromolecules 2009, 42, PBO/PPA Solution Spinning. Macromolecules 2002, 35, 433−439.
4751−4757. (76) Marconi, U.; Puglisi, A.; Rondoni, L.; Vulpiani, A. Fluctuation−
(54) Su, F.; Ji, Y.; Meng, L.; Wang, Z.; Qi, Z.; Chang, J.; Ju, J.; Li, L. dissipation: Response Theory in Statistical Physics. Phys. Rep. 2008,
Coupling of Multiscale Orderings during Flow-Induced Crystallization 461, 111−195.
of Isotactic Polypropylene. Macromolecules 2017, 50, 1991−1997. (77) Jarzynski, C. Nonequilibrium Equality for Free Energy
(55) Yamamoto, T. Molecular Dynamics of Crystallization in a Helical Differences. Phys. Rev. Lett. 1997, 78, 2690−2693.
Polymer Isotactic Polypropylene from the Oriented Amorphous State. (78) Lenstra, T. A. J.; Dogic, Z.; Dhont, J. K. G. Shear-Induced
Macromolecules 2014, 47, 3192−3202. Displacement of Isotropic-Nematic Spinodals. J. Chem. Phys. 2001, 114,
(56) Stephens, J. S.; Chase, D. B.; Rabolt, J. F. Effect of the 10151−10162.
Electrospinning Process on Polymer Crystallization Chain Conforma- (79) Turnbull, D.; Fisher, J. C. Rate of Nucleation in Condensed
tion in Nylon-6 and Nylon-12. Macromolecules 2004, 37, 877−881. Systems. J. Chem. Phys. 1949, 17, 71−73.
(57) Ajji, A.; Guevremont, J.; Cole, K. C.; Dumoulin, M. M. (80) Chen, Y.-H.; Fang, D.-F.; Lei, J.; Li, L.-B.; Hsiao, B. S.; Li, Z.-M.
Orientation and Structure of Drawn Poly(ethylene Terephthalate). Shear-Induced Precursor Relaxation-Dependent Growth Dynamics
Polymer 1996, 37, 3707−3714. and Lamellar Orientation of β-Crystals in β-Nucleated Isotactic
(58) Chai, C. K.; Dixon, N. M.; Gerrard, D. L.; Reed, W. Rheo-Raman Polypropylene. J. Phys. Chem. B 2015, 119, 5716−5727.
Studies of Polyethylene Melts. Polymer 1995, 36, 661−663. (81) Nie, Y.; Gao, H.; Hu, W. Variable Trends of Chain-Folding in
(59) Hsiao, B.; Yang, L.; Somani, R.; Avila-Orta, C.; Zhu, L. Separate Stages of Strain-Induced Crystallization of Bulk Polymers.
Unexpected Shish-Kebab Structure in a Sheared Polyethylene Melt. Polymer 2014, 55, 1267−1272.
Phys. Rev. Lett. 2005, 94, No. 117802. (82) Kim, Y. H.; Pincus, P. Nematic Polymers - Excluded-Volume
(60) Sakellarides, S. L.; McHugh, A. J. Formation of Fibrous Crystals Effects. Biopolymers 1979, 18, 2315−2322.
in Flowing Blends of Polyethylene Melts. Rheol. Acta 1987, 26, 64−77. (83) Li, L. B.; de Jeu, W. H. Flow-Induced Mesophases in
(61) Tashiro, K.; Sasaki, S.; Kobayashi, M. Structural Investigation of Crystallizable Polymers. Adv. Polym. Sci. 2005, 181, 75−120.
Orthorhombic-to-Hexagonal Phase Transition in Polyethylene Crystal: (84) Katayama, K.; Amano, T.; Nakamura, K. Structural Formation
The Experimental Confirmation of the Conformationally Disordered during Melt Spinning Process. Colloid Polym. Sci. 1968, 226, 125−134.
Structure by X-Ray Diffraction and Infrared/Raman Spectroscopic (85) Bonart, R. Parakristalline Strukturen in Polyathylenterephthalat
Measurements. Macromolecules 1996, 29, 7460−7469. (Pet). Colloid Polym. Sci. 1966, 213, 1.
(62) Uehara, H.; Kanamoto, T.; Kawaguchi, A.; Murakami, S. Real- (86) Blundell, D. J.; MacKerron, D. H.; Fuller, W.; Mahendrasingam,
Time X-Ray Diffraction Study on Two-Stage Drawing of Ultra-High A.; Martin, C.; Oldman, R. J.; Rule, R. J.; Riekel, C. Characterization of
Molecular Weight Polyethylene Reactor Powder above the Static Strain-Induced Crystallization of Poly(ethylene Terephthalate) at Fast
Melting Temperature. Macromolecules 1996, 29, 1540−1547. Draw Rates Using Synchrotron Radiation. Polymer 1996, 37, 3303−
(63) Hikosaka, M.; Rastogi, S.; Keller, A.; Kawabata, H. Investigations 3311.
on the Crystallization of Polyethylene under High Pressure: Role of (87) Welsh, G. E.; Blundell, D. J.; Windle, A. H. A Transient
Mobile Phases, Lamellar Thickening Growth, Phase Transformations, Mesophase on Drawing Polymers Based on Polyethylene Tereph-
and Morphology. J. Macromol. Sci., Part B: Phys. 1992, 31, 87−131. thalate (PET) and Polyethylene Naphthoate (PEN). J. Mater. Sci. 2000,
(64) Meier, R. J. Studying the Length of Trans Conformational 35, 5225−5240.
Sequences in Polyethylene Using Raman Spectroscopy: A Computa- (88) Welsh, G. E.; Blundell, D. J.; Windle, A. H. A Transient Liquid
tional Study. Polymer 2002, 43, 517−522. Crystalline Phase as a Precursor for Crystallization in Random Co-
(65) Sasaki, S.; Tashiro, K.; Kobayashi, M.; Izumi, Y.; Kobayashi, K. Polyester Fibers. Macromolecules 1998, 31, 7562−7565.
Microscopically Viewed Structural Change of PE during the Isothermal (89) Matsuba, G.; Kaji, K.; Nishida, K.; Kanaya, T.; Imai, M.
Crystallization from the Melt. Polymer 1999, 40, 7125−7135. Conformational Change and Orientation Fluctuations of Isotactic
(66) Onsager, L. The Effects of Shape on the Interaction of Colloidal Polystyrene prior to Crystallization. Polym. J. 1999, 31, 722−727.
Particles. Ann. N. Y. Acad. Sci. 1949, 51, 627−659. (90) Liedauer, S.; Eder, G.; Janeschitz-Kriegl, H.; Jerschow, P.;
(67) Doi, M.; Edwards, S. F. The Theory of Polymer Dynamics; Geymayer, W.; Ingolic, E. On the Kinetics of Shear-Induced
Clarendon Press: Oxford, U.K., 1986. Crystallization in Polypropylene. Int. Polym. Process. 1993, 8, 236−244.
(68) Kanaya, T.; Polec, I. A.; Fujiwara, T.; Inoue, R.; Nishida, K.; (91) Kumaraswamy, G.; Verma, R. K.; Kornfield, J. A. Novel Flow
Matsuura, T.; Ogawa, H.; Ohta, N. Precursor of Shish-Kebab above the Apparatus for Investigating Shear-Enhanced Crystallization and

1881 DOI: 10.1021/acs.chemrev.7b00500


Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

Structure Development in Semicrystalline Polymers. Rev. Sci. Instrum. (110) Schultz, J. M.; Lin, J. S.; Hendricks, R. W.; Petermann, J.; Gohil,
1999, 70, 2097−2104. R. M. Annealing of Polypropylene Films Crystallized from a Highly
(92) Liu, Y.; Zhou, W.; Cui, K.; Tian, N.; Wang, X.; Liu, L.; Li, L.; Extended Melt. J. Polym. Sci., Polym. Phys. Ed. 1981, 19, 609−620.
Zhou, Y. Extensional Rheometer for in Situ X-Ray Scattering Study on (111) Ma, Z.; Balzano, L.; Portale, G.; Peters, G. W. M. Flow Induced
Flow-Induced Crystallization of Polymer. Rev. Sci. Instrum. 2011, 82, Crystallization in Isotactic Polypropylene during and after Flow.
045104. Polymer 2014, 55, 6140−6151.
(93) Zhang, C. G.; Hu, H. Q.; Wang, D. J.; Yan, S.; Han, C. C. In Situ (112) Wang, Z. G.; Hsiao, B. S.; Fu, B. X.; Liu, L.; Yeh, F.; Sauer, B. B.;
Optical Microscope Study of the Shear-Induced Crystallization of Chang, H.; Schultz, J. M. Correct Determination of Crystal Lamellar
Isotactic Polypropylene. Polymer 2005, 46, 8157−8161. Thickness in Semicrystalline Poly(ethylene Terephthalate) by Small-
(94) Kim, S.; Yu, J. W.; Han, C. C. Shear Light Scattering Photometer Angle X-Ray Scattering. Polymer 2000, 41, 1791−1797.
with Optical Microscope for the Study of Polymer Blends. Rev. Sci. (113) Wang, Z. G.; Hsiao, B. S.; Sirota, E. B.; Agarwal, P.; Srinivas, S.
Instrum. 1996, 67, 3940−3947. Probing the Early Stages of Melt Crystallization in Polypropylene by
(95) Seki, M.; Thurman, D. W.; Oberhauser, J. P.; Kornfield, J. A. Simultaneous Small- and Wide-Angle X-Ray Scattering and Laser Light
Shear-Mediated Crystallization of Isotactic Polypropylene: The Role of Scattering. Macromolecules 2000, 33, 978−989.
Long Chain-Long Chain Overlap. Macromolecules 2002, 35, 2583− (114) Heeley, E. L.; Maidens, A. V.; Olmsted, P. D.; Bras, W.;
2594. Dolbnya, I. P.; Fairclough, J. P. A.; Terrill, N. J.; Ryan, A. J. Early Stages
(96) Hobbs, J. K.; Humphris, A. D. L.; Miles, M. J. In-Situ Atomic of Crystallization in Isotactic Polypropylene. Macromolecules 2003, 36,
Force Microscopy of Polyethylene Crystallization. 1. Crystallization 3656−3665.
from an Oriented Backbone. Macromolecules 2001, 34, 5508−5519. (115) Pogodina, N. V.; Lavrenko, V. P.; Srinivas, S.; Winter, H. H.
(97) Cui, K.; Meng, L.; Tian, N.; Zhou, W.; Liu, Y.; Wang, Z.; He, J.; Rheology and Structure of Isotactic Polypropylene near the Gel Point:
Li, L. Self-Acceleration of Nucleation and Formation of Shish in Quiescent and Shear-Induced Crystallization. Polymer 2001, 42, 9031−
Extension-Induced Crystallization with Strain beyond Fracture. 9043.
Macromolecules 2012, 45, 5477−5486. (116) Zhang, C.; Hu, H.; Wang, X.; Yao, Y.; Dong, X.; Wang, D.;
(98) Ma, Z.; Balzano, L.; van Erp, T.; Portale, G.; Peters, G. W. M. Wang, Z.; Han, C. C. Formation of Cylindrite Structures in Shear-
Short-Term Flow Induced Crystallization in Isotactic Polypropylene: Induced Crystallization of Isotactic Polypropylene at Low Shear Rate.
How Short Is Short? Macromolecules 2013, 46, 9249−9258. Polymer 2007, 48, 1105−1115.
(99) Su, F.; Zhou, W.; Li, X.; Ji, Y.; Cui, K.; Qi, Z.; Li, L. Flow-Induced (117) Gutierrez, M. C. G.; Alfonso, G. C.; Riekel, C.; Azzurri, F.
Precursors of Isotactic Polypropylene: Anin SituTime and Space Spatially Resolved Flow-Induced Crystallization Precursors in Isotactic
Resolved Study with Synchrotron Radiation Scanning X-Ray Micro- Polystyrene by Simultaneous Small- and Wide-Angle X-Ray Micro-
diffraction. Macromolecules 2014, 47, 4408−4416. diffraction. Macromolecules 2004, 37, 478−485.
(100) Pratt, L. R.; Hsu, C. S.; Chandler, D. Statistical-Mechanics of (118) Azzurri, F.; Alfonso, G. C. Insights into Formation and
Small Chain Molecules in Liquids. I Effects of Liquid Packing on Relaxation of Shear-Induced Nucleation Precursors in Isotactic
Conformational Structures. J. Chem. Phys. 1978, 68, 4202−4212. Polystyrene. Macromolecules 2008, 41, 1377−1383.
(101) Imai, M.; Mori, K.; Mizukami, T.; Kaji, K.; Kanaya, T. Structural (119) Cavallo, D.; Azzurri, F.; Balzano, L.; Funari, S. S.; Alfonso, G. C.
Formation of Poly (Ethylene Terephthalate) during the Induction Flow Memory and Stability of Shear-Induced Nucleation Precursors in
Period of Crystallization: 1. Ordered Structure Appearing before Isotactic Polypropylene. Macromolecules 2010, 43, 9394−9400.
Crystal Nucleation. Polymer 1992, 33, 4451−4456. (120) Balzano, L.; Kukalyekar, N.; Rastogi, S.; Peters, G. W.;
(102) Matsuba, G.; Kaji, K.; Kanaya, T.; Nishida, K. Detailed Analysis Chadwick, J. Crystallization and Dissolution of Flow-Induced
of the Induction Period of Polymer Crystallization by Depolarized Precursors. Phys. Rev. Lett. 2008, 100, No. 048302.
Light Scattering. Phys. Rev. E: Stat. Phys., Plasmas, Fluids, Relat. (121) Polec, I. A.; Fujiwara, T.; Kanaya, T.; Deng, C. Simultaneous
Interdiscip. Top. 2002, 65, No. 061801. SAXS/WAXS Experiments on Shear Induced iPP Crystallization near
(103) Imai, M.; Kaji, K.; Kanaya, T.; Sakai, Y. Chain Conformation in Nominal Melting Temperature. Polymer 2012, 53, 3540−3547.
the Induction Period of Crystallization of Poly(ethylene Terephtha- (122) Somani, R. H.; Hsiao, B. S.; Nogales, A.; Srinivas, S.; Tsou, A.
late). Phys. B 1995, 213-214, 718−720. H.; Sics, I.; Balta-Calleja, F. J.; Ezquerra, T. A. Structure Development
(104) Imai, M.; Kaji, K.; Kanaya, T.; Sakai, Y. Ordering Process in the during Shear Flow-Induced Crystallization of I-PP: In-Situ Small-Angle
Induction Period of Crystallization of Poly(Ethylene-Terephthalate). X-Ray Scattering Study. Macromolecules 2000, 33, 9385−9394.
Phys. Rev. B: Condens. Matter Mater. Phys. 1995, 52, 12696−12704. (123) Liu, D.; Tian, N.; Huang, N.; Cui, K.; Wang, Z.; Hu, T.; Yang,
(105) Imai, M.; Mori, K.; Mizukami, T.; Kaji, K.; Kanaya, T. Structural H.; Li, X.; Li, L. Extension-Induced Nucleation under near-Equilibrium
Formation of Poly(ethylene Terephthalate) during the Induction Conditions: The Mechanism on the Transition from Point Nucleus to
Period of Crystallization: 2. Kinetic Analysis Based on the Theories of Shish. Macromolecules 2014, 47, 6813−6823.
Phase Separation. Polymer 1992, 33, 4457−4462. (124) Zhao, Y.; Hayasaka, K.; Matsuba, G.; Ito, H. In Situ
(106) Ryan, A. J.; Terrill, N. J.; Fairclough, J. P. A. A Scattering Study Observations of Flow-Induced Precursors during Shear Flow. Macro-
of Nucleation Phenomena in Homopolymer Melts. In Scattering from molecules 2013, 46, 172−178.
Polymers; Cebe, P., Hsiao, B. S., Lohse, D. J., Eds.; American Chemical (125) Somani, R. H.; Yang, L.; Sics, I.; Hsiao, B. S.; Pogodina, N. V.;
Society: Washington, DC, 1999; pp 201−217; DOI: 10.1021/bk-2000- Winter, H. H.; Agarwal, P.; Fruitwala, H.; Tsou, A. Orientation-Induced
0739.ch013. Crystallization in Isotactic Polypropylene Melt by Shear Deformation.
(107) Ryan, A. J.; Fairclough, J. P. A.; Terrill, N. J.; Olmsted, P. D.; Macromol. Symp. 2002, 185, 105−117.
Poon, W. C. K. A Scattering Study of Nucleation Phenomena in (126) Somani, R. H.; Yang, L.; Hsiao, B. S. Precursors of Primary
Polymer Crystallisation. Faraday Discuss. 1999, 112, 13−29. Nucleation Induced by Flow in Isotactic Polypropylene. Phys. A 2002,
(108) Terrill, N. J.; Fairclough, P. A.; Towns-Andrews, E.; 304, 145−157.
Komanschek, B. U.; Young, R. J.; Ryan, A. J. Density Fluctuations: (127) Balzano, L.; Ma, Z.; Cavallo, D.; Van Erp, T. B.; Fernandez-
The Nucleation Event in Isotactic Polypropylene Crystallization. Ballester, L.; Peters, G. W. M. Molecular Aspects of the Formation of
Polymer 1998, 39, 2381−2385. Shish-Kebab in Isotactic Polypropylene. Macromolecules 2016, 49,
(109) Cakmak, M.; Teitge, A.; Zachmann, H. G.; White, J. L. Online 3799−3809.
Small-Angle and Wide-Angle X-Ray-Scattering Studies on Melt- (128) Kornfield, J. A.; Kumaraswamy, G.; Issaian, A. M. Recent
Spinning Poly(Vinylidene Fluoride) Tape Using Synchrotron Advances in Understanding Flow Effects on Polymer Crystallization.
Radiation. J. Polym. Sci., Part B: Polym. Phys. 1993, 31, 371−381. Ind. Eng. Chem. Res. 2002, 41, 6383−6392.

1882 DOI: 10.1021/acs.chemrev.7b00500


Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

(129) Persikov, A. V.; Xu, Y. J.; Brodsky, B. Equilibrium Thermal (149) Janeschitz-Kriegl, H.; Ratajski, E.; Stadlbauer, M. Flow as an
Transitions of Collagen Model Peptides. Protein Sci. 2004, 13, 893− Effective Promotor of Nucleation in Polymer Melts: A Quantitative
902. Evaluation. Rheol. Acta 2003, 42, 355−364.
(130) Janeschitz-Kriegl, H.; Ratajski, E. Kinetics of Polymer (150) Mykhaylyk, O. O.; Chambon, P.; Graham, R. S.; Fairclough, J.
Crystallization under Processing Conditions: Transformation of P. A.; Olmsted, P. D.; Ryan, A. J. The Specific Work of Flow as a
Dormant Nuclei by the Action of Flow. Polymer 2005, 46, 3856−3870. Criterion for Orientation in Polymer Crystallization. Macromolecules
(131) Janeschitz-Kriegl, H. Some Remarks on Flow Induced 2008, 41, 1901−1904.
Crystallization in Polymer Melts. J. Rheol. 2013, 57, 1057. (151) Mykhaylyk, O. O.; Chambon, P.; Impradice, C.; Fairclough, J. P.
(132) Liedauer, S.; Eder, G.; Janeschitz-Kriegl, H. On the Limitations A.; Terrill, N. J.; Ryan, A. J. Control of Structural Morphology in Shear-
of Shear-Induced Crystallization in Polypropylene Melts. Int. Polym. Induced Crystallization of Polymers. Macromolecules 2010, 43, 2389−
Process. 1995, 10, 243−250. 2405.
(133) Sun, X.; Li, H.; Wang, J.; Yan, S. Shear-Induced Interfacial (152) D’Haese, M.; Mykhaylyk, O. O.; Van Puyvelde, P. On the Onset
Structure of Isotactic Polypropylene (iPP) in iPP/fiber Composites. of Oriented Structures in Flow-Induced Crystallization of Polymers: A
Macromolecules 2006, 39, 8720−8726. Comparison of Experimental Techniques. Macromolecules 2011, 44,
(134) Cui, K.; Meng, L.; Ji, Y.; Li, J.; Zhu, S.; Li, X.; Tian, N.; Liu, D.; 1783−1787.
Li, L. Extension-Induced Crystallization of Poly(ethylene Oxide) (153) Keum, J. K.; Zuo, F.; Hsiao, B. S. Formation and Stability of
Bidisperse Blends: An Entanglement Network Perspective. Macro- Shear-Induced Shish-Kebab Structure in Highly Entangled Melts of
molecules 2014, 47, 677−686. UHMWPE/HDPE Blends. Macromolecules 2008, 41, 4766−4776.
(135) Hobbs, J. K.; Farrance, O. E.; Kailas, L. How Atomic Force (154) Kumaraswamy, G.; Kornfield, J. A.; Yeh, F. J.; Hsiao, B. S. Shear-
Microscopy Has Contributed to Our Understanding of Polymer Enhanced Crystallization in Isotactic Polypropylene. 3. Evidence for a
Crystallization. Polymer 2009, 50, 4281−4292. Kinetic Pathway to Nucleation. Macromolecules 2002, 35, 1762−1769.
(136) Hobbs, J. K.; Miles, M. J. Direct Observation of Polyethylene (155) van Meerveld, J.; Peters, G. W. M.; Hutter, M. Towards a
Shish-Kebab Crystallization Using in-Situ Atomic Force Microscopy. Rheological Classification of Flow Induced Crystallization Experiments
Macromolecules 2001, 34, 353−355. of Polymer Melts. Rheol. Acta 2004, 44, 119−134.
(137) Roozemond, P. C.; van Drongelen, M.; Ma, Z.; Spoelstra, A. B.; (156) Housmans, J. W.; Steenbakkers, R. J. A.; Roozemond, P. C.;
Hermida-Merino, D.; Peters, G. W. M. Self-Regulation in Flow-Induced Peters, G. W. M.; Meijer, H. E. H. Saturation of Pointlike Nuclei and the
Structure Formation of Polypropylene. Macromol. Rapid Commun. Transition to Oriented Structures in Flow-Induced Crystallization of
2015, 36, 385−390. Isotactic Polypropylene. Macromolecules 2009, 42, 5728−5740.
(138) Mykhaylyk, O. O.; Fernyhough, C. M.; Okura, M.; Fairclough, J. (157) Wang, M. X.; Hu, W. B.; Ma, Y.; Ma, Y. Q. Orientational
P. A.; Ryan, A. J.; Graham, R. Monodisperse Macromolecules − A Relaxation Together with Polydispersity Decides Precursor Formation
Stepping Stone to Understanding Industrial Polymers. Eur. Polym. J.
in Polymer Melt Crystallization. Macromolecules 2005, 38, 2806−2812.
2011, 47, 447−464. (158) Bai, H.; Deng, H.; Zhang, Q.; Wang, K.; Fu, Q.; Zhang, Z.; Men,
(139) Fernandez-Ballester, L.; Thurman, D. W.; Zhou, W.; Kornfield,
Y. Effect of Annealing on the Microstructure and Mechanical Properties
J. A. Effect of Long Chains on the Threshold Stresses for Flow-Induced
of Polypropylene with Oriented Shish-Kebab Structure. Polym. Int.
Crystallization in iPP: Shish Kebabs vs Sausages. Macromolecules 2012,
2012, 61, 252−258.
45, 6557−6570.
(159) Ning, N.; Luo, F.; Pan, B.; Zhang, Q.; Wang, K.; Fu, Q.
(140) Shen, B.; Liang, Y.; Kornfield, J. A.; Han, C. C. Mechanism for
Observation of Shear-Induced Hybrid Shish Kebab in the Injection
Shish Formation under Shear Flow: An Interpretation from an in Situ
Morphological Study. Macromolecules 2013, 46, 1528−1542. Molded Bars of Linear Polyethylene Containing Inorganic Whiskers.
(141) Zhang, B.; Chen, J. B.; Ji, F. F.; Zhang, X. L.; Zheng, G. Q.; Shen, Macromolecules 2007, 40, 8533−8536.
C. Y. Effects of Melt Structure on Shear-Induced Beta-Cylindrites of (160) Xu, H.; Zhong, G.-J.; Fu, Q.; Lei, J.; Jiang, W.; Hsiao, B. S.; Li,
Isotactic Polypropylene. Polymer 2012, 53, 1791−1800. Z.-M. Formation of Shish-Kebabs in Injection-Molded Poly(L -Lactic
(142) Zhang, B.; Chen, J. B.; Cui, J.; Zhang, H.; Ji, F. F.; Zheng, G. Q.; Acid) by Application of an Intense Flow Field. ACS Appl. Mater.
Heck, B.; Reiter, G.; Shen, C. Y. Effect of Shear Stress on Crystallization Interfaces 2012, 4, 6774−6784.
of Isotactic Polypropylene from a Structured Melt. Macromolecules (161) Yang, H.-R.; Lei, J.; Li, L.; Fu, Q.; Li, Z.-M. Formation of
2012, 45, 8933−8937. Interlinked Shish-Kebabs in Injection-Molded Polyethylene under the
(143) Zhang, B.; Chen, J.; Zhang, X.; Shen, C. Formation of β- Coexistence of Lightly Cross-Linked Chain Network and Oscillation
Cylindrites under Supercooled Extrusion of Isotactic Polypropylene at Shear Flow. Macromolecules 2012, 45, 6600−6610.
Low Shear Stress. Polymer 2011, 52, 2075−2084. (162) Xu, H.; Xie, L.; Chen, Y.-H.; Huang, H.-D.; Xu, J.-Z.; Zhong, G.-
(144) Liu, D.; Tian, N.; Cui, K.; Zhou, W.; Li, X.; Li, L. Correlation J.; Hsiao, B. S.; Li, Z.-M. Strong Shear Flow-Driven Simultaneous
between Flow-Induced Nucleation Morphologies and Strain in Formation of Classic Shish-Kebab, Hybrid Shish-Kebab, and Trans-
Polyethylene: From Uncorrelated Oriented Point-Nuclei, Scaffold- crystallinity in Poly(lactic acid)/Natural Fiber Biocomposites. ACS
Network, and Microshish to Shish. Macromolecules 2013, 46, 3435− Sustainable Chem. Eng. 2013, 1, 1619−1629.
3443. (163) Ma, G.; Li, D.; Sheng, J. Shear-Induced Crystallization in Phase-
(145) Balzano, L.; Rastogi, S.; Peters, G. W. M. Crystallization and Separated Blends of Isotactic Polypropylene with Ethylene-Propylene-
Precursors during Fast Short-Term Shear. Macromolecules 2009, 42, Diene Terpolymer. Chin. J. Polym. Sci. 2015, 33, 1538−1549.
2088−2092. (164) An, F.; Gao, X.; Lei, J.; Deng, C.; Li, Z.; Shen, K. Vibration
(146) Somani, R. H.; Yang, L.; Hsiao, B. S.; Sun, T.; Pogodina, N. V.; Assisted Extrusion of Polypropylene. Chin. J. Polym. Sci. 2015, 33, 688−
Lustiger, A. Shear-Induced Molecular Orientation and Crystallization 696.
in Isotactic Polypropylene: Effects of the Deformation Rate and Strain. (165) Jiang, Z.; Tang, Y.; Rieger, J.; Enderle, H.-F.; Lilge, D.; Roth, S.
Macromolecules 2005, 38, 1244−1255. V.; Gehrke, R.; Wu, Z.; Li, Z.; Li, X.; et al. Structural Evolution of Melt-
(147) Kimata, S.; Sakurai, T.; Nozue, Y.; Kasahara, T.; Yamaguchi, N.; Drawn Transparent High-Density Polyethylene during Heating and
Karino, T.; Shibayama, M.; Kornfield, J. A. Molecular Basis of the Shish- Annealing: Synchrotron Small-Angle X-Ray Scattering Study. Eur.
Kebab Morphology in Polymer Crystallization. Science 2007, 316, Polym. J. 2010, 46, 1866−1877.
1014−1017. (166) Li, X.-J.; Li, Z.-M.; Zhong, G.-J.; Li, L.-B. Steady-Shear-Induced
(148) Balzano, L.; Cavallo, D.; Van Erp, T. B.; Ma, Z.; Housmans, J.- Isothermal Crystallization of Poly(L-Lactide) (PLLA). J. Macromol.
W.; Fernandez-Ballester, L.; Peters, G. W. M. Dynamics of Fibrillar Sci., Part B: Phys. 2008, 47, 511−522.
Precursors of Shishes as a Function of Stress. IOP Conf. Ser.: Mater. Sci. (167) Yang, S.-G.; Zhang, Z.; Zhou, D.; Wang, Y.; Lei, J.; Li, L.; Li, Z.-
Eng. 2010, 14, No. 012005. M. Flow and Pressure Jointly Induced Ultrahigh Melting Temperature

1883 DOI: 10.1021/acs.chemrev.7b00500


Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

Spherulites with Oriented Thick Lamellae in Isotactic Polypropylene. Polymers  I. Isotactic Polypropylene. Colloid Polym. Sci. 1984, 262,
Macromolecules 2015, 48, 5834−5844. 294−300.
(168) Zhang, R.-C.; Xu, Y.; Lu, A.; Cheng, K.; Huang, Y.; Li, Z.-M. (192) Koike, Y.; Cakmak, M. Atomic Force Microscopy Observations
Shear-Induced Crystallization of Poly(phenylene Sulfide). Polymer of the Structural Development during the Uniaxial Stretching of
2008, 49, 2604−2613. Crosslinked Low-Density Polyethylene in Partial and Fully Molten
(169) Yang, S.-G.; Zhang, Z.; Zhang, L.-Q.; Zhou, D.; Wang, Y.; Lei, J.; States. J. Polym. Sci., Part B: Polym. Phys. 2004, 42, 2228−2237.
Li, L.; Li, Z.-M. Unexpected Shear Dependence of Pressure-Induced γ- (193) Hashimoto, T. Mechanics” of Molecular Assembly: Real-Time
Crystals in Isotactic Polypropylene. Polym. Chem. 2015, 6, 4588−4596. and In-Situ Analysis of Nano-to-Mesoscopic Scale Hierarchical
(170) Qin, Y.; Xu, Y.; Zhang, L.; Zheng, G.; Dai, K.; Liu, C.; Yan, X.; Structures and Nonequilibrium Phenomena. Bull. Chem. Soc. Jpn.
Guo, J.; Guo, Z. Shear-Induced Interfacial Sheath Structure in Isotactic 2005, 78, 1−39.
Polypropylene/glass Fiber Composites. Polymer 2015, 70, 326−335. (194) Murase, H.; Kume, T.; Hashimoto, T.; Ohta, Y. Time Evolution
(171) Zhang, C.; Wang, B.; Yang, J.; Ding, D.; Yan, X.; Zheng, G.; Dai, of Structures under Shear-Induced Phase Separation and Crystal-
K.; Liu, C.; Guo, Z. Synergies among the Self-Assembled β-Nucleating lization in Semidilute Solution of Ultrahigh Molecular Weight
Agent and the Sheared Isotactic Polypropylene Matrix. Polymer 2015, Polyethylene. Macromolecules 2005, 38, 8719−8728.
60, 40−49. (195) Kawakami, D.; Ran, S. F.; Burger, C.; Avila-Orta, C.; Sics, I.;
(172) Ziabicki, A.; Kedzierska, K. Studies on the Orientation
Chu, B.; Hsiao, B. S.; Kikutani, T. Superstructure Evolution in
Phenomena by Fiber Formation from Polymer Melts. Part I.
Poly(ethylene Terephthalate) during Uniaxial Deformation above
Preliminary Investigations on Polycapronamide. J. Appl. Polym. Sci.
1959, 2, 14−23. Glass Transition Temperature. Macromolecules 2006, 39, 2909−2920.
(173) Ziabicki, A. Studies on Orientation Phenomena by Fiber (196) Van der Beek, M. H. E.; Peters, G. W. M.; Meijer, H. E. H.
Formation from Polymer Melts. Part II. Theoretical Considerations. J. Influence of Shear Flow on the Specific Volume and the Crystalline
Appl. Polym. Sci. 1959, 2, 24−31. Morphology of Isotactic Polypropylene. Macromolecules 2006, 39,
(174) Peterlin, A. Hydrodynamics of Macromolecules in a Velocity 1805−1814.
Field with Longitudinal Gradient. J. Polym. Sci., Part B: Polym. Lett. (197) Matsuba, G.; Sakamoto, S.; Ogino, Y.; Nishida, K.; Kanaya, T.
1966, 4, 287−291. Crystallization of Polyethylene Blends under Shear Flow. Effects of
(175) Mitsuhashi, S. On Polyethylene Crystals Grown from Flowing Crystallization Temperature and Ultrahigh Molecular Weight
Solutions in Xylene. Bull. Text. Res. Inst. Jpn. 1963, 66, 1−10. Component. Macromolecules 2007, 40, 7270−7275.
(176) Vanderheijde, H. B. Whisker-Like Growth of Polyoxymethylene (198) Zhao, Y.; Matsuba, G.; Nishida, K.; Fujiwara, T.; Inoue, R.;
from Solution. Nature 1963, 199, 798−799. Polec, I.; Deng, C.; Kanaya, T. Relaxation of Shish-Kebab Precursor in
(177) Pennings, A. J.; Kiel, A. M. Fractionation of Polymers by Isotactic Polystyrene after Short-Term Shear Flow. J. Polym. Sci., Part B:
Crystallization from Solution, III. On Morphology of Fibrillar Polym. Phys. 2011, 49, 214−221.
Polyethylene Crystals Grown in Solution. Colloid Polym. Sci. 1965, (199) Stribeck, N.; Androsch, R.; Funari, S. S. Nanostructure
205, 160−162. Evolution of Homogeneous Poly(ethylene-Co-1-Octene) as a Function
(178) Reneker, D. H. Localized Deformation of Lamellar Poly- of Strain. Macromol. Chem. Phys. 2003, 204, 1202−1216.
ethylene Crystals. J. Polym. Sci., Part A: Gen. Pap. 1965, 3, 1069−1077. (200) Hill, M. J.; Keller, A. Hairdressing Shish-Kebabs by Melting.
(179) Blackadder, D. A.; Schleinitz, H. M. Effect of Ultrasonic Colloid Polym. Sci. 1981, 259, 335−341.
Radiation on Crystallization of Polyethylene from Dilute Solution. (201) Liu, T. X.; Tjiu, W. C.; Petermann, J. Transmission Electron
Nature 1963, 200, 778−779. Microscopy Observations on Fine Structures of Shish-Kebab Crystals
(180) Pennings, A. J.; van der Mark, J. M. A. A.; Kiel, A. M. of Isotactic Polystyrene by Partial Melting. J. Cryst. Growth 2002, 243,
Hydrodynamically Induced Crystallization of Polymers from Solution. 218−223.
III. Morphology. Colloid Polym. Sci. 1970, 237, 336−358. (202) Isayev, A. I.; Chan, T. W.; Shimojo, K.; Gmerek, M. Injection-
(181) Pennings, A. J.; Mark, J. M. A. A.; Booij, H. C. Hydrodynami- Molding of Semicrystalline Polymers. I Material Characterization. J.
cally Induced Crystallization of Polymers from Solution. II. the Effect of Appl. Polym. Sci. 1995, 55, 807−819.
Secondary Flow. Colloid Polym. Sci. 1970, 236, 99−111. (203) Yamazaki, S.; Hikosaka, M.; Toda, A.; Wataoka, I.; Yamada, K.;
(182) Frank, F. C. Strength and Stiffness of Polymers. Proc. R. Soc. Tagashira, K. Nucleation and Morphology of Polyethylene Under Shear
London, Ser. A 1970, 319, 127−136. Flow. J. Macromol. Sci., Part B: Phys. 2003, 42, 499−514.
(183) Zwijnenburg, A.; Pennings, A. J. Longitudinal Growth of (204) Keller, A.; Willmouth, F. M. Some Macroscopic Properties of
Polymer Crystals from Flowing Solutions II. Polyethylene Crystals in Stirring-Induced Crystals of Polyethylene. J. Macromol. Sci., Part B:
Poiseuille Flow. Colloid Polym. Sci. 1975, 253, 452−461.
Phys. 1972, 6, 493−537.
(184) Frank, F. C.; Keller, A.; Mackley, M. R. Polymer Chain
(205) Kanaya, T.; Matsuba, G.; Ogino, Y.; Nishida, K.; Shimizu, H.
Extension Produced by Impinging Jets and Its Effect on Polyethylene
M.; Shinohara, T.; Oku, T.; Suzuki, J.; Otomo, T. Hierarchic Structure
Solution. Polymer 1971, 12, 467−473.
(185) Mackley, M. R.; Frank, F. C.; Keller, A. Flow-Induced of Shish-Kebab by Neutron Scattering in a Wide Q Range.
Crystallization of Polyethylene Melts. J. Mater. Sci. 1975, 10, 1501− Macromolecules 2007, 40, 3650−3654.
1509. (206) De Gennes, P. G. Coil-Stretch Transition of Dilute Flexible
(186) Mackley, M. R.; Keller, A. Flow Induced Crystallization of Polymers under Ultrahigh Velocity-Gradients. J. Chem. Phys. 1974, 60,
Polyethylene Melts. Polymer 1973, 14, 16−20. 5030−5042.
(187) Bashir, Z.; Hill, M. J.; Keller, A. Comparative-Study of Etching (207) Keller, A.; Kolnaar, J. W. H. Chain Extension and Orientation:
Techniques for Electron-Microscopy Using Melt Processed Poly- Fundamentals and Relevance to Processing and Products. Prog. Colloid
ethylene. J. Mater. Sci. Lett. 1986, 5, 876−878. Polym. Sci. 1993, 92, 81−102.
(188) Keller, A.; Odell, J. A. The Extensibility of Macromolecules in (208) Mackley, M. Stretching Polymer Chains. Rheol. Acta 2010, 49,
Solution; A New Focus for Macromolecular Science. Colloid Polym. Sci. 443−458.
1985, 263, 181−201. (209) Mackley, M. R.; Keller, A. Flow Induced Polymer-Chain
(189) Barham, P. J.; Keller, A. High-Strength Polyethylene Fibers Extension and Its Relation to Fibrous Crystallization. Philos. Trans. R.
from Solution and Gel Spinning. J. Mater. Sci. 1985, 20, 2281−2302. Soc., A 1975, 278, 29−66.
(190) Hill, M. J.; Barham, P. J.; Keller, A. On the Hairdressing of (210) Smith, D. E.; Babcock, H. P.; Chu, S. Single-Polymer Dynamics
Shish-Kebabs. Colloid Polym. Sci. 1980, 258, 1023−1037. in Steady Shear Flow. Science 1999, 283, 1724−1727.
(191) Schultz, J. M.; Petermann, J. Transmission Electron Microscope (211) Smith, D. E.; Chu, S. Response of Flexible Polymers to a Sudden
Observations of Fibrillar-to-Lamellar Transformations in Melt-Drawn Elongational Flow. Science. 1998, 281, 1335−1340.

1884 DOI: 10.1021/acs.chemrev.7b00500


Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

(212) Dukovski, I.; Muthukumar, M. Langevin Dynamics Simulations (232) Okura, M.; Mykhaylyk, O. O.; Ryan, A. J. Effect of Matrix
of Early Stage Shish-Kebab Crystallization of Polymers in Extensional Polymer on Flow-Induced Nucleation in Polymer Blends. Phys. Rev.
Flow. J. Chem. Phys. 2003, 118, 6648−6655. Lett. 2013, 110, No. 087801.
(213) Chang, H.; Lee, K. G.; Schultz, J. M. Structure Development of (233) Humphris, A. D. L.; Hobbs, J. K.; Miles, M. J. Ultrahigh-Speed
Polyethylene Terephthalate Fibers during Postspinning Annealing. J. Scanning near-Field Optical Microscopy Capable of over 100 Frames
Macromol. Sci., Part B: Phys. 1994, 33, 105−127. per Second. Appl. Phys. Lett. 2003, 83, 6−8.
(214) Hristov, H. A.; Schultz, J. M. Thermal Response and Structure (234) Humphris, A. D. L.; Miles, M. J.; Hobbs, J. K. A Mechanical
of PET Fibers. J. Polym. Sci., Part B: Polym. Phys. 1990, 28, 1647−1663. Microscope: High-Speed Atomic Force Microscopy. Appl. Phys. Lett.
(215) Petermann, J.; Gleiter, H. Electron Microscopic Observations 2005, 86, No. 034106.
on the Crystallization of Row Structures in Strained Melts. J. Polym. Sci., (235) Hobbs, J. K. In Situ Atomic Force Microscopy of the Melting of
Polym. Lett. Ed. 1977, 15, 649−654. Melt-Crystallized Polyethylene. Polymer 2006, 47, 5566−5573.
(216) Keller, A.; Kolnaar, H. W. H. Flow-Induced Orientation and (236) Petermann, J.; Miles, M.; Gleiter, H. Crystalline Core of the
Structure Formation. In Materials Science and Technology, Part II. Row Structures in Isotactic Polystyrene. I Nucleation and Growth. J.
Structure Development During Processing; Wiley−VCH: Weinheim, Polym. Sci., Polym. Phys. Ed. 1979, 17, 55−62.
Germany, 2006; DOI: 10.1002/9783527603978.mst0210. (237) Ma, Z.; Balzano, L.; Peters, G. W. M. Dissolution and Re-
(217) Yan, T. Z.; Zhao, B. J.; Cong, Y. H.; Fang, Y. Y.; Cheng, S. W.; Li, Emergence of Flow-Induced Shish in Polyethylene with a Broad
L. B.; Pan, G. Q.; Wang, Z. J.; Li, X. H.; Bian, F. G. Critical Strain for Molecular Weight Distribution. Macromolecules 2016, 49, 2724−2730.
Shish-Kebab Formation. Macromolecules 2010, 43, 602−605. (238) Flory, P. J. Thermodynamics of Crystallization in High
(218) Hoffman, J. On the Formation of Polymer Fibrils by Flow- Polymers. I. Crystallization Induced by Stretching. J. Chem. Phys.
Induced Crystallization. Polymer 1979, 20, 1071−1077. 1947, 15, 397−408.
(219) Hoffman, J. D. Theory of Flow-Induced Fibril Formation in (239) Evans, R. D.; Mighton, H. R.; Flory, P. J. Thermodynamics of
Polymer-Solutions. J. Res. Natl. Bur. Stand. (U. S.) 1979, 84, 359−384. Crystallization in High Polymers. III. Dependence of Melting Points of
(220) Smook, J.; Pennings, A. J. Elastic Flow Instabilities and Shish- Polyesters on Molecular Weight and Composition. J. Chem. Phys. 1947,
Kebab Formation during Gel-Spinning of Ultrahigh Molecular-Weight 15, 685.
Polyethylene. J. Mater. Sci. 1984, 19, 31−43. (240) Yeh, G. S. Y.; Hong, K. Z. Strain-Induced Crystallization, Part
(221) Zuo, F.; Keum, J. K.; Yang, L.; Somani, R. H.; Hsiao, B. S. III. Theory. Polym. Eng. Sci. 1979, 19, 395−400.
Thermal Stability of Shear-Induced Shish-Kebab Precursor Structure (241) Somani, R. H.; Yang, L.; Hsiao, B. S.; Agarwal, P. K.; Fruitwala,
from High Molecular Weight Polyethylene Chains. Macromolecules H. A.; Tsou, A. H. Shear-Induced Precursor Structures in Isotactic
2006, 39, 2209−2218. Polypropylene Melt by in-Situ Rheo-SAXS and Rheo-WAXD Studies.
(222) Murase, H.; Ohta, Y.; Hashimoto, T. A New Scenario of Shish- Macromolecules 2002, 35, 9096−9104.
Kebab Formation from Homogeneous Solutions of Entangled (242) Mahendrasingam, A.; Martin, C.; Fuller, W.; Blundell, D. J.;
Polymers: Visualization of Structure Evolution along the Fiber Oldman, R. J.; MacKerron, D. H.; Harvie, J. L.; Riekel, C. Observation
Spinning Line. Macromolecules 2011, 44, 7335−7350. of a Transient Structure prior to Strain-Induced Crystallization in
(223) Mackley, M. R.; Moggridge, G. D.; Saquet, O. Direct Poly(ethylene Terephthalate). Polymer 2000, 41, 1217−1221.
Experimental Evidence for Flow Induced Fibrous Polymer Crystal- (243) Kanaya, T.; Takayama, Y.; Ogino, Y.; Matsuba, G.; Nishida, K.
lisation Occurring at a Solid/melt Interface. J. Mater. Sci. 2000, 35, Precursor of Primary Nucleation in Isotactic Polystyrene Induced by
5247−5253. Shear Flow. In Progress in Understanding of Polymer Crystallization;
(224) Smook, J.; Pennings, J. Influence of Draw Ratio on Springer: Berlin and Heidelberg, Germany, 2007; pp 87−96; DOI:
Morphological and Structural Changes in Hot-Drawing of UHMW 10.1007/3-540-47307-6_5.
Polyethylene Fibres as Revealed by DSC. Colloid Polym. Sci. 1984, 262, (244) Hashimoto, T.; Murase, H.; Ohta, Y. A New Scenario of Flow-
712−722. Induced Shish-Kebab Formation in Entangled Polymer Solutions.
(225) Ogino, Y.; Fukushima, H.; Matsuba, G.; Takahashi, N.; Nishida, Macromolecules 2010, 43, 6542−6548.
K.; Kanaya, T. Effects of High Molecular Weight Component on (245) McHugh, A. J.; Forrest, E. H. A Discussion of Nucleation and
Crystallization of Polyethylene under Shear Flow. Polymer 2006, 47, Growth in Flow-Induced Crystallization from Solution and an
5669−5677. Improved Model for the Growth Process. J. Macromol. Sci., Part B:
(226) Heeley, E. L.; Fernyhough, C. M.; Graham, R. S.; Olmsted, P. Phys. 1975, 11, 219−238.
D.; Inkson, N. J.; Embery, J.; Groves, D. J.; McLeish, T. C. B.; (246) Onuki, A. Dynamic Scattering and Phase-Separation in
Morgovan, A. C.; Meneau, F.; et al. Shear-Induced Crystallization in Viscoelastic 2-Component Fluids. J. Non-Cryst. Solids 1994, 172-174,
Blends of Model Linear and Long-Chain Branched Hydrogenated 1151−1157.
Polybutadienes. Macromolecules 2006, 39, 5058−5071. (247) Doi, M.; Onuki, A. Dynamic Coupling between Stress and
(227) Cui, K.; Ma, Z.; Wang, Z.; Ji, Y.; Liu, D.; Huang, N.; Chen, L.; Composition in Polymer-Solutions and Blends. J. Phys. II 1992, 2,
Zhang, W.; Li, L. Kinetic Process of Shish Formation: From Stretched 1631−1656.
Network to Stabilized Nuclei. Macromolecules 2015, 48, 5276−5285. (248) Wang, Z.; Ju, J.; Yang, J.; Ma, Z.; Liu, D.; Cui, K.; Yang, H.;
(228) Yang, H.; Liu, D.; Ju, J.; Li, J.; Wang, Z.; Yan, G.; Ji, Y.; Zhang, Chang, J.; Huang, N.; Li, L. The Non-Equilibrium Phase Diagrams of
W.; Sun, G.; Li, L. Chain Deformation on the Formation of Shish Flow-Induced Crystallization and Melting of Polyethylene. Sci. Rep.
Nuclei under Extension Flow: An in Situ SANS and SAXS Study. 2016, 6, No. 32968.
Macromolecules 2016, 49, 9080−9088. (249) Ju, J.; Wang, Z.; Su, F.; Ji, Y.; Yang, H.; Chang, J.; Ali, S.; Li, X.;
(229) Wang, Z.; Su, F.; Ji, Y.; Yang, H.; Tian, N.; Chang, J.; Meng, L.; Li, L. Extensional Flow-Induced Dynamic Phase Transitions in
Li, L. Transition from Chain- to Crystal-Network in Extension Induced Isotactic Polypropylene. Macromol. Rapid Commun. 2016, 37, 1441−
Crystallization of Isotactic Polypropylene. J. Rheol. 2017, 61, 589−599. 1445.
(230) Yang, L.; Somani, R. H.; Sics, I.; Hsiao, B. S.; Kolb, R.; Fruitwala, (250) Wang, Z.; Ju, J.; Meng, L.; Tian, N.; Chang, J.; Yang, H.; Ji, Y.;
H.; Ong, C. Shear-Induced Crystallization Precursor Studies in Model Su, F.; Li, L. Structural and Morphological Transitions in Extension-
Polyethylene Blends by in-Situ Rheo-SAXS and Rheo-WAXD. Induced Crystallization of poly(1-Butene) Melt. Soft Matter 2017, 13,
Macromolecules 2004, 37, 4845−4859. 3639−3648.
(231) Zhao, B. J.; Li, X. Y.; Huang, Y. J.; Cong, Y. H.; Ma, Z.; Shao, C. (251) Liu, D.; Cui, K.; Huang, N.; Wang, Z.; Li, L. The
G.; An, H. N.; Yan, T. Z.; Li, L. B. Inducing Crystallization of Polymer Thermodynamic Properties of Flow-Induced Precursor of Poly-
through Stretched Network. Macromolecules 2009, 42, 1428−1432. ethylene. Sci. China: Chem. 2015, 58, 1570−1578.

1885 DOI: 10.1021/acs.chemrev.7b00500


Chem. Rev. 2018, 118, 1840−1886
Chemical Reviews Review

(252) Coppola, S.; Grizzuti, N.; Maffettone, P. L. Microrheological Deformation Including Chain Stretch and Convective Constraint
Modeling of Flow-Induced Crystallization. Macromolecules 2001, 34, Release. J. Rheol. 2003, 47, 1171.
5030−5036. (274) Zuidema, H.; Peters, G. W. M.; Meijer, H. E. H. Development
(253) Marrucci, G.; Grizzuti, N. The Free-Energy Function of the and Validation of a Recoverable Strain-Based Model for Flow-Induced
Doi-Edwards Theory - Analysis of the Instabilities in Stress-Relaxation. Crystallization of Polymers. Macromol. Theory Simul. 2001, 10, 447−
J. Rheol. 1983, 27, 433−450. 460.
(254) Acierno, S.; Coppola, S.; Grizzuti, N. Effects of Molecular (275) Custodio, F. J. M. F.; Steenbakkers, R. J. A.; Anderson, P. D.;
Weight Distribution on the Flow-Enhanced Crystallization of poly(1- Peters, G. W. M.; Meijer, H. E. H. Model Development and Validation
Butene). J. Rheol. 2008, 52, 551−566. of Crystallization Behavior in Injection Molding Prototype Flows.
(255) Acierno, S.; Coppola, S.; Grizzuti, N.; Maffettone, P. L. Macromol. Theory Simul. 2009, 18, 469−494.
Coupling between Kinetics and Rheological Parameters in the Flow- (276) Rutledge, G. C. Computer Modeling of Polymer Crystallization.
Induced Crystallization of Thermoplastic Polymers. Macromol. Symp. In Handbook of Polymer Crystallization; John Wiley & Sons, Inc.:
2002, 185, 233−241. Hoboken, NJ, 2013; pp 197−214; DOI: 10.1002/9781118541838.ch6.
(256) Bushman, A. C.; McHugh, A. J. A Continuum Model for the (277) Haudin, J.-M. Crystallization in Processing Conditions. In
Dynamics of Flow-Induced Crystallization. J. Polym. Sci., Part B: Polym. Handbook of Polymer Crystallization; John Wiley & Sons, Inc.:
Phys. 1996, 34, 2393−2407. Hoboken, NJ, 2013; pp 433−462; DOI: 10.1002/
(257) Titomanlio, G.; Lamberti, G. Modeling Flow Induced 9781118541838.ch15.
Crystallization in Film Casting of Polypropylene. Rheol. Acta 2004, (278) Eder, G.; Janeschitz-Kriegl, H. Crystallization. In Materials
43, 146−158. Science and Technology, Part II. Structure Development during Processing;
(258) Lamberti, G. Flow-Induced Crystallization during Isotactic Wiley−VCH: Weinheim, Germany, 2006; DOI: 10.1002/
Polypropylene Film Casting. Polym. Eng. Sci. 2011, 51, 851−861. 9783527603978.mst0211.
(259) Pantani, R.; Coccorullo, I.; Speranza, V.; Titomanlio, G. (279) McHugh, A. J.; Guy, R. K.; Tree, D. A. Extensional Flow-
Modeling of Morphology Evolution in the Injection Molding Process of Induced Crystallization of a Polyethylene Melt. Colloid Polym. Sci.
Thermoplastic Polymers. Prog. Polym. Sci. 2005, 30, 1185−1222. 1993, 271, 629−645.
(260) Tian, N.; Zhou, W.; Cui, K.; Liu, Y.; Fang, Y.; Wang, X.; Liu, L.; (280) Steenbakkers, R. J. A.; Peters, G. W. M. A Stretch-Based Model
Li, L. Extension Flow Induced Crystallization of Poly(ethylene Oxide). for Flow-Enhanced Nucleation of Polymer Melts. J. Rheol. 2011, 55,
Macromolecules 2011, 44, 7704−7712. 401−433.
(261) Tian, N.; Liu, D.; Li, X.; Wang, Z.; Zhu, S.; Cui, K.; Zhou, W.; Li, (281) van Erp, T. B.; Roozemond, P. C.; Peters, G. W. M. Flow-
L. Relaxation Propelled Long Period Change in the Extension Induced Enhanced Crystallization Kinetics of iPP during Cooling at Elevated
Crystallization of Polyethylene Oxide. Soft Matter 2013, 9, 10759− Pressure: Characterization, Validation, and Development. Macromol.
10767. Theory Simul. 2013, 22, 309−318.
(262) Tian, N.; Liu, D.; Meng, L.; Zhou, W.; Hu, T.; Li, X.; Li, L. How (282) Roozemond, P. C.; van Drongelen, M.; Ma, Z.; Hulsen, M. A.;
Flow Affects Crystallization in a Heterogeneous Polyethylene Oxide Peters, G. W. M. Modeling Flow-Induced Crystallization in Isotactic
Melt. RSC Adv. 2014, 4, 9632−9638. Polypropylene at High Shear Rates. J. Rheol. 2015, 59, 613−642.
(263) Elmoumni, A.; Gonzalez-Ruiz, R. A.; Coughlin, E. B.; Winter, H. (283) Pantani, R.; Nappo, V.; De Santis, F.; Titomanlio, G. Fibrillar
H. Isotactic Poly(propylene) Crystallization: Role of Small Fractions of Morphology in Shear-Induced Crystallization of Polypropylene.
High or Low Molecular Weight Polymer. Macromol. Chem. Phys. 2005, Macromol. Mater. Eng. 2014, 299, 1465−1473.
206, 125−134.
(264) Boukany, P. E.; Tapadia, P.; Wang, S.-Q. Interfacial Stick-Slip
Transition in Simple Shear of Entangled Melts. J. Rheol. 2006, 50, 641−
654.
(265) Fang, Y.; Wang, G.; Tian, N.; Wang, X.; Zhu, X.; Lin, P.; Ma, G.;
Li, L. Shear Inhomogeneity in Poly(ethylene Oxide) Melts. J. Rheol.
2011, 55, 939−949.
(266) Cong, Y.; Liu, H.; Wang, D.; Zhao, B.; Yan, T.; Li, L.; Chen, W.;
Zhong, Z.; Lin, M. C.; Chen, H. L.; et al. Stretch-Induced
Crystallization through Single Molecular Force Generating Mecha-
nism. Macromolecules 2011, 44, 5878−5882.
(267) Ko, M. J.; Waheed, N.; Lavine, M. S.; Rutledge, G. C.
Characterization of Polyethylene Crystallization from an Oriented Melt
by Molecular Dynamics Simulation. J. Chem. Phys. 2004, 121, 2823−
2832.
(268) Yamamoto, T. Molecular Dynamics Simulations of Polymer
Crystallization in Highly Supercooled Melt: Primary Nucleation and
Cold Crystallization. J. Chem. Phys. 2010, 133, No. 034904.
(269) Romanos, N. A.; Theodorou, D. N. Crystallization and Melting
Simulations of Oligomeric α1 Isotactic Polypropylene. Macromolecules
2010, 43, 5455−5469.
(270) Hu, W. B.; Frenkel, D.; Mathot, V. B. F. Simulation of Shish-
Kebab Crystallite Induced by a Single Prealigned Macromolecule.
Macromolecules 2002, 35, 7172−7174.
(271) Xu, G.; Lin, H.; Mattice, W. L. Configuration Selection in the
Simulations of the Crystallization of Short Polyethylene Chains in a
Free-Standing Thin Film. J. Chem. Phys. 2003, 119, 6736−6743.
(272) Graham, R. S.; Olmsted, P. D. Coarse-Grained Simulations of
Flow-Induced Nucleation in Semicrystalline Polymers. Phys. Rev. Lett.
2009, 103, No. 115702.
(273) Graham, R. S.; Likhtman, A. E.; McLeish, T. C. B.; Milner, S. T.
Microscopic Theory of Linear, Entangled Polymer Chains under Rapid

1886 DOI: 10.1021/acs.chemrev.7b00500


Chem. Rev. 2018, 118, 1840−1886

You might also like