Vous êtes sur la page 1sur 11

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/332947966

Thermal and dielectric behavior of polyamide-6/clay nanocomposites

Article · April 2019

CITATIONS READS

0 120

5 authors, including:

Imen Hammami Helmi Hammami


University of Sfax University of Sfax
1 PUBLICATION   0 CITATIONS    10 PUBLICATIONS   142 CITATIONS   

SEE PROFILE SEE PROFILE

Mourad Arous A. Kallel


University of Sfax University of Sfax
71 PUBLICATIONS   808 CITATIONS    119 PUBLICATIONS   1,008 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Hybrid Materials for Optical and Optoelectronic Applications View project

Adhesion analysis of Alfa fibers in unsaturated polyester matrix. View project

All content following this page was uploaded by Imen Hammami on 13 May 2019.

The user has requested enhancement of the downloaded file.


Materials Chemistry and Physics 232 (2019) 99–108

Contents lists available at ScienceDirect

Materials Chemistry and Physics


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

Thermal and dielectric behavior of polyamide-6/clay nanocomposites


Imen Hammami a, *, Helmi Hammami a, J�er�emie Soulestin b, Mourad Arous a, Ali Kallel a
a
LaMaCoP, Faculty of Sciences of Sfax, Department of Physic, University of Sfax, 3018, Sfax, Tunisia
b
TPCIM, IMT Lille Douai, Department of Polymers and Composites Technology and Mechanical Engineering, University of Lille, 59500, Douai, France

A R T I C L E I N F O A B S T R A C T

Keywords: In this work, the influence of the incorporation of nanoparticles (organo-modified montmorillonite Cloisite 30B)
Polymer in polyamide 6 (PA6) on rigid amorphous fraction (RAF) formation had been explored employing Differential
Composite Scanning Calorimetry (DSC), Flash Differential Scanning Calorimetry (Flash DSC) and Broadband Dielectric
Interface
Spectroscopy (BDS) techniques. The existence of a RAF in PA6-montmorillonite nanocomposite films is available
from specific heat capacity measurement at the glass transition region of the nanocomposites. It was shown that
at high C30B content, this fraction becomes larger. Using Flash DSC, it was possible not only to measure the heat
capacity step at the glass transition of the materials, but also to provide quantitative knowledge on the kinetics of
crystallization and nucleation of PA6-based nanocomposites. The dielectric relaxation spectroscopy measure­
ment was investigated, in the frequency range 0.1–106 Hz and varying temperature from 20 to 200 � C, which
highlight different relaxation phenomena: the α dipolar relaxation, the αc relaxation and Max­
well–Wagner–Sillars (MWS) interfacial polarizations. As C30B content increases, a MWS relaxation emerges in
the nanocomposites, thus revealing the increase of RAF in the nanocomposite with high C30B content.

1. Introduction balance cost/performance [2]. The presence of clay nanoparticles in the


polyamide 6 induces an improvement in its properties and in particular
Polymer nanocomposites have attracted a great deal of attention in those related to barrier against water and gases.
recent years due to their exceptional properties. Indeed, filling polymers In this work, the interfacial effects in polymer PA6/MMT nano­
with nanoparticles, particularly those containing high aspect ratio composites were investigated by several techniques. Differential Scan­
nanofillers, is used to enhance physical characteristics, such as increased ning Calorimetry (DSC) is such a technique, and it is used to study the
mechanical strength, resistance to dielectric breakdown, and improved interfacial effects in polymer nanocomposites by focusing on glass
thermal behavior [1]. transition region. Indeed, Privalko et al. [3,4] recognized the impor­
Among the Numerous types of nanosized fillers, clay and layered tance of specific heat capacity measurements for the identification of a
silicates are the most commonly used nanofiller because of their natural fraction of reduced mobility in nanocomposites. Following these ideas,
abundance, their low cost and their high aspect ratio (greater than 100). the existence of a rigid amorphous fraction (RAF) in poly(methyl
To optimize and to control the montmorillonite (MMT) filled nano­ methacrylate) (PMMA)/SiO2 nanocomposites was shown by applying
composites performance, it is essential to increase as much as possible heat capacity measurements at the glass transition of the polymer as
the degree of nanoparticles exfoliation and dispersion within polymer described by Wunderlich et al. [5,6]. Then, a ‘3-phase-model’ was
matrix for creating large interaction surfaces promoting the formation of established for semicrystalline polymers where the polymer matrix
interphase. These interphases are formed due to the existence of an consists of the mobile amorphous fraction (MAF), the crystalline fraction
amorphous polymer fraction immobilized around the nanoparticles. (CF), and the RAF.
Several parameters influence the state of dispersion and exfoliation such Complementary to thermal analysis, Broadband Dielectric spectros­
as the matrix type, the clay treatment and the processing conditions. copy (BDS) is one of the most powerful methods used to evaluate the
Polyamide 6 (PA6), also called Nylon 6, have been widely used in different molecular mobilities of complex nanostructured systems and
packaging applications due to their good mechanical resistance, the high provide an experimental access to the interfacial effects. The paper aims,
barrier against gas and water, the fire resistance and the excellent using the correlation between thermal and dielectric analysis, to

* Corresponding author.
E-mail address: imen.hammami1992@gmail.com (I. Hammami).

https://doi.org/10.1016/j.matchemphys.2019.04.048
Received 23 January 2019; Received in revised form 3 April 2019; Accepted 16 April 2019
Available online 20 April 2019
0254-0584/© 2019 Elsevier B.V. All rights reserved.
I. Hammami et al. Materials Chemistry and Physics 232 (2019) 99–108

characterize the RAF induced by clay nanoparticles. dynamics of the processes involved. In this study, stochastic TMDSC runs
were carried out on a Mettler-Toledo DSC1 instrument. The measure­
2. Experimental section ments were made using samples with a mass of about 20 mg, at under­
lying heating rate of 1 � Cmin-1 between 60 � C and 250 � C. The
2.1. Materials amplitude of the temperature pulse was set at � 0.5 � C and the switching
time range from 25 to 50 s.
A commercially available polyamide 6 (Akulon, F132E1, DSM, NL)
with a molecular weight of 66000 g/mol and density of 1.13 g/cm3, was 2.3.3. Flash differential scanning calorimetry (Flash-DSC)
used in this study. As reinforcement, the cloisite 30B (C30B) organo- In order to evaluate the mobility of macromolecular chains in un­
modified, contains the cation methyl tallow bis-2-hydroxyethyl, seems filled films and composite films, Flash DSC experiments were performed
to be a good choice. In fact, the cloisite 30B, as already shown before, using a Flash DSC 1 instrument provided by Mettler-Toledo [9].
leads to a high exfoliation degree in PA6 [7]. The presence of hydrogen Contrary to conventional DSC instrument, a tiny fragment of sample
bond between the hydroxyl groups and the amide groups of the polymer (with a mass in the order of nanograms) was placed onto a MultiSTAR
chains facilitates the delamination. UFS 1 chip sensor, based on MEMS (Micro-Electro-Mechanical Systems)
sensor technology. Heating-cooling cycles in dry nitrogen gas are
2.2. Processing of PA6 nanocomposite films essential to ensure that the sample has established a good thermal
contact to the sensor.
The nanocomposites PA6/C30B were obtained using a two-step Measuring the weight of such low masses is not feasible with a
treatment: the first step called masterbatch step, corresponded to the common balances. Therefore, the sample masses were estimated by
preparation of a mixture containing the PA6 with 12% by weight of applying an identical protocol on the sample placed on the conventional
C30B using a twin-screw extruder type Clextral BC45 at a rotation speed DSC and on another sample inserted in the Flash DSC. Finally, the
of 50 rpm and a temperature ranging from 240 � C to 270 � C. The second evaluation of the samples masses can be easily done via the melt
step was the dilution of this masterbatch into a Haake Buchler Rheocord enthalpy determined from the two DSC curves. For the PA6 matrix, the
40 single-screw extruder at a temperature and screw speed of 230 � C and mass was estimated at 106 ng, while for the filled PA6, it was estimated
50 rpm. at 27 ng for PA6þ3%C30B, 131 ng for PA6þ5%C30B and 201 ng for
Nanocomposites films of 200–250 μm in thickness with different clay PA6þ7%C30B.
content (3%, 5% and 7%) are finally obtained using single extrusion. To study the kinetics of crystallization and nucleation, Flash DSC
Prior to processing into films, the PA6 polymer and the nano-fillers measurements were made in a temperature range of 90 � C to 250 � C
C30B are dried at 80 � C under vacuum overnight to remove residual using different heating and cooling rates (from 10 � Cs-1 to 1000 � Cs-1).
water moisture. After carrying out the calorimetric measurements of the nano­
composite films, the RAF can be determined by applying the method
described by Xenopoulos and Wunderlich [10] for semi-crystalline
2.3. Experimental procedure polymers. From the heat capacity measurements at the glass transition
region, they led to establish a model that makes it possible to distinguish
2.3.1. Transmission electron microscopy (TEM) between the crystalline fraction CF, the rigid amorphous fraction RAF
The morphologies of the nanocomposite films were analyzed by and the mobile amorphous fraction MAF. This model is called a
transmission electron microscopy (TEM). three-phase model.
The specimens of 70 nm nominal thickness were cut with a diamond From the results obtained by the TMDSC and the DFSC, the RAF can
knife on an ultracut microtome (Leica UCT, Switzerland), at cryo tem­ be determined using the following equation:
perature ( 120 � C).
Bright-field TEM images of nanocomposites were obtained at 300 kV Xrað%Þ ¼ 1 ðXc þ xmaÞ (2)
under low dose conditions with an electronic microscope (Philips
CM30), using a Gatan CCD camera (Gatan, USA). where Xc is the crystallinity degree, calculated using eqn (1), and Xma is
the mobile amorphous phase fraction, deduced from eqn (3).
2.3.2. Differential scanning calorimetry (DSC) ΔCp
Differential Scanning Calorimetry (DSC) was performed to study the Xmað%Þ ¼ (3)
ΔCp ðamÞ
effect of C30B incorporation on PA6 melting and crystallization
behavior, hence the percentage of RAF. Δcp and Δcp(am) are the measured heat capacity increment at the
The DSC experiments were realized on a Mettler Toledo DSC-1 in­ glass transition region of the semi-crystalline and 100% amorphous
strument under a nitrogen atmosphere at heating and cooling rate of polymer respectively. Δcp(am) can be determined experimentally by the
50 � C min-1 and 10 � C min-1 respectively. In all cases, the samples, with Flash DSC measurements or taken theoretically from ATHAS database
a mass ranging between 2 and 4 mg, were heated from 90 � C to 250 � C, [11].
cooled from 250 � C to 90 � C, then followed by a second heating step
from 90 � C to 250 � C. 2.3.4. Dielectric Spectroscopy (DS)
The crystallinity index can be calculated from melting enthalpy by Dielectric Spectroscopy measurements were performed on a Novo­
the following relation: control Impedance Spectrometer based on an Alpha analyzer and a
Quatro temperature controller in the temperature ranged from to
XC ¼
ðΔHm ΔHc Þ
(1) 20 � C–200 � C on heating at a rate of 5� Cmin-1 and in a broad frequency
ΔHmo ð1 ϕÞ window from 0.1 Hz to 1 MHz.
For the dielectric analysis, an alternating voltage with amplitude 1V
where ΔHm is the melting enthalpy, ΔHc is the cold crystallization was applied to a sample placed between two parallel plate electrodes.
enthalpy, ΔHm0 is the theoretical melting enthalpy of 100% crystalline The measures dielectric permittivity data were collected and evaluated
PA6 (ΔHm0 ¼ 190 J/g [8]) and ϕ is the mass fraction of C30B intro­ using a WinDETA impedance analysis software.
duced into PA6. The dielectric behavior is investigated with the complex dielectric
Temperature Modulated Differential Scanning Calorimetry (TMDSC) constant ε* and the complex electric modulus M * expressed as [12,13]:
is a method that allows a separation of reversing and non-reversing
processes. Hence, this technique provides more information about the

100
I. Hammami et al. Materials Chemistry and Physics 232 (2019) 99–108

Fig. 1. TEM micrographs of (a, b) PA6 þ 3% C30B, (c, d) PA6 þ 5% C30B and (e, f) PA6 þ 7% C30B [14].

ε* ¼ ε’ jε} (4)

1 ε’
M* ¼ ¼ M’ þ iM} ¼ (5)
ε* ε’2 þ ε}2

where ε ’ is the real and ε ’’ is the imaginary parts of the dielectric


constant, M’ and M00 are, respectively, the real and imaginary parts of the
electric modulus.

3. Results and discussion

3.1. Morphological analysis

Fig. 1 shows TEM images of PA6 matrix and PA6/C30B nano­


composites taken at two magnifications (100 n.m. and 200 n.m.)
The TEM observation reveal that the PA6/C30B nanocomposites
contained individual nanoplatelets with more or less associated platelets
observed at law filler content. Therefore, PA6/C30B films present a high
degree of exfoliation of C30B nanoplatelets for the law and hight C30B
content. Moreover, the TEM images show a high degree of orientation of Fig. 2. DSC thermograms of PA6 matrix and PA6/C30B nanocomposites.
the nanoclay with an orientation parallel to the surface of the film. This
effect is associated to the processing condition, in other word to the presence of the nanoclay reinforcements. Therefore, it is necessary to
shearing forces induced in during the extrusion. These result are more determine the effect of the incorporation of these MMT nanolayers on
exploited in alix et al. research work [14]. the crystallinity degree. Fig. 2 shows thermograms obtained from ther­
mal analysis using differential scanning calorimetry (DSC). We note that
all the samples display an endothermic peak located in the region of
3.2. Crystalline structure and crystallinity
215 � C, associated with the fusion of the α-form crystals of PA6. We also
observe a second peak at 175 � C for the neat PA6 while it is centered at
In semi-crystalline polymers, the formation of the RAF can be
195 � C for the nanocomposites. This peak is due to the fusion of γ-form
affected by a change in the crystalline state of the material due to the

101
I. Hammami et al. Materials Chemistry and Physics 232 (2019) 99–108

Table 1
Crystallinity index of PA6 matrix and PA6/C30B
nanocomposites.
Samples Xc (%)

PA6 27.7%
PA6þ3% C30B 30.9%
PA6þ5% C30B 31.8%
PA6þ7% C30B 31.4%

Fig. 4. Molar heat capacity of PA6 and PA6/C30B matrix in the glass transi­
tion region.

contributes therefore to the heat capacity step at Tg. According to the


DSC measurements, the RAF is considered completely rigid and does not
contribute to the glass transition. In order to determine the RAF, precise
measurements of heat capacity performed by the TMDSC technique in
the temperature range of the glass transition of the polymer nano­
composite are required.
Fig. 3. Molar heat capacity of PA6 in the glass transition region: measured Fig. 3 shows the evolution of the specific heat capacity of the PA6
curve for the semi-crystalline PA6; determined by TMDSC (TOPEM) measurement.
amorph (Cp am) and cristal (Cp cr) reference curves from ATHAS The standard method evaluates the glass transition by using tangents
database; expected molar heat capacity from a 2-phase model; molar at the heat capacity curve above and below the glass transition region.
heat capacity calculated from a 3-phase model; the tangent on the The glass transition temperature (Tg) is the temperature at which the
measured curve above the glass transition region.
measured curve is equidistant between tangents. Then, the step height at
the glass transition Δcp is specified as the difference between the upper
crystals. In fact, according to the literature [15–17], the PA6 exhibits a and lower tangents at Tg.
polymorphic structure that presents two major crystal forms: the α-form In Fig. 3, the tangent below the glass transition coincides with the
and the γ-form. This polymorphic behavior is dependent on the pro­ heat capacity of the crystal (obtained from ATHAS database), while the
cessing conditions, thermal history, crystallization conditions and me­ tangent traced above the glass transition is shown in dashed lines. In
chanical stress. The most thermodynamically stable phase is the α-form. order to fit the measured curve above Tg, a 2-phase model was applied
This phase has a monoclinic structure, in which the anti parallel polymer using the following equation:
chains are linked by hydrogen bonding. The γ-form crystal structure is
monoclinic or pseudo-hexagonal, in which hydrogen bonds are formed Cpsc(2-phase model) ¼ Cpam � (1 CF)þCpcr � CF (6)
between parallel polymer chains. Furthermore, the α and γ structures
We note that the heat capacity measured just above the Tg is lower
coexist in the presence of MMTs in which the γ structure occurs near the
than that expected from the 2-phase model. Consequently, a 3-phase
interfaces of the nanolayers. For the α structure, it develops where the
model, taking into account the RAF, has to be applied according to
nanoplates do not affect the conformation of the chain and the folding of
Equation (7).
the PA6. The cause of this is that MMT hampers the displacement of the
Hydrogen bonds associated with the transformation from γ structure to α Cpsc(3-phase model) ¼ Cpam � MAF þ Cpcr � CF þ Cpra � RAF (7)
structure.
Therefore, the difference in melting temperatures of the γ-form It is highlighted that for semi-crystalline polymers, two glass tran­
crystal between the pure PA6 and the nanocomposites may be due to a sitions can be detected. The first transition results from the MAF and the
growth in formation of crystal nuclei caused by the presence of MMT in second transition arises from the RAF [18].
the nanocomposites. Above the glass transition of the MAF, the RAF is still in the vitreous
From the two endothermic peaks associated with the fusion of the state and it can be considered as solid and subsequently its heat capacity
α-form crystals and the fusion of the γ-form observed in the DSC ther­ is equal to the solid Cp.
mograms (Fig. 2) and using the equation (1), the crystallinity index Xc of Equation (7) can be described as follows:
the samples could be deduced. These values are reported in Table 1.We Cpsc (3-phase model) ¼ Cpam � MAF þ Cpcr � SF (8)
notice a slight increase in crystallinity index following the introduction
of MMT. This increase confirms the formation of new crystal nuclei Where SF ¼ CF þ RAF: solid fraction.
caused by the presence of MMT. The comparison between the slope of the tangent traced above the
glass transition (dashed line) and the slope of the semi-crystalline heat
3.3. Determination of the rigid amorphous fraction capacity shows a non-negligible contribution added to the measured
thermal capacity. This contribution can be either an excess heat ca­
For semi-crystalline polymers, a RAF is broadly discussed [5]. This pacity, or a continued extension of the glass transition to higher tem­
fraction is defined as the polymer, which behaves liquid like, e.g., peratures (successive devitrification of the RAF) [19].
vitrified or devitrified in the common glass transition region (Tg) and The same formalism can be applied for nanocomposites taking into

102
I. Hammami et al. Materials Chemistry and Physics 232 (2019) 99–108

Table 2 Table 3
The thermal parameters of PA6 and its nanocomposites: glass transition tem­ The thermal parameters of PA6 and its nanocomposites determined from Flash
perature (Tg), heat capacity step (ΔCp), crystalline fraction (CF), mobile DSC measurements at a heating rate of 1000 � C.s-1: glass transition temperature
amorphous fraction (MAF) and rigid amorphous fraction (RAF). (Tg), heat capacity step (ΔCp), crystalline fraction (CF), mobile amorphous
Sample Tg ΔCp (J CF (%) MAF RAF
fraction (MAF) and rigid amorphous fraction (RAF).
(� C) mol 1K 1) (%) (%) Sample Tg (� C) ΔCp (J CF MAF RAF
mol 1K 1) (%) (%) (%)
PA6 52 18.35 27.7 34.1 38.2
PA6þ3 wt% 51 15.63 30.9 29.1 40.0 PA6 60.5 36.93 21.8 68.7 9.5
C30B PA6þ3 wt% 61.5 32.7 28.1 60.8 11.1
PA6þ5 wt% 52 14.78 31.8 27.5 40.7 C30B
C30B PA6þ5 wt% 59.5 31.68 30.0 59.0 11.0
PA6þ7 wt% 53 14.07 31.4 26.2 42.4 C30B
C30B PA6þ7 wt% 60.5 28.14 33.6 52.4 14.0
C30B

Fig. 6. Diagram of time-temperature profile of the Flash DSC experiments used


to investigate the nucleation and crystallization in PA6 and PA6/C30B films.

Table 3. The values of ΔCp and ΔHf are determined from the Flash DSC
Fig. 5. Flash DSC thermograms of PA6 matrix and PA6/C30B nanocomposites.
curves (Fig. 5). Subsequently, a correction of these values with respect to
the masses of the samples was made.
account the crystalline fraction from MMT incorporation [20]. To The results observed in Table 3 confirm that obtained from the
determine the RAF, the heat capacity measurements, using the TOPEM TMDSC measurements despite the remarkable difference in the per­
method, in the temperature range of the glass transition were performed centages of MAF and RAF. This difference is due to the fact that in these
for the different samples of PA6/C30B. The results are shown in Fig. 4. two techniques we did not use the same heating and cooling rates.
As shown in Fig. 4, the heat capacity curves are shifted towards lower
values by increasing filler concentration. The lowering is due to a part of 3.4. Nucleation and crystallization kinetics
the polymer immobilized by the C30B which does not contribute to the
glass transition. As known, Flash DSC is the ideal technique for studying nucleation
From the obtained curves, the heat capacity step at Tg of PA6 and its and crystallization kinetics in PA6 and PA6/C30B nanocomposites
nanocomposites can be determined. Eventually, the MAF and the RAF during cooling. A simple analysis protocol is schematized in Fig. 6.
can be deducted. The value of ΔCp(am) used for the calculation of the The advantage of this analysis is its ability to acquire additional in­
MAF is determined from the ATHAS database. This value corresponds to formation on the crystallization, nucleation and stability of objects
53.7 J.K-1.mol. The ΔCp, MAF, CF and MAF are available and given in formed by controlled cooling. In fact, information on the interaction of
Table 2. From the Table 2, it can be noticed that ΔCp is directly related the growing objects and the surrounding melt can be determined from
to MAF. This result means that MAF decreases in the presence of MMT. changes in the increase in thermal capacity at Tg. Then, the available
Therefore, the increase in the percentage of MMT introduced in the PA6 mobile and crystallizable material is given by the analysis of the cold
causes the increase of the RAF. The incorporation of the nanoparticles in crystallization peak. In addition, in noncrystalline samples, the cold
the matrix affects the polymer layer adjacent to these additives by crystallization enthalpy appears to be a good measure of the number of
stopping the movement of this layer. Consequently, this reduction in available nuclei, and the temperature of the cold crystallization range is
molecular mobility tends to increase RAF. In addition, the RAF is related related to changes in the type of nuclei present. Finally, the melting
to the degree of dispersion of C30B in the matrix [21]. In fact, the per­ temperature of the different species can be related to the stability of the
centage of RAF increases for nanocomposites with a higher degree of objects and the change of total enthalpy at heating provides information
dispersion and higher clay content. This is the case of PA6þ7 wt% C30B the total crystallinity present in the sample before the heating scan [24].
nanocomposite which is exfoliated and has the higest clay content [7]. It A set of heating curves after cooling at different rates is shown in
can be expected that the RAF will increase with increasing the filler Fig. 7.
content. However, at a high filler content, it is known that the C30B For PA6, the melting peak decreases and becomes zero at cooling
tended to aggregate and that will lead to the decrease of the RAF [22, rates above 500 � C.s-1. Then, no crystallization occurs in this case.
23]. At low cooling rates, the nanocomposites show at the glass transition
To confirm the results obtained by the TMDSC technique, the Flash range a relatively small and broadened increment, no cold crystalliza­
DSC was used in this study. To calculate the RAF, the same approach tion and a large melting peak. While for higher cooling rates, the
described previously was used. The results were calculated and shown in increment in the glass transition increases and becomes more noticeable

103
I. Hammami et al. Materials Chemistry and Physics 232 (2019) 99–108

Fig. 8. Temperature dependence of real (a) and imaginary (b) parts of the
dielectric permittivity for the PA6 matrix.

due to the onset of a cold crystallization peak. This peak develops and its
intensity increases for high cooling rates. This increase is explained by a
growth in the formation of crystalline nuclei. In addition, the melting
peak decreases as a result of low crystallinity developed during cooling.
It can be deduced that, depending on the cooling rates, different
states of the materials can be detected allowing modeling of their
thermal histories. Indeed, a slow cooling leads to a thickening and better
stability of the crystallites. This thickening occurred through homoge­
neous and heterogeneous nucleation followed by growth, i.e., the
transformation of low-perfection crystallites into thicker crystalline
lamellae with a higher degree of perfection. By increasing the cooling
rate, the sample cannot complete its crystallization during cooling.
Initially, a homogeneous and heterogeneous nucleation occurs but fewer
crystals grow. The remaining nuclei induce cold crystallization.at higher
rates of cooling the homogeneous and heterogeneous nuclei in their
growth do not reach a perfect size, i.e., the amount of crystals continues
to decrease, which causes the decrease of the melting peak. On the other
hand, the amount of nuclei continues to increase causing the increase of
the peak of cold crystallization.
Furthermore, the Fig. 7 can provide information regarding the
Fig. 7. Apparent heat capacity of: (a) PA6, b) PA6þ3%C30B, c) PA6þ5%C30B, enthalpy relaxation of the polymer, which is also characteristic of a
d) PA6þ7%C30Bfrom Flash DSC on heating with 1000 � C/s after cooling with physical aging. It is known that the rate at which you cool through Tg
rates between 50 and 1000 � C/s. has a considerable effect on the resulting ageing kinetics [25,26]. A
slower cooling rate gives more time for the molecular chains to find
energetically favorable orientations. For the PA6, it can be observed a
peak of enthalpic relaxation at the glass transition region, which is more
remarkable for the slowest cooling rate. In factthe PA6 has the time to

104
I. Hammami et al. Materials Chemistry and Physics 232 (2019) 99–108

Fig. 10. The dielectric loss data ε" and the corresponding derivative data
ε00 deriv at 150 � C for the PA6 matrix.

Fig. 9. Frequency dependence of imaginary of the dielectric permittivity (a)


and imaginary (b) for the PA6 matrix.

undergo a physical aging at and below the Tg. The shorter the time (fast
speed) the less physical aging.
For the nanocomposites this phenomenon is less observable when the
C30B content introduced in PA6 increases until disappearing at 7% of
C30B content. This could be related to the increase of RAF with
increasing the filler content. Indeed, the presence the RAF will affect the
behavior of the mobile amorphous phase. Consequently, the phenome­
non of enthalpy relaxation decreases with the increase of the rigid
amorphous phase [27,28].

3.5. Dielectric study

Fig. 8 depicts the temperature dependence of the dielectric permit­


tivity ε0 and the loss factor ε00 , respectively, for the PA6 matrix. As it can
be seen, three relaxations are observed:

- The primary relaxation, α-relaxation, appears for temperatures be­


tween 40 and 140 � C and frequencies between 102 and 106 Hz. It is
associated to the glass transition of PA6 matrix [29–31]. Fig. 11. Imaginary part of the dielectric modulus M00 versus frequency, at 70 � C
(a) and at 150 � C (b), for the PA6 matrix and PA6/C30B nanocomposites.
- The αc-relaxation occurs between 60 and 200 � C (between 10–1 and
104 Hz). According to Laredo et al. [32], in the case of PA6, this kind
of relaxation may be related to the presence of crystalline phases or large increase in both the real and imaginary parts of the dielectric
to the diffusion of carriers such as protons and other impurities. function [33].
- The ionic conduction process observed at high temperature and a
low-frequency ranges, which appears from the increase in the elec­ In the same way, the variations of the loss factor ε ’’ can be presented,
tric charges mobility in the polymer with temperature, resulting a in isothermal conditions (Fig. 9a). These variations show a linear in­
crease of ε ’’ especially for the high temperatures and the low

105
I. Hammami et al. Materials Chemistry and Physics 232 (2019) 99–108

Fig. 13. Relaxation map corresponding to the temperature dependence of the


relaxation time for α, αcand MWS processes for PA6 matrix and PA6/C30B
nanocomposites.

αc relaxation and interfacial polarization appear in the frequency win­


dows at higher temperatures (Fig. 11b at 150 � C).
The effect of clay-nanoparticles on the segmental mobility is further
analyzed by performing an analysis based on fitting appropriate model
functions. The electric modulus spectra were fitted by a sum of
Havriliak-Negami (HN) model function terms of the form. The equation
is an approximation of the HN equation [40]:
2 3
X n
6 Msi M∞i 7
*
M ðωÞ ¼ 4M∞i þ h �αi iβi 5 (10)
1
i¼1 1 þ ð jðωτi Þ
Fig. 12. Imaginary part of the electric modulus M00 versus frequency for the
PA6þ7%C30B nanocomposites. at 70 � C (a) and at 150 � C (b). where ω ¼ 2π f the angular pulsation, τ the relaxation time and α and β
the symmetric and asymmetric broadening factor.
frequencies. In these domains the slope of ε ’’ is close to 1, which in­ To better distinguish the different mechanisms observed, a separa­
dicates the presence of DC conductivity effect [34,35]. tion of overlapping relaxation regions via the deconvolution of the
This phenomenon can be detected by presenting the frequency imaginary part of the electric modulus M00 is presented in the Fig. 12.
dependence of the AC conductivity. As it can be seen from Fig. 9.b, the We note a good agreement between the experimental data and the
presence of a horizontal plateau at low frequencies corresponds to the obtained modelizations.
DC conductivity effect. Moreover, we notice a slight decrease of AC From the data obtained by the adjustment, the maximum relaxation
conductivity at low frequencies. This behavior is related to the elec­ times versus the reciprocal of the temperature can be traced as can be
trodes polarization mechanism [36,37]. seen in Fig. 13.
To further investigate the phenomena that occur at low frequencies The temperature dependence of the relaxation rate for the MWS-
region, we use the first derivative of the real part of the dielectric relaxations were described by the Arrhenius equation [41]:
permittivity [38]: �
Ea

τ ¼ τ0 exp (11)
π ∂ε’ kB T
ε’’deriv ¼ (9)
2 ∂lnω
where Ea is the activation energy, τ 0 the pre-exponential time,
The ε’’deriv method is useful for systems that exhibit low frequency kB ¼ 1.38 � 10–23 J K-1 is the Boltzmann constant and T the absolute
relaxations with appreciable ohmic conductivity, since the values of ε temperature.
0
(ω) are unaffected by ohmic conductivity, and according to Kramers- The α and αc relaxations were better described by a VFT
Kroning relationships its derivative is proportional to the loss factor ε (Vogel–Fulcher–Tammann) equation [42–44]:
’’ (ω). Indeed, in the derivative formalism, the relaxation processes that � �
take place in the dielectric loss variations appear as sharper peaks and τðTÞ ¼ τ0 exp
DT0
(12)
without contribution of conductivity [39]. T T0
In the Fig. 10, we compare the dielectric loss data ε’’ , ε’’deriv at 150 � C.
where τ0 is a pre-exponential factor, T0 is Vogel temperature (ideal glass
From the variations of ε’’deriv , and as it can be seen, we note the presence
transition), it was found to lie between 30 and 70 K below glass transi­
of three relaxation phenomena: relaxation αc, interfacial polarization tion temperature Tg [37]. D, the so-called strength parameter [45], is a
MWS that appears due to the differences in conductivity values of the measure of the fragility factor describing a deviation of the temperature
crystalline and amorphous phases and the electrodes polarization PE. dependence of τ (T) from the Arrhenius behavior.
The frequency dependence of M00 for PA6 and the nanocomposites at The curves in Fig. 11 showed that the relaxation time decreases with
two different temperatures is display in Fig. 11. At low temperature increasing temperature. In addition, one can note that each relaxation
(70 � C) the α relaxation is observed for the four samples (Fig. 11a). The process manifests itself in the same temperature range for the different

106
I. Hammami et al. Materials Chemistry and Physics 232 (2019) 99–108

Table 4 [2] A. Xenopoulos, E.S. Clark, in: M.I. Kohan (Ed.), Nylon Plastics Handbook, Hanser
Calculated activation energies EA and strength parameter D. Publishers, Munich, 1995.
[3] V. Privalko, Y.S. Lipatov, Y.Y. Kercha, Calorimetric study of the phase boundary
Samples EA (eV) D effect on oligo-ethylene glycol adipate (oega) properties, Polym. Sci. 12 (1970)
1520–1529. https://doi.org/10.1016/0032-3950(70)90084-5.
α-Relaxation PA6 – 3.85 [4] Y.S. Lipatov, V. Privalko, Glass transition in filled polymer systems, Polym. Sci. 14
PA6-3%C30B – 5.56 (1972) 1843–1848. https://doi.org/10.1016/0032-3950(72)90286-9.
PA6-5%C30B – 5.64 [5] B. Wunderlich, Reversible crystallization and the rigid–amorphous phase in
PA6-7%C30B – 5.76 semicrystalline macromolecules, Prog. Polym. Sci. 28 (2003) 383–450. https://doi.
αc-Relaxation PA6 – 10.05 org/10.1016/S0079-6700(02)00085-0.
PA6-3%C30B – 10.51 [6] H. Suzuki, J. Grebowicz, B. Wunderlich, Glass transition of poly (oxymethylene),
PA6-5%C30B – 12.61 Polym. Int. 17 (1985) 1–3. https://doi.org/10.1002/pi.4980170101.
PA6-7%C30B – 14.32 [7] J. Soulestin, B.J. Rashmi, S. Bourbigot, M.F. Lacrampe, P. Krawczak, Mechanical
Interfacial polarization MWS PA6-7%C30B 1.209 – and optical properties of polyamide 6/clay nanocomposite cast films: influence of
the degree of exfoliation, Macromol. Mater. Eng. 297 (2012) 444–454. https://doi.
org/10.1002/mame.201100202.
[8] I. Campoy, M. Gomez, C. Marco, Structure and thermal properties of blends of
samples, except the interfacial polarization which appears in the same Nylon 6 and a liquid crystal Copolyester1, Polymer 39 (1998) 6279–6288.
temperature range of αc relaxation. This may reveal a similarity of these https://doi.org/10.1016/S0032-3861(98)00181-5.
two processes, i.e., αc can be considered as an interfacial polarization. [9] N. Bosq, N. l. Guigo, E. Zhuravlev, N. Sbirrazzuoli, Nonisothermal crystallization of
polytetrafluoroethylene in a wide range of cooling rates, J. Phys. Chem. B 117
This result is in good agreement with those found by Laredo et al. [32] (2013) 3407–3415, https://doi.org/10.1021/jp311196g.
who considered that the relaxation αc is related to the interfaces be­ [10] A. Xenopoulos, B. Wunderlich, Thermodynamic properties of liquid and
tween the amorphous and crystalline regions. The resulting parameters semicrystalline linear aliphatic polyamides, J. Polym. Sci. B Polym. Phys. 28
(1990) 2271–2290. https://doi.org/10.1002/polb.1990.090281209.
for the observed relaxation processes are listed in Table 4. We find that [11] B. Wunderlich, The athas database on heat capacities of polymers, Pure Appl.
the values of the fragility parameters D for the relaxation α and αc in­ Chem. 67 (1995) 1019–1026. https://doi.org/10.1351/pac199567061019.
crease considerably as a function of the reinforcement rate indicating [12] H. Rekik, Z. Ghallabi, I. Royaud, M. Arous, G. Seytre, G. Boiteux, A. Kallel,
Dielectric relaxation behaviour in semi-crystalline polyvinylidene fluoride (Pvdf)/
the reinforcing effect of the clays, which limits the mobility of the PA6
Tio2 nanocomposites, Compos. B Eng. 45 (2013) 1199–1206. https://doi.org/10
chains. The interfacial polarization can be assigned to the interfaces .1016/j.compositesb.2012.08.002.
between the filler and the polymer. This can be confirmed by the [13] A. Ladhar, M. Arous, S. Boufi, A. Kallel, Molecular dynamics of poly (styrene-Co-2-
Ethyl hexylacrylate) copolymer/cellulose nanocrystals nanocomposites
quasi-inexistence of this relaxation for the PA6 matrix and the nano­
investigated by dielectric relaxation spectroscopy: effect of the silane content,
composite with low filler content. This relaxation intensity increases J. Mol. Liq. 224 (2016) 515–525. https://doi.org/10.1016/j.molliq.2016.10.031.
with the filler content. [14] S. Alix, N. Follain, N. Tenn, B. Alexandre, S. Bourbigot, J. Soulestin, S. Marais,
For the PA6-7%C30B nanocomposite, the activation energy is very Effect of highly exfoliated and oriented organoclays on the barrier properties of
polyamide 6 based nanocomposites, J. Phys. Chem. C 116 (2012) 4937–4947,
high and comparable to that found by other authors [46]. This important https://doi.org/10.1021/jp2052344.
value reflects the strong interactions between clay nanoparticles and [15] A. Yebra-Rodríguez, P. Alvarez-Lloret, C. Cardell, A.B. Rodríguez-Navarro,
PA6 matrix. Crystalline properties of injection molded polyamide-6 and polyamide-6/
montmorillonite nanocomposites, Appl. Clay Sci. 43 (2009) 91–97. https://doi.
org/10.1016/j.clay.2008.07.010.
4. Conclusions [16] A. Yebra-Rodríguez, P. Alvarez-Lloret, A.B. Rodríguez-Navarro, J.D. Martín-Ramos,
C. Cardell, Thermo-xrd and differential scanning calorimetry to trace epitaxial
crystallization in Pa6/montmorillonite nanocomposites, Mater. Lett. 63 (2009)
In summary, thermal and dielectric properties of PA6/MMT nano­ 1159–1161. https://doi.org/10.1016/j.matlet.2009.02.027.
composites were investigated. DSC analyzes allowed to highlight the [17] W. Steinmann, S. Walter, T. Gries, G. Seide, G. Roth, Modification of the
existence of an immobilized fraction through measurements of the mechanical properties of polyamide 6 multifilaments in high-speed melt spinning
with nano silicates, Textil. Res. J. 82 (2012) 1846–1858. https://doi.org/10.1177
thermal capacity during the glass transition. At higher C30B content
/0040517512456756.
incorporated in the PA6 matrix, this fraction becomes larger. This means [18] C. Schick, in: Study Rigid Amorphous Fraction in Polymer Nano-Composites by
that two different parts of the RAF must be considered by analyzing the Stepscan and Hyperdsc, PerkinElmer Application Note, 2009.
[19] R. Androsch, B. Wunderlich, Reversible crystallization and melting at the lateral
heat capacity of the semi-crystalline nanocomposites, one immobilized
surface of isotactic polypropylene crystals, Macromolecules 34 (2001) 5950–5960,
at the interface between the crystallites and the amorphous polymer and https://doi.org/10.1021/ma010260g.
the other at the interface between the filler and the polymer. [20] A. Wurm, M. Ismail, B. Kretzschmar, D. Pospiech, C. Schick, Retarded
In addition, the Flash DSC analyzes provide new quantitative crystallization in polyamide/layered silicates nanocomposites caused by an
immobilized interphase, Macromolecules 43 (2010) 1480–1487, https://doi.org/
knowledge on the kinetics of crystallization and nucleation of PA6 and 10.1021/ma902175r.
its nanocomposites. We have noticed that crystallization at high cooling [21] J. Bandyopadhyay, S.A. Al-Thabaiti, S.S. Ray, S.N. Basahel, M. Mokhtar, Unique
rates may be related to an increase in nucleation density; while for low cold-crystallization behavior and kinetics of biodegradable poly [(Butylene
succinate)- Co adipate] nanocomposites: a high speed differential scanning
cooling rates the crystal growth rate is faster than the nucleation rate. calorimetry study, Macromol. Mater. Eng. 299 (2014) 939–952. https://doi.
Concerning the dielectric study, different relaxation processes have org/10.1002/mame.201300359.
been detected for PA6 matrix and nanocomposites. These processes were [22] V. Mittal, Nanocomposites with Biodegradable Polymers: Synthesis, Properties,
and Future Perspectives, Oxford University Press, 2011.
mainly associated to the relaxation α associated with the glass transition [23] K.L. Mittal, Polyimides and Other High Temperature Polymers: Synthesis,
of the amorphous mobile phase of polyamide 6 and the relaxation αc Characterization and Applications, CRC Press, 2003.
related to the interfaces created between the amorphous phase and the [24] E. Zhuravlev, J.W. Schmelzer, B. Wunderlich, C. Schick, Kinetics of nucleation and
crystallization in poly (Ɛ-Caprolactone)(Pcl), Polymer 52 (2011) 1983–1997.
crystalline phase. Moreover, the incorporation of MMT into PA6 matrix
https://doi.org/10.1016/j.polymer.2011.03.013.
gave rise to another relaxation associated to MWS interfacial polariza­ [25] J.M. Hutchinson, Physical aging of polymers, Prog. Polym. Sci. 20 (1995) 703–760.
tion. This latter arises from the trapping of electric charges at the in­ [26] R. Surana, A. Pyne, M. Rani, R. Suryanarayanan, Measurement of enthalpic
relaxation by differential scanning calorimetry—effect of experimental conditions,
terfaces between the MMT nanoparticles and the PA6 matrix. This later
Thermochim. Acta 433 (2005) 173–182. https://doi.org/10.1016/j.tca.2005.02.01
is well defined for the nanocomposite with 7% reinforcement. Thus, lead 4.
to the same results obtained by the TMDSC analysis. [27] J.D. Menczel, The rigid amorphous fraction in semicrystalline macromolecules,
J. Therm. Anal. Calorim. 106 (2011) 7–24. https://doi.org/10.1007/s10973-0
11-1496-7.
References [28] J. Menczel, B. Wunderlich, Heat capacity hysteresis of semicrystalline
macromolecular glasses, J. Polym. Sci., Polym. Lett. Ed. 19 (1981) 261–264. htt
[1] W. Sun, L. Li, E.A. Stefanescu, M.R. Kessler, N. Bowler, Dynamics of poly (methyl ps://doi.org/10.1002/pol.1981.130190506.
methacrylate)–montmorillonite nanocomposites: a dielectric study, J. Non-Cryst. [29] E. Laredo, M. Hernandez, Moisture effect on the low-and high-temperature
Solids 410 (2015) 43–50. https://doi.org/10.1016/j.jnoncrysol.2014.11.030. dielectric relaxations in nylon-6, J. Polym. Sci. B Polym. Phys. 35 (1997)

107
I. Hammami et al. Materials Chemistry and Physics 232 (2019) 99–108

2879–2888. https://doi.org/10.1002/(SICI)1099-0488(199712)35:17<2879:: [37] P. Gonon, T.P. Hong, O. Lesaint, S. Bourdelais, H. Debruyne, Influence of high
AIDPOLB11>3.0.CO;2-4. levels of water absorption on the resistivity and dielectric permittivity of epoxy
[30] Y.-H. Lee, A.J. Bur, S.C. Roth, P.R. Start, Accelerated А relaxation dynamics in the composites, Polym. Test. 24 (2005) 799–804. https://doi.org/10.1016/j.polymert
exfoliated nylon-11/clay nanocomposite observed in the melt and semicrystalline esting.2005.02.001.
state by dielectric spectroscopy, Macromolecules 38 (2005) 3828–3837, https:// [38] F. Kremer, A. Sch€ onhals, Broadband Dielectric Spectroscopy, Springer-Verlag,
doi.org/10.1021/ma048052m. Berlin, 2003.
[31] N. Noda, Y.-H. Lee, A.J. Bur, V.M. Prabhu, C.R. Snyder, S.C. Roth, M. McBrearty, [39] M. Wübbenhorst, J. van Turnhout, Analysis of complex dielectric spectra. I. One-
Dielectric properties of Nylon 6/clay nanocomposites from on-line process dimensional derivative techniques and three-dimensional modelling, J. Non-Cryst.
monitoring and off-line measurements, Polymer 46 (2005) 7201–7217. https:// Solids 305 (2002) 40–49. https://doi.org/10.1016/S0022-3093(02)01086-4.
doi.org/10.1016/j.polymer.2005.06.046. [40] G. Floudas, in: K.M. M€ oller (Ed.), 32-Dielectric Spectroscopy, Polymer Science: A
[32] E. Laredo, M. Grimau, F. Sanchez, A. Bello, Water absorption effect on the dynamic Comprehensive Reference, Elsevier, Amsterdam, 2012, pp. 825–845.
properties of Nylon-6 by dielectric spectroscopy, Macromolecules 36 (2003) [41] P.J.W. Debye, Polar Molecules, Chemical Catalog Company, Incorporated, 1929.
9840–9850, https://doi.org/10.1021/ma034954w. [42] H. Vogel, The law of the relation between the viscosity of liquids and the
[33] A. Ladhar, M. Arous, H. Kaddami, M. Raihane, A. Kallel, M. Graça, L. Costa, Ionic temperature, Physik, Zeit 22 (1921) 645–646.
hopping conductivity in potential batteries separator based on natural [43] G.S. Fulcher, Analysis of recent measurements of the viscosity of glasses, J. Am.
rubber–nanocellulose green nanocomposites, J. Mol. Liq. 211 (2015) 792–802. Ceram. Soc. 8 (1925) 339–355. https://doi.org/10.1111/j.1151-2916.1925.tb1
https://doi.org/10.1016/j.molliq.2015.08.014. 6731.x.
[34] A. Kanapitsas, P. Pissis, R. Kotsilkova, Dielectric studies of molecular mobility and [44] G. Tammann, W. Hesse, Die Abh€ angigkeit der Viscosit€
at von der Temperature bie
phase morphology in polymer–layered silicate nanocomposites, J. Non-Cryst. unterkühlten Flüssigkeiten, Z. Anorg. Allg. Chem, 156, https://doi.org/10.100
Solids 305 (2002) 204–211. https://doi.org/10.1016/S0022-3093(02)01107-9. 2/zaac.19261560121.
[35] H. Hammami, M. Arous, M. Lagache, A. Kallel, Study of the interfacial mws [45] C.A. Angell, K.L. Ngai, G.B. McKenna, P.F. McMillan, S.W. Martin, Relaxation in
relaxation by dielectric spectroscopy in unidirectional pzt fibres/epoxy resin glassforming liquids and amorphous solids, J. Appl. Phys. 88 (2000) 3113–3157.
composites, J. Alloy. Comp. 430 (2007) 1–8. https://doi.org/10.1016/j.jallcom. https://doi.org/10.1063/1.1286035.
2006.04.048. [46] E. Nikaj, I. Stevenson-Royaud, G. Seytre, L. David, E. Espuche, Dielectric properties
[36] M. Samet, V. Levchenko, G. Boiteux, G. Seytre, A. Kallel, A. Serghei, Electrode of polyamide 6-montmorillonite nanocomposites, J. Non-Cryst. Solids 356 (2010)
polarization vs. Maxwell-Wagner-Sillars interfacial polarization in dielectric 589–596. https://doi.org/10.1016/j.jnoncrysol.2009.06.047.
spectra of materials: characteristic frequencies and scaling laws, J. Chem. Phys.
142 (2015) 194703. https://doi.org/10.1063/1.4919877.

108

View publication stats

Vous aimerez peut-être aussi