You are on page 1of 40

Polymer

Synthesis, characterization, thermal and mechanical behavior of polypropylene matrix


embedded with 2wt.% CaCO3 and x wt.% graphene nano-platelets (GNPs) hybrid
composites for structural applications
--Manuscript Draft--

Manuscript Number: POLYMER-24-518

Article Type: Research Paper

Section/Category: Polymer Physics and Processing

Keywords: composite; Polypropylene; mechanical and thermal properties

Corresponding Author: Dr.Rasu Karunanithi


B S Abdur Rahman Crescent Institute of Science & Technology
INDIA

First Author: Mrs R.Daulath Banu, B.E.,M.E., (Ph.D)/.

Order of Authors: Mrs R.Daulath Banu, B.E.,M.E., (Ph.D)/.

Dr.Rasu Karunanithi

Dr.S. Sivasankaran, B.E.,M.Tech.,Ph.D.,

Dr.M. Subramanian, Ph.D.,

Dr.Abdullah A. Alhomidand, B.Sc.,M.S., Ph.D.,

Abstract: In this study, ultra-fine graphene nanoplatelets (GNPs) were employed as nanofillers to
reinforce the polypropylene (PP) matrix of composites. This was done in conjunction
with polypropylene grafted maleic anhydride (PP-MAH) compatibilizers and calcium
carbonate (CaCO3), with the aim of improving the mechanical properties of the
resulting hybrid composites. Composite formulations were fabricated by compounding
the PP matrix with varying weight percentages of GNPs (x = 0.5, 1, 1.5, 2.0), 2 wt.%
CaCO3, and 5 wt.% PP-MAH using a twin-screw extruder followed by injection
molding. This research thoroughly investigates the mechanical and thermal
characteristics. XRD, XPS, and FTIR results confirm the successful development of
hybrid composite polymers. Thermal stability, crystallization temperature, melting
temperature, tensile strength, flexural strength, and impact resistance were evaluated
through DSC, TGA, universal testing machine, and low-velocity impact test,
respectively. The results indicated a significant improvement in the tensile strength of
the PP matrix with the addition of GNPs, with the highest enhancement observed at 1.5
wt.% GNP loading, where tensile strength reached a maximum of 40.54 MPa. This
improvement was attributed to the proper interconnection, bonding, and compounding
of PP with GNPs, leading to an increase in load transfer efficiency.

Powered by Editorial Manager® and ProduXion Manager® from Aries Systems Corporation
Cover Letter

Date: 12.02.2024
To
The Editor-in-Chief,
Polymer,

Dear Editor,

Sub: Submission of Technical Research Paper –- for Publication in your esteemed Journal –
Requisition- Reg.

This is for your kind information that our technical research paper entitled “Synthesis,
characterization, thermal and mechanical behavior of polypropylene matrix embedded
with 2wt.% CaCO3 and x wt.% graphene nano-platelets (GNPs) hybrid composites for
structural applications” is a novel work and unpublished one. Numerous experimental work
includes the synthesis of five thermoplastic embedded with 2wt.% CaCO3/x wt.% GNPs of
hybrid composite polymers, detailed characterization using XRD, XPS, FTIR, DSc, DTA, and
HRSEM, mechanical behavior by tensile, compression, flexural, and low velocity impact test
were carried out and investigated in this study. This work is under the scope of your esteemed
journal “Polymer".

I write on behalf of myself and all co-authors to confirm that the results reported in the
manuscript are original and neither the entire work, nor any of its parts have been previously
published. The author(s) confirms that the research in their work is original, and that all the
data given in the article are real and authentic. We strongly believe that the contribution of this
work warrants its publication in your esteemed journal “Polymer”. We are always welcoming
blind peer review process (without showing the country and authors details) to avoid biased
decision by the experts who are working in this field. Kindly consider this paper for its
publication in your esteemed journal after peer review. We are ready to work/revise the
manuscript based on potential reviewer’s/editors suggestions/comments. Looking positive
response.

With kind regards and thanking you,

Yours truly,

R. Karunanithi, Associate Professor.


Manuscript Click here to access/download;Manuscript;Manuscript includes
Figures and Tables.docx
Click here to view linked References

Synthesis, characterization, thermal and mechanical behavior of polypropylene matrix


embedded with 2wt.% CaCO3 and x wt.% graphene nano-platelets (GNPs) hybrid
composites for structural applications
R.Daulath Banua, R. Karunanithia,*, S.Sivasankaranb, M. Subramanianc, Abdullah A.
Alhomidand
1a
Department of Polymer Engineering, B. S. Abdur Rahman Crescent Institute of Science and
Technology, Chennai – 600048, India.
a
*Department of Mechanical Engineering, B. S. Abdur Rahman Crescent Institute of Science
and Technology, Chennai – 600048, India.
b
Department of Mechanical Engineering, College of Engineering, Qassim University, Buraydah
51452, Saudi Arabia.
c
Electroplating and Metal Finishing division, CSIR, Central Electro Chemical Research
Institute, Karaikudi 600 003, Tamilnadu, India
d
Department of Mechanical Engineering, College of Engineering, University of Akron, Ohio
44325-3903, United States
* Correspondence: karunanithi@crescent.education
Abstract:

In this study, ultra-fine graphene nanoplatelets (GNPs) were employed as nanofillers to reinforce
the polypropylene (PP) matrix of composites. This was done in conjunction with polypropylene
grafted maleic anhydride (PP-MAH) compatibilizers and calcium carbonate (CaCO3), with the aim
of improving the mechanical properties of the resulting hybrid composites. Composite
formulations were fabricated by compounding the PP matrix with varying weight percentages of
GNPs (x = 0.5, 1, 1.5, 2.0), 2 wt.% CaCO3, and 5 wt.% PP-MAH using a twin-screw extruder
followed by injection molding. This research thoroughly investigates the mechanical and thermal
characteristics. XRD, XPS, and FTIR results confirm the successful development of hybrid
composite polymers. Thermal stability, crystallization temperature, melting temperature, tensile
strength, flexural strength, and impact resistance were evaluated through DSC, TGA, universal
testing machine, and low-velocity impact test, respectively. The results indicated a significant
improvement in the tensile strength of the PP matrix with the addition of GNPs, with the highest
enhancement observed at 1.5 wt.% GNP loading, where tensile strength reached a maximum of
40.54 MPa. This improvement was attributed to the proper interconnection, bonding, and
compounding of PP with GNPs, leading to an increase in load transfer efficiency.

Keywords: Thermoplastic matrix; GNPs, Composite; polypropylene; mechanical and thermal


properties

1
List of Abbreviations

PP: Polypropylene TGA: thermogravimetric analysis

PP–MAH: Maleic anhydride grafted PP DSc: Differential scanning calorimetry

PET: polyethylene terephthalate XRD: x-ray diffraction

GNPs: graphene nano-platelets XPS: X-ray photoelectron spectroscopy

CNTs: carbon nanotubes FEG-HRSEM: Field emission gun high-


resolution SEM

HDPE: high-density polyethylene EDS: Energy-dispersive X-ray spectrum

FTIR: Fourier transform infrared spectroscopy

SEM: scanning electron microscopy

1. Introduction
Polypropylene (PP) has seen a significant surge in applications, spanning from household items to
various industrial sectors, thanks to its remarkable mechanical strength and favorable thermal and
electrical properties. Despite its advantages, such as abundant raw materials, non-toxicity, ease of
processing, and cost-effectiveness [1], PP encounters strength challenges in diverse contexts.
Research initiatives aim to enhance the performance and thermal characteristics of neat polymer
blends, often involving the incorporation of nano-fillers [2–4]. Common fillers like talc, calcium
carbonate (CaCO3), mica, wollastonite, and kaolin have become standard in the plastics industry,
aiming to reduce production costs of molded components. These fillers play a pivotal role in
improving working properties, including flexural strength, rigidity, durability, and hardness [5].
Typically, these materials are composite structures with an effective polymeric matrix where small
filler and/or fiber particles are systematically dispersed. Drawing insights from prior studies on
thermoplastic polymers incorporated with GNPs [6–11,11–13], it is evident that the performance
of GnPs-based composites depends on multiple factors. These factors encompass the aspect ratio
of the filler, the dispersion and orientation of GnPs within the matrix, the interfacial interaction
between GnPs and the matrix, the selected processing method, and the choice of matrix.

Polymer blending offers a cost-effective means to create materials with desired properties, despite
the thermodynamic immiscibility of many polymers [14]. However, this immiscibility can lead to
poor interfacial adhesion, resulting in inferior properties. Stabilizing morphology through

2
compatibilizers is crucial for achieving enhanced performance [15]. To achieve stabilized
morphology and enhanced properties, compatibilization, either non-reactive or reactive, is crucial.
Non-reactive methods involve adding small amounts of third components, such as block or graft
co-polymers, to blends, promoting interaction with each polymer component. Reactive
compatibilization involves in-situ copolymer formation during processing [16]. Another effective
strategy involves nanofillers for compatibilization in immiscible blends [17]. Non-reactive method
of maleic anhydride grafted polypropylene (PP–MAH) notably improves interfacial adhesion
within the matrix. Compatibilized composites display higher yield strength establishing a converse
relationship compared to unmodified composites [18]. Recently, graphene has gained significant
attention for its addition of GNPs, proven to improve both mechanical and thermal properties.
Wang et al. [19] demonstrated a synergistic effect of GNPs and CNTs when incorporated into
HDPE as the base polymer, resulting in improved electrical properties but a slight reduction in
mechanical properties. To reinforce the mechanical and thermal performance of PP products, this
study explores the use of surface-modified kaolin in PP-MAH/M-kaolin composites.
Characterization involves FTIR [20], SEM, and TGA [21].

Meena et al. [22] conducted a comprehensive study to explore the impact of fly ash on the thermo-
mechanical and mechanical behavior of injection-molded PP matrix composites. Concerning
thermo-mechanical behavior, the addition of fly ash significantly reduced the coefficient of
thermal expansion (CTE) in the composites. Quantitatively, CTE decreased from 10.82 to 7.15
µm/m°C with increasing fly ash content, indicating a clear influence on the dimensional stability
of the composites. Additionally, fly ash inclusion substantially enhanced the flexural modulus,
increasing it from 3200 MPa for pure PP to 4500 MPa with fly ash, highlighting a substantial
enhancement in stiffness. Zafar [23] studied the impact of microwave power on hole characteristics
in microwave-drilled kenaf/PP composites. Results showed a non-linear growth in hole diameter
as microwave power increased. Tirlangi et al. [24] explored the effect of different reinforcing loads
(glass fiber and CNTs) on the mechanical properties of virgin recycled PP composites. Glass fiber
significantly increased tensile strength from 22.6 MPa to 47.3 MPa, while CNTs enhanced flexural
strength from 56.8 MPa to 68.4 MPa in neat recycled PP.

Jan et al. [25] produced composite specimens using green wood-based fillers in both virgin and
recycled PP matrices, examining tribological behavior. Results showed distinct variations in the
coefficient of friction (COF), and wear rates in recycled PP composites were notably lower than

3
in virgin PP composites. Görbe et al. [26] explored nanoparticle embedding in PP matrix-formed
composite foams, revealing significant mechanical property improvements with increased
nanoparticle content. Tensile strength and Young's modulus notably increased, e.g., tensile
strength rose from 3.2 MPa for neat PP foam to 6.7 MPa with added nanoparticles. Leong et al.
[27] studied the mechanical behavior of PP embedded with talc fillers and cellulose fibers in hybrid
composites, reporting an increase in neat PP's tensile strength from 30 MPa to 50 MPa with these
reinforcing agents. Ahmed et al. [28] developed PP embedded with waste tire-based composites,
finding increased hardness from 60 Shore A for pure PP to 80 Shore A with added waste tire
rubber. Anandakumar et al. [29] fabricated PP matrix embedded with short/continuous fibers via
injection molding and conducted a low-velocity impact test. The impact strength of neat PP was
12.5 J, while composite specimens exhibited higher impact strengths, ranging from 18 J to 31 J,
depending on fiber reinforcement type and proportion. Balogun et al. [30] developed PP matrix
embedded with jute and Tetracarpidium conophorum composites, demonstrating an increase in
neat polypropylene's tensile strength from 28 MPa to 50 MPa with added jute and Tetracarpidium
conophorum, signifying a significant enhancement in tensile performance. Basilia et al. [23]
investigated the impact of CaCO3 on thermoplastic polymers, including PP, polyethylene (PE),
and polyvinyl chloride (PVC). Their findings indicated that incorporating 3 wt.% CaCO3 improved
properties, but there was minimal improvement beyond this concentration. Therefore, in this study,
a fixed amount of 2 wt.% CaCO3 was employed, and the variation was solely focused on GNPs
loading. In addition, 5 wt. PP-MAH as a non-reactive compatiblizer was used in this work [15].
Previously, several researchers attempted to develop PP based composites reinforced with single
filler. However, polymeric composites embedded with more than one filler is expected to get more
positive hybrid effect obtain from two different fillers or matrix. In this work, CaCO3 with 2 wt.%
as primary filler was fixed based on previous literature [23] and presence of CaCO3 in polymer
can improve mechanical properties of pipes and high resistance to external blow [31]. GNPs was
chosen as secondary variable fillers with PP matrix to enhance service temperature, mechanical
and thermal properties [32]. Numerous studies explored variations in process conditions and filler
combinations in composite materials, aiming to deduce the impacts of compatibilizers on PP/GNPs
nanocomposites. However, an unaddressed gap exists in research concerning PP matrices
embedded with CaCO3 and ultra-fine GNP-based hybrid composites. Our primary focus was
formulating PP/CaCO3/GNPs hybrid composites through the incorporation of PP-MAH
compatibilizers via twin-screw extrusion followed by injection molding. The key aim was to

4
comprehensively investigate the influence of GNPs loading on the mechanical and thermal
properties of PP/2CaCO3/GNPs-x wt.% (x = 0, 0.5, 1.0, 1.5, & 2.0) hybrid composite materials.

2.0 Experimental work

2.1. Materials

PP homopolymer (Granules: 2.5-3.5 mm, SABIC, Saudi Arabia) was utilized as the base polymer.
The as-received PP granules have a density of 0.92 g/cm3, and the melt flow index is 7 g/ min. The
compatibilizer, PP–MAH (NG2002 grade, 5wt.% [15]), and Calcium carbonate (CaCO3, 2wt.%
[23], 0.64 m) were commercially sourced in the Saudi market (Riyadh). The GNPs, in dry powder
form, have an average diameter of 5 µm and a thickness of 8 nm used in this study. The GNPs
were purchased from M/s Nanografi, Çankaya/Ankara, Turkey, with a purity exceeding 99.9%. In
Figure 1 (a)-(c), the HR-SEM microstructure illustrates the distinct layers of the as-received GNPs
[33]. Additionally, Figure 1d presents the schematic diagram of the hexagonal arrangement of
GNPs.

Figure 1.(a)-(c) HR-SEM micrograph of as-received GNPs at different magnification, inset of (c)
show the thickness of GNP; (d) schematic diagram representing the arrangement of atoms in
GNPs as hexagonal topography

5
2.2 Development of PP/2CaCO3/GNPs-x wt.% (x = 0, 0.5, 1.0, 1.5, and 2.0) hybrid composite
polymers

The as-received PP matrix granules, PP–MAH granules, CaCo3, and GNPs underwent drying in
an oven at 80°C for 1 h to eliminate moisture. Different weight percentages of GNPs were used to
reinforce the pure PP matrix along with compatibilizers. Materials composition, sample ID, and
reinforcement percentages are detailed in Table 1. All materials were weighed and mechanically
blended for approximately 45 minutes to achieve homogenization. Prior to this, compatibilizers
and GNPs were dispersed in ethanol (Concentration of ethanol:materials ratio: 100:15) using a
magnetic stirrer (1000 ml capacity, LKTC-B1-T, KATLY, China) for 30 minutes to minimize
GNPs' agglomeration over the base matrix. The mixture was then dried in an oven at 353 K for 2
hours. Mechanically blended base PP, PP–MAH, CaCO3, and GNPs were processed in a twin-
screw extruder (Screw diameter: 21 mm, Input voltage: 220V, Power: 1.1 kW, Output range: 100-
3000 g/hr, production linear speed: <1000 cm/min, machine size: 1800x200x1000mm, Made: M/s
Dongguan Junxin Plastic & Metal Co. Ltd., Qiaotou Town, Dongguan, China) for melt
compounding. The extrusion process involved three heating zones with set temperatures of
195±2°C, 205±2°C, and 215±2°C for zones 1, 2, and 3, respectively. After material loading, a 10-
minute holding period to attain uniform temperature preceded the start of extrusion. The extruded
filament, with a 3.5 mm diameter, was immediately cooled in water and then fed into a chopper to
produce pellets. A screw speed of 80 rpm was maintained during extrusion. Figure 2 illustrates the
schematic diagram of the twin-screw extruder's operation followed by pelletizing into granules.
Pelletized granules were utilized in a vertical injection molding machine (30 Ton capacity,
supplied by M/s Dongguan Junxin Plastic & Metal Co. Ltd., Qiaotou Town, Dongguan, China)
for fabricating bulk samples.

6
Figure 2. Schematic diagram representing present research work includes twin extrusion,
pelletizing/chopping, vertical injection moulding, materials characterization, mechanical testing,
and photograph of developed polymers

Table 1. Sample designation, sample ID, percentage of PP, GNPs, CaCO3, and PP-MAH used in
this study
PP GNPs PP-
CaCO3,
Sample designation Sample ID MAH,
(wt. %) (wt. %) wt. %)
wt. %

PP/2CC/GN
PP+0GNP (as-received PP granules) 100 - - -
Ps-0

PP/2CC/GN
PP+0.5GNP+2CaCO3+5PP–MAH 92.5 0.5 2 5
Ps-0.5

PP/2CC/GN
PP+1GNP+2CaCO3+5PP–MAH 92 1 2 5
Ps-1.0

PP/2CC/GN
PP+1.5GNP+2 CaCO3+5PP–MAH 91.5 1.5 2 5
Ps-1.5

PP/2CC/GN
PP+2GNP+2CaCO3+5PP–MAH 91 2 2 5
Ps-2.0

2.3 Characterization of developed PP/CC/GNPs hybrid composite polymers using FTIR,


DSc, TGA, XRD, XPS, and FEG-HRSEM

The FTIR test, following the ASTM E168-16 standard [24], covered the range of 4000 to 380
cm-1 wavelength using a JASCO BSA spectrometer with a 4 cm-1 resolution. FTIR analyses of
PP/2CC/GNPs-0, PP/2CC/GNPs-1.0, and PP/2CC/GNPs-2.0 were conducted to identify chemical

7
compounds and crystalline properties. The high-spectral infrared data captured by the FTIR
machine underwent Fourier transform for compound identification. DSc analyses of
PP/2CC/GNPs-0, PP/2CC/GNPs-1.0, and PP/2CC/GNPs-2.0 polymeric hybrid composite samples
covered a temperature range from 27°C to 550°C, utilizing a heating rate of 10 K/min under N2
atmosphere. A NETZSCH DSC 204 F1 thermal analyzer was employed, following ASTM 3418
standards [24]. The samples were cooled to room temperature at 10 °C min⁻ ¹ under nitrogen.
Specimens weighing approximately 10 mg were prepared with precision to ensure flatness and
smoothness for accurate DSC runs. The net heat flow, recorded concerning a high-purity annealed
aluminum reference, was plotted against temperature. Baseline correction and drift correction were
meticulously performed for accurate measurement of heat evolution and absorption. DSC
thermograms were repeated to confirm reproducibility. Thermal properties, including melting
enthalpy (∆Hm), degree of crystallinity (Xc), crystallization temperature (Tc), and melting
temperature (Tm), were calculated using Eq. (1) based on the melting enthalpy results.
∆𝐻
% Crystallinity, 𝑋𝑐 = 𝑓×∆𝐻𝑚° (1)
𝑚

Here, ∆Hm represents the observed melting enthalpy, while f denotes the weight fraction of the
°
PP matrix phase. The reference value for the enthalpy of 100% crystalline PP, denoted as ∆𝐻𝑚 ,
is fixed at 209 J/g [34].

For TGA, a TA instrument (New Castle, DE, USA) was employed. The sample underwent heating
from 25 °C to 700 °C, with a heating rate of 10 °C per minute, all within a nitrogen atmosphere.
Essential TGA parameters, including the onset temperature, were derived from the TGA data,
representing the temperature when initial weight loss commences. XRD analysis was performed
on PP/2CC/GNPs-0, PP/2CC/GNPs-1.0, and PP/2CC/GNPs-2.0 to examine different phases
within the polymer composites. PANalytical X’Pert Pro instrument with CuKα radiation was used
for scanning from a 5° to a 90° diffraction angle at a speed of 2° per minute. XPS ( K surface
analysis, M/s Thermo Fisher Scientific, Altrincham, Cheshire , U.K) analyzed elemental
composition and chemical properties of the samples, providing insights into the overall structure.
A sample size of around 10×10×10 mm was used for XPS examination. Before, the sample was
cleaned and ensured free from contaminants. FEG-HRSEM using the Apreo instrument assessed
surface characteristics and fracture surfaces of samples subjected to mechanical testing, including
tensile and impact tests. The operational voltage was 30 keV, and the resolution was 1.3 nm

8
2.4 Mechanical Testing of Tensile, Compression, 3P bending, and low velocity impact tests

Tensile tests, following ASTM D638 (Figure 3a), were conducted using a universal testing
machine (M/s MTS System Corporation, model No: 370.25, 250kN capacity, Texas, USA).
Mechanical properties, including yield strength, ultimate strength, and elongation at rupture, were
determined with a loading rate of 1 mm/min. Each material composition underwent five trials, and
the average values were used for interpretation. Tensile stress (σt) was calculated using Eq. (2):
𝐹𝑡
𝜎𝑡 = (2)
𝐴

where Ft represents applied tensile load (in N) and A is the specimen's cross-sectional area (in
mm²). Tensile strain (εt) was determined by Eq. (3):
𝛿
𝜀𝑡 = 𝐿 (3)

where δ signifies deformation (in mm) and L is the initial length of the specimen (in mm).

Flexural testing, according to ASTM D790 standard (Figure 3b), utilized the same MTS universal
testing machine with a special fixture. The sample had a span of 40 mm between supports, and the
load was applied at the center. Flexural stress, flexural strain, and flexural modulus were
determined with a testing speed of 1 mm/min. Each material composition underwent five trials in
this test. Flexural stress (σb) was calculated using Eq. (4):
8𝐹𝐿
𝜎𝑏 = 𝜋𝑑3 (4)

where F is bending load (N), L is support distance (mm), and d is the bending sample diameter
(mm). Flexural strain (εb) was determined using Eq. (5) considering deflection (δ) at the span
center (mm):
6𝛿𝑑
𝜀𝑏 = (5)
𝐿2

The flexural modulus (Eb) was calculated via Eq. (6), utilizing the first linear slope (m) from the
bending force-deflection curve (between 0.25Fmax and 0.75 Fmax) [35].

4𝐿3 𝑚
𝐸𝑏 = (6)
3𝜋𝑑4

9
Figure 3. Photographs of fabricated samples for mechanical testing of: (a) Tensile samples; (b)
compressive samples; (c) 3P bending test samples; and (d) low velocity impact test samples

Figure 4. Schematic diagram showing various mechanical testing (tensile, flexural, and
compressive) using MTS universal testing machine, photograph of MTS machine and magnified
photos showing samples under different types of loadings
The uniaxial compression test, following ASTM D695 standard (Figure 3c), utilized a 15 mm
diameter and 22.5 mm length specimen. At least five trials were conducted for each composition.
Compressive stress (σc) was determined using Eq. (7):
𝐹𝑐
𝜎𝑐 = (7)
𝐴

where Fc is the applied compressive load (in N), and A is the cross-sectional area of the specimen
(in mm²). Compressive strain (εc) was calculated with Eq. (8):
𝐹𝑐
𝜎𝑐 = (8)
𝐴

with δ representing deformation (in mm) and L as the initial length of the specimen (in mm).

10
Similarly, the low-velocity impact test (LVI), adhering to ASTM D7136 standard (Figure 3d), used
60 mm square samples with a 10 mm thickness. Figure 4 illustrates the schematic diagram of
various mechanical tests (tensile, flexural, and compressive) performed on an MTS universal
testing machine, including a photograph of the MTS machine and magnified photos showing
samples under different loading conditions. Additionally, Figure 5 presents the schematic diagram
of the low-velocity impact test conducted on an INSTRON CEAST 9350 machine (M/s INSTRON
CEAST factory, 90 Ton capacity, Milan city, Italy) featuring a photograph of the machine and
magnified photos showing samples under impact load. Peak impact force, absorbed impact energy,
velocity, and displacement profiles were computed using equations (9), (10), and (11) [36,37]:
𝑡
𝐹(𝑡)
𝑣(𝑡) = 𝑣𝑖 + ∫ 𝑑𝑡 (9)
0 𝑚

𝑡 𝑡
𝑔𝑡 2 𝐹(𝑡)
𝐷(𝑡) = 𝐷𝑖 + 𝑣𝑖 𝑡 + − ∫ (∫ ) 𝑑𝑡 (10)
2 0 0 𝑚

𝑚[𝑣𝑖2 − 𝑣(𝑡)2 ]
𝐸(𝑡) = + 𝑚𝑔𝐷(𝑡) (11)
2

Where v is the velocity in m/s, vi is the incident velocity of the impactor in m/s, t is the time in ms,
g is the gravity in m/s2, F(t) is the vertical force in N, m is the mass in kg, D(t) is the displacement
in mm, and Di is the initial position in mm.

Figure 5. Schematic diagram showing low velocity impact test in a drop tower testing machine,
photograph of INSTRON CEAST 9350 machine and magnified photos showing samples under
impact load

11
3.0 Results and discussions

3.1 Characterization of Morphological Features using FEG-HRSEM

The dispersion of GNPs and CaCO3 within the PP matrix was examined through HRSEM using a
cross-sectioned injection-molded sample, as shown in Figure 6. The injection-molded sample,
broken and mounted with acrylic resin, underwent gold coating via a sputtering machine before
HRSEM scanning. Figure 6a displays the HRSEM topography of the pure PP matrix, showing a
clean and more PP domain. Figure 6b depicts the presence of a slightly reduced PP domain, GNPs,
and CaCO3, indicating the interfacial tension between incorporated GNPs and CaCO3 with the PP
matrix [15]. With an increase in GNPs loading, as shown in Figure 6c-f, the observed PP domain
started to decrease due to the large interfacial tension of the PP matrix with GNPs and CaCO 3,
confirming the effective compatibilizing effect given by PP-MAH. This was attributed to the
increased viscosity of the PP matrix and the barrier effect produced by GNPs and CaCO 3.
Furthermore, the presence of GNPs and CaCO3 over the PP matrix domain confirms the proper
bonding of embedded GNPs and CaCO3 with the PP matrix. However, as per Figure 6e, non-
uniform distribution of GNPs over the PP matrix begins to occur, and more agglomeration of GNPs
and CaCO3 occurs in higher loading of GNPs, which may be expected to decrease the mechanical
properties. EDS analyses of PP/2CC/GNP-1.0 hybrid composite polymer are shown in Figure 6d,
with insets presenting the observed elemental composition. The presence of calcium (Ca), silicon
(Si), sodium (Na), aluminum (Al), and sulfur (S) in the composite polymer (Figure 6d) indicates
successful compounding of incorporated compatibilizers along with GNPs [22]. The formed Si,
Na, Al, and S elements are expected during the processing of incorporated GNPs and CaCO3 filler
[38].

12
Figure 6. HRSEM cross-sectional analyses of injection moulded samples of: (a) pure PP matrix;
(b) PP/2CC/GNP-0.5 (c) PP/2CC/GNP-1.0 (d) EDS results of PP/2CC/GNP-1.0 (e)
PP/2CC/GNP-1.5 and (f) PP/2CC/GNP-2.0 hybrid composite polymers

3.2. Investigation of Crystalline Structure via XRD Analysis

Figure 7 presents the XRD peak profiles of the pure PP polymer matrix, PP/2CC/GNPs-1.0, and
PP/2CC/GNPs-2.0 hybrid composite polymers. In the pure PP matrix polymer, XRD peaks
centered at 2θ angles of 14.3°, 16.7°, 18.7°, 22.6°, and 25.4° were observed, corresponding to (1
1 0), (0 4 0), (1 3 0), (2 02), and (0 6 0) planes, respectively, associated with the α-form of PP
matrix crystals. No additional peaks or oxides were noticed, indicating purity and process
capability. The XRD profiles of hybrid-composite polymers at a 2θ angle of 26.4° and 26.8° were
related to the incorporated CaCO3 and GNPs with the corresponding plane being (1 0 4) and (0 0
2). The results clearly demonstrate a decrease and broadening of PP matrix crystal peaks with the
addition of ultra-fine GNPs, and the peak corresponding to GNPs increased with their content,
confirming successful composite polymer formation. The peak positions of the α-form PP matrix

13
shifted with an increase in GNPs loading. This shift was attributed to the large interfacial tension
between the incorporated GNPs and the PP matrix. This significant interfacial tension occurs due
to the increasing viscosity of the PP matrix with the addition of GNPs. Niu et al [38] and Xie et al
[39] reported that the addition of PP–MAH compatibilizer in PP matrix can suppress the formation
of β-form of PP crystals. Therefore, in the present XRD results, no β-form of PP crystals was
observed. The relative amount of GNPs in the PP matrix, represented by a factor (K), is determined
by the ratio of XRD peak intensity of GNPs (IGNPS) at (0 0 2) to the sum of intensities of α-form
PP matrix crystals (Iα1, Iα2, Iα3, and Iα4) at four different diffraction angles of (1 1 0), (0 4 0), (1 3
0), and (2 0 2), calculated using the following equation (12) [38,39]:
𝐼𝐺𝑁𝑃𝑠
𝐾 = (𝐼 (12)
𝛼1 +𝐼𝛼2 +𝐼𝛼3 +𝐼𝛼4 +𝐼𝐺𝑁𝑃𝑠 )

The calculated K values were 0 for pure PP matrix, 0.10 for PP/2CC/GNPs-1.0, and 0.16 for
PP/2CC/GNPs-2.0 composite polymers. The K value signifies GNPs crystal formation, with
PP/2CC/GNPs-2.0 showing a higher K, indicating more GNPs crystals in this sample. This result
confirms the successful development of composite polymers.

Figure 7. XRD peak profiles of pure PP matrix, PP/2CC/GNPs-1.0, and PP/2CC/GNPs-2.0


composite polymers
3.3 Examination of Chemical Properties through XPS Analysis

Figure 8 illustrates the XPS survey profiles of pure PP matrix, PP/2CC/GNPs-1.0, and
PP/2CC/GNPs-2.0 hybrid composite polymers, elucidating the chemical nature and composition.
The atomic concentrations of various elements present in the developed polymers are detailed in

14
Table 2. The XPS survey scanning (Figure 8) indicates an increased peak intensity of C1s with
rising GNPs (0 to 2 wt.%), signifying successful development of composite polymers. Conversely,
the O1s peak intensity decreased with escalating GNPs, confirming proper bonding and nucleation
of GNPs in the PP matrix [40,41]. Detailed peak deconvolution was examined in the XPS results
at C1s and O1s levels, as illustrated in Figure 9 and Figure 10, respectively (

Table 3). At the C1s level, three de-convoluted peaks at 282, 283, and 284 eV represented CHn/C-
C (Alkyl group, [42,43]), C-OH (Alcoholic group, [43]), and C=O (Alkyl group, [44]). The
increase in alkyls (CHn/C-C) and alcoholics (C-OH) peaks with rising GNPs confirms effective
bonding in the polymer chain. At the O1s level (Figure 10), peaks at 533 and 535 eV represented
carboxylic acid group (O=C-O) and ketone group (O=C) [44], associated with nucleation and
bonding effects, which increased with higher GNPs content, and affirming the successful
formation of hybrid composite polymers.

Figure 8. X-ray photoelectron spectroscopy (XPS) of pure PP matrix, PP/2CC/GNPs-1.0 and


PP/2CC/GNPs-2.0 composite polymers obtained from extrusion

Table 2. Chemical composition of different elements in the developed pure PP matrix,


PP/2CC/GNPs-1.0 and PP/2CC/GNPs-2.0 hybrid composite polymers measured using XPS
Atomic concentrations, %

15
Name of
C1s O1s Ca2p
Sample

Pure PP matrix 86.89 13.11 -

PPGNPs-1.0 88.62 10.26 1.12

PPGNPs-2.0 92.49 6.61 0.9

Table 3. XPS peak position for the corresponding observed C1s, and O1s of developed pure PP
matrix, PP/2CC/GNPs-1.0 and PP/2CC/GNPs-2.0 hybrid composite polymers
C1s O1s Ca2p

-
Binding
CHn/C- C-OH C=O C=O O=C-O
energy
C CaCO3
(eV)
Alkyl Alcoholics Ketone Ketone carboxylic
group group group group acid group

HEP0Gr 284.8 283.8 281.6 535.3 533.6 -

HEP1.0Gr 284.6 283.6 281.9 535.5 533.7 350.0

HEP2.0Gr 284.2 283.4 282.1 535.7 533.8 350.1

16
Figure 9. XPS spectra examined at the centre of C1s level for pure PP matrix, PP/2CC/GNPs-1.0
and PP/2CC/GNPs-2.0 hybrid composite polymers

Figure 10. XPS spectra examined at the centre of O1s level for pure PP matrix, PP/2CC/GNPs-
1.0 and PP/2CC/GNPs-2.0 hybrid composite polymers
3.4. Evaluation of Molecular Structure using FTIR Analysis

The molecular formation and bonding structure of the as-received GNPs, developed pure PP
matrix, PP/2CC/GNPs-1.0, and PP/2CC/GNPs-2.0 hybrid composite polymers were examined
using FTIR, as depicted in Figure 11. For as-received GNPs (Figure 11a), the observed peaks at
3438 and 2923 cm-1 are related stretching of O-H bond, whereas, the observed peaks at 1731, 1635,
and 1075 cm-1 are attributed to stretching vibrations of C=C, C-OH, and C=O, respectively [45,46].
For the developed composite polymers (Figure 11b), asymmetric vibrations within the wavelength

17
range of 2903-2945 cm-1 (point A) and symmetric vibrations within 2815-2962 cm-1 (point B)
confirm the copolymer nature and robust bonding of incorporated GNPs in the PP matrix [47,48].
The range of 2903-2945 cm-1 (point A) corresponds to asymmetric stretching vibrations of the CH3
group of the PP matrix [48]. The range of 2815-29625 cm-1 (point B) indicates symmetric
stretching vibrations corresponding to the CH2 group of the PP matrix. The observed peak depth
at point A increases with GNPs up to 1 wt.% in the PP matrix (PP/2CC/GNPs-1.0), then decreases
in PP/2CC/GNPs-2.0, signifying alterations in internal structure and the occurrence of
compounding due to nucleation and strong bonding of incorporated GNPs in the PP matrix [46]
(Figure 11c). At 1380 cm-1 (point C), umbrella bending vibrations indicate C-H bonds in all
samples, and at 1460 cm-1 (point D), symmetrical bending vibrations corresponding to the CH3
group are present. These FTIR results confirm the successful development of composite polymers.

Figure 11. Fourier transform infrared spectroscopy (FTIR) profile of: (a) as-received pure GNPs;
(b) developed pure PP matrix, PP/2CC/GNPs-1.0 and PP/2CC/GNPs-2.0 hybrid composite
polymers; (c) Magnified view of (b)

18
3.5. Assessment of Thermal Properties via DSC Analysis

Figure 12a illustrates the DSC analyses of the developed polymers (pure PP, PP/2CC/GNP-1.0,
and PP/2CC/GNP-2.0) during heating (Figure 12b) and cooling (Figure 12c) cycles. The thermal
properties, including recrystallization temperature (Tc), melting temperature (Tm), change in
enthalpy of recrystallization (ΔHc), change in enthalpy of melting (ΔHm), crystallinity at room
temperature, and percentage crystallinity (Xc), were determined, as listed in

Table 4. The results clearly demonstrate that incorporating GNPs into the PP matrix significantly
Name of Sample ∑(∆𝐻𝑐 + ∆𝐻𝑚),
Tc, °C ∆𝐻𝑐 , J/g Tm, °C ∆𝐻𝑚, J/g Xc
J/g

Pure PP matrix 124.8±0.56 166.2±0.32 26.271±0.20 25.086±1.52


-26.159±1.55 52.430±2.21
PP/2CC/GNPs-1.0 121.2±0.64 171.4±0.14 26.492±0.11 30.095±0.95
-34.520±1.82 61.013±1.58
PP/2CC/GNPs-2.0 118.6±0.71 172.1±0.47 33.530±0.87 36.889±0.84
-40.486±2.04 74.016±2.66
influences thermal properties. For the pure PP matrix polymer, the calculated ΔHm and ΔHc were
52.43 J/g and -26.15 J/g, respectively. In contrast, PP/2CC/GNP-2.0 composite polymer exhibited
ΔHm and ΔHc of 74.01 J/g and -40.48 J/g, respectively. PP/2CC/GNP-1.0 composite polymer
showed 16.4% more ΔHm and 31.2% more ΔHc compared to pure PP matrix, while PP/2CC/GNP-
2.0 exhibited 41.11% more ΔHm and 54.76% more ΔHc. These results clearly indicate varied
thermal properties with the incorporation of GNPs into the PP matrix. The percentage of
crystallinity also increased significantly with the addition of GNPs, attributed to effective bonding
between GNPs and PP, resulting in the formation of more α-PP crystals. The introduction of GNPs
caused a noteworthy change in melting temperature (Tm, increased from 166.2 to 172.1°C) and
recrystallization temperature (Tc, decreased from 124.8° to 118.6°), highlighting the strong
bonding effect of GNPs with the α-PP matrix [49]. The observed decrease in Tc and the increase
in Tm can be attributed to presence of GNPs and CaCO3 in the composites. The decrease in Tc
might be associated with the hindrance of the crystallization process caused by the presence of
GNPs and CaCO3, which could act as nucleation sites but also affect the overall crystallization
kinetics. The increase in Tm could be related to changes in the crystalline structure induced by the
interaction between PP and the incorporated fillers [49].

19
Figure 12. DSC curves of developed pure PP matrix, PP/2CC/GNP-1.0 and PP/2CC/GNP-2.0
composite polymers of: (a) Combined heating & cooling curves (b) magnified view of heating
curve; (b) magnified view of cooling curve

Table 4. Thermal properties obtained from DSc results for the developed composite
polymers
Name of Sample ∑(∆𝐻𝑐 + ∆𝐻𝑚),
Tc, °C ∆𝐻𝑐 , J/g Tm, °C ∆𝐻𝑚, J/g Xc
J/g

Pure PP matrix 124.8±0.56 166.2±0.32 26.271±0.20 25.086±1.52


-26.159±1.55 52.430±2.21
PP/2CC/GNPs-1.0 121.2±0.64 171.4±0.14 26.492±0.11 30.095±0.95
-34.520±1.82 61.013±1.58
PP/2CC/GNPs-2.0 118.6±0.71 172.1±0.47 33.530±0.87 36.889±0.84
-40.486±2.04 74.016±2.66

3.6. Analysis of Thermal Stability using TGA Analysis

Figure 13 displays the thermal degradation profiles of the developed pure PP matrix,
PP/2CC/GNPs-1.0, and PP/2CC/GNPs-2.0 hybrid composite polymers obtained through melt
extrusion via twin extruder. This provides insights into the thermal stability of all fabricated
samples. The calculated thermal degradation temperatures at different weight loss percentages
(5%, 10%, 20%, and 50%) are listed in Table 5. Examining Figure 13 and Table 5 reveals a
substantial improvement in thermal stability with the incorporation of GNPs compared to the pure
PP matrix, attributed to the strong nucleating and bonding effects of GNPs over the PP matrix. For

20
instance, the onset temperature on weight loss of pure PP matrix at T5%, T10%, T20%, and T50% was
371.89°C, 406.49°C, 433.99°C, and 460.35°C, respectively. In contrast, the onset temperature on
weight loss of PP/2CC/GNP-2.0 composite polymer at T5%, T10%, T20%, and T50% was 388.02 °C,
422.81 °C, 444.36 °C, and 466.40 °C, respectively. The enhanced thermal stability in
PP/2CC/GNPs hybrid composite polymers can be attributed to the insulating and barrier effects of
the introduced GNPs, leading to increased thermal resistance compared to the pure PP matrix [18].
At a maximum temperature of 500°C, the residual weight of the pure PP matrix, PP/2CC/GNPs-
1.0, and PP/2CC/GNPs-2.0 hybrid composite polymers was found to be 0%, 1.5%, and 2.45%,
respectively. These results indicate an improvement in the thermal stability of the GNPs-loaded
composite polymers.

Table 5. Thermo-gravimetric analyses (TGA) of developed pure PP matrix, PP/2CC/GNPs-1.0


and PP/2CC/GNPs-2.0 hybrid composite polymers listing the degradation temperature with the
function of weight loss
Thermal degradation temperature, (°C)

Weight
Name of Sample Weight loss Weight loss
loss at Weight loss at
at 10%, T10 at 50%, T50
5%, T5 20%, T20 (°C)
(°C) (°C)
(°C)

Pure PP matrix 371.89 406.49 433.99 460.34

PP/2CC/GNPs-1.0 377.61 412.37 437.53 462.72

PP/2CC/GNPs-2.0 388.02 422.81 444.36 466.40

21
Figure 13. Thermo-gravimetric analysis (TGA) of pure PP matrix, PP/2CC/GNPs-1.0 and
PP/2CC/GNPs-2.0 hybrid composite polymers obtained from melt extrusion via twin extruder
3.7. Evaluation of Mechanical Properties

3.7.1. Tensile Testing

The mechanical properties, encompassing yield strength, strain at the yield point, ultimate strength,
strain at the ultimate point, fracture strength, strain at fracture, modulus of elasticity, toughness,
and resilience, were evaluated via tensile testing using engineering stress-strain curves, depicted
in Figure 14. Corresponding mechanical properties are detailed in Table 6. The results in Figure
14 and Table 6 unmistakably reveal a substantial increase in the tensile strength of the pure PP
matrix with an escalating percentage of GNPs. For example, the ultimate tensile strength of pure
PP matrix, PP/2CC/GNPs-0.5, PP/2CC/GNPs-1.0, PP/2CC/GNPs-1.5, and PP/2CC/GNPs-2.0 was
35.829, 37.474, 40.076, 40.546, and 38.426 MPa, respectively. The enhanced tensile strength
compared to the pure PP matrix was 4.59%, 11.85%, 13.16%, and 7.25% for PP/2CC/GNPs-0.5,
PP/2CC/GNPs-1.0, PP/2CC/GNPs-1.5, and PP/2CC/GNPs-2.0, respectively. This improvement is
attributed to the nucleating effect of GNPs in the PP matrix, proper bonding of incorporated GNPs,
and the uniform dispersion of GNPs, facilitating effective load transfer between PP matrix and
GNPs fillers [18,50]. As depicted in Table 6, there is a notable decrease in tensile strain at the
ultimate point with the increasing content of GNPs in the PP polymer matrix, attributed to an
escalation in defects upon GNPs introduction. For instance, the percentage reduction in tensile

22
strain for developed composite polymers compared to pure PP matrix was 19.23%, 23.07%,
25.64%, and 34.65% for PP/2CC/GNPs-0.5, PP/2CC/GNPs-1.0, PP/2CC/GNPs-1.5, and
PP/2CC/GNPs-2.0, respectively. This outcome signifies a substantial reduction in
elongation/ductility, indicating hindered chain mobility within the polymer matrix due to the
increased GNPs content in the PP matrix. Consequently, this results in heightened stiffness in the
composite polymers, aligning with previously reported findings in the literature [18,51].

Figure 14. Engineering stress-strain curves of pure PP matrix, PP/2CC/GNPs-0.5,


PP/2CC/GNPs-1.0, PP/2CC/GNPs-1.5, and PP/2CC/GNPs-2.0 hybrid composite polymers
obtained from uniaxial tensile test

Table 6. Mechanical properties of developed pure PP matrix, PP/2CC/GNPs-0.5, PP/2CC/GNPs-


1.0, PP/2CC/GNPs-1.5, and PP/2CC/GNPs-2.0 hybrid composite polymers obtained from
uniaxial tensile test
Ultimate Tensile strain
Density of Modulus of Tensile Yield Tensile Strain
Sample ID Tensile stress at ultimate
samples Elasticity (MPa) Stress (MPA) at yield point
(MPa) point

Pure PP 0.918±0.002 1921.869±3.56 20.896±0.65 0.016±0.0021 35.829±0.38 0.078±0.0011

PP/2CC/GNPs-0.5 0.895±0.001 2600.221±2.86 22.189±0.34 0.013±0.0016 37.474±0.42 0.063±0.0020

PP/2CC/GNPs-1.0 0.862±0.003 2965.183±1.95 25.137±0.42 0.015±0.0011 40.076±0.44 0.060±0.0014

PP/2CC/GNPs-1.5 0.827±0.002 2923.035±2.05 26.790±0.25 0.014±0.0008 40.546±0.52 0.058±0.0013

PP/2CC/GNPs-2.0 0.797±0.003 2860.220±1.31 25.896±0.51 0.014±0.0014 38.426±0.22 0.051±0.0016

23
3.7.2. Compression testing
The analysis of mechanical properties from a compression test, including compressive yield
strength, compressive strain at the yield point, ultimate compressive strength, and compressive
strain at the ultimate point, was conducted through engineering compressive stress-strain curves
(Figure 15). These properties are detailed in Table 7. Upon analyzing Figure 15 and Table 7, it's
clear that the compressive strength of the composite polymers significantly increases with GNPs
incorporation into the pure PP matrix. For example, the percentage improvement in ultimate
compressive strength compared to pure PP matrix was 9.75%, 14.44%, 21.27%, and 16.57% for
PP/2CC/GNPs-0.5, PP/2CC/GNPs-1.0, PP/2CC/GNPs-1.5, and PP/2CC/GNPs-2.0, respectively.
The compressive strength depends on factors such as the compatibility between polar and nonpolar
components in the polymer blend. The observed enhancement (Figure 15 and Table 7) is attributed
to nucleating effects of GNPs in the PP matrix, effective dispersion, and load transfer between
GNPs and PP matrix. Additionally, the introduced percentage of PP-MAH and CaCO3 as
compatibilizers may have contributed to the improvement in compressive properties during the
uniaxial compressive test. PP/2CC/GNPs-1.5 demonstrated the highest mechanical properties
among the developed composite polymers. Mechanical properties started to decrease beyond
1.5%, leading to the decision to halt the addition of GNPs at 2.0%

24
Figure 15. Compressive stress-strain curves of pure PP matrix, PP/2CC/GNPs-0.5,
PP/2CC/GNPs-1.0, PP/2CC/GNPs-1.5, and PP/2CC/GNPs-2.0 hybrid composite polymers
obtained from uniaxial compression test

Table 7. Mechanical properties of developed pure PP matrix, PP/2CC/GNPs-0.5, PP/2CC/GNPs-


1.0, PP/2CC/GNPs-1.5, and PP/2CC/GNPs-2.0 hybrid composite polymers obtained from
uniaxial compression test

Compressive Compressive Ultimate Compressive


Sample ID Yield Stress Strain at compressive strain at
(MPa) yield point stress (MPa) ultimate point

Pure PP 15.324±1.85 0.0217±0.005 45.1699±0.85 0.1504±0.003

PP/2CC/GNPs-0.5 18.345±0.95 0.0189±0.006 49.5740±0.78 0.1537±0.005

PP/2CC/GNPs-1.0 21.367±1.10 0.0227±0.004 51.6940±1.05 0.1504±0.004

PP/2CC/GNPs-1.5 24.891±0.86 0.0287±0.007 54.7760±0.98 0.1508±0.005

PP/2CC/GNPs-2.0 23.256±1.74 0.0219±0.003 52.6540±0.69 0.1504±0.004

3.7.3 Flexural testing


The flexural properties, encompassing flexural stress, strain, modulus, and toughness, were
determined through a three-point bending test. These properties, derived from engineering flexural
stress-strain curves (Figure 16), are detailed in

25
Table 8. Reviewing the results in Figure 16.and

Table 8 reveals a substantial increase in flexural strength for PP/2CC/GNPs-based hybrid


composite polymers with GNPs incorporation. This improvement is attributed to effective GNP
interaction with the PP matrix, uniform GNP dispersion, and the barrier effect of GNPs, along with
other factors described earlier.

Figure 16. Flexural stress-strain curves of pure PP matrix, PP/2CC/GNPs-0.5, PP/2CC/GNPs-


1.0, PP/2CC/GNPs-1.5, and PP/2CC/GNPs-2.0 composite polymers obtained from 3P bending
test

Table 8. Flexural properties of developed pure PP matrix, PP/2CC/GNPs-0.5,


PP/2CC/GNPs-1.0, PP/2CC/GNPs-1.5, and PP/2CC/GNPs-2.0 composite polymers obtained
from 3P bending test
Flexural Stress Flexural Strain Flexural Modulus of
Sample
(MPa) (mm/mm) modulus (MPa) Toughness (J/mm3)
Pure PP 46.572 0.3655 371.323 30.979
PP/2CC/GNPs-0.5 49.707 0.4249 382.215 49.707
PP/2CC/GNPs-1.0 55.437 0.4600 425.585 55.437
PP/2CC/GNPs-1.5 57.118 0.5152 448.723 57.118
PP/2CC/GNPs-2.0 56.894 0.4488 439.186 56.841

3.7.4 Low velocity impact test

26
Low-velocity impact testing aimed to evaluate the impact performance of the newly developed
composite polymers was conducted using a CEAST 9350 Instron impact testing machine (Figure
5). This machine comprises a weight dropping system with an impactor head housing the insert,
an anti-rebound mechanism, a lubrication system, a photocell, and a high-energy configuration
positioning system. A hemispherical impactor insert (40 mm length, 20 mm diameter) was
employed. The system included an anti-rebound mechanism, a precise photocell for measurement,
and a lubrication system. The testing chamber, maintained at room temperature, contained a
specimen holder, clamping mechanism, and temperature control. The specimen was clamped with
a pressure of 5.5 bars. Parameters for low-velocity impact tests are detailed in in Table 9. Impact
velocity denotes the speed at which the impactor insert contacts the specimen surface. Based on
the findings presented in Figure 17, it becomes evident that the peak impact force and energy
observed in the developed composite polymers significantly increased with the incorporation of
2wt.% CaCO3/x wt.% GNPs. Furthermore, the residual velocity decreased considerably with
increasing the content of 2wt.% CaCO3/x wt.% GNPs in PP matrix. The reasons for the improved
impact properties were attributed to nucleating effects of incorporated GNPs in PP matrix, strong
interactions between GNPs & CaCO3 with PP matrix, and effective dispersion of GNPs as
discussed in earlier section.

27
Figure 17. Low velocity behaviour of pure PP matrix, PP/2CC/GNPs-0.5, PP/2CC/GNPs-1.0,
PP/2CC/GNPs-1.5, and PP/2CC/GNPs-2.0 composite polymers: (a) force Vs displacement
profile; (b) energy Vs displacement profile; and (c) velocity Vs displacement profile.

Table 9: Low-velocity impact testing parameters used in the testing of developed high-entropy
polymers

Parameter Value Parameter Value

Insert
Applied impact energy 75 J 20 mm
diameter

Environme
Impact velocity 3.11 m/s ntal 20 oC
temperature

Striker load
Falling height 493 mm, 90 KN
capacity

Impactor head mass 15.5 kg

28
3.8 Analysis of Fracture Surface Topography

Figure 18. SEM microstructure of fractured area after tensile test of: (a1-a3) pure PP matrix; (b1-
b3) PP/2CC/GNPs-1.0, and (c1-c3) PP/2CC/GNPs-2.0 hybrid composite polymers. Yellow
dashed circle represents the location of GNPs/CaCO3 fillers over PP matrix
Fracture surface analysis is a pivotal technique in studying polymer composites, providing crucial
insights into behavior and failure mechanisms by examining the fractured surfaces post-
mechanical testing. SEM microstructures after tensile and impact tests were captured and are
depicted in Figure 18 and Figure 19, respectively. Ductile SEM morphology with regular wave-
like patterns characterized the pure PP matrix due to proper compounding [52] ( Figure 18a1-a3).
Notably, the pure PP matrix showed an absence of cracks, voids, and inclusions. In contrast,
PP/2CC/GNPs-1.0 exhibited a rough surface with uniform GNPs pull-outs, indicating effective
load-transfer between GNPs & CaCO3 with PP matrix (Figure 18b1-b3). However, PP/2CC/GNPs-

29
2.0 showed a very rough surface with more voids, cracks, cleavage patterns, and non-uniform
GNPs pull-outs due to agglomeration and increased defects (Figure 18c1-c3), resulting in lower
mechanical properties (Table 6, Table 7, and Table 8). Beyond 1 wt.% of GNPs, the presence of
more defects, cleavage patterns, and non-uniform dispersion led to brittle fracture behavior,
deteriorating mechanical properties. Similarly, after the impact test, a homogeneous ductile
fracture surface was obtained in the pure PP matrix (Figure 19a1-a3). PP/2CC/GNPs-1.0 showed a
slightly rough surface with a brittle flat surface (Figure 19b1-b3), while PP/2CC/GNPs-2.0
exhibited a severely rough, completely brittle fracture surface with more damages (Figure 19c1-
c3), indicating the presence of defects and reduced bonding strength (Figure 17)

Figure 19. SEM microstructure of fractured area after impact test of: (a1-a3) pure PP matrix; (b1-
b3) PP/2CC/GNPs-1.0, and (c1-c3) PP/2CC/GNPs-2.0 composite polymers. Yellow dashed circle
represents the location of GNPs/CaCO3 fillers over PP matrix

30
5.0 Conclusions
 Five hybrid composite polymers, including pure PP matrix, PP/2CC/GNPs-0.5, PP/2CC/GNPs-
1.0, PP/2CC/GNPs-1.5, and PP/2CC/GNps-2.0, were successfully synthesized using melt
compounding followed by injection molding techniques.
 Macroscopic and microscopic examinations confirmed the purity and morphology of the
synthesized polymers. HRSEM morphological analyses revealed an average size of agglomerated
GNPs at 6±21 µm with a thickness of 5 nm.
 Characterization of the composite polymers using FTIR, XRD, DSC, TGA, and XPS techniques
indicated the presence of certain links between the incorporated polymer blend and GNPs.
 XRD analysis showed a varying K value, with PPGNPs-2.0 exhibiting a higher value, indicating
more GNPs crystals.
 FTIR results confirmed symmetric vibrations in the wave length range of 2963-3015 cm-1, and
symmetric vibrations in the range of 2815-29625 cm-1, indicating the copolymer nature and strong
bonding of incorporated GNPs in the PP matrix.
 DSC analyses revealed significant changes in melting and recrystallization temperatures,
suggesting a strong nucleation effect with GNPs introduction.
 Mechanical properties from tensile, compression, and three-point bending tests showed improved
behavior in composite samples with GNPs incorporation, attributed to strong nucleating and
bonding effects
 Low-velocity impact tests demonstrated enhanced performance with GNPs incorporation,
attributed to effective dispersion, interconnection, and bonding between GNPs and PP matrix.
 HRSEM fracture surface analyses revealed microstructural features, with PP/2CC/GNPs-1.0
showing improved characteristics compared to other composites.
 Based on the findings, PP/2CC/GNPs-1.0 hybrid composite is recommended for structural and
other applications due to its improved thermal and mechanical properties

Author Contributions: Conceptualization, D.B., R.K., S.S., A.E.; validation, D.B., R.K., S.S.,
M.S; formal analysis, D.B., A.A., R.K. ; investigation, R.K., S.S., M.S., A.A.; resources, R.K.,
S.S., A.A.; data curation, R.K., S.S., M.S., A.A.; writing—original draft preparation, R.K., S.S.,
M.S., A.A., writing—review and editing, S.S., M.S., A.A.; supervision, R.K., S.S., M.S., All
authors have read and agreed to the published version of the manuscript.

Funding: The authors received no specific funding for this study.

31
Conflicts of Interest: The authors declare no conflict of interest.

Declaration of generative AI in scientific writing: During the preparation of this work, the
author(s) have not used any AI tool.

References

[1] P. Hu, H. Yang, Polypropylene filled with kaolinite-based conductive powders, Appl.
Clay Sci. 83 (2013) 122–128.

[2] P. Cataldi, A. Athanassiou, I.S. Bayer, Graphene nanoplatelets-based advanced materials


and recent progress in sustainable applications, Appl. Sci. 8 (2018) 1438.

[3] J.R. Potts, D.R. Dreyer, C.W. Bielawski, R.S. Ruoff, Graphene-based polymer
nanocomposites, Polymer (Guildf). 52 (2011) 5–25.

[4] Y. Geng, J. Li, S.J. Wang, J.K. Kim, Amino functionalization of graphite nanoplatelet, J.
Nanosci. Nanotechnol. 8 (2008) 6238–6246.

[5] I.M. Inuwa, A. Hassan, S.A. Samsudin, M.K.M. Haafiz, M. Jawaid, K. Majeed, N.C.A.
Razak, Characterization and mechanical properties of exfoliated graphite nanoplatelets
reinforced polyethylene terephthalate/polypropylene composites, J. Appl. Polym. Sci. 131
(2014).

[6] M. Karevan, K. Kalaitzidou, Formation of a complex constrained region at the graphite


nanoplatelets-polyamide 12 interface, Polymer (Guildf). 54 (2013) 3691–3698.

[7] K. Kalaitzidou, H. Fukushima, L.T. Drzal, Mechanical properties and morphological


characterization of exfoliated graphite–polypropylene nanocomposites, Compos. Part A
Appl. Sci. Manuf. 38 (2007) 1675–1682.

[8] A.J. Duguay, J.W. Nader, A. Kiziltas, D.J. Gardner, H.J. Dagher, Exfoliated graphite
nanoplatelet-filled impact modified polypropylene nanocomposites: influence of particle
diameter, filler loading, and coupling agent on the mechanical properties, Appl. Nanosci.
4 (2014) 279–291.

[9] J.N. Coleman, U. Khan, W.J. Blau, Y.K. Gun’ko, Small but strong: a review of the
mechanical properties of carbon nanotube–polymer composites, Carbon N. Y. 44 (2006)
1624–1652.

[10] V.A. Beloshenko, A. V Voznyak, Y. Voznyak, L.A. Novokshonova, V.G. Grinyov, Effect

32
of simple shear induced orientation process on the morphology and properties of
polyolefin/graphite nanoplates composites, Compos. Sci. Technol. 139 (2017) 47–56.

[11] C. Vilaverde, R.M. Santos, M.C. Paiva, J.A. Covas, Dispersion and re-agglomeration of
graphite nanoplates in polypropylene melts under controlled flow conditions, Compos.
Part A Appl. Sci. Manuf. 78 (2015) 143–151.

[12] B.Z. Jang, A. Zhamu, Processing of nanographene platelets (NGPs) and NGP
nanocomposites: a review, J. Mater. Sci. 43 (2008) 5092–5101.

[13] F.-A. He, H.-J. Wu, X.-L. Yang, K.-H. Lam, J.-T. Fan, L.-W.H. Chan, Novel exfoliated
graphite nanoplates/syndiotactic polystyrene composites prepared by solution-blending,
Polym. Test. 42 (2015) 45–53.

[14] I. Taraghi, A. Fereidoon, S. Paszkiewicz, Z. Roslaniec, Nanocomposites based on polymer


blends: enhanced interfacial interactions in polycarbonate/ethylene-propylene copolymer
blends with multi-walled carbon nanotubes, Compos. Interfaces. 25 (2018) 275–286.

[15] A.A. Ramachandran, L.P. Mathew, S. Thomas, Effect of MA-g-PP compatibilizer on


morphology and electrical properties of MWCNT based blend nanocomposites: New
strategy to enhance the dispersion of MWCNTs in immiscible poly (trimethylene
terephthalate)/polypropylene blends, Eur. Polym. J. 118 (2019) 595–605.

[16] A.F. Jolfaei, J.N. Gavgani, A. Jalali, F. Goharpey, Effect of organoclay and
compatibilizers on microstructure, rheological and mechanical properties of dynamically
vulcanized EPDM/PP elastomers, Polym. Bull. 72 (2015) 1127–1144.

[17] M. Yousfi, S. Livi, A. Dumas, J. Crépin‐ Leblond, M. Greenhill‐ Hooper, J. Duchet‐


Rumeau, Compatibilization of polypropylene/polyamide 6 blends using new synthetic
nanosized talc fillers: Morphology, thermal, and mechanical properties, J. Appl. Polym.
Sci. 131 (2014).

[18] M.A. Al-Saleh, A.A. Yussuf, S. Al-Enezi, R. Kazemi, M.U. Wahit, T. Al-Shammari, A.
Al-Banna, Polypropylene/graphene nanocomposites: Effects of GNP loading and
compatibilizers on the mechanical and thermal properties, Materials (Basel). 12 (2019)
3924.

[19] S. Wang, A. Ajji, S. Guo, C. Xiong, Preparation of microporous polypropylene/titanium

33
dioxide composite membranes with enhanced electrolyte uptake capability via melt
extruding and stretching, Polymers (Basel). 9 (2017) 110.

[20] R. Alaburdaite, E. Paluckiene, S. Grevys, Comparison of the surface characteristics of


polyethylene and polypropylene films and polyester textile coated with electroconductive
copper sulphide thin films, Chalcogenide Lett. 13 (2016) 529–536.

[21] N. Yang, Z.-C. Zhang, N. Ma, H.-L. Liu, X.-Q. Zhan, B. Li, W. Gao, F.-C. Tsai, T. Jiang,
C.-J. Chang, Effect of surface modified kaolin on properties of polypropylene grafted
maleic anhydride, Results Phys. 7 (2017) 969–974.

[22] R. Meena, A.W. Hashmi, S. Ahmad, F. Iqbal, H. Soni, A. Meena, A.A. Al-Kahtani, B.
Pandit, H. Kamyab, H. Payal, Influence of fly ash on thermo-mechanical and mechanical
behavior of injection molded polypropylene matrix composites, Chemosphere. (2023)
140225.

[23] S. Zafar, Effect of microwave power on the hole characteristics in microwave-drilled


kenaf/polypropylene composites, J. Manuf. Process. 102 (2023) 218–230.

[24] S. Tirlangi, M.J.S. Mohamed, K. Karthik, A.G.M. Gandhi, K. Thiruselvam, Influence on


the mechanical properties of virgin recycled polypropylene composites with different
types of reinforcing loads, Mater. Today Proc. (2023).

[25] P. Jan, S. Matkovič, M. Bek, L.S. Perše, M. Kalin, Tribological behaviour of green wood-
based unrecycled and recycled polypropylene composites, Wear. 524 (2023) 204826.

[26] Á. Görbe, L.J. Varga, T. Bárány, Development of nanoparticle-filled polypropylene-based


single polymer composite foams, Heliyon. 9 (2023).

[27] Y.W. Leong, M.B. Abu Bakar, Z.A.M. Ishak, A. Ariffin, B. Pukanszky, Comparison of
the mechanical properties and interfacial interactions between talc, kaolin, and calcium
carbonate filled polypropylene composites, J. Appl. Polym. Sci. 91 (2004) 3315–3326.

[28] K.Z. Ahmed, M. Faizan, F. Azam, A. Faheem, Hardness assessment of novel waste tire
rubber-polypropylene composite, Mater. Today Proc. (2023).

[29] P. Anandakumar, M.V. Timmaraju, R. Velmurugan, Low-velocity impact behavior of


injection over-molded short/continuous fiber reinforced polypropylene composites, Mater.
Today Proc. (2023).

34
[30] O.A. Balogun, O.O. Daramola, A.A. Adediran, A.A. Akinwande, O.S. Bello,
Investigation of Jute/Tetracarpidium conophorum reinforced polypropylene composites
for automobile application: Mechanical, wear and flow properties, Alexandria Eng. J. 65
(2023) 327–341.

[31] R. Rothon, C. Paynter, Calcium Carbonate Fillers BT - Fillers for Polymer Applications,
in: R. Rothon (Ed.), Springer International Publishing, Cham, 2017: pp. 149–160.
https://doi.org/10.1007/978-3-319-28117-9_35.

[32] S.G. Prolongo, R. Moriche, A. Jiménez-Suárez, M. Sánchez, A. Ureña, Advantages and


disadvantages of the addition of graphene nanoplatelets to epoxy resins, Eur. Polym. J. 61
(2014) 206–214.

[33] S.-Y. Yang, W.-N. Lin, Y.-L. Huang, H.-W. Tien, J.-Y. Wang, C.-C.M. Ma, S.-M. Li, Y.-
S. Wang, Synergetic effects of graphene platelets and carbon nanotubes on the mechanical
and thermal properties of epoxy composites, Carbon N. Y. 49 (2011) 793–803.

[34] L.B. Fitaroni, J.A. de Lima, S.A. Cruz, W.R. Waldman, Thermal stability of
polypropylene–montmorillonite clay nanocomposites: Limitation of the
thermogravimetric analysis, Polym. Degrad. Stab. 111 (2015) 102–108.

[35] J.M. Chacón, M.A. Caminero, E. García-Plaza, P.J. Núñez, Additive manufacturing of
PLA structures using fused deposition modelling: Effect of process parameters on
mechanical properties and their optimal selection, Mater. Des. 124 (2017) 143–157.
https://doi.org/10.1016/j.matdes.2017.03.065.

[36] F.A. Shishevan, H. Akbulut, M.A. Mohtadi-Bonab, Low velocity impact behavior of
basalt fiber-reinforced polymer composites, J. Mater. Eng. Perform. 26 (2017) 2890–2900.

[37] A. Standard, D7136: Standard test method for measuring the damage resistance of a fiber-
reinforced polymer matrix composite to a drop-weight impact event, ASTM Int. West
Conshohocken. (2005).

[38] P. Niu, B. Liu, X. Wei, X. Wang, J. Yang, Study on mechanical properties and thermal
stability of polypropylene/hemp fiber composites, J. Reinf. Plast. Compos. 30 (2011) 36–
44.

[39] X.L. Xie, K.L. Fung, R.K.Y. Li, S.C. Tjong, Y. Mai, Structural and mechanical behavior

35
of polypropylene/maleated styrene‐ (ethylene‐ co‐ butylene)‐ styrene/sisal fiber
composites prepared by injection molding, J. Polym. Sci. Part B Polym. Phys. 40 (2002)
1214–1222.

[40] J.S. Blocker, Alcohol and temperance in modern history: An international encyclopedia,
ABC-CLIO, 2003.

[41] V. Gutmann, Halogen chemistry, Elsevier, 2012.

[42] Z. Yang, X. Wang, J. Wang, Y. Yao, H. Sun, N. Huang, Pulsed‐ Plasma Polymeric
Allylamine Thin Films, Plasma Process. Polym. 6 (2009) 498–505.

[43] J. Dai, X. Liu, Y. Xiao, J. Yang, P. Qi, J. Wang, Y. Wang, Z. Zhou, High hydrophilicity
and excellent adsorption ability of a stretched polypropylene/graphene oxide composite
membrane achieved by plasma assisted surface modification, RSC Adv. 5 (2015) 71240–
71252.

[44] Z. Yang, J. Wu, X. Wang, J. Wang, N. Huang, Inspired Chemistry for a Simple but Highly
Effective Immobilization of Vascular Endothelial Growth Factor on Gallic Acid‐
functionalized Plasma Polymerized Film, Plasma Process. Polym. 9 (2012) 718–725.

[45] T. Xu, Y. Li, J. Chen, H. Wu, X. Zhou, Z. Zhang, Improving thermal management of
electronic apparatus with paraffin (PA)/expanded graphite (EG)/graphene (GN) composite
material, Appl. Therm. Eng. 140 (2018) 13–22.

[46] G. Moradkhani, M. Fasihi, T. Parpaite, L. Brison, F. Laoutid, H. Vahabi, M.R. Saeb,


Phosphorization of exfoliated graphite for developing flame retardant ethylene vinyl
acetate composites, J. Mater. Res. Technol. 9 (2020) 7341–7353.

[47] K.I. Hadjiivanov, D.A. Panayotov, M.Y. Mihaylov, E.Z. Ivanova, K.K. Chakarova, S.M.
Andonova, N.L. Drenchev, Power of infrared and raman spectroscopies to characterize
metal-organic frameworks and investigate their interaction with guest molecules, Chem.
Rev. 121 (2020) 1286–1424.

[48] J. Fang, L. Zhang, D. Sutton, X. Wang, T. Lin, Needleless melt-electrospinning of


polypropylene nanofibres, J. Nanomater. 2012 (2012).

[49] H.U. Zaman, P.D. Hun, R.A. Khan, K.-B. Yoon, Polypropylene/clay nanocomposites:
Effect of compatibilizers on the morphology, mechanical properties and crystallization

36
behaviors, J. Thermoplast. Compos. Mater. 27 (2014) 338–349.

[50] K. Gaska, X. Xu, S. Gubanski, R. Kádár, Electrical, mechanical, and thermal properties of
LDPE graphene nanoplatelets composites produced by means of melt extrusion process,
Polymers (Basel). 9 (2017) 11.

[51] M. El Achaby, F. Arrakhiz, S. Vaudreuil, A. el Kacem Qaiss, M. Bousmina, O. Fassi‐


Fehri, Mechanical, thermal, and rheological properties of graphene‐ based polypropylene
nanocomposites prepared by melt mixing, Polym. Compos. 33 (2012) 733–744.

[52] G.B. Olowojoba, S. Eslava, E.S. Gutierrez, A.J. Kinloch, C. Mattevi, V.G. Rocha, A.C.
Taylor, In situ thermally reduced graphene oxide/epoxy composites: thermal and
mechanical properties, Appl. Nanosci. 6 (2016) 1015–1022.

37
Declaration of Interest Statement

Declaration of interests

☐ The authors declare that they have no known competing financial interests or personal relationships
that could have appeared to influence the work reported in this paper.

☐The authors declare the following financial interests/personal relationships which may be considered
as potential competing interests:

You might also like