Vous êtes sur la page 1sur 31

Progress in Polymer Science 37 (2012) 14251455

Contents lists available at SciVerse ScienceDirect

Progress in Polymer Science


journal homepage: www.elsevier.com/locate/ppolysci

Realizing the enhancement of interfacial interaction in semicrystalline


polymer/ller composites via interfacial crystallization
Nanying Ning, Sirui Fu, Wei Zhang, Feng Chen, Ke Wang , Hua Deng,
Qin Zhang, Qiang Fu
College of Polymer Science and Engineering, State Key Laboratory of Polymer Materials Engineering, Sichuan University, Chengdu 610065, PR China

a r t i c l e

i n f o

Article history:
Received 11 August 2011
Received in revised form
17 December 2011
Accepted 21 December 2011
Available online 27 December 2011
Keywords:
Interfacial crystallization
Hybrid crystalline structure
Crystalline mechanism
Interfacial interaction
Interfacial enhancement
Mechanical properties

a b s t r a c t
Polymer/ller composites have been widely used in various areas. One of the keys to
achieve the high performance of these composites is good interfacial interaction between
polymer matrix and ller. As a relatively new approach, the possibility to enhance polymer/ller interfacial interaction via crystallization of polymer on the surface of llers, i.e.,
interfacial crystallization, is summarized and discussed in this paper. Interfacial crystallization has attracted tremendous interest in the past several decades, and some unique
hybrid crystalline structures have been observed, including hybrid shishkebab and hybrid
shishcalabash structures in which the ller served as the shish and crystalline polymer
as the kebab/calabash. Thus, the manipulation of the interfacial crystallization architecture
offers a potential highly effective route to achieve strong polymer/ller interaction. This
review is based on the latest development of interfacial crystallization in polymer/ller
composites and will be organized as follows. The structural/morphological features of
various interfacial crystallization fashions are described rst. Subsequently, various inuences on the nal structure/morphology of hybrid crystallization and the nucleation and/or
growth mechanisms of crystallization behaviors at polymer/ller interface are reviewed.

Abbreviations: 2d, two dimensional; AD-MWNTs, multi wall carbon nanotubes synthesized by arc discharge method; AFM, atomic force microscopy;
b, thickness of polymer coating layer; CF, carbon ber; CNF, carbon nanober; CNT, carbon nanotube; CNTs, carbon nanotubes; CTE, coefcient of thermal expansion; CVD, chemical vapor deposition; CVD-MWNTs, multi wall carbon nanotubes synthesized by CVD method; DBS, 1,3:2,4-dibenzylidene
glucitol; Df , ber diameter; DMA, dynamic mechanical analysis; DMAc, N,N-dimethyl acetamide; DMF, N,N-dimethyl formamide; DMSO, dimethyl sulfoxide; DPIM, dynamic packing injection molding technology; DSC, differential scanning calorimetry; Fmax , maximum pullout force; FTIR, Fourier-transform
infrared spectroscopy; GF, glass ber; GONPs, graphite oxide nanoplatelets; HDPE, high density polyethylene; HDT, thermal distortion temperature; HMCF,
ultrahigh-modulus carbon ber; HMW-PE, high molecular weight PE; HMW-PP, high molecular weight polypropylene; HOPG, highly oriented pyrolytic
graphite; HSC, hybrid shishcalabash; HSK, hybrid shishkebab; H-T equation, Halpin-Tsai equation; HTCF, high-tenacity carbon ber; IFSS, interfacial
shear strength; IMCF, intermediate-modulus carbon ber; iPP, isotactic polypropylene; K-BrBz, potassium 4-bromobenzoate; lc /D, critical aspect ratio; lc ,
critical effective length; lemb , ber embedded length; LLDPE, linear low density polyethylene; lm , mean fragment length of ber; LMW-PE, low molecular
weight PE; LMW-PP, low molecular weight polypropylene; MAPP, maleic anhydride grafted polypropylene; MD, molecular dynamics; MoS2 , molybdenum
disulde; MWNT, multi wall carbon nanotube; MWNTs, multi wall carbon nanotubes; NF, natural ber; NHSK, nanohybrid shishkebab; P3HT, poly (3hexylthiophene); PA, polyamide; PA-12, polyamide-12; PA-6, polyamide-6; PAN, polyacrylonitrile; PBT, polybutylece terephthalate; PCL, polycaprolactone;
PE, polyethylene; PE-b-PEO, polyethylene-b-poly ethylene oxide; PEEK, poly (ether ether ketone); PEO, poly ethylene oxide; PET, poly(ethylene terepthalate); PHBV, poly(hydroxybutyrate-co-hydroxyvalerate); PLLA, poly(l-lactide); PP, polypropylene; PPDT, poly (p-phenylene terephthalamide); Pp-g-MA,
polypropylene grafted maleic anhydride; PPS, poly (phenylene sulde); PVA, poly(vinyl alcohol); PVDF, poly(vinylidene uoride); rf , ber radius; SC CO2 ,
supercritical CO2 ; SEM, scanning electron microscopy; SMCW, SiO2 MgOCaO whisker; sPP, syndiotactic polypropylene; sPS, syndiotactic polystyrene;
SWNT, single wall carbon nanotube; SWNTs, single wall carbon nanotubes; TC, transcrystallinity; TEM, transmission electron microscopy; UHMCF, ultrahigh
modulus carbon ber; UHMWPE, ultrahigh molecular weight polyethylene; Vf , ber volume fraction; VGCF, vapor grown carbon bers; WAXD, wide-angle
X-ray diffraction; 0 , orientation efciency factor of ber; l , length efciency factor of ber;  c , composite strength;  f , ber tensile strength;  m , basal
polymer strength;  s , shear strength at the edge of the interfacial layer region;  i , interfacial shear strength.
Corresponding author. Tel.: +86 28 85461795.
Corresponding author.
E-mail addresses: wkestar@scu.edu.cn (K. Wang), qiangfu@scu.edu.cn (Q. Fu).
0079-6700/$ see front matter 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.progpolymsci.2011.12.005

1426

N. Ning et al. / Progress in Polymer Science 37 (2012) 14251455

Then recent studies on interfacial crystallization induced interfacial enhancement ascertained by different research methodologies are addressed, including a comparative analysis
to highlight the positive role of interfacial crystallization on the resultant mechanical reinforcement. Finally, a conclusion, including future perspectives, is presented.
2011 Elsevier Ltd. All rights reserved.

Contents
1.
2.

3.

4.

5.

6.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1426
Hybrid crystalline structures in semicrystalline polymer composites. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1427
2.1.
Transcrystallinity (TC) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1428
2.2.
Hybrid shishkebab (HSK) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1430
2.3.
Hybrid shishcalabash structure (HSC) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1432
2.4.
Other hybrid crystalline structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1432
The factors controlling the hybrid crystalline structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1434
3.1.
Effect of llers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1434
3.1.1.
Surface chemical and physical characteristic of llers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1434
3.1.2.
Geometry and size of llers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1435
3.2.
Effect of polymer matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1435
3.2.1.
Molecular weight . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1435
3.2.2.
Molecular chain structure and conformation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1436
3.2.3.
Functional groups of polymer matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1436
3.3.
Effect of external eld . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1436
Formation mechanisms of hybrid crystalline structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1438
4.1.
Epitaxy and soft epitaxy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1438
4.2.
Chemisorption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1439
4.3.
Stress or strain induced interfacial crystallization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1439
Interfacial and mechanical enhancement induced by interfacial crystallization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1440
5.1.
Interfacial enhancement of crystalline interface estimated by different research strategies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1440
5.2.
Analysis and discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1446
5.3.
Mechanical properties of composites containing special interfacial crystallization structures . . . . . . . . . . . . . . . . . . . . . . . . . . . 1447
Conclusions and nal remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1449
Acknowledgement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1451
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1451

1. Introduction
It is well known that polymer/ller composites are
among the most important materials in industry, and have
been widely used in chemical engineering, sports goods,
automobiles, aerospace and weapons, etc. Millions of tons
of polymer/ller composites are consumed in numerous applications every year. In addition to homogeneous
dispersion of llers, strong interfacial adhesion between
polymer matrix and ller is essentially crucial to obtain
high-performance polymer/ller composites. In the past,
a great effort has been made to enhance the interaction
between polymer and llers, with most attention on the
strengthening of polymer/ller interface focused on four
strategies: (1) chemical or physical surface modication
of the ller [17], (2) functionalization of the polymer
matrix [8], (3) adding compatibilizer [9,10], and (4) preparing polymer composites via in situ polymerization method
[1114]. Interfacial crystallization offers another possible means to enhance polymer/ller interfacial interaction
in composite systems composed of semicrystalline polymer and ller with high aspect ratio. That is through the
enhanced interaction between polymer and ller caused
by the crystallization of polymer on the ller surface.
These ller particles can act as nuclei and induce the polymer lamellae grow on the ller surface. These as-formed

crystalline superstructures are generally denoted as hybrid


crystalline structure or hybrid crystal, in contrast to conventional supermolecular crystalline structures consisting
of only polymer species (Fig. 1).
Interfacial crystallization has attracted tremendous
interest in the past several decades, not only due to
its crystallography interests, but also due to the fact
that it may be a novel strategy to enhance interfacial adhesion and realize the full potential of llers to
reinforce the mechanical performance of composites. Obviously, the formation of an interfacial crystalline layer can
offer a good interfacial combination between ller and
polymer matrix. Since the origins of ller/polymer interaction for other interfacial-connecting fashions, such as
macromolecular chain wrapping, covalent bonding and
chain-grafting attachment are distinct or well identied
[5,15,16], the positive roles played by these physical or
chemical interfacial combinations on improving load transfer efciency are anticipated. The nature of ller/polymer
linking arising from interfacial crystalline structure is,
however, difcult to elucidate clearly, and there is still
a question of whether interfacial crystallization can ultimately bring effective interfacial enhancement or not has
been argued for a long time. Obvious contradictions in
the experimental work have been reported by different
researchers in the literature [1720]. For micrometer-scale

N. Ning et al. / Progress in Polymer Science 37 (2012) 14251455

1427

Fig. 1. Schematic representation of various hybrid crystalline structures.

ller-reinforced composites, a number of studies show that


interfacial crystallization offers little effect or plays a negative role on interfacial enhancement. However, most recent
advances in nanometer-scale ller-reinforced semicrystalline/polymer composites demonstrate that the appearance of interfacial crystallization favors a strong interfacial
adhesion and high load transfer efciency between
polymer matrix and nanoller. The methodologies of
mechanical model tting analysis and computational
molecular dynamic simulation have provided some quantitative data about the shear strength of crystalline interface
layer, which could be one order of magnitude higher
than that for amorphous interface in nanoller-reinforced
composites. To the best of our knowledge, there is no
systemic and detailed review on interfacial enhancement
and mechanical reinforcement of polymer/ller composites via interfacial crystallization, though there are several
general reviews related to the mechanical reinforcement
in polymer/ller composites [2123]. Thus, this review
paper is based on the latest developments of interfacial

crystallization in polymer/ller composites and it is organized according to the following topics: (1) hybrid crystalline structures/morphologies achieved in the presence
of micro- and/or nano-llers; (2) factors controlling hybrid
crystalline structures/morphologies; (3) formation mechanisms of hybrid crystalline structures/morphologies;
(4) interfacial and mechanical enhancement induced by
interfacial crystallization; (5) conclusions and future perspectives.
2. Hybrid crystalline structures in semicrystalline
polymer composites
For semicrystalline polymer composites, the llers can
act as a nucleating agent and have the potential to induce
or alter polymer crystallization. Thus, an increase in crystallization rate and crystallization temperature as well as
a decrease in crystal size is generally observed. Furthermore, various hybrid crystalline structures may be formed
in these composites, as schematically shown in Fig. 1.

1428

N. Ning et al. / Progress in Polymer Science 37 (2012) 14251455

For polymer/spherical particle composites, the hybrid


spherulites with much smaller size than bulk spherulites
will be formed due to the nucleation effect of spherical particles. For polymer/lamellar ller composites, the hybrid
crystalline structure will be polymer lamellae decorating
the surface of the ller. For polymer/brous-ller composites, the transcrystallinity (TC) or hybrid shishkebab
(HSK) or hybrid shishcalabash (HSC) could be obtained. If
the brous ller can initiate a high density of active nuclei
on its surface, which can hinder the free radial growth of
spherulites and therefore force the lamellae to grow in one
direction, the TC structure will be formed. If the brous
ller can initiate a medium density of active nuclei, an
HSK structure with polymer crystal lamellae (kebab) periodically decorating its surface and aligning approximately
perpendicular to its long axis will be formed. If the brous
ller can only initiate a few nuclei, which can develop into
large polymer spherulites without hindrance, the peculiar
HSC structure with brous ller serves as shish and polymer spherulites as calabash will be formed. Among these
hybrid crystalline structures, TC, HSK and HSC are of particular interest and have attracted tremendous attention.
2.1. Transcrystallinity (TC)
The earliest reported hybrid crystalline structure is socalled TC, which often emerges at polymer/ber interfaces.
The formation of this unique TC layer is due to a high density of active nuclei on the ber/substrate surface, which
hinder the full extension of spherulites and therefore force
the crystal growth in one direction, namely perpendicular to the ber/substrate. A typical TC structure in isotactic
polypropylene (iPP)/carbon nanotube (CNT) bers composite is shown in Fig. 2(a) (SEM image) and Fig. 2(b) (PLM
image) [24]. It is quite clear that TC grows perpendicular
to the long axis of the ber. Since it was rst reported in
1952 by Jenckel et al. [25], TC has attracted tremendous
attention, because it may be an effective and economical method to improve the interfacial adhesion between
polymer matrix and ber. To date, TC has been reported
to occur at the interface between several semicrystalline
polymers, such as polypropylene (PP) [2634], polyamide
(PA) [35,36], polyethylene (PE) [37], poly (phenylene sulde) (PPS) [38], poly (ether ether ketone) (PEEK) [3941],
poly(l-lactide) (PLLA) [4245], etc., and various llers with
high aspect (length/diameter) ratio, such as glass bers (GF)
[35,4648], carbon bers (CF) [49], aramid ber, natural
bers (NF) [34], carbon nanotubes (CNTs) [50], talc [51],
copper [52], aluminum [53], etc., as summarized in Table 1.
Here, we will briey summarize the recent studies on TC,
as there is already a comprehensive review on this topic
[54].
Among the llers cited previously, glass ber (GF) is
quite attractive as a reinforcement material for a polymer matrix due to its high mechanical properties, very low
cost, good heat resistance, and high electrical resistivity.
Therefore, TC on a GF surface has been widely investigated during the past few decades. Generally, it is very
difcult to induce TC on a GF surface under common, quiescent crystallization conditions, since the structure of the
GF is completely amorphous. Hence a few methods such as

Fig. 2. (a) The SEM image and (b) PLM image of a typical transcrystalline
structure in iPP/CNT ber composite [24] ( 2008 Elsevier Ltd).

modifying GF surface by nucleating or coupling agents, fast


cooling the samples to generate thermal stress, or exerting an external stress eld on polymer/GF interface during
crystallization, have been proposed and successfully utilized to obtain a TC structure on GF surface, as summarized
in a review paper [54]. In recent years, there were still some
studies focused on TC structure at polymer/GF interface
obtained by using the above methods [5557]. For example, Zhou and coworkers [56] reported that TC appeared
on GF surface when adding 10 wt% MAPP in a PP matrix
and simultaneously modifying GF with a common silane
coupling agent or a di-block copolymer coupling agent.
However, TC could not be induced at the interface when
the GF were treated with a kind of tri-block copolymer
coupling agent, even under fast cooling or shearing condition. The reason may be that the exible interlayer formed
by the tri-block copolymer could relax even under a stress
eld [56]. Furthermore, Zheng and coworkers [58] reported
that the acid-corroded GF exhibited anomalous nucleating ability to induce ringed alpha nuclei, and beta form TC
could be developed unexpectedly from these nuclei during
isothermal crystallization, which provides a new approach
to induce TC on a GF surface [58].
Compared with GF, CF is very expensive, but it has
an outstanding weight-to-strength ratio. In terms of this
aspect, polymer/CF composites are considered as the most
attractive material that could be produced in appreciable quantities. In the past few decades, the interfacial
TC structure in polymer/CF composites has been widely

N. Ning et al. / Progress in Polymer Science 37 (2012) 14251455

1429

Table 1
Summary of different types of llers that can induce TC structure in various crystalline polymers.
Filler*

Polymer

Refs.

GF
CF
Aramid ber
NF
NF
NF
NF
NF
NF
NF
NF
CNTs bers
Talc
Copper
Aluminum

iPP, PA 66, PA-6, PPS


PEEK, PA 66, iPP, sPP, sPS
iPP, PA 66, PA-6, PEEK
iPP
iPP
iPP, PLA
HDPE, PLA, PP
iPP, PLA
iPP
iPP
PHBV
iPP
iPP
iPP
iPP

[35,38,4648,5558]
[3941,5962]
[29,6573]
[7476]
[77]
[45,7880]
[37,44,81]
[34,43,83]
[82]
[84]
[85]
[24,50,93,94]
[51]
[52]
[53]

Cellulose
Cotton
Flax
Sisal
Jute
Wood
Kenaf
Hemp

GF, glass ber; CF, carbon ber, NF, natural ber; CNT, carbon nanotube.

investigated. For example, CF is often used to induce TC


for PEEK because it shows good nucleation ability for this
polymer [54]. CF has also been used to induce TC for other
matrixes, such as PA 66 [59], iPP [60], sPP [61] and sPS [62].
Regarding polymer/CF composites, it is quite interesting
that the high modulus CF (HMCF) generally shows much
better nucleation ability than high strength CF and high
tenacity CF. Therefore HMCF is usually used to induce TC
in these polymer composites, as reported in some studies
[60,62]. In recent years, less work has been focused on TC
structure at polymer/CF interface, with only a few studies
on the interfacial TC structure of semicrystalline polymers
on the surface of vapor grown carbon bers (VGCF) as a
ller [63,64].
Furthermore, aramid ber has the characteristic of
extremely high tensile modulus/high strength, and thus
become very popular as a reinforcing ber for polymer materials. A considerable amount of work has also
been focused on the interfacial TC structure of aramid
ber/polymer composites in the recent decade [29,6573].
For example, Assouline et al. [29] reported that TC of
gamma orthorhombic iPP could be induced on aramid
bers under high pressure. It was determined that the c
axes of the gamma iPP lamellae (the growth axes) were
distributed radially about the ber and that the lamellar
ab face was randomly oriented on the ber surface [29].
Feldman et al. [69,73], found that a double TC layer of PA
66 was generated on the surface of treated aramid (Kevlar
49) ber using a saturated aqueous bromine solution, in
which the lamellar a* axis is nearly perpendicular and at
an angle of about 12 to the ber in the outer and inner
TC layer, respectively. By comparison, only one regular TC
layer could be induced by pristine aramid ber under the
same processing conditions.
Various natural bers (NF) have received increasing
attention as reinforcement materials for polymer composites due to their various attractivities, such as low
cost, low density, high specic strength and modulus,
biodegradability and derivation from renewable resources.
A large number of studies have been focused on the investigation of TC structure in various NF (cellulose [7476],
cotton [77], ax ber [45,7880], sisal [37,44,81], jute
[82], wood [34,43,83], kenaf [84], hemp [85]) reinforced

polymer composites. In recent years, NF has attracted


much more attention due to its environmental benets.
For example, it was reported that TC of iPP could be
induced on cellulose nanocrystal surfaces [74]. In addition, Joseph et al. reported that the interfacial free energy
difference for nucleation of PP on cotton ber is smaller
than that in the bulk PP, which favors the formation and
growth of TC. Furthermore, ber surface roughness and
thermal stresses facilitated the growth of TC on cotton
ber [77]. On the other hand, sisal bers untreated or
treated with alkali or silane all had a nucleating ability to induce TC in PLLA matrix [44]. Hermida and Mega
reported that hemp bers could act as a heterogeneous
nucleation agent for a biodegradable semicrystalline polymer poly(hydroxybutyrate-co-hydroxyvalerate) (PHBV)
and facilitate the formation of TC structure at the
PHBV/hemp ber interface [85]. They reported that the
growth rate of TC was the same as that of spherulites in the
bulk, indicating that nucleation and crystal growth were
two independent processes, as has been reported in some
studies [86,87].
CNTs, with large aspect ratio, extremely high strength,
stiffness and exibility, have been considered as a highly
attractive reinforcing agents in a polymer matrix for
future structural materials. Furthermore, CNTs show
very good nucleation ability for many kinds of polymer
matrix. Therefore, the interfacial crystallization behavior of
polymer/CNTs nanocomposites has attracted tremendous
attention in recent years. For many crystalline polymer systems, CNTs can induce a unique NHSK structure
(see Section 2.2) [8892]. However, Zhang et al. found
that a typical TC of iPP could be induced by CNTs ber
under appropriate condition [24]. Microstructure analysis showed that CNT ber could induce the growth of
both and TC, and TC dominated at the iPP/CNT
ber interface region [24]. Furthermore, TC of iPP was
also observed around the individual CNTs during quiescent melting crystallization of iPP/CNT nanocomposites
formed by inltrating iPP into nanotube aerogel bers.
In this case, the -form TC of iPP dominated the overall interfacial crystalline morphology, perhaps due to the
geometric connement of CNTs with the very small intertube spacing (10100 nm) [93]. Moreover, in the work of

1430

N. Ning et al. / Progress in Polymer Science 37 (2012) 14251455

Fig. 3. (a) TEM image of a transcrystallinity structure of -iPP around individual MWNT in ultrathin lms of iPP/MWNT composites, (b) corresponding
selected-area electron diffraction pattern and sketch indicating the (h k l)-indices of the Bragg-reection; only (h k 0) reections are presented. (c) Sketch
explaining the possible nucleation mechanism of iPP on the surface of a CNT, iPP macromolecules initially are partly wrapped around the CNT (brighter
macromolecule in the sketch, dark rod represents a CNT) and form a nuclei with crystallographic c-axis perpendicular to the long axis of the CNT [50].
2008 American Chemical Society.

Loos and coworkers [50,94], TC of -iPP around individual


CNTs was obtained in ultrathin lms of iPP/CNTs composites, as shown in Fig. 3(a). The corresponding selected-area
electron diffraction (ED) pattern of transcrystalline grown
iPP lamellae around CNTs and sketch indicating the (h k l)indices of the Bragg-reection are shown in Fig. 3(b). Only
(h k 0) reections such as 0 4 0, 0 6 0, 1 1 0, 1 3 0, 2 0 0, and
2 2 0, of the -phase were present, indicating that the caxis orientation (chain orientation) of the transcrystalline
iPP lamellar crystal was perpendicular to the substrate
plane and perpendicular to the long axis of CNTs. This
was obviously different from the assumptions proposed in
other studies, and the authors ascribed this to iPP macromolecules wrapped around rather than aligned along the
CNTs, leading to the formation of a nuclei with crystallographic c-axis perpendicular to the long axis of the CNT,
as schematically shown in Fig. 3(c). However, in our opinion, there is another possibility: the molecular chains of iPP
are oriented parallel to the long axis of CNT near the CNT
surface, while they will twist by 90 at a distance away
from the CNT surface due to the self-epitaxy effect, as has
been reported in many studies on iPP crystallization in the

presence of bers [28,95,96]. Nevertheless, the iPP lamellae


with molecular chains aligned along the CNT surface may
be too thin or too small, thus the corresponding ED pattern
is too weak to be detected. More careful study should be
carried out to better understand the formation mechanism.
2.2. Hybrid shishkebab (HSK)
The hybrid shishkebab (HSK) structure is another
attractive and widely investigated interfacial crystal morphology, rst observed by Thierry et al. in 1990 [97].
So far, the HSK structure has been obtained in various cases such as solution crystallization [89,92,98104],
physical vapor deposition [90,105], solvent evaporation
and thin-lm crystallization [106], in situ polymerization [107,108], injection molding [109], and melt
spinning [110], at the interface between various semicrystalline polymers such as PE [98], PE block copolymer
[102,111] polyamide-6 [112], nylon 66 [89,91], polyamide11 [104], poly(vinyl alcohol) (PVA) [92], poly(vinylidene
uoride) (PVDF) [90], poly(butylene terephthalate) (PBT)
[113,114] and inorganic or organic llers such as

N. Ning et al. / Progress in Polymer Science 37 (2012) 14251455

1431

Table 2
Summary of various crystalline polymer/ller systems that can form HSK structure using various preparation methods.
Filler

Polymer

Preparation methods

Refs.

CNTs
CNTs
CNTs
CNTs
CNTs
CNTs
CNTs
CNTs
CNTs
CNTs
CNTs
CNTs
CNTs
CNTs
Whiskers
Whiskers
DBS ber
Clay

HDPE
HDPE
HDPE
HDPE
HDPE
HDPE
UHMWPE
LLDPE
PE block copolymer
Nylon 66
PVA
PA11
PVDF
PBT
HDPE
LLDPE
PE
Nylon-6

Solution crystallization
Physical vapor deposition
Solvent evaporation and thin-lm crystallization
In situ polymerization
Injection molding
Melt spinning
Melt compounding followed by extrusion-stretching
Injection molding
Solution crystallization
Solution crystallization
Solution crystallization
Solution crystallization
Physical vapor deposition
Compression molding
Injection molding and melt spinning
Injection molding
Physical vapor deposition
In situ polymerization

[89,92,98104]
[90,105]
[106]
[107,109]
[109]
[110]
[124]
[202]
[102,110]
[89,91]
[92]
[104]
[90]
[113,114]
[119121,128]
[132]
[118]
[122]

CNTs [89,90,98,101,102,104106,109111,113,115117],
1,3:2,4-dibenzylidene glucitol (DBS) ber [118], whisker
[119121], clay [122], etc., as summarized in Table 2.
For example, Thierry et al. investigated the crystallization
behavior of polyolen induced by DBS ber and observed
that the polyolen crystalline lamellae epitaxially grew on
the surface of DBS ber [118]. This epitaxial crystalline
structure is actually HSK structure, in which DBS ber
served as shish and polyolen crystal lamellae formed the
kebab [97,118]. Bai et al. [123], reported a special kind of
nucleating agent was used to enhance the crystallization
of PLLA. It was found that the nucleating agent could be
self-organized into ne brils prior to PLLA crystallization.
These brils could then serve as shish to induce the growth
of PLLA lamellar (kebab-like structure) approximately perpendicular to the long axis of the brils. Maiti and Okamoto
[122] reported that a so-called shishkebab superstructure
could be formed in nylon-6/clay nanocomposites prepared
by in situ polymerization method, in which clay acted as
shish and nylon-6 lamellae formed the kebabs.
Recently, a novel nanohybrid shishkebab (NHSK)
structure with CNT acting as shish and polymer crystal lamellae forming kebab has attracted much attention
from many researchers due to its potential use in periodical functionalizing CNTs. This NHSK superstructure was
rst observed by Li et al. in PE and Nylon 66 solution
crystallization in the presence of SWNT, MWNT, as well
as carbon nanober (CNF) [89,98]. A typical NHSK structure is shown in Fig. 4, where it can be clearly observed
that inorganic ller (CNT) with a high length to diameter ratio acts as shish and induces polymer (PE) crystal
lamellae (kebab) periodically decorating its surface and
aligning approximately perpendicular to its long axis. The
thickness of the kebab and the periodicity of the polymer
lamellae can be readily controlled by varying crystallization conditions. It was also reported that the NHSK
structure could be obtained with selected crystalline block
copolymers, such as polyethylene-b-poly(ethylene oxide)
(PE-b-PEO), periodically decorated along CNTs, which led
to amphiphilic, alternating patterns [111]. Furthermore,
a unique 2d NHSK structure with periodic polymer (PE,

Nylon 66, PVDF and poly(l-lysine)) oligomer patterning on


individual CNTs has been achieved using a facile physical
vapor deposition method by Lis group [90,105]. Recently,
a simple and rapid yet effective approach to produce the
NHSK structure in PE/CNTs system using solvent evaporation and thin-lm crystallization was reported [106].
Rastogi and coworkers [124] sprayed an aqueous solution of SWNTs directly onto ne UHMWPE powder and
then melt compounded the CNTs-coated UHMWPE and
extrusion-stretched the composite into a sheet to obtain
a ne NHSK structure with SWNTs acting as shish and
UHMWPE forming the kebab. More recently, Bucknall
and coworkers [125,126] reported that CNT ber based
nanocomposites with aligned HSK nanostructures could
be obtained by inltrating CNT arrays into PE/p-xylene

Fig. 4. A typical NHSK structure (TEM image and schematic representation) in PE/MWNT system [89].
2006 Wiley Periodicals.

1432

N. Ning et al. / Progress in Polymer Science 37 (2012) 14251455

Fig. 5. (a) SEM images of NHSK structure in PE/inorganic whisker (SiO2 MgOCaO whisker) composites obtained by injection molding [128] (Copyright
2011, Elsevier Ltd.); (b) and (c) SEM images of NHSK structure in PE/CNTs composites using (b) injection molding [109] (2009 American Chemical Society)
and (c) melt spinning method [110].
2010 American Chemical Society.

solutions and controlling the isothermal crystallization


temperature and crystallization time. Seo and coworkers [107] observed a NHSK structure with high density
polyethylene (HDPE) crystal lamellae growing perpendicular to MWNT surface in HDPE/MWNTs nanocomposites
prepared by in situ polymerization method. Furthermore,
Xu and coworkers successfully obtained a typical NHSK
structures in PE/CNTs [100,101], PE-b-PEO/CNTs [102] and
PVA/CNTs [92] via solution crystallization with the assistance of supercritical CO2 (SC CO2 ) as an anti-solvent. In
that work, the solubility of the polymer matrix in the
solvent could be varied by the introduction of SC CO2 ,
thus the NHSK structure, the size of the lamellae, and the
interval between them along the CNT was controlled by
varying a series of experimental conditions such as different solvents, polymer concentration, CNTs concentration,
and supercritical CO2 pressure.
The methods described above to obtain the NHSK superstructure were primarily based on solution crystallization,
physical vapor deposition, or in situ polymerization. While
the NHSK structure was obtained, these methods are not
readily extendable for large-scale industrial processing.
With this impetus, signicant effort has been devoted in
our group to obtain a hybrid crystalline structure via scalable industrial processing method. For example, the NHSK
superstructure (see Fig. 5) has indeed been observed for the
rst time in PE/inorganic whisker and PE/CNTs composites
using both injection molding and melt spinning methods
[109,110,119121,127,128]. In these situations, the formation of NHSK structure could potentially bring signicant
mechanical reinforcement in these composites due to the
enhancement of interfacial adhesion between polymer and
CNTs (whiskers) caused by the formation of interfacial HSK
structure. This is discussed in more detail in Section 5.

brous-shaped llers served as shish and polymer


spherulites served as calabash. A typical HSC structure
in iPP/CF composite is shown in Fig. 6(a) [30]. It may be
clearly observed that this HSC structure is different from
the HSK structure, for the polymer crystals in HSK are
lamellae, while they are spherulites in HSC. A schematic
representation of the formation process of the HSC structure is represented in Fig. 6(b) [129]. In this case, the
nucleation density is much less for the formation of HSC
structure than that for the formation of HSK structure. The
brous-shaped ller can only initiate a few nuclei, which
develop into large polymer spherulites without hindrance,
and nally guided by the brous-shaped ller, form the
peculiar HSC structure. It should be noted that this kind
of hybrid crystalline structure was called a TC structure in
some studies [30,130]. However, we think that it is actually
different from TC, since in a TC structure, the high density of active nuclei on the ber surface will hinder the
full extension of spherulites and therefore force the crystal growth in one direction, while in the HSC structure,
only a few nuclei are induced and can fully develop into
large separated polymer spherulites on the surface of a
ller without hindrance. HSC structure has been observed
in single berpolymer system [30]. In addition, a HSC
structure with MAPP spherulites serving as calabash and
ceramic bers acting as shish has also been reported for
a PP gel system [131]. Furthermore, this interfacial crystalline morphology has also been successfully obtained in
the injection-molded bar of iPP/SiO2 MgOCaO whisker
(SMCW) composites and linear low density polyethylene
(LLDPE)/SMCW composites, as shown in Fig. 6(c) and (d)
[129,132].
2.4. Other hybrid crystalline structures

2.3. Hybrid shishcalabash structure (HSC)


In addition to TC and HSK, there exists another kind
of hybrid shishcalabash (HSC) structure, in which

Besides TC, HSK and HSC, many other kinds of hybrid


crystalline structure can be induced on various kinds of
llers, as reported in a number of studies [133136]. For

N. Ning et al. / Progress in Polymer Science 37 (2012) 14251455

1433

Fig. 6. (a) A typical HSC structure in iPP/CF composite [30] (Copyright 1999, Elsevier Ltd.); (b) a schematic representation of the formation process of the
HSC structure [129] (Copyright 2009, Elsevier Ltd.); (c) and (d) HSC structure obtained in injection-molded bar of (c) iPP/SMCW composites [129] (Copyright
2009, Elsevier Ltd.) and (d) LLDPE/SMCW composites [132].
2011 Wiley Periodicals.

example, Prokhorow and Nitta recently reported that the


interfacial crystalline morphology of ultrahigh molecular
weight polyethylene (UHMWPE) on mica was small, single
molecule rod-like nanocrystallites and isolated block-type
edge-on nanolamellae comprising several PE molecules,
whereas it was isolated lamellae and lamellar domains of
a monolayer height of UHMWPE on the surface of graphite
[133]. In addition, Brinkmann et al. reported that P3HT

could be epitaxially grown on the surface of an aromatic


salt (potassium 4-bromobenzoate) (K-BrBz), which led to
highly oriented and nanotextured P3HT lms with a regular network of interconnected semicrystalline domains
oriented along two preferential in-plane directions [134].
Moreover, Tracz and coworkers reported that the interfacial crystalline morphology of PE on highly oriented
pyrolytic graphite (HOPG) and molybdenite (MoS2 ) was

1434

N. Ning et al. / Progress in Polymer Science 37 (2012) 14251455

similar to the fractured surface of extended-chain crystals


of PE crystallized under high pressure [137,138].
3. The factors controlling the hybrid crystalline
structure
Hybrid crystalline structure in polymer/ller composites has been investigated for more than half a century.
Many factors, such as surface chemical and physical
characteristics of the llers, molecular weight, chain conformation and functional groups of polymer matrix, and
various kinds of external elds will affect the formation
and growth of interfacial crystalline structures, which have
attracted tremendous attention. Here we summarize the
most important progress in this area.
3.1. Effect of llers
3.1.1. Surface chemical and physical characteristic of
llers
It is well known that the intrinsic surface chemical and
physical characteristics of various llers have a signicant
effect on the nucleation activity of a given polymer matrix,
and can signicantly affect the interfacial crystallization of
polymer/llers composites. First of all, the specic surface
chemical structure of ller can play a crucial role in the formation of hybrid crystalline structure at the polymer/llers
interface. For example, Maiti and Okamoto [122] investigated the unique crystallization behavior controlled by
silicate surfaces in nylon-6/clay nanocomposites and found
that the silicon hydride chemical bonds on the surface of
clay and amino bonds of nylon-6 could form hydrogen
bonding, which facilitated the formation of a pseudohexagonal lamellar packing of nylon-6 on both sides of clay. As
a result, the exclusive formation of the -phase of nylon-6
grown on the surface of clay particles was observed; and
a so-called shishkebab structure with clay acting as shish
and nylon-6 -phase lamella as kebab was obtained.
Second, the crystal lattice structure of the llers can
play a key role in the formation of interfacial crystal morphology of polymer/ller composites. For example, Hobbs
investigated the nucleation ability of two types of polyacrylonitrile (PAN) based CF with different graphitic structures
for iPP matrix and the interfacial crystalline behavior of the
composites [139]. The results showed that CF composed of
small (about 2.5 nm) graphitic nuclei with a high degree of
disorientation acted as a very poor nucleating agent for iPP
and almost no interfacial crystallites could be observed at
the iPP/CF interface. However, CF composed of much larger
(>10 nm) graphitic planes with a high degree of orientation
displayed a strong nucleation ability for iPP crystallization
and resulted in a thick TC layer along the whole ber axis.
The extreme variation of nucleation ability of these two CFs
on iPP could be ascribed to the difference in graphite crystal lattice structure. Recently, Li et al. [105] investigated
the impacting of side-wall structure of CNTs on the formation of NHSK in the PE decorated CNTs situation. They
observed that PE crystal rods could periodically decorate
both CVD-MWNTs and AD-MWNTs using solution crystallization, but the PE crystal lamella on CVD-MWNTs surface
was not as uniform as those on the AD-MWNTs. In some

locations, CVD-MWNTs were only partially decorated by PE


crystal lamella. However, when using solvent evaporation
and thin-lm crystallization, only AD-MWNTs could induce
NHSK due to relatively fast solvent evaporation [106]. This
could be due to the crystal lattice defect of the CVD-MWNT.
Obviously, these results demonstrated that the crystal lattice structure of llers played a crucial role in the formation
of hybrid crystalline structure.
Third, it has been well established that surface roughness of llers also plays a very important role in the
formation of interfacial crystallization structure. For example, Wang and Liu [30] observed an inverse relation
between induction time and nucleation rate for polytetrauoroethylene (PTFE) ber and carbon ber systems;
whereas this inverse relation was not applicable to Kevlar
and poly(ethylene terepthalate) (PET) bers. This was due
to the non-uniformity of surface roughness of Kevlar and
PET bers. Furthermore, Lin et al. [53] also investigated
the effect of surface roughness of copper (Cu) sheet on the
heterogeneous nucleation and interfacial crystallization
behavior of iPP at the iPP/Cu interface, and observed that
Cu surface with higher surface roughness will induce more
nuclei of iPP, resulting in a thicker TC layer in the interfacial region upon supercooling over a certain temperature
range. This result was consistent with the investigation of
aluminum and talc surfaces [52,53], which showed that
surface roughness, instead of chemical factors or surface
energy, played a dominant role in interfacial crystallization at the polymer/ber interface. Joseph and coworkers
[77] also reported that the surface roughness of cotton ber
could facilitate the growth of TC.
Fourth, surface treatment of llers is an important
method to change the surface characteristics of llers
and is widely used to improve the interfacial adhesion
of polymer/ller composites. For semicrystalline polymer
composites, surface treatment of llers could also be used
to change the nucleation ability of the llers, and control
the interfacial crystallization behavior of composites. For
example, Li et al. [105] investigated the effect of surface
modication of MWNT by octadecylamine (C18-MWNT) on
the crystallization behavior at the PE/CNTs interface. The
results showed that no interfacial PE crystallization was
induced on the C18-MWNT surface, indicating that alkanemodication of the MWNT surface prohibited the PE single
crystal growth on the CNTs. Therefore, the authors concluded that a uniform, smooth, graphene-like surface was
preferred for PE crystal formation on the surface of CNTs.
Mondragon and coworkers [140] investigated the effect of
ber surface treatments on interfacial crystallization of ax
ber/PP composites. It was observed that TC of iPP could
be induced by both untreated and MAPP-treated ber at
the isothermal crystallization temperature of 134 C, but
MAPP-treated ber composites showed lower TC density
than composites with untreated bers due to the better
nucleation ability of untreated ber on PP crystallization.
With increasing crystallization temperature to 140 C, the
untreated ber still induced TC of iPP whereas MAPPtreated ax ber did not. These results indicated that the
interfacial crystal morphology could be strongly inuenced
by surface treatment of the llers. On the other hand, for
some ller with very poor nucleation ability, appropriate

N. Ning et al. / Progress in Polymer Science 37 (2012) 14251455

1435

Fig. 7. (a) The typical NHSK structure in PE/CNT (with a diameter of about 10 nm) system with PE single crystal lamellae perpendicular to CNTs axis; (b)
PE crystal lamellae are randomly decorated on CNF surface in PE/CNF (with a diameter of about 300 nm) system [89].
2006 Wiley Periodicals.

surface treatment is necessary to improve the nucleation


ability of the llers to polymer matrix. A typical example
is the iPP/GF composite systems: GF is usually coated with
some nucleation agents to enhance its nucleation ability for
iPP matrix. As reported by Assouline et al. [141], TC of beta
form of iPP could be induced on the GF surface when the
bers were coated with appropriate beta nucleating agent.
Furthermore, the surface topography structure, surface
roughness, thermal conductivity mismatch, and thermal expansion mismatch could also favor heterogeneous
nucleation and has a signicant inuence on interfacial
crystallization of polymer llers composites, which has
been well reviewed in ref. [54].

the larger wetting force. Therefore, the nucleation ability of small-diameter whiskers was increased, resulting
in a more dense interfacial HDPE crystal lamellae grown
on this whisker. Furthermore, Li et al. [89] reported that
the diameter of brous carbon ller played a major role
in the interfacial crystallization morphology. As shown in
Fig. 7(a), the typical NHSK structure in PE/CNT (with a diameter of about 10 nm) system with PE single crystal lamellae
perpendicular to CNTs axis were obtained. However, for
larger diameters of CNF (with a diameter of about 300 nm),
the interfacial PE crystal lamellae on the surface of CNF
shows a variety of orientations, as clearly shown in Fig. 7(b).
3.2. Effect of polymer matrix

3.1.2. Geometry and size of llers


The above describes the surface chemical and physical
characteristics and surface modication of llers on the
interfacial crystallization behavior of polymer/ller composites. It is also well established that the geometry and
size of llers also have signicant inuence on hybrid crystalline structure of polymer composites. As an example, in
our previous work [128] for the same whisker but with
different diameters, the small-diameter whisker showed
better nucleation ability. As a result, signicant interfacial HDPE crystal lamellae were epitaxially grown on the
surface of small-diameter whiskers, while only little HDPE
crystal lamellae decorated the surface of large-diameter
ones. The reason could be that the melt pressure exerted
on the small-diameter whisker was larger, which led to

3.2.1. Molecular weight


The molecular weight of a polymer matrix dominates
the molecular chain mobility of the matrix, and thus can
also affect the crystal growth at the polymer/ller interface.
Therefore, the interfacial crystallization is signicantly
inuenced by the molecular weight of the polymer matrix.
For example, Folkes and Hardwick [142] observed that the
lowest molecular weight PP/single ber system showed the
highest nucleation density and the densest TC layer. With
increasing molecular weight of the PP matrix, the TC layers became less uniform, and even absent on some parts
of the ber. For the highest molecular weight PP system,
almost no TC layer was observed at the PP/ber interface. The above phenomenon could be explained as follows.

1436

N. Ning et al. / Progress in Polymer Science 37 (2012) 14251455

On the one hand, adsorption of the polymer molecules


onto the ber surface (nucleation) is primarily occurred
by attachments of the molecule chain ends. Therefore, the
probability of an attachment increases as the molecular
weight of the polymer decreases due to the increase of
the number of chain ends. On the other hand, the lateral
development of the TC layer (crystal growth) depends on
the chain mobility of the polymer matrix. For low molecular weight polymer, the higher chain end concentration
and higher chain mobility led to higher nucleation density
and the denser TC layer. Similar results were reported in
single GF/PP composites by Moon [143]. Furthermore, in
our previous work [120] the effect of PE molecular weight
on the formation of HSK structure in injection-molded
HDPE/SMCW composites was investigated. An obvious
HSK structure, with PE crystal lamellae closely packed
on the surface of SMCW, was observed in samples with
a relatively low molecular weight PE (LMW-PE) matrix.
However, in samples with a relatively high molecular
weight PE (HMW-PE) matrix, an incomplete HSK structure
with PE crystal lamellae loosely decorated on the surface
of SMCW was observed. Moreover, the molecular weight
dependence of interfacial crystal structure in injectionmolded HDPE/mica composites was also investigated in
our group [144]. The results also showed that a LMW-PE
system can facilitate the growth of TC layer on mica surface,
whereas no obvious TC layer could be detected in HMW-PE
system. These results could also be ascribed to the comparatively high chain mobility and more chain ends of LMW-PE
matrix.

crystal structure was only observed in the injectionmolded HDPE/whiskers composites, whereas a unique HSC
interfacial crystal structure with whiskers served as shish
and polymer spherulites formed calabash was observed
in the injection-molded iPP/whiskers and LLDPE/whiskers
composites [119,120,128,129,132].
3.2.3. Functional groups of polymer matrix
The functional groups of a polymer matrix have the
potential to form some special interactions with the surface of llers, and thus also have a signicant effect on
interfacial crystallization. For example, it was reported
that chemisorption of the polar groups of nylon-6 and/or
nylon 66 onto the graphite surface could lead to the strong
interactions between the macromolecules and the graphite
crystallites, and thus contribute to nucleation and lateral interfacial crystallization [146,147]. Furthermore, for
iPP/GF system, no interfacial TC could be formed on the
amorphous surface of GF under normal quiescent conditions, whereas for polyamide matrix, interfacial TC could be
successfully induced on GF under the quiescent conditions.
This observation could also be due to the chemisorption
of the polar chemical groups ( CO and NH ) of the
polyamide matrix on GF surface [148]. Moreover, as already
mentioned above, the amino bonds of nylon-6 and silicon
hydride chemical bonds on the surface of clay could form
hydrogen bonding, which facilitated the interfacial crystallization of nylon-6 on clay surface [122].
3.3. Effect of external eld

3.2.2. Molecular chain structure and conformation


Molecular chain structure and conformation could dramatically inuence the adsorption of polymer molecules
onto the surface of llers and lateral crystal growth, and
thus can also affect the interfacial crystallization of polymer/ller composites. For example, Zheng and Xu [145]
performed a comparison study of interfacial crystallization behavior of two types of semicrystalline polymers (PE
and PEO) with different molecular chain architectures and
conformations on SWNTs using a solution crystallization
method. For PE, the molecular chains could easily adsorb
onto the SWNT surface and further crystallize to form a
unique NHSK architecture, due to its planar zigzag conformation and favorable physical interactions with SWNTs.
However, in the case of PEO system, SWNTs were wrapped
by a thin amorphous polymer coating, because of the distorted helical conformation and unfavorable interaction
with SWNTs. Typically, the ordered NHSK structure has
been reported in the cases of PE [98], nylon-6 [112], nylon
66 [89,91], and PVA [92]. For these, the chain conformation in the crystal unit-cell of these polymers is either a
zigzag conformation or a planar zigzag conformation. In
the case of iPP with 31 helical conformation, TC instead of
NHSK structure was formed on the surface of CNTs [50,94].
This indicated that molecular chain structure and conformation did have a remarkable inuence on interfacial
crystallization. More evidence for the importance of chain
structure and conformation on interfacial crystallization
comes from our recent work on the crystallization of HDPE,
iPP and LLDPE as induced by whisker. The HSK interfacial

It is well known that external extension/shearing


elds can strongly impact the crystallization behaviors
of semicrystalline polymer. Generally, the crystallization
kinetics can be signicantly promoted with extensional
deformation or shearing ow [19,149], and some highly
anisotropic crystal superstructures, such as shishkebab,
can be easily obtained from the oriented polymer melts
[150]. In the case of polymer/ller composites, the external extension/shearing eld can induce novel interfacial
crystallization superstructure, such as HSK and TC. Many
studies on external shear eld induced TC have been performed in single berpolymer systems [54]. Furthermore,
our group has investigated the effect of shear or extension
elds on hybrid crystal structures (interfacial TC, HSK, or
other epitaxial crystalline structure) of polyolen/brousller composites via melt processing methods, such as
injection molding and melt spinning. By using a so-called
dynamic packing injection molding technology (DPIM), in
which an oscillatory shear eld was imposed on the gradually cooled melt during the solidication stage, we have
successfully achieved a clear-cut TC of iPP on GF using
practical processing approaches [55]. In contrast, no TC
could be observed for the sample prepared via traditional
injection molding. Obviously, shear plays a key role in the
formation of TC in the injection-molded bar of iPP/GF composites. Furthermore, a direct formation of HSK structure,
in which brous whiskers served as shish while PE lamellae acted as the kebab, has been successfully obtained in
the injection-molded bar of HDPE/whiskers composites via

N. Ning et al. / Progress in Polymer Science 37 (2012) 14251455

1437

Fig. 8. (a) Schematic of the anti-solvent process of SC CO2 and the growth of PE crystal lamellae on CNTs; (b) and (c) TEM images of PE/MWNT NHSK
structure formed in the same PE concentration and MWNT concentration, but at different SC CO2 conditions: (b) 120 C/9 MPa; (c) 100 C/9 MPa, for 3 h
[100].
2007 American Chemical Society.

the above-mentioned DPIM technology. Nevertheless, for


samples prepared via common injection molding, the PE
crystal lamella were randomly decorated on the surface of
the whiskers [119]. This indicates again that external shear
eld plays a key role in the formation of the interfacial
crystal structure. Moreover, in the case of PE/CNTs systems
[127], a hierarchical structure with changed PE lamellar
orientation along the thickness direction of molded bar
prepared by DPIM technology was detected. The lamellae
were randomly distributed on the surface of CNTs in the
skin and the core layers. However, in the sheared layer,
the typical NHSK structure with disk-shaped lamellae periodically decorated along the CNTs axis was developed.
This could be due to the combined effect of the external
shearing eld and the temperature eld. Furthermore, the
typical HSK structure could also be successfully induced
in PE/CNTs (SMCW) composites by stretching during melt
spinning [110,121].

The preceding describes the effect of external shearing


(stretching) elds on the formation of interfacial crystal
morphology under melt processing conditions. For solution
crystallization, the solvent and temperature usually play
a key role in the formation of interfacial crystal structure.
Furthermore, supercritical CO2 (SC CO2 ) also has an obvious
effect on the formation of interfacial crystal morphology in
solution systems, as schematically shown in Fig. 8(a). Xu
and coworkers have achieved some interesting results in
this area [100,101]. They found that the nanohybrid crystal structure at polymer/CNT interface in solution systems
could be successfully adjusted by changing the solvent,
temperature and pressure of SC CO2 [100,101]. For example, at the same SC CO2 pressure, the average PE lamellar
crystals (kebab) diameters, thicknesses and periodicities
decorated on the surface of CNTs all showed an obvious
increase as the SC CO2 temperature decreased from 120 C
to 100 C, as shown in Fig. 8(b) and (c). This could be due to

1438

N. Ning et al. / Progress in Polymer Science 37 (2012) 14251455

the stronger anti-solvent effect of SC CO2 under relatively


lower temperature. In this situation, more PE molecules
could be separated and absorbed onto the surface of CNTs,
which favored the formation of larger lamella (kebab) size.
On the other hand, with the pressure of SC CO2 increasing
from 9 MPa to 11 MPa, the amount and speed of CO2 dissolved in the p-xylene solvent increased and the solvent
power of p-xylene decreased. Therefore, more PE could
be deposited on the CNT surface and the diameter of the
kebab increased. However, with further increasing the
CO2 pressure to 13 MPa, both the amount and rate of the PE
precipitation were greatly increased and could obviously
increase the nucleation number, leading to the formation
of more PE lamellae (kebab).
Furthermore, they also investigated the effect of solvent on the formation of interfacial crystal structure [151].
For PVDF/SWNTs systems, no interfacial crystallization
occurred on the surface of CNTs, when using DMF and
DMAc as the solvent. Nevertheless, the helical structure of
PVDF wrapping on CNTs surface was obtained when using
DMSO as the solvent. Moreover, for PE-b-PEO/CNTs system, the regular NHSK structure could be formed on CNTs
surface, when DCB and p-xylene were used as the solvent
[102]. This could be due to the favorable molecular interactions between the CH groups and systems for PE-b-PEO
and CNTs in DCB and p-xylene, to facilitate the adsorption
of molecular chains on CNTs and the lateral nucleation and
interfacial crystal growth. By contrast, using DMAc as the
solvent, which is more selective for PEO, only thin polymer
coatings were observed on the CNTs, ascribed to the fact
that the PE block was in poorly solvated, perhaps globular
state in DMAc, suppressing absorption of PE segments on
the CNTs. On the other hand, the unfavorable OH interactions between PE-b-PEO molecules and CNTs in DMAc
could hinder the nucleation of PEO on CNTs.
4. Formation mechanisms of hybrid crystalline
structure
It is well known that the enhanced nucleation on the
surface of llers plays a key role in the formation of various
hybrid crystalline structures. Since the 1950s, the origin
of interfacial crystallization has been extensively investigated. Nevertheless, the formation mechanism for each
kind of hybrid crystalline structure is still not quite clear.
Furthermore, to date, little work has been focused on developing a universal formation mechanism for these hybrid
crystalline structures. Typically, the proposed formation
mechanisms are all associated with the enhanced nucleation on the surface of the llers and may be summarized
as follows.
4.1. Epitaxy and soft epitaxy
Many kinds of llers demonstrate a heterogeneous
nucleating effect on semicrystalline polymer matrix and
can induce lateral interfacial crystallization. This is often
due to specic surface physical and chemical characteristic of these llers and polymers. Epitaxial crystallization
of polymer matrix on the surface of these llers is often
observed for many crystalline llers, and is widely believed

to be the formation mechanism of interfacial crystallization. As we know, epitaxy is most generally dened as
the oriented overgrowth of one phase (guest crystal) on
the surface of a crystal of another phase (host crystal)
[152,153]. The epitaxial match could be due to a unit-cell
dimensional match or crystal structure similarity between
the crystalline ller and polymer matrix. Pioneering work
regarding polymer epitaxial crystallization on llers was
rst carried out in 1958 by Willems and Fischer [153,154].
Since then, it has been widely reported that polymer interfacial crystallization on the surface of various kinds of llers
(such as graphite [133,138,154], talc [51,155157], mica
[133], pitch-based CF [139] and so on) obeyed the epitaxial relationship. For example, the dimension of graphite
crystal unit-cell lattice (2.46 A) is very close to the dimension of the PE crystal cell along the c-axis (2.55 A), which
could result in the epitaxial crystallization of PE on the surface of graphite, with the PE chain direction (the <0 0 1> of
the PE crystal) parallel to <2 1 10> direction of graphite
[133].
However, for some llers with a very small diameter,
such as CNTs, the formation mechanism of interfacial crystallization is soft epitaxy (geometric connement). In this
case, the epitaxial mechanism no longer exists. This is due
to the curvature of small-diameter CNTs. If the interfacial
crystallization of polymers on CNTs still obeys the epitaxial
relationship, the curved surface of CNTs would lead to curve
polymer crystals with distorted lattice, which presumably are not stable. Therefore, the interfacial crystallization
mechanism of polymers on the surface of CNTs is geometric
connement (soft epitaxy) as proposed by Li et al. [89].
Furthermore, for some llers unique surface topography
structures such as edge planes or grooves, also results in
a geometric connement effect on the polymer molecules
absorbed on the surface, and the formation mechanism of
interfacial crystallization could also be concluded as soft
epitaxy, as observed in many cases [148,158].
To explain clearly the mechanism of epitaxy and soft
epitaxy, a schematic representation is shown in Fig. 9. For
two dimensional (2d) lamellar llers (see Fig. 9(a)) or bers
with a diameter much larger than radius of gyration (Rg ) of
the polymer (see Fig. 9(b)) [88], the interfacial crystallization of polymer matrix on such llers may strictly obey the
epitaxial relationship, as shown in Fig. 9(a) and (b). At rst,
the polymer molecular chains will adsorb onto the surface
of the 2d lamellar llers or bers due to physical interactions, then these molecular chains will be orientated and
nucleated on the surface of llers obeying epitaxial relationship. As a result, the various crystal orientations on the
surface of llers lead to different orientations of the polymer crystal lamellae. However, for llers with a diameter
similar to the polymer Rg , such as CNTs, as shown in Fig. 9(c)
the polymer chains are exclusively parallel to the long axis
of these kinds of llers due to the geometric connement.
As a result, a novel NHSK structure with polymer crystal lamella perpendicular to the long axis of such llers is
obtained, as described in Section 2.2. Regarding the soft epitaxy mechanism in semicrystalline polymer/CNTs systems,
there is another possibility: CNTs rst induce a homogeneous coating of polymer chains on its surface with few
subglobules, then the polymer chains expand from these

N. Ning et al. / Progress in Polymer Science 37 (2012) 14251455

1439

Fig. 9. A schematic representation of epitaxy and soft epitaxy: (a) the interfacial crystallization process of two dimensional lamellar llers obeying epitaxy
mechanism; (b) bers with a diameter much larger than the radius of gyration (Rg ) of the polymer obeying epitaxy mechanism; and (c) llers with a
diameter similar to the radius of gyration (Rg ) of the polymer obeying soft epitaxy mechanism.

subglobules and form lamellar crystals perpendicular to the


CNT axis, as reported in some studies [99,145].
4.2. Chemisorption
The chemisorption of atomic hydrogen, carbon, nitrogen and oxygen on the surface of llers has also been
proposed as a nucleating and interfacial crystallization mechanism [146,147]. This interfacial crystallization
mechanism generally works in polar polymers and
inorganic llers system [146,148]. In this case, the
chemisorption of the polar groups of a polymer matrix
onto the inorganic llers surface could lead to strong interactions between the macromolecules and the inorganic
llers, and thus facilitate the nucleation and lateral interfacial crystallization. A typical example of the chemisorption
induced interfacial crystallization is shown schematically
in Fig. 10. It may be seen that the chemisorption of the
amino groups of PA-6 onto the surface of clay (with silicon
hydride groups) could result in a formation of hydrogen
bonding, to stabilize the molecular orientation and facilitate nucleation. As one molecular layer is nucleated on the
clay surface, other molecules could form hydrogen bonds
to the molecules already attached on the clay surface. As a
result, the discrete lamellar structure on both sides of the
clay is formed [122].

4.3. Stress or strain induced interfacial crystallization


Enhanced nucleation on the surface of llers is the key
factor for the formation of various hybrid crystalline structures. The nucleation ability of llers on crystalline polymer
matrix can lead to heterogeneous nucleation and lateral
interfacial crystallization, as described in the preceding.
Stress or strain generated at the polymer/ller interface
could lead to an orientation of macromolecular chains and
a decrease of the nucleation barrier, and thus can also contribute to the enhancement of nucleation on the surface
of llers and the formation of various hybrid crystalline
structures. For example, in some single berpolymer systems, TC was not induced on the ber surface under normal
quiescent conditions due to the poor nucleation activity of
the ber. However, imposition of an external shear eld
to the ber/matrix interface by pulling the ber in the
polymer melt, facilitated the orientation of polymer molecular chains along the long axis of ber, and TC could be
successfully induced on the ber surface. The formation
mechanism of shear induced hybrid crystalline structure
is as following, taking the iPP/GF system as an example. First, the iPP chains will align along the GF long axis
under the shearing effect, followed by the formation of row
nuclei composed of these aligned iPP chains on the GF surface. Dense active nuclei will then be induced following a

1440

N. Ning et al. / Progress in Polymer Science 37 (2012) 14251455

Fig. 10. A schematic representation of the chemisorption induced interfacial crystallization mechanism.

homogeneous nucleation of these row nuclei, promoting


the growth of closely packed iPP crystals from both sides
of GF, to form an interfacial TC structure. It should be noted
here that the homogeneous nucleation (self-epitaxy) of iPP
from those row nuclei could result in the change of molecular chain direction by 80 . Therefore, a unique lamellar
branching structure (cross-hatching) composed of parent
lamellae (whose c-axes are parallel to the ber surface)
and daughter lamellae (whose c-axes are perpendicular to
both the ber surface and the parent lamellae), is generally observed near the ber surface, whereas these lamellae
will twist by 90 at a distance away from the ber surface.
This phenomenon has been often reported in the literature
[24,28,54,95,96].
Furthermore, under real processing conditions, the formation mechanism of external shear or extension eld
induced hybrid crystalline structure is similar to that
described in the preceding. A typical example is the formation mechanism of shear induced NHSK structure in the
injection-molded bar of HDPE/CNTs composites via DPIM
technology [109]. In this case, the formation mechanism
of NHSK structure could be summarized in the following
three steps: (1) shear induced disentanglement and orientation of both HDPE chains and CNTs, (2) formation of PE
folded-chain lamellae directly nucleated on CNTs surface
and/or formation of PE extended-chain shish rst directly
on CNTs surface followed by nucleation of folded-chain
lamellae, and (3) the growth of PE kebabs.
The preceding describes the external stress or strain
induced interfacial crystallization mechanism. The internal stress or strain caused by fast cooling the sample from

melt or thermal expansion coefcient mismatch or thermal conductivity mismatch between llers and polymers
could also induce the pre-alignment of polymer molecular
chains on the llers surface, and thus decrease the nucleation barrier, enhance the nucleation of polymer matrix on
the ller surface, and lateral interfacial crystallization, as
observed in some cases [54,148]. In such cases, the formation mechanism of hybrid crystalline structure is similar
to the above-mentioned external stress or strain induced
interfacial crystallization.
5. Interfacial and mechanical enhancement
induced by interfacial crystallization
5.1. Interfacial enhancement of crystalline interface
estimated by different research strategies
Micromechanical parameters are commonly invoked
to evaluate the magnitude of interfacial interaction and
the level of load transfer efciency, such as interfacial
shear strength (IFSS,  i ) [159161], interfacial adhesion/bonding strength [162], critical effective length (lc )
or critical aspect ratio (lc /D) [163,164]. The assessment
of interfacial enhancement behaviors is directly or indirectly determined for these crucial micromechanical
parameters. Among them, the IFSS is most important
and most commonly used because its value is strictly
proportional to the interfacial interaction and load transfer
ability of the system. Meanwhile, there are a number of
experimental means to identify the occurrence of interfacial enhancement, roughly divided into ve strategies:

N. Ning et al. / Progress in Polymer Science 37 (2012) 14251455

micromechanical tests, such as the pullout [165] and


fragmentation tests [163]; mechanical model prediction
using the role of mixtures [166] or the Halpin-Tsai equations [167]; computational molecular dynamic simulation
[168]; microstructureproperty relation analysis [125];
and in situ spectroscopic analysis such as Raman spectroscopy [169]. As suggested by Coleman et al., the strong
interfacial adhesion and highly efcient load transfer
between ller and polymer are critical requirements for
a ller-reinforced composite are the most requirements
for a ller-reinforced composite [170]. Therefore, the
mechanical behavior of the interface is essential for
optimum reinforcement. It is logically deduced that for
a non-covalent bonding the mechanical strength and
modulus of an ordered crystalline interface is better than
that of amorphous one, since the crystalline interface
layer can bear larger shear stress before its failure and
hence lead to higher load transfer efciency. However, as
discussed in the following, this assumption is not always
observed on assessing the critical parameters such as the
IFSS with the interfacial crystallization structure.
The most straightforward means to evaluate IFSS is
by performing various micromechanical tests. For brousshaped llers, the single ber pullout test was frequently
used, in which a ber was embedded in a polymer matrix
and the force to pullout the ber was measured [161].
According to a general force-balance rule, the maximum
pullout force Fmax is related to the IFSS  i via. According to
a general force-balance rule, the maximum pullout force,
Fmax is equal to the  i , multiplying the ber embedded area
[159]:
Fmax = i 2rf lemb

(1)

where rf is the ber radius and lemb is the ber embedded length. Therefore the linear slope of Fmax versus lemb
or embedded area may be used to estimate  i [165,171].
Gati and Wagner [18] purposely changed the lemb in transcrystalline poly(caprolactone) (PCL) or amorphous PCL
droplets, and performed the pullout test. The  i values
obtained from the plots of Fmax against lemb as shown
in Fig. 11 indicated that the presence of TC offers little impact on the load transfer ability of the interface, as
the IFSS for both transcrystalline and amorphous PCL was
around 3 MPa. In addition, Folkes and Wong [172] found
that the formation of TC in the PP/GF composites leads to
a reduction in the IFSS. A somewhat different result has
been achieved by Wu et al. [173]. They manipulated the
microstructure of TC using PP as matrix, and measured
 i by using the pullout test. Although the highest IFSS
of transcrystalline interface was prominently lower than
the tensile strength of pure PP, the value of IFSS changed
remarkably, from 1 MPa to 5.5 MPa, by altering the structural feature of TC, implying a possibility that TC tailoring
enhanced interfacial adhesion/bonding. The IFSS of PP transcrystalline interface was also measured using the pullout
test in other studies [174,175], to obtain values generally
lower than 5 MPa. Nevertheless, for the cases of strong
interfacial adhesion, bers could be easily occur broken and
fail under external force effect; the fragmentation test is
more appropriate than the pullout test to estimate the IFSS

1441

[164,176,177]. For the ber fragmentation test, the classical


Kelly-Tyson mode [178] is used to determine  i :
i =

f
2(lc /Df )

(2)

where  f is the ber tensile strength, Df the ber diameter


and lc the critical effective length. Since the distribution of
fragment lengths is related to the lc , the mean fragment
length lm is expressed as Klc , where K is a correction factor
and usually assumed to be 3/4 for random orientation of
the bers. Therefore an empirical equation can be derived
as:
i =

3Df f
8lm

(3)

Obviously, precise values of  f are crucial for calculating


the IFSS. In the study of Nagae et al., the interaction between
GF and iPP was enhanced by coating maleic anhybridePP (MA-PP) around GF, and the IFSS was measured using
the fragmentation test [179]. The IFSS was dramatically
improved from 6 MPa to 24 MPa after the formation of
interfacial transcrystalline layer. Wu et al. [61] utilized Eq.
(3) to obtain the IFSS of interfacial transcrystalline layer in
the CF-reinforced sPP and iPP composites, respectively. The
IFSS of sPP/CF, 12.7 MPa, was three times for that of iPP/CF,
4.3 MPa, indicating obviously higher interfacial adhesion
due to the formation of sPP TC structure. This was attributed
to the difference in lamellar orientation behavior of TC layer
between sPP/CF and iPP/CF. The formation of transcrystalline layer coating on CF induced interfacial enhancement
was also observed by other authors [180,181].
The interfacial adhesion/bonding strength is another
type of micromechanical parameter to assess the mechanical enhancement behavior of the crystalline interface.
Chen and Hsiao [17] measured the debonding forces of
the polymer/ber interfaces with or without TC, by using
the microdebonding test. The debonding force may be
improved by 40% with the appearance of a transcrystalline interface. Karger-Kocsis [182] also found that the
appearance of transcrystalline layer enhanced the interfacial adhesion strength. To study the case for which a
transcrystalline layer occurred on a layered substrate, Cho
et al. [183] utilized a cleavage mode test, such as, the
asymmetric double cantilever beam test, to estimate the
interfacial adhesion strength between the TC of PP and silicon substrate. The interfacial adhesion strength increased
dramatically with the perfection of transcrystalline interface, as shown in Fig. 12(a); a microstructural analysis
revealed that for high transcrystalline level the breakage
of interface occurred across the brillation of iPP lamellae in TC, resulting in strong interfacial bonding, whereas
for low transcrystalline level cracking initiated at the weak
boundary layer between TC and spherulites, as schematically presented in Fig. 12(b).
For nanometer-scale bers, such as CNTs, the micromechanical tests of pullout and fragmentation were also
successfully utilized to estimate the IFSS [163,165], interfacial fracture energy [184], and critical ber length [163]
etc. To attain the IFSS in the CNTs-reinforced composites
with the presence of interfacial crystallization, the predictive mechanics models such as the prediction models of

1442

N. Ning et al. / Progress in Polymer Science 37 (2012) 14251455

Fig. 11. Plots of maximum pullout force versus ber embedded length for PCL transcrystalline and quasi-amorphous droplets [18].
1997 American Chemical Society.

mechanics like the Cox-Krenchels rule of mixtures and the


Halpin-Tsai (H-T) equations have been considered [170]. In
a brous-shaped ller-reinforced polymer composite, if an
ideal interfacial bonding exists and the mean length of ber
lm is longer than the critical ber length lc the bers may
take place rupture under tension deformation, as schematically presented in Fig. 13(a). For such a circumstance, the
rule of mixtures is expressed as:
c = 0 1 f Vf (1 Vf ) m

(4)

where  c ,  f and  m are the strength for composite, ber


and polymer matrix, respectively, Vf is the ber volume
fraction, 0 is the orientation efciency factor and l the
length efciency factor. The IFSS can be calculated through
Eq. (2) or Eq. (3) after  f has been identied. For a short
ber reinforced composite for which interfacial interaction

is weak, the failure model is described as ber pullout and


debonding occurs at the polymer/ber interface (Fig. 13(c)).
In this case an alternative of the rule of mixtures is given
as:
c =

 l 
i m

Df

Vf (1 Vf ) m

(5)

From this equation, one may note that the composite strength is determined by the IFSS,  i , and the slope
of linear tting of  c versus Vf is closely related to the
IFSS. For some special cases of CNTs-reinforced composites [185,186], the mechanical strength of interfacial layer
is extremely stronger than that of polymer bulk, and a
polymer sheath coating around the pullout CNTs could be
observed, as schematically shown in Fig. 13(b). The rule of
mixtures is further modied with taking into account the

N. Ning et al. / Progress in Polymer Science 37 (2012) 14251455

1443

Fig. 12. (a) Changes of interfacial adhesion strength with varying level of crystalline structure in the TC region; (b) schematic indicating that a crack
propagates across transcrystallinities or the boundary between transcrystallinities and spherulites [183].
2003 American Chemical Society.

Fig. 13. Three failure modes in ber-enhanced polymer composite: (a) ber breakage/rupture; (b) debonding between interfacial coating layer and polymer
matrix; (c) debonding at ber/polymer matrix interface.

1444

N. Ning et al. / Progress in Polymer Science 37 (2012) 14251455

Fig. 14. (a) Experimental evidence of PVA crystalline layer coating on CNT surface after pull out test; (b) stressstrain curves of PVA/CNT nanocomposites
and a linear tting between composite strength,  c , and CNT volume fraction, Vf [187].
2004 Wiley Periodicals.

thickness of the polymer sheath, b, and the shear strength


( s ) at the edge of the interfacial layer region, where R is
the sum of rf and b.
c =

1+

b
rf

 l

2rf

s 1 +

b
R

m Vf + m

(6)

By using Eq. (6), Coleman et al. [187] theoretically elucidated the contribution of the interfacial crystallization
layer to mechanical strengthening in PVA/CNTs composites. As shown in Fig. 14(a), the linear increment of
crystallinity with the CNT volume fraction [187] and the
exact microscopic analysis on the pullout CNT [167,187],
demonstrated that a thick crystalline layer coating around
the pullout CNTs. A dramatic enhancement achieved by
adding CNTs is presented in Fig. 14(b), and the linear tting of  c versus Vf is also included. Subsequently, the  s
could be calculated by correlating the slope of  c versus Vf
with Eq. (6). It was about 95 MPa, obviously higher than
the tensile strength of bulk PVA, which is 81 MPa. They
also demonstrated the positive inuences of the interfacial crystallization layer on the enhancement in Youngs
modulus by using the role of mixtures and/or the H-T equation [167,188]. The occurrence of interfacial crystallization
induces an interfacial layer with high mechanical strength,
as compared to bulk polymer, and enhances interfacial
bonding and polymerCNT stress transfer. Thus the formation of stiff ordered crystalline coating around CNT indeed

plays a dominant role in the reinforcement of semicrystalline PVA/CNTs nanocomposites.


Other researchers have attempted to estimate the
interfacial enhancement using computational molecular
dynamic simulations. Liao and coworkers [189,190] simulated the CNT pullout from a CNTpolymer system without
specic chemical bonding; the IFSS achieved was about
100 MPa, which is one order of magnitude higher than that
of microber (GF, CF) reinforced polymer composites. They
also suggested that if there is no chemical bonding, the
CNT/polymer adhesion arises from (1) electrostatic and van
der Waals interactions, and (2) mismatch in the coefcient
of thermal expansion (CTE) between CNT and polymer,
among other factors. An extremely strong non-covalent
bonding energy existing between polymer and CNT was
suggested by Panhuis et al. through a combination of computational simulation and experimental evidence [191].
Wei et al. [192] studied the interface between semicrystalline PE and CNT. Their results indicated that ordered
PE adsorption layers wrapped around the CNT and the
higher structural order parameter favors the enhancement
of composite modulus. A comparative analysis of IFSS and
lc for crystalline PE/CNT and amorphous PE/CNT composites was conducted by Frankland et al. [193]. The IFSS was
only 3 MPa for both of crystalline and amorphous interfaces under a non-bonded circumstance; whereas, even
for a very small amount cross-linking, the IFSS remarkably increased to 30 MPa for the amorphous interface and

N. Ning et al. / Progress in Polymer Science 37 (2012) 14251455

to 110 MPa for the crystalline interface. It should be noted


that the value of 110 MPa is about two times that for the
IFSS, 47 MPa, in a non-crystalline copolymer/CNT composite, measured directly through the pullout test [165].
A methodology that is reasonably straightforward to
determine whether any load is transferred to the CNT is
in situ spectroscopic analysis of the CNTpolymer composite under tension. Fluorescence [194], Raman [169,195]
and X-ray [196] spectroscopy are all suitable as candidates. In particular, for Raman spectroscopy, the G band
of CNT, located around 1620 cm1 , is sensitive to stress in
the CNT. This peak position will shift to a lower wavenumber under tensile deformation, and the slope of this peak
shift as a function of strain should be proportional to
the modulus of the composite [197] or the strength of
the CNTpolymer interface [198,199]. These features have
been utilized to indicate load transfer from polymer matrix
to CNT. In situ Raman measurements conrmed that efcient load transfer indeed occurred through the interface
of semicrystalline polymer and CNT in the PP/CNT [199]
and PVA/CNT [200] composites. In addition, Wang et al.
[196] utilized synchrotron X-ray spectroscopy to detect
prominent structural changes during deformation in the
semicrystalline PAN/CNT composite bers. The occurrence
of structural change was attributed to the fact that CNTs
enhance the load transferred to PAN crystals, implying
strong interfacial adhesion between crystalline PAN and
CNT lament.
A number of studies have attempted to manipulate the
crystalline interface between semicrystalline polymer and
ller, and then establish the microstructure-macroscopic
performance relation to deduce whether load transfer was
enhanced by the presence of interfacial crystallization.
An example is presented in Fig. 15, in which a comparative analysis of tensile behavior between spherulitic PP,
transcrystallized PP and PP transcrystallized on CNT ber
was performed by Zhang et al. [24]. The tensile strength
and Youngs modulus of PP transcrystallized on CNT ber
are signicantly higher than that of transcrystallized PP,

Fig. 15. Stressstrain curves for (i) PP isotropic spherulites, (ii) transcrystallized PP and (iii) PP transcrystallized on CNT ber [24] ( 2008 Elsevier
Ltd).

1445

indicating the effective load transfer realized by transcrystalline interface layer. In other work, the individual CNT
laments were combined together through PE crystallized on CNT to form the interlocking hybrid shishkebab
and/or transcrystalline nanostructures [125]. The dramatic
mechanical enhancements, especially for Youngs modulus, were attributed to the formation of hybrid crystalline
structures increased the load transfer from polymer crystals to the CNT laments. Minus et al. [201] used TEM to
reveal an extended-chain crystalline layer coating around
the CNTs in PVA/CNT composite bers. The interfacial
crystallization layer resulted in strong interfacial adhesion, which appeared to maximize the CNT-reinforcement
effect.
In the authors group, by imposing intensive shear
elds, a variety of interfacial crystallization superstructures were attained in large-scale, melt-processed, molded
articles. For the injection-molded bars of iPP/GF composites, the tensile strength increased continuously with the
GF content upon the appearance of iPP transcrystallized on
GF, whereas the strength was almost independent on the
GF content without the TC layer [56]. Hybrid shishkebabs
were achieved in HDPE/inorganic whisker composites by
using the DPIM technology; the exact structure of such
HSK could be varied by adjusting the molecular weight
of PE matrix [120] and the nuclei density on the whisker
[128]. The relationship between composite strength versus
whisker content indicated that a high level of perfection
of HSK results in excellent whisker reinforcement. For a
polymer/whisker system, a novel HSC superstructure was
observed. The crystal density [129] and perfection extent
[132] of the calabash spherulite dominated the interfacial
enhancement level between the polymer matrix and the
whisker ller. For PE/CNT composites, the formation of
HSK superstructure induced more obvious interfacial
reinforcement, compared to the composites without the
HSK structure [109,202]. Mai et al. [110] modulated the
exact structure of HSK structure in the melt-spun PE/CNT
ber by varying the ber spinning speed; a faster spinning
speed led to a ner HSK structure, correspondingly the
best mechanical enhancement was realized. These ndings
strongly suggest that the ultimate mechanical reinforcement provided by the ller can be sensitive to the state
of interfacial crystallization in the composite, which is
due to interfacial crystallization effectively enhancing the
interfacial adhesion and load transfer between polymer
and ber. Regarding to polymer/lamellar ller composites,
there are also some examples showing that interfacial
enhancement is induced by interfacial crystallization. The
TC superstructure is often detected between semicrystalline polymer and lamellar llers, such as mica [203]
and talc [204]. Xiang et al. [203] demonstrated that the
formation of TC structure on mica layer might enhance
interaction between PE and mica and result in a distinct
increment in mechanical strength. Kim et al. [205] found
that polyamide-12 (PA-12) transcrystallized on the surface
of nanoclay, with nanoclay layers also connected with each
other through the PA-12 lamellae perpendicularly grown
on nanoclay layer, yielding a rigid nanostructure network.
This nanostructure network resulting from interfacial
crystallization might impact the failure mechanisms under

1446

N. Ning et al. / Progress in Polymer Science 37 (2012) 14251455

tension and lead to mechanical enhancement of the composite [206]. Recently, Cai and coworkers [207] developed
a method to yield a polycaprolactone (PCL) crystalline layer
on the surface of graphite oxide nanoplatelets (GONPs); it
was suggested that the non-covalent interface is enhanced
by the crystalline interface layer and is favorable to load
transfer between polymer and GONPs.
5.2. Analysis and discussion
The inuence of interfacial crystallization on the interfacial properties such as interfacial adhesion and load transfer efciency has been well demonstrated above, based on
a signicant number of relevant investigations. By comparing various reports of interfacial crystallization, it can
be concluded that for micrometer-scale ller-reinforced
composites the issue of interfacial crystallization induced
interfacial enhancement is highly uncertain, whereas in
many nanometer-scale ller-reinforced crystalline polymer composites interfacial crystallization leads to enhance
interfacial interaction. So the question as to whether interfacial crystallization may induce interfacial enhancement
at all should be discussed, respectively, for these two kinds
of crystalline polymer/inorganic ller composites.
The interfacial crystallization fashion in microllerreinforced composites is typically TC. A transcrystalline
layer formed on GF and/or CF can be seen in many crystalline polymer/ber composites, which is often expected
to enhance the interfacial adhesion and load transfer ability. However, the increment in the IFSS originated from TC
is, in general, very limited. For many polymer/microber
composites, although the IFSS was increased prominently
with the appearance of transcrystalline interface, its absolute value was still lower than the tensile strength of
bulk polymer. For example, the IFSS of PP transcrystalline
interface layer on GF or CF is ranged from 5 MPa to
10 MPa, which is only 2030% of the tensile strength of
bulk PP [62,173]. This may explain why in many studies transcrystalline interface layer was considered to play
a negative role on interfacial enhancement. Furthermore,
it has been proposed that the interfacial enhancement
induced by transcrystalline layer will be impacted by
many factors, such as basal polymer properties, ber type,
ber mechanical strength, lamellae arrangement and orientation mode on ber [62,208], etc. Among these, we
emphasize two factors which have often been overlooked
in the literature. First is the question if the nucleation
mechanism of polymer transcrystallized on ber is due
to heterogeneous nucleation. For some semicrystalline
polymer/GF systems, TC does not occur upon quiescent
condition. In order to induce the occurrence of TC, a
shear eld is usually required, which could lead to a
thin layer of row nuclei coated on GF, and the subsequent growth of TC occurred on this row nuclei layer
is indeed followed by self-nucleation mechanism [209].
Therefore, in such a circumstance, the interfacial adhesion effect arising from heterogeneous nucleation might
be absent. Second, depression of the boundary formed
between the transcrystallinities and the spherulites may
dominate the extent of interfacial enhancement originated from TC. In a transcrystalline region, the brillation

of polymer crystals [183,208] will depend on nucleating


site density on ber surface. However, in a bulk crystalline region polymer commonly crystallizes into isotropic
spherulites. Due to the notable difference in structural
feature between TC and spherulites, a distinct boundary appears between transcrystalline region and bulk
crystalline region [210]. The adhesion at the boundary
of transcrystallites and spherulites is weak, and failure/breakage preferentially occurs at this location [183].
Experimental data supporting this suggestion is seen in
Moons work [175], in which the formation of interfacial
TC improved the interfacial strength between PP and GF
when no spherulite existed in the PP matrix, whereas the
interfacial strength was decreased when the spherulites
were well developed in the PP matrix even with the appearance of TC. To optimize the interfacial enhancement effect
of TC, cracking should occur across the crystalline brils in TC, and a blurry boundary of transcrystallites and
spherulites is necessary. This aim could be realized by
increasing nucleating site density on the ber surface,
reducing the distance between bers, or enlarging the
transcrystalline region to ll the entire composite volume.
On the other hand, interfacial crystallization fashion
in nanoller-reinforced polymers has received increased
attention. The nanometer-scale ller, such as CNT, possesses ultrahigh heterogeneous nucleation ability to
facilitate polymer crystallization on the nanoller surface. The IFSS measured in CNT-reinforced composites
is typically within a range from 50 MPa to 100 MPa,
which is one order of magnitude higher than that in
microber-reinforced composites, indicating extremely
strong interfacial adhesion and effective load transfer
between polymer and CNT. Moreover, the IFSS of crystalline interface is often signicantly larger than that of
an amorphous interface. Since the high values of IFSS are
close to or higher than the tensile strength of bulk polymer,
the interfacial enhancement effect is more easily detected
in CNT-reinforced composites, as compared to microberreinforced ones.
In a CNT-reinforced composite, even there is no chemical bonding between polymer and CNT, strong interfacial
adhesion can still be achieved. The origins of non-covalent
bonding are, as suggested by Liao and Li [189], mainly
from (1) electrostatic and van der Waals interactions,
(2) mismatch in the CTEs between CNT and polymer. For
the former item, an interfacial crystalline layer yielded
on the graphitic surface of CNT could enhance the electrostatic matching between graphite crystallites and the
segments in a special crystalline-chain conformation [211].
For example, the interfacial crystalline layer of PVDF coating on CNT is -modication, in which crystalline chains
adopt zigzag all-trans conformation; this conformation
favors to the generation of electro-transfer type F C bonding between PVDF and CNT [208]. For semicrystalline
polyolen without any polar atoms, a similar result may
occur after the appearance of crystalline coating which
may enhance the nonspecic CH interaction as suggested by Baskaran et al. [15]. As for differences in CTE,
since the thermal expansion coefcient of a polymer crystalline phase is smaller than its amorphous counterpart,

N. Ning et al. / Progress in Polymer Science 37 (2012) 14251455

1447

Fig. 16. Different interfacial crystallization superstructures induced remarkable mechanical reinforcements in crystalline polymer/inorganic ber composites [55] ( 2006 American Chemical Society); [119] ( 2007 American Chemical Society); [129] ( 2009 American Chemical Society).

mismatch in the CTEs between CNT and polymer might be


enlarged with the appearance of crystalline interface layer.
This would lead to higher compressive thermal residual
stress and a stronger mechanical interlocking effect [190],
compared to the case of amorphous polymer adsorbed
on the CNT surface. Therefore, the non-covalent bonding
between polymer and CNT would be strengthened significantly if interfacial crystallization occurred. On the other
hand, compared to an amorphous interface layer, the crystalline layer is stronger and stiffer and can bear a larger
shear stress and hence offer a more effective load transfer.
Since the adhesion between CNT and crystalline interface
layer is strong and the mechanical strength in crystalline
interface layer is evidently higher than that of the polymer bulk, failure would occur at the edge of the crystalline
interface layer rather than the interface between CNT
and polymer, as observed with some semicrystalline polymer/CNT composites [187]. When the CNT is tightly coated
with an ordered stiff crystalline layer, one could regard
this as an increase in the effective volume fraction of CNT.
Moreover, the distinct mismatch between stiff nanoller
and soft polymer could be looked as a barrier for effective load transfer, and the crystalline interface layer acts
as an intermediate to reduce this mismatch and thus
prominently enhance the load transfer efciency [212].
According to the reasons mentioned above, it is not surprising that interfacial crystallization can induce obvious
interfacial enhancement in nanoller-reinforced polymer
composites.

5.3. Mechanical properties of composites containing


special interfacial crystallization structures
Fillers with ultrahigh mechanical strength are regarded
as ideal reinforcement elements to tailor the mechanical
properties of polymers. The mechanical reinforcement
behavior in various polymer/microller or nanoller
composites has been well reviewed [170,213217]. Here
we devote our attention on the mechanical reinforcements
in the non-covalently bonded polymer/ller composites
containing special interfacial crystallization structures.
The effects of interfacial crystallization on mechanical
reinforcement will be particularly emphasized. For examples, different interfacial crystallization structures, such
as TC, hybrid shishkebab, and hybrid shishcalabash,
resulting in enhancements in tensile strength are demonstrated in Fig. 16. The dynamic samples were obtained
using the specic DPIM technique, while the static samples
were from the conventional injection-molding technique.
Since intensive shearing was introduced during the
DPIM process, hybrid shishkebab, hybrid shishcalabash
and TC superstructures were formed in HDPE/whisker,
iPP/whisker and iPP/GF composites, respectively, resulting
in dramatic reinforcements in the dynamic samples. In
particular, the increase of tensile strength dependent
on ller content was more signicant for the dynamic
samples, indicating that the appearance of interfacial
crystallization brings a higher efciency of reinforcement,
compared to samples without such interfacial structure.

1448

N. Ning et al. / Progress in Polymer Science 37 (2012) 14251455

Fig. 17. Time-resolved FTIR spectra and the corresponding normalized peak intensities at various characteristic bands as a function of crystallization time
of pure PLLA (a) and PLLA/CNTs nanocomposites (b); schematic representation of conformational ordering in PLLA and PLLA/CNTs nanocomposites (c)
[223].
2009 American Chemical Society.

The effects of transcrystallization on mechanical


strength have been intensively studied, and a summary of
TC-induced mechanical enhancement is seen in a review
paper [55]. Only a few examples will be presented as
follows. As mentioned in Section 2.1, CF possesses high
efciency for inducing PEEK TC, and correspondingly interfacial TC improved the mechanical strength of PEEK/CF
composites [39]. In a CF-reinforced semicrystalline thermoplastic system, the formation of interfacial TC induced a
100% increase in the yield stress versus the quenched control sample [218]. For injection-molded articles of iPP/GF
composite, a transcrystalline interface layer yielded a tensile strength increase of 40% with 30 wt% GF; whereas,
without TC the tensile strength changed for only 10% by
adding the same content of GF [218]. A comparative analysis (see Fig. 15) indicated that when PP transcrystallized
on CNT ber, the tensile strength and Youngs modulus
were all three times that found for isotropic PP spherulites;
however, the increases of strength and modulus were only
100% and 33%, respectively, for transcrystallized PP, compared to the case of isotropic PP spherulites [192]. Zhang
et al. [125] fabricated CNT/PE 70/30 wt% bers containing
a special interlocking transcrystalline nanostructure; the

strength and modulus increased 100% and 300%, respectively, as compared to the neat CNT ber.
Hybrid shishkebab (HSK) superstructures have been
achieved in some microber- and/or nanober-reinforced
composites in our previous work. We found that the
mechanical enhancement extent in PE/whisker composites was strongly dependent on the perfection of HSK; the
increment in tensile strength due to HSK was 300% for a
ne HSK structure, while it was only 200% for an imperfect HSK [120]. For an iPP/whisker system, the strength
and modulus were increased for 41% and 200%, respectively, upon the appearance of hybrid shishcalabash [129].
Yang et al. [109] obtained hybrid shishkebab superstructures in PE/multiwalled carbon nanotube (MWNT)
composites; for a 5 wt% MWNT content, the formation of
HSK resulted in 170% and 80% enhancement in strength
and modulus, respectively. The structural feature of HSK
might impact strongly the CNT-reinforcement effect in
melt-spun PE/MWNT bers [110]. When MWNT content
was improved from 0 wt% to 5 wt%, the tensile strength
was increased 200% upon the formation of ne HSK superstructure, while the increased amplitude was only 30%
upon the appearance of imperfect HSK. Ruan et al. [219]

N. Ning et al. / Progress in Polymer Science 37 (2012) 14251455

1449

Finally, for lamellar nanoller-reinforced polymer composites, Maiti and Okamoto [122] revealed that the
interfacial crystalline layer of polyamide-6 (PA-6) bonded
on the surface of nanoclay through hydrogen bonding
between N H groups on PA-6 and O Si groups on nanoclay, which was responsible for a dramatic enhancement
in elastic storage modulus, which increased by 115% after
adding 3.7 wt% nanoclay. In another study [212], the PCL
crystalline layer formed on the layered surface of GONPs
leaded to a 1.5 times increase in the Youngs modulus upon
adding 2 wt% GONPs.
6. Conclusions and nal remarks

Fig. 18. Molecular dynamic simulations of PE molecules at the interface


of CNT (5, 5) at T = 50 K (a); snapshots of the nal conformation of PE with
different chain lengths on SWNT (10, 10) (b) [228]. ( 2007 American
Institute of Physics); [232] ( 2006 American Chemical Society).

also observed an HSK superstructure in a UHMWPE/MWNT


composite lm; 40% and 50% increases in modulus and
yield strength were obtained by adding 1 wt% MWNT.
For MWNT-reinforced PVA composites, debonding
occurs at the edge of the crystalline interface layer;
correspondingly, a comprehensive mechanics analysis
found that the modulus, strength and toughness were
increased for 370%, 430% and 170%, respectively, by adding
less than 1 wt% MWNT [187]. A comparative investigation
between semicrystalline PVA/CNT composite and amorphous poly(9-vinyl carbazole) (PVK)/CNT composite has
been done by Cadek et al. [167]. The presence of CNT
nucleated the PVA crystallization and resulted in an obvious enhancement as the Youngs modulus was increased
by 180% after adding 1 wt% CNT; whereas, for the noncrystalline PVK/CNT system, the heterogeneous nucleation
effect was absent and the modulus was increased only by
110% after adding approximate 1.5 wt% CNT. Obviously,
interfacial crystallization is more favorable for realizing
the ultimate mechanical enhancement in polymer/CNT
composites. In gel-spun PVA/single wall carbon nanotube
(SWNT) composite bers, the addition of 1 wt% SWNT
resulted in a 60% increase in tensile strength with the
occurrence of interfacial crystallization [201]. Wang et al.
[220] adopted the rule of mixtures to estimate that the
effective strength of SWNT was as high as 88 GPa, which
approaches the real failure strength of SWNT measured
by Yu et al. [221], indicating an extraordinary reinforcing
efciency due to the addition of SWNTs.

In summary, interfacial crystallization has attracted


tremendous interest over the past several decades, not
only due to its crystallography interests, but also due to
the fact that it may be a novel strategy for enhancing the
interfacial adhesion and reinforcing the mechanical performances of polymer/ller composites. Numerous studies
have been devoted to the structural/morphological features of different kinds of interesting hybrid crystalline
superstructure (TC, HSK, HSC, etc.) in various semicrystalline polymer/ller (with high aspect ratio) composite
systems. Furthermore, tremendous effort has been made
to assess various inuences, such as the physical and
chemical characteristic of llers, molecular chain structure
and conformation of polymer matrix, molecular weight,
and external elds, on the nal structure/morphology of
interfacial crystallization. Meanwhile, the nucleation and
growth mechanisms of crystallization behaviors at polymer/ller interface were clearly demonstrated. For many
crystalline llers, such as graphite, talc, or mica, the epitaxial crystallization of polymer matrix on the surface of
these llers is widely believed to be the formation mechanism of interfacial crystallization. For some llers with
very small diameter (similar to the radius of gyration of
the polymer), such as CNTs, the interfacial crystallization
mechanism is most likely soft epitaxy (geometric connement). For polar polymer/inorganic llers systems, the
chemisorption of the polar groups of the polymer onto
the inorganic ller surface could lead to strong interactions between the macromolecules and the inorganic llers
and thus can facilitate nucleation and lateral interfacial
crystallization. Furthermore, stress or strain generated at
the polymer/llers interface may lead to the orientation of
macromolecular chains and a decrease of the nucleation
barrier, and thus contribute to the enhancement of nucleation on the surface of llers and the formation of various
hybrid crystalline structures.
Systematic studies of the crystallography and crystallization mechanisms of interfacial crystallization (hybrid
crystalline superstructure) in polymer/ller composites
provide opportunities for the manipulation of interfacial
crystalline architecture during the fabrication and processing of composites, and have stimulated great interest to
establish and/or modulate some specic interfacial crystalline structures/morphologies to lead to novel interfacial
enhancement. For micrometer-scale ller-reinforced composites, the as-generated interfacial crystalline fashion is
typically TC. The IFSS of interfacial transcrystalline layer

1450

N. Ning et al. / Progress in Polymer Science 37 (2012) 14251455

is typically at a level of approximately 10 MPa, which is


signicantly lower than the shear strength of a bulk polymer bulk (such as PP). Accordingly, a number of studies
have suggested that TC offers little effect (or in fact plays
a negative role) on interfacial enhancement. On the other
hand, recent advances in nanometer-scale ller-reinforced
semicrystalline/polymer composites suggest a prevailing
viewpoint that the appearance of interfacial crystallization
favors strong interfacial adhesion and high load transfer efciency between polymer matrix and nanoller. In
addition, computational molecular dynamic simulations
suggest that the shear strength of crystalline interface
layer and the IFSS between semicrystalline/polymer and
nanoller is at a level of approximately 100 MPa. This value
is one order of magnitude higher than that of interfacial TC layer and much larger than that for an amorphous
interface in nanoller-reinforced composites. Meanwhile,
by elucidating the microstructure-macroscopic performance relations, the interfacial enhancement induced
by interfacial crystallization has been conrmed for
many semicrystalline polymer/nanoller composite systems. More importantly, in a number of comparative
studies contrasting different levels of interfacial crystallization, the best mechanical reinforcement was always
found in the composites with high perfection, high volume fraction of interfacial crystalline superstructure,
strongly indicating that the manipulation of interfacial crystalline structure/morphology is a promising
approach to effectively achieve excellent properties in real
processed articles of semicrystalline polymer/ller composites.
Lastly, some perspectives based on the authors comprehensions are provided as follows. (1) Despite the extensive
investigation of interfacial crystallization, the exact formation mechanism is still not quite clear. To obtain a complete
understanding of the origin of interfacial crystallization,
what happens in the early stage of the crystallization is
a key. By using time-resolved Fourier-transform infrared
spectroscopy (FTIR), the detailed information on chain conformational ordering induced by the surface of llers and
the lateral crystallization kinetics could be obtained, providing a very easy and effective method to understand the
origin of interfacial crystallization in the early stages at the
molecular level. A few reports have indeed revealed the
origin of CNT-induced polymer crystallization at an early
stage with the help of in situ FTIR [222,223]. For example,
the time-resolved FTIR spectra and the corresponding normalized peak intensities at various characteristic bands as
a function of crystallization time of pure PLLA (Fig. 17(a))
and PLLA/CNTs nanocomposites (Fig. 17(b)) revealed that
CNTs could provide templates for the conformational
ordering of PLLA. The schematic representation of conformational ordering in PLLA (upper part) and PLLA/CNTs
nanocomposites (bottom part) is clearly shown in Fig. 17(c)
[223]. This work provides an interesting exploration of
the early stages of CNTs nanocomposite interfacial crystallization. In the future, additional work focused on the
early stages of interfacial crystallization mechanism using
this in situ FTIR technique is recommended. Meanwhile,
molecular dynamics (MD) simulations have been widely
regarded as a powerful tool for investigating the dynamics

of polymer crystallization as well as the physical interactions between polymer chain and the surface of llers
[224233], offering huge potential to ascertain the adsorption and preorientation processes of polymer molecules
on the surface of llers, which would helpful for understanding the mechanisms of nucleation and interfacial
crystallization. Some work has demonstrated that the PE
molecule chains will be aligned along CNT axis in extended
conformations under certain conditions using MD simulation, as shown in Fig. 18(a) [232]. In addition, some
work reported that the attractive van der Waals interactions controlled the adsorption and preorientation of
PE on SWNT, and the single PE chains with different
chain lengths will be indeed aligned parallel to the SWNT
axis under isothermal crystallization process, as clearly
shown in Fig. 18(b) [228]. (2) The control of interfacial
crystalline structure/morphology via solution processing
has been well addressed in past literature. An in-depth
understanding about how to exactly adjust the interfacial
crystallization behavior in melt processing requires more
studies, and a facile and effective methodology for achieving ne interfacial crystalline superstructure in large-scale
melt processing articles has not yet been established.
Thus more work is necessary to realize the interfacial
crystallization and the resultant interfacial enhancement
in real processing fabrications of semicrystalline polymer/ller composites. (3) The relations between interfacial
crystalline superstructure and interfacial interaction are
still qualitative. Quantitative dependences of structural
issues of interfacial crystallization, such as crystallinity,
lamellar thickness and crystallographic periodicity, on the
interfacial adhesion level are urgently desired, as such
investigations are necessary to understand the fundamental reasons of interfacial crystallization induced interfacial
enhancement. By using novel controlled-crystallization
method like SC CO2 -processing, manipulation of interfacial crystalline structure/morphology and volume fraction
become easy and effective, providing huge potential to
resolve this subject. (4) While mechanical models relating
interfacial crystallization to macroscopic mechanics have
focused on brous-ller-reinforced semicrystalline polymer composites; models for layered-ller composites are
lacking. Nevertheless, the authors believe that, with the rise
of graphene and/or graphene oxide nanosheet-reinforced
polymer composites, the relevant investigations might
appear soon and develop rapidly. (5) Except for mechanical reinforcement, the effects of interfacial crystallization
on other macroscopic properties need to be addressed.
For example, a dramatic improvement in thermal distortion temperature (HDT) was detected in nylon-6/nanoclay
composites [122], and the formation of interfacial crystalline layer of nylon-6 on the surface of nanoclay sheet
has been revealed clearly [122]. However, the connection
between the increased HDT and the formation of interfacial crystalline layer is not clear. The situation is similar for
the improved gas permeability in layered-ller-reinforced
semicrystalline polymer composites. It is highly expected
that the appearance of interfacial crystallization can tailor not only mechanical properties but also functional
performances, even bring certain new functionalities in
polymer/ller composites.

N. Ning et al. / Progress in Polymer Science 37 (2012) 14251455

Acknowledgement
We express our sincere thanks to the National Natural
Science Foundation of China (NNSFC) for nancial support
(Grant nos.: 21034005, 51121001). K.W. greatly appreciates the nancial support from NNSFC (50903048) for
partial researches cited in this paper.
References
[1] Yuan W, Feng JL, Judeh Z, Dai J, Chan-Park MB. Use of polyimidegraft-bisphenol a diglyceryl acrylate as a reactive noncovalent
dispersant of single-walled carbon nanotubes for reinforcement of
cyanate ester/epoxy composite. Chem Mater 2010;22:654254.
[2] Ramanathan T, Abdala AA, Stankovich S, Dikin DA, Herrera-Alonso
M, Piner RD, Adamson DH, Schniepp HC, Chen X, Ruoff RS,
Nguyen ST, Aksay IA, Prudhomme RK, Brinson LC. Functionalized
graphene sheets for polymer nanocomposites. Nat Nanotechnol
2008;3:32731.
[3] Deng Y, Li YJ, Dai J, Lang MD, Huang XY. An efcient way to functionalize graphene sheets with presynthesized polymer via ATNRC
chemistry. J Polym Sci, Part A Polym Chem 2011;49:158290.
[4] Cao YW, Lai ZL, Feng JC, Wu PY. Graphene oxide sheets covalently
functionalized with block copolymers via click chemistry as reinforcing llers. J Mater Chem 2011;21:92718.
[5] Yang BX, Pramoda KP, Xu GQ, Goh SH. Mechanical reinforcement
of polyethylene using polyethylene-grafted multiwalled carbon
nanotubes. Adv Funct Mater 2007;17:20629.
[6] Vaz CM, Reis RL, Cunha AM. Use of coupling agents to enhance
the interfacial interactions in starch-EVOH/hydroxylapatite composites. Biomaterials 2002;23:62935.
[7] Das B, Prasad KE, Ramamurty U, Rao CNR. Nano-indentation studies
on polymer matrix composites reinforced by few-layer graphene.
Nanotechnology 2009;20:125705/1125705.
[8] Salavagione HJ, Martinez G. Importance of covalent linkages in the
preparation of effective reduced graphene oxide-poly(vinyl chloride) nanocomposites. Macromolecules 2011;44:268592.
[9] Kim HS, Lee BH, Choi SW, Kim S, Kim HJ. The effect of types of
maleic anhydride-grafted polypropylene (MAPP) on the interfacial
adhesion properties of bio-our-lled polypropylene composites.
Compos Part A-Appl S 2007;38:147382.
[10] Dominkovics Z, Renner K, Pukanszky Jr B, Pukanszky B. Quantitative
characterization of the structure of PP/layered silicate nanocomposites at various length scales. Macromol Symp 2008;267:526.
[11] Hu HT, Wang XB, Wang JC, Wan L, Liu FM, Zheng H, Chen R, Xu CH.
Preparation and properties of graphene nanosheets-polystyrene
nanocomposites via in situ emulsion polymerization. Chem Phys
Lett 2010;484:24753.
[12] Raee MA, Raee J, Srivastava I, Wang Z, Song H, Yu ZZ,
Koratkar N. Fracture and fatigue in graphene nanocomposites.
Small 2010;6:17983.
[13] Sun T, Garces JM. High-performance polypropylene-clay nanocomposites by in situ polymerization with metallocene/clay catalysts.
Adv Mater 2002;14:12830.
[14] Liu K, Chen L, Chen Y, Wu J, Zhang W, Chen F, Fu Q. Preparation of polyester/ reduced graphene oxide composites via in situ
melt polycondensation and simultaneous thermo-reduction of
graphene oxide. J Mater Chem 2011;21:86127.
[15] Baskaran D, Mays JW, Bratcher MS. Noncovalent and nonspecic
molecular interactions of polymers with multiwalled carbon nanotubes. Chem Mater 2005;17:338997.
[16] Moniruzzaman M, Chattopadhyay J, Billups WE, Winey KI. Tuning
the mechanical properties of SWNT/nylon 6, 10 composites with
exible spacers at the interface. Nano Lett 2007;7:117885.
[17] Chen EJH, Hsiao BS. The effects of transcrystalline interphase in
advanced polymer composites. Polym Eng Sci 1992;32:2806.
[18] Gati A, Wagner H. Stress transfer efciency in semicrystallinebased composites comprising transcrystalline interlayers. Macromolecules 1997;30:39335.
[19] Stadlbauer M, Janeschitz-Kriegl H, Eder G, Ratajski E. New extensional rheometer for creep ow at high tensile stress. Part II.
Flow induced nucleation for the crystallization of iPP. J Rheol
2004;48:6319.
[20] Stern T, Teishev A, Marom G. Composites of polyethylene reinforced with chopped polyethylene bers: effect of transcrystalline
interphase. Compos Sci Technol 1997;57:100915.

1451

[21] Coleman JN, Khan U, Gunko YK. Mechanical reinforcement of polymers using carbon nanotubes. Adv Mater 2006;18:689706.
[22] Njuguna J, Pielichowski K, Desai S. Nanoller-reinforced polymer
nanocomposites. Polym Adv Technol 2008;19:94759.
[23] Ku H, Wang H, Pattarachaiyakoop N, Trada M. A review on the
tensile properties of natural ber reinforced polymer composites.
Compos Part B-Eng 2011;42:85673.
[24] Zhang S, Minus ML, Zhu L, Wong CP, Kumar S. Polymer transcrystallinity induced by carbon nanotubes. Polymer 2008;49:1356
64.
[25] Jenckel E, Teege E, Hinrichs W. Transkristallisation in hochmolekularen Stoffen. Colloid Polym Sci 1952;129:1924.
[26] Yan BW, Wu H, Jiang GJ, Guo SY, Huang JA. Interfacial crystalline structures in injection over-molded polypropylene and bond
strength. ACS Appl Mater Inter 2010;2:302336.
[27] Assouline E, Grigull S, Marom G, Wachtel E, Wagner HD. Morphology of alpha-transcrystalline isotactic polypropylene under tensile
stress studied with synchrotron microbeam X-ray diffraction. J
Polym Sci, Part B Polym Lett 2001;39:201621.
[28] Assouline E, Wachtel E, Grigull S, Lustiger A, Wagner HD, Marom
G. Lamellar twisting in alpha isotactic polypropylene transcrystallinity investigated by synchrotron microbeam X-ray diffraction.
Polymer 2001;42:62317.
[29] Assouline E, Wachtel E, Grigull S, Lustiger A, Wagner HD,
Marom G. Lamellar orientation in transcrystalline gamma isotactic polypropylene nucleated on aramid bers. Macromolecules
2002;35:4039.
[30] Wang C, Liu CR. Transcrystallization of polypropylene composites:
nucleating ability of bres. Polymer 1999;40:28998.
[31] Wang C, Liu FH, Huang WH. Electrospun-ber induced transcrystallization of isotactic polypropylene matrix. Polymer
2011;52:132636.
[32] Feller JF, Grohens Y. Coupling ability of silane grafted poly(propene)
at glass bers/poly(propene) interface. Compos Part A-Appl S
2004;35:110.
[33] Cao Y, Feng J, Wu P. DSC and morphological studies on the
crystallization behavior of nucleated isotactic polypropylene
composites lled with Kevlar bers. J Therm Anal Calorim
2011;103:33945.
[34] Lee BG, Lee S, Via BK. Inuence of surface morphology of the kraft
pulp bers on the growth of the transcrystalline layer of polypropylene. J Appl Polym Sci 2010;116:195866.
[35] Cartledge HCY, Baillie CA. Studies of microstructural and mechanical properties of nylon/glass composite-Part I-the effect of thermal
processing on crystallinity, transcrystallinity and crystal phases. J
Mater Sci 1999;34:5099111.
[36] Gong Y, Yang G. Manufacturing and physical properties of allpolyamide composites. J Mater Sci 2009;44:463944.
[37] Choudhury A. Isothermal crystallization and mechanical behavior of ionomer treated sisal/HDPE composites. Mater Sci Eng A
2008;491:492500.
[38] Zhang R, Huang Y, Min M, Gao Y, Yu X, Lu A, Lu Z. Isothermal crystallization of pure and glass ber reinforced poly (phenylene sulde)
composites. Polym Compos 2009;30:4606.
[39] Jeng CC, Chen M. Flexural failure mechanisms in injectionmoulded carbon bre/PEEK composites. Compos Sci Technol
2000;60:186372.
[40] Gao SL, Kim JK. Cooling rate inuences in carbon bre/PEEK composites. Part 1. Crystallinity and interface adhesion. Compos Part
A-Appl S 2000;31:51730.
[41] Yuan Q, Bateman SA, Friedrich K. Thermal and mechanical properties of PAN-and pitch-based carbon ber reinforced PEEK
composites. J Thermoplast Compos 2008;21:32336.
[42] Ninomiya N, Kato K, Fujimori A, Masuko T. Transcrystalline structures of Poly(L-lactide). Polymer 2007;48:487482.
[43] Mathew AP, Oksman K, Sain M. The effect of morphology and chemical characteristics of cellulose reinforcements on the crystallinity
of polylactic acid. J Appl Polym Sci 2006;101:30010.
[44] Wang YM, Tong BB, Hou SJ, Li M, Shen CY. Transcrystallization
behavior at the poly(lactic acid)/sisal bre biocomposite interface.
Compos Part A-Appl S 2011;42:6674.
[45] Le Duigou A, Davies P, Baley C. Interfacial bonding of ax bre/poly
(l-lactide) bio-composites. Compos Sci Technol 2010;70:2319.
[46] Na K, Park HS, Won HY, Lee JK, Lee KH, Nam JY, Jin BS.
SALS study on transcrystallization and ber orientation in glass
ber/polypropylene composites. Macromol Res 2006;14:499503.
[47] Cartledge HCY, Baillie CA. Effects of crystallinity, transcrystallinity
and crystal phases of GF/PA on friction and wear mechanisms. J
Mater Sci 2002;37:300522.

1452

N. Ning et al. / Progress in Polymer Science 37 (2012) 14251455

[48] Cartledge HCY, Baillie CA. Studies of microstructural and mechanical properties of nylon/glass composite-Part II-the effect of
microstructures on mechanical and interfacial properties. J Mater
Sci 1999;34:511326.
[49] Ramanathan T, Schulz E, Subramanian K. Determination of microhardness and elastic modulus in the transcrystalline zone of carbon
bre/PPS composites using a knife edge indentor. Compos Sci Technol 2005;65:17.
[50] Lu KB, Grossiord N, Koning CE, Miltner HE, van Mele B, Loos J.
Carbon nanotube/isotactic polypropylene composites prepared by
latex technology: morphology analysis of CNT-induced nucleation.
Macromolecules 2008;41:80815.
[51] Naiki M, Fukui Y, Matsumura T, Nomura T, Matsuda M. The effect of
talc on the crystallization of isotactic polypropylene. J Appl Polym
Sci 2001;79:1693703.
[52] Lin CW, Ding SY, Hwang YW. Interfacial crystallization of isotactic
polypropylene molded against the copper surface with various surface roughnesses prepared by an electrochemical process. J Mater
Sci 2001;36:49438.
[53] Lin CW, Lai YC, Liu SS. Effect of the surface roughness of sulfuric acid-anodized aluminum mold on the interfacial crystallization
behavior of isotactic polypropylene. J Adhes Sci Technol 2001;15:
92944.
[54] Quan H, Li ZM, Yang MB, Huang R. On transcrystallinity in
semi-crystalline polymer composites. Compos Sci Technol 2005;
65:9991021.
[55] Wang K, Guo M, Zhao DG, Zhang Q, Du RN, Fu Q, Dong X, Han CC.
Facilitating transcrystallization of polypropylene/glass ber composites by imposed shear during injection molding. Polymer 2006;
47:83749.
[56] Li Y, Chen LX, Zhou XD. Interfacial crystalline behavior in glassber/polypropylene composites modied by block copolymer
coupling agents. J Mater Sci 2008;43:508391.
[57] Li Y, Lin Q, Zhou X. Thermal treatment effect on transcrystalline interphase and interfacial shear strengths of glass
bre/polypropylene composites. Iran Polym J 2011;20:50311.
[58] Li Q, Zheng G, Dai K, Xie M, Liu C, Liu B, Zhang X, Wang B, Chen J,
Shen C. -transcrystallinity developed from the novel ringed nuclei
in the glass ber/isotactic polypropylene composite. Mater Lett
2011;65:22747.
[59] Clark RL, Kander RG, Sauer BB. Nylon 66/poly(vinyl pyrrolidone)
reinforced composites: 1. Interphase microstructure and evaluation of ber-matrix adhesion. Compos Part A-Appl S 1999;30:
2736.
[60] Wu CM, Chen M, Karger-Kocsis J. Transcrystallization in syndiotactic polypropylene induced by high-modulus carbon bers. Polym
Bull 1998;41:23945.
[61] Wu CM, Chen M, Karger-Kocsis J. Interfacial shear strength and failure modes in sPP/CF and iPP/CF microcomposites by fragmentation.
Polymer 2001;42:12935.
[62] Wang C, Chen CC. Surface-induced crystallization of syndiotactic polystyrene on high modulus carbon bers. Polym Bull
1999;43:43340.
[63] Nesher G, Serror M, Avnir D, Marom G. Silver coated vapor-growncarbon-nanobers for effective reinforcement of polypropylenepolyaniline. Compos Sci Technol 2011;71:1529.
[64] Yudin VE, Svetlichnyi VM, Shumakov AN, Schechter R, Harel H,
Marom G. Morphology and mechanical properties of carbon ber
reinforced composites based on semicrystalline polyimides modied by carbon nanobers. Compos Part A-Appl S 2008;39:8590.
[65] Shi HF, Zhao Y, Dong X, He CC, Wang DJ, Xu DF. Transcrystalline
morphology of nylon 6 on the surface of aramid bers. Polym Int
2004;53:16726.
[66] Larin B, Marom G, Avila-Orta CA, Somani RH, Hsiao BS. Orientated
crystallization in discontinuous aramid ber/isotactic polypropylene composites under shear ow conditions. J Appl Polym Sci
2005;98:11138.
[67] Feldman AY, Wachtel E, Zafeiropoulos NE, Schneider K, Stamm M,
Davies RJ, Weinberg A, Marom G. In situ synchrotron microbeam
analysis of the stiffness of transcrystallinity in aramid ber
reinforced nylon 66 composites. Compos Sci Technol 2006;66:
200915.
[68] Larin B, Avila-Orta CA, Somani RH, Hsiao BS, Marom G. Combined
effect of shear and brous llers on orientation-induced crystallization in discontinuous aramid ber/isotactic polypropylene
composites. Polymer 2008;49:295302.
[69] Feldman A, Gonzalez MF, Marom G. Transcrystallinity in surface
modied aramid ber reinforced nylon 66 composites. Macromol
Mater Eng 2003;288:8616.

[70] He CC, Dong X, Zhang XQ, Wang DJ, Xu DF. Morphology investigation of transcrystallinity at polyamide 66/aramid ber interface. J
Appl Polym Sci 2004;91:29803.
[71] Korbakov N, Harel H, Feldman Y, Marom G. Dielectric-response
of aramid ber-reinforced PEEK. Macromol Chem Phys 2002;203:
226772.
[72] Feldman A, Wachtel E, Vaughan G, Weinberg A, Marom G.
The brill transition in transcrystalline nylon-66. Macromolecules
2006;39:44559.
[73] Feldman AY, Gonzalez MF, Wachtel E, Moret MP, Marom G. Transcrystallinity in aramid and carbon ber reinforced nylon 66:
determining the lamellar orientation by synchrotron X-ray micro
diffraction. Polymer 2004;45:723945.
[74] Gray DG. Transcrystallization of polypropylene at cellulose
nanocrystal surfaces. Cellulose 2008;15:297301.
[75] Son SJ, Lee YM, Im SS. Transcrystalline morphology and mechanical properties in polypropylene composites containing cellulose
treated with sodium hydroxide and cellulase. J Mater Sci 2000;
35:576778.
[76] Amash A, Zugenmaier P. Morphology and properties of isotropic
and oriented samples of cellulose bre-polypropylene composites.
Polymer 2000;41:158996.
[77] Jose T, Joseph A, Skrifvars M, Thomas S, Joseph K. Thermal
and crystallization behavior of cotton-polypropylene commingled
composite systems. Polym Compos 2010;31:148794.
[78] Garkhail S, Wieland B, George J, Soykeabkaew N, Peijs T. Transcrystallisation in PP/ax composites and its effect on interfacial and
mechanical properties. J Mater Sci 2009;44:5109.
[79] Zafeiropoulos NE, Baillie CA, Matthews FL. The effect of transcrystallinity on the interface of green ax/polypropylene composite
materials. Adv Compos Lett 2001;10:22936.
[80] Zafeiropoulos NE, Baillie CA, Matthews FL. A study of transcrystallinity and its effect on the interface in ax bre
reinforced composite materials. Compos Part A-Appl S 2001;32:
52543.
[81] Joseph PV, Joseph K, Thomas S, Pillai CKS, Prasad VS, Groeninckx G,
Sarkissova M. The thermal and crystallisation studies of short sisal
bre reinforced polypropylene composites. Compos Part A-Appl S
2003;34:25366.
[82] Gassan J, Mildner I, Bledzki AK. Transcrystallization of polypropylene on different modied jute bers. Compos Interfaces
2001;8:44352.
[83] Borysiak S. A study of transcrystallinity in polypropylene in the
presence of wood irradiated with gamma rays. J Therm Anal
Calorim 2010;101:43945.
[84] Sanadi AR, Cauleld DF. Transcrystalline interphases in natural
ber-PP composites: effect of coupling agent. Compos Interfaces
2000;7:3143.
[85] Hermida EB, Mega VI. Transcrystallization kinetics at the
poly(3-hydroxybutyrate-co-3-hydroxyvalerate)/hemp bre interface. Compos Part A-Appl S 2007;38:138794.
[86] Ning NY, Yin QJ, Luo F, Zhang Q, Du R, Fu Q. Crystallization behavior
and mechanical properties of polypropylene/halloysite composites. Polymer 2007;48:737484.
[87] Canetti M, De Chirico A, Audisio G. Morphology, crystallization and
melting properties of isotactic polypropylene blended with lignin.
J Appl Polym Sci 2004;91:143542.
[88] Li LY, Li B, Hood MA, Li CY. Carbon nanotube induced polymer
crystallization: the formation of nanohybrid shish-kebabs. Polymer
2009;50:95365.
[89] Li LY, Li CY, Ni CY. Polymer crystallization-driven, periodic patterning on carbon nanotubes. J Am Chem Soc 2006;128:16929.
[90] Li LY, Li B, Yang GL, Li CY. Polymer decoration on carbon nanotubes
via physical vapor deposition. Langmuir 2007;23:85225.
[91] Li LY, Li CY, Ni CY, Rong LX, Hsiao B. Structure and crystallization
behavior of Nylon 66/multi-walled carbon nanotube nanocomposites at low carbon nanotube contents. Polymer 2007;48:
345260.
[92] Zhang F, Zhang H, Zhang ZW, Chen ZM, Xu Q. Modication of carbon nanotubes: water-soluble polymers nanocrystal wrapping to
periodic patterning with assistance of supercritical CO2 . Macromolecules 2008;41:451923.
[93] Zhang SJ, Lin W, Zhu LB, Wong CP, Bucknall DG. Gamma-form
transcrystals of poly(propylene) induced by individual carbon
nanotubes. Macromol Chem Phys 2010;211:134854.
[94] Miltner HE, Grossiord N, Lu KB, Loos J, Koning CE, Van Mele B. Isotactic polypropylene/carbon nanotube composites prepared by latex
technology. Thermal analysis of carbon nanotube-induced nucleation. Macromolecules 2008;41:575362.

N. Ning et al. / Progress in Polymer Science 37 (2012) 14251455


[95] Kitayama T, Utsumi S, Hamada H, Nishino T, Kikutani T, Ito
H. Interfacial properties of PP/PP composites. J Appl Polym Sci
2003;88:287583.
[96] Dean DM, Rebenfeld L, Register RA, Hsiao BS. Matrix molecular orientation in ber-reinforced polypropylene composites. J Mater Sci
1998;33:4797812.
[97] Thierry A, Straupe C, Lotz B, Wittmann JC. Physical gelation: a path
towards ideal dispersion of additives in polymers. Polym Commun
1990;31:299301.
[98] Li CY, Li LY, Cai WW, Kodjie SL, Tenneti KK. Nanohybrid Shishkebabs. Periodically functionalized carbon nanotubes. Adv Mater
2005;17:1198.
[99] Zhang L, Tao T, Li CZ. Formation of polymer/carbon nanotubes
nano-hybrid shish-kebab via non-isothermal crystallization. Polymer 2009;50:383540.
[100] Yue J, Xu Q, Zhang ZW, Chen ZM. Periodic patterning on carbon nanotubes: supercritical CO2 -induced polyethylene epitaxy.
Macromolecules 2007;40:88216.
[101] Zhang ZW, Xu Q, Chen ZM, Yue J. Nanohybrid shish-kebabs:
supercritical CO2 -induced PE epitaxy on carbon nanotubes. Macromolecules 2008;41:286873.
[102] Yu N, Zheng XL, Xu Q, He LH. Controllable-induced crystallization
of PE-b-PEO on carbon nanotubes with assistance of supercritical
CO2 : effect of solvent. Macromolecules 2011;44:395865.
[103] Han WJ, Zheng GQ, Liang YY, Dai K, Liu CT, Chen JB, Shen CY,
Peng XF, Fu P, Cao W, Li Q. HDPE solution crystallization induced
by electrospun PA66 nanober. Colloid Polym Sci 2011;289:843
8.
[104] Mago G, Kalyon DM, Fisher FT. Nanocomposites of polyamide-11
and carbon nanostructures: development of microstructure and
ultimate properties following solution processing. J Polym Sci, Part
B Polym Lett 2011;49:131121.
[105] Li LY, Yang Y, Yang GL, Chen XM, Hsiao BS, Chu B, Spanier JE, Li
CY. Patterning polyethylene oligomers on carbon nanotubes using
physical vapor deposition. Nano Lett 2006;6:100712.
[106] Li L, Wang W, Laird ED, Li CY, Defaux M, Ivanov DA.
Polyethylene/carbon nanotube nano hybrid shish-kebab obtained
by solvent evaporation and thin-lm crystallization. Polymer
2011;52:36338.
[107] Kim J, Hong SM, Kwak S, Seo Y. Physical properties of nanocomposites prepared by in situ polymerization of high-density
polyethylene on multiwalled carbon nanotubes. Phys Chem Chem
Phys 2009;11:108519.
[108] Kim J, Kwak S, Hong SM, Lee JR, Takahara A, Seo Y. Nonisothermal crystallization behaviors of nanocomposites prepared by in situ
polymerization of high-density polyethylene on multiwalled carbon nanotubes. Macromolecules 2010;43:1054553.
[109] Yang JH, Wang CY, Wang K, Zhang Q, Chen F, Du RN, Fu Q. Direct
formation of nanohybrid Shish-Kebab in the injection molded bar
of polyethylene/multiwalled carbon nanotubes composite. Macromolecules 2009;42:701623.
[110] Mai F, Wang K, Yao MJ, Deng H, Chen F, Fu QA. Superior reinforcement in melt-spun polyethylene/multiwalled carbon nanotube
ber through formation of a Shish-Kebab structure. J Phys Chem
B 2010;114:10693702.
[111] Li B, Li LY, Wang BB, Li CY. Alternating patterns on single-walled
carbon nanotubes. Nat Nanotechnol 2009;4:35862.
[112] Brosse AC, Tenc-Girault S, Piccione PM, Leibler L. Effect of
multi-walled carbon nanotubes on the lamellae morphology of
polyamide-6. Polymer 2008;49:46806.
[113] Garcia-Gutierrez MC, Nogales A, Rueda DR, Domingo C, GarciaRamos JV, Broza G, Roslaniec Z, Schulte K, Ezquerra TA.
X-ray microdiffraction and micro-Raman study on an injection
moulding SWCNT-polymer nanocomposite. Compos Sci Technol
2007;67:798805.
[114] Garcia-Gutierrez MC, Hernandez JJ, Nogales A, Pantine P, Rueda
DR, Ezquerra TA. Inuence of shear on the templated crystallization of poly(butylene terephthalate)/single wall carbon nanotube
nanocomposites. Macromolecules 2008;41:84451.
[115] Li MJ, Wang XB, Tian R, Wan L, Li SQ, Li Q. Carbon nanotubes periodically decorated by high-density polyethylene crystalline using
solution crystallization. Sci Chin Ser B 2009;52:90510.
[116] Zhang F, Xu Q, Zhang H, Zhang ZW. Polymer supermolecular
structures built on carbon nanotubes via a supercritical carbon
dioxide-assisted route. J Phys Chem C 2009;113:185315.
[117] Wang WR, Xie XM, Ye XY. Crystallization induced block copolymer
decoration on carbon nanotubes. Carbon 2010;48:16803.
[118] Thierry A, Straupe C, Wittmann JC, Lotz B. Organogelators and polymer crystallisation. Macromol Symp 2006;241:10310.

1453

[119] Ning NY, Luo F, Pan BF, Zhang Q, Wang K, Fu Q. Observation


of shear-induced hybrid shish kebab in the injection molded
bars of linear polyethylene containing inorganic whiskers. Macromolecules 2007;40:85336.
[120] Ning NY, Luo F, Wang K, Zhang Q, Chen F, Du RN, An CY, Pan BF, Fu
Q. Molecular weight dependence of hybrid shish kebab structure in
injection molded bar of polyethylene/inorganic whisker composites. J Phys Chem B 2008;112:141408.
[121] Su R, Wang K, Ning NY, Chen F, Zhang Q, Wang CY, Fu Q, Na B.
Orientation in high-density polyethylene/inorganic whisker composite bers as studied via polarized Fourier transform infrared
spectroscopy. Compos Sci Technol 2010;70:68591.
[122] Maiti P, Okamoto M. Crystallization controlled by silicate surfaces in nylon 6-clay nanocomposites. Macromol Mater Eng
2003;288:4405.
[123] Bai HW, Zhang WY, Deng H, Zhang Q, Fu Q. Control of crystal
morphology in poly(L-lactide) by adding nucleating agent. Macromolecules 2011;44:12337.
[124] Zhang QH, Lippits DR, Rastogi S. Dispersion, rheological aspects
of SWNTs in ultrahigh molecular weight polyethylene. Macromolecules 2006;39:65866.
[125] Zhang SJ, Lin W, Wong CP, Bucknall DG, Kumar S. Nanocomposites
of carbon nanotube bers prepared by polymer crystallization. ACS
Appl Mater Inter 2010;2:16427.
[126] Zhang SJ, Lin W, Yu XF, Wong CP, Cheng SZD, Bucknall DG.
Surface-induced polymer crystallization in high volume fraction
aligned carbon nanotube-polymer composites. Macromol Chem
Phys 2010;211:100311.
[127] Yang JH, Wang K, Deng H, Chen F, Fu Q. Hierarchical structure of
injection-molded bars of HDPE/MWCNTs composites with novel
nanohybrid shish-kebab. Polymer 2010;51:77482.
[128] Ning NY, Deng H, Luo F, Wang K, Zhang Q, Chen F, Fu Q. Effect of
whiskers nucleation ability and shearing function on the interfacial
crystal morphology of polyethylene (PE)/raw whiskers composites.
Compos Part B-Eng 2011;42:6317.
[129] Ning NY, Luo F, Wang K, Du RN, Zhang Q, Chen F, Fu Q.
Interfacial enhancement by shish-calabash crystal structure in
polypropylene/inorganic whisker composites. Polymer 2009;50:
38516.
[130] Wang W, Qi Z, Jeronimidis G. Studies on interface structure and
crystal texture of poly (ether-ether-ketone)-carbon bre composite. J Mater Sci 1991;26:591520.
[131] Machida S, Matsuda Y, Tasaka S. Gelation of a composite of ceramic
bers and polypropylene modied with maleic anhydrate in naphthenic oil. Chem Lett 2009;38:701.
[132] Ning NY, Deng H, Luo F, Zhang Q, Wang K, Chen F, Fu Q. The interfacial enhancement of LLDPE/whisker composites via interfacial
crystallization. Polym Adv Technol, doi:10.1002/pat.1894; in press.
[133] Prokhorow VV, Nitta K. The AFM observation of linear chain and
crystalline conformations of ultrahigh molecular weight polyethylene molecules on mica and graphite. J Polym Sci, Part B Polym Lett
2010;48:76677.
[134] Brinkmann M, Contal C, Kayunkid N, Djuric T, Resel R.
Highly oriented and nanotextured lms of regioregular poly
(3-hexylthiophene) grown by epitaxy on the nanostructured
surface of an aromatic substrate. Macromolecules 2010;43:
760410.
[135] Minus ML, Chae HG, Kumar S. Observations on solution crystallization of poly (vinyl alcohol) in the presence of single-wall carbon
nanotubes. Macromol Rapid Commun 2010;31:3106.
[136] Yu N, He LH, Ren YY, Xu Q. High-crystallization polyoxymethylene
modication on carbon nanotubes with assistance of supercritical
carbon dioxide: molecular interactions and their thermal stability.
Polymer 2011;52:47280.

[137] Tracz A, Kucinska


IK, Jeszka JK. Formation of highly ordered, unusually broad polyethylene lamellae in contact with atomically at
solid surfaces. Macromolecules 2003;36:101302.
[138] Takenaka Y, Miyaji H, Hoshino A, Tracz A, Jeszka JK, Kucinska I. Interface structure of epitaxial polyethylene crystal grown on HOPG and
MoS2 substrates. Macromolecules 2004;37:96679.
[139] Hobbs S. Row nucleation of isotactic polypropylene on graphite
bres. Nature 1971;234:123.
[140] Arbelaiz A, Fernandez B, Ramos JA, Mondragon I. Thermal and
crystallization studies of short ax bre reinforced polypropylene matrix composites: effect of treatments. Thermochim Acta
2006;440:11121.
[141] Assouline E, Pohl S, Fulchiron R, Gerard JF, Lustiger A, Wagner HD,
Marom G. The kinetics of alpha and beta transcrystallization in
bre-reinforced polypropylene. Polymer 2000;41:784354.

1454

N. Ning et al. / Progress in Polymer Science 37 (2012) 14251455

[142] Folkes MJ, Hardwick ST. The molecular weight dependence of transcrystallinity in bre reinforced thermoplastics. J Mater Sci Lett
1984;3:10713.
[143] Moon CK. Effect of molecular weight and ber diameter on the
interfacial behavior in glass ber/PP composites. J Appl Polym Sci
1998;67:11917.
[144] Chen L, Xiang YF, Wang K, Zhang Q, Du RN, Fu Q. Effects of matrix
molecular weight on structure and reinforcement of high density polyethylene/mica composites. Chin J Polym Sci 2011;29:377
89.
[145] Zheng XL, Xu Q. Comparison study of morphology and crystallization behavior of polyethylene and poly(ethylene oxide) on
single-walled carbon nanotubes. J Phys Chem B 2010;114:9435
44.
[146] Bessell T, Shortall JB. The crystallization and interfacial bond
strength of nylon 6 at carbon and glass bre surfaces. J Mater Sci
1975;10:203543.
[147] Frayer P, Lando J. Polymerization in electric elds of hexanethylenediammonium adipate crystallized on graphite bers. J Polym Sci,
Part B Polym Lett 1972;10:2934.
[148] Cai YQ, Petermann J, Wittich H. Transcrystallization in berreinforced isotactic polypropylene composites in a temperature
gradient. J Appl Polym Sci 1997;65:6775.
[149] Janeschitz-Kriegl H, Ratajski E, Stadlbauer M. Flow as an effective
promotor of nucleation in polymer melts: a quantitative evaluation.
Rheol Acta 2003;42:35564.
[150] Fukushima H, Ogino Y, Matsuba G, Nishida K, Kanaya T. Crystallization of polyethylene under shear ow as studied by time resolved
depolarized light scattering. Effects of shear rate and shear strain.
Polymer 2005;46:187885.
[151] Li Y, Pickering K, Farrell R. Determination of interfacial shear
strength of white rot fungi treated hemp bre reinforced
polypropylene. Compos Sci Technol 2009;69:116571.
[152] He LH, Zheng XL, Xu Q. Modication of carbon nanotubes using
poly(vinylidene uoride) with assistance of supercritical carbon dioxide: the impact of solvent. J Phys Chem B 2010;114:
525762.
[153] Wittmann JC, Lotz B. Epitaxial crystallization of polymers on
organic and polymeric substrates. Prog Polym Sci 1990;15:90948.
[154] Li H, Yan S. Surface-induced polymer crystallization and
the resultant structures and morphologies. Macromolecules
2011;44:41728.
[155] Sano M, Sasaki DY, Kunitake T. Structural study of polyethers
on graphite by scanning tunneling microscopy. Polymerizationinduced epitaxy. Macromolecules 1992;25:69619.
[156] Haubruge H, Daussin R, Jonas A, Legras R, Wittmann J, Lotz B.
Epitaxial nucleation of poly (ethylene terephthalate) by talc: structure at the lattice and lamellar scales. Macromolecules 2003;36:
44526.
[157] Gafur M. Structures and properties of the compression-molded
istactic-polypropylene/talc composites: effect of cooling and
rolling. Polym Degrad Stabil 2010;95:181825.
[158] Ferrage E, Martin F, Boudet A, Petit S, Fourty G, Jouffret F,
Micoud P, De Parseval P, Salvi S, Bourgerette C, Ferret J, SaintGerard Y, Buratto S, Fortune JP. Talc as nucleating agent of
polypropylene: morphology induced by lamellar particles addition
and interface mineral-matrix modelization. J Mater Sci 2002;37:
156173.
[159] Binsbergen FL. Natural and articial heterogeneous nucleation in
polymer crystallization. J Polym Sci Polym Sym 1977;59:1129.
[160] Duncan RK, Chen XG, Bult J, Brinson L, Schadler L. Measurement of the critical aspect ratio and interfacial shear strength
in MWNT/polymer composites. Compos Sci Technol 2010;70:
599605.
[161] Haque A, Ramasetty A. Theoretical study of stress transfer in carbon
nanotube reinforced polymer matrix composites. Compos Struct
2005;71:6877.
[162] Cho K, Li F. Reinforcement of amorphous and semicrystalline
polymer interfaces via in situ reactive compatibilization. Macromolecules 1998;31:7495505.
[163] Wagner H, Lourie O, Feldman Y, Tenne R. Stress-induced fragmentation of multiwall carbon nanotubes in a polymer matrix. Appl
Phys Lett 1998;72:18890.
[164] Peron B, Lowe A, Baillie C. The effect of transcrystallinity on the
interfacial characteristics of polypropylene/alumina single bre
composites. Compos Part A-Appl S 1996;27:83945.
[165] Barber AH, Cohen SR, Wagner HD. Measurement of carbon
nanotube-polymer interfacial strength. Appl Phys Lett 2003;82:
41402.

[166] Young K, Blighe FM, Vilatela JJ, Windle AH, Kinloch IA, Deng L, Young
RJ, Coleman JN. Strong dependence of mechanical properties on
ber diameter for polymer-nanotube composite bers: differentiating defect from orientation effects. ACS Nano 2010;4:698997.
[167] Cadek M, Coleman J, Barron V, Hedicke K, Blau W. Morphological and mechanical properties of carbon-nanotube-reinforced
semicrystalline and amorphous polymer composites. Appl Phys
Lett 2002;81:51235.
[168] Gou J, Liang Z, Zhang C, Wang B. Computational analysis of
effect of single-walled carbon nanotube rope on molecular interaction and load transfer of nanocomposites. Compos Part B-Eng
2005;36:52433.
[169] Ajayan PM, Schadler LS, Giannaris C, Rubio A. Single-walled carbon
nanotube-polymer composites: strength and weakness. Adv Mater
2000;12:7503.
[170] Coleman JN, Khan U, Blau WJ, Gunko YK. Small but strong: a review
of the mechanical properties of carbon nanotube-polymer composites. Carbon 2006;44:162452.
[171] Cooper CA, Cohen SR, Barber AH, Wagner HD. Detachment of nanotubes from a polymer matrix. Appl Phys Lett 2002;81:38735.
[172] Folkes M, Wong W. Determination of interfacial shear
strength in bre-reinforced thermoplastic composites. Polymer
1987;28:130914.
[173] Wu CM, Chen M, Karger-Kocsis J. Effect of micromorphologic
features on the interfacial strength of iPP/Kevlar ber microcomposites. Polymer 2001;42:199208.
[174] Hoecker F, Karger-Kocsis J. Effects of crystallinity and supermolecular formations on the interfacial shear strength and adhesion in
GF/PP composites. Polym Bull 1993;31:70714.
[175] Moon C. The effect of interfacial microstructure on the interfacial strength of glass ber/polypropylene resin composites. J Appl
Polym Sci 1994;54:7382.
[176] Shioya M, Takaku A. Estimation of bre and interfacial shear
strength by using a single-bre composite. Compos Sci Technol
1995;55:339.
[177] Wagner H, Eitan A. Interpretation of the fragmentation phenomenon in single-lament composite experiments. Appl Phys Lett
1990;56:19657.
[178] Kelly A, Tyson W. Tensile properties of bre-reinforced metals:
copper/tungsten and copper/molybdenum. J Mech Phys Solids
1965;13:32938, in1in2, 33950.
[179] Nagae S, Otsuka Y, Nishida M, Shimizu T, Takeda T, Yumitori S. Transcrystallization at glass bre/polypropylene interface and its effect
on the improvement of mechanical properties of the composites. J
Mater Sci Lett 1995;14:12346.
[180] Zhang M, Xu J, Zhang Z, Zeng H, Xiong X. Effect of transcrystallinity on tensile behaviour of discontinuous carbon bre
reinforced semicrystalline thermoplastic composites. Polymer
1996;37:51518.
[181] Amitay C, Sadovsky E, Wagner HD. Hardness and Youngs modulus
of transcrystalline polypropylene by Vickers and Knoop microindentation. J Polym Sci, Part B Polym Phys 1999;37:52330.
[182] Karger-Kocsis J. Interphase with lamellar interlocking and amorphous adherent-a model to explain effects of transcrystallinity. Adv
Compos Lett 2000;9:2257.
[183] Cho K, Kim D, Yoon S. Effect of substrate surface energy
on transcrystalline growth and its effect on interfacial adhesion of semicrystalline polymers. Macromolecules 2003;36:7652
60.
[184] Barber AH, Cohen SR, Kenig S, Wagner HD. Interfacial fracture
energy measurements for multi-walled carbon nanotubes pulled
from a polymer matrix. Compos Sci Technol 2004;64:22839.
[185] Ding W, Eitan A, Fisher F, Chen X, Dikin D, Andrews R, Brinson LC, Schadler LS, Ruoff RS. Direct observation of polymer
sheathing in carbon nanotube-polycarbonate composites. Nano
Lett 2003;3:15937.
[186] Blond D, Barron V, Ruether M, Ryan KP, Nicolosi V, Blau WJ, Coleman JN. Enhancement of modulus, strength, and toughness in poly
(methyl methacrylate)-based composites by the incorporation of
poly (methyl methacrylate)-functionalized nanotubes. Adv Funct
Mater 2006;16:160814.
[187] Coleman JN, Cadek M, Blake R, Nicolosi V, Ryan KP, Belton C, Fonseca
A, Nagy JB, Gunko YK, Blau WJ. High performance nanotubereinforced plastics: understanding the mechanism of strength
increase. Adv Funct Mater 2004;14:7918.
[188] Coleman JN, Cadek M, Ryan KP, Fonseca A, Nagy JB, Blau WJ, Ferreira
MS. Reinforcement of polymers with carbon nanotubes. The role of
an ordered polymer interfacial region. Experiment and modeling.
Polymer 2006;47:855661.

N. Ning et al. / Progress in Polymer Science 37 (2012) 14251455


[189] Liao K, Li S. Interfacial characteristics of a carbon nanotubepolystyrene composite system. Appl Phys Lett 2001;79:42257.
[190] Wong M, Paramsothy M, Xu X, Ren Y, Li S, Liao K. Physical interactions at carbon nanotube-polymer interface. Polymer
2003;44:775764.
[191] in het Panhuis M, Maiti A, Dalton AB, van den Noort A, Coleman JN,
McCarthy B, Blau WJ. Selective interaction in a polymer-single-wall
carbon nanotube composite. J Phys Chem B 2003;107:47882.
[192] Wei C, Srivastava D, Cho K. Structural ordering in nanotube polymer
composites. Nano Lett 2004;4:194952.
[193] Frankland S, Caglar A, Brenner D, Griebel M. Molecular simulation of the inuence of chemical cross-links on the shear
strength of carbon nanotube-polymer interfaces. J Phys Chem B
2002;106:30468.
[194] Leeuw TK, Tsyboulski DA, Nikolaev PN, Bachilo SM, Arepalli S, Weisman RB. Strain measurements on individual single-walled carbon
nanotubes in a polymer host: structure-dependent spectral shifts
and load transfer. Nano Lett 2008;8:82631.
[195] Schadler L, Giannaris S, Ajayan P. Load transfer in carbon nanotube
epoxy composites. Appl Phys Lett 1998;73:38424.
[196] Wang W, Murthy NS, Chae HG, Kumar S. Structural changes during deformation in carbon nanotube-reinforced polyacrylonitrile
bers. Polymer 2008;49:213345.
[197] Ma W, Liu L, Zhang Z, Yang R, Liu G, Zhang T, An X, Yi X, Ren Y,
Niu Z. High-strength composite bers: realizing true potential of
carbon nanotubes in polymer matrix through continuous reticulate architecture and molecular level couplings. Nano Lett 2009;9:
285561.
[198] Eitan A, Fisher F, Andrews R, Brinson L, Schadler L. Reinforcement
mechanisms in MWCNT-lled polycarbonate. Compos Sci Technol
2006;66:116273.
[199] Chang T, Jensen LR, Kisliuk A, Pipes R, Pyrz R, Sokolov A. Microscopic mechanism of reinforcement in single-wall carbon nanotube/polypropylene nanocomposite. Polymer 2005;46:43944.
[200] Blighe FM, Young K, Vilatela JJ, Windle AH, Kinloch IA, Deng L, Young
RJ, Coleman JN. The effect of nanotube content and orientation on
the mechanical properties of polymer-nanotube composite bers:
separating intrinsic reinforcement from orientational effects. Adv
Funct Mater 2011;21:36471.
[201] Minus ML, Chae HG, Kumar S. Interfacial crystallization in gel-spun
poly (vinyl alcohol)/single-wall carbon nanotube composite bers.
Macromol Chem Phys 2009;210:1799808.
[202] Liang S, Wang K, Chen D, Zhang Q, Du R, Fu Q. Shear enhanced interfacial interaction between carbon nanotubes and polyethylene and
formation of nanohybrid shish-kebabs. Polymer 2008;49:49259.
[203] Xiang YF, Hou ZC, Su R, Wang K, Fu Q. The effect of shear on mechanical properties and orientation of HDPE/mica composites obtained
via dynamic packing injection molding (DPIM). Polym Adv Technol
2010;21:4854.
[204] Choi WJ, Kim SC. Effects of talc orientation and non-isothermal
crystallization rate on crystal orientation of polypropylene in
injection-molded polypropylene/ethylene-propylene rubber/talc
blends. Polymer 2004;45:2393401.
[205] Kim GM, Lee DH, Hoffmann B, Kressler J, Stppelmann G.
Inuence of nanollers on the deformation process in layered silicate/polyamide-12 nanocomposites. Polymer 2001;42:
1095100.
[206] Sheng N, Boyce M, Parks D, Rutledge G, Abes J, Cohen R. Multiscale
micromechanical modeling of polymer/clay nanocomposites and
the effective clay particle. Polymer 2004;45:487506.
[207] Dasari A, Yu Z-Z, Mai Y-W, Cai G, Song H. Roles of graphite oxide,
clay and POSS during the combustion of polyamide 6. Polymer
2009;50:157787.
[208] Klein N, Marom G, Wachtel E. Microstructure of nylon 66 transcrystalline layers in carbon and aramid bre reinforced composites.
Polymer 1996;37:54938.
[209] Varga J, Karger-Kocsis J. The occurence of transcrystallization or
row-nucleated cylindritic crystallization as a result of shearing
in a glass-ber-reinforced polypropylene. Compos Sci Technol
1993;48:1918.
[210] Alon Y, Marom G. On the transition in high density polyethylene: the effect of transcrystallinity. Macromol Rapid Commun
2004;25:138791.

1455

[211] Li TQ, Zhang MQ, Zeng HM. Strong interaction at interface of carbon
ber reinforced aromatic semicrystalline thermoplastics. Polymer
1999;40:430713.
[212] Cai D, Song M. A simple route to enhance the interface between
graphite oxide nanoplatelets and a semi-crystalline polymer for
stress transfer. Nanotechnology 2009;20:315708/1315708.
[213] Byrne MT, Gunko YK. Recent advances in research on carbon
nanotube-polymer composites. Adv Mater 2010;22:167288.
[214] Ray S, Okamoto M. Polymer/layered silicate nanocomposites:
a review from preparation to processing. Prog Polym Sci
2003;28:1539641.
[215] Kim H, Abdala AA, Macosko CW. Graphene/polymer nanocomposites. Macromolecules 2010;43:651530.
[216] Pavlidou S, Papaspyrides C. A review on polymer-layered silicate
nanocomposites. Prog Polym Sci 2008;33:111998.
[217] Sengupta R, Bhattacharya M, Bandyopadhyay S, Bhowmick AK. A
review on the mechanical and electrical properties of graphite and
modied graphite reinforced polymer composites. Prog Polym Sci
2010;36:63870.
[218] Incardona S, Migliaresi C, Daniel Wagner H, Gilbert AH, Marom G.
The mechanical role of the bre/matrix transcrystalline interphase
in carbon bre reinforced J-polymer microcomposites. Compos Sci
Technol 1993;47:4350.
[219] Ruan S, Gao P, Yang X, Yu T. Toughening high performance ultrahigh
molecular weight polyethylene using multiwalled carbon nanotubes. Polymer 2003;44:564354.
[220] Wang Z, Ciselli P, Peijs T. The extraordinary reinforcing efciency
of single-walled carbon nanotubes in oriented poly (vinyl alcohol)
tapes. Nanotechnology 2007;18:455709/1455709.
[221] Yu MF, Files BS, Arepalli S, Ruoff RS. Tensile loading of ropes of single
wall carbon nanotubes and their mechanical properties. Phys Rev
Lett 2000;84:55525.
[222] Xu JZ, Chen T, Yang CL, Li ZM, Mao YM, Zeng BQ, Hsiao BS.
Isothermal crystallization of poly(L-lactide) induced by graphene
nanosheets and carbon nanotubes: a comparative study. Macromolecules 2010;43:50008.
[223] Hu X, An H, Li ZM, Geng Y, Li L, Yang C. Origin of carbon nanotubes
induced poly (L-lactide) crystallization: surface induced conformational order. Macromolecules 2009;42:32158.
[224] Yang H, Zhao XJ, Lu ZY, Yan FD. Temperature inuence
on the crystallization of polyethylene/fullerene nanocomposites: molecular dynamics simulation. J Chem Phys 2009;131:
234906-17.
[225] Yang H, Zhao XJ, Li ZS, Yan FD. Molecular dynamics simulations on
crystallization of polyethylene/fullerene nanocomposites. J Chem
Phys 2009;130:074902-16.
[226] Liu W, Yang CL, Zhu YT, Wang MS. Interactions between
single-walled carbon nanotubes and polyethylene/polypropylene/
polystyrene/poly (phenylacetylene) /poly (p-phen ylenevinylene)
considering repeat unit arrangements and conformations: a molecular dynamics simulation study. J Phys Chem B 2008;112:180311.
[227] Zheng QB, Xue QZ, Yan KO, Hao LZ, Li Q, Gao XL. Investigation of molecular interactions between SWNT and polyethylene/polypropylene/polystyrene/polyaniline molecules. J Phys
Chem B 2007;111:462835.
[228] Yang H, Chen Y, Liu Y, Cai WS, Li ZS. Molecular dynamics simulation of polyethylene on single wall carbon nanotube. J Chem Phys
2007;127:094902/194902.
[229] Wang XL, Lu ZY, Li ZS, Sun CC. Molecular dynamics simulation
study on controlling the adsorption behavior of polyethylene by
ne tuning the surface nanodecoration of graphite. Langmuir
2007;23:8028.
[230] Zhang XB, Li ZS, Yang H, Sun CC. Molecular dynamics simulations on
crystallization of polyethylene copolymer with precisely controlled
branching. Macromolecules 2004;37:7393400.
[231] Guo HX, Yang XZ, Li T. Molecular dynamics study of the behavior
of a single long chain polyethylene on a solid surface. Phys Rev E
2000;61:418593.
[232] Wei C. Radius and chirality dependent conformation of polymer
molecule at nanotube interface. Nano Lett 2006;6:162731.
[233] Zheng Q, Xue Q, Yan K, Gao X, Li Q, Hao L. Effect of chemisorption on
the interfacial bonding characteristics of carbon nanotube-polymer
composites. Polymer 2008;49:8008.

Vous aimerez peut-être aussi