Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

Polymer Bulletin (2021) 78:6777–6795

https://doi.org/10.1007/s00289-020-03454-3

ORIGINAL PAPER

Effect of waste fillers addition on properties of high‑density


polyethylene composites: mechanical properties, burning
rate, and water absorption

S. Z. M. Rasib1 · M. Mariatti1   · H. Y. Atay2

Received: 12 July 2020 / Revised: 28 September 2020 / Accepted: 1 November 2020 /


Published online: 10 November 2020
© Springer-Verlag GmbH Germany, part of Springer Nature 2020

Abstract
Generally, fillers are added in thermoplastic polymers to enhance their mechanical
properties, reduces cost, and improve appearance of the composites. In this paper,
high-density polyethylene (HDPE) was compounded with different waste fillers such
as silica, kaolin, calcium carbonate, and fly ash with an aim to compare the com-
posite performance and effects of different filler loading. Tensile test, impact test,
burning test, and water absorption were used to characterize the HDPE composites.
Results show that addition of fillers in HDPE improved the tensile modulus, reduced
14% of the burning rate, and increased water absorption over time, compared to
unfilled HDPE. However, tensile strength, impact strength, and elongation at break
reduced by 30–50%, 50%, and 98% compared to the neat HDPE, respectively. Addi-
tion of calcium carbonate in HDPE shows the best fire retardant, the highest tensile
strength, and the longest elongation at break as compared to composites with silica,
fly ash, and kaolin.

Keywords  Polymer composites · HDPE · Waste filler · Tensile · Impact

Introduction

There has been an increased interest in composite materials over the last two dec-
ades, especially composites produced from polymeric matrices. But most of the
commonly used resins are obtained from the petroleum industry, which generally
generate large volumes of wastes and contaminates [1, 2]. This often pose signifi-
cant threats to the environment, but efforts have been concerted on development of

* M. Mariatti
mariatti@usm.my
1
School of Materials and Mineral Resources Engineering, Universiti Sains Malaysia,
14300 Nibong Tebal, Penang, Malaysia
2
Department of Material Science and Engineering, İzmir Katip Çelebi University, Izmir, Turkey

13
Vol.:(0123456789)
6778 Polymer Bulletin (2021) 78:6777–6795

bio-composites, where at least one of the components is derived from renewable


resources [3, 4]. Usually, the resin may be derived from renewable resources, or
reinforcing fillers derived from renewable resources may be incorporated into petro-
leum-based resins to produce environmentally benign composites. Nevertheless, the
performance of composites in different applications is largely dependent on factors
such as the intrinsic properties of individual components, their interactions, and fab-
rication processes [5].
Currently, different thermoplastics are being investigated, and reinforcing fillers
are commonly incorporated into these plastics to enhance their physical, chemi-
cal, thermal, and dynamic mechanical performance [6–11]. However, reports have
indicated that poor compatibility between these thermoplastics which are inherently
hydrophobic and hydrophilic bio-fillers often result in reduced mechanical perfor-
mance due to aggregate formation during processing [3, 12]. Therefore, the recent
trend of research is the incorporation of fine mineral powders to increase the contact
surface area of matrices and fillers in different applications. Interestingly, the surface
of these particles can be modified to enhance their compatibility with polymeric
matrices [5].
Particularly, the global desire for sustainable development have motivated the
addition of waste mineral fillers into thermoplastics, as a way of utilizing wastes
for economic values, and to partially solve the environmental challenges associated
with waste generation [13, 14]. Notably, these waste fillers may be obtained as by-
products or deviations from certain industrial processes and manufacturing methods
[5]. Some of the notable fillers obtained as by-products or deviations from the indus-
trial processes are fly ash, silica, kaolin, and calcium carbonate. Fly ash is the resi-
due generated from coal in the burning chamber of an electrical power plant, while
silica waste is generated from the smelting process in the silicon, ferrosilicon, and
glassmaking industries. In contrast, kaolin waste is obtained from the paper-coating
industry, whereas calcium carbonate is a waste from the glossy paper and cultured
marble industries. Being mere wastes, these materials could create a nuisance to the
environment if not properly utilized or managed. Interestingly, these materials can
be used as fillers in polymeric composites to fabricate products for different applica-
tions. Table 1 shows a summary of previous work on various type of mineral fillers
used with thermoplastics. Notably, different types of thermoplastics were used such
as low-density polyethylene (LDPE) [15], high-density polyethylene (HDPE) [16],
polycarbonate (PC) [17], PE [18], nylon 6 [19], and polypropylene (PP) [20].
Among the thermoplastic materials, HDPE exhibits excellent characteristics such
as lightweight, flexible, low cost, chemical-resistant to most acidic and chemical
agents, high resistance to the impact, translucent but not clear and easy to process
[25, 26]. These salient properties of HDPE make it suitable for various applications
especially as a protection of electrical power, such as in TV cables and telecommu-
nication cables like fibre-optic cable. HDPE is also used for daily applications such
as water bottle, toys, chemical containers, and pipe systems. However, HDPE has
low toughness, has poor weather resistance, and is flammable, which tends to limit
its extended application. As summarized in Table 1, some of these properties may be
improved by the incorporation of different mineral fillers. However, to the best of the
authors knowledge, there is no extensive report on the comparison of performance

13
Table 1  Summary of previous works on the addition of particulate fillers into thermoplastic
Type of Thermoplastic Properties enhanced Properties reduced References
fillers

Talc LDPE Stabilization torques, tensile modulus, and tensile strength, Tensile strain and impact strength [15]
Calcium LDPE impact strength and stabilization torques tensile modulus, tensile strain, tensile
carbon- strength,
ate
Polymer Bulletin (2021) 78:6777–6795

Wollas- Recycle PC/ tensile strength, and flexural strength elongation at break, notched Izod impact [17]
tonite virgin PC strength
Talc Isotactic Tensile strength, tensile modulus, flexural modulus, and impact strength Elongation at break, glass transition tem- [18]
Mica PE perature

Talc/mica
Kaolin Nylon 6 Tensile modulus, flexural modulus, tensile strength, flexural strength, and impact strength Notched impact strength [19]
Talc PP Tensile Tensile [20, 21]
Bentonite Modulus, flexural strength, elongation
strength, and flexural at break
modulus
Silica HDPE Tensile strength, Young’s modulus, and flexural strength Heat of fusion, percent of crystallinity [22]
Calcium HDPE Melting temperature, percent of crystallinity, heat of fusion, impact resistance, yield stress, and – [16]
carbon- Young’s modulus
ate
Calcium HDPE Young’s modulus and yield stress impact strength [23]
carbon-
ate
Fly ash HDPE Thermal, tensile strength, Young’s modulus, and flexural properties Elongation and impact resistance [24]
6779

13
6780 Polymer Bulletin (2021) 78:6777–6795

among the notable mineral waste fillers, when incorporated into HDPE to produced
reinforced composites. Therefore, in this contribution, waste fillers such as silica,
kaolin, calcium carbonate, and fly ash were compounded in HDPE at varying filler
loading. The performance of each filled HDPE composites was investigated through
various testing, including tensile, impact, burning test, and water absorption. In
addition, observation of the samples fractured surfaces, particle shape, and particle
size was used to correlate the characterization results with the intrinsic properties of
the filled HDPE composites.

Materials and methods

Materials

High-density polyethylene (HDPE) resin, blowing grade (MW 50.944—0.955  g/


cm3), was supplied by Lotte Chemical Titan (M) Sdn. Bhd. The HDPE has a melt
flow index (MFI) of 4 g/10 min @ 5 kg/190°. Waste fillers such as silica, kaolin, fly
ash, and calcium carbonate were used in this study. Fly ash powders were supplied
by Tenaga Nasional Berhad Power Station while, silica, kaolin, and calcium carbon-
ate were obtained from Silbelco Malaysia Sdn Bhd, Tinex Kaolin Corporation Sdn
Bhd, and Malaysian Calcium Corporation Sdn. Bhd, respectively.

Particle size analysis

The particle size distribution of mineral filler was measured by dry dispersion
(Rodos system, Sympatec GmbH, Clausthal-Zellerfeld, Germany) and laser dif-
fraction measurements (Helos system, Sympatec GmbH, Clausthal-Zellerfeld, Ger-
many). The mean particle diameter or d50 was taken based on the cumulative distri-
bution graph, and the span value was calculated based on Eq. 1 [27].
Dv0.9 − Dv0.1
Span = (1)
Dv0.5

where Dv0.9 is the particle diameter at 90% cumulative size, Dv0.1 is the particle
diameter at 10% cumulative size, and Dv0.5 is the particle diameter at 50% cumula-
tive size. For example, if the Dv0.9 is 20 nm, this indicates that 90% of the filler has
a size of 20 nm or smaller. Hence, span is basically an indication of how far the 90%
and 10% points are apart.

Composite preparation

The HDPE composites were prepared by melt mixing, using a heated Labtech LRM-
S-110/3E + W Scientific Laboratory Two Roll Mixer at 190 °C with the rotor speed
of 10 rpm. The HDPE pellets were placed between the two rolls for 10 min with the
nip gap of 1 mm. The filler was added at varying volume ratio, as shown in Table 2.

13
Polymer Bulletin (2021) 78:6777–6795 6781

Table 2  Designation of the prepared HDPE reinforced with waste filler specimens


Sample designation Neat Silica (Vol%) Fly ash (Vol%) Calcium car- Kaolin (Vol%)
HDPE bonate (Vol%)
(Vol%)

HDPE 100 – – – –
HDPE/S5 95 5 – – –
HDPE/S15 85 15 – – –
HDPE/F5 95 – 5 – –
HDPE/F15 85 – 15 – –
HDPE/C5 95 – – 5 –
HDPE/C15 85 – – 15 –
HDPE/K5 95 – – – 5
HDPE/K15 85 – – – 15

The compound was sheeted out at room temperature at a nip gap of 0.5 mm. The
sheeted compound was cut, preheated for 6 min and pressed for 2 min in a GOTECH
(GT-7014-A30C) hot press machine at 190  °C. Different mould types were used
according to the standard for each testing. The pressed composite was cooled for
another 3 min before being removed from the mould.

Mechanical properties

Tensile properties were evaluated according to the ASTM D638 standard, using an
Instron Universal Testing Machine (model 3366). Five specimens were prepared for
each compound by cutting into a dumbbell shape. A crosshead speed of 50 mm/min
was used for the testing. Tensile strength, elongation at break, and tensile modulus
results were recorded from the test.
Notched Izod impact tests were carried out using a Zwick Impact Tester according
to ASTM D256-92 standard. The testing was run at room temperature by using 7.5 J
pendulum hammer. Ten specimens with dimension of 60 mm × 12 mm x 3 mm (long x
width x thickness) were prepared for the testing.

Morphological characterization

The tensile-fracture surface, impact-fracture surface and the morphology of the waste
filler particle such as particle size and shape were observed using field emission scan-
ning electron microscope (FESEM) (LEO SUPRA 35VP, Carl Zeiss, and Germany).
The aspect ratio of the waste fillers was calculated as the average value of 100 selected
particles from the SEM micrograph, using an ImageJ-1.47c software.

13
6782 Polymer Bulletin (2021) 78:6777–6795

Burning test

Five specimens were cut from the sheet material according to ASTM D 635 with
the bar specimens of 125 mm × 12.5 mm × 3.0 mm (long x wide x thickness). Two
lines were marked at 25 mm and 100 mm from the end and ignited. The specimen
and burner were set up according to the standard with the height of the blue flame
is 20 mm. The specimens were exposed to the flame for 30 s without changing the
burner position and were removed from the burner after 30 s or once the combustion
front reached the 25 mm mark (for the case less than 30 s). Time was taken when the
combustion front reached the 25 marks. The linear burning rate, V, in millimeters
per minute was calculated using Eq. 2:
V = 60L∕t (2)
in which V is the linear burning rate in mm/minute, L is the damaged length, in
millimeters, t is time, in seconds. Since the flame front passed the 100 mm mark for
every specimen, the L is 75 mm.

Water absorption

Three specimens for neat HDPE and HDPE composites were used for measurement
of water absorption according to ASTM D570. Specimens were dried at 80 °C over-
night before immersed in distilled water at room temperature (23 °C) and 100 °C. At
different time interval, specimens were dried with tissue and weight using Mettler
balance type AJ150 as soon as specimens were taken out from water. The percent-
age of water absorption ( Mt ) was calculated by Eq. 3:
Wt − Wo
Mt = × 100% (3)
Wo

where Wo is the initial weight of the sample prior to immersion, while Wt represents
the sample weight, after immersion.

Results and discussion

Characterization of waste fillers

The properties of thermoplastic composites can be significantly influenced by the


characteristics of the reinforcing fillers. Specifically, features such as particle size,
span, and shape of the fillers are among the notable characteristics that could influ-
ence the composites properties. The features and general properties of the fillers used
in this study (fly ash, calcium carbonate, silica, and kaolin) are presented in Table 3.
Generally, particle size could play a role in improving the interaction between the
filler and the matrix. Usually, smaller particle sized fillers will exhibit larger surface
area, thereby facilitating better interaction between the filler and matrix [24]. The
result obtained from the average particle size (d50) indicates that the particle sizes of

13
Polymer Bulletin (2021) 78:6777–6795 6783

Table 3  General properties of Filler Particle Span ratio Shape Colour


the fillers used in this study size, d50
(µm)

Fly ash 4.78 2.30 Spherical Light


brown-
ish grey
Calcium carbonate 9.37 1.59 Irregular White
Silica 8.90 3.10 Angular White
Kaolin 13.42 1.66 Plate–like White

all the fillers fall within the micron range, with values below 15 µm. Notably, fly ash
has the smallest particle size (d50 = 4.78 µm), whereas kaolin has the largest particle
size (d50 = 13.42 µm). The particle size distribution illustrated in Fig. 1 shows that
silica exhibits a bimodal distribution, as confirmed by the two distinct peaks in the
particle size distribution curve of silica in Fig. 1. On the other hand, the single peak
curves of fly ash, calcium carbonate, and kaolin indicates that they exhibit unimodal
particle size distribution. This shows that their particle sizes are distributed over a
small particle size range. Furthermore, the particle size of fly ash showed a positive
skewness distribution in the range of 1.5 µm to 23 µm. The skewness indicates that
there are higher amounts of fine particles that are below 5 µm. Notably, the particle
size of calcium carbonate presented in Fig. 1 (1.5 µm to 27 µm) is almost in a similar
range with fly ash (1.5 µm to 23). However, calcium carbonate which has an average
particle size of 9.37 µm has the smallest span, as presented in Table 3. This indicates
that calcium carbonate has homogeneous particle size distribution. The span values

1.6
Fly Ash
1.4 silica
Calcium Carbonate
Density Distribution, dQ3 (X)

1.2
Kaolin
1.0

0.8

0.6

0.4

0.2

0.0
0 10 20 30 40 50
Particle Size (µm)

Fig. 1  Particle size distribution of waste filler

13
6784 Polymer Bulletin (2021) 78:6777–6795

of the fillers presented in Table 3 show that silica and fly ash have larger span val-
ues. This suggests that their particle size distribution is not uniform.
Scanning electron microscope (SEM) micrographs of the fillers incorporated into
the HDPE are shown in Fig. 2. The image of fly ash in Fig. 2a is generally spherical,
with various particle sizes, while Fig. 2b shows the SEM image of irregular calcium
carbonate with sharp edges. Figure 2c shows the angular shape of silica particle with
sharp indentations, whereas kaolin particles with coarser morphology compared to
the other samples is shown in Fig. 2d. The higher magnification image of the kaolin
particles reveals that there are flat particles stacked together, which forms a stiff hex-
agonal plate with flat-faced edges called a booklet of kaolin, with a size larger than
10 µm [28].

Mechanical properties

Tensile test was used to determine the tensile properties of the neat HDPE and filled
HDPE composites. Figure  3 shows plastic deformation through the entire length
of the specimens during the tensile testing. Specifically, stress whitening can be
observed on the neat HDPE and all the filled HDPE composites. Neck propagation
extended to a certain length, and whitening zone was observed along the narrow
length for neat HDPE, HDPE/F5, HDPE/C5, HDPE/S5, and HDPE/C15 specimens.

5 µm 5 µm
(a) (b)

5 µm 5 µm
(c) (d)

Fig. 2  Filler particle a fly ash, b calcium carbonate, c silica, and d kaolin. The inserted figure indicates
high magnification image

13
Polymer Bulletin (2021) 78:6777–6795 6785

HDPE

HDPE/F5

HDPE/F15

HDPE/C5

HDPE/C15

HDPE/S5

HDPE/S15

HDPE/K5

HDPE/K15

Fig. 3  Photograph of stress whitened neat HDPE and HDPE/mineral composites strained at a crosshead
speed of 50 mm/min

Generally, whitening occurs as a result of plastic deformation when load is applied


to the sample. However, it had also been reported that whitening could be due to
the presence of cracks, voids, or exposure of filler [29]. As can be seen in Fig. 3,
the specimens were able to stretch when less amount of filler was added into the
HDPE, and they were continuously elongated by the propagation of necking area
along with the gauge which is an indication of a stable necking zone [30]. In con-
trast, shorter whitened regions were observed at the break and necking area within
the gauge length for the rest of the specimens containing higher (15%) filler loading,
except HDPE/C15. This indicates that higher filler loading generally obstructed the
stretching of the composite. Hence, stress whitening can be mainly observed around
the break and necking area [31].
Figure  4 shows the stress–strain curves for neat HDPE, calcium carbonate, and
kaolin filled HDPE composites. It should be noted that composite samples with
calcium carbonate and kaolin are chosen for stress–strain comparison based on the
obvious contrast in the samples of these composites as shown in Fig. 3. In Fig. 4,
large area under the curve indicates higher toughness of neat HDPE as compared to
filled HDPE composites. At lower filler loading (5 wt%), calcium carbonate filled
HDPE composite shows higher toughness than that of 15 wt%, and the toughness
was reduced the filler loading was increased. Stress drop after the yield point and
remained constant where the propagation occurs through the gauge length. Then
stress gradually increased until the specimen rupture. In contrast, a different trend
can be seen when kaolin was added into the HDPE, as the area under the curve
became smaller especially when the amount of filler was increased. After the yield
point, stress drops drastically, and fracture strain decreased as the filler content
increased. This shows that addition of filler into the HDPE reduced the strength and
elasticity of HDPE composite and consequently decreased the toughness.

13
6786 Polymer Bulletin (2021) 78:6777–6795

25
HDPE
40 HDPE/C5
HDPE/C15 20

HDPE/K5
30 HDPE/K15
15

Stress (MPa)
Stress (MPa)

20
10

10 5

0 0
0 10 20 30 40
0 200 400 600 800
Strain (%)
Strain (%)

(a) (b)

Fig. 4  Stress–strain curves of a HDPE and filled HDPE composite with 5 and 15% of calcium carbonate
and kaolin waste filler and b enlarged portion of the stress–strain curve at the strain of about 40%

The toughness of the HDPE correlates to a brittle–ductile characteristic, which


is confirmed by the value of elongation at break. Higher elongation is an indi-
cation that the polymer is more ductile, whereas brittle polymers show lower
elongation [32]. Figure 5 shows the elongation at break of neat HDPE and filled
HDPE composites. Addition of kaolin fillers into HDPE reduced more than 97%
of the elongation at break compared with the neat HDPE (868.5 ± 8.5%). This
suggests that the large particle size and plate-like structure of kaolin (Table 3 and
Fig.  2d) might have reduced the flexibility of the composite, even at lower per-
centage of filler. Generally, for all the composite categories, addition of 15% filler
in HDPE reduced more than 50% of the elongation, except for calcium carbon-
ate filler. Notably, the elongation at break of calcium carbonate filled HDPE only
reduced by about 25%. Siddique et al. [33] reported that increased filler loading
decreased the value of elongation at break due to the rigid interphase between
the filler and HDPE. This resulted in incredibly low plastic deformation, and the
composite breaks at low strain as observed herein for HDPE/F15 and HDPE/S15
and HDPE/K15 composites.
Maximum amount of stress to withstand stretching before break determines the
tensile strength of the polymer. Addition of different fillers into HDPE resulted

13
Polymer Bulletin (2021) 78:6777–6795 6787

Elongation at Break
1000 Tensile Strength 40

35
800
30
Elongation at Break (%)

Tensile Strength (MPa)


600 25

20
400
15

200
10

5
0

0
PE % % % % % % % %
HD Ash 5 Ash 15 onate 5 nate 15 Silica 5 ilica 15 aolin 5 olin 15
l y y r b o S K a
F Fl Ca Car
b K
cium ium
Cal Calc
Filler Loading (Vol%)

Fig. 5  Effect of filler loading on the elongation at break of HDPE and filled HDPE composite

in reduction of tensile strength compared to the neat HDPE (36.2 ± 0.9 MPa) as


shown in Fig. 5. Among the composites, the sample with 5 wt% calcium carbon-
ate loading shows the highest tensile strength value (25.9 ± 1.0  MPa) followed
by silica, fly ash, and kaolin. The filled HDPE composites show further tensile
strength reduction at 15 wt % filler loading. Reduction of the tensile strength is
mostly due to the poor dispersion and adhesion between the filler and polymer
matrix especially as the filler loading increased. This is because poor wettability
and adhesion generally tend to reduce the filler-matrix contact, thereby disrupting
the effective transfer of stress within the composites [34]. This suggests that sur-
face modification of mineral fillers is important to facilitate good dispersion and
adhesion in mineral filler filled polymer composites.
Figure 6 shows the fractured surface morphology of HDPE/mineral filler com-
posites. As can be seen in the figure, calcium carbonate fillers were debonded and
pulled-out from the HDPE matrix during the tensile test (as shown in Fig.  6a).
This is an indication of weak interaction between the filler and matrix which is
believed to have contributed to the decrease in tensile strength of the HDPE/C5
composite, compared to neat HDPE [35]. On the other hand, the fractured surface
morphology of the HDPE/C15 composite is presented in Fig. 6b. It can be seen in
the image that there are crack strips and tails between the filler and matrix. Con-
tinuous elongation seems to have induced tail structure in the direction of crack
propagation, which indicates that poor interfaces allowed the matrix to be eas-
ily stretched. This explains the reason for the high elongation at break in HDPE/

13
6788 Polymer Bulletin (2021) 78:6777–6795

Filler pull-out
Tail structure

1 µm

(a) (b)

(c) (d)

Fig. 6  SEM micrographs showing the fractured surface morphology of a HDPE/C5, b HDPE/C15, c


HDPE/S15, and d HDPE/K15

C15 as shown in Fig. 5. On the other hand, the poor interface and sharp edges of
calcium carbonate and silica fillers might have contributed to observed reduction
in tensile strength of 15 wt% calcium carbonate and 5 wt% silica composites. Fig-
ure 6c, d shows the fractured surfaces of HDPE/S15 and HDPE/K15. It is evident
from both images that the samples do not exhibit sufficient plastic deformation
prior to failure, especially compared to HDPE/C5. Significantly, the surface of
the HDPE/K15 composite is smoother, compared to HDPE/S15 which indicates
brittle fracture of HDPE/K15. This resulted in low strain at break of HDPE/K15,
compared to HDPE/S15. Notably, HDPE/S15 shows slightly stretchable behav-
iour as confirmed by the deformation lines on the fractured surface of HDPE/S15.
As such, the elongation at break of HDPE/S15 is 10% higher compared to HDPE/
K15.
Although the tensile strength of the filled HDPE composites is lower than
the neat HDPE, the tensile modulus of the filled HDPE composites increased
as shown in Fig. 7. This shows that filled HDPE composites could stand higher
stress over the amount of strain at the early stage of stress application to the com-
posite materials. Increase in the modulus could be due to the rigidity of fillers

13
Polymer Bulletin (2021) 78:6777–6795 6789

Impact Strength
70 Tensile Modulus 1400

60 1300
Impact Strength (J/m2)

Tensile Modulus (MPa)


50
1200
40
1100
30
1000
20

10 900

0 800
E % % % % % % % %
HDPy Ash 5 Ash 15 onate 5 nate 15 Silica 5 ilica 15 aolin 5 olin 15
Fl Fly m Carb Carbo S K Ka
iu um
Calc Calci
Filler Loading (Vol%)

Fig. 7  Effect of filler loading on the tensile modulus and impact strength (notched Izod Charpy test) of
HDPE and filled HDPE composite

added into the HDPE. This is because the addition of rigid filler improves the
stiffness of HDPE composites, as the fillers tend to restrict the mobility and plas-
tic deformation of the HDPE [36]. Hence, the tensile modulus increased with
increasing amount of filler content. This generally follow the simple rule of mix-
ture for composite materials [34].
Figure 7 shows the results of notched Izod impact test of neat HDPE and filled
HDPE composites. Obviously, the addition of filler into the HDPE results in a
decrease in the impact strength value, compared to neat HDPE (54.0 ± 9.2 kJ/m2).
HDPE/F5 specimen shows the highest impact strength with a value of 21.6 ± 6.3 kJ/
m2 while the lowest value was obtained for HDPE/K5 (9.5 ± 1.3 kJ/m2). In general,
interfacial adhesion and filler properties affect the impact strength. Obvious reduc-
tion of impact strength as compared to the neat HDPE shows poor adhesion between
the filler and matrix which results in less resistance to crack propagation through the
filler and matrix interface [37]. Fillers that are weakly bonded create voids between
the interface, thereby limiting proper transfer stress [38]. This will lead to undesir-
able initiation and propagation of crack, which will invariably result in fracture and
reduced energy absorption. In another vein, the particle size and the particle shape
can significantly influence the impact properties of the composite.
Figure 8 shows the fractured surfaces of selected impact samples. Obviously,
morphology of impact fracture surface is brittle compared to that of the ten-
sile fractured surface (Fig.  6). The higher impact strength shown by HDPE/F5
is probably due to the spherical shape and the small particle size of the fly ash
(Table  3), which might have facilitated good dispersion of fly ash in the poly-
mer matrix. Small particle size as shown in Fig. 8a results in better interaction
between the filler and matrix. Increased filler loading produced several voids

13
6790 Polymer Bulletin (2021) 78:6777–6795

Better interaction of small filler Filler pull-out Debonding

20 µm Debonding of large particle 20 µm Filler aggregation

(a) (b)
Stress concentration point

Crack propagation

20 µm
Sphere shape filler Sharp-edge filler
(c) (d)

Fig. 8  SEM micrographs showing the notched Izod Charpy fractured surface morphology of a HDPE/
F5, b HDPE/F15, c HDPE/K15, and d illustration of a fracture profile during crack propagation

and increased debonding sites. Hence, in filler–matrix interface separation, fill-


ers pull-out significant aggregation which can be observed in the fractured sur-
face in Fig.  8b, which might have contributed to the observed impact strength
reduction [39, 40]. Nevertheless, it is noteworthy that no significant fracture was
observed on the filler.
In this impact test, notch has already been used to create cracks such that
the stress can be propagated, and the crack is driven until it encounters the fill-
ers. As illustrated in Fig. 8d, flat and sharp-edged particles, in this case calcium
carbonate, kaolin, and silica, respectively, acted as flaws that create weak areas
to transfer the stress, with high tendency to initiate cracks. The sharp edges of
the particles are the weakest point where it creates stress concentration points.
The stress concentrated around the filler or at the end of the filler would lead to
the crack deflection, debonding on inclusion/matrix interface, and cracking the
matrix. Therefore, this will trigger undesirable failure via micro-crack forma-
tion, which will rapidly propagate to macroscopic failure [41]. Figure 8c shows
that the fractured surface of HDPE/K15 reveals less fibril formation compared
to HDPE/C5, which indicates higher brittle behaviour as filler loading increases.

13
Polymer Bulletin (2021) 78:6777–6795 6791

Table 4  Summary of burning Sample Samples flame Flame reach Burning rate


rate for HDPE and filled HDPE reach 25 mm (y/n) 100 mm (y/n) (mm/min)
composite
HDPE Y Y 23.86 ± 1.83
HDPE/S15 Y Y 20.45 ± 1.12
HDPE/F15 Y Y 23.41 ± 2.20
HDPE/C5 Y Y 21.13 ± 0.25
HDPE/C15 Y Y 18.40 ± 0.29
HDPE/K15 Y Y 19.25 ± 0.97

Burning Test

The results of burning rate for HDPE and selected filled HDPE composites in a hori-
zontal position are shown in Table 4. In this measurement, time was taken once the
flame reached 25  mm mark and when the flame reaches the 100  mm mark for all
the specimens. The sample with the fastest burning rate was found to be neat HDPE
followed by HDPE/F15, and the lowest was found to be HDPE/C15. Faster burning
rate indicates poor flammability of polymer and vice versa. The result in Table  4
indicates that addition of filler improved the flammability of HDPE. Fly ash filler
can dilute polymeric fuel. The chemical inert nature of the fly ash hinders the fly
ash from achieving solid-phase intumescent or gas-phase flame retardancy which
makes it as a poor filler for flame retardance [42]. HDPE/C15 shows the slowest rate
of burning. Being a carbonate material, the calcium carbonate ­(CaCO3) decomposed
and produced copper oxide (CaO) and carbon dioxide (­CO2) gas. The C ­ O2 then
dilutes the flammable gas and delay the decomposition of the composite [43]. Com-
paring the burning rate between HDPE/C5 and HDPE/C15, it is found that the burn-
ing rate became slower as calcium carbonate filler loading increased. This result is
similar with the report by Jancar et al. [44] where they suggested the use of higher
calcium carbonate filler loading as an effective way to improve the flammability of
thermoplastic.

Fig. 9  High flame characteristic of a neat HDPE, b HDPE/C15, and c HDPE/K15

13
6792 Polymer Bulletin (2021) 78:6777–6795

Figure  9 shows the burning of neat HDPE and filled HDPE composite. All the
HDPE and filled HDPE composites show intense burning with high flames. Based
on the observation during the burning process, neat HDPE burned with fast drip
whereas for the filled HDPE composites, the drips are much slower which indicates
that the fillers helped to delay the flame spread. In combustion front zone, there are
black area which suggests char formation to prevent the composite from continuous
burning.

Water absorption

The water absorption data is important to understand the  performance of materi-


als  in water or humid environments during application. Figure  10 shows the per-
centage of water absorption of neat HDPE and 5 wt% filled HDPE composites at
temperatures of 23  °C and 100  °C. As shown in the figure, water absorption of
filled HDPE composites increased by about 29–57% compared to the neat HDPE
at 23 °C. This shows that addition of fillers increased the rate of water penetration
into the composite structure. It is found that HDPE/K15 composite absorbed more
water followed by HDPE/F15, HDPE/C15 and HDPE/S15 composite. Poor interac-
tion between filler and HDPE matrix due to incomplete wettability produced higher
voids. This could lead to high water accumulate at the filler-matrix interface [45].
As shown in Sect. 3.3, poor interaction between the filler and matrix also reduced
the tensile strength and impact strength of the composites.
Furthermore, agglomeration of waste filler particles results in the occurrence of
more defects between the filler and matrix which permits significant water penetra-
tion [46]. As expected, water absorption percentage of composites was far greater
in the case of elevated temperature (100  °C) compared to the room temperature

1.6
1.4 23ºC 100ºC
Water Absorpon (%)

1.2
1.0
0.8
0.6
0.4
0.2
0.0

Sample

Fig. 10  Water absorption percentage of neat HDPE and filled HDPE composite at 23 °C and 100 °C

13
Polymer Bulletin (2021) 78:6777–6795 6793

(23 °C). At high temperature, the debonding between filler and matrix occurs at the
interface due to swelling, and this invariably generate voids which act as a reservoir
for water absorption. Neat HDPE shows small increment in water absorption per-
centage since the structure does not contain any chemical bond that can be easily
hydrolysed. Hence, neat HDPE absorbed little water and essentially unaffected by
aging in water.

Conclusions

Efforts to utilize waste material of fly ash, kaolin, calcium carbonate, and silica as a
filler in thermoplastic materials can be realized by mixing these waste materials into
HDPE using a two-roll mill method. The investigation on addition of waste filler
into HDPE properties showed the following conclusions:

1 Addition of different amount of waste fillers into HDPE transform the polymer
from ductile to brittle behaviour.
2 Addition of fillers into HDPE reduced the tensile strength, impact strength, and
elongation at break of the neat HDPE with reduction of 30–50%, 50%, and 98%,
respectively. Particle shape, particle size, interaction between filler and matrix,
and amount of loading effect the properties of the filled HDPE composite.
3 Tensile modulus of filled HDPE composite improved up to 36% from the original
modulus of neat HPDE by enhancement rigidity and restricted the mobility of
polymer by addition of the fillers. The modulus of filled HDPE composite also
followed the trend of Rule of Mixture as increase the filler loading.
4 Burning rate of the filled HDPE composite reduced more than 14% compared
to the neat HDPE. Calcium carbonate is the most efficient filler in reducing the
flame spread. Increase in the calcium carbonate fillers loading enhances the flame
retardancy.
5 Addition of waste fillers in HDPE composite increased water absorption at 23 °C
and 100 °C. Neat HDPE shows small increment in water absorption percentage
if compared to composite samples.

Acknowledgements  We are very grateful to the Malaysian Ministry of Education for awarding us a Fun-
damental Research Grant (MRSA with grant no. 6071284) and Universiti Sains, Malaysia, that made this
study possible.

References
1. Akindoyo JO et al (2020) Simultaneous impact modified and chain extended glass fiber reinforced
poly (lactic acid) composites: mechanical, thermal, crystallization, and dynamic mechanical perfor-
mance. J Appl Polym Sci 138(5):49752
2. Akindoyo JO, Ismail NH, Mariatti M (2019) Performance of poly(vinyl alcohol) nanocomposite
reinforced with hybrid TEMPO mediated cellulose-graphene filler. Polym Test 80:106140

13
6794 Polymer Bulletin (2021) 78:6777–6795

3. Beg MDH et  al (2018) Characterization of polyamide 6.10 composites incorporated with micro-
crystalline cellulose fiber: Effects of fiber loading and impact modifier. Adv Polym Technol
37(8):3412–3420
4. Orue A et al (2016) The effect of alkaline and silane treatments on mechanical properties and break-
age of sisal fibers and poly(lactic acid)/sisal fiber composites. Compos Part A Appl Sci Manuf
84:186–195
5. Essabir H et al (2017) A comparison between bio- and mineral calcium carbonate on the properties
of polypropylene composites. Constr Build Mater 134:549–555
6. Akindoyo JO et  al (2018) Impact modified PLA-hydroxyapatite composites–thermo-mechanical
properties. Compos Part A Appl Sci Manuf 107:326–333
7. Akindoyo JO et al (2019) Oxidative induction and performance of oil palm fiber reinforced polypro-
pylene composites—effects of coupling agent and UV stabilizer. Compos Part A Appl Sci Manuf
125:105577
8. López OV et al (2018) Processing–properties–applications relationship of nanocomposites based on
thermoplastic corn starch and talc. Polym Compos 39(4):1331–1338
9. Passaretti MG et  al (2019) Biocomposites based on thermoplastic starch and granite sand quarry
waste. J Renew Mater 7(4):393–402
10. Kamerling S, Schlarb AK (2018) Enhancing polymer based tribocompounds using mineral fillers
with an energy absorbing material transformation. In Proceedings of Asia International Conference
on Tribology 2018. Malaysian Tribology Society
11. Sormunen P, Kärki T (2019) Compression molded thermoplastic composites entirely made of recy-
cled materials. Sustainability 11(3):631
12. Yi M et al (2016) Flexible fiber-reinforced composites with improved interfacial adhesion by mus-
sel-inspired polydopamine and poly(methyl methacrylate) coating. Mater Sci Eng C 58:742–749
13. Yao Z et al (2014) Mechanical and thermal properties of polypropylene (PP) composites filled with
CaCO3 and shell waste derived bio-fillers. Fibers Polym 15(6):1278–1287
14. Väisänen T et al (2016) Utilization of agricultural and forest industry waste and residues in natural
fiber-polymer composites: a review. Waste Manag 54:62–73
15. Shamsuri AA, Sumadin ZA (2018) Influence of hydrophobic and hydrophilic mineral fillers on pro-
cessing, tensile and impact properties of LDPE/KCF biocomposites. Compos Commun 9:65–69
16. Tanniru M, Misra R (2005) On enhanced impact strength of calcium carbonate-reinforced high-
density polyethylene composites. Mater Sci Eng A 405(1–2):178–193
17. Tarade RS, Mahanwar PA (2019) Mechanical Thermal and morphologicalproperties of recy-

cled and virgin PC/wollastonite composite and its compatibilization by SBC. J Mater Environ Sci
10:357–366
18. Mittal P et al (2019) Polypropylene composites reinforced with hybrid inorganic fillers: morphologi-
cal, mechanical, and rheological properties. J Thermoplast Compos Mater 32(6):848–864
19. Baioumy MGS et al (2019) Surface-treated kaolin minerals as a complement or substitute to glass
fibers in thermoplastics. Polym Eng Sci 59(S1):E330–E338
20. Sarikanat M et al (2019) The effect of various minerals on sound transmission loss and mechanical
properties of polypropylene. Acta Phys Pol, A 135(5):1055–1057
21. Rothon R (1999) Mineral fillers in thermoplastics: filler manufacture and characterisation. Springer,
Berlin, pp 67–107
22. Ahmed T, Mamat O (2011) The development and characterization of HDPE-silica sand nanoparti-
cles composites. In: 2011 IEEE Colloquium on Humanities, Science and Engineering. IEEE
23. Lazzeri A et al (2005) Filler toughening of plastics. Part 1—The effect of surface interactions on
physico-mechanical properties and rheological behaviour of ultrafine CaCO3/HDPE nanocompos-
ites. Polym 46(3):827–844
24. Ahmad I, Mahanwar PA (2010) Mechanical properties of fly ash filled high density polyethylene. J
Min Mater Charact Eng 9(03):183
25. D’Amato M et  al (2012) High performance polyethylene nanocomposite fibers. Exp Polym Lett
6(12):954–964
26. Sewda K, Maiti S (2013) Dynamic mechanical properties of high density polyethylene and teak
wood flour composites. Polym Bull 70(10):2657–2674
27. Christensen K, Pedersen G, Kristensen H (2001) Preparation of redispersible dry emulsions by
spray drying. Int J Pharm 212(2):187–194
28. Ivanić M et  al (2015) Mineralogy, surface properties and electrokinetic behaviour of kaolin clays
from the naturally occurring pegmatite deposits. Geol Croati 68(2):139–145

13
Polymer Bulletin (2021) 78:6777–6795 6795

29. Wong M et al (2004) Study of surface damage of polypropylene under progressive loading. J Mater
Sci 39(10):3293–3308
30. Schrauwen BAG (2003) Deformation and failure of semicrystalline polymer systems: influence of
micro and molecular structure. PhD Thesis, Eindhoven University of Technology, Eindhoven, The
Netherlands, ISBN-90-386-2914-1
31. Mazidi MM, Aghjeh MR, Abbasi F (2012) Relationship between microscopic deformations and
macroscopic mechanical response of SAN/PB-g-SAN blends via interparticle distance concept. J
Polym Res 19(8):9928
32. Shebani A et  al (2018) The Influence of LDPE Content on the Mechanical Properties of HDPE/
LDPE Blends. Res Dev Mater Sci 7(5):791–797
33. Siddique S et al (2019) Structural and thermal degradation behaviour of reclaimed clay nano-rein-
forced low-density polyethylene nanocomposites. J Polym Res 26(6):154
34. Nurdina A, Mariatti M, Samayamutthirian P (2009) Effect of single-mineral filler and hybrid-
mineral filler additives on the properties of polypropylene composites. J Vinyl Addit Technol
15(1):20–28
35. Thomas S, Zaikov GE (2008) Polymer nanocomposite research advances. Nova Publishers, New
York
36. Wypych G (2010) Handbook of fillers, 4th edn. Chem Tec Publishing, Toronto
37. Shebani A et al (2016) Effects of Libyan kaolin clay on the impact strength properties of high den-
sity polyethylene/clay nanocomposites. Int J Compos Mater 6(5):152–158
38. Muthuraj R et al (2016) Influence of processing parameters on the impact strength of biocompos-
ites: A statistical approach. Compos Part A Appl Sci Manuf 83:120–129
39. Fu S-Y et al (2008) Effects of particle size, particle/matrix interface adhesion and particle loading
on mechanical properties of particulate–polymer composites. Compos Part B Eng 39(6):933–961
40. Fairuz A et al (2016) Effect of filler loading on mechanical properties of pultruded kenaf fibre rein-
forced vinyl ester composites. J Mech Eng Sci 10(1):1931–1942
41. DeArmitt C (2011) 26 - Functional Fillers for Plastics. In: Kutz M (ed) Applied Plastics Engineer-
ing Handbook. William Andrew Publishing, Oxford, pp 455–468
42. Dong Y, Jow J, S-y Lai Fly Ash based Fillers and Flame Retardants
43. Depeng L et al (2018) Synergistic effects of intumescent flame retardant and nano-CaCO3 on foama-
bility and flame-retardant property of polypropylene composites foams. J Cell Plast 54(3):615–631
44. Jancar J et  al (1999) Mineral fillers in thermoplastics I: raw materials and processing. Springer,
Berlin
45. Mustafa SN (2012) Effect of kaolin on the mechanical properties of polypropylene/polyethylene
composite material. Diyala J Eng Sci 5(2):162–178
46. Daramola OO (2019) Flexural properties and water absorption characteristics of high density poly-
ethylene-based siloceous composites. Acta Technica Corviniensis-Bull Eng 12(3):57–62

Publisher’s Note  Springer Nature remains neutral with regard to jurisdictional claims in published
maps and institutional affiliations.

13

You might also like