You are on page 1of 10

Article

Polymers and Polymer Composites


2021, Vol. 29(8) 1 212­–1221
Mechanical, thermal and dynamic- ª The Author(s) 2020
Article reuse guidelines:

mechanical studies of functionalized sagepub.com/journals-permissions


DOI: 10.1177/0967391120965115
journals.sagepub.com/home/ppc
halloysite nanotubes reinforced
polypropylene composites

Prakash Krishnaiah1,2 , Sivakumar Manickam2,


Chantara Thevy Ratnam3, MS Raghu4, L Parashuram4,
S Prasanna Kumar5 and Byong-Hun Jeon1

Abstract
Mechanical, dynamic-mechanical and thermal performance of polypropylene (PP) composites which are composed of
(3-Aminopropyl) triethoxysilane (APTES) functionalized Halloysite nanotubes (HNTs) were investigated. Functionalization
of HNTs was confirmed by the presence of amine stretching peaks in the FTIR spectrum. A decrease in the agglomeration
and high dispersion of APTES-HNTs across the PP matrix was confirmed by scanning electron micrographs (SEM). The
mechanical properties of APTES-HNT-PP polymer composites were superior over their unmodified counterparts. Tensile
properties such as maximum strength, Young’s modulus and impact strength were significantly enhanced by 28%, 45% and
60% respectively, with 6 wt% incorporation of surface-modified HNTs into PP matrix. A drastic improvement of stiffness
and thermal stability of composites was noted with the incorporation of APTES modified HNTs into PP polymer. Dif-
ferential scanning calorimetry (DSC) analysis showed a total increase of 22% in the crystallinity of clay polymer nano-
composite after filled with surface-modified HNTs. Overall, the outcome of this research confirms the modification of the
surface of HNTs with a silane coupling agent, which enhances the mechanical and thermal performance of PP composites
incorporated HNTs.

Keywords
Functionalization, halloysite nanotubes, (3-aminopropyl) triethoxysilane, polymer composite, polypropylene

Introduction
Generation of polymer nanocomposites with the reinforcement of nanomaterials has an important deal of attention due to
that nanomaterials can enhance the mechanical, physical and thermal characteristics even at a lower addition of nano-
filler.1–3 The expansion of mechanical and thermal performance of resultant polymer nanocomposites could be governed
by several factors including the type and aspect ratio of nano-filler, degree of dispersion, interfacial adhesion of nano-filler
and polymer matrix as well as the orientation of polymer matrix.4,5 Among these factors, the degree of dispersion and
interfacial bonding between nano-filler and polymer play a critical role in the mechanical and thermal performance of the
resultant polymer nanocomposites.6,7
HNTs are form of multi-walled hollow tubular clay mineral having a similar stoichiometric composition of kaolinite
with the structural formula of Al2Si2O5(OH)4.nH2O which are occurring naturally. The outer surface of HNTs has SiO2
bonds, while Al2O3 bonds are present in the inner lumen of the nanotubes.8–10 In recent years, HNTs have received great

1
Department of Earth Resources and Environmental Engineering, Hanyang University, Seoul, Republic of Korea
2
Department of Chemical and Environmental Engineering, University of Nottingham Malaysia, Semenyih, Malaysia
3
Malaysian Nuclear Agency, Selangor, Malaysia
4
Department of Chemistry, New Horizon College of Engineering, Bangalore, India
5
Indian Academy Degree College, Bangalore, India

Corresponding author:
Prakash Krishnaiah, Department of Earth Resources and Environmental Engineering, Hanyang University, 222, Wangsimni-ro, Seongdong-gu, Seoul,
04763, Republic of Korea.
Email: ksp.shine@gmail.com
Krishnaiah
2 et al. 1213
Polymers and Polymer Composites XX(X)

attention by the researchers as a new type of nano-fillers for the generation of polymer composites owing to their good
aspect ratio, easy availability and biocompatibility.11,12 Most commonly HNTs are found with fine tubular structure with
an approximate tube length of 300–1500 nm and an inner and outer diameter around 15–100 nm and 40–120 nm,
respectively.13
Among the wider range of conventional polymer matrices, polypropylene (PP) is the widely used polymer for a range
of applications such as packaging, home appliances, automotive, etc. due to its easy processability, economical, low
density and solvent resistant properties.9,14 However, PP has limitations in specific engineering applications due to its
moderate mechanical properties. Incorporation of nano-fillers such as HNTs into PP could enhance the mechanical and
thermal properties.15,16 However, similar to other nano-fillers, HNTs too have drawbacks of poor dispersion with the
polymer matrix which causes a significant drop in the mechanical properties of the resultant polymer nanocomposites.17
Similar to other nano-fillers, HNTs demonstrate difficulties in the dispersion and found to show significant agglom-
eration while obtaining clay polymer nanocomposites. The main reason beyond agglomeration of HNTs is the restriction
of proper dispersion due to the presence of a large number of hydroxyl groups and hence agglomeration occurs in the non-
polar polymer matrices.18 Surface functionalization or modification of nano-fillers is often the used technique to modify
the surface of nanomaterials to reduce the agglomeration and enhance the dispersion of nanomaterials in the polymer
matrices.19
Sikora et al.20 studied on the improved physical, and thermal properties of HNTs filled polyethylene (PE) matrix
composites. They reported that the ultimate tensile strength increased by 16 MPa. However, they also noticed that strain at
ultimate stress, impact strength and hardness of the examined specimens slightly decreased. This is mainly due to the
agglomeration of nano-fillers within the polymer matrix. Also, the lack of interfacial interaction between nano-fillers and
polymer could be the reason for the reduction of mechanical properties. Surface functionalization of nano-fillers improves
the interfacial adhesion between nano-fillers and polymer matrices which in turn enhances the thermal and mechanical
performance of polymer composites.19,21 Many reports suggest that organosilanes could be more effective for the
inorganic clay nanomaterials such as nano-silicates, HNTs etc.22,23 Albdiry et al.24 studied about the outcome of the
surface-modified HNTs on the impact fracture behavior of HNTs reinforced unsaturated polyester nanocomposites. They
noticed a significant enhancement of impact strength for the surface-modified HNTs nanocomposites as compared to the
unmodified ones. Functionalization of HNTs with silane coupling agent can enhance the diffusion and interfacial bonding
between the polymer matrix and nano-fillers which leads to increased mechanical and thermal characteristics.22,25,26 Ng
and Chow27 investigated the surface-modified HNTs filled PP and PA66 polymer blends. They reported that flexural
strength and modulus improved significantly due to the reinforcing and compatibilization effect of surface-modified
HNTs. Thermal stability of PP/PA66 composites increased significantly with the addition of surface-modified HNTs. This
is mainly due to the heat barrier effects of surface-modified HNTs in the polymer composites. In the present work, we have
functionalized HNTs with APTES coupling agent, which is one of the best silane coupling agents, especially for inorganic
nano-fillers. We have achieved a good mechanical (28% higher tensile strength as compared to pure PP) and thermal
stability (increased by 17 C) with a small quantity (6 wt%) of HNTs with APTES modification. Besides, a significant
increment in the impact strength (40% higher as compared to pure PP) was also noted.
The aim of this study is to examine and compare the effects of functionalization of HNTs with APTES on the surface
morphology, tensile, impact, dynamic-mechanical and thermal performance of PP nanocomposites reinforced with HNTs.
For this, polymer nanocomposites of altered weight percentages of unmodified and APTES modified HNTs reinforced
with PP (HNTs-PP) were prepared, and their morphology, tensile, impact, visco-elastic and thermal performances were
investigated.

Experimental
Materials
Polypropylene (Titanpro 6331), a type of homopolymer with injection molding grade was used. The melt flow index and
the density of PP were 14 kg.m3 and 0.9 kg.m3, respectively. Halloysite nanotubes and (3-Aminopropyl) triethoxysi-
lane (APTES) were obtained from Sigma-Aldrich, Malaysia. Typical HNTs with the diameter, length, pore size and
surface area of 30–70 nm, 1–3 mm, 1.26–1.34 ml.g1 and 64 m2.g1, respectively were used. Calcium chloride and toluene
were obtained from SD Fine Chemicals and were used as obtained without further purification.

Modification of HNTs
Surface functionalization of HNTs was carried out with APTES as reported earlier.23 Calculated quantity of APTES was
dissolved in 25 ml of toluene and was subjected to ultrasonication for 10 min. 0.6 g of halloysites was transferred to
APTES mixer, and the suspension was dispersed ultrasonically (ultrasound probe; Cole Parmer, USA; 20 kHz; 750 W)
using a tapered microtip (6.35 mm diameter) for 30 min. Then the above suspension was subjected to reflux at 120 C for
20 h under constant stirring. For the dry environment, a calcium chloride drying tube was attached to the end. The resultant
1214
Krishnaiah et al. Polymers and Polymer Composites 29(8)
3

solid mixture was filtered and washed with toluene to ensure the removal of excess APTES. Finally, the slurry was then
dried at 120 C for 12 h for further curing.

Preparation of HNTs/PP polymer composites


The preparation method of PP composites was the same as described in our previous work.28 Previously dried and
weighed both the unmodified and surface-modified HNTs with different weight percentage were mixed with PP granules
by using an internal mixer (Brabender Plasticorder PL2000-6 with a volumetric capacity of 69 cm3 with the rotor speed
and time of mixing were 50 rpm and 10 min, respectively) at 180 C. Materials obtained from the internal mixer were
pelletized. After cooling them to room temperature they were molded by using compression machine (LP-S-50 Scientific
Hot and Cold Press).

Characterization
FTIR analysis was used to check the structural modification of HNTs before and after surface modifications. For this, the
pellets of unmodified and surface-modified HNTs were prepared by grounding and pressing well with KBr of definite
proportion. FTIR test was carried out from the wavelength of 400 to 4000 cm1 with 0.85 cm1 resolution by FTIR
Spectrophotometer (FTIR 8300 Shimadzu, Japan). The effect of surface modification on the morphology was studied by
fractured surfaces of PP composites filled with HNTs employing FE-SEM (Quanta 400 FE-SEM). Mechanical properties
such as tensile (Instron 5980) and impact tests (Model CE UM-636) were conducted according to ASTM standards D638
and D256-10, respectively for virgin PP and its composites filled with unmodified and surface-modified HNTs with a
constant crosshead speed of 20 mm.min1 at room temperature. For both tensile and impact properties, a total of six
specimens were tested, and at least five replicate specimens were presented as an average of tested specimens. DMA test
(Perkin-Elmer DMA8000) was carried out and the samples were exposed to a cyclic tensile strain with the force amplitude
of 0.1 N at a frequency of 1.0 Hz. Storage modulus and damping factor (tan delta) were determined from room temperature
to 120 C at a heating rate of 3 C min1. TGA analysis (Perkin-Elmer STA6000 TA instrument) was carried out between
room temperature and 600 C at a heating rate of 10 C.min1 with the constant flow of nitrogen at a flow rate of 20
ml.min1. DSC was carried out for HNTs (before and modification with APTES) filled PP polymer composites using DSC
equipment (Mettler Toledo DSC 1-32) to measure the melting and crystallization temperatures and also the crystallinity.
Investigations were carried out by heating the samples in an enclosed aluminum pan from ambient temperature to 120 C at
the rate of 10 C.min1. The samples were subjected to cooling to room temperature to eliminate the thermal history.
Second heating was carried out with the same procedure to obtain heat flow vs temperature thermograms. All tests were
accomplished using an inert atmosphere (nitrogen flow with 20 ml.min1).

Results and discussion


FTIR analysis
The surface modification of HNTs with APTES was confirmed by FTIR analysis, and the FTIR spectra of unmodified and
APTES modified HNTs are represented in Figure 1. As seen from this Figure 1, unmodified and APTES modified HNTs
show the characteristic broad peaks at around 3620 cm1 which represents the inner O-H stretching vibrations of Al-OH
groups of HNT nanotubes. The peaks at 1652 cm1 exhibit O-H deformation and stretching of water molecules and the
peaks at 1141 cm1 and 1025 cm1 represent Si-O-Si stretching vibration and in-plane stretching of Si-O on the surface of
HNTs.29,30
The FTIR spectra of APTES modified HNTs exhibit new peaks as compared to unmodified HNTs, as shown in Figure 1.
The new peaks at 2987 cm1 and 2940 cm1 exhibit the stretching band of aliphatic C-H groups. The deformation vibration
of NH2 groups shows the peaks at 1641 cm1 and 1556 cm1 which could be noted in the APTES functionalized HNTs,
however these peaks disappeared in the FTIR spectrum of unmodified HNTs. These new peaks of APTES modified HNTs
correlated to the chemical bonds of APTES and confirmed the surface modification of HNTs with APTES.

Morphological studies
The surface morphology of the HNTs and impact fractured samples of PP composites filled with HNTs (with and without
surface functionalized) was investigated. Figures 2A, 2B and 2C demonstrate the morphology of HNTs, PP composites
incorporated by unmodified and surface-modified HNTs, respectively. From Figure 2B, it can be seen that the accumula-
tion and poor distribution of unmodified HNTs throughout the PP matrix. The agglomeration was due to poor interfacial
bonding between unmodified HNTs and polymer matrix, which results in the formation of a stress concentration point.
These stress concentration points lead to lowering the mechanical properties of HNTs filled PP polymer nanocompo-
sites.31 However, in the case of HNT filled polymer nanocomposites filled with surface-modified HNTs (6 wt%), the
Krishnaiah
4 et al. 1215
Polymers and Polymer Composites XX(X)

Tranmittance %
mHNT
2940
2987
1422
1556
1641

uHNT

1652
1141 3620
1025

500 1000 1500 2000 2500 3000 3500 4000


-1
Wavenumber cm

Figure 1. FTIR spectra of unmodified (uHNT) and APTES modified HNTs (mHNT).

Figure 2. FE-SEM images of (A) pure HNTs (B) fractured samples from the impact test of 6 wt% of unmodified and (C) surface-modified
HNTs filled PP polymer composites.

surface-modified HNTs were well distributed throughout the polymer. Comparatively, fewer agglomeration might be seen
than the unmodified HNTs/PP composites (Figure 2C).
From Figure 2C, it is clear that after surface modification of HNTs, interfacial adhesion between HNTs and PP polymer
significantly improved which leads to a drastic increase in the mechanical and thermal properties of the resultant clay
polymer nanocomposites. Furthermore, the modification of HNTs with APTES considerably reduces the stress concen-
tration points due to well distribution of HNTs throughout the PP polymer which results in the improvement of the
mechanical performance of polymer composites.22,27
1216
Krishnaiah et al. Polymers and Polymer Composites 29(8)
5

40

35

30

Stress (MPa)
25

20

15
Pure PP
10 6uHNT
6mHNT
5 8uHNT
8mHNT
0
0 4 8 12 16 20 24 28
Strain (%)

Figure 3. Stress–strain graph of unmodified and surface-modified HNTs-PP polymer composites.

Table 1. Mechanical properties of HNTs (unmodified and APTES modified) filled PP nanocomposites.

Sl. No. Samples Tensile strength (MPa) Elongation at break (%) Modulus (MPa) Izod Impact strength (J m1)

1 Pure PP 29.2 + 1.2 10.9 + 1.3 1280 + 22 20 + 1.4


2 PP2uHNT 30.6 + 1.0 12.4 + 0.8 1360 + 23 22.2 + 1.8
3 PP2mHNT 31.7 + 0.6 14.5 + 0.8 1420 + 32 23.5 + 1.5
4 PP4uHNT 31.5 + 0.7 16.2 + 0.5 1450 + 24 25 + 1.4
5 PP4mHNT 33.3 + 1.0 16.9 + 0.8 1510 + 21 27.5 + 1.5
6 PP6uHNT 35.2 + 0.8 18.5 + 0.4 1560 + 22 29.2 + 1.3
7 PP6mHNT 37.9 + 1.1 23.2 + 0.2 1620 + 23 31.8 + 1.3
8 PP8uHNT 35.0 + 1.2 26.7 + 0.3 1740 + 24 29.0 + 3.5
9 PP8mHNT 36.1 + 1.0 21.5 + 0.5 1840 + 15 30.5 + 1.8

Effect of APTES functionalized HNTs on the mechanical properties


Figure 3 shows the results obtained from stress–strain curves of PP and samples containing different weight percentage of
unmodified and APTES functionalized HNTs. It could be seen from Figure 3 that a substantial enhancement in the tensile
strength could be noticed by the incorporation of HNTs regardless of their surface modification as compared to virgin PP
polymer. Compared to virgin PP polymer, the incorporation of 6 wt% of unmodified and surface-modified HNTs into PP
polymer drastically enhanced the tensile strength by 20% and 28% respectively. It is observed that the surface-modified
HNTs with APTES modified HNT doped polymer composites exhibited good tensile properties which agree well with the
results of FTIR and SEM. This improved tensile properties due to the efficient load transfer from the polymer matrix to
filler proves the effective distribution of HNTs across the polymer matrix. In addition, regarding FTIR results, increasing
hydrogen bond interaction between surface-modified HNTs and the PP polymer enhances the compatibility between them
and leads to better tensile strength.
Further, the surface modification of the high aspect ratio of HNTs stimulates filler-polymer interface, further increasing
the mechanical performance of HNTs-PP polymer composites. These findings are well relate to the literature reports.13,19
Guo et al.19 reported that a good enhancement of tensile properties for 6 wt% surface-modified HNTs filled with
Polyamide 6 (PA6) composites than pure PA6 resin.
However, PP polymer composites by the addition of 8 wt% of HNTs showed an insignificant weakening of tensile
strength because of the increased filler concentration above the optimal content (above 6 wt%). Also, due to the weak
interfacial adhesion of filler as well as polymer also increased the stress concentration points. Similar to tensile strength,
modulus also meaningfully improved by the addition of HNTs into the polymer. This trend again agrees well with the
literature reports. Albdiry and Yousif27 observed that the tensile properties were likely to decline due to further addition of
HNTs above the optimum content (3 wt%) into the polymer. However, tensile modulus enhanced by 45% with the
incorporation of higher weight percentage (above 8 wt%) of surface-modified HNTs-PP polymer composites than the
virgin PP. This conclusion again claimed a good distribution of surface-modified HNTs with APTES within the PP matrix.
Steady improvement of impact strength might be witnessed with an increase in HNTs into PP polymer.
The incorporation of a small quantity of highly dispersed and surface-modified HNT in PP composites behaves like
plasticizer which helps dissipating the impact energy across the composites. However, a reduction of toughness might be
detected for the polymer composites with a further addition of HNTs above an optimal range (6 wt%). Other investigators
also reported similar outcomes.13,17 The impact strength of HNTs (unmodified and APTES modified) filled PP composites
were also studied and are shown in Table 1. It can be seen that the impact strength increased with an increase in the
addition of HNTs into PP composites.
Krishnaiah
6 et al. 1217
Polymers and Polymer Composites XX(X)

9 Pure PP
5x10
PP6uHNT
PP6mHNT

Storage modulus (MPa)


9 PP8uHNT
4x10
PP8mHNT

9
3x10

9
2x10

9
1x10

0
-40 -20 0 20 40 60 80 100

Temperature (oC)

Figure 4. Storage modulus (E’) of with and without the surface-modified HNTs/PP polymer composites.

8
2.4x10
Pure PP
2.2x10
8
PP6uHal
8 PP6sHal
2.0x10
PP8uHal
Loss modulus (MPa)

1.8x10
8
PP8sHal
8
1.6x10
8
1.4x10
8
1.2x10
8
1.0x10
7
8.0x10
7
6.0x10

-40 -20 0 20 40 60 80 100

Temperature (oC)

Figure 5. Loss modulus of with and without surface-modified HNTs-PP polymer composites.

The maximum impact strength obtained was 31.8 J.m1 for the surface-modified HNT filled PP composites. It was also
noticed that the impact strength was much higher for the PP composites with the incorporation of surface-modified HNTs
as compared to unmodified and virgin PP. The obtained results correlate with the results reported in the literature. Vahedi
and Pasbakhsh32 studied the impact and fracture properties of surface-modified HNTs and epoxy nanocomposites. It was
reported that the impact strength of epoxy nanocomposites significantly increased with the addition of surface-modified
HNTs into the polymer matrix due to the increased dispersion of nanotubes across the polymer matrix and better
interaction between the HNTs and polymer.

DMA analysis
DMA was conducted to examine the outcome of the surface-modified HNTs doped polymer composites on the visco-
elastic performance. This study is beneficial for the assessment of visco-elastic performance of the nanocomposite with
varying temperature and mechanical stress.33 Figure 4 displays the curves of storage modulus (E’) of HNTs filled PP
nanocomposites. It could be seen from Figure 4 that the storage modulus increased at a lesser temperature of 50 C and
after that a steady reduction with a gradual enhancement of temperature was observed. At low temperature (51 C), the
storage modulus (E’) was greater for the unmodified HNTs doped polymer composites than the virgin PP, and it improved
further with the incorporation of surface-modified HNTs.
It has been noticed that the loading of 6 wt% of the surface-modified HNTs into PP polymer enhanced the storage
modulus to 5.55 GPa, which is 28% greater than the virgin PP. A higher storage modulus for the surface-modified HNTs
incorporated polymer composites than the virgin PP displays a good distribution and interfacial bonding between the
surface-modified HNTs and PP polymer, which correlate with the observations of SEM and FTIR. Other researchers also
reported similar outcomes.11
Lecouvet et al.34 reported on the flammability and thermal performance of HNTs filled polyethersulfone (PES)
composites which were synthesized by melt processing. They found that the storage modulus drastically improved by
40% with the addition of 16 wt% HNTs into polymer than the virgin PES polymer. This is due to the greater intrinsic
toughness of HNTs.
Figure 5 illustrates the loss modulus (E”) of the PP composites with the incorporation of different weight percentage of
unmodified and the surface-modified HNTs.
1218
Krishnaiah et al. Polymers and Polymer Composites 29(8)
7

0.10
Pure PP
PP6uHal
PP6sHal
0.08 PP8uHal
PP8sHal

Tan delta
0.06

0.04

0.02

-40 -20 0 20 40 60 80 100

Temperature (oC)

Figure 6. Tan d of with and without surface-modified HNTs-PP polymer composites.

100
90 Pure PP
6uHNT
80
6mHNT
Relative mass (wt%)

70 8uHNT
8mHNT
60
50
40
30
20
10
0
100 200 300 400 500 600
o
Temperature ( C)

Figure 7. TGA curves of virgin PP, with and without surface-modified HNTs-PP composites.

Table 2. TGA and DSC values of crystallinity (Xc), melting (Tm) and cold crystallization temperatures (Tc),

Samples Tm (� C) Tcc (� C) Xc (%) T 50% loss (� C)

Pure PP 166.65 123.52 38.5 442.13


PP þ 6uHNT 168.57 122.8 42.3 437.29
PP þ 6mHNT 168.8 121.56 46.8 445.24
PP þ 8uHNT 169.21 121.5 44.71 427.79
PP þ 8mHNT 169.83 121.02 47.1 432.66

A wider loss modulus peak has been noted with the peak height was comparatively higher for virgin PP polymer than
the HNTs loaded polymer composites. The loss modulus curves of the surface-modified HNT incorporated PP composites
showed lesser peak height which confirms the better modulus and declined loss factor. This is due to good distribution of
HNTs throughout the PP matrix as well as good interfacial bonding between polymer and filler.35 It could be observed
from Figure 6 that there is a substantial raise in the Tg with the addition of surface-modified HNTs doped polymer
composites than pure PP. The Tg for pure PP was at 6.8� C, whereas Tg with the incorporation of surface-modified HNTs of
6 wt% and 8 wt% were at 7.2� C and 8.3� C, respectively. This could be due to the constraint of segmental movement in the
polymer chain after the incorporation of HNTs, which improved the modulus and leads to larger Tg of the polymer
composite.36

TGA analysis
The thermal stability of polymer composites filled with and without surface-modified HNTs was examined using TGA.
TGA graph of virgin PP and its polymer composites incorporated with and without surface-modified HNTs have been
shown in Figure 7. A small rise in the thermal steadiness could be observed with the incorporation of HNTs into PP
polymer. Table 2 shows the thermal stability of the samples with different wt% loading of with and without surface-
modified HNTs and their proportions of char remaining at the extreme temperature (600� C). It is good to indicate that the
thermal stability is significantly enhanced by 18� C with the incorporation of surface-modified HNTs into PP polymer
composites than virgin PP. The improved thermal stability displays good interfacial bonding between the surface-
modified HNTs and PP polymer. However, a substantial reduction in the thermal stability for unmodified HNTs filled
PP polymer composites at the maximum weight loss point than virgin PP could be noted. Similar outcomes have been
Krishnaiah
8 et al. 1219
Polymers and Polymer Composites XX(X)

PP8mHNT
PP8uHNT
PP6mHNT
PP6uHNT
Pure PP

Heat flow (a. u.)


40 60 80 100 120 140 160 180 200
Temperature (° C)

Figure 8. DSC melting thermograms of virgin PP and its polymer composites with the incorporation of HNTs with and without surface
functionalization.

PP8mHNT
PP8uHNT
PP6mHNT
PP6uHNT
Pure PP
Heat flow (a. u.)

40 60 80 100 120 140 160 180


Temperature (°C)

Figure 9. DSC cooling curves of virgin PP, with and with surface-functionalized HNTs-PP polymer composites.

described in the literature.37 It is also noticed that a considerable reduction of thermal stability for the samples of PP
composites filled HNTs (higher than 6 wt%, an optimum quantity). These results are in agreement with the mechanical
properties observed in this study and also other reports in the literature.11 The reason for the reduction of thermal stability
is due to the increased number of agglomeration with a further addition of HNT fillers.
It is significant to note that the charred remains of HNTs doped polymer composites considerably greater than virgin
PP. This outcome illustrates the amount of HNTs incorporated into polymer composites and hence an enhanced thermal
stability of the HNTs-PP polymer composites.

DSC analysis
DSC was conducted to explore the effect of incorporation of with and without surface-modified HNTs into PP polymer
composites. Figures 8 and 9 display the melting as well as crystallinity thermograms of HNTs-PP polymer composites.
Table 2 displays the crystallinity (Xc), melting (Tm), and cold crystallization temperatures (Tc) for the HNTs (with
and without surface modified) incorporated PP composites. The melting temperature (Tm) peak displays a substantial
rise in the melting temperature (Figure 8). The peaks of melting temperature moved to 169.83 C with the incorporation
of 8 wt% surface-functionalized HNTs, while the peak of melting temperature for virgin PP was at 166.65 C. An
increase in the melting temperature could be due to the strengthening effect, which limits the unrestricted movement of
polymer chain fragments in the HNTs incorporated polymer composites.15 It is described that the incorporation of
HNTs prompted the nucleating effects in the composites which leads to the improved crystallinity of HNTs doped
polymer composites.13,37
The cold crystalline temperature (Tcc) of HNTs-PP polymer composites moved to lesser temperature after the incor-
poration of HNTs. The Tcc for the virgin PP was at 123.52 C, and with 8 wt% of HNTs loading into PP matrix, Tcc moved
to 121.02 C. The lesser Tcc as compared to virgin PP specifies the incorporation of HNTs supporting the kinetics of
nucleating effect in the HNTs-PP polymer composites and thus the better crystallinity.38–40
1220
Krishnaiah et al. Polymers and Polymer Composites 29(8)
9

Conclusions
In this investigation, the influence of the surface-modified HNTs on the mechanical, dynamic-mechanical and thermal
performance of HNTs-PP polymer composites has been examined. SEM results show that the APTES modified HNTs
were well isolated within the PP polymer than the unmodified ones, due to reduced agglomeration. The tensile strength,
modulus and impact properties were improved by 28%, 45% and 60% respectively with the incorporation of 6 wt% of the
surface-modified HNTs doped polymer nanocomposites than the virgin PP polymer. The thermal investigation displays
that the thermal stability and crystallinity improved by 18 C and 22% respectively for the HNTs-PP polymer composites
with the surface modification by APTES than virgin PP. DMA analysis exhibited an enhanced dynamic modulus by 28%
than virgin PP. The surface-modified HNTs reinforced PP polymer composites find wider applications not only in
packaging but also in automotive, structural and home appliances.

Declaration of conflicting interests


The author(s) declared no potential conflicts of interest with respect to the research, authorship, and/or publication of this article.

Funding
The author(s) received no financial support for the research, authorship, and/or publication of this article.

ORCID iD
Prakash Krishnaiah https://orcid.org/0000-0002-1567-7618

References
1. Ferreira JAM, Reis PNB, Costa JDM, et al. A study of the mechanical properties on polypropylene enhanced by surface treated
nanoclays. Compos Part B Eng 2011; 42: 1366–1372.
2. Venkatesh GS, Deb A, Karmarkar A, et al. Effect of nanoclay content and compatibilizer on viscoelastic properties of montmor-
illonite/polypropylene nanocomposites. Mater Des 2012; 37: 285–291.
3. Awad WH, Beyer G, Benderly D, et al. Material properties of nanoclay PVC composites. Polymer (Guildf) 2009; 50: 1857–1867.
4. Cheng S, Cairncross RA, Hsuan YG, et al. Clay orientation effect on the thermal stability of polyethylene-nanoclay nanocompo-
sites. Polymer (United Kingdom) 2013; 54: 5016–5023.
5. Matadi Boumbimba R, Bouquey M, Muller R, et al. Dispersion and morphology of polypropylene nanocomposites: characteriza-
tion based on a compact and flexible optical sensor. Polym Test 2012; 31: 800–809.
6. Chen Y, Guo Z and Fang Z. Relationship between the distribution of organo-montmorillonite and the flammability of flame
retardant polypropylene. Polym Eng Sci 2012; 47: 21–25.
7. Allahverdi A, Ehsani M, Janpour H, et al. The effect of nanosilica on mechanical, thermal and morphological properties of epoxy
coating. Prog Org Coat 2012; 75: 543–548.
8. Farzadnia N, Abang Ali AA, Demirboga R, et al. Effect of halloysite nanoclay on mechanical properties, thermal behavior and
microstructure of cement mortars. Cem Concr Res 2013; 48: 97–104.
9. Liu M, Guo B, Du M, et al. Halloysite nanotubes as a novel b-nucleating agent for isotactic polypropylene. Polymer (Guildf) 2009;
50: 3022–3030.
10. Yuan P, Tan D and Annabi-Bergaya F. Properties and applications of halloysite nanotubes: recent research advances and future
prospects. Appl Clay Sci 2015; 112–113: 75–93.
11. Prashantha K, Lacrampe MF and Krawczak P. Processing and characterization of halloysite nanotubes filled polypropylene
nanocomposites based on a masterbatch route: effect of halloysites treatment on structural and mechanical properties. Express
Polym Lett 2011; 5: 295–307.
12. Voon HC, Bhat R, Easa AM, et al. Effect of addition of halloysite nanoclay and SiO2 nanoparticles on barrier and mechanical
properties of bovine gelatin films. Food Bioprocess Technol 2012; 5: 1766–1774.
13. Prashantha K, Schmitt H, Lacrampe MF, et al. Mechanical behaviour and essential work of fracture of halloysite nanotubes filled
polyamide 6 nanocomposites. Compos Sci Technol 2011; 71: 1859–1866.
14. Liu M, Jia Z, Liu F, et al. Tailoring the wettability of polypropylene surfaces with halloysite nanotubes. J Colloid Interface Sci
2010; 350: 186–193.
15. Wang B and Huang H-X. Effects of halloysite nanotube orientation on crystallization and thermal stability of polypropylene
nanocomposites. Polym Degrad Stab 2013; 98: 1601–1608.
16. Liu X and Wu Q. PP/clay nanocomposites prepared by grafting-melt intercalation. Polymer (Guildf) 2001; 42: 10013–10019.
17. Liu M, Zhang Y and Zhou C. Nanocomposites of halloysite and polylactide. Appl Clay Sci 2013; 75–76: 52–59.
18. Tzounis L, Herlekar S, Tzounis A, et al. Halloysite nanotubes noncovalently functionalized with SDS anionic surfactant and PS-b-
P4VP block copolymer for their effective dispersion in polystyrene as UV-blocking nanocomposite films. J Nanomater 2017;
2017: 1–11. DOI: 10.1155/2017/3852310.
Krishnaiah
10 et al. 1221
Polymers and Polymer Composites XX(X)

19. Guo B, Zou Q, Lei Y, et al. Structure and performance of polyamide 6/halloysite nanotubes nanocomposites. Polym J 2009; 41:
835–842.
20. Sikora JW, Gajdoš I and Puszka A. Polyethylene-matrix composites with halloysite nanotubes with enhanced physical/thermal
properties. Polymers 2019; 11: 1–11.
21. He Y, Kong W, Wang W, et al. Modified natural halloysite/potato starch composite films. Carbohydr Polym 2012; 87: 2706–2711.
22. Carli LN, Daitx TS, Soares G V, et al. The effects of silane coupling agents on the properties of PHBV/halloysite nanocomposites.
Appl Clay Sci 2014; 87: 311–319.
23. Yuan P, Southon PD, Liu Z, et al. Functionalization of halloysite clay nanotubes by grafting with 3-aminopropyltriethoxysilane. J
Phys Chem C 2008; 112: 15742–15751.
24. Albdiry MT, Ku H and Yousif BF. Impact fracture behaviour of silane-treated halloysite nanotubes-reinforced unsaturated
polyester. Eng Fail Anal 2013; 35: 718–725.
25. Pasbakhsh P, Ismail H, Fauzi MNA, et al. EPDM/modified halloysite nanocomposites. Appl Clay Sci 2010; 48: 405–413.
26. Albdiry MT and Yousif BF. Role of silanized halloysite nanotubes on structural, mechanical properties and fracture toughness of
thermoset nanocomposites. Mater Des 2014; 57: 279–288.
27. Ng CL and Chow WS. Multifunctional halloysite nanotube–reinforced polypropylene/polyamide binary nanocomposites. Polym
Polym Compos 2019; 28: 1–8.
28. Krishnaiah P, Thevy C and Manickam S. Development of silane grafted halloysite nanotube reinforced polylactide nanocomposites
for the enhancement of mechanical, thermal and dynamic-mechanical properties. Appl Clay Sci 2017; 135: 583–595.
29. Wang LF and Rhim JW. Functionalization of halloysite nanotubes for the preparation of carboxymethyl cellulose-based nano-
composite films. Appl Clay Sci 2017; 150: 138–146.
30. Ibrahim GPS, Isloor AM, Moslehyani A, et al. Bio-inspired, fouling resistant, tannic acid functionalized halloysite nanotube
reinforced polysulfone loose nanofiltration hollow fiber membranes for efficient dye and salt separation. J Water Process Eng
2017; 20: 138–148.
31. Krishnaiah P. Development of polylactide and polypropylene composites reinforced with sisal fibres and halloysite nanotubes for
automotive and structural engineering applications. PhD Thesis, University of Nottingham. 2017. http://eprints.nottingham.ac.uk/
43498/ (accessed 3 April 2018).
32. Vahedi V and Pasbakhsh P. Instrumented impact properties and fracture behaviour of epoxy/modified halloysite nanocomposites.
Polym Test 2014; 39: 101–114.
33. Saba N, Jawaid M, Alothman OY, et al. A review on dynamic mechanical properties of natural fibre reinforced polymer compo-
sites. Constr Build Mater 2016; 106: 149–159.
34. Lecouvet B, Sclavons M, Bailly C, et al. A comprehensive study of the synergistic flame retardant mechanisms of halloysite in
intumescent polypropylene. Polym Degrad Stab 2013; 98: 2268–2281.
35. Khonakdar HA. Dynamic mechanical analysis and thermal properties of LLDPE/EVA/modified silica nanocomposites. Compos
Part B Eng 2015; 76: 343–353.
36. Russo P, Acierno D, Rosa R, et al. Mechanical and dynamic-mechanical behavior and morphology of polystyrene/perovskite
composites: Effects of filler size. Surf Coat Technol 2014; 243: 65–70.
37. Wang B, Huang H and Wang Z. Composites: Part B In situ fibrillation of polymeric nucleating agents in polypropylene and
subsequent transcrystallization propelled by high-pressure water penetration during water-assisted injection molding. Compos Part
B 2013; 51: 215–223.
38. Evagelia K, Michael N, Panayiotis G, et al. Comparative study of PLA nanocomposites reinforced with clay and silica nanofillers
and their mixtures. J Appl Polym Sci 2011; 122: 1519–1529.
39. Wang H and Qiu Z. Crystallization behaviors of biodegradable poly(l-lactic acid)/graphene oxide nanocomposites from the
amorphous state. Thermochim Acta 2011; 526: 229–236.
40. Lecouvet B, Sclavons M, Bourbigot S, et al. Thermal and flammability properties of polyethersulfone/halloysite nanocomposites
prepared by melt compounding. Polym Degrad Stab 2013; 98: 1993–2004.

You might also like