You are on page 1of 11

Composites: Part A 40 (2009) 11991209

Contents lists available at ScienceDirect

Composites: Part A
journal homepage: www.elsevier.com/locate/compositesa

Optimization of the mechanical properties of polypropylene-based


nanocomposite via the addition of a combination of organoclays
K.S. Santos a, S.A. Liberman b, M.A.S. Oviedo b, R.S. Mauler a,*
a
b

Instituto de Qumica, Universidade Federal do Rio Grande do Sul, Av Bento Gonalves 9500, Porto Alegre-RS 91501-970, Brazil
Braskem S/A, III Plo Petroqumico, Via Oeste Lote 5, Passo Raso, Triunfo, Brazil

a r t i c l e

i n f o

Article history:
Received 17 November 2008
Received in revised form 5 May 2009
Accepted 9 May 2009

Keywords:
A. Thermoplastic resin
A. Nano-structures
B. Mechanical properties

a b s t r a c t
The properties of polypropylene (PP) nanocomposites are dependent on the quaternary ammonium salt
in the montmorillonite (MMT). A nanocomposite with C-15A, which has a high cation exchange capacity
(CEC), exhibits an increase in its impact properties, while one prepared with C-20A, which has a low CEC,
shows an increase in the exural modulus. In order to obtain enhancements in both properties, PP nanocomposites were prepared using a combination of 1:1 of C-15A/C-20A. X-ray, TEM, thermal properties,
dynamical mechanical analysis (DMA), and mechanical tests were used to evaluate the properties of this
novel mixture. Nanocomposites of partially exfoliated morphology were obtained, especially when 5 wt%
of poly(propylene-graft-maleic anhydride) (PP-g-MA) was used. The mechanical tests showed that the
use of a 1:1 mixture of C-15A/C-20A caused a simultaneous gain of approximately 12% in exural modulus and a ve times higher impact strength. In addition, the dispersion of the clay was more homogeneous, with the absence of agglomerated structures that were present when either the individual
C-15A or C-20A was used. The DMA results showed that while the organoclay improved the modulus
of PP, the Tg was decreased slightly.
2009 Elsevier Ltd. All rights reserved.

1. Introduction
Polymer-layered silicate nanocomposites, a new class of materials, often exhibit a remarkable improvement in both their physical and mechanical properties when compared to pristine polymer
or conventional composites due to the presence of nanometerscale dispersion in the former [14]. Polypropylene (PP) is one of
the most widely used polyolens in large-scale production, while
montmorillonite (MMT) is an environmentally friendly, naturally
abundant and cheap clay. Consequently, PP/clay nanocomposites,
which combine the advantages of both types of materials, have
found widespread use in general household goods, various electronic product, and automobile parts [5,6].
Layered silicates, such as MMT, have a layer thickness on the order of 1 nm and a high aspect ratio. Hence, only a small amount of
layered silicates is necessary to provide a much higher surface area
for facilitating polymernanoller interactions [7]. There are generally three methods to obtain dispersed polymerclay nanocomposites: intercalation of a suitable monomer followed by
polymerization, polymer intercalation from solution, and direct
polymer melt intercalation in which the polymer chains are diffused into the space between the silicate layers [8,9].

* Corresponding author. Tel.: +55 51 33086296; fax: +55 51 33086304.


E-mail address: raquel.mauler@ufrgs.br (R.S. Mauler).
1359-835X/$ - see front matter 2009 Elsevier Ltd. All rights reserved.
doi:10.1016/j.compositesa.2009.05.009

The entropic and enthalpic balances determine the dispersion of


the layered silicates in the polymer matrix. As the polymer chains
become constrained between clay sheets, the entropy decreases. In
order to obtain an effective dispersion, the enthalpic balanced, that
is interaction, between the clay and the polymer must be favorable
[4,10,11]. Therefore, the use of polar groups is necessary to achieve
effective nanoscale dispersion in intercalated or exfoliated structure of silicates sheets in a non-polar polymer matrix, such as
maleic anhydride or silanes [4,5]. In addition, the clay must be
organically modied in order to reduce the interactions between
its sheets [12]. This organic modication causes the clay surface
to be more hydrophobic, and expands the spaces between the silicate layers [13]. The most common method to modify the clay is
through ionic exchange of the cations between the clay platelets
using ammonium cations with long alkyl tails, such as the quaternary ammonium salts [14]. Furthermore, the use of a processing
aid of higher molecular weight can facilitate the dispersion of clay
sheets in the matrix, which decreases the number of agglomerated
structures, hence improving the overall properties of the nanocomposites [15].
Ammonium cations make the clay more organophilic. Besides
stabilizing the exfoliated sheets, they also prevent the interaction
of polymer chains or the compatibilizer agents with the polar
surfaces of the clay [11]. The strength of these interactions is
determined by the size and quantity of the ammonium. The greater
the size and quantity of the surfactant, the stronger are the

1200

K.S. Santos et al. / Composites: Part A 40 (2009) 11991209

interactions between the polymer/surfactant and the clay/surfactant components, leading consequently to a weaker polymer/clay
interaction [13,1517]. Besides, the organic modier has low
molecular weight and the rigidity of the matrix is reduced due
its smaller size and higher amount. This decrease in rigidity enhances the impact strength of the composite, but decreases its
capacity for reinforcement [15]. Previous studies showed that the
use of Cloisite 15A improves the impact strength while Cloisite
20A enhances the exural modulus [15]. It is thus believed that
the combined use of two clays, one with a higher and the other
with a lower amount of organic modier, will promote a simultaneous gain in both the reinforcement capacity and the impact
strength.
The objective of this work is to evaluate a 1:1 combination of
two commercial modied clays (Cloisite 20A and 15A) and the effect of varying the amount of quaternary ammonium salt on the
morphology and thermal and mechanical properties of the PP
nanocomposites. The use of a compatibilizer agent will also be
evaluated.
2. Experimental
2.1. Materials
Polypropylene homopolymer with a melt ow index (MFI
(230 C/2.16 kg)) of 3.5 g/10 min was obtained from Braskem S.A.
Poly(propylene-graft-maleic anhydride) (PP-g-MA) with 0.2 wt%
of reactive MA modier and MFI (230 C/2.16 kg) of 7 g/10 min
was supplied by Crompton. Irganox B215 an antioxidant from
Ciba was added to PP. The solvent methylethylketone (MEK) with
density of 0.95 g/cm3, boiling point of 72.11 C from Nuclear,
Poly(propylene glycol) PPG, with Mn = 1000 g/mol, viscosity of
190 cP, density of 1.005 g/cm3 from Aldrich were used to prepare
a stable suspension with the organophilic montmorillonites: Cloisite 15A (CEC: 125 meq/100 g) and 20A (CEC: 98 meq/100 g) from
Southern Clay Products.
2.2. Melt processing
PP nanocomposites with 2 wt% and 5 wt% of a combination of
1:1 C-15A/C-20A were obtained using a twin screw co-rotating
extruder (Haake H-25, model Rheomex PTW 16/25, L/D = 25) operating at 80 rpm. A stable suspension of MMT (combination of 1:1
C-15A/C-20A) and 1 wt% PPG in MEK (300 mL) was added into
the second feed point, and residual solvent was removed in the degassed zone as described elsewhere [18]. A temperature range between 170 and 190 C was chosen for the preparation and
processing of the nanocomposites in order to minimize possible
degradation of the organic modier and the matrix. The nanocomposites were injection molded for ASTM D-638 type 1 using the
Battenfeld Plus 350/075 injection-molding machine. The temperature of the cylinders was kept between 220 and 230 C, and the
mold was maintained at 60 C. Films with 47 mm of thickness were
obtained by compression heating the polymer up to 190 C, and the
temperature was then maintained for 2 min in order to obtain a
complete melting of the pellets before a pressure of 6 lbs was
applied for 3 min. The samples were then cooled to room temperature at a cooling rate of 20 C/min to be used in measures of
X-ray, TGA and DSC analyses.
2.3. Characterization
The WAXD measurements were performed using a Siemens D500 diffractometer. Films (PP nanocomposites) and powder (organoclays) samples were scanned in the reection mode using an

incident X-ray of Cu Ka with wavelength of 1.54 , at a step width


of 0.05/min from 2h = 1 to 10. The dispersion of the layers in the
nanocomposites, as well as the basal spacing of the clays, was estimated from the (0 0 1) diffraction. The morphologies of the specimens were examined by TEM (JEOL JEM-120 Ex II), which operated
at an accelerating voltage of 80 kV. Ultra thin specimens (70 nm)
were cut from the middle section of injection-molded specimens
in a direction perpendicular to the ow of the melt during the
injection process. Cutting operations were carried out under cryogenic conditions with a Leica Ultracut UCT microtome equipped
with a glass or diamond knife at 80 C, and the lm was retrieved
onto 300 mesh Cu grids. The TEM images, which were typically
magnied 30,000 times, were quantied using Adobe Photoshop,
where the dispersed platelets and/or agglomerate structures captured through TEM were traced over onto an overlapped blank
layer. To ensure the accurate measurements of the particle length
and thickness, the images were sufciently magnied so that a
majority of the particles, including single platelets, were counted
by the Image Tools program.
Thermal properties were determined using a DSC Thermal Analyst 2100/TA Instruments. All measurements were carried out under a nitrogen atmosphere. The samples were heated from 50 to
200 C at a heating and cooling rate of 10 C/min. The measurements were made in the second heating and cooling cycle. The degree of crystallinity was determined using DH0m 190 J/g for PP
[19]. The non-isothermal crystallization kinetics (t1/2 values) were
calculated from the integral of the DSC curves. The DSC instrument
was calibrated with indium before use.
Thermogravimetric analyses were carried out using a TA model
QA-50 instrument to obtain the inorganic and organic residues, together with the decomposition proles of both the clays, and the
PP nanocomposites. The samples (10.0 0.3 mg in lm form) were
heated from 80 C to 900 C at a rate of 20 C/min under nitrogen
ow.
A TA model QA 800 instrument was used for dynamic mechanical analyses (DMA) and the measurement of heat deection temperature (HDT) behavior of the materials at a xed frequency of
1 Hz. DMA analyses were performed in the single cantilever mode,
and the injection-molded samples were heated from 30 C to
130 C at a rate of 3 C/min. The conditions for the HDT analyses
were adapted from the ASTM D-648 standard, and carried out in
the three point bending mode, using specimens with approached
dimensions of 50.00  3.2  12.77 mm, and a force of 0.79 N. The
temperature range was from 30 C to 120 C using a heating rate
of 2 C/min. The HDT was recorded when the sample had suffered
a deection of 0.25 mm.
The exural modulus was measured at room temperature using
an Instron 4466 testing machine according to the ASTM D-790
standard, at a crosshead speed of 13 mm/min. The notched Izod
impact strength was measured using a pendulum-type Ceast
6545 at 23 C with an impact speed of 3.46 m/s according to the
ASTM D-256 standard. The reported values were averaged over
10 measurements.
3. Results and discussion
3.1. Clay characterization
Fig. 1 and Table 1 show the characteristic diffraction peak corresponding to the 001 plane, and the interlayer distances of the
platelets. The peak at 6.9 suggests that a small percentage of the
Na+-MMT was not modied by the quaternary ammonium salt
[2,20]. Both the C-15A and C-20A clays were modied with a quaternary ammonium salt derived from hydrogenated tallow with
two long alkyl groups having a distribution of 65% C18, 30% C16,

1201

K.S. Santos et al. / Composites: Part A 40 (2009) 11991209

ganic modier present, the narrower the distribution of weight


loss.

10000

2.6

3.3

Intensity

8000

C-15A
C-20A

3.2. Morphology of the PP/C-15AC-20A nanocomposites

6000

4000

2000

6.9
0

10

2
Fig. 1. WAXD patterns for the organophilic clays. The arrows indicate the values of
the silicate reection of the organophilic clays.

Table 1
Interlayer distance galleries and residues of the organophilic clays.
Clay

d(1 peak)
(nm)

d(2 peak)
(nm)

Organic
residuea (wt%)

Inorganic
residue (wt%)

C-15A
C-20A

3.39
2.68

1.3
1.3

42.8
37.2

57.2
62.8

a
% Calculated by TGA regarding of thermal decomposition range of the ammonium salt (200500 C) Ref. [21].

All PP nanocomposites exhibited displacement to lower angles


(Figs. 3 and 4 and Table 2), indicating the intercalation of the
PPG between clay sheets that was also reected by a d002 peak that
was assigned as a second order peak [4,15,22]. The PPG might penetrate into the silicate sheets due to its higher polar character and
lower molecular weight, promoting the sliding of silicate sheets
over another, which facilitated the diffusion of the PP and/or PPg-MA chains between the platelets. The combination of the diffusion of polymer chains into the clay and the shear on the platelet
surface carry on a peel in individual platelet, resulting in an exfoliated structure [4,15]. However, the intercalation of the PPG between the clay sheets reached a maximum of expansion at about
3.68 nm, independent of the clay used or the presence of PP-gMA. Beyond that interlamelar distance, the clay sheets lost their
ordered arrangement, and the d001 peak disappeared from the Xray spectrum. Since the d001 value of C-20A was lower than that
of C-15A, there was higher intercalation of PPG until the platelets
reached their maximum separation distance before their interactions ceased.

(a)

C-15A POWDER
2% C-15A+1% PPG
2% C-15A+5% PP-g-MA+1% PPG
5% C-15A+1% PPG

35000

30000

25000

INTENSITY (Cps)

and 5% C14. The amount of ammonium salt in the C-15A is higher


than that in C-20A (around 5.6 wt%), thereby increasing the interlayer space in C-15A from 2.68 to 3.39 nm.
Fig. 2 shows the DTG scan of pristine organoclays, which display
four distinct regions: the elimination of the free and physisorbed
water in the internal layer of the clay, and the dis-aggregation or
melting of the long chain tails of the surfactant molecules occurred
at temperatures up to 200 C; surfactant decomposition occurred
between 200 and 500 C; dehydroxylation of the clay occurred between 500 and 700 C; and the decomposition of carbonaceous
matter with the release of CO2 occurred between 700 and
1000 C [21]. It was observed that the lower the amount of the or-

20000

15000

10000

5000

0
1

(b) 35000
C-15A
C-20A

C-20A POWDER
2% C-20A+1% PPG
2% C-20A+5% PP-g-MA+1% PPG
5% C-20A+1% PPG

30000

0.20

25000

INTENSITY (Cps)

DERIV. WEIGHT (%/ C)

0.25

0.15

0.10

20000

15000

10000

0.05

5000
0.00

0
0

100

200

300

400

500

600

TEMPERATURE

700

800

( C)

Fig. 2. DTG of the organophilic clays: C-15A and C-20A.

900

1000

2
Fig. 3. WAXD patterns of the PP/C-15A (a) and PP/C-20A; (b) nanocomposites.

1202

K.S. Santos et al. / Composites: Part A 40 (2009) 11991209

INTENSITY (Cps)

30000

25000

50000

C-15A Powder
C-20A Powder
2,5% C-15A + 2,5% C-20A+1% PPG
2,5% C-15A + 2,5% C-20A + 5% PP-g-MA+1% PPG
1% C-15A + 1% C-20A+1% PPG
1% C-15A + 1% C-20A + 5% PP-g-MA+1% PPG

C-15A
C-20A
C-15A/C-20A

45000
40000

INTENSITY (Cps)

35000

20000

15000

35000
30000
25000

(a)

20000
15000
10000

10000
5000
0

5000

(b)

-5000
2

0
2

2
Fig. 5. WAXD patterns of the PP nanocomposites: (a) 2% MMT + 1% PPG and (b) 2%
MMT + 5% PP-g-MA + 1% PPG.

Fig. 4. WAXD patterns of the PP/C-15A/C-20A nanocomposites.

Table 2
Interlayer distance of the PP/C-15A, PP/C-20A and PP/C-15A/C-20A nanocomposites.
Samples

2h
2h
d(1 peak) d(2 peak)
(nm)
(1 peak) (2 peak) (nm)

C-15A
C-20A
PP/C-15A/PPG (97/2/1)
PP/C-20A/PPG (97/2/1)
PP/C-15A/PP-g-MA/PPG (92/2/5/1)
PP/C-20A/PP-g-MA/PPG (92/2/5/1)
PP/C-15A/PPG (94/5/1)
PP/C-20A/PPG (94/5/1)
PP/C-15AC-20A/PPG (94/5/1)
PP/C-15AC-20A/PP-g-MA/PPG (89/5/5/1)
PP/C-15AC-20A/PPG (97/2/1)
PP/C-15AC-20A/PP-g-MA/PPG (92/2/5/1)

2.6
3.3
2.3
2.3
2.5
2.4
2.4
2.4
2.4
2.4
2.4
2.4

4.6
4.6
5.0
4.8
4.8
4.8
4.8
4.8
4.8
4.8

3.39
2.68
3.84
3.84
3.53
3.68
3.68
3.68
3.68
3.68
3.68
3.68

1.93
1.93
1.77
1.84
1.84
1.84
1.84
1.84
1.84
1.84

The systems with 5 wt% MMT but without PP-g-MA showed a


higher d001 peak intensity indicating a higher number of agglomerate structures. The systems with 2 wt% C-20A without PP-g-MA
and 2 wt% C-15A with 5% PP-g-MA showed a decreased in the
d001 peak intensity, which indicated a partially exfoliated structure
(Fig. 3b). A similar behavior was also observed for the nanocomposites of the C-15AC-20A clays combination (Fig. 4). The d001
peak intensity was smaller for a lower MMT content (2 wt% MMT
mixture). The use of PP-g-MA reduced the intensity of the d001
peak, suggesting a breakdown of the agglomerated platelet and/
or partial exfoliation, especially when 1:2.5 (MMT:PP-g-MA) was
used.
To compare the behavior between clay combinations and individual clay, the d001 peaks were overlaid. The combination of 2%
C-20A-C-15A with 1% PPG exhibited an intermediary intensity between the d001 peaks of the nanocomposites obtained with either
C-20A or C-15A (Fig. 5a). The use of PP-g-MA decreased the d001
peak intensity from C-15A but not that from C-20A, while the clay
combination behaved similarly to the C-15A nanocomposites
(Fig. 5b).
A better understanding of the state of dispersion and orientation of the clay platelets in the PP matrix [2326] can be gleaned
from TEM analysis. This technique enables the evaluation of the aspect ratio of clay particles, which is one of the important factors
that determine the reinforcement efciency of clay nanocomposites [23,2730]. The maximum efciency of reinforcement is

reached when the parallel-exfoliated platelets are uniformly oriented [31,32]. Besides the evaluation of the orientation, the number of agglomerates, and the spacing between the MMT platelets
can also be obtained using TEM analysis.
The morphology of the nanocomposite using either the individual or combination of clays exhibited typical aggregates of MMT
platelets in the absence of PP-g-MA (Figs. 6a, b and 7a). The number of particles increased as the MMT concentration was increased,
with the particles remaining as large aggregates or tactoids. When
PP-g-MA was added, better dispersion of the clay was achieved, as
can be seen in Figs. 6c, d, 7b and 8. The morphological structures
were a combination of tactoids, intercalated, and partially exfoliated particles [33,1]. Due to the sledding of the sheets promoted
by PPG into the layers, elongated structures were formed. Systems
with 5 wt% MMT is more homogeneous than systems with 2 wt%
MMT. This is due to the presence of a higher amount of clay, which
increased the viscosity of the system, thereby increasing the shear
forces that led to the exfoliation of the platelets in the matrix [34].
The degree of exfoliation clearly increased as the ratio of PP-g-MA/
organoclay was increased, in a manner independent of the total
clay content. Furthermore, TEM images indicated the exibility of
MMT, which reinforces the matrix while simultaneously improving
its modulus and impact strength properties. However if the orientation is not achieved due to the greater exibility of MMT, the nal improvement in the properties of the matrix can be reduced.
The clay particles were better aligned in the ow direction at a
lower clay concentration. An exception is the PP/C-20A/PPG (97/
2/1) nanocomposite in which a prolonged intercalated structure
was absent, because this clay had the smallest basal distance,
which led to an interaction between layers that was larger than
that between the PPGclay interface (Fig. 6a). The nanocomposites
obtained with the combination of clay using 2 wt% MMT and 5 wt%
PP-g-MA yielded more highly oriented exfoliated structures than
the ones obtained with individual clays.
Statistical analyses of the TEM images were used to evaluate the
dependency between particle length/thickness and the use of PP-gMA in both the individual clays and the combination of clays [23].
In the absence of PP-g-MA, the particle length and thickness were
larger, which resulted in a higher number of the agglomerate structures with a higher aspect ratio. The use of combined clay (2 wt%)
led to a longer particle length (0.8 lm) than when the individual
clays (0.3 lm) were used, while similar thickness (0.06 lm) was
observed in both systems, resulting in a 2.67 times higher aspect

K.S. Santos et al. / Composites: Part A 40 (2009) 11991209

1203

Fig. 6. TEM images of PP nanocomposites: 2% MMT + 1% PPG: (a) C-20A and (b) C-15A; 2% MMT + 5% PP-g-MA + 1% PPG: (c) C-20A and (d) C-15A.

ratio in combined clay than that in individual clays (Figs. 9 and 11).
When PP-g-MA was added into the PP nanocomposites, the length/
thickness were smaller, indicating better dispersion in the matrix.
The nanocomposites using the combination of clay had a 4.1 times
higher aspect ratio than that of the individual clays: a length of
0.35 lm and a thickness of 0.015 lm, versus a length of 0.2 lm
and a thickness of 0.035 lm, respectively (Figs. 10 and 12). The
use of either a 2 or 5 wt% of combined MMT with 5 wt% PP-g-MA
resulted in the same distribution of particle (length/thickness)
(Figs. 12 and 13).
3.3. Relationship between morphology and mechanical properties
As shown in Table 3, the presence of MEK in the process
caused a slight increase in the impact strength of PP (from 34
to 51 J/m) because MEK increased the free volume between the
polymer chains, thus reducing the Tg and the exural modulus
of the matrix. On the other hand, the use of MEK and PPG as a
processing aid did not affect the mechanical properties of PP,
while their combination with the organoclay increased the impact
strength but did not inuence the exural modulus. When PP-gMA was used, the exural modulus was 7% higher for the system
with C-20A than with C-15A. The PP/MMT exhibited better impact strength when 2 wt% MMT was used instead of 5 wt%, especially for the C-15A systems, independent of the use of PP-g-MA
[15]. When a combination of 1:1 C-15A/C-20A was used, the addition of the PP-g-MA improved the clay dispersion and the orientation of its sheets in PP matrix while decreasing the number of
agglomerated structures. These agglomerated structures can act
as tensions concentrators, and their decrease induced a superior
simultaneous gain in both the exural modulus (12% more than

pristine PP) and the impact strength (1.5 times) than when individual clays were used.
The behavior of the storage modulus was similar to that of the
exural modulus although the variations were more signicant
(Table 3). The b transition temperatures (Tg) of the nanocomposites
were slightly smaller than that for the pristine PP (Fig. 14). The b
and a transition in the tan d curve gives an indication of the interactions in the matrixller interface, which yields an interphase or
a region in the ller surroundings that may change the physical
properties of the nanocomposite. It has been reported that a high
llermatrix interaction would lower the b transition peak intensity due to reduced friction at the matrixller interface [35,36].
However, the intensity of the b transition peak for polymers without the compatibilizer agent was similar to that in the pure polymer. The PP/C-15AC-20A nanocomposites displayed weak clayPP matrix interaction as indicated by the increase of the b transition peak intensity. An exception was the system with 2 wt%
MMT and 5 wt% PP-g-MA, which had a lower b transition intensity
due to the interface adhesion between matrix and clay, was promoted by the PP-g-MA/MMT.
The a transition peak, approximately at 86 C, is related to the
relaxations in the intracrystalline amorphous chains of the PP: the
higher the a temperature, the larger the spherulite size [37]. The
a peak intensity is also related to the number, length and orientation of amorphous segments in the crystal [38]. The reduction
in the spherulite size was indicated by the decrease in the a temperature of the combined clay. The amount of amorphous segments in the crystal interface increased in all systems, leading
to the higher impact strength of the PP nanocomposites mixture
[39], which was independent of the use of PP-g-MA or the
amount of clay used. However, an increase in interaction between

1204

K.S. Santos et al. / Composites: Part A 40 (2009) 11991209

Fig. 7. TEM images of PP/1% C-15A + 1% C-20A + 1% PPG mixture nanocomposites: (a) absence PP-g-MA (b) 5% PP-g-MA.

Fig. 8. TEM images of PP/MMT mix nanocomposites: 2.5% C-15A + 2.5% C-20A + 5% PP-g-MA + 1% PPG.

PP and the combined clay (2 wt% MMT mixture and 5 wt% PP-gMA) caused a slight decrease in the number of amorphous segments in the crystal interface. Furthermore, the a peak distributions of the combined clay were narrower than that of the
pristine PP, indicating more uniform crystals sizes in the combined clay.

3.4. Relationship between morphology and thermal properties


3.4.1. Crystallization and melting of PP nanocomposites
Mineral llers are known to be a nucleation agent for polymer
crystallization. The llers increase the crystallization temperature
(Tc) and thus promote changes in the crystal that may inuence

1205

K.S. Santos et al. / Composites: Part A 40 (2009) 11991209

(a)

20

60
55
50
45
40

Total N : 106
St. dev.: 0.38 m

10

Frequency

Frequency

15

35

Total N :106
St. dev.: 0.03 m

30
25
20
15
10
5

0
0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

1.8

2.0

2.2

0
0.00

2.4

0.02

0.04

Particle Length (m)

0.06

0.08

0.10

0.12

0.14

0.16

0.18

0.20

0.18

0.20

Particle Thickness ( m)

(b)

60

20

55
50
45

15

Total N : 123
St. dev.: 0.04 m

Total N : 123
St. dev.: 0.23 m

Frequency

Frequency

40

10

35
30
25
20
15

10
5
0
0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

1.8

2.0

2.2

2.4

0
0.00 0.02 0.04 0.06 0.08 0.10

0.12

0.14

0.16

Particle Thickness ( m)

Particle Length ( m)

Fig. 9. Histogram of MMT particle length and thickness in the PP/MMT nanocomposites: (a) 2% C-15A + 1% PPG (b) 2% C-20A + 1% PPG.

the reinforcement effects of the silicates in the nanocomposites [40].


Consequently, more crystallization nuclei may be formed when a
higher amount of clay with a better dispersion in matrix is used
[41]. The presence of clay, the addition of PPG as a processing aid,
and the use of PP-g-MA did not inuence the melt temperature
(Tm) or the relative degree of crystallinity (Xc), However there was
an increase in Tc when PP-g-MA was used (Table 4) due to both the
increase of the clay platelets distribution in matrix, and the enhancement of a strong interaction between the clay and the matrix.
The relative degree of crystallinity (Xc) measured as a function of
time (t) in a non-isothermal process may give an indication of the
crystallization rate. t1/2 is the time required for the PP to reach 50%
in the relative degree of crystallinity. t1/2 may be used to evaluate
the increase or the decrease in the crystallization rate of the different
systems [13,42]. Table 4 shows the t1/2 values obtained with a heating rate of 10 C/min. The use of 5 wt% of PP-g-MA caused a slight increase in Tc, and a decrease in t1/2. This decrease in t1/2 indicates an
increase in the crystallization rate in the initial stages when either
the individual or clay combination was used. However, when
2 wt% of C-20A/C-15A was used, the t1/2 values were smaller than
that for both the pristine PP and the individual clays.

formed a protecting barrier that impeded the release of gases from


the decomposition, hence decreasing the degradation temperature
of the material [43,44]. The use of PP-g-MA caused an increase in
the initial decomposition temperature of the PP nanocomposites.
The increase is due to the improved claypolymer adhesion, which
also facilitated the partial exfoliation of the platelets in the matrix,
thereby lengthening the tortuous path that hindered the diffusion
of gases from the decomposition [13]. However, the nal decomposition temperature (T50%) of all the nanocomposites were not
improved due to both the degradation of the ammonium salt
(210410 C), which led to an increased loss in the PP nanocomposites, and the presence of agglomerated layers in the nal morphology
[21,43,45].
The addition of clay into the PP matrix caused a signicant increase in the HDT from 6 C to 26 C for the individual clay (Table
4). For the clay combination, the highest values were obtained only
when higher amount of clay was used probably because the HDT
depend on both the mechanical and thermal stability. As shown
in Table 4, the initial thermal stability (T10%) of the individual clays
was superior to that of the combined clays and they reached the
same values when 5 wt% C-20A/C-15A system was used.

3.4.2. Thermal stability and heat deection temperature (HDT) of PP


nanocomposites
The incorporation of clay into the PP matrix increased the initial
decomposition temperature (T10%) for all the PP nanocomposites
(Table 4). This is due to the migration of MMT to the surface, which

4. Conclusions
PP nanocomposites were prepared using montmorillonite 1:1
(C-15A/C-20A) with different amounts of organic modiers.
The gain in impact strength was more evident when prolonged

1206

K.S. Santos et al. / Composites: Part A 40 (2009) 11991209

(a)

40
110
35
100
30

90
0

Total N : 145
St. dev.: 0.11 m

25

80
0

Frequency

Frequency

70
20

15

Total N : 145
St. dev.: 0.01m

60
50
40

10

30
20

5
10
0
0.00

0
0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

1.1

1.2

0.02

0.04

0.06

0.08

0.10

0.12

0.14

0.16

0.18

Particle Thickness (m)

Particle Length (m)

(b) 40

110

35
100
30

Total N : 123
St. dev.: 0.15 m

20

90

Frequency

Frequency

25

15

Total N : 123
St. dev.: 0.02 m

14
12
10
8

10

6
4

2
0
0.00

0
0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

1.1

1. 2

0.02

0.04

0.06

0.08

0.10

0.12

0.14

0.16

0.18

Particle Thickness (m)

Particle Length (m)

Fig. 10. Histogram of MMT particle length and thickness in the PP/MMT nanocomposites: (a) 2% C-15A + 5% PP-g-MA + 1% PPG (b) 2% C-20A + 5% PP-g-MA + 1% PPG.

(a)

(b) 110

24
22

100

20
90

18

14
12
10

80

Frequency

Frequency

16

Total n : 13
St. dev.: 0.42 m

70
0

Total n : 13
St. dev.: 0.02 m

60
50
40

8
6

30

20

10
0

0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 1.4

Particle Length (( m)

0.00

0.01

0.02

0.03

0.04

0.05

0.06

0.07

0.08

0.09

0.10

Particle Thickness ( m)

Fig. 11. Histogram of MMT particle length (a) and thickness (b) in the PP/MMT combination nanocomposites: 1% C-15A + 1% C-20A + 1% PPG.

intercalated structures were present. The modulus values were


also larger when partially exfoliated with orientation were obtained, the formation of which was favored when PP-g-MA was
used. However, the increase in modulus was not more signicant

due to the decrease in particles length induced by PP-g-MA and


the plasticizer nature of MEK. The use of 1:1 C-15A/C-20A led to
a simultaneous improvement in both the mechanical (the exural
modulus and impact strength) and morphological (the absence of

1207

K.S. Santos et al. / Composites: Part A 40 (2009) 11991209

(a)

(b)

24
22

100

20

80

18

60

16

40

12

Frequency

Frequency

Total n : 81
St. dev.: 0.12 m

14

10
8

Total n : 81
St. dev.: 0.04 m

20

4
3

6
2

2
0

0
0.00

0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 1.4

0.01

0.02

0.03

0.04

0.05

0.06

0.07

0.08

0.09

0.10

Particle Thickness (m)

Particle Length (m)

Fig. 12. Histogram of MMT particle length (a) and thickness (b) in the PP/MMT mixture nanocomposites: 1% C-15A + 1% C-20A + 5% PP-g-MA + 1% PPG.

(b)

(a) 24

100

22
20

80
0

18

Total n : 126
St. dev.: 0.15 m

40

14

Frequency

Frequency

16

Total n : 126
St. dev.: 0.01m

60

12
10
8

20

6
5
4

0
0.00

0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 1.4

0.01

0.02

0.03

0.04

0.05

0.06

0.07

0.08

0.09

0.10

Particle Thickness (m)

Particle Length ( m)

Fig. 13. Histogram of MMT particle length (a) and thickness (b) in the PP/MMT mixture nanocomposites: 2.5% C-15A + 2.5% C-20A + 5% PP-g-MA + 1% PPG.

Table 3
Mechanical and dynamic mechanical properties of the PP/clay mixture nanocomposites.
Sample

Flexural modulus (MPa)

Neat
Neat PP

1416 17

PP + MEK
Neat PP + MEK
1% PPG
2%C-15A + 1% PPG
2%C-20A + 1% PPG
2%C-15A + 5% PP-g-MA + 1% PPG
2%C-20A + 5% PP-g-MA + 1% PPG
5%C-15A + 1% PPG
5%C-20A + 1% PPG

1251 28
1459 5
1474 20
1482 20
1390 14
1512 37
1349 23
1502 29

1:1 C-20A/C-15A + 1% PPG


2.5% C-15A + 2.5%C-20A
1% C-15A + 1%C-20A

1530 13
1526 10

1:1 C-20A/C-15A + 1% PPG + 5% PP-g-MA


2.5% C-15A + 2.5%C-20A
1538 24
1% C-15A + 1%C-20A
1580 29

Storage modulus E 23 C (MPa)

b Transition Tg (C)

a Transition (C)

34 2

1565

14

86

51 6
49 7
92 7
68 9
95 12
73 10
70 4
66 6

1324
1592
1666
1864
1630
1728

1728

10
12
9
12
9
10

10

86
89
84
84
83
85

85

89 4
100 4

1678

10

81

115 5
164 22

1732
1753

12
10

78
77

Impact Izod 23 C (J/m)

agglomerated structure) properties of PP nanocomposites formed.


The properties were all superior to the cases in which individual
clay systems were used separately. The behavior of the storage

modulus was similar to that of the exural modulus although the


variations were more signicant. Tg of the nanocomposites was
slightly lower than that of the pristine PP. Nonetheless, Tm and Xc

1208

K.S. Santos et al. / Composites: Part A 40 (2009) 11991209


0.13
0.12
0.11
0.10

tan

0.09
0.08
0.07

PP+MEK
Neat PP
2,5% C20A+2,5% C15A+5% PP-g-MA+1% PPG
1% C20A+1% C15A+1% PPG
1% C20A+1% C15A+5% PP-g-MA+1% PPG

0.06
0.05
0.04
-20

20

40

60

80

100

120

TemperatureC
Fig. 14. The tan d peak: b transition and a transition of the C-20A-C-15A mixture/PP
nanocomposites.

Table 4
Thermal properties (TGA, DSC and HDT data) of the PP/clay mixture nanocomposites.
Sample

T10%a
(C)

T50%a
(C)

Tma
(C)

Tca
(C)

t
(min)

Xcb
(%)

HDTc
(C)

401
377
452
443
420

457
442
462
454
449

164
163
163
164
163

113
113
113
113
115

2.23
2.25
2.44
2.34
1.23

53
57
53
56
54

86
83
107
92
94

Neat
Neat PP
PP + MEK
2%C-15A + 1% PPG
2%C-20A + 1% PPG
2%C-15A + 5% PP-gMA + 1% PPG
2%C-20A + 5% PP-gMA + 1% PPG
5%C-15A + 1% PPG
5%C-20A + 1% PPG

410

450

163

115

1.53

55

103

427
448

456
462

163
165

115
114

2.18
2.35

54
58

101
112

1:1 C-20A/C-15A + 1% PPG


2.5% C-15A + 2.5%C-20A
1% C-15A + 1%C-20

420
404

448
445

163
162

113
113

1.99
1.01

55
57

107
86

1:1 C-20A/C-15A + 1% PPG + 5% PP-g-MA


2.5% C-15A + 2.5%C-20A
431
456
1% C-15A + 1%C-20A
429
451

163
164

116
115

1.54
1.71

56
57

107
88

T10% temperature at which 10% of weight loss occurs.


T50% temperature at which 50% of weight loss occurs.
a
Standard deviation 1 C.
b
Standard deviation 10%.
c
Standard deviation 2 C.

were not affected by the presence of clay, PP-g-MA, and any processing aid. There was however, a slight nucleation effect in the
crystallization of PP, that was reected by an increasing HDT and
thermal stability when the 5 wt% PP-g-MA and 5 wt% MMT mixture was used.
Acknowledgments
The authors are grateful to CAPES, CNPq, Finep, and FAPERGS/
PRONEX for their nancial support.
References
[1] Dong Y, Bhattacharyya D. Effects of clay type, clay/compatibiliser content and
matrix viscosity on the mechanical properties of polypropylene/organoclay
nanocomposites. Composites: Part A 2008;39:117791.

[2] Calcagno CIW et al. The role of the MMT on the morphology and mechanical
properties of the PP/PET blends. Compos Sci Technol 2008;68:2193200.
[3] Wilkinson AN, Man Z, Stanford JL, Matikainen P, Clemens ML, Lees GC, et al.
Tensile properties of melt intercalated polyamide 6 montmorillonite
nanocomposites. Compos Sci Technol 2007;67:33608.
[4] Paul DR, Robeson LM. Polymer nanotechnology: nanocomposites. Polymer
2008;49:3187204.
[5] Ray SS, Okamoto M. Polymer/layered silicate nanocomposites: review from
preparation to processing. Prog Polym Sci 2003;28:1539641.
[6] Hussain F, Hojjati M, Okamoto M, Gorga R. Review article: polymermatrix
nanocomposites, processing, manufacturing, and application: an overview. J
Compos Mater 2006;40:151175.
[7] Esteves ACC, Timmons AB, Trindade T. Nanocompsitos de matriz
polimrica: estratgias de sntese de materiais hbridos. Qum Nova
2004;27(5):798806.
[8] Zhu L, Xanthos MJ. Effects of processing conditions and mixing protocols on
structure of extruded polypropylene nanocomposite. Appl Polym Sci
2004;93:18919.
[9] Lpez-Quintanilla ML, Snchez-Valds S, Ramos de Valle LF, Miranda RG.
Preparation
and
mechanical
properties
of
PP/PP-g-MA/Org-MMT
nanocomposites with different MA content. Polym Bull 2006;57:38593.
[10] Zeng QH, Yu AB, Lu GQ, Paul DR. Clay-based polymer nanocomposites:
research
and commercial
development.
J
Nanosci
Nanotechnol
2005;5:157492.
[11] Meneghetti P, Qutubuddin S. Application of mean-eld model of polymer melt
intercalation in organo-silicates for nanocomposites. J Colloids Interf Sci
2005;288(2):3879.
[12] Ton-That MT, Perrin-Sarazin F, Cole KC, Bureau MN, Denault. Polyolen
nanocomposites: formulation and development. J Polym Eng Sci
2004;44(7):12129.
[13] Calcagno CIW, Mariani CM, Teixeira SR, Mauler RS. The effect of organic
modier of the clay on morphology and crystallization properties of PET
nanocomposites. Polymer 2007;48:96674.
[14] Leszczynska A, Njuguna J, Pielichowski K, Banerjee JR. Polymer/
montmorillonite nanocomposites with improved thermal properties. Part I.
Factors inuencing thermal stability and mechanisms of thermal stability
improvement. Thermochim Acta 2007;453:7596.
[15] Santos KS, Liberman AS, Oviedo MAS, Mauler RS. Polyolen-based
nanocomposite: the effect of organoclay modier. J Polym Sci: Part B: Polym
Physics 2008;46:251931.
[16] Zhao Z, Tang T, Qin Y, Huang B. Effects of surfactant loadings on the dispersion
of clays in maleated polypropylene. Langmuir 2003;19:71579.
[17] Cui L et al. Effect of organoclay purity and degradation on nanocomposite
performance. Part 1: surfactant degradation. Polymer 2008;49:375161.
[18] Liberman S, Da Silva L, Pelegrine T, Barbosa R, Mauler R. Patent numbers:
W02007009200-A2; BR200503777-A; 2007.
[19] Amash A, Zugenmaier P. Thermal and dynamic mechanical investigations on
ber-reinforced polypropylene composites. J Appl Polym Sci 1997;63:1143.
[20] Kim NH, Malhotra SV, Xanthos M. Modication of cationic nanoclays with
ionic liquids. Micropor Mesopor Mater 2006;96:2935.
[21] Lewin M, Pearce ME, Levon K, Mey-Marom A, Zammarano M, Wilkie CA, et al.
Nanocomposites at elevated temperatures: migration and structural changes.
Polym Adv Technol 2006;17:22634.
[22] Lee EC, Mielewski DF, Baird R. Exfoliation and dispersion enhancement in
polypropylene nanocomposites by in-situ melt phase ultrasonication. J Polym
Eng Sci 2004;44(9):177382.
[23] Kim DH et al. Structure and properties of polypropylene-based nanocomposites:
effect of PP-g-MA to organoclay ratio. Polymer 2007;48:530823.
[24] Garcia-Lopez D, Picazo O, Merino JC, Pastor JM. Polypropyleneclay
nanocomposites: effect of compatibilizing agents on clay dispersion. Eur
Polym J 2003;39(5):94550.
[25] Vermogen A, Masenelli-Varlot K, Seguela R, Duchet-Rumeau J, Boucard S, Prele
P. Evaluation of the structure and dispersion in polymer-layered silicate
nanocomposites. Macromolecules 2005;38:96619.
[26] Morgan AB, Gliman JW. Characterization of polymer-layered silicate (clay)
nanocomposites by transmission electron microscopy and X-ray diffraction: a
comparative study. J Appl Polym Sci 2003;87:132938.
[27] Fornes TD, Hunter DL, Paul DR. Effect of sodium montmorillonite source on
nylon 6/clay nanocomposites. Polymer 2004;45:232131.
[28] Yoon PJ, Fornes TD, Paul DR. Thermal expansion behavior of nylon 6
nanocomposites. Polymer 2002;43:672741.
[29] Chavarria F, Paul DR. Comparison of nanocomposites based on nylon 6 and
nylon 66. Polymer 2004;45:850115.
[30] Lee H-S, Fasulo PD, Rodgers WR, Paul DR. TPO based nanocomposites. Part 1.
Morphology and mechanical properties. Polymer 2005;46:1167389.
[31] Kanny K, Moodley VK. Characterization of polypropylene nanocomposite
structures. J Eng Mater Technol 2007;129:10512.
[32] Szzdi L, Abranyi A, Puknszky Jr B, Vancso JG, Puknszky B. Morphology
characterization of PP/clay nanocomposites across the length scales of the
structural architecture. Macromol Mater Eng 2006;291:85868.
[33] Kim DH et al. Effect of the ratio of maleated polypropylene to organoclay on
the structure and properties of TPO-based nanocomposites. Part I: morphology
and mechanical properties. Polymer 2007;48:596078.
[34] Wong S-C, Lee H, Qu S, Mall S, Chen L. A study of global vs. local properties for
maleic anhydride modied polypropylene nanocomposites. Polymer
2006;47:747784.

K.S. Santos et al. / Composites: Part A 40 (2009) 11991209


[35] Correa CA, Razzino CA, Hage EJRJ. Role of maleated coupling agents on the
interface adhesion of polypropylenewood composites thermoplastic. Compos
Mater 2007;20:32339.
[36] Chua PS. Dynamic mechanical analysis studies of the interphase. Polym
Compos 1987;8:30813.
[37] Lim JW, Hassan A, Rahmat AR, Wahit MU. Morphology, thermal and
mechanical behavior of polypropylene nanocomposites toughened with
poly(ethylene-co-octene). Polym Int 2006;55:20415.
[38] Tiemblo P, Gomez EJM, Beltran SG, Matisova-Rychla L, Rychly J. Melting and a
relaxation effects on the kinetics of polypropylene thermooxidation in the
range 80170 C. Macromolecules 2002;35:59226.
[39] Tanniru M, Yuan Q, Misra RDK. On signicant retention of impact strength in
clay-reinforced high-density polyethylene (HDPE) nanocomposites. Polymer
2006;47:213346.
[40] Szzdi L et al. Factors and processes inuencing the reinforcing effect of
layered silicates in polymer nanocomposites. Eur Polym J 2007;43:34559.

1209

[41] Krump H, Luyt AS, Hudec I. Effect of different modied clays on the thermal
and physical properties of polypropylenemontmorillonite nanocomposites.
Mater Lett 2006;60:287780.
[42] Jain S, Goossens H, Duin VM, Lemstra P. Effect of in situ prepared silica nanoparticles on non-isothermal crystallization of polypropylene. Polymer
2005;46:880518.
[43] Tang Y, Lewin M. Maleated polypropylene OMMT nanocomposite: annealing,
structural changes, exfoliated and migration. Polym Degrad Stabil
2007;92:5360.
[44] Zhang S, Horrocks AR. A review of ame retardant polypropylene bres. Prog
Polym Sci 2003;28:151738.
[45] Golebiewski J, Galeski A. Thermal stability of nanoclay polypropylene
composites by simultaneous DSC and TGA. Compos Sci Technol
2007;67:34427.

You might also like