Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Materials & Design 131 (2017) 1–11

Contents lists available at ScienceDirect

Materials & Design


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

High thermal insulation and compressive strength polypropylene foams MARK


fabricated by high-pressure foam injection molding and mold opening of
nano-fibrillar composites
Jinchuan Zhaoa,c, Qingliang Zhaoa, Chongda Wangc, Bing Guoa,⁎, Chul B. Parkc,⁎,
Guilong Wangb,c,⁎
a
Centre for Precision Engineering, School of Mechatronics Engineering, Harbin Institute of Technology, Harbin 150001, China
b
Key Laboratory for Liquid-Solid Structural Evolution and Processing of Materials (Ministry of Education), Shandong University, Jinan, Shandong 250061, China
c
Microcellular Plastics Manufacturing Laboratory, Department of Mechanical and Industrial Engineering, University of Toronto, Toronto, Ontario M5S 3G8, Canada

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

Keywords: Polypropylene (PP) foams with a thermal conductivity of as low as 36.5 mW m− 1 K− 1 are fabricated by high-
Foam injection molding pressure foam injection molding followed by mold-opening with CO2 as a blowing agent. Regular PPs are not
Mold-opening suitable for foaming due to their poor melt strength. To improve melt strength, the in-situ fibrillated blends of
Polypropylene PP/polytetrafluoroethylene (PTFE) are prepared using a regular co-rotating twin-screw extruder. The micro-
PTFE fibrils
morphology characterized by SEM shows that fibrillated nanoscale PTFE fibers disperse very well in PP matrix.
Thermal insulation
Compressive strength
The DSC, dynamic shear rheology, and extensional rheology measurements demonstrate that the in-situ
fibrillated PTFE fibers can significantly improve crystallization, visco-elastic performance, and strain-hardening
behaviors, respectively. All these factors confirm that PTFE fibers are very effective to improve melt strength and
thus foaming ability of PP. The foam injection molding results show that the PP foam's cell size reduces by nearly
one order of magnitude while its expansion ratio increases by approximately three times in presence of PTFE
fibers. Compared to PP foams, PP/PTFE foams show significantly improved thermal insulation performance due
to the increased expansion ratio, as well as unique cell wall structures with micro-holes and/or nano-fibrils.
Moreover, it demonstrates that smaller cell size leads to improved compressive strength.

G RA P H I C A L AB S T R A C T


Corresponding authors.
E-mail addresses: guobing@hit.edu.cn (B. Guo), park@mie.utoronto.ca (C.B. Park), guilong@sdu.edu.cn (G. Wang).

http://dx.doi.org/10.1016/j.matdes.2017.05.093
Received 23 March 2017; Received in revised form 29 May 2017; Accepted 29 May 2017
Available online 31 May 2017
0264-1275/ © 2017 Elsevier Ltd. All rights reserved.
J. Zhao et al. Materials & Design 131 (2017) 1–11

1. Introduction foaming behavior in high-pressure foam injection molding with mold


opening.
In recent decades, world energy consumption has increased more In this study, the in-situ fibrillation process based on a twin-screw
than two times from 1973 to 2014 [1], where the share of total world compounding was conducted to prepare nano-fibrillar PP/PTFE com-
energy production from fossil fuels is up to 81.4%. The burning of fossil posite with improved melt strength and foaming ability. To clarify PTFE
fuels releases billions of polluting emissions, damaging the environment fibrils' role in the crystallization behavior of PP, the non-isothermal
[2]. Renewable energy is an efficient solution to the pollution, but it crystallization of PP and PP/PTFE composites were investigated by a
contributes only 19.2% to the global energy consumption reported by differential scanning calorimeter (DSC). Furthermore, the dynamic
REN21 of 2016 [3]. In this context, energy conservation through frequency sweep tests, dynamic temperature sweep tests, and exten-
enhancing energy efficiency shows great potential to achieve global sional viscosity tests were implemented using a rheometer to study the
energy sustainability. Thermal insulation materials play a pivotal role effect of PTFE fibrils on the rheological performance of PP melt.
in energy conservation [4], of which the most commonly used ones are Moreover, thermal conductivity and compressive strength were mea-
mineral wools, fibreglass, cellulose, and polymeric foams. These sured to evaluate the performance of PP/PTFE foams prepared for
common insulation materials present a typical thermal conductivity thermal insulation applications. The mechanisms of cellular structure's
of higher than 35 mW m− 1 K− 1 in the absence of special insulation gas effect on thermal conductivity and compressive strength were dis-
at standard conditions. Although silica-based aerogels exhibit much cussed. The prepared PP foams showed excellent thermal insulation
better thermal insulation performance than common insulation materi- performance with a thermal conductivity of as low as
als, their application are still very limited due to high cost, difficulty of 36.5 mW m− 1 K− 1.
processing, intrinsic brittle feature, as well as toxic substances involved
in their fabrication procedures [5]. 2. Experimental
Among common insulation materials, polymeric foams currently
ranking second wide applications exhibit many superior properties, 2.1. Materials and sample preparation
such as better insulation, easier handling and installing, lower cost, and
no water absorption [5–8]. General technologies to fabricate polymeric The polymer used was a linear isotactic polypropylene (Novatec-PP
foams include batch foaming, bead foaming, foam extrusion, and foam FY4), a commercial homo-polymer supplied by Japan Polypropylene
injection molding. Among these technologies, foam injection molding Corporation, with a MFR of 5 g/10 min (at 230 °C/2.16 kg load).
introduces several advantages, such as a shorter cycle time, less Polytetrafluoroethylene powders, Metablen-PTFE A-3000, were sup-
material usage, greater dimensional stability, and lower energy con- plied by Mitsubishi Rayon. The blowing agent used was supercritical
sumption [9]. Moreover, it has potential to produce porous products CO2 with a purity of higher than 99%, which was purchased from Linde
with improved mechanical properties. However, the traditional foam Gas Inc.
injection molding process can only produce foams with small expansion Four PP blends with various PTFE weight concentrations of 100/0,
ratios, typically < 1.5-fold [9–11]. To achieve excellent thermal in- 99/1, 97/3, and 95/5 were prepared. A co-rotational twin-screw
sulation foams, a large expansion ratio is essential to minimize the heat extruder, manufactured by Toshiba Machine Co. Ltd. (Trade name:
transfer contributed by solid conduction [12]. To increase the foam's TEM-26SS), was used for compounding. The temperatures from the
expansion ratio, high-pressure foam injection molding [13,14] with hopper zone to the die were set to be 50 °C, 160 °C, 180 °C, 200 °C,
mold-opening was developed [15–17], which can not only increase 200 °C, 200 °C, 200 °C, 200 °C, 200 °C, 200 °C, and 190 °C, respectively,
expansion ratios but also improve mechanical performances including as shown in Fig. S1 of the supporting information. Compounding run at
fracture toughness, impact and compressive strength [18,19]. a discharge rate and a screw speed of 20 kg/h and 200 rpm, respec-
Polypropylene (PP), a common semi-crystalline polymer, has be- tively.
come the second most important plastic with a global market share up To prepare samples for dynamic frequency sweep tests and dynamic
to 19% by 2020 [20], due to its superior robustness to chemical and temperature step tests, PP/PTFE composites in pellet forms were
moisture attack, mechanical performance, manufacturability, recycl- molded to round plates (i.e., Φ25 mm × 1.2 mm) by hot compression.
ability, and versatile applications. In particular, foamed PP exhibits To prepare samples for extensional viscosity tests, the composites were
remarkable impact strength and toughness, durability, and strength to molded to rectangle shapes (i.e., 10 mm × 18 mm × 0.7 mm), with a
weight ratio [21,22]. Due to its unique and advantageous character- duration of 4 min at 180 °C.
istics, PP is an ideal candidate to prepare high-performance thermal
insulation foams combined with excellent mechanical properties. 2.2. PP/PTFE blends morphology characterization
However, pure PP generally exhibits low melt strength and weak
strain-hardening behavior due to its linear structure of molecules as A scanning electron microscope (SEM, JEOL 6060) was used to
well as the crystalline phase [23,24], which leads to a poor foaming characterize the morphology of PP/PTFE composites. To observe the
ability. One solution is to modify the polymer chain by cross-linking or morphology and dispersion of PTFE fibrils, the sample was cryogeni-
long-chain branching, which can remarkably enhance the strain-hard- cally fractured in liquid nitrogen, and the fracture section was then
ening response yet accompanying with high cost of manufacture exposed to the xylene bath for 2 min at 130 °C. Since xylene can
[25,26]. Another solution is to improve crystallization using nucleating dissolve the PP matrix while preserve PTFE fibrils which can thus be
agents because crystallization helps to improve the melt strength and observed clearly. Thereafter, the samples were dried after being cleaned
thus foaming ability of PP. A variety of nucleating agents such as nano-/ by acetone. Finally, the etched section was coated with a 10–20 nm
micron-sized talc [27], clay [28,29], and silica [30,31], as well as fibrils layer of platinum using a sputter coater for SEM observation. The SEM
(e.g. cellulosic fibrils [32], and polytetrafluoroethylene (PTFE) fibrils images were analyzed to acquire the cellular structure information of
[33]) have exhibited excellent potential on improving PP's crystal- foams using the software, Image J.
lization and hence foaming ability. Rizvi et al. [34,35] have revealed
that PTFE fibrils can dramatically increase the bubble density of PP 2.3. DSC characterization
foam by up to three orders of magnitude in foam extrusion process. The
mechanism is that the in-situ fibrillated PTFE fibrils create physical A differential scanning calorimeter (DSC-Q2000, TA instruments),
networks of entanglements in PP matrix, resulting in significantly was used to measure the endothermic peak caused by crystal dissolu-
enhanced strain-hardening behavior and melt strength. However, to tion, for investigating the crystallization processes of the PP/PTFE
the best of our knowledge, there is few studies involving the PP/PTFE's blends. In a typical measurement, a small amount of material (around

2
J. Zhao et al. Materials & Design 131 (2017) 1–11

Mold closing Filling Dwelling Mold-opening Cooling


Fig. 1. Schematic of the mold-opening based foam injection molding (MOFIM) process.

10 mg) was first packaged with an aluminium pan and cap, and then dwelling (packing) time directly influences the temperature and crystal-
put into a chamber filled with nitrogen. First, the blend was heated lization behavior of the melt while the mold-opening distance impacts
from 25 °C to 230 °C at 10 °C/min, and kept in the isothermal condition the cell growth significantly, which are the reasons for choosing the
of 230 °C for 5 min, to eliminate the thermal history of the blends. After dwelling time and mold-opening distance as variables.
that, the melt was cooled to 25 °C at 10 °C/min, and thereafter heated
to 230 °C again at 10 °C/min, to attain the thermal properties of the 2.6. Foam characterization
blends. To finish the measurement, the blend was cooled to 25 °C at
25 °C/min. A water displacement method (ASTM D792-00) was used to
measure the density of the PP/PTFE blends and foamed samples, and
2.4. Rheological characterizations the foam's expansion ratio (ϕ) is the density ratio of foams to blends. It
should be noted that the solid skins of the samples were removed before
The dynamic frequency sweep experiments were conducted using a the measurement. The cell size was measured based on the SEM
rotational rheometer, ARES, TA Instruments, to characterize the micrographs, and at least 100 cells were measured for each case to
rheology behaviors of PP and PP/PTFE composites. To identify the obtain the average cell size. The cell nucleation density (Nn) was
strain limits of the linear viscoelastic regions, all dynamic strain sweep calculated as follows [12]:
tests were implemented at a pre-set strain of 1% over a frequency range ⎛ n ⎞3 2
from 0.1 rad/s to 100 rad/s at 190 °C. The dynamic temperature step Nn = ⎜, ⎟ × ϕ
⎝ A⎠ (1)
tests were also conducted to analyze the rheological behavior of blends
as the temperature decreased from 220 °C to 100 °C at 2 °C/min with a where n and A are the measured cell number of the micrograph and the
frequency of 0.63 rad/s. area of the micrograph, respectively.
Extensional viscosity of the PP/PTFE blends was conducted using
the same rheometer equipment (ARES, TA Instruments) to examine 2.7. Thermal insulation characterization
whether the strain-hardening behavior occurs in the presence of PTFE
fibrils. In measurement, the extension rate was set at three levels: A transient plane source hot disk thermal constant analyzer (TPS
0.01 s− 1, 0.1 s− 1, and 1 s− 1, and the temperature selected was 170 °C. 2500, Them Test Inc., Sweden) was used to measure the thermal
For each condition, experiments were repeated by at least three times, conductivity of the solid and foamed samples [36,37]. In a standard
and the average values were reported. measurement, a round planar sensor made of a double spiral nickel foil,
acting simultaneously as a heating device and as a temperature sensor,
2.5. Mold-opening process and conditions is first placed between two pieces of the sample. Afterwards, current

A 50-ton Arburg Allrounder 270 injection molding machine,


equipped with a Mucell SCF delivery system (Trexel Inc.,
Massachusetts), was used in experiments. The mold used was a two-
plate one with a fan gate and a rectangular mold cavity (132 mm
108 mm × 3.2 mm). Fig. 1 illustrates the schematic of MOFIM proce-
dures. First, polymer/gas solution in barrel is injected to fill mold
cavity. Second, as the mold cavity is fully filled, a high packing pressure
is maintained for a certain of dwelling time to compress the cells
formed in filling stage back into polymer melt. Third, the injection mold
is opened to trigger foaming. Finally, foam's cellular structure is
stabilized by cooling. CO2 was pressurized to 24 MPa at the delivery
system and injected into the barrel of the machine at 8 wt% of the
molten PP. The temperatures of melt and mold were kept at 220 °C and
80 °C, respectively. The injection speed was 100 cm3/s. and the shot
size was 60 cm3. The packing pressure was set to 30 MPa. The values of
dwelling time were set to be 32.5 s, 35 s, 37.5 s, and 40 s, respectively.
The values of mold-opening distance were set to be 12 mm, 15 mm,
18 mm, 21 mm, and 24 mm, respectively. The high packing pressure
was used to extinguish the bubbles formed in the filling stage. The Fig. 2. SEM image of the PP/PTFE (5 wt%) fracture surface etched by xylene.

3
J. Zhao et al. Materials & Design 131 (2017) 1–11

a 0
b 2.0

Heat Flow (W/g)


Heat Flow (W/g)
-1 1.5

-2 1.0
0 wt.% PTFE 0 wt.% PTFE
-3 1 wt.% PTFE 0.5 1 wt.% PTFE
3 wt.% PTFE 3 wt.% PTFE
5 wt.% PTFE 5 wt.% PTFE
-4 0.0
100 120 140 160 155 160 165 170 175
Temperature (°C) Temperature (°C)
Fig. 3. DSC curves of PP/PTFE blends of (a) cooling graphs at a cooling rate of 10 °C/min; and (b) the second heating graphs at a heating rate of 10 °C/min.

power is supplied to the sensor for step-wise heating of the sample. PP/PTFE blends as a function of PTFE content at cooling/heating trace
During heating, the sensor can record its temperature response as a of DSC. It is observed from Fig. 3a that the addition of PTFE fibrils
function of time. Based on the recorded temperature response curve, promoted a remarkable shift of crystallization peak to the higher
the sample's thermal conductivity can be derived by using the built-in temperature, which clearly demonstrated that PTFE fibrils can promote
Hot Disk Thermal Analyser software. The diameters of the sensor and crystallization significantly due to their heterogeneous nucleation. In
sample were 4 mm and 20 mm, respectively, and the samples was the heating graphs as Fig. 3b shows, the peak of melting temperature of
measured at room temperature (23 ± 2 °C). The power output and test PP/PTFE blends is higher than that of PP, indicating that PTFE fibrils
duration were set at 0.005 W and 10 s, respectively. At least three improved the quality of crystalline structure, such as quantity, size, and
specimens were measured for each condition, and the average value distribution [38–40]. In order to quantitatively estimate the effect of
was recorded. PTFE fibrils on the crystallization behavior of PP, the key parameters
from DSC scans were summarized in Table 1. We noticed that the
2.8. Mechanical properties crystallization peak was increased by > 10 °C, the melting temperature
was increased by 1.5–3 °C, and the degree of crystallinity was increased
A computer controlled Instron mechanical system (INSTRON 2710- from 36.14% to 40.18% in the presence of PTFE fibrils. All of these
102) was used to investigate the compressive strength of the PP/PTFE factors clearly demonstrate that FTFE fibrils are very effective hetero-
foams, according to ASTM D1621-2010. Cubic samples with a side geneous nucleating agents in promoting crystallization of the PP
length of 15 mm, which were cut from the fabricated PP/PTFE foam, matrix.
were used for compression testing. The compression speed used in
measurement was 1.5 mm/min. For each case, at least three samples 3.3. Rheological behavior of the PP/PTFE blends
were tested and the average value was reported.
The rheological behavior of material plays a critical role in both
3. Results and discussion bubble nucleation and growth of the foaming process. The melt linear
viscoelastic responses of PP and its blends were characterized to
3.1. Fibrillated PP/PTFE blends investigate the effect of PTFE fibrils on the rheological behavior.
Fig. 4 plots the storage (G′) and loss (G″) modulus, complex viscosity
Fig. 2 shows the cryo-fractured surface of PP with 5 wt% of PTFE (η*), and tangent of the phase angle shift (tan δ) against frequency (ω)
after xylene etched. It can be found that the reticular fibrillated PTFE for PP and its blends, respectively. It is observed that G′ and G″
dispersed uniformly in the PP matrix. The average diameter of the PTFE increased at different PTFE concentrations as ω increasing from
fibril was around 300 nm. It is believed the reticular PTFE fibrils could 0.1 rad/s to 100 rad/s in Fig. 4a–b. Furthermore, the increments of G′
dramatically improve the melt strength of PP. Yamaguchi et al. and G″ at low frequencies (< 1 rad/s) were larger than that at high
investigated the mechanism of fibrillation of solid PTFE in PP melt frequencies (> 100 rad/s), which was resulted from the constraint
[33] and, furthermore, Rizvi et al. achieved similar interdigitated generated by the spatially linked structure at low frequency [41–43].
networks of PTFE fibrils for improving the foaming ability of PP in Besides, increasing the fibril concentration resulted in a gradual
extrusion foaming process [34,35]. increase of both G′ and G″, due to the increase of characteristic
relaxation time with the fiber concentration [44,45]. The longer
3.2. Crystallization behavior of the PP/PTFE blends characteristic relaxation time is available to generate higher local stress
for cell nucleation in the foaming process [21]. As Fig. 4c shows, η*
Fig. 3 shows the non-isothermal crystallization behavior of PP and values decreased with the increase of ω, since the increasing frequency
aligned the fibrils preferentially along the shear direction, thereby
Table 1 decreasing the viscosity. The observation that η* values increased with
Key parameters for the crystallization behavior of PP/PTFE blends. an increase in the PTFE concentration, indicates the formation of PTFE-
fibril network structure and its ability to improve the viscosity,
Crystallization Enthalpy- Melt Crystallinity (%)
especially at a low frequency. Fig. 4d plots the loss tangent, tan δ, to
peak (°C) heating (J/ temperature
g) (°C) characterize the elastic response of the polymer melt. The tan δ values
decreased with an increase in the fibril concentration, showing a lower
0 wt% PTFE 117.8 ± 0.1 74.8 ± 1.5 163.8 ± 0.1 36.1 ± 0.7 viscous/elastic ratio as the fibril concentration was increased. At high
1 wt% PTFE 126.9 ± 0.1 81.6 ± 2.1 165.4 ± 0.1 39.8 ± 1.0
PTFE-fibril concentrations (over 3 wt%), the value of tan δ became
3 wt% PTFE 128.2 ± 0.0 80.7 ± 1.1 166.1 ± 0.2 40.2 ± 0.5
5 wt% PTFE 128.9 ± 0.1 75.4 ± 1.8 165.8 ± 0.1 38.3 ± 0.9 almost independent of the frequency, attributing to the network
structure formed by fibril-fibril interactions (shown as Fig. 3) and/or

4
J. Zhao et al. Materials & Design 131 (2017) 1–11

a 105 b 105

104 104

G' (Pa)
103

G" (Pa)
103
5 wt.% PTFE
102 5 wt.% PTFE 102 3 wt.% PTFE
3 wt.% PTFE 1 wt.% PTFE
101 1 wt.% PTFE 101 0 wt.% PTFE
0 wt.% PTFE

100 100
0.1 1 10 100 0.1 1 10 100
ω (rad/s) ω (rad/s)
c 105 d 25

104 20 0 wt.% PTFE


1 wt.% PTFE
3 wt.% PTFE
η∗ (Pa⋅s)

103 15 5 wt.% PTFE

tan δ
102 5 wt.% PTFE 10
3 wt.% PTFE
101 1 wt.% PTFE 5
0 wt.% PTFE
100 0
0.1 1 10 100 0.1 1 10 100
ω (rad/s) ω (rad/s)
Fig. 4. Rheological behavior of PP/PTFE fibrillar blends with four different PTFE fibril content at 190 °C over a frequency range from 0.1 rad/s to 100 rad/s; (a) Frequency dependence of
storage modulus G′ (ω); (b) Frequency dependence of loss modulus G″ (ω); (c) Frequency dependence of complex viscosity η* (ω); (d) Frequency dependence of loss tangent, tan δ.

107 fibril-polymer chain interactions [42,45].


To characterize the viscoelastic response in the extension flow, the
106 growth curves of uniaxial extensional viscosity (ηE+) over various strain
rates were measured at 170 °C shown as Fig. 5. The transient exten-
5 wt.% PTFE sional viscosity, ηE+, described as the tensile-stress growth coefficient,
ηΕ+ (Pa⋅s)

105
is related with the tensile stress growth function (σE+ (t , ε )̇ ) and exten-
sional strain rate (ε̇) illustrated as Eq. (2)
104
0 wt.% PTFE
1.00 s-1 σE+ (t , ε )̇
3 ηE+ (t , ε )̇ =
10 0.10 s-1 ε̇ (2)
0.01 s-1
102 It is obvious from Fig. 5 that the addition of PTFE fibrils improved
0.01 0.1 1 10 100 1000 the extensional viscosity by nearly one order of magnitude. Further-
Time (s) more, PTFE fibrils also effectively enhanced the strain hardening
response for the PP matrix over the applied strain rates, including
Fig. 5. Uniaxial extensional viscosity of PP/PTFE (5 wt%) and PP, ηE+ (t , ε )̇ , measured at a
1 s− 1, 0.1 s− 1, and 0.01 s− 1. This enhancement in extensional viscos-
temperature of 170 °C and at strain rates of 0.01 s− 1, 0.1 s− 1, and 1 s− 1.
ity and strain hardening is attributed to the physical network formed by
the flexible PTFE fibrils [33,34], as illustrated by the SEM micrograph
in Fig. 2 and the platform-shaped curve in Fig. 4d. Besides, the degree
of strain hardening depended upon the applied strain rate, showing as
the larger the strain rate, the higher the degree of strain hardening. The
origin of the strain hardening behavior depending on the strain rates is
106
that strain hardening generally relates to the inability of the macro-
0 wt.% PTFE molecules to disentangle quickly enough and follow the deformation
1 wt.% PTFE [46–49].
105
3 wt.% PTFE As for semi-crystal polymers, temperature critically influences the
η∗ (Pa⋅s)

5 wt.% PTFE rheological properties due to the rapid change in viscosity as the
crystallization behavior occurs. To investigate the rheological proper-
104
ties of blends as a function of temperature, dynamic temperature step
tests were carried out with temperature decreasing from 220 °C to
120 °C at 2 °C/min, and the complex viscosity (η*) was illustrated in
103
Fig. 6. η* increased with the decrease of temperature, and a rapid
increase of η* occurred at a temperature lower than the crystallization
temperature owing to the formation of interlaced crystals during the
102
100 150 200 250 crystallization process. Moreover, the PTFE fibrils effectively improved
the crystallization temperature η* over all the applied temperature. The
Temperature (°C) higher the PTFE concentration was, the larger the η* value was, which
Fig. 6. Rheological behavior of PP/PTFE fibrillar blends with four different PTFE fibril was consequent with the DSC and rheological results in Fig. 3 and
content at a frequency of 0.63 rad/s over a temperature range from 220 °C to120 °C. Fig. 4c, respectively.

5
J. Zhao et al. Materials & Design 131 (2017) 1–11

Fig. 7. Geometry of the foam product showing two different sections for SEM observation.

3.4. Foaming behavior and cellular morphology (expansion ratio was around 8-fold). Obviously, when the dwelling time
was 25 s, there is no foam obtained between the solid skins, a hollow
Fig. 7 shows a schematic view of the injection molding product, balloon-like sample as shown in the first row of Fig. 8. At early dwelling
including the basic size, the typical morphological structure of injec- times (< 25 s), the melt temperature was relatively higher [17], and
tion-molded foams, and the method to prepare samples for the SEM hence, the melt strength was too weak to withstand the biaxial or
observation in two different sections, perpendicular and parallel to the uniaxial stretch due to bubble growth because there were few crystals
mold-opening direction. Note that the cells were clearly stretched in the foamed in the high temperature PP matrix. With an increase in the
mold-opening direction, compared with that in the direction perpendi- dwelling time, well foamed samples were achieved as in the middle
cular to mold opening, and a classical sandwich structure was region of Fig. 8, resulting from the increased crystals which improved
generated in the mold-opening direction. The cell structures in the the melt strength and accordingly enhanced the foaming ability of PP
subsequent characterization were all in the section perpendicular to the matrix. Furthermore, it can be deduced that the PTFE fibrils improve
mold-opening direction. the melt strength of PP mainly by perfecting its crystallization behavior,
To investigate the effect of dwelling time as well as the in-situ rather than the self-structure of fibrils. However, when the dwelling
fibrillated PTFE fibrils on the foaming behavior of PP matrix, the time was excessively longer (> 45 s), the superabundant crystals
MOFIM experiments were carried out for the PP blends with four foamed network structure, significantly enhancing the rigidity of the
different concentrations of PTFE. Fig. 8 illustrates the SEM micrographs PP melt, which greatly suppressed cell nucleation and growth, and thus,
of the foams fabricated with MOFIM by changing the dwelling time partly foamed samples was attained in the last row of Fig. 8. Besides, it
from 25 s to 45 s and setting the mold-opening distance to 12 mm is also noticed that the addition of PTFE fibrils effectively decreased the

Fig. 8. SEM micrographs of PP foam structures with four different PTFE concentrations over a variation of dwelling time from 25 s to 45 s as the foam's expansion ratios fluctuate around
8-fold.

6
J. Zhao et al. Materials & Design 131 (2017) 1–11

a 900 b 10
9
0 wt.% PTFE
750 1 wt.% PTFE 8

Cell Density (cell/cm )


10

3
600 3 wt.% PTFE
5 wt.% PTFE

Cell Size ( μm)


450 10
7

5 wt.% PTFE
75 10
6
3 wt.% PTFE
60 1 wt.% PTFE
5 0 wt.% PTFE
45 10
30 4
10
15
3
0 10
27 30 33 36 39 42 45 27 30 33 36 39 42 45
Dwelling Time (s) Dwelling Time (s)
Fig. 9. Foam characterization: comparison of (a) the cell size and (b) the cell density of PP and PP/PTFE fibrillar blends as a function of the dwelling time.

cell size. With an increase in the PTFE concentration, the cell size also cm3. As the PTFE concentration increases from 1 wt% to 5 wt%, the cell
had a tendency to decrease. The decreased cell size was owing to the densities were increased from 107 cells/cm3 to 109 cells/cm3. Besides,
effect of PTFE fibrils on the crystallization and rheological behavior, the cell densities of PP/PTFE foams also presented sensitivity to
which can be verified by the results of DSC and rheological tests. dwelling time, and the larger the PTFE concentration was, the higher
To quantitatively analyze the role of dwelling time in promoting the sensitivity was. The addition of PTFE promoted the crystallinity and
foaming ability of PP matrix, the average cell sizes and cell nucleation melt strength, and thus, improved the cell nucleation and simulta-
densities in the section which is perpendicular to the mold-opening neously suppressed the collapse and coalescence of cells, leading to the
direction were calculated based on the SEM micrographs. Fig. 9a plots significant enhancement of cellular morphology [33,34].
the cell sizes of PP and PP/PTFE blends as a function of the dwelling To further increase the expansion ratio, the experiments with larger
time, where the aspect ratio is around 3 which is the ratio of the cell mold-opening distances were carried out using pure PP and PP blends
diameter in the mold-opening direction to that in its perpendicular with 5 wt% PTFE. It is found that the expansion ratio of PP foams
section. The cell size decreased with an increased dwelling time over cannot increase further with the mold-opening distance, the maximum
different concentrations of PTFE. By adding only 1% PTFE, the cell sizes expansion ratio that can be achieved for pure PP foam was < 10-fold.
decreased by one order of magnitude, from around 600 μm to 60 μm, In contrast, the maximum expansion ratio that can be achieved for PP/
and as the PTFE concentration increased, the cell sizes further PTFE blend can be higher than 18-fold. The results could be explained
decreased from 60 μm to 30 μm. PTFE fibrils dramatically promoted by the effect of PTFE fibrils on foaming ability as follows. PTFE
the crystallization kinetics of the PP matrix, i.e., melt strength, and a efficiently improved the crystallinity and visco-elastic properties sev-
higher PTFE concentration more effectively influenced the crystallinity erally confirmed by Fig. 4 and Fig. 5–7, and thus promoted the cell
by supplying more crystal nucleation agents [34], corresponding with heterogenous nucleation and suppressed the collapse and coalescence
the DSC results. Furthermore, we noticed that the cell size with a high of cells. Furthermore, the enhanced melt strength reduced the loss of
PTFE concentration was more sensitive to dwelling time than that with gas, leading to the improvement of expansion by supplying sufficient
a low PTFE concentration, which was illustrated by the variation gas.
tendency of cell sizes with PTFE concentration varying from 1 wt% to To study the role of the mold-opening distance on the cellular
5 wt% (Fig. 9a). It indicated that the variation of melt temperature by morphology, Fig. 10a–b plot the cell sizes and cell densities of PP blends
changing dwelling time, impacted more significantly on melt strength with 5 wt% PTFE over different mold-opening distances, respectively.
of the blend with high-concentration PTFE than that with low- The cell sizes and cell densities had been kept almost constant around a
concentration PTFE, resulting from the stronger interaction of higher value with an increased mold-opening distance over different dwelling
crystalline blends [33]. times. To clarify their correlation, SEM micrographs were detected in
Fig. 9b shows the foam cell nucleation densities of PP and PP/PTFE two directions: parallel (Fig. 11a–c) and perpendicular (Fig. 11d–f) to
blends corresponding to different dwelling times. The cell nucleation the mold-opening direction. We noticed that as the mold-opening
densities increased gradually with an increase in the dwelling time. distance increases, cells were significantly stretched along the mold-
Notably, by blending PTFE into PP, the cell densities were dramatically opening direction (Fig. 11a–c). However, the cell structures in the
increased by three orders of magnitude, from 104 cells/cm3 to 107 cells/ section perpendicular to the mold-opening direction were almost the

a 50 b 1010
Dwelling Time-35.0s Dwelling Time-37.5s
Cell Density (cells/cm3)

Dwelling Time-37.5s Dwelling Time-40.0s


40 Dwelling Time-40.0s Dwelling Time-35.0s
Cell Size ( μm)

109

30

108
20

10 107
12 14 16 18 20 22 24 12 14 16 18 20 22 24
Mold-opening Distance (mm) Mold-opening Distance (mm)
6 8 10 12 14 16 18 6 8 10 12 14 16 18
Expansion Ratio Expansion Ratio
Fig. 10. Variation of the (a) cell size and (b) cell density of PP/PTFE (5 wt%) foams with different mold-opening distances.

7
J. Zhao et al. Materials & Design 131 (2017) 1–11

a b c

d e f

Fig. 11. SEM micrographs of PP/PTFE (5 wt%) foams using 37.5 s as the dwelling time, (a) and (d) 12 mm, (b) and (e) 18 mm, (c) and (f) 24 mm in the mold-opening direction in two
directions, parallel and perpendicular to the mold-opening direction, respectively.

80 crystallinity could generate high-stiffness melt, which could be harmful


0 wt.% PTFE for cell nucleation due to the increased energy barrier [21].
5 wt.% PTFE-Micro holes
70 5 wt.% PTFE-Micro holes
λtot (mW m-1 K-1)

and Nano fibrils 3.5. Thermal property characterization of PP and PP/PTFE foams
5 wt.% PTFE-Nano fibrils
60
To investigate the thermal insulation properties of the prepared PP
and PP/PTFE (5 wt%) foams, their thermal conductivities were mea-
50
sured using the TPS technique. Fig. 12 shows the variation of the
thermal conductivity of the PP foams and the PP/PTFE foams as a
40 function of the expansion ratio. The thermal conductivity gradually
decreased with an increase in the expansion ratio for both PP and PP/
30 PTFE foams. The minimum thermal conductivity of the PP foam was
5 10 15 20
about 60 mW m− 1 K− 1 corresponding to the maximum expansion ratio
Expansion Ratio of 10-fold while the minimum thermal conductivity of the PP/PTFE
Fig. 12. Variation of thermal conductivity of foams with four different PTFE content over foam was as small as 36.5 mW m− 1 K− 1 corresponding to the max-
an expansion ratio range from 6 to 18-fold. imum expansion ratio of 18-fold. The dramatically enhanced thermal
insulation property of the PP/PTFE foams should be mainly owing to its
same, Fig. 11d–f show. significantly increased expansion ratio. Notably, PP/PTFE foams had
It is also observed from Fig. 10a that cell size first decreased and much smaller thermal conductivity than the PP foams even under the
then increased with the increase of dwelling time regardless of the same expansion ratio. It means that the increased expansion ratio was
mold-opening distance. We noticed that the optimum dwelling time not the only reason leading to the improved thermal insulation property
was 37.5 s, which decreased the cell sizes by > 40%, from around of the PP/PTFE foams. It is observed from Fig. 12 that a longer dwelling
35 μm to < 20 μm. The reason for existing this optimum dwelling time benefited the thermal insulation property of the PP/PTFE foams.
could be explained from the crystallization's role in cell nucleation. A For better understanding the measured thermal conductivity of the
longer dwelling time can reduce the melt temperature and produce PP foams and the PP/PTFE foams, a classical theoretical model was
more crystals, both of which help to improve the melt strength, thus employed to quantitatively model the heat transfer through the
producing more cells with smaller sizes. However, excessively high polymer foams. Basically, conduction, convection, and radiation are
the three fundamental modes of heat transfer. Heat transfer through

a b c

Fig. 13. SEM micrographs of PP and PP/PTFE (5 wt%) foams with a 10-fold expansion ratio: (a) PP foams, (b) PP/PTFE foams for 35 s dwelling time, and (c) PP/PTFE foams for 40 s
dwelling time.

8
J. Zhao et al. Materials & Design 131 (2017) 1–11

foams consists of four contributions, and the total thermal conductivity, be clearly observed that there were a mass of micro-holes and micro/
λtot, can be expressed as follows [50]: nano fibrils in the cell walls. Such micro/nano structures in the cell wall
λtot = λgas + λsol + λrad + λcon destroyed the continuity of the polymer matrix, significantly increased
(3)
the tortuosity of the heat transfer path, and also increased the phonon
where λgas, λsol are the thermal conductivities contributed by the gas scattering at the nano structure boundaries [50]. Thus, they could
phase and the polymer phase in the heat transfer mode of thermal increase the thermal resistance and hence reduce the thermal con-
conduction, which are proportional to their respective volume fraction ductivity. This should be the main reason why PP/PTFE foams had
in the polymer foams. λrad is the equivalent thermal conductivity much smaller thermal conductivities than the PP foams even if they had
contributed by thermal radiation. λcon is the equivalent thermal the same expansion ratio (Fig. 12). With an increase in the dwelling
conductivity contributed by thermal convection. Owing to the restraint time from 35 s to 40 s, it is observed that there were more open cells
of gas movement inside < 4 mm cells, the convection could be dis- with micro/nano holes and fibrils (Fig. 13). This can be used to explain
regarded [51]. Accordingly, Eq. (3) can be rewritten as follows: the improved thermal insulation property of the PP/PTFE foams with a
λtot = kgas Vgas + ζksol Vsol + λrad longer dwelling time. Moreover, the PTFE fibers can also significantly
(4)
reduce the foam's cell size and hence increase the number of cell walls.
where Vgas and Vsol are the volume fraction of the air phase and polymer The increased cell walls can block more thermal radiations [52], which
phase in the foams, respectively. ζ is a parameter related to the inherent also contributed to the smaller thermal conductivity of the PP/PTFE
tortuosity of the cell structure, which can be assumed that ζ = 1 to foams.
reduce Eq. (4) to the simple mixture rule [13]. kgas is the air's thermal
conductivity which is around 26 mW m− 1 K− 1 at the standard condi-
tions, and ksol is the polymer's thermal conductivity which is 3.6. Compressive strength characterization of PP/PTFE foams
231 mW m− 1 K− 1 measured by TPS. Notably, PTFE fibers show very
little effect on the thermal conductivity of PP (Fig. S2 in the Supporting The compression tests were conducted to investigate the effect of
information). the cellular morphology on the prepared PP/PTFE (5 wt%) foams'
Since the foam's expansion ratio is not large, the total thermal mechanical properties. Fig. 14a shows the stress-strain curves of the
conductivity is mainly contributed by the thermal conduction of the PP/PTFE foams with various expansion ratios. It clearly demonstrates
polymer phase due to the relatively small gaseous thermal conduction that the foam's compression modulus as well as the compressive
and simultaneously the weak thermal radiation [50]. In this situation, it strength reduced gradually with an increase in the expansion ratio.
will be very effective to reduce the foam's total thermal conductivity by Fig. 14b plots the dependence of the compressive strength on the
increasing the foam's expansion ratio due to the reduced thermal expansion ratio. The compressive strength reduced significantly with an
conductivity contributed by the polymer phase. Because the PP/PTFE increase in the expansion ratio but the reducing rate became slower. As
foam can have larger expansion ratios than the PP foam, the former one the expansion ratio increased from 8-fold to 18-fold with an increment
should therefore exhibit much better thermal insulation property than by 2.25 times, the compressive strength dramatically decreased from
the later one (Fig. 12). It is also noticed that the PTFE fibers can not 1.2 MPa to < 0.2 MPa with an decrement by > 6 times. Fig. 14c shows
only increase the expansion ratio but also opening in the cell walls. the stress-strain curves of the PP/PTFE foams with a constant expansion
Fig. 13 shows the typical cellular structure of the PP/PTFE foams. It can ratio of 8-fold. It clearly indicates that a smaller cell size helped to

a b
Compressive Strength (MPa)

1.6 1.6

ER:8
1.2 1.2
Stress (MPa)

ER:10
0.8 0.8
ER:12 ER:15
0.4 0.4
ER:18

0.0 0.0
0.0 0.1 0.2 0.3 0.4 0.5 5 10 15 20
Strain (mm/mm) Expansion Ratio
c d
Compressive Strength (MPa)

1.6 1.6

38 μm 41μ m 47 μm 43μm 50 μm
1.2 1.2
Stress (MPa)

0.8 0.8

0.4 0.4

0.0 0.0
0.0 0.1 0.2 0.3 0.4 0.5 36 40 44 48 52
Strain (mm/mm) Cell Size ( μm)
Fig. 14. Strain-stress curves and compressive strength of PP/PTFE (5 wt%) foams as a function of: (a) the strain and (b) the expansion ratio varying from 8-fold to 18-fold for a fixed cell
density of 108 / cm3; (c) the strain and (d) the cell size varying from 38 μm to 50 μm for a fixed 8-fold expansion ratio.

9
J. Zhao et al. Materials & Design 131 (2017) 1–11

increase the foam's compression modulus and strength. Fig. 14d plots Acknowledgement
the correlation between the cell size and the compressive strength. It is
observed that the compressive strength nearly increased linearly with This work was supported by the National Natural Science
the decrease of the cell size in the research scope. Specifically, the Foundation of China (NSFC, 51405267), Shandong Provincial Natural
compressive strength increased from 0.81 MPa to 1.24 MPa with an Science Foundation (ZR2014EEQ017), the Korean Ministry of Trade,
increment by 53.1% as the cell size reduced from 50 μm to 38 μm. Industry and Energy (MOTIE), and Korea Evaluation Institute of
According to the theoretical model developed by Gibson and Ashby, Industrial Technology (KEIT) as part of their Global Convergence ATC
the Young's modulus of the closed cell foam, E, can be evaluated by the project (ATC-10053160). The China Scholarship Council (CSC) that
following equation [53,54], enabled the first author to study at the University of Toronto is also
appreciated.
Es ⎡ ξ3 ⎤
E= ⎢1 + ⎥
ϕ2 (1 + ξ) ⎣
2 ϕ (1 + ξ ) ⎦ (5) Appendix A. Supplementary data

where Es is the Young modulus for the foam and the solid polymer. ϕ is Supplementary data to this article can be found online at http://dx.
the expansion ratio of foams. ξ is the volume ratio of the material in the doi.org/10.1016/j.matdes.2017.05.093.
cell wall to that in the cell strut. Based on Eq. (5), it clearly confirms
that the foam's compressive modulus reduces gradually with an References
increase in expansion ratio. But the decrease rate of compressive
modulus becomes slower with the increase of expansion ratio, which [1] U.S. Energy Information Administration, Annual Energy Outlook 2015 With
Projections to 2040. www.eia.gov/forecasts/aeo, (2017) (accessed 31.05.17).
is consistent with the experimental data shown in Fig. 14b.
[2] D.M. Ibrahiem, Evaluating cost of air pollution from using fossil fuels in some
To clarify the effect of foam's cell size on its compressive strength, industries in Egypt, Adv. Manag. Appl. Econ. 5 (2015) 27–39.
the cell strut's compressive strength was also investigated.The depen- [3] Paris: REN21 Secretariat, Renewables 2016 Global Status Report, REN21, (2016).
dence of the cell strut's compressive strength, σ, on the cell size can be [4] B.P. Jelle, Traditional, state-of-the-art and future thermal building insulation
materials and solutions-properties, requirements and possibilities, Energ. Buildings
described by the following expression [55], 43 (2011) 2549–2563.
[5] G. Wang, J. Zhao, L.H. Mark, G. Wang, K. Yu, C. Wang, C.B. Park, G. Zhao, Ultra-
σ = 36PLD (9πD 4 − 16bh3) (6) tough and super thermal-insulation nanocellular PMMA/TPU, Chem. Eng. J. 325
(2017) 632–646.
[6] S. Wong, J.W.S. Lee, H.E. Naguib, C.B. Park, Effect of processing parameters on the
where P is the fracture load; D is the outside diameter of the cell; and b mechanical properties of injection molded thermoplastic polyolefin (TPO) cellular
and h are the base and height of the triangular hole within the cell wall, foams, Macromol. Mater. Eng. 293 (2008) 605–613.
respectively. Based on Eq. (6), it is clear that the compressive strength [7] V. Bernardo, E. Laguna-Gutierrez, A. Lopez-Gil, M.A. Rodriguez-Perez, Highly
anisotropic crosslinked HDPE foams with a controlled anisotropy ratio: production
of a cell strut is significantly dependent on the cell size, and increases and characterization of the cellular structure and mechanical properties, Mater.
with the decrease of the cell size, which is in accordance with Fig. 14d. Des. 114 (2017) 83–91.
Therefore, it is an effective method to improve the compressive strength [8] J.W.S. Lee, C.B. Park, S.G. Kim, Reducing material costs with microcellular/fine-
celled foaming, J. Cell. Plast. 43 (2007) 297–312.
by decreasing the cell size while keeping an identical foam's expansion [9] J. Hou, G. Zhao, G. Wang, G. Dong, J. Xu, A novel gas-assisted microcellular
ratio. injection molding method for preparing lightweight foams with superior surface
appearance and enhanced mechanical performance, Mater. Des. 127 (2017)
115–125.
[10] G. Wang, G. Zhao, J. Wang, L. Zhang, Research on formation mechanisms and
4. Conclusion control of external and inner bubble morphology in microcellular injection
molding, Polym. Eng. Sci. 55 (2015) 807–835.
PP foams with thermal conductivity of as low as 36.5 mW m− 1 K− 1 [11] D. Jahani, A. Ameli, P.U. Jung, M.R. Barzegari, C.B. Park, H. Naguib, Open-cell
cavity-integrated injection-molded acoustic polypropylene foams, Mater. Des. 53
were fabricated by mold-opening foam injection molding with CO2 as (2014) 20–28.
the blowing agent, combining in situ fibrillation of PTFE in PP matrix to [12] P. Gong, P. Buahom, M.P. Tran, M. Saniei, C.B. Park, P. Pötschke, Heat transfer in
improve melt strength. The fibrillar morphology, crystallinity, rheolo- microcellular polystyrene/multi-walled carbon nanotube nanocomposite foams,
Carbon 93 (2015) 819–829.
gical properties, cellular morphology, thermal insulation and mechan- [13] V. Shaayegan, G.L. Wang, C.B. Park, Study of the bubble nucleation and growth
ical properties were characterized. The correlation between cellular mechanisms in high-pressure foam injection molding through in-situ visualization,
structure and the performance of PP/PTFE composite foams were Eur. Polym. J. 76 (2016) 2–13.
[14] V. Shaayegan, G.L. Wang, C.B. Park, Effect of foam processing parameters on pubble
clarified. nucleation and growth dynamics in high-pressure foam injection molding, Chem.
The fibrillated PTFE fibrils improved the cell morphology dramati- Eng. Sci. 155 (2016) 27–37.
cally by enhancing the melt strength, strain hardening behavior, and [15] A. Ameli, D. Jahani, M. Nofar, P.U. Jung, C.B. Park, Development of high void
fraction polylactide foams using injection molding: mechanical and thermal
crystallinity of PP due to their sub-micro fibrillar structure and good insulation properties, Compos. Sci. Technol. 90 (2014) 88–95.
dispersion in the matrix, which was represented as the remarkable drop [16] D. Jahani, A. Ameli, P.U. Jung, M.R. Barzegari, C.B. Park, H. Naguib, Open-cell
of the cell size by 30 times, from 600 μm to 20 μm, and the enhance- cavity-integrated injection-molded acoustic polypropylene foams, Mater. Des. 53
(2014) 20–28.
ment of the expansion ratio from 10-fold to around 20-fold, up to 2
[17] T. Ishikawa, M. Ohshima, Visual observation and numerical studies of polymer
times. foaming behavior of polypropylene/carbon dioxide system in a core-back injection
The PP/PTFE foams exhibited lower thermal conductivity than PP molding process, Polym. Eng. Sci. 51 (2011) 1617–1625.
foams. The improved thermal insulation property was majorly owing to [18] K.F. Karlsson, B. Tomasåström, Manufacturing and applications of structural
sandwich components, Compos. Part A 28 (1997) 97–111.
the dramatic increase of the expansion ratio. Moreover, PTFE caused [19] M.A. Dweiba, B. Hub, A. O'Donnella, H.W. Shentonb, R.P. Woola, All natural
the formation of micro-holes and/or nanoscale fibrils in cell walls, composite sandwich beams for structural applications, Compos. Struct. 63 (2004)
which further decreased the thermal conductivity due to phonon 147–157.
[20] Ceresana, Market Studies: Polypropylene. http://www.ceresana.com/en/, (2017)
scattering effect. Furthermore, PTFE significantly increased the quan- (accessed 31.05.17).
tity of cell walls, which could block more thermal radiation, and thus [21] A. Wong, Y. Guo, C.B. Park, Fundamental mechanisms of cell nucleation in
contributed to reducing the thermal conductivity. polypropylene foaming with supercritical carbon dioxide - effects of extensional
stresses and crystals, J. Supercrit. Fluids 79 (2013) 142–151.
As expected, the compressive strength of polymer foams reduces [22] K. Wang, F. Wu, W.T. Zhai, W.G. Zheng, Effect of polytetrafluoroethylene on the
significantly with the increase of expansion ratio. Reducing cell size led foaming behaviors of linear polypropylene in continuous extrusion, J. Appl. Polym.
to improved compressive strength of polymer foams because of the Sci. 129 (2013) 2253–2260.
[23] D.J. Fu, F. Chen, T.R. Kuang, D.C. Li, X.F. Peng, D.Y. Chiu, C.S. Lin, L.J. Lee,
reduced bending moment.

10
J. Zhao et al. Materials & Design 131 (2017) 1–11

Supercritical CO2 foaming of pressure-induced-flow processed linear polypropy- electrical and thermal properties of MWCNT/polyamide 6,6 composites prepared
lene, Mater. Des. 93 (2016) 509–513. by melt mixing, Carbon 50 (2012) 3694–3707.
[24] J.B. Bao, A.N. Juniora, G.S. Weng, J. Wang, Y.W. Fang, G.H. Hu, Tensile and impact [40] D.M. Lincoln, R.A. Vaia, Z.G. Wang, B.S. Hsiao, Secondary structure and elevated
properties of microcellular isotactic polypropylene (PP) foams obtained by super- temperature crystallite morphology of nylon-6/layered silicate nanocomposite,
critical carbon dioxide, J. Supercrit. Fluids 111 (2016) 63–73. Polymer 42 (2001) 1621–1631.
[25] H.E. Park, J.M. Dealy, Effects of long-chain branching, concentration of super- [41] S.S. Ray, K. Yamada, M. Okamoto, K. Ueda, Control of biodegradability of
critical CO2 and pressure on the viscosity of linear and branched polypropylenes, polylactide via nanocomposite technology, Macromol. Mater. Eng. 288 (2003)
Soc. Plast. Eng. 4 (2008) 2517–2521. 203–208.
[26] Y. Luo, C. Xin, D. Zheng, Z. Li, W. Zhu, S.K. Wu, Q.T. Zheng, Y.D. He, Effect of [42] G. Galgali, C. Ramesh, A. Lele, A rheological study on the kinetics of hybrid
processing history on the rheological properties, crystallization and foamability of formation in polypropylene nanocomposites, Macromolecules 34 (2001) 852–858.
branched polypropylene, J. Polym. Res. 22 (2015) 1–13. [43] V. Khoshkava, M.R. Kamal, Effect of cellulose nanocrystals (CNC) particle mor-
[27] J.S. Colton, Nucleation of microcellular foams in semi-crystalline thermoplastics, phology on dispersion and rheological and mechanical properties of polypropylene/
Mater. Manuf. Process. 42 (1989) 253–262. CNC nanocomposites, ACS Appl. Mater. Interfaces 6 (2014) 8146–8157.
[28] X.L. Jiang, J.B. Bao, T. Liu, L. Zhao, Z.M. Xu, W.K. Yuan, Microcellular foaming of [44] X.Y. Zhu, B. Kundukad, J.R.C. van der Maarel, Viscoelasticity of entangled λ-phage
polypropylene/clay nanocomposites with supercritical carbon dioxide, J. Cell. DNA solutions, J. Chem. Phys. 129 (2008) 185103.
Plast. 45 (2009) 515–538. [45] A.J. Hsieh, P. Moy, F.L. Beyer, P. Madison, E. Napadensky, J. Ren,
[29] M. Okamoto, P.H. Nam, P. Maiti, T. Kotaka, T. Nakayama, M. Takada, M. Ohshima, R. Krishnamoorti, Mechanical response and rheological properties of polycarbonate
A. Usuki, N. Hasegawa, H. Okamoto, Biaxial flow-induced alignment of silicate layered-silicate nanocomposites, Polym. Eng. Sci. 44 (2004) 825–837.
layers in polypropylene/clay nanocomposite foam, Nano Lett. 1 (2001) 503–505. [46] W.D. Ding, T. Kuboki, A. Wong, C.B. Park, M. Sain, Rheology, thermal properties,
[30] S. Jain, G. Han, M.V. Duin, P. Lemstra, Effect of in situ prepared silica nano- and foaming behavior of high D-content polylactic acid/cellulose nanofiber
particles on non-isothermal crystallization of polypropylene, Polymer 46 (2005) composites, RSC Adv. 5 (2015) 91544–91557.
8805–8818. [47] T. Yokohara, S. Nobukawa, M. Yamaguchi, Rheological properties of polymer
[31] A. Fereidoon, S. Memarian, A. Albooyeh, S. Tarahomi, Influence of mesoporous composites with flexible fine fibers, J. Rheol. 55 (2011) 1205–1218.
silica and hydroxyapatite nanoparticles on the mechanical and morphological [48] M. Yamaguchi, K. Fukuda, T. Yokohara, M.A.B.M. Ali, S. Nobukawa, Modification
properties of polypropylene, Mater. Des. 57 (2016) 201–210. of rheological properties under elongational flow by addition of polymeric fine
[32] H.S. Yang, A. Kiziltas, D.J. Gardner, Thermal analysis and crystallinity study of fibers, Macromol. Mater. Eng. 297 (2012) 654–658.
cellulose nanofibril-filled polypropylene composites, J. Therm. Anal. Calorim. 113 [49] A. Rizvi, C.B. Park, B.D. Favis, Tuning viscoelastic and crystallization properties of
(2013) 673–682. polypropylene containing in-situ generated high aspect ratio polyethylene ter-
[33] M.A.B.M. Ali, K. Okamoto, M. Yamaguchi, T. Kasai, A. Koshirai, Rheological ephthalate fibrils, Polymer 68 (2015) 83–91.
properties for polypropylene modified by polytetrafluoroethylene, J. Polym. Sci. B [50] G. Wang, C. Wang, J. Zhao, G. Wang, C.B. Park, G. Zhao, Modelling of thermal
Polym. Phys. 47 (2009) 2008–2014. transport through the nanocellular polymer foam: toward the generation of a new
[34] A. Rizvi, A. Tabatabaei, M.R. Barzegari, S.H. Mahmood, C.B. Park, In situ superinsulating material, Nanoscale 9 (2017) 5996–6009.
fibrillation of CO2-philic polymers: sustainable route to polymer foams in a [51] J. Kuhn, H.P. Ebert, M.C. Arduini-Schuster, D. Büttner, J. Fricke, Thermal transport
continuous process, Polymer 54 (2013) 4645–4652. in polystyrene and polyurethane foam insulations, Int. J. Heat Mass Transf. 35
[35] A. Rizvi, R.K.M. Chu, J.H. Lee, C.B. Park, Superhydrophobic and oleophilic open- (1992) 1795–1801.
cell foams from fibrillar blends of polypropylene and polytetrafluoroethylene, ACS [52] P. Ferkl, R. Pokorný, J. Kosek, Multiphase approach to coupled conductioneradia-
Appl. Mater. Interfaces 6 (2014) 21131–21140. tion heat transfer in reconstructed polymeric foams, Int. J. Therm. Sci. 3 (2014)
[36] M. Antunes, V. Realinho, J.I. Velasco, E. Solórzano, M.Á. Rodríguez, J.A. de Saja, 68–79.
Thermal conductivity anisotropy in polyfpropylene foams prepared by supercritical [53] L.J. Gibson, M.F. Ashby, The mechanics of three-dimensional cellular materials, R.
CO2 dissolution, Mater. Chem. Phys. 136 (2012) 268–276. Soc. London 382 (1982) 43–59.
[37] S.A. Al-Ajlan, Measurements of thermal properties of insulation materials by using [54] J. Andersons, M. Kirpluks, L. Stiebra, U. Cabulis, Anisotropy of the stiffness and
transient plane source technique, Appl. Therm. Eng. 26 (2006) 2184–2191. strength of rigid low-density closed-cell polyisocyanurate foams, Mater. Des. 92
[38] T.D. Fornes, D.R. Paul, Crystallization behavior of nylon 6 nanocomposites, (2016) 836–845.
Polymer 44 (2003) 3945–3961. [55] R. Brezny, D.J. Green, C.Q. Dam, Evaluation of strut strength in open-cell ceramics,
[39] C. Caamaño, B. Grady, D.E. Resasco, Influence of nanotube characteristics on J. Am. Ceram. Soc. 72 (1988) 885–889.

11

You might also like