You are on page 1of 12

Materials Science and Engineering A 527 (2010) 812–823

Contents lists available at ScienceDirect

Materials Science and Engineering A


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

Thermal and dynamic mechanical characterization of thermoplastic


polyurethane/organoclay nanocomposites prepared by melt compounding夽
A.K. Barick, D.K. Tripathy ∗
Rubber Technology Centre, Indian Institute of Technology, Kharagpur 721302, West Bengal, India

a r t i c l e i n f o a b s t r a c t

Article history: Thermoplastic polyurethane (TPU) nanocomposites based on organically modified layered silicate (OMLS)
Received 1 May 2009 were prepared by melt intercalation process followed by compression molding. Different percent-
Received in revised form 25 August 2009 age of organoclays was incorporated into the TPU matrix in order to examine the influence of the
Accepted 30 October 2009
nanoscaled fillers on nanostructure morphology and material properties. The microscopic morphol-
ogy of the nanocomposites was evaluated by wide angle X-ray diffraction (WAXD), transmission
electron microscopy (TEM), and atomic force microscopy (AFM). The observation revealed that both
Keywords:
nanoclay–polymer interactions and shear stress developed during melt mixing are responsible for the
Thermoplastic polyurethane
Organoclay
effectively organoclay dispersion in TPU matrix resulting intercalated/exfoliated morphology. Thermal
Nanocomposites stability of the nanocomposites measured by thermogravimetric analysis (TGA) was improved signifi-
Morphology cantly with the addition of nanoclay. The differential scanning calorimetry (DSC) analysis reveals that
Thermal properties melting point of the nanocomposites increased with incorporation of nanoclay. The dynamic mechani-
Dynamic mechanical analysis cal properties of the TPU nanocomposites were analyzed using a dynamic mechanical thermal analyzer
(DMTA), which indicates that the storage modulus (E ), loss modulus (E ), and glass transition temperature
(Tg ) are significantly increased with increasing nanoclay content.
© 2009 Elsevier B.V. All rights reserved.

1. Introduction on TPU nanocomposites prepared by melt intercalation technique


in spite of the obvious advantages of this approach [12–14].
Polymer nanocomposites based on layered silicates have The objectives of the present work are to investigate the mor-
become a prominent area of current research and development phology and thermo-mechanical properties of nanocomposites
in the emerging field of nanoscience and nanotechnology. This based on TPU prepared by melt intercalation technique. The mor-
area has gained momentum with the recognition that exfoliated phology of the nanocomposites was evaluated by WAXD, TEM,
clays could yield significant material properties. The applicability and AFM. TGA was performed under different heating condi-
of organosilicate based polymer nanocomposites to biomedi- tions, including temperature, time, and heating environments to
cal/biotechnological, electrical/electronic/optoelectronic, and fuel examine the effect of organoclay modifier structure and hard seg-
cell field is a rapidly growing area of developments [1]. ment concentration on the thermal decomposition process of TPU
The literature survey have shown that the addition of a very low nanocomposites. The activation energy (Ea ) of the degradation of
percentage (≤10 wt%) of layered silicates to TPU matrix can lead to pure TPU and its nanocomposites in the thermal degradation pro-
a significant enhancement in many important properties, such as cesses has been investigated using kinetic methods. The effects of
tensile strength, modulus and elongation at break [2], tear strength layered silicates on the glass transition temperature, melting point,
[3], flame retardancy [4], gas barrier property [5], electrical prop- and dynamic mechanical properties of TPU nanocomposites have
erty [6], thermal stability [7], heat distortion temperature (HDT) [8], been studied through DSC and DMTA.
biodegradability [9], shape memory behavior [10], and drugs deliv-
ery property [11]. Very few research works has been carried out
2. Materials

Commercial medical grade aliphatic, polyether-based ther-


夽 Contract grant sponsor: Council of Scientific & Industrial Research (CSIR), New moplastic polyurethane (Tecoflex® EG 80A injection grade)
Delhi, INDIA; contract grant number: 22(0410)/06/EMR-II, dated 13 September selected for this work was procured from Lubrizol Advanced
2006.
∗ Corresponding author. Tel.: +91 3222 283196; fax: +91 3222 255303. Materials, ThermedicsTM Inc. Polymer Products, USA. Tecoflex®
E-mail addresses: akbarick@gmail.com (A.K. Barick), EG 80A (around 35% of hard segments) has Shore A Hard-
dkt@rtc.iitkgp.ernet.in (D.K. Tripathy). ness = 72 A, specific gravity = 1.04 and its constituent formulation

0921-5093/$ – see front matter © 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.msea.2009.10.063
A.K. Barick, D.K. Tripathy / Materials Science and Engineering A 527 (2010) 812–823 813

contains methylene bis(cyclohexyl) diisocyanate (HMDI) hard seg- temperature range of –100 to 300 ◦ C at a scan rate of 10 ◦ C/min
ment, polytetramethylene glycol (PTMG) soft segment (molecular in liquid nitrogen atmosphere. Dynamic mechanical properties
weight = 1000 g/mol) and chain extender 1,4 butane diol (BD). Com- of nanocomposites were measured by dynamic mechanical ther-
mercial nanoclay used in this study was Cloisite 15A procured mal analyzer (DMTA model 2980 V1.7B, TA Instrument). The test
from the Southern Clay Products, Inc., USA. The Cloisite 15A is was carried out in tensile mode at a constant frequency of 1 Hz,
an organosilicate clay based on a natural montmorillonite hav- static force of 0.01 N, and amplitude of 10 ␮m in the temperature
ing density = 1.66 g/cm3 , d0 0 1 = 31.5 Å, cation exchange capacity range of −100 to 100 ◦ C and a scan rate 3 ◦ C/min in liquid nitrogen
(CEC) = 1.25 meq/gm with dimethyl dehydrogenated tallow ammo- atmosphere. Strain sweep of the nanocomposites was studied at
nium modifier (2M2HT) of tallow composition as ∼65% C18; ∼30% frequency of 10 Hz at room temperature and frequency sweep was
C16; ∼5% C14. also carried out at strain amplitude of 10 ␮m at room temperature
(30 ◦ C). The size of the rectangular test samples was approximately
3. Manufacturing of nanocomposites 25 mm × 6.5 mm × 2 mm.

Polymer nanocomposites based on TPU and Cloisite 15A was


5. Results and discussion
prepared by melt intercalation technique using Thermo Scientific
HAAKE PolyLab OS Rheomix (Thermo Electron Corporation, MA,
5.1. Morphological analysis
USA) at 185 ◦ C with a rotor speed of 100 rpm and mixing time of
6 min. Varying amount of Cloisite 15A nanoclay (1, 3, 5, 7, and 9 wt%)
5.1.1. Wide angle X-ray diffraction (WAXD)
was added to the TPU matrix. Prior to mixing, nanoclay and TPU
The WAXD lower angular patterns (2 = 2–10◦ ) of pure TPU,
were dried in a vacuum oven at 80 ◦ C for 12 h for the evapora-
organoclay, and TPU nanocomposites are shown in Fig. 1. It is seen
tion of moisture content (if any in supplied materials). TPU and
from Fig. 1 that Cloisite 15A clay shows two prominent peaks at
its nanocomposite sheets were prepared by compression molding
2 values of 2.408◦ and 5.352◦ corresponding to d-spacing of 36.64
(Moore Press, Birmingham, UK) at 180 ◦ C and samples for mechan-
and 16.5 Å of d0 0 1 and d0 0 2 plane reflections, respectively. TPU
ical testing were punched from the compression molded sheets.
nanocomposites prepared by taking 1 and 3 wt% organoclay did
Virgin thermoplastic polyurethane sample was coded as TPU and
not show any distinguishable peaks in WAXD plot, which refers to
TPU nanocomposite samples were designated as TPUXA. Where ‘X’
the disordered exfoliated nanostructure because molecular diffu-
values are 1, 3, 5, 7, and 9 wt% of organoclay loaded to 100 wt%
sion and shear played an important role in the exfoliation process
of TPU matrix and ‘A’ stand for Cloisite 15A types of organically
by melt compounding. The disappearance of diffraction peaks at
modified montmorillonite (OMMT).
low organoclay loading may be explained either by the high dis-
order state or the exfoliation of the silicate layers. The d-spacing
4. Experimental procedure
of clay layers is too high in case of exfoliated nanocomposites
and fail to produce Bragg diffraction peaks as consequence it can
4.1. Microstructure characterization
not be detected by the WAXD methods. The WAXD peak posi-
tion shifts towards left with decrease in peak height and the peak
X-ray diffraction (XRD) was performed with a high resolution
becomes broader in case of TPU nanocomposites above 5 wt% nan-
X-ray diffractometer (model X’Pert PRO, Philips PANalytical B.V.,
oclay loading, which indicate that delamination of silicate layers
Almelo, The Netherlands) with Cu K␣ ( = 1.54 Å) radiation source
increase. The peak position at d0 0 1 = 16.5 Å and d0 0 2 = 36. 64 Å of
operated at voltage of 40 kV and 30 mA electric current. The sam-
the clay is shifted to 19.5 and 40.5 Å, respectively in 9 wt% loaded
ples were scanned from 1 to 10◦ (2) at a scan rate of 2◦ /min. The
nanocomposites. The low peak intensities and range of basal spac-
samples for TEM analysis were prepared using an ultramicrotomy
ings suggest that TPU nanocomposites contain both intercalated
with a Leica Ultracut UCT (Leica Mikrosystems GmbH, Vienna, Aus-
and exfoliated portions of clay platelets. Polymer chains entering
tria). A clean diamond knife with cutting edges of 45◦ was used
the clay layer galleries pushing the platelets apart and causing
to obtain cryosections of 50–70 nm thickness specimens at ambi-
intercalation nanostructure may explain the changes seen in the
ent temperature of −70 ◦ C. The distribution of clay particles into
WAXD. As more polymers enter the galleries, two possible changes
the polymer matrix was studied using a high resolution transmis-
can occur; firstly, the platelets can loose their ordered crystalline
sion electron microscopy (model JEM 2100, JEOL Limited, Tokyo,
Japan) operated at an accelerated voltage of 200 keV. Atomic force
microscopy topographic images were obtained using a NanoScope
IV MultiMode scanning probe microscope (SPM) (Veeco Instru-
ments Inc., Santa Barbara, CA, USA) using the tapping mode probe
with commercial Si3 N4 integral tips with constant amplitude of
40 mV at the resonance frequency of 463 kHz of the cantilever
with a scan rate of 1 Hz and a resolution of 512 samples per
line.

4.2. Thermal measurement

Thermal stability and composition of nanocomposites were


measured by thermogravimetric analysis (TGA model Q50 V6.1
series, TA Instruments, New Castle, DE, USA) from room temper-
ature to 600 ◦ C with a scan rate of 20 ◦ C/min in both nitrogen
(N2 ) and oxygen (O2 ) gas atmosphere. TGA was also carried out
in inert atmosphere at scan rate of 5, 10, 15, and 20 ◦ C/min to
evaluate the thermal degradation kinetics. The glass transition tem-
perature, melting point, and crystallinity of nanocomposites were Fig. 1. WAXD lower angular patterns of pure TPU and TPU/Cloisite 15A nanocom-
evaluated by DSC (model Q100 V8.1 series, TA Instruments) in the posites.
814 A.K. Barick, D.K. Tripathy / Materials Science and Engineering A 527 (2010) 812–823

layered structure and become disordered with the platelets being


no longer parallel. This results in broadening of WAXD peak into
the baseline formation of intercalated disordered morphology. Sec-
ondly, the polymer that enters the galleries pushes the platelets far
apart such that the platelet separation exceeds the sensitivity of
WAXD leading to formation of exfoliation/intercalation nanocom-
posites [15].

5.1.2. Transmission electron microscopy (TEM)


The state of dispersion of the clay was investigated by TEM. The
TEM analysis not only tends to support the findings from WAXD
but also shows that the clay is well dispersed on the nanoscale
in all compositions. The efficiency of the clay in modifying the
properties of the polymer is primarily determined by the degree
of its dispersion. The TEM microphotographs of the cross-sectional
views of Cloisite 15A based TPU nanocomposites containing dif-
ferent wt% of organoclays loading are presented in Fig. 2(a–e).
The dark lines in the microphotographs are the intersections of
about 10 Å thick clay layers and the spaces between the dark lines
are interlayer spaces. TEM microphotograph shows that most clay
layers are parallel to the surface of the films and are dispersed
homogeneously into the polymer matrix, and small clusters of
intercalated silicates and a portion of exfoliated silicates are also
visible in all images. Some clusters or agglomerated clay particles
are observed due to higher clay loaded TPU nanocomposites. It
appear that significantly mixed morphology, i.e. regions of both
partially exfoliated and intercalated nanostructures were achieved
via melt compounding. This result can be attributed to the higher
shear stresses associated with the melt processing, which pro-
mote the fracturing and separation of layered silicate stack into
individual clay units. In case of the Cloisite 15A, there are no func-
tional groups that can cause some specific interaction with the
TPU matrix. The quaternary ammonium ion modifier of Cloisite
15A has larger hydrogenated tallow alkyl chains and also, it has a
large amount of modifier concentration (cation exchange capacity
of 125 meq/100 gm of nanoclay). Cloisite 15A has more modifiers
so that the majority of the outer clay surface may be covered up by
some excess hydrophobic organic modifiers and thus the affinity of
the modified clay to TPU matrix may decreases. Thus Cloisite 15A
clay may be regarded as almost hydrophobic in nature since only
a small portion of neat inorganic clay surface may be exposed to
polymer matrix.

5.1.3. Atomic force microscopy (AFM)


The nanophase surface topography of TPU and its nanocom-
posites is presented in Fig. 3(a–c). It was imaged in real space
using phase and topographical information from tapping mode of
AFM techniques. The surface of the neat TPU was very smooth
whereas the surface roughness was found to increase with addi-
tion of organoclay wt% to TPU matrix. The aggregation of nanoclay
starts rapidly at higher filler loading and makes the nanocompos-
ites surface rough. It can be seen from Fig. 3(a–c) that there are
many nanoclay layers on the surface and it is well distributed
both in soft and hard domain of TPU matrix. Fig. 3(a–c) indi-
cate nearly uniform matrix with some small lighter regions within
the matrix, which may arise from the more elastic soft segment
domains and will also possess some crystallinity whereas the
darker areas are indicative of more particulate hard material region.
It is most likely that the soft segment, which is more elastic Fig. 2. TEM images of TPU nanocomposites (a) 1 wt%, (b) 3 wt%, (c) 5 wt%, (d) 7 wt%,
compared to the hard segment, will protrude to higher eleva- and (e) 9 wt% of Cloisite 15A nanoclays loadings at 10,000× (left) and 25,000× (right)
tion. The aggregates of hard domains with the spherical structures magnifications.
were observed whose size is about 800 nm in the absence of clay.
When the clay was incorporated, the size of the aggregates of
hard domain was reduced to approximately 500 nm. The reduc-
tion of the size of the aggregates with the addition of clay indicates
that introduction of clay affects the aggregation behavior of hard
A.K. Barick, D.K. Tripathy / Materials Science and Engineering A 527 (2010) 812–823 815

Fig. 3. AFM topography represents height image (left) and 3D image (right) of (a) pure TPU and TPU nanocomposites with (b) 3 wt% and (c) 5 wt% of Cloisite 15A nanoclays
loadings.

domains. Previous study have suggested that microphase separa- However, the factors to decide the final morphology and the size
tion in TPU can produce hard microdomains, but that can approach of the aggregates are still not clear. Although the phase separa-
each other and coalesce, which is termed as “secondary order- tion is severely disrupted, the sample surface is still covered with
ing” [16]. The aggregation process occurs because there is strong a thin layer of smooth soft phase. It is observed that even this
interaction due to the hydrogen bonding among hard domains surface layer was disrupted by the addition of nanoclay in the
of TPU. Also, the polyol as soft segments is a flexible long chain. nanocomposites.
816 A.K. Barick, D.K. Tripathy / Materials Science and Engineering A 527 (2010) 812–823

Fig. 4. TGA and DTG thermograms of Cloisite 15A nanoclays in (a) nitrogen and (b)
oxygen gas atmosphere. Fig. 5. TGA and DTG thermograms of pure TPU and TPU/Cloisite 15A nanocompos-
ites in (a) nitrogen and (b) oxygen gas atmosphere.

5.2. Thermal properties carried out by TGA under nitrogen gas atmosphere as shown in
Fig. 5(a). It is found that all specimens displayed two degrada-
5.2.1. Thermogravimetric analysis (TGA) tion processes from room temperature to 600 ◦ C at a ramp rate
5.2.1.1. Thermal stability study. TGA and DTG (derivative thermo- of 20 ◦ C/min. To study the degradation thoroughly, the onset tem-
gravimetry) curves of Cloisite 15A nanoclay in nitrogen and oxygen perature (Td ) of the two degradation processes and the respective
atmosphere are shown in Fig. 4(a) and (b), respectively. Ther- temperature of maximum degradation rate (DTGmax ) are con-
mal decomposition of the alkyl quaternary ammonium modified sidered as shown in Fig. 5(a). It can be seen that, compared to
montmorillonite has taken place in two regions. The first stage TPU matrix, the first onset temperature (Td1 ) of the nanocom-
degradation at around 200 ◦ C refers to the evolution of interlayer posites with 1–9 wt% of clay is found to be higher and DTG
absorbed water, CO2 , and long chain alkyl fragments whereas sec- thermograms show the first maximum degradation rate peak
ond stage of thermal degradation was shown in between 200 and (DTG1max) . The second onset temperature (Td2 ) and second max-
500 ◦ C, which is due to the evolution of water from dehydroxylation imum degradation rate peak (DTG2max ) of all the nanocomposites
of CO2 and organic substances [17]. For the organic montmoril- show significant increase compared to the TPU matrix. The degra-
lonite (Cloisite 15A), the initial mass loss is dominated by water. dation of the TPU molecular backbone is dominant at DTG2max . It
For Cloisite group of clays, release of water begins around 40 ◦ C can also be seen that the nanocomposites with 5 wt% nanoclay
and continues till about 200 ◦ C. Weakly bound, physically absorbed exhibit maximum DTG peak degradation temperature (T50wt% )
and free water pockets within the aggregate structure evolve at value than the other nanocomposites prepared from 1, 3, 7, and
the lowest temperatures, whereas water within the interlayer and 9 wt% clay, which tally with the WAXD result. The end degrada-
strongly bonded water of hydration (Na+ and alkyl ammonium tion temperature (T95wt% ) and residue obtained at 500 ◦ C (r500wt% )
ions (R4 N+ )) evolve at progressively higher temperatures. At tem- increases with increasing clay percentage, which confirms the
peratures between 200 and 500 ◦ C, the organic constituent in the compatibility of clay with the TPU matrix. It is reported that
organically modified layered silicate begins to decompose. Three the first degradation process corresponds to the release of the
overlapping DTG peaks indicate that the release of organic sub- smaller molecules or unstable side chains, which usually degrade
stances is staged and arises from different mechanisms. In general, at lower temperature. The introduction of OMMT layers can sub-
all of the organically modified montmorillonites decompose in a stantially improve the thermal properties of the polymer matrix.
similar fashion. They exhibit three to four DTG peaks, a lower tem- The results show that the degradation rates of the nanocomposites
perature event resulting in a sharp DTG peak around 250 ◦ C, and become significantly slower compared to that of TPU, indicat-
two to three closely overlapping higher temperature events pro- ing improvement of thermal stability of TPU. This is because the
ducing a broad DTG peak between 300 and 400 ◦ C. inorganic material can prevent the heat to expand quickly and
TGA and DTG thermograms of the pristine TPU and its nanocom- limits further degradation [18]. The OMMT itself contains some
posites prepared from organically modified Cloisite 15A clay was smaller molecules that release at lower temperature [19], more of
A.K. Barick, D.K. Tripathy / Materials Science and Engineering A 527 (2010) 812–823 817

these molecules are likely to impair the thermal stability of the


nanocomposites.

5.2.1.2. Thermo-oxidative degradation study. TGA and DTG ther-


mograph of pristine TPU and its nanocomposites in oxygen
atmosphere is shown in Fig. 5(b). Thermo-oxidative degradation of
pristine TPU and its nanocomposites were studied from room tem-
perature to 600 ◦ C at scan rate of 20 ◦ C/min. Thermogram of neat
TPU shows four stages of degradation in the presence of oxygen
whereas two steps in case of nitrogen atmosphere. The significant
low temperature degradation peak is associated with the splitting
of urethane linkage to form alcohols and isocyanates (first stage).
The second stage of TPU decomposition is due to oxidative decom-
position of the polyol chain in oxygen gas to yield volatile gaseous
products such as aldehydes, ketones, carbon dioxide, and water.
This sufficiently retards the formation of metastable products such
as isocyanurate, carbodiimide, substituted urea, and stable isocya-
nurate that causes the reduction of DTG peak area associated with Fig. 6. Isothermal TGA curves of pure TPU and TPU/Cloisite 15A nanocomposites at
their decomposition. Complete volatilization of resulting alcohols 300 ◦ C for 30 min in nitrogen gas atmosphere.
and isocyanates chain fragments is prevented by dimerisation of
isocyanates to carbodiimides, which react with the alcohol groups posites (1–9 wt%) are much higher compared to the pure TPU. The
to give relatively stable substituted urea that decompose in the TPU nanocomposites show a delayed decomposition compared
third stage. Despite the faster low temperature thermal decompo- to the pristine TPU due to the homogeneous distribution of the
sition, TGA profile of TPU in oxygen shows more quantity of charred silicate sheets which increase the total path of the volatile gases
residue at around 400 ◦ C as compared to the degradation profile by acting as a gas barrier, preventing small gaseous molecules
TPU in nitrogen condition. This is probably due to formation of from permeating out of the TPU nanocomposites during thermal
a metastable oxidation intermediate product in the temperature treatment. At 300 ◦ C in nitrogen atmosphere, the decomposition
range of 400–450 ◦ C. The decomposition of this thermally oxidized rate of the TPU nanocomposites is lower than that of the pure TPU
residual chars occurred at 500 ◦ C in the DTG obtained under oxygen and the weight loss is slowed down in nanocomposites compared
atmosphere [4]. to the pure TPU. TPU nanocomposites filled with 5 wt% Cloisite 15A
As shown in Fig. 5(b), DTG curve for nanocomposites indicate have greater thermal resistance than the other nanocomposite
completely different thermal decomposition behavior under oxy- systems. However, for 9 wt% filler loading, thermal stability of
gen atmosphere than that of TPU counterpart. The low temperature the nanocomposites decreases slightly but remains higher with
degradation of polyurethane chain completely diminishes by the respect to neat TPU. Such a behavior is probably associated with the
addition of organoclay and a demarkable sharp DTG peak is clearly morphological changes in proportion with the exfoliated and inter-
visible at around 430 ◦ C associated with the decomposition of both calated species in the nanocomposites. At lower nanoclay loading
substituted urea and isocyanurate stabilized char residue liberated (1 wt%), exfoliation dominates, whereas the amount of exfoliated
at low temperature. The thermal oxidation of TPU nanocomposites nanolayer is not sufficient enough to exhibit strong barrier effect
result at higher temperature because of the well-known ‘Labyrinth and enhance the thermal stability of the TPU nanocomposites.
effect’ of the OMMT platelets dispersed on the nanometer scale in When Cloisite 15A concentration is increased to 3 and 5 wt%,
the TPU matrix caused by shielding of the material from oxygen relatively more Intercalation/exfoliated layers are formed thereby
via excellent barrier and strong interaction between the OMMT leading to excellent barrier effect and higher thermal stability. At
platelets and TPU chains. This enhancement of thermal stability higher filler loading (above 5 wt%), the formation of aggregated
exhibited by nanoclay is directly related to the higher gallery spac- Cloisite 15A tactoid layers dominates which probably does not
ing between the clay platelets and the more uniform dispersion of contribute to thermal stability of TPU matrix.
agglomerated tactoids in TPU nanocomposites.
5.2.1.4. Thermal degradation kinetics. Fig. 7(a–c) shows the linear
5.2.1.3. Isothermal thermogravimetric analysis. The catalytic 2 ) versus 1/T , and ln(ˇ/T 2 )
plots of log ˇ against 1/T50 , ln(ˇ/T50 50 p
decomposition and barrier effect of the silicate layers are clearly versus 1/Tp for pristine TPU and various compositions of TPU
evident from the isothermal thermogravimetric curves as shown nanocomposites from the Flynn–Wall–Ozawa (F–W–O), modified
in Fig. 6 for neat TPU and Cloisite 15A based TPU nanocomposites Coats–Redfern, and Kissinger methods, respectively. It is seen
at 300 ◦ C for 30 min in nitrogen atmosphere. The initial decompo- from Fig. 7(a–c) that the lines are nearly parallel, which indicates
sition of the TPU is accompanied by scissioning of the urethane approximate activation energies at different conversions and con-
linkage. The thermal stability of the TPU/Cloisite 15A nanocom- sequently implies the possibility of single reaction mechanism or

Table 1
Apparent activation energy (Ea ) of neat TPU and TPU/Cloisite 15A nanocomposites evaluated by three kinetic methods.

Kinetic methods Kissinger method Flynn–Wall–Ozawa (F–W–O) method Modified Coats–Redfern method
2 2
Sample code Activation energy (kJ/mol) R Activation energy (kJ/mol) R Activation energy (kJ/mol) R2

TPU 127.50 0.9992 110.96 0.9999 103.40 0.9991


TPU1A 162.63 0.9989 163.47 0.9999 161.99 0.9998
TPU3A 156.42 0.9959 158.07 0.9979 144.61 0.9949
TPU5A 146.30 0.9963 149.37 0.9979 145.74 0.9976
TPU7A 164.73 0.9940 165.48 0.9936 162.64 0.9926
TPU9A 169.20 0.9948 168.62 0.9955 167.28 0.9950
818 A.K. Barick, D.K. Tripathy / Materials Science and Engineering A 527 (2010) 812–823

Fig. 8. DSC thermograms in (a) 1st heating scan and (b) 2nd heating scan of pure
TPU and TPU/Cloisite 15A nanocomposites.

conversional methods. Table 1 shows that the activation energy


values (Ea ) evaluated using the above methods agree on the similar
change trend that the activation energies of the thermal degrada-
tion for the TPU nanocomposite are much higher than that of the
pure TPU matrix. The phenomenon indicates an important role of
clay in improving the thermal stability of the polymer matrix. This
increasing tendency coincides with the thermal analysis results
that the polymer/nanoclay nanocomposite has a higher thermal
stability. The nano-dispersed lamellae of clay (exfoliation or inter-
calation) in polymer matrix may change the decomposition process
of polymer since the dispersed silicate layers act as thermal hinder
in polymer matrix. The nano-dispersed silicate layers slowdown
Fig. 7. Linear plots of (a) Flynn–Wall–Ozawa (F–W–O) method, (b) modified the decomposition rate and increase the temperature of degrada-
Coats–Redfern method, and (c) Kissinger method for pure TPU and TPU/Cloisite 15A tion (especially measures at the point of 50% mass loss) by acting
nanocomposites. as an excellent thermal insulator and mass transport barrier [20].
The activation energy alone may not provide an integral prediction
the unification of multiple reaction mechanisms at various con- or modeling for thermal decomposition process of nanocompos-
version rates. Table 1 shows the apparent activation energy values ites. As the apparent activation energy calculated from different
calculated by Kissinger, Flynn–Wall–Ozawa (F–W–O), and modi- methods related to the energy barrier for nanocomposites degrada-
fied Coats–Redfern method for neat TPU and TPU nanocomposites. tion mechanism, which provide the information about the critical
Kissinger’s method is a special case for determining activation energy needed to initiate a combustion process. The combustion
energy (Ea ) from a certain conversion rate at the temperature of activation energy obtained in this study can help to understand
DTG peak, which may not display overall trend of activation energy thermal decomposition and stability of organoclay reinforced poly-
(Ea ). It is observed that the activation energy (Ea ) was stable in mer nanocomposites.
a conversion range of 5–60 wt% according to the aforementioned
results by iso-conversional methods (Flynn–Wall–Ozawa (F–W–O) 5.2.2. Differential scanning calorimetry (DSC)
and modified Coats–Redfern methods). That is why the results DSC thermogram curves of the pure TPU and TPU nanocompos-
from Kissinger’s method are meaningful for determining activation ites are shown in Fig. 8(a) and (b). Summary of the DSC results
energy (Ea ) in this case. It is observed from Table 1 that the values such as glass transition temperature (Tg(soft) ), soft and hard seg-
obtained by Kissinger’s method are nearly equal to that of the iso- ment melting points (Tm(soft) , Tm(hard) ), heat of fusion of soft and
A.K. Barick, D.K. Tripathy / Materials Science and Engineering A 527 (2010) 812–823 819

Table 2
Differential scanning calorimetry (DSC) results of neat TPU and TPU/Cloisite 15A nanocomposites.

Sample codes Tg(soft) (◦ C) Cp (W/g) Tm(soft) (◦ C) Tm(hard) (◦ C) Hm(soft) (J/g) Hm(hard) (J/g) Xc,s (%) Xc,h (%)

TPU −71 0.52 21.91 94.50 14.42 9.91 8.39 5.76


TPU1A −71 0.47 22.72 97.40 33.91 15.29 19.73 8.90
TPU3A −69 0.46 28.78 97.00 44.86 15.14 26.10 8.80
TPU5A −70 0.45 28.50 97.84 43.08 15.75 25.06 9.16
TPU7A −70 0.47 28.82 95.65 42.44 16.67 24.70 9.70
TPU9A −70 0.48 28.17 95.75 41.71 10.49 24.26 6.10

hard segment melting (Hm(soft) , Hm(hard) ), percentage soft and the transition temperature. As a result, the crystallinity due to the
hard segment crystallinity (Xc,s , Xc,h ), and change in heat capacity soft segments is one of the necessary conditions to show shape
of soft segment at Tg (Cp(soft) ) are given in Table 2. As the struc- memory behavior [10].
ture of TPU includes two segments such as hard and soft, generally The pristine TPU has a strong soft segment glass transition at
two melting temperatures are expected. The DSC curve shows two approximately −71 ◦ C. The glass transition temperature of the of
very small peaks from DSC first heating scan and these endothermic the soft segment of TPU nanocomposite is at around −69 ◦ C, which
peaks are associated with the melting temperature of the soft and is slightly higher than that of neat TPU because DSC cannot probe
hard segment domains present in TPU matrix. The virgin TPU also the glass transition temperature of polymer chains close to the lay-
gives rise to a higher temperature small melting endotherm peak ered silicates. This may be attributed to the confined thin polymers
at about 253 ◦ C that may be attributed to the predominantly dis- in between interlayer spacing of organoclay tending to lower the Tg
ruption of specific hard segment structures and length populations. and enhancing the segmental dynamics of the polymer chains. The
The limitation of specific hard segment length populations to the glass transition temperature of TPU nanocomposites is referred to
assigned endotherm is because of the heterogeneity nature of the the layered silicates free soft segment phases of TPU matrix and it is
TPU. The small fraction of hard segments melting above 200 ◦ C pro- found to be same as that of the pure TPU. The effect of small amounts
vided an strong evidence to suggests that the hard segments were of dispersed silicate layers on the free volume of TPU is insignif-
able to form crystal structures with long range order as a result well icant to influence the glass transition temperature of pure TPU,
phase separated hard domains arise in TPU matrix. It is suggested which may be attributed to the interactions between the organic
that in case of ether based TPU this endotherm could be attributed and inorganic phases. These interactions enhance the rigidity of the
to the disordering of nonuniformly packed hard segment hydrogen soft segments and limit the movement of the soft segments. There-
bonding between amine (–NH) and carbonyl (>C O) groups in the fore, the introduction of nanoparticles results in a slight increase
interfacial region between the soft and hard phases. The absence in phase transition temperatures of the soft segments. The amor-
of this endotherm in the melt compounded TPU nanocomposites phous state environment is clearly influenced by the concentration
provides evidence of more narrower and ordered interfacial region of soft and hard segments and also by the clay content. The glass
resulting improved phase separation among hard and soft phases transition of the hard segment phase of the pristine TPU was not
[21]. The peak associated with the melting temperature of the hard detectable in the DSC measurements because of low hard segment
segment of TPU nanocomposites is observed at about 100 ◦ C in the content and a rather small heat capacity difference at its Tg [26].
thermogram plot and peak positions were not appreciably affected As shown in Table 2, the change in the heat capacity associ-
by the nanoclay particles that may be attributed to the disruption ated with glass transition temperature (Cp ) of the nanocomposite
and distribution of various degrees of short range hard segment decreases as the amount of the modified clay increases from 1 to
orders. The small endotherm peak for hard segment melting may 9 wt%. The Cp , which is the magnitude of the transition at the
be due to the inactive movement of the hard segment, which has Tg , is related to the amount of mobility of the polymer chains. The
a small heat capacity change [22] and to the widely uniform dis- partial exfoliated/intercalated clay layers may provide some steric
persion of hard segment microdomains within the TPU matrix [23]. hindrance to the TPU chains [27]. However, the study of the glass
The melting transition temperatures of the soft segments from DSC transition of either the soft or the hard segments is better per-
second heating scan for pristine TPU and its nanocomposites are formed by dynamic mechanical analysis and will be discussed in
found to be around 22 and 30 ◦ C, respectively, which indicate that the following section.
crystallization of polytetramethylene glycol (PTMG) is taken place
in both neat TPU and TPU nanocomposites. This suggests that the 5.2.3. Dynamic mechanical thermal analysis (DMTA)
nanoclay affects the crystal structure of the soft segments and tends 5.2.3.1. Effect of temperature. DMTA results are expressed by three
to increase the melting temperature of the same with the addition main parameters: (a) the elastic response to the deformation
of organoclay due to their nucleating effect [24]. This increase in termed as the storage modulus (E ), (b) the plastic response to
the soft segment melting temperature and enthalpy change can the deformation corresponding to the loss modulus (E ), and (c)
be attributed to the effects produced by the rigid nanoclay in the the tan ı (the ratio of E /E ), a measure of the damping behavior
soft segment domains resulting in the increase in degree of crys- responsible for determining the occurrence of molecular mobil-
tallinity of the TPU soft segments. The value of enthalpy associated ity transitions such as the glass transition temperature (Tg ). DMTA
with the melting in each case is also too small, which indicates that analysis was carried out to monitor the temperature dependence
only a small fraction of urethane linkages formed phase separated of storage modulus (E ), loss tangent (tan ı), and loss modulus (E )
domains. Moreover, soft and hard segment crystallinity values are of neat TPU and its nanocomposites based on Cloisite 15A as shown
found to increase with nanoclay concentrations up to 5 wt% beyond in Fig. 9(a)–(c), respectively. The significant enhancement of stor-
which it decreases. This behavior is attributed to the reduction in age modulus (E ) in the investigated temperature range for all TPU
the effective surface area caused by filler platelet agglomeration nanocomposites over that of TPU matrix indicates that nanoclay
at high organoclay contents [25]. The soft segment is especially has a strong effect on the elastic properties of the neat TPU due the
important because the transition temperature of sample originates restricted movement of TPU chains resulted by the dispersed clay
from the melting of soft segments and the corresponding crys- layers. The significant increase in storage modulus (E ) is marked
tallinity affects a shape recovery when the sample is heated above above the glass transition region because the relative reinforcing
820 A.K. Barick, D.K. Tripathy / Materials Science and Engineering A 527 (2010) 812–823

Fig. 10. DMA curves of storage modulus (E ) of pure TPU and TPU/Cloisite 15A
nanocomposites as a function of strain amplitude.

hindering molecular motion exhibited due to the nanoscopic dis-


tribution of delaminated organosilicate clay layers. Furthermore,
it is found that the height of the tan ı decreases and the curve
broadens with OMMT content in the vicinity of Tg . Since the three
parameters storage modulus (E ), loss modulus (E ), and tan ı are
related by tan ı = E /E , the low tan ı values for the nanocomposites
are mainly because of larger changes of the storage modulus (E ) in
the Tg region than that of the loss modulus (E ).

5.2.3.2. Effect of strain. The dependence of storage modulus (E ) on


strain for TPU nanocomposites at room temperature is shown in
Fig. 10. The plots of storage modulus (E ) with double strain ampli-
tude (DSA) show almost similar trend for different filler loadings.
A decrease in modulus is observed for the neat TPU, which is pre-
sumably associated with the orientation of hard segment domains
of the TPU matrix [29]. Low strain causes a slight modulus vari-
ation indicating a linear response of materials, which drastically
drops after certain strain amplitude and finally converges towards
a single point. The linear viscoelastic range up to which the stor-
age modulus (E ) is independent of strain amplitude decreases
with the filler loadings. The E for the nanoclay filled TPU shows
a non-linear behavior with the dynamic strain amplitude, which is
known as ‘Payne effect’. As strain increases, the dynamic behavior
is dominated by the Payne effect resulting in a considerable reduc-
tion in the storage modulus (E ), which is due to the existence of
filler–filler networks in the polymer matrix above the percolation
threshold, i.e. above 7 wt% nanoclay loading [30]. The initial mod-
ulus value in the nanocomposites is the result of important factors,
Fig. 9. DMA curves of (a) storage modulus (E ), (b) dissipation factor (tan ı), and (c) such as types of polymer matrix and fillers, amount of fillers in the
loss modulus (E ) of pure TPU and TPU/Cloisite 15A nanocomposites as a function nanocomposites, stiffness and rigidity of the filler systems in poly-
of temperature. mer matrix and the force of attraction among filler and polymer.
All the compositions show highest storage modulus (E ) value at
effect of rigid nanoclay on TPU matrix is enhanced over this tem- lower strains irrespective of weight percentage of nanoclay. This
perature change. The result is also supported by the contribution can be explained by the fact that the secondary structure of the
of hydrodynamic reinforcing effect produced due to the dispersion system gradually breaks down with increase of strain amplitude
of the nanosized filler in the TPU matrix, which is well controlled and results in dynamic modulus reduction. This effect is significant
by the shape factor and filler volume fraction of the organoclay at higher filler loadings as compare to the lower clay loadings. At
[28]. A sharp fall in storage modulus (E ) after −75 ◦ C confirms the lower strains, the three-dimensional filler–filler and filler–polymer
Tg region as modulus decreases drastically in going from the glassy conformation acts as a rigid unit against the imposed strain leading
state to the rubbery state. A drop of modulus has taken place around to higher modulus. The strain input associated at low strains is not
25 ◦ C in case of the neat TPU matrix because of the disordering of sufficient to cause any significant change in the network structure.
the PTMG microcrystalline domains of the soft segments. The TPU Fig. 11(a) shows effect of nanoclay loadings on storage modulus
matrix that is free from inorganic component shows rapid mobi- (E ) of nanocomposites at different strain amplitudes. The storage
lization of the amorphous soft domains. The tan ı curves are shifted modulus (E ) increases monotonically with filler content due to the
significantly to higher temperatures on addition of OMMT, which high stiffness and rigidity of the finely dispersed nanoclays in TPU
means that Tg has shifted to higher temperatures because of the matrix as well as existence of strong interactions among filler and
A.K. Barick, D.K. Tripathy / Materials Science and Engineering A 527 (2010) 812–823 821

polymer. Further layered filler decreases the entanglement den-


sity of amorphous zone and the network structure becomes weak,
which does not resist large deformation like its pristine counter-
part. On application of dynamic strains amplitude to the clay filled
system, the molecules having smaller chain lengths in soft seg-
ments between the densely packed network points get directional
orientation and form crystallites that causes strain induced crys-
tallization whereas the molecules of much longer chain lengths
remain in random coil configuration.
The variation of loss modulus (E ) with filler loadings at different
double strain amplitudes is shown in Fig. 11(b). The plot indicates
that loss modulus (E ) is gradually enhanced with strain amplitude
at different wt% of nanoclay. At lower filler loading (up to 3 wt%),
the modulus value does not change appreciably with strain because
of the rigidity and high stiffness of the material. However, at higher
filler loading there is a steady increase of modulus value with nan-
oclay content, which may be attributed to the interaction between
filler and TPU matrix. This result can also be explained on the basis
of secondary structure of the agglomerates. The breakdown of nan-
oclay particles takes place with the increase in strain and hence
loss modulus (E ) possesses a maximum value where most of the
agglomerates have broken down [31].
Fig. 11(c) shows the variation in tan ı as a function of increasing
concentrations of filler loadings with different strain amplitudes
in nanoclay reinforced TPU nanocomposites. At low filler load-
ing (up to 1 wt%) there is no appreciable change in tan ı with
strain amplitude. Beyond 1 wt% filler loading tan ı value shows an
increasing trend, which is more pronounced at high filler loading.
This can be explained on the basis of polymer–filler interaction,
the desorption and reabsorption of soft segments surrounding
the filler particles or breaking and reforming of effective network
structure in the polymer matrix forming transition zone between
the polymer–filler network and the bulk matrix. Another factor
responsible for non-linearity in viscoelastic properties in the filled
polymer systems is the sensitivity of the relaxation of polymer
chains to the polymer–filler network structure, because the charac-
teristic long chain molecules are sensitive to the local environment.

5.2.3.3. Effect of frequency. The frequency dependence of storage


(E ), loss modulus (E ), and damping factor (tan ı) of TPU/organoclay
nanocomposites and their pristine counterpart at room tempera-
ture are shown in Fig. 12(a–c). The storage and loss moduli of the
nanocomposites are substantially higher than that of their pristine
counterpart for all frequency because of the strong dimensional sta-
bility and rigidity of the nanoscopically dispersed nanoclay in TPU
matrix. The nanocomposites containing higher filler concentration
show higher modulus at all measured frequencies as organoclay
may restrain the relaxation phenomenon of soft and hard seg-
ments in TPU chains as a result modulus increases. Both the storage
Fig. 11. DMA curves of (a) storage modulus (E ), (b) loss modulus (E ), and (c) dis- modulus (E ) and loss modulus (E ) increases with the increase in
sipation factor (tan ı) of pure TPU and TPU nanocomposites as a function of Cloisite frequency for all the samples. Also the rate of increase of storage
15A nanoclay loadings at different strain amplitudes.
modulus (E ) and loss modulus (E ) values at higher filler loading
is more pronounced than that exhibited at lower filler loading for
polymer matrix. At low strain amplitude the storage modulus (E ) all frequency. The tan ı value is found to increase with increase
exhibits higher values, which shows a decreasing trend with the in filler loadings at all frequencies as a result of the stiffening
increase in strain. This is due to the fact that the material at lower effect produced by the intercalated/exfoliated nanoclays on TPU
strain gets sufficient time to undergo molecular rearrangement in matrix.
order to minimize the localized stresses than at higher strain. The
decrease in the storage modulus (E ) with the increase in strain 5.2.3.4. Cole–Cole plot. The phase behavior and structural changes
amplitude in case of filled system is greatly affected by disruption of taking place after addition of nanofiller to polymeric systems
the continuous network structure of filler that interpenetrates into can be studied using the Cole–Cole plot, where the loss mod-
the polymer matrix. The network structure of the sample emanates ulus (E ) are plotted as a function of the storage modulus (E ).
from the presence of entanglements of amorphous soft segment The dynamic mechanical properties when examined as a func-
and ordered crystalline hard segment domain of TPU matrix. Net- tion of temperature, frequency, and strain are represented on the
work structure of the unfilled matrix sustains large deformation Cole–Cole complex plane. Fig. 13(a) and (b) shows the Cole–Cole
while formation of new nanostructure occurs in nanoclay filled plot from temperature and strain sweep, respectively. The nature
822 A.K. Barick, D.K. Tripathy / Materials Science and Engineering A 527 (2010) 812–823

Fig. 13. Cole–Cole plots (plot of E versus E ) of the pure TPU and TPU/Cloisite
15A nanocomposites as a function of (a) temperature, (b) strain amplitude, and (c)
frequency.

Fig. 12. DMA curves of (a) storage modulus (E ), (b) loss modulus (E ), and (c) dis-
sipation factor (tan ı) of pure TPU and TPU nanocomposites as a function of Cloisite
sweep are shown in Fig. 13(c). The storage modulus (E ) ver-
15A nanoclay loadings at different frequencies.
sus loss modulus (E ) plot is almost a straight line having the
slope of about 2 for both virgin TPU and TPU nanocomposites
because clay loadings do not affect the morphological state of
of the plot is reported to be indicative of the nature of the poly-
nanocomposites.
mer filler system. A semicircular diagram is reported in case
of homogeneous polymeric systems whereas two-phase systems
show two modified semicircles [32]. The Cole–Cole diagrams 6. Conclusions
presented in Fig. 13(a) and (b) are imperfect semicircles plot indi-
cating heterogeneity of the system. However, the shape of the In this research work, the TPU nanocomposites were suc-
curve points towards the relatively good clay polymer interac- cessfully prepared via melt intercalation technique. Morphology,
tions. Therefore, it can be concluded that TPU and TPU/OMMT thermal stability, and mechanical properties of nanocomposites
nanocomposites systems maintain the homogeneities for differ- were evaluated controlling the exfoliation/intercalation of organ-
ent amount of organoclay loadings. Cole–Cole plots from frequency oclay in TPU matrix. Morphological analysis by WAXD, TEM, and
A.K. Barick, D.K. Tripathy / Materials Science and Engineering A 527 (2010) 812–823 823

AFM reveals that organoclays are more prone towards the soft seg- References
ment domains. TEM microphotographs and WAXD patterns shows
that Cloisite 15A filled nanocomposites display the partially exfo- [1] S. Pavlidou, C.D. Papaspyrides, Prog. Polym. Sci. 33 (2008) 1119–1198.
[2] H. Chen, M. Zheng, H. Sun, Q. Jia, Mater. Sci. Eng. A 445–446 (2007) 725–730.
liated/intercalated structures as revealed by XRD and TEM. AFM [3] A. Pattanayak, S.C. Jana, Polymer 46 (2005) 3394–3406.
topographic images show that clay layers are significantly influ- [4] M. Berta, C. Lindsay, G. Pans, G. Camino, Polym. Degrad. Stab. 91 (2006)
enced the soft and hard segment ordering and distribution in 1179–1191.
[5] J.M. Herrera-Alonso, E. Marand, J.C. Little, S.S. Cox, J. Membr. Sci. 337 (2009)
TPU matrix. According to TGA results the thermal stability of TPU 208–214.
nanocomposites shift to higher temperature compared to pure TPU [6] H.T. Lee, L.H. Lin, Macromolecules 39 (2006) 6133–6141.
because of the barrier effect produced by the nanofillers. Ther- [7] A. Rehab, N. Salahuddin, Mater. Sci. Eng. A 399 (2005) 368–376.
[8] S.M. Liff, N. Kumar, G.H. McKinley, Nat. Mater. 6 (2007) 76–83.
mal degradation kinetics based on the three well-known methods [9] E.H. Jeong, J. Yang, H.S. Lee, S.W. Seo, D.H. Baik, J. Kim, J.H. Youk, J. Appl. Polym.
was thoroughly analyzed, which shows that activation energy of Sci. 107 (2008) 803–809.
combustion is prominently increases with clay loading. TGA mea- [10] I.S. Gunes, F. Cao, S.C. Jana, Polymer 49 (2008) 2223–2234.
[11] G.R.D. Silva, E. Ayres, R.L. Orefice, S.A.L. Moura, D.C. Cara, A.D.S. Cunha Jr., J. Drug
surements in oxidative atmosphere proved to be very sensitive
Target 17 (2009) 374–383.
to the type of organoclay used. The thermal stability was con- [12] F. Chavarria, D.R. Paul, Polymer 47 (2006) 7760–7773.
trolled essentially by the dispersion state of the clay. DSC shows [13] B.C. Chun, T.K. Cho, M.H. Chong, Y.C. Chung, J. Chen, D. Martin, R.C. Cieslinski, J.
soft and hard segment melting points as well as Tg of soft segments Appl. Polym. Sci. 106 (2007) 712–721.
[14] X. Meng, X. Du, Z. Wang, W. Bi, T. Tang, Compos. Sci. Technol. 68 (2008)
of TPU matrix. DMA temperature sweep measurements illustrated 1815–1821.
that the glass transition temperature of the TPU/OMMT nanocom- [15] H.R. Dennis, D.L. Hunter, D. Chang, S. Kim, J.L. White, J.W. Cho, D.R. Paul, Polymer
posite is improved, which is correspond to the restriction of the 42 (2001) 9513–9522.
[16] B.D. Kaushiva, G.L. Wilkes, Polymer 41 (2000) 6981–6986.
soft segments of TPU. DMA result with different strain amplitude [17] W. Xie, Z. Gao, K. Liu, W.P. Pan, R. Vaia, D. Hunter, A. Singh, Thermochim. Acta
reveals that at low strain the reinforcing effect of nanocomposites 367–368 (2001) 339–350.
with high filler loading seems to be improved by the physi- [18] J. Xiong, Y. Liu, X. Yang, X. Wang, Polym. Degrad. Stab. 86 (2004) 549–555.
[19] M. Tortora, G. Gorrasi, V. Vittoria, G. Galli, S. Ritrovati, E. Chiellini, Polymer 43
cal interaction among the ammonium modifiers residing inside (2002) 6147–6157.
the interlayer spacing of the Cloisite 15A with the carbonyl and [20] J. Ma, S. Zhang, Z. Qi, J. Appl. Polym. Sci. 82 (2001) 1444–1448.
ether group of the TPU matrix. DMA study carried out at vari- [21] B. Finnigan, K. Jack, K. Campbell, P. Halley, R. Truss, P. Casey, D. Cookson, S. King,
D. Martin, Macromolecules 38 (2005) 7386–7396.
ous frequencies indicate that both modulus and tan ı of the TPU [22] T.K. Chen, J.Y. Chui, T.S. Shieh, Macromolecules 30 (1997) 5068–5074.
nanocomposites increases with increase in filler loading and always [23] Y. Li, T. Gao, J. Liu, K. Linliu, C.R. Desper, B. Chu, Macromolecules 25 (1992)
show higher value compared to TPU matrix at all the measured 7365–7372.
[24] T.K. Chen, Y.I. Tien, K.H. Wei, J. Polym. Sci. A: Polym. Chem. 37 (1999)
frequencies.
2225–2233.
[25] C.H. Wang, M.L. Auad, N.E. Marcovich, S. Nutt, J. Appl. Polym. Sci. 109 (2008)
Acknowledgements 2562–2570.
[26] L.M. Leung, J.T. Koberstein, Macromolecules 19 (1986) 706–713.
[27] W.J. Seo, Y.T. Sung, S.J. Han, Y.H. Kim, O.H. Ryu, H.S. Lee, W.N. Kim, J. Appl. Polym.
The authors would like to thankfully acknowledge the finan- Sci. 101 (2006) 2879–2883.
cial assistance provided by Extra Mural Research Division (EMR-II), [28] M. Siliani, M.A. López-Manchado, J.L. Valentín, M. Arroyo, A. Marcos, M. Khayet,
Human Resource Development Group (HRDG), Council of Scien- J.P.G. Villaluenga, J. Nanosci. Nanotechnol. 7 (2007) 634–640.
[29] D.W. Schaefer, J. Zhao, H. Dowty, M. Alexander, E.B. Orler, Soft Matter 4 (2008)
tific and Industrial Research (CSIR), New Delhi 110 012, India vide 2071–2079.
sanction no. 22(0410)/06/EMR-II, dated 13-09-2006. Authors also [30] A.R. Payne, J. Appl. Polym. Sci. 6 (1962) 57–63.
thank Miku Traders, Vadodara 390 005, Gujarat, India, for providing [31] L. Ibarra, A. Macías, E. Palma, J. Appl. Polym. Sci. 57 (1995) 831–842.
[32] C. Wisniewski, G. Marin, Ph. Monge, Eur. Polym. J. 21 (1985) 479–484.
Tecoflex® TPU for the research work.

You might also like