You are on page 1of 15

Designing tough and fracture resistant polypropylene/multi wall

carbon nanotubes nanocomposites by controlling stereo-complexity


and dispersion morphology
Dibyendu Das, Bhabani K. Satapathy

Centre for Polymer Science and Engineering, Indian Institute of Technology Delhi, Hauz Khas, New Delhi 110016, India
a r t i c l e i n f o
Article history:
Received 7 June 2013
Accepted 18 August 2013
Available online 29 August 2013
Keywords:
Polypropylene
Multi wall carbon nanotubes
Tacticity
Toughness
Fracture
Crack propagation
a b s t r a c t
A remarkable toughness enhancement (>330%) of multi wall carbon nanotubes (MWCNT) lled stereo-
complex polypropylene (PP) matrix i.e. blend of isotactic-PP and syndiotactic-PP (70:30) with differences
in stereo-regularity has been observed. The enhancement has been correlated to quantiable morpholog-
ical parameters such as free-space lengths concerning dispersion and relatively greater reduction in crys-
tallite size/lamellar thickness. Systematic analysis of glass transition data and estimation of multi wall
carbon nanotubes induced reduction in interfacial polymer chain immobilization reiterates susceptibility
of polymer segments to ready-mobility. The extent of toughening has quantitatively been analyzed by
fracture-energy partitioning, essential work of fracture (EWF), approach enabling the detection of a
semi-ductile-to-tough-to-quasi-brittle transition in the MWCNT lled stereo-complex polypropylene.
Real-time fracture kinetics analysis revealed toughening mechanism to be primarily blunting-assisted;
an aspect also corroborated by extensive plastic ow without much energy dissipation in the inner frac-
ture process zone. Thus the study establishes a new pathway of tacticity-dened matrix modication to
toughen nanocomposites.
2013 Elsevier Ltd. All rights reserved.
1. Introduction
Difference in stereo-regularity (tacticity) induced matrix modi-
cation enabling the manipulation of the polymer nanocomposites
structure favorably towards enhanced energy dissipation (tough-
ening mode) is a theme not so outrightly attempted. This is be-
cause molecular level interaction involving topological
constraints, van der Waals forces, Lennard-Jones potential and Lon-
don dispersion forces may be very complex to explain the precise
nature of adhesive/cohesive interactions in understanding tough-
ening mechanisms. The topological attributes to the mechanism
of toughening has the signicance in the sense that the conformal
proximity of the zig-zag arrangements and the wrapping-up phe-
nomenon via the helical conformation of the polymer chain cong-
urations facilitate in effectively enhancing the stress transfer
mechanism by increasing the interfacial interaction. The size scale
advantage in this sense plays a crucial role since smaller dimension
of the hollow cylinders extend a larger amount of interfacial area
for molecular level interactions. This is because the effective vol-
ume fraction of the polymer chains that potentially can be ad-
sorbed per unit surface area of the reinforcing second phase is
much lower and hence may lead to promoting efcient polymer-
nanotube interaction. Such interactions involving less of loosely
entangled polymer chains renders greater extent of chain immobi-
lization enabling the interfacial polymer-nanotube network to act
as a composite phase and hence larger energy dissipation and
toughness [13]. The pioneering work of Lordi and Yao [4] postu-
lating the role of helical conformations in increasing the effective
binding of the polymer-nanotube interface and subsequently the
counter approach view point in evolving the theory of a strong
polymer-nanotube interface for enhanced toughening by Jiang
and Penn [5] reiterates the above understanding pertaining to
nanoscale reinforcement especially that of carbon nanotubes.
Wong et al. [6] have reported the determining role of polymer-
nanotube interfacial characteristics using molecular mechanics
simulations and elasticity calculations. Relatively higher interfacial
shear stresses in case of nanotubes than their micron sized coun-
terparts have reportedly been attributed to intimate solid phase
contact between polymer and nanotube at the molecular level.
Lee et al. in another fundamental work concerning the role of dis-
persion and exfoliation of single wall carbon nanotubes (SWCNT)
and multi wall carbon nanotubes (MWCNT) in polypropylene
(PP) matrix have postulated on the temperature dependence of
interfacial strength across polymer/nanotube interface and the
building-up of a three dimensional percolation network [7,8]. Cole-
mann et al. [9] in their pioneering work emphasized the role of an
0261-3069/$ - see front matter 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.matdes.2013.08.067

Corresponding author. Tel.: +91 11 26596043; fax: +91 011 26591421.


E-mail address: bhabani@polymers.iitd.ac.in (B.K. Satapathy).
Materials and Design 54 (2014) 712726
Contents lists available at ScienceDirect
Materials and Design
j our nal homepage: www. el sevi er . com/ l ocat e/ mat des
ordered interfacial stiff polymer region in enhancing the reinforce-
ment efciency, particularly with respect to enhancements in
modulus. Schaefer and Justice [10] while commenting on the elu-
sive nano effects have put forth an explanation based on small
angle X-ray scattering (SAXS), light scattering and electron imaging
studies for the discrepancy pertaining to mechanical property in-
crease in terms of ubiquitous large scale disorder of the ller dis-
persion morphology. Lin et al. [11] have reported regarding the
nanoplastic ow of glassy polymer chains interacting with
MWCNT in polymer nanocomposites where intrinsic differences
in the fundamental behavior of the entangled polymer chains with
respect to crazing and shear yielding mechanism determining the
overall extensibility/ductility of the nanocomposites were high-
lighted. Interestingly, the entangled network responses in polymer
nanocomposites have been proved not only to be responsible for
the mode of deformation but also their interaction with individual
nanotubes. Mu and Winey [12] have inferred the possible relation-
ship between enhanced load-transfer effectiveness and increased
polymer molecular weight with their molecular level implications
to the larger size of the polymer chain (radius of gyration-R
g
) rela-
tive to that of the nanoller. It was elaborated to the extent that
the criteria 2R
g
< D (D diameter of nanoller) for poor interaction
and 2R
g
> D for stronger interfacial interaction between Polymer/
carbon nanotube were semi-empirically proposed to qualitatively
explain the nature of interface.
Comprehensive investigations by Satapathy et al. [13] and Ganss
et al. [14] on the morphology and fracture behavior relationship in
isotactic polypropylene (i-PP)/MWCNT nanocomposites have re-
vealed a ductile-to-semi-ductile transition in the nanotube loading
range of 0.51.5 wt.%. Further the details of such a transition were
corroborated by fracture kinetics and localization of strain contours
analysis. Qualitatively, such fracture transitions have been experi-
mentally characterized by a transition from stable non-steady to
steady state crack tip opening displacement rate and thereby caus-
ing a delayed yielding phenomenon. Extending these observations,
further, to understand the MWCNT induced toughening effects in
i-PP/MWCNT nanocomposites the temperature dependence of
creep response was also intensively discussed by Ganss et al. [15].
Bao and Tjong [16] have assessed the strain rate and temperature
sensitivity of i-PP/MWCNT nanocomposites where the increase in
heat deection temperature and glass transition temperatures have
been reported to increase with MWCNT content and the yielding
phenomenon could be explained by a time dened Eyrings equa-
tion. In consequence to such observations by Bao and Tjong [16]
the reinforcing effects were observed to enhance at elevated tem-
peratures. Karevan et al. [17] have reported the dominant role of
interphase in dening the stress-transfer efciency while validating
the experimentallyobtainedtensilemoduli tothat of the predictions
from HalpinTsai model. The interrelationship of ber strength,
interface strengthandcritical lengthof the carbonnanotubes as pro-
posed in the KellyTyson model (arguably valid for nanocompos-
ites) further indicates that toughness as a bulk nanomaterial
response is substantially dependent on interface strength assuming
nanotube related parameters remain unchanged [18]. Prashantha
et al. [19] have reported on the enhanced impact resistance for
notched samples where nanotubes effectively limits the crack prop-
agation of PP/MWCNT nanocomposites prepared by masterbatch
dilutionroute. This was accompaniedwithanincrease inyieldstress
and lesser extent of reduction in ductility when compared to the
classical carbon ber reinforced PP composites.
In an effort to understand the mechanism of increase in
strength in polymercarbon nanotube nanocomposites interfacial
characteristics including restrained relaxation of polymer chains
in the interphase, tailored interfaces via surface modication of
MWCNT, role of nanoparticle size, loading and shape and evolution
of interface across polymer-carbon nanotubes have also been
discussed in the literature [2022]. Kovalchuk et al. [23] have
investigated on the puried and alkyl functionalized MWCNT lled
i-PP and syndiotactic polypropylene (s-PP) nanocomposites where
signicant improvement in nanocomposite plasticity and modulus
of s-PP was achieved. However, the possible role of varying the tac-
tility/stereo regularity of the polymer matrix as blend of the two
stereo-forms of PP on the mechanical performance of nanocompos-
ites has not been attempted yet, despite the theoretical possibility
that syndiotacticity of PP may enhance the toughness of PP/
MWCNT nanocomposites by interfering with the nature of the
interface because of the differences in the helical conformation of
isotactic and syndiotactic polypropylenes. Therefore, the present
paper deals with the assessment of the morphological and fracture
toughness properties of PP/MWCNT nanocomposites where the
conventional matrix (i-PP) based system is modied and the ma-
trix is replaced by a 70:30 blend of i-PP to s-PP giving rise to a ma-
trix with asymmetric stereo-complexity.
2. Experimental details
2.1. Materials and composites preparation
The details of the rawmaterials selected and the processing con-
ditions are given in Table 1 and Table 2. The polymer nanocompos-
ites viz. IPNC (i-PP/PP-g-MA/MWCNT) and ISPNC (i-PP/s-PP(70:30)/
PP-g-MA/MWCNT) were preparedby melt mixing of i-PP, s-PP, poly-
propylene graftedmaleic anhydride (PP-g-MA) withthe commercial
master batchPlasticyl PP2001 containing20 wt.%MWCNTina co-
rotating type twinscrewextruder (Steer Omega 20) at a screwspeed
of 250 rpm. The continuous strands obtained from the twin screw
extruder were later chopped in a granulator and dried in a vacuum
oven at 80 C for 2 h before the injection molding on an L&T Demag
injection molding machine (Model PFY 40-LNC4P) to obtain
80 mm 80 mm square plates of 1.5 mm thickness. The process-
ing conditions with temperature proles in twin screw extruder
and injection molding are given in Table 2. The composites designa-
tion and compositions are given in Table 3.
2.2. Morphology of the nanocomposites
Transmission electron microscopy (TEM) was carried out on
cryo-microtomed specimen sections of thickness of about 70
90 nm. The samples were cut from the central portion of the injec-
tion molded sample using a Leica Ultracut EM UC6/EM FC6 ultra-
microtome (Model Leica Mikrosysteme GmbH, Wien, Austria).
The ultramicrotome is equipped with a diamond knife with a cut
angle of 35. The TEM used is an EM 912 (LEO, Oberkochen, Ger-
many) operated at 120 kV.
2.3. Structural characterization by 2D wide angle X-ray (WAXD)
diffraction
WAXD measurements were done with an X-ray diffractometer
model number PW304060 Xpert PRO (Netherland) with 40 kV
voltage, 30 mA current and Cu Ka = 1.54 , in a 2h range from
550, to evaluate the crystallinity of the nanocomposites. The
crystallinity was calculated by applying the peak-area integration
method in the range of 2h = 540 (as typical for i-PP). The amor-
phous scattering curve was drawn by an approximation based on
standard experimental and theoretical principles.
2.4. Light microscopic measurement
The spherulitic morphology of i-PP, s-PP, IPNC and ISPNC are
studied using a Instec HCS-302 (Meiji Techno-Japan) microscope.
D. Das, B.K. Satapathy / Materials and Design 54 (2014) 712726 713
The sample was placed between a microscope glass slide and a
cover slip. The sample was heated up to 200 C and then kept for
2 min to ensure complete melting of the samples. Then samples
are cooled down to 130 C at a cooling rate of 5 C per minute. After
complete crystallization the images were captured, using a
2 mega-pixel camera attached to the microscope, with a resolution
of 20-fold magnication.
2.5. Thermal characterization
2.5.1. Differential scanning calorimetry (DSC)/Thermo gravimetric
analysis (TGA)
Differential scanning calorimetry (DSC) measurements were
carried out (in TA instruments, Q200) at a heating rate of 10 C/
min to obtain the crystallinity of the composites. The samples were
rst heated up to 220 C then allowed to stand still at that temper-
ature for 1 min (to relieve its thermal history) followed by cooling
down to 20 C at a cooling rate 10 C/min and then subsequently
heated up to 220 C at a heating rate 10 C/min. Apparent enthal-
pies of fusion have been calculated from the area under the exo-
thermic peaks and the crystallinity (%) of the i-PP, i-PP/s-PP
blends and its MWCNT lled nanocomposites have been deter-
mined using following equations:
X
c;iPP
DH
m;iPP
=DH
0
m;iPP
wt:%
iPP
1
X
c;sPP
DH
m;sPP
=DH
0
m;sPP
wt:%
sPP
2
X
c;blends
X
c;iPP
wt:%
iPP
X
c;sPP
wt:%
sPP
3
where DH
m
is the heat of fusion of i-PP, s-PP in the MWCNT nano-
composites and DH
o
m
is the heat of fusion of 100% crystalline i-PP
i.e., taken as 209 J/g and that of s-PP as 196 J/g [24]. The crystallite
sizes have been determined following Scherrers equation [7,25].
The referred equation for the crystallite size may be given as,
Table 1
Data sheet of raw materials and their characteristics.
Raw materials Grade Supplier Characteristics
Isotactic polypropylene homopolymer REPOL
H033MG
Reliance
Industries
Limited
Melt ow index of 3.30 g/10 min. at 230 C and 2.26 kg load; T
m
= 165 C
Syndiotactic polypropylene
homopolymer
452149 SigmaAldrich Melt ow index of 2.20 g/10 min. at 230 C and 2.26 kg load; T
m
= 125 C
PP-g-MA OPTIM P-
405
PLUSS Polymer Density (q) = 0.91 g/ml; T
m
= 163 C; melt ow index of 20 g/10 min at 190 C and 2.16 kg
load; Maleic anhydride content range 1.62.5%
Isotactic polypropylene/MWCNT
commercial master batch
PlastiCyl
PP 2001
Nanocyl 20 wt.% MWCNT in master batch, average diameter = 9.5 nm; average length = 1.5 lm and
purity > 90%
Table 2
Processing parameters used in extrusion and injection molding machine.
Extrusion temperature prole
Zones Barrel 2 Barrel 3 Barrel 4 Barrel 5 Barrel 6 Die adaptor Die head
Temperature C 180 190 200 210 220 225 230
Temperature prole set in injection molding machine
Feed (C) Z-I (C) Z-II (C) Z-III (C) Nozzle (C)
40 190 210 220 230
Processing conditions used in injection molding machine
Process parameter Value
Injection pressure 1059 bar
Injection time 4 s
Cooling time 25 s
Table 3
Details of the compositions and designation of i-PP/PP-g-MA/MWCNT nanocomposites (IPNC) and i-PP/s-PP/PP-g-MA/MWCNT nanocomposites (ISPNC).
Designation Compositions
i-PP (wt.%) i-PP/s-PP
a
(70:30) (wt.%) PP-g-MA (wt.%) MWCNT (wt.%)
i-PP 100.00 0.00 0.0
IPNC-0 (iPP/PP-g-MA) 95.00 5.00 0.0
IPNC-0.5 (iPP/PP-g-MA/0.5 wt.% MWCNT) 94.53 4.97 0.5
IPNC-1.0 (iPP/PP-g-MA/1.0 wt.% MWCNT 94.05 4.95 1.0
IPNC-1.5 (iPP/PP-g-MA/1.5 wt.% MWCNT) 93.58 4.92 1.5
IPNC-2.0 (iPP/PP-g-MA/2.0 wt.% MWCNT) 93.10 4.90 2.0
IPNC-3.0 (iPP/PP-g-MA/3.0 wt.% MWCNT) 92.15 4.85 3.0
ISPNC-0 (i-PP/s-PP/PP-g-MA) 95.00 5.00 0.0
ISPNC-0.5 (iPP/s-PP/PP-g-MA/0.5 wt.% MWCNT) 94.53 4.97 0.5
IPSNC-1.0 (iPP/s-PP/PP-g-MA/1.0 wt.% MWCNT) 94.05 4.95 1.0
IPSNC-1.5 (iPP/s-PP/PP-g-MA/1.5 wt.% MWCNT) 93.58 4.92 1.5
IPSNC-2.0 (iPP/s-PP/PP-g-MA/2.0 wt.% MWCNT) 93.10 4.90 2.0
IPSPN-3.0 (iPP/s-PP/PP-g-MA/3.0 wt.% MWCNT) 92.15 4.85 3.0
a
Stereo-complex Polypropylene matrix (70:30-Asymmetric composition).
714 D. Das, B.K. Satapathy / Materials and Design 54 (2014) 712726
L
crystallite

Kk
B2h cos h
4
where L is the crystallite size, k is wave length of radiation, B is full
width half maxima (FWHM) and K is proportionality constant (0.94
for FWHM of spherical crystals with cubic symmetry).
Furthermore the non-oxidative thermal stability of the nano-
composites has been measured using a Pyris 6 TGA instrument,
Perkin-Elmer. The thermo gravimetric analysis (TGA) were carried
out in nitrogen atmosphere at a heating rate 10 C/min in the tem-
perature range 30700 C.
2.6. Dynamic mechanical analysis (DMA)
Nanocomposite specimens with dimensions of 20 8 1.5 mm
3
were subjected to dynamic mechanical analysis (DMA) measure-
ments in single cantilever mode. The measurements were done in
the temperature range of 30 C to 165 C at a frequency of 1.0 rad/
s and heating rate of 5 K/min on an Q800 machine (TA Instruments,
USA) to characterize loss tangent (tand) for the qualitative estimation
of damping/energy dissipation characteristics and for quantitatively
ascertaining shift (if any) in glass transition temperature (T
g
) of the
MWCNT lled virgin isotactic (IPNC systems) and stereo-complex
(ISPNC system) polypropylene matrices.
2.7. Determination of fracture behavior
The fracture behavior of the nanocomposites has been mea-
sured following the plane stress essential work of fracture (EWF)
approach [26,27]. Rectangular injection molded bars of
80 mm 20 mm 1.5 mm dimension were used as specimens
for the preparation of double edge notch tension (DENT) specimens
for the measurement of fracture parameters, EWF and Non- essen-
tial work of fracture (N-EWF). The specimens were prepared in
such a manner that the injection molding direction and the direc-
tion of the application of uni-axial tensile force onto the bars re-
mained identical. The samples were pre-notched with different
ligament lengths varying from 210 mm. The fracture of these
pre-notched specimens was carried out using a universal tensile
testing machine (Zwick Z250) under constant extension speed
(2 mm/min) to obtain individual loaddisplacement curves. The
plane-stress essential work of fracture (EWF) method has been
used since conceptually it enables the distinguishing between
two terms representing the resistance to crack initiation (EWF: w
e
)
and resistance to crack propagation (N-EWF: bw
p
) corresponding
to inner fracture process zone (IFPZ) and outer plastic deformation
zone (OPDZ) respectively. The precondition for the validity of EWF
approach has been demonstrated by the self-similar nature of the
loaddisplacement diagrams of these nanocomposites. In this
study, the fracture mechanical tests (test speed: 2 mm/min, i.e.
e
9
= 0.033 mm/s, room temperature) were performed on double-
edge-notched-tension (DENT) specimen by a universal testing ma-
chine with mechanical grips. The clamp distance was 40 mm. For
notching, a special device with fresh razor blades (notch tip radius
of 0.20 lm) was used to realize that both notches are similar
sized. For each material, at least 8 specimens with different liga-
ment lengths were tested.
The total work of fracture W (kJ/m
2
) dissipated in a notched
specimen under plane-stress conditions can be partitioned into
two components, W
e
and W
p
characterizing the inner fracture
process zone (IFPZ) and outer plastic deformation zone (OPDZ)
respectively, as schematically shown in Fig. 1. Therefore,
W W
e
W
p
w
e
B l bw
p
B:l
2
5
where B, l and b are specimen thickness, ligament length and shape
factor of the plastic zone respectively. The specic work of fracture
w may be obtained on normalizing (dividing) W by the ligament
(notched) area i.e., B l . The relationship (with the quantities w
e
and bw
p
are given in N/mm and N/mm
2
units respectively) may
be represented as,
w w
e
bw
p
l 6
Based on the fact that the intrinsic fracture process takes place
in the inner fracture process zone (IFPZ), the term EWF, the essen-
tial work of fracture, is experimentally determined by extrapola-
tion of w as a function of l to zero ligament length. For the
quantitative determination of these fracture parameters (w
e
and
bw
p
), several similar sized specimens with different ligament
lengths are monotonically loaded to obtain several data points in
the plot of w versus l, the intercept and the slope of which gives
rise to w
e
(EWF: resistance to crack initiation) and bw
p
(N-EWF:
resistance to crack propagation) respectively [28].
2.8. Fractured surface morphology
The post-yield fracture surface morphologies of the DENT spec-
imens of MWCNT lled isotactic and stereo-complex (iso-syndio
blend) polypropylene matrices based nanocomposites have been
investigated using scanning electron microscopy (SEM) on a Zeiss
EVO-50 electron microscope to analyze the associated fracture mi-
cro mechanisms involving micro-deformation, micro-brillation,
shear yielding and layer peeling off characteristics. The failed
nanocomposite surfaces were gold sputter-coated prior to exami-
nation to make the surfaces conductive.
3. Results and discussion
3.1. Morphology of the nanocomposites
The distribution and dispersion of the MWCNT in i-PP and in
stereo-complex PP matrices (i-PP/s-PP: 70/30) are shown in
Fig. 2. It was observed that the MWCNT formed partially agglomer-
ated domains at the distribution level that are uniformly spaced in
the PP matrix. The state of dispersion involving entangled, curved
and partially looped MWCNT that are randomly spaced apart from
each other could be observed. The free space length (l
f
) in the ma-
trix are estimated by resorting to the principles proposed by Burris
[29] with the exception that the sides of all the maximum square
spaces are averaged linearly. Subsequently the inltrated free-
space lengths (l
inf
) have been estimated by locating the maximum
Fig. 1. Double edge notch tension (DENT) specimen showing inner fracture process
zone (IFPZ) and outer fracture process zone (OPDZ).
D. Das, B.K. Satapathy / Materials and Design 54 (2014) 712726 715
possible free spaces in between the individual MWCNTs while
excluding the squares that have sides less than or equal to the
diameter of the incorporated MWCNT. The diameter of the
MWCNT used, in the present study is in the range of 810 nm.
The ratio l
inf
/l
f
is taken as a quantitative measure to ascertain the
extent of dispersion. The higher the ratio, the better is the disper-
sion. From the Table 4, it could be well observed that the dispersion
of the MWCNT in the stereo-complex PP matrix is fairly uniform
where inltrated PP rich region could be observed in the inter-tube
spacing spanning with a dimension more than that of the MWCNT.
The increase in MWCNT content has led to the appearance of mas-
terbatch rich regions indicating inefcient inltration of PP chains.
This may be attributed to a phenomenon resembling agglomera-
tion of MWCNT (as seen in ISPNC-3.0 and is discussed in the
SEM micrographs). Interestingly the distributive mixing of IPNC
has been observed to be better than that of ISPNC since isomorph-
icity of the matrix PP may facilitate the easy distribution of the
MWCNT-rich regions of PP/MWCNT masterbatch via physical
entanglements of the loose PP chains at melt temperatures. How-
ever, the incorporation of PP with a difference in stereo-regularity
facilitates the dispersion of the MWCNT through the inherent ten-
dency of the stereo-complex PP blend tending to demix due to a
higher repulsive interaction parameter operating across the two
stereo regular forms of the PP matrices. The enthalpy of demixing
of s-PP and i-PP may well be construed from their differences in
solubility parameters (d) and their repulsive interaction parame-
ters (v). The solubility parameters of i-PP and s-PP at a tempera-
ture 453 K are reported to be 5.64 (cal/cm
3
)
1/2
and 5.76 (cal/
cm
3
)
1/2
respectively [30]. The solubility parameter of i-PP/s-PP
(stereo-complex) blend may be calculated, by assuming that the
segmental volume change is negligible due to mixing, using the
following equation:
d
iPP=sPP
/
1
d
iPP
1 /
1
d
sPP
7
The critical interaction parameter may theoretically be esti-
mated, while assuming monodispersity of the two components,
by the following equation [31]:
v 0:5N
0:5
1
N
0:5
2

2
; 8
where N
1
and N
2
are the degrees of polymerization of i-PP and s-PP
respectively. The v value so obtained is 10.02 10
4
which is in
striking contrast to reported value of v of 3.54 10
4
correspond-
ing to 50:50 blend composition [30]. It must also be noted that the
Fig. 2. Dispersion and distribution of MWCNT: TEM images of the nanocomposites (a) IPNC-0.5 at low magnication, (b) IPNC-0.5 at high magnication, (c) IPNC-1.0 at high
magnication, (d) ISPNC-0.5 at low magnication, (e) ISPNC-1.0 at low magnication, and (f) ISPNC-1.0 at higher magnication.
Table 4
The free space length in the matrix and inltrated free-space lengths of
nanocomposites.
l
f
(nm) l
inf
(nm) l
inf
/l
f
IPNC-0.5 911.01 36.68 0.040
IPNC-1.0 797.75 29.51 0.037
ISPNC-0.5 774.22 55.83 0.072
ISPNC-1.0 483.07 77.89 0.161
716 D. Das, B.K. Satapathy / Materials and Design 54 (2014) 712726
segmental dynamics of s-PP is reported to be 1.7 times lower than
that of i-PP at 500 K [32]. These estimations corroborate the fact
that the two stereo-regular phases of PP retain the differences in
their dynamics of miscibility and that may lead to a difference in
the stereo specic relaxation response causing an enhancement in
the extent of dispersion of MWCNT.
3.2. Structural characterization by 2D wide angle X-ray diffraction
(WAXD)
The structural characteristics of the IPNC (i-PP/PP-g-MA/
MWCNT nanocomposites) and ISPNC (i-PP/s-PP/PP-g-MA/MWCNT
nanocomposites) have been obtained from WAXD of the injection
molded specimens. The X-ray diffraction plots in terms of intensity
(I) versus 2h measured in the 2h range of 550 are shown in Fig. 3.
Four distinct peaks at 2h of 14.08, 16.83, 18.45 and 21.7 corre-
sponding to the (110), (040), (130) and (111)/(041) net planes
have been observed in both, i-PP and its nanocomposites. These
peaks crystallographically correspond to the monoclinic a-crystal-
line phase with complete absence of the b-crystalline phase, which
shows two strong peaks at 2h of 16.2 and 21.2[33,34]. But in case
of ISPNC two additional peaks at 2h of 12.23 and 24.70 have been
observed which indicate the characteristic peaks of (200), (310/
400) plane of helical form I of s-PP [35,36]. The crystallinity results
obtained from WAXD and DSC (2nd heating curve) are given in Ta-
ble 5. It was observed that the percentage crystallinity of ISPNC-0
and its nanocomposites are much lower than IPNC-0 and its
nanocomposites.
3.3. Light optical microscopy
The optical micrographs of the virgin polymer matrix and nano-
composite melts cooled to a temperature of 130 C are shown in
Fig. 4. It has been observed that incorporation of MWCNT into i-
PP matrix led to a distinct reduction in the spherulites radius, i.e.
from 24.48 lm in IPNC-0 to 14.36 lm in IPNC-0.5. On increas-
ing the MWCNT content to IPNC-1.0 and IPNC-3.0 the spherulites
sizes tended to reduce further. In contrast the incorporation of
MWCNT into the stereo-complex matrix i.e. ISPNC-0 has led to sig-
nicant changes in the spherulitic dimensions, where the spheru-
lites have appeared to be much ner with worm-like structures
distributed uniformly in the relatively darker amorphous PP phase.
Such alterations in the bulk spherulitic morphology may be attrib-
uted to the relatively slower crystallization process of the syndio-
tactic phase of the stereo-complex PP matrix of ISPNC based
systems, when compared to the virgin i-PP matrix of IPNC. The
substantial reduction in spherulitic dimensions observed in ISPNC
may be construed to have an analogy with relatively stiffer crystal-
line phase of PP dispersed in a softer amorphous phase of PP. Such
a scenario leads to the understanding that stiffer crystalline micro-
domains dispersed in a randomly entangled amorphous PP matrix,
may lead to a morphology that is conducive to toughening.
3.4. Thermal and mechanical properties of the nanocomposites
3.4.1. Differential scanning calorimetry (DSC)
The effects of MWCNT on the melting behavior of i-PP, in IPNC
and ISPNC are shown in Fig. 5. It has been observed that in IPNC the
melt temperature and crystallinity remains nearly unaffected with
the addition of MWCNT indicating the absence of MWCNT induced
changes in crystalline packing of the base matrix, i.e. i-PP. Interest-
ingly the nature of the melting peaks corresponding to IPNC and
ISPNC showed characteristic difference in their corresponding
endotherms. In IPNC a single melting peak at 165 C is observed
without any appreciable MWCNT-induced shift in the T
m
. On the
other hand ISPNC showed two distinct endothermic transitions at
120125 C and 155165 C corresponding to stable helical-1
(2:1) [37] and double helix (3:1) conformations of s-PP and i-PP
respectively. Furthermore on increasing the MWCNT content
above 0.5 wt.% MWCNT, a shift of the melting endothermic peak
to a lower temperature of 155 C could be observed. Such a de-
crease by 10 C in the T
m
may be attributed to faster rate of crys-
tallization indicating possible role of MWCNT as a-nucleating
agents. Typically faster nucleation process facilitates the reduction
of the crystallite size as has already been observed from the crys-
tallite sizes determined by using Scherrers equation based on
WAXD data for ISPNC nanocomposites [38]. Such a reduction
may eventually lead to closer crystalline packing that may be con-
strued as a consequence of stereo-complexity. Since such a phe-
nomenon of a-nucleation could not be observed in IPNC unlike
ISPNC, the possibility of higher extent of dispersion of MWCNT in
ISPNC may be conceptually presumed. Theoretically, when two
isomorphic phases are melt blended with differences in their melt-
ing points (T
m
), the possibility of a nanoscopic inclusion getting
dispersed becomes a diffusion controlled process. Such a diffusion
controlled process may facilitate by the partial melting of s-PP
phase around 120 C and thereby may improve the dispersion
of MWCNT. Thus these calorimetric observations reiterate the rel-
atively better state of dispersion morphology in ISPNC as compared
to IPNC systems that are already discussed (Fig. 1).
3.4.2. Dynamic mechanical analysis of the nanocomposites
3.4.2.1. Loss-tangent (tand) response and T
g
. The results from solid
state dynamic mechanical analysis in terms of variation of loss tan-
gent (tand) with temperature of IPNC and ISPNC are shown in
Fig. 6a and b respectively. From the gure it could be clearly ob-
served that in case of both IPNC and ISPNC transitions occur at
around 20 C to 25 C, which correspond to the glass-rubber tran-
5 10 15 20 25 30 35 40 45 50
IPNC-3.0
IPNC-2.0
IPNC-1.5
IPNC-1.0
IPNC-0.5
I
n
t
e
n
s
i
t
y

(
a
.
u
.
)
2
i-PP
IPNC-0
5 10 15 20 25 30 35 40 45 50
ISPNC-3.0
ISPNC-2.0
ISPNC-1.5
ISPNC-1.0 I
n
t
e
n
s
i
t
y

(
a
.
u
.
)
ISPNC-0.5
ISPNC-0
2
(a) (b)
Fig. 3. Wide angle X-ray diffractogram (WAXD) measurement of nanocomposites: variation of intensity with 2h (a) IPNC (b) ISPNC.
D. Das, B.K. Satapathy / Materials and Design 54 (2014) 712726 717
sition of the i-PP. The magnitude of tand increases signicantly
with the incorporation of 0.5 wt.% MWCNT in the IPNC-0. But with
the further loading of MWCNTs the tand values decreases substan-
tially. In the entire composition range the tand values remained in
between that of IPNC-0.5 and i-PP with an exception of IPNC-3.0.
This increment qualitatively attributed to a better dispersion of
MWCNT up to 0.5 wt.% loading. Similarly in case of ISPNC the mag-
nitude of tand is highest in case of 1.0 wt.% MWCNT content
(ISPNC-1.0) and it decreases with the further loading of MWCNTs.
Qualitatively this indicated that in case of ISPNC dispersion of
MWCNT is better up to a higher percentage loading. Interestingly
it could be also observed from Fig. 6 that in case of IPNC the
glass-rubber transition i.e. glass transition temperature (T
g
) re-
mained almost unaffected whereas in case of ISPNC the T
g
is shifted
slightly to relatively lower values. Theoretically, the T
g
of the ma-
trix should be expected to increase with the incorporation of the
nanollers, since the nanollers mostly inltrate into the amor-
phous region imposing restrictions to the chain segmental mobil-
ity. Similar results for reduction in T
g
based on dynamic
mechanical properties in case of polycarbonate/MWCNT nanocom-
posites has been reported and attributed to processing induced
thermal degradation of polymer chain [39]. But in our case the on-
set degradation temperature obtained from the TGA (listed in Ta-
ble 5) is increased by 8 C for IPNC and by 20 C for ISPNC.
These results however do not support the assumption of process-
ing induced thermal degradation of i-PP matrix.
3.4.2.2. Inuence of interface and MWCNT content on T
g
. The promi-
nent reduction in T
g
of the ISPNC systems relative to the IPNC
nanocomposites may be attributed to an overall reduction in the
percentage crystallinity estimated from WAXD and supported by
the crystallinity data obtained by calorimetric (DSC) measure-
ments. To further analyze the phenomenon two complementary
approaches were adopted by resorting to the assumptions that T
g
(of the nanocomposites) may get affected by the corresponding
thermal responses (a) of the nanoinclusion with respect to matrix
and (b) of the immobilized matrix polymer chains by dispersed
MWCNT. In the former approach the reduction in T
g
equivalent
due to the incorporated MWCNT (T
g2
), that is responsible for the
manifestation of the overall decrease in T
g
of the nanocomposite,
may theoretically be estimated by Fox equation [40]. The Fox equa-
tion may be stated as below
1
T
g

w
1
T
g1

w
2
T
g2
9
where T
g
, T
g1
and T
g2
correspond to the glass transition temperature
of the nanocomposites, matrix without MWCNT, reduced T
g
-equiv-
alent due to incorporated MWCNT respectively, w
1
and w
2
are the
weight fraction of the matrix and nanotube respectively.
However, in the later approach the immobilized volume frac-
tion (
T
) may be estimated from Kerners equation [41]. The
T
may be construed as responsible for storage modulus (E
0
) enhance-
ment. The
T
is taken into account in order to quantify approxi-
mately the interfacial polymer chain immobilization inuence on
the overall T
g
of the nanocomposites. The rearranged Kerners
equation for composite to estimate the immobilized volume frac-
tion of polymer chains (
T
) may be given as below.
/
T

E
c
T E
m
T
E
m
T aE
c
T
10
where E
m
(T) and E
c
(T) are the storage moduli of the matrix (IPNC-0
and ISPNC-0 matrices) and their nanocomposites respectively at a
reference temperature (i.e. at T = 25 C) and
a
24 5t
m

7 5t
m
11
where m
m
is Poissons ratio of the composites and is taken as 0.30 in
the present case. For the estimation of T
g
of the immobilized phase
in IPNC-0.5 and IPNC-1.0, IPNC-0 has been taken as the matrix,
whereas for the same compositions of ISPNC series ISPNC-0 has
been taken as the matrix. Since theoretically the unaffected T
g
in
a nanocomposite system indicates the unaffected bulk relaxation
phenomenon of the polymer chains, hence the nanocomposites
with T
g
identical to that of the unlled matrix, may be assumed
to be the system with virtual absence of any immobilization/mobi-
lization effects. The T
g
estimates obtained by adhering to the above
two approaches are given in Table.5. It could be clearly observed
that the T
g
of the immobilized PP phase in IPNC in the composition
range of IPNC-0.5 to IPNC-1.0 remained at 10 C. Such observa-
tions indicated the formation of an ordered interphase that is
dynamically (unaltered) resembling i-PP chains. In contrast in
ISPNC the T
g
of the immobilized phase remained in the range of
12 C to +4.7 C. These indicate that the incorporation of
MWCNT into ISPNC matrix readily enhances the segmental mobil-
ity. This fundamentally reiterates the possibility of MWCNT induced
nano-structural reorganization facilitating an increase in the overall
amorphous free volume space (fraction); which in turn, microme-
chanically, may lead to corresponding ductile toughening effects.
Table 5
Nanotube induced immobilized volume fraction at the interface, thermal property and crystalline morphological properties.

T
@ 25 C

T
g
(C) T
g
#
(C) T
g
##
(C) TGA Crystallite size
$
(nm) Crystallinity from
WAXD (%)
Crystallinity from
DSC (%)
L
lamellar
(nm)
T
onset
(C) (040)
i-PP
IPNC-0 0 23 443.6 3.85 55.10 53.78 1.997
IPNC-0.5 0.103 25 2 10 449.7 3.80 56.03 54.06 1.983
IPNC-1.0 0.076 21 2 10 451.5 3.83 56.84 55.52 1.991
IPNC-1.5 0.056 23 450.8 3.70 56.83 55.21 1.993
IPNC-2.0 0.102 23 449.6 3.79 56.94 56.63 1.994
IPNC-3.0 0.166 23 450.3 3.77 58.46 58.02 1.990
ISPNC-0 0 24 416.1 3.66 35.10 30.67 1.987
ISPNC-0.5 0.026 26 2 12 440.9 3.47 34.61 25.57 1.927
ISPNC-1.0 0.049 20 1 5 443.5 3.12 34.64 30.20 1.930
ISPNC-1.5 0.053 22 3 9 425.2 3.49 35.83 29.20 1.933
ISPNC-2.0 0.071 24 422.2 4.25 35.94 29.61 1.924
ISPNC-3.0 0.113 23 10 17 422.2 3.90 35.46 31.10 1.925

T
= Volume fraction of immobilized polymer chains at 25 C; T
g
= glass transition temperature of the composites, T
g
#
= Reduced equivalent glass transition temperature due
to incorporation of MWCNT; T
g
##
= glass transition temperature of immobilized volume fraction of polymer chain (immobilized volume fraction of polymer are converted to
corresponding weight fraction of the polymer to apply Fox equation), $ = crystallite size estimated from Scherrers equation, L
lamellar
= lamellar thickness deduced from the
Thomson Gibbs equation.
718 D. Das, B.K. Satapathy / Materials and Design 54 (2014) 712726
3.5. Fracture behavior of the nanocomposites
3.5.1. Load displacement diagram and the validity of the EWF
approach
The loaddisplacement diagram of i-PP, IPNC-0 (i-PP/PP-g-MA),
ISPNC-0 (i-PP/s-PP/PP-g-MA) and its MWCNT lled nanocompos-
ites are shown in Fig. 7. All the composites viz. IPNC and ISPNC
and their neat components have not only shown thermoplastic
behavior but also have shown self-similar nature of the force
displacement diagrams, indicating pre-conditional validity of
EWF approach. It is also clear from the loaddisplacement diagram
that displacement becomes higher in case of i-PP/PP-g-MA as
compared to i-PP although maximum load required for the stable
crack propagation remain almost unaffected. In case of IPNC-0.5
(i-PP/PP-g-MA/0.5 wt.% MWCNT) the displacement is higher as
compared to the neat component, but further addition of
MWCNT affect the displacement as well as the maximum load
for the stable crack propagation. Interestingly, in the stereo-com-
plex matrix system, ISPNC-0 the displacement becomes nearly
double as compared to its neat counterpart (i.e. IPNC-0) and re-
mains unaffected up to ISPNC-1.5. The plane stress criteria for
the applicability of post-yield fracture mechanics (PYFM) concept
is ensured by the Hills analysis [42] as shown in Fig. 8. The anal-
ysis revealed that the net section stress (r
n
) remained indepen-
dent of the ligament length (l). The full yielding of the entire
ligament length region in the DENT specimens occurred at max-
imum load (F
max
) and prior to the resumption of crack propaga-
tion, which is visually ensured. The total works of fractures (W)
for various nanocomposites were obtained by integration of the
total area under the loaddisplacement diagrams. After normali-
Fig. 4. Light optical micrographs of (a) i-PP, (b) s-PP (c) IPNC-0, (d) ISPNC-0, (e) IPNC-0.5, (f) ISPNC-0.5, (g) IPNC-1.0, (h) ISPNC-1.0.
D. Das, B.K. Satapathy / Materials and Design 54 (2014) 712726 719
zation by ligament area (B. l) the specic work of fracture (w) is
obtained.
3.5.2. Work of fracture-composition relationship and transition in
fracture modes
The specic work of fracture as a function of ligament length (l)
for i-PP, IPNC and ISPNC are shown in Fig. 9. The linear t of the
data points across the valid range of l leads to the determination
of slope as non-essential work of fracture (N-EWF; bw
p
) and inter-
cept as essential work of fracture (EWF; w
e
) for each composition.
The variation of w
e
and bw
p
as a function of MWCNT content is
shown in Fig. 10. It has been observed that with the incorporation
of 5 wt.% PP-g-MA into i-PP as the matrix, i.e. in IPNC-0, w
e
in-
creased by 125% with respect to i-PP which is followed by a sharp
drop of 70% in w
e
value in the nanocomposites containing
0.5 wt.% MWCNT (IPNC-0.5). On further addition MWCNT into
the nanocomposites, w
e
values are found to remain broadly unaf-
fected till 1.5 wt.% of MWCNT followed by a linear increase till
IPNC-3. In contrast blending of i-PP matrix with 30 wt.% of s-PP
(ISPNC-0) lead to an enhancement in w
e
by 135%, whereas on
incorporation of MWCNT into ISPNC as the matrix till ISPNC-1.5
a consistent reduction in w
e
could be observed. However, despite
the reduction in w
e
of ISPNC the gross magnitude of the same re-
mained well above their corresponding IPNC counterparts. These
inevitably indicate an intrinsic enhancement in the resistance to
crack initiation of the nanocomposites based on stereo-complex
PP-matrix, i.e. ISPNC when compared to i-PP based IPNC systems.
Mechanistically, these observations indicate the non-linear depen-
dency of the crack growth prior to failure. The resistance to crack
propagation (bw
p
) increased by 9% and 77% for IPNC-0 and
IPNC-0.5 respectively, when compared to i-PP matrix (Fig. 10b).
Such a remarkable toughening may be attributed to an improved
state of dispersion of MWCNT in the polymer as evident from the
TEM micrographs (Fig. 1). These ndings are in striking contrast
to the earlier reports where an increase in bw
p
by 15% was re-
ported in case of i-PP/MWCNT nanocomposites in the absence of
PP-g-MA [13]. The role of PP-g-MA in improving the dispersion
of MWCNT in polymer matrix was reported independently by Pra-
santha et al. [43] and Yang et al. [44] while investigating the
mechanical and toughness properties of nylon and PP nanocom-
posites. However upon further increasing the amount of MWCNT
up to 1.5 wt.% as in IPNC-1.5, the bw
p
decreases by 57% compared
to IPNC-0.5. Following such a reduction, upon further addition of
MWCNT the magnitude of bw
p
remained nearly unaffected.
Interestingly in ISPNC-0.5 and ISPNC-1.0 the bw
p
increased by
123% and 339% respectively when compared to the ISPNC-0
matrix. The maxima in bw
p
in ISPNC-1.0 indicate the composition
with maximum resistance to crack propagation that essentially ac-
counts for the energy dissipation in the outer-plastic deformation
zone (OPDZ). However, on further increasing the MWCNT content
to ISPNC-1.5 and ISPNC-2.0 a sharp reduction in bw
p
by 60% was
observed. The magnitude of w
e
showed a nearly linear decrease
with the increase in MWCNT content from ISPNC-0.5 to ISPNC-
1.5 indicating a systematic decrease in the resistance to crack ini-
tiation of the nanocomposites. Theoretically resistance to crack ini-
tiation has a qualitative correspondence to the energy dissipated in
80 90 100 110 120 130 140 150 160 170 180
IPNC-3.0
IPNC-2.0
IPNC-1.5
IPNC-1.0
IPNC-0.5
IPNC-0
i-PP
Temperature ( C)
H
e
a
t

f
l
o
w

(
e
n
d
o
)

[


H
(
J
/
g
)
]

110 120 130 140 150 160 170
Temperature (

C)
Helical form I
melting of s-PP
Shift of T
m
a- ISPNC-0
b- ISPNC-0.5
c- ISPNC-1.0
d- ISPNC-1.5
e- ISPNC-2.0
f- ISPNC-3.0
a
b
c
d
e
f
H
e
a
t

f
l
o
w

(
e
n
d
o
)

[


H
(
J
/
g
)
]

(a) (b)
Fig. 5. Differential scanning calorimetry (DSC) heat scans for the nanocomposites (a) IPNC (b) ISPNC.
-25 0 25 50 75 100 125 150
0.00
0.02
0.04
0.06
0.08
0.10
0.12
0.14
0.16
0.18
Temperature ( C) Temperature ( C)
i-PP
IPNC-0
IPNC-0.5
IPNC-1.0
IPNC-1.5
IPNC-2.0
IPNC-3.0
t
a
n

-25 0 25 50 75 100 125 150


0.00
0.02
0.04
0.06
0.08
0.10
0.12
0.14
0.16
0.18
0.20
0.22
0.24

t
a
n

ISPNC-0
ISPNC-0.5
ISPNC-1.0
ISPNC-1.5
ISPNC-2.0
ISPNC-3.0
(a) (b)
Fig. 6. Variation of loss tangent (tand) with temperature (T) of the nanocomposites (a) IPNC (b) ISPNC.
720 D. Das, B.K. Satapathy / Materials and Design 54 (2014) 712726
0
0
50
100
150
200
250
300
350
400
450
500
550
600
650
700
750
L
O
A
D

(
N
)
DISPLACEMENT (mm)
3.89 mm
4.67 mm
6.49 mm
7.46 mm
9.56 mm
i-PP
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 0 1 2 3 4 5 6 7 8 9
0
50
100
150
200
250
300
350
400
450
500
550
600
650
700
DISPLACEMENT (mm)
L
O
A
D

(
N
)
3.65 mm
4.28 mm
5.91 mm
6.74 mm
7.78 mm
9.39 mm
IPNC-0
1 2 3 4 5 6 7 8 9 10 11
0
100
200
300
400
500
600
DISPLACEMENT (mm)
L
O
A
D

(
N
)
3.81 mm
4.86 mm
5.74 mm
6.79 mm
7.32 mm
9.48 mm
IPNC-0.5
0 1 2 3 4 5 6
0
75
150
225
300
375
450
525
600
675
L
O
A
D

(
N
)
DISPLACEMENT (mm)
3.89 mm
4.99 mm
5.98 mm
7.34 mm
9.69 mm
IPNC-3.0
0 4 8 12 16 20 24
0
75
150
225
300
375
450
DISPLACEMENT (mm)
L
O
A
D

(
N
)
1.87 mm
3.66 mm
4.67 mm
5.94 mm
6.61 mm
7.38 mm
ISPNC-0
0 2 4 6 8 10 12 14 16 18 20
0
50
100
150
200
250
300
350
400
450
500
550
L
o
a
d

(
N
)
Displacement (mm)
2.35 mm
3.87 mm
4.83 mm
6.65 mm
8.34 mm
ISPNC-1.0
Fig. 7. Self-similarity of loaddisplacement diagrams.
0
20
40
60
80
100
i-PP
IPNC-0
IPNC-0
IPNC-1.0
IPNC-1.5
IPNC-2.0
IPNC-3.0
N
e
t

s
e
c
t
i
o
n

s
t
r
e
s
s

(
N
/
m
m
2
)
Ligament length (mm)
(a)
3 4 5 6 7 8 9 10 1 2 3 4 5 6 7 8 9 10 11
0
10
20
30
40
50
60
70
80
90
100
Ligament length (mm)
N
e
t

s
e
c
t
i
o
n

s
t
r
e
s
s

(
N
/
m
m
2
)
ISPNC-0
ISPNC-0.5
ISPNC-1.0
ISPNC-1.5
ISPNC-2.0
ISPNC-3.0
(b)
5 .
Fig. 8. Hills analysis plot: variation of net section stress (r
n
) with ligament length (l) (a) IPNC (b) ISPNC.
D. Das, B.K. Satapathy / Materials and Design 54 (2014) 712726 721
the inner fracture process zone (IFPZ). On a critical look into the
data, it was observed that ISPNC-0 and ISPNC-2.0 showed w
e
-max-
ima and w
e
-minima respectively. In contrast, ISPNC-1.0 showed an
intermediate value of w
e
that is 60% lower than that of ISPNC-0.5
and 40% higher than that of ISPNC-1.5. These observations imper-
atively indicate the existence of a semi-ductile-to-tough-to-quasi-
brittle transition in ISPNC in the composition range of 0.5
1.5 wt.% of MWCNT, unlike the failure transition involving defor-
mation with plastic-ow to deformation without plastic-ow in
IPNC. The semi-ductile, tough and quasi-brittle natures of the
respective ISPNC compositions are characterized by the high-w
e
/
low-bw
p
, high-w
e
/high-bw
p
and low-w
e
/low-bw
p
combina-
tions of the fracture mechanics parameters. Comprehensively such
a nature of the failure transition in ISPNC underlines the role of tac-
ticity-dened modication of the polymer matrix on the fracture
toughness of MWCNT lled polymer nanocomposites. Furthermore
the fracture surface morphology of these compositions by SEM
have revealed a systematic transition in the nature of their failure
characteristics and are discussed in a subsequent section.
3.6. Morphology and fracture toughness correlation
The observed signicant and moderate enhancements in tough-
ness due to tacticity differences in the stereo-complex PP matrix
(ISPNC) and in the conventional (single form crystalline matrix)
i-PP (IPNC) may be correlated to morphological parameters; qual-
itatively with extent of dispersion and quantitatively with the
lamellar thickness (L
lamellar
) of the PP chains. Several authors have
attributed toughness and stiffness enhancements to dispersion and
distribution uniformity of nanotubes [32,45,46]. The length-scale
(radius of gyration of polymer chains and nanotube dimensions)
and time-scale (segmental relaxation of polymer and mobility of
nanotube within the framework of entangled polymer chains) cor-
respondences in nanocomposites are theoretically well demon-
strated and was reported that a fundamentally different
toughening mechanism operates in these class of nano-structured
materials where the singular mobility of the nano-inclusion in the
polymer matrix under deformation denes the energy dissipation
capability of the material [47]. Interestingly the preceding argu-
ment seems to be in agreement to the fact that the singular mobil-
ity of MWCNT may get accentuated by the increase in the overall
amorphous nature of the matrix. This is well in agreement to our
observations manifested in the form of signicant reduction in T
g
and reduction in crystallite size (as obtained from Scherrers equa-
tion) upon blending of s-PP, i.e. upon rendering the nature of ma-
trix-polymer stereo-complex. To further correlate the
morphological parameters, composition specic lamellar thick-
nesses (L
lamellar
) was estimated following the ThomsonGibbs
equation and was analyzed vis--vis the fracture parameters w
e
and bw
p
values. The ThomsonGibbs equation may be given as,
L
lamellar

2rT
0
m
DH
m
T
0
m
T
m

12
where r is lamellar surface free energy (0.122 J m
2
), DH
v
is melt-
ing enthalpy of lamellar with innite thickness (192.28 10
6
-
J m
3
), T
0
m
is equilibrium melting temperature [48,49]. The
calculated lamellar thickness for the investigated IPNC and ISPNC
3 4 5 6 7 8 9 10
25
50
75
100
125
150
175
200
225
250
275
300
i-PP
IPNC-0
IPNC-0.5
IPNC-1.0
IPNC-1.5
IPNC-2.0
IPNC-3.0
S
p
e
c
i
f
i
c

w
o
r
k

o
f

f
r
a
c
t
u
r
e

Ligament length (mm)
(a)
1 2 3 4 5 6 7 8 9 10
50
100
150
200
250
300
350
400
450
500
550
600
650
Ligament length (mm)
S
p
e
c
i
f
i
c

w
o
r
k

o
f

f
r
a
c
t
u
r
e

(
N
/
m
m
)
ISPNC-0
ISPNC-0.5
ISPNC-1.0
ISPNC-1.5
ISPNC-2.0
ISPNC-3.0
(b)
(
N
/
m
m
)
Fig. 9. Variation of specic work of fracture (w) with ligament length for (a) IPNC (b) ISPNC.
Fig. 10. Variation of plane stress fracture parameters with MWCNT content (a) essential work of fracture (EWF; w
e
) and (b). non-essential work of fracture (N-EWF; bw
p
).
722 D. Das, B.K. Satapathy / Materials and Design 54 (2014) 712726
compositions are shown in Table 5. It could be clearly observed that
the w
e
-essential work of fracture has a direct correspondence to
lamellar thickness indicating that the resistance to stable crack ini-
tiation has a direct correlation to lamellar thickness and crystallite
arrangement [50]. However, the inverse correlation to bw
p
may the-
oretically be construed as lamellar break-up/destruction phenome-
non leading to an increase in resistance to stable crack propagation.
The scenario may be simplistically explained as the availability of
more of random polymer chains segments in the larger frame of
the material architecture facilitating energy dissipation mechanism
by promoting the singular mobility of the embedded MWCNT in the
entangled polymer chain networks with substantially lesser
hindrance.
3.7. Kinetics of crack propagation
The investigated nanocomposites showed ductile yielding,
though the nature of the ductile response remained dependent
on MWCNT content and therefore this observation encouraged
the authors to probe into the nature of such transitions in failure
dynamics in real-time framework following the kinetics of crack
growth approach. In both type of nanocomposites the matrix-dom-
inant ductile-yielding behavior has been observed to get altered
with the state of dispersion/distribution of nanoscopic inclusions
and hence with their consequent effects on morphological changes.
The determination of kinetics parameters of crack growth has been
done by using an optical video monitoring system in combination
with digital image correlation (DIC) techniques. The real-time
crack growth has further been monitored with periodic recording
of crack extension (Da), crack tip opening displacement (CTOD,
d) and CTOD rate (dd/dt) data at various deformation stages as
shown in Fig. 11. The systems i-PP (IPNC-0) and IPNC-2.0/IPNC-
3.0 have undergone much faster crack extension than the other
IPNC systems while the slowest crack extension was registered
for IPNC-0.5. In contrast a drastic retardation (>5-fold) of the crack
extension could be observed in ISPNC when compared to i-PP
(IPNC-0), an observation solely attributed to tacticity induced ste-
reo-complexity of the matrix. On incorporation of MWCNT to
ISPNC-0 matrix the crack extension rates have been observed to
get further retarded, especially at larger time scales, with ISPNC-
0.5 exhibiting the slowest crack extension rate. Interestingly
ISPNC-2.0/ISPNC-3.0 has also been observed to undergo failure at
much faster rate than ISPNC-0 matrix, an observation resembling
IPNC-2.0/IPNC-3.0. These observations corroborate the intrinsic
correspondence between poor dispersion and faster crack exten-
sion leading to catastrophic failure. The CTOD data, indicating the
inherent blunting efciency of material, of IPNC revealed a similar
trend to that of Da showing the lowest CTOD-rate of IPNC-0.5. In
contrast the CTOD-rate of ISPNC showed nearly identical values
irrespective of the extent of MWCNT incorporation. Since isolated
analysis of the kinetic parameters may lead to discrepancies quan-
titatively, the analysis of the fracture parameters has been carried
out in a cooperative fashion by analyzing the plots of dd/dt versus
Da.
The variations of CTOD-rate (dd/dt) with crack extension (Da)
for IPNC and ISPNC are shown in Fig. 11. It has been observed that
barring i-PP and IPNC-0 all the investigated nanocomposites failed
within Da 2.0 mm, i.e. at one third of the ligament length
l 6 mm. Interestingly such a range of Da when compared to the
displacement measured in the loaddisplacement diagrams for
l 6 mm remains well beyond the load-maxima in the failure of
DENT specimens. However, the maximum crack extension for i-
PP and IPNC-0 exceeds the generally observed extent of Da to fail-
ure for the nanocomposites irrespective of tacticity modication
i.e. stereo-complexity. Further the dd/dt has been observed to be
leveled off within dd/dt 0.04 mm/s. It must be mentioned that
the leveling off limit of dd/dt is little (1.5 times) higher than the
strain rate (e
9
= 0.033 mm/s) at which the fracture experiments
have been conducted. Based on these observations, the kinetically
dened failure map of the nanocomposites may be divided into
four failure-regimes as indicated in Fig. 11. The four macro-failure
regimes and their associated deformation attributes may be sum-
marized as given in Table 6.
It was observed that in IPNC and ISPNC systems the failure map
remained mostly conned in regime-II indicating the possibility of
MWCNT induced toughening, though the extent of toughening and
their underlying mechanistic attributes may not be easy to quan-
tify. However, from the plots it is amply evident that in ISPNC,
the CTOD-rate has been found to be more uniform as compared
to IPNC systems. Additionally the nature of clustering of the mea-
sured data points in regime-II essentially indicates the inherent
dependence of the macroscopic kinetic parameters on the morpho-
logical attributes of the systems. For ISPNC the less random clus-
tering of the data points may imply a better state of dispersion
unlike their IPNC counterparts. A more comprehensive analysis
based on qualitative estimation of J-integral and tear modulus re-
sponses may lead to further details of the inuence of stereo-com-
plexity (tacticity-dened blending) of matrix on fracture toughness
response of these nanocomposites, an aspect which will be fol-
lowed up in future.
3.8. Fracture surface morphology by SEM
The fracture surface morphology responsible for the associated
micro-deformation and crack propagation mechanisms of the i-PP
and the stereo-complex PP blend (i-PP/s-PP: 70/30) i.e. ISPNC-0 as
the matrices and their MWCNT lled nanocomposites have been
evaluated by scanning electron microscopy and is shown in
0.00
0.02
0.04
0.06
0.08
0.10
0.12
Regime-I
Regime-IV
Regime-III Regime-II
Steady state d/dt for i-PP/ PP-g-MA
Steady state d/dt for nanocomposites
C
T
O
D

r
a
t
e

d


/
d
t

[
m
m
/
s
]
Crack extension a (mm)
i-PP
IPNC-0
IPNC-0.5
IPNC-1.0
IPNC-1.5
IPNC-2.0
IPNC-3.0
Steady state d/dt for i-PP
0 1 2 3 4 5 6 0 1 2 3 4 5 6
0.00
0.02
0.04
0.06
0.08
0.10
0.12
Regime-I
Regime-IV
Regime-III
Regime-II
Steady state d/dt for
nanocomposites
Steady state d/dt for i-PP/ PP-g-MA
C
T
O
D

r
a
t
e

d


/
d
t

[
m
m
/
s
]
Crack extensiona (mm)
IPNC-0
ISPNC-0
ISPNC-0.5
ISPNC-1.0
ISPNC-1.5
ISPNC-2.0
ISPNC-3.0
(a) (b)
Fig. 11. Kinetics of crack propagation of the nanocomposites crack tip opening displacement (CTOD) rate dependence on crack growth (a) IPNC nanocomposites (b) ISPNC
nanocomposites.
D. Das, B.K. Satapathy / Materials and Design 54 (2014) 712726 723
Fig. 10. It was clearly observed that IPNC-0 showed a homogeneous
surfacemorphologywhereas IPNC-0.5showeddistinct signs of peel-
ing-off tendencies of fracture surface layers indicating the fracture
resistance tobe controlledmore bycrackpropagatingalong the frac-
tureplane. The observationis alsocorroboratedbythe fact that there
is a small decrease in the w
e
(essential work of fracture process) of
IPNC-0.5 as compared to IPNC-0 as shown in Fig. 12. Interestingly
the failure surface morphology of the stereo-complex blend matrix
i.e., ISPNC-0, the surface showed signs of alternately patterned
peeled-off strips indicating the presence of a quasi-co-continuous
morphologydue to asymmetric stereocomplexity. Onincorporating
0.5 wt.% of MWCNT i.e., the nanocomposite ISPNC-0.5 revealed
intensive brillation process involving thin brils and briler strips
to be involved in the course of fracture under uni-axial tension. On
further increasing the MWCNT content to 1.0 wt.% i.e., in ISPNC-
1.0, the nanocomposite showed relatively distinct and dense popu-
lationof brillatedstructures spanningacross the crackalsoreferred
as crack-bridging, highly stretched brils (0.10.3 lm) indicating
strong interfacial interaction between PP and MWCNT due to ste-
reo-complexity and very long brillar strips (210 lm) undergo-
ing fragmentation to lead to smaller strips (24 lm) and
brils(0.81.0 lm) at their tips prior to failure. On increasing the
MWCNT content to 3.0 wt.%, as in ISPNC-3.0, the fracture surface re-
vealed i-PP/MWCNT (80:20) masterbatch rich domains as the re-
gions responsible for failure initiation. Such a failure mechanism
may be attributed to highly localized morphology with improper
dispersion and distribution since the MWCNT rich region does not
get inltrated effectively by the excessively added i-PP or s-PP so
as to lead to a state of real dilution. Hence the fracture surface in
these regions showed the relative absence of any signs of plastic-
deformation/shear-yielding prior to failure causing quasi-brittle/
ductility-arrested failure characteristics.
Table 6
The macro-failure regimes and their associated deformation.
Regimes dd/dt Da Deformation attributes
Regime-I High (>0.04 mm/s) Low (<2.0 mm) Ductile yielding/plastic ow
Regime-II Low (<0.04 mm/s) Low (<2.0 mm) Tough
Regime-III Low (<0.04 mm/s) High (>2.0 mm) Resistant to crack propagation
Regime-IV High (>0.04 mm/s) High (>2.0 mm) Catastrophic failure
Fig. 12. SEM investigation fracture surfaces of nanocomposites (a) IPNC 0.0, (b) IPNC 0.5, (c) ISPNC 0.0, (d) ISPNC 0.5, (e) ISPNC 1.0, (f) ISPNC 3.0.
724 D. Das, B.K. Satapathy / Materials and Design 54 (2014) 712726
4. Conclusions
A new phenomenon of matrix stereo-complexity enhancing the
fracture toughness by >330% and improvement of the extent of dis-
persion of MWCNT in polypropylene has been observed. Dispersion
morphology was quantitatively explained in terms of the ratio of
free space length to inltrated free space lengths. Such morpholog-
ical consequences leading to a decrease in T
g
of MWCNT lled ste-
reo-complex or tacticity modied blends of polypropylene relative
to the nanocomposites based on isotactic-only matrix by 28 C
has been observed. The fracture toughness performances of these
nanocomposites have been observed to be signicantly affected
as estimated by essential work of fracture approach. A semi-duc-
tile-to-tough-to-quasi-brittle transition in ISPNC in the composi-
tion range of 0.51.5 wt.% of MWCNT has been observed that
was primarily attributed to a corresponding transition in the nat-
ure of deformation occurring with (in ISPNC) and without (in IPNC)
plastic ows respectively. A set of distinctly evolved and novel cri-
teria for the semi-ductile, tough and quasi-brittle natures ISPNC
compositions have emerged that are characterized by high-w
e
/
low-bw
p
, high-w
e
/high-bw
p
and low-w
e
/low-bw
p
combina-
tions of the fracture mechanics parameters respectively. A vefold
retardation of the crack extension in tacticity modied relative to
the unmodied virgin isotactic PP matrix could be observed. The
enhanced dispersion in stereocomplex PP matrix based nanocom-
posite have shown a direct bearing on the CTOD rate as reected
in the blunting response. Fracture surface morphology revealed
signicant crack-bridging, brils stretching across a growing crack
accompanied by extensive brillation and formation of microstrips
getting terminally fragmented into micro-brils as the possible
toughening mechanism in stereocomplex PP matrix based MWCNT
lled nanocomposites.
Our study demonstrates a new pathway for toughening of car-
bon nanotube based thermoplastic polymer nanocomposites by
blending of two different stereo-regular forms of the chemically
identical polymer matrix. The conceptual validation has success-
fully been established with regard to PP/MWCNT nanocomposites,
where remarkable toughness enhancement has been corroborated
by signicant (nearly vefold) retardation of the crack-growth
phenomenon assisted by efcient crack-blunting mechanism.
Acknowledgements
This paper is dedicated to late Prof. Roland Weidisch. Authors
gratefully acknowledge the nancial support, vide Grant No.SR/
S3/ME/0010/2008 by Department of Science and Technology New
Delhi, India.
References
[1] Jancar J, Douglas JF, Starr FW, Kumar SK, Cassagnau P, Lesser AJ, et al. Current
issues in research on structureproperty relationships in polymer
nanocomposites. Polymer 2010;51:332143.
[2] Zerda AS, Lesser AJ. Intercalated clay nanocomposites: morphology,
mechanics, and fracture behavior. J Polym Sci Part B: Polym Phys
2001;39:113746.
[3] Andrews R, Weisenberger MC. Carbon nanotube polymer composites. Curr
Opin Solid State Mater Sci 2004;8:317.
[4] Lordi V, Yao N. Molecular mechanics of binding in carbon-nanotubepolymer
composites. J Mater Res 2000;15:27709.
[5] Jiang KR, Penn LS. Improved analysis and experimental evaluation of the single
lament pull-out test. Compos Sci Technol 1992;45:89103.
[6] Wong M, Paramsothy M, Xu XJ, Ren Y, Li S, Liao K. Physical interactions at
carbon nanotube-polymer interface. Polymer 2003;44:775764.
[7] Lee G-W, Jagannathan S, Chae HG, Minus ML, Kumar S. Carbon nanotube
dispersion and exfoliation in polypropylene and structure and properties of the
resulting composites. Polymer 2008;49:183140.
[8] Lee SH, Cho E, Jeon SH, Youn JR. Rheological and electrical properties of
polypropylene composites containing functionalized multi-walled carbon
nanotubes and compatibilizers. Carbon 2007;45:281022.
[9] Coleman JN, Cadek M, Ryan KP, Fonseca A, Nagy JB, Blau WJ, et al.
Reinforcement of polymers with carbon nanotubes. The role of an
ordered polymer interfacial region. Experiment and modeling. Polymer
2006;47:855661.
[10] Schaefer DW, Justice RS. How Nano Are Nanocomposites? Macromolecules
2007;40:850117.
[11] Smith Jr JG, Connell JW, Delozier DM, Lillehei PT, Watson KA, Lin Y, et al. Space
durable polymer/carbon nanotube lms for electrostatic charge mitigation.
Polymer 2004;45:82536.
[12] Mu M, Winey KI. Improved load transfer in nanotube/polymer composites
with increased polymer molecular weight. J Phys Chem C 2007;111:179237.
[13] Satapathy BK, Gan M, Weidisch R, Ptschke P, Jehnichen D, Keller T, et al.
Ductile-to-semiductile transition in PP-MWNT nanocomposites. Macromol
Rapid Commun 2007;28:83441.
[14] Gan M, Satapathy BK, Thunga M, Weidisch R, Ptschke P, Jehnichen D.
Structural interpretations of deformation and fracture behavior of
polypropylene/multi-walled carbon nanotube composites. Acta Mater
2008;56:224761.
[15] Gan M, Satapathy BK, Thunga M, Weidisch R, Ptschke P, Janke A.
Temperature dependence of creep behavior of PPMWNT nanocomposites.
Macromol Rapid Commun 2007;28:162433.
[16] Bao SP, Tjong SC. Mechanical behaviors of polypropylene/carbon nanotube
nanocomposites: the effects of loading rate and temperature. Mater Sci Eng A
2008;485:50816.
[17] Bhuiyan MA, Pucha RV, Worthy J, Karevan M, Kalaitzidou K. Dening the lower
and upper limit of the effective modulus of CNT/polypropylene composites
through integration of modeling and experiments. Compos Struct
2013;95:807.
[18] Wichmann MHG, Schulte K, Wagner HD. On nanocomposite toughness.
Compos Sci Technol 2008;68:32931.
[19] Prashantha K, Soulestin J, Lacrampe MF, Krawczak P, Dupin G, Claes M.
Masterbatch-based multi-walled carbon nanotube lled polypropylene
nanocomposites: assessment of rheological and mechanical properties.
Compos Sci Technol 2009;69:175663.
[20] Coleman JN, Khan U, Blau WJ, Gunko YK. Small but strong: a review of the
mechanical properties of carbon nanotubepolymer composites. Carbon
2006;44:162452.
[21] Du J-H, Bai J, Cheng H-M. The present status and key problems of carbon
nanotube based polymer composites. eXPRESS Polym Lett 2007;1(5):
25373.
[22] Desai AV, Haque MA. Mechanics of the interface for carbon nanotubepolymer
composites. Thin-Wall Struct 2005;43:1787803.
[23] Kovalchuk AA, Shevchenko VG, Shchegolikhin AN, Nedorezova PM, Klyamkina
AN, Aladyshev AM. Effect of carbon nanotube functionalization on the
structural and mechanical properties of polypropylene/MWCNT composites.
Macromolecules 2008;41:753642.
[24] Zhang X, Zhao Y, Wang Z, Zheng C, Dong X, Su Z, et al. Morphology and
mechanical behavior of isotactic polypropylene (iPP)/syndiotactic
polypropylene (sPP) blends and bers. Polymer 2005;46:595665.
[25] Yang L, Zhang Z, Wang X, Chen J, Li H. Effect of ultrasonic irradiation on the
microstructure and the electric property of PP/CPP/MWNT composites. J Appl
Polym Sci 2013;128:151020.
[26] Brny T, Czigny T, Karger-Kocsis J. Application of the essential work of
fracture (EWF) concept for polymers, related blends and composites: a review.
Prog Polym Sci 2010;35:125787.
[27] Clutton EQ. ESIS TC4 experience with the essential work of fracture method.
ESIS Publ 2000;27:18799.
[28] Benkhenafou F, Chemingui M, Fayolle B, Verdu J, Ferouani AK, Lefebvre JM.
Fracture behaviour of a polypropylene lm. Mater Des 2011;32:15159.
[29] Khare HS, Burris DL. A quantitative method for measuring nanocomposite
dispersion. Polymer 2010;51:71929.
[30] Clancy TC, Ptz M, Weinhold JD, Curro JG, Mattice WL. Mixing of isotactic and
syndiotactic polypropylenes in the melt. Macromolecules 2000;33:945263.
[31] Vukovic R, Bogdanic G, Karasz FE, MacKnight WJ. Phase behavior and
miscibility in binary blends containing polymers and copolymers of styrene,
of 2,6-dimethyl-1,4-phenylene oxide, and of their derivatives. J Phys Chem Ref
Data 1999;28:85168.
[32] Lippow SM, Qiu X, Ediger MD. Effect of tacticity on the segmental dynamics of
polypropylene melts investigated by 13C-nuclear magnetic resonance. J Chem
Phys 2001;115:49615.
[33] Logakis E, Pollatos E, Pandis C, Peoglos V, Zuburtikudis I, Delides CG, et al.
Structure-property relationships in isotactic polypropylene/multi-walled
carbon nanotubes nanocomposites. Compos Sci Technol 2010;70:32835.
[34] Ma L-F, Wang W-K, Bao R-Y, Yang W, Xie B-H, Yang M-B. Toughening of
polypropylene with b-nucleated thermoplastic vulcanizates based on
polypropylene/ethylenepropylenediene rubber blends. Mater Des 2013;51:
53643.
[35] Gregoriou VG, Kandilioti G, Bollas ST. Chain conformational transformations in
syndiotactic polypropylene/layered silicate nanocomposites during
mechanical elongation and thermal treatment. Polymer 2005;46:1134050.
[36] Lovinger AJ, Lotz B, Davis DD. Interchain packing and unit cell of syndiotactic
polypropylene. Polymer 1990;31:22539.
[37] Gatos KG, Kandilioti G, Galiotis C, Gregoriou VG. Mechanically and thermally
induced chain conformational transformations between helical form I and
trans-planar form III in syndiotactic polypropylene using FT-IR and Raman
spectroscopic techniques. Polymer 2004;45:445364.
D. Das, B.K. Satapathy / Materials and Design 54 (2014) 712726 725
[38] Fereidoon A, Ahangari MG, Saedodin S. Thermal and structural behaviors of
polypropylene nanocomposites reinforced with single-walled carbon nanotubes
by melt processing method. J Macromol Sci Part B 2008;48:196211.
[39] Ptschke P, Bhattacharyya AR, Janke A, Goering H. Melt mixing of
polycarbonate/multi-wall carbon nanotube composites. Compos Interfaces
2003;10:389404.
[40] Rocco AM, Pereira RP, Felisberti MI. Miscibility, crystallinity and
morphological behavior of binary blends of poly(ethylene oxide) and
poly(methyl vinyl ethermaleic acid). Polymer 2001;42:5199205.
[41] Puknszky B. Interfaces and interphases in multicomponent materials: past,
present, future. Eur Polymer J 2005;41:64562.
[42] Hill R. On discontinuous plastic states, with special reference to localized
necking in thin sheets. J Mech Phys Solids 1952;1:1930.
[43] Prashantha K, Soulestin J, Lacrampe MF, Claes M, Dupin G, Krawczak P. Multi-
walled carbon nanotube lled polypropylene nanocomposites based on
masterbatch route: Improvement of dispersion and mechanical properties
through PP-g-MA addition. eXPRESS Polym Lett 2008;2:73545.
[44] Yang B-X, Shi J-H, Pramoda KP, Goh SH. Enhancement of the mechanical
properties of polypropylene using polypropylene-grafted multiwalled carbon
nanotubes. Compos Sci Technol 2008;68:24907.
[45] Alig I, Ptschke P, Lellinger D, Skipa T, Pegel S, Kasaliwal GR, et al.
Establishment, morphology and properties of carbon nanotube networks in
polymer melts. Polymer 2012;53:428.
[46] Kutvonen A, Rossi G, Puisto SR, Rostedt NKJ, Ala-Nissila T. Inuence of
nanoparticle size, loading, and shape on the mechanical properties of polymer
nanocomposites. J Chem Phys 2012;137. 214901/1-/8.
[47] Gersappe D. Molecular mechanisms of failure in polymer nanocomposites.
Phys Rev Lett 2002;89:058301.
[48] Sun X, Shen H, Xie B, Yang W, Yang M. Fracture behavior of bimodal
polyethylene: effect of molecular weight distribution characteristics. Polymer
2011;52:56470.
[49] Viana JC, Cunha AM, Billon N. The thermomechanical environment and the
microstructure of an injection moulded polypropylene copolymer. Polymer
2002;43:418596.
[50] Lach R, Grellmann W. Time-and temperature-dependent fracture mechanics of
polymers: general aspects at monotonic quasi-static and impact loading
conditions. Macromol Mater Eng 2008;293:55567.
726 D. Das, B.K. Satapathy / Materials and Design 54 (2014) 712726

You might also like