Published at : 10 Jul 2024
Volume : IJtech
Vol 15, No 4 (2024)
DOI : https://doi.org/10.14716/ijtech.v15i4.5746
Dwi Setyaningsih | 1 Department of Agroindustrial Technology, IPB University, Bogor 16680, West Java, Indonesia 2 Surfactant and Bioenergy Research Center, IPB University, Bogor 16144, West Java, Indonesia |
Muhammad Syukur Sarfat | 1 Department of Agroindustrial Technology, IPB University, Bogor 16680, West Java, Indonesia 2 Surfactant and Bioenergy Research Center, IPB University, Bogor 16144, West Java, Indonesia |
Farah Fahma | Department of Agroindustrial Technology, IPB University, Bogor 16680, West Java, Indonesia |
Nastiti Siswi Indrasti | Department of Agroindustrial Technology, IPB University, Bogor 16680, West Java, Indonesia |
This research aims to identify the performance and mechanical characteristics of polypropylene-based antistatic bio-nano composites reinforced with 2% mono-diacylglycerols (M-DAG) and 0, 0.5, 2.5% cellulose nanocrystals (CNC) and compared to pure PP. Based on the results of SEM on cross-section, there was an agglomeration of M-DAG and CNC on the PP matrix. XRD diffractogram pattern of antistatic bio-nano composites showed peaks representing the diffraction structure of glycerol monostearate and cellulose I. The FTIR spectrum of each formulation was very similar to the characteristic peaks of PP and showed three distinct peaks compared to pure PP. The melting temperature of antistatic composites without CNC (176.54oC) was higher than pure PP (175.44oC). Thermal stability of antistatic bio-nano composites with 0, 0.5, and 2.5% CNC were 472.07, 470.25, and 475.15 oC, respectively, higher than pure PP (468.27oC). The best mechanical properties were 2.5% CNC with 11.071 MPa flexural modulus, 30.376 MPa tensile strength, 23.796% tensile elongation, 1.659 J/m2 impact strength, which were higher than pure PP, and it generated antistatic activity of 1010 - 1012 /sq resistivity.
Antistatic bio-nano composites; Biopolymers synthesis; Cellulose nanocrystals; Mono-diacylglycerols; Polypropylene
The trend of using synthetic polymer-based materials is predicted to increase in terms of fulfilling human needs. Synthetic polypropylene (PP) is known for having a high softening point or glass transition, high resistance to bending stress, low water absorption, good electrical resistance, light dimensional stability, high impact strength, non-toxicity properties, and the degree of crystallinity ranges from 40 to 60% with the melting temperature range from 130 to 171°C (Shubhra, Alam, and Quaiyyum, 2013). But PP is susceptible to high temperatures, flammable, prone to UV degradation, susceptible to oxidation, difficult to paint, and harmful to the environment due to its non-degradable nature (Purnomo, Baskoro, and Muslim, 2021; Shieddieque et al., 2021), but PP is recyclable (Jain and Tiwari, 2015). Therefore, to overcome the weakness, it is necessary to modify PP into bio-nano composites. Antistatic bio-nano composites are synthesized by adding natural antistatic materials.
Nanocomposites are multicomponent materials consisting of several
different phases in which at least one phase size is in the nanoscale (less
than 100 nm) (Sandri et
al., 2016). The antistatic bio-nano
composites synthesized using mono-diacylglycerols (M-DAG) as an antistatic agent (Salsabila et
al., 2021), cellulose nanocrystals (CNC) as a reinforcement, and PP
as a thermoplastic matrix (Clemons and
Rick, 2020; Sabaruddin, Md-Tahir, and Lee, 2019; Gwon et al., 2018).
A combination of M-DAG and CNC is expected to
produce a synergistic effect to improve the quality
of the antistatic bio-nano composites. The
addition of M-DAG and CNC to the PP matrix had a positive impact on
the characteristics of the resulting bio-nano composites and antistatic bio-nano composites (Sabaruddin, Md-Tahir, and Lee, 2019; Gwon et al.,
2018).
M-DAG
is an ester of glycerol and free fatty acid (FFA)
which has unreacted or free hydroxyl groups. MAG, or
monoglyceride, has a single fatty acyl chain, while DAG, or diacylglyceride,
has two fatty acyl chains (Sarfat et al., 2022).
This free hydroxyl group makes M-DAG a non-ionic surfactant that is degradable
and bio-compatible, so it is widely used in the food, cosmetic, and
pharmaceutical industries. This free hydroxyl group allows M-DAG to be used as
a stabilizer and an antistatic agent in bio composites or plastics (Salsabila et al., 2021).
M-DAG used in this research was produced from palm fatty acid distillates
(PFAD) from the refining process of crude palm oil (CPO) (Setyaningsih, Suwarna, and Muna, 2020; Setyaningsih et
al., 2020).
CNC
is a cellulose-based nanomaterial that has better mechanical characteristics
such as tensile strength (7.5 GPa) (Tang et al., 2017),
tensile modulus (100 - 140 GPa) (Tang et al., 2017),
high surface area (569 m²/g) (Brinkmann et al., 2016),
with a diameter average of 8 nm (Rochardjo et al., 2021) compared
to other cellulose-based nanomaterials such as cellulose nanofiber (CNF) which
has a tensile
strength of 0.3833 GPa (Kafy et al., 2017),
a tensile modulus of 23.9 GPa (Kafy et al., 2017) and
surface area of 430 m²/g (Moser, Henriksson,
and Lindström, 2016).
However, there are disadvantages of CNC, namely low stability starting from
283.55 0C, which causes CNC to be very
susceptible to high heat treatment when used as a reinforcing material in
polymer matrices. M-DAG can be used as a lubricant and stabilizer to protect CNC from thermal degradation during processing.
Therefore,
a combination of M-DAG and CNC as additive materials for the synthesis of
PP-based antistatic bio-nano composites has a prospect as next-generation
material
that is more flexible in use and exhibits superior
performance and mechanical characteristics compared to pure PP. There has never
been researched that combines M-DAG and CNC simultaneously as reinforcement in
the PP matrix. Therefore, this study aims to evaluate the performance and
mechanical characteristics of PP-based antistatic bio-nano composites
reinforced with varying concentrations of M-DAG and CNC, and compare them to
those of pure PP.
2.1. Raw
Materials
The raw materials for synthesizing
antistatic bio-nano composites were PP, M-DAG, and CNC. PP (PT Chandra Asri Petrochemical Tbk) has a melt
flow index of 10 g/10 minute and a density of 0.903 g/cm3.
M-DAG (SBRC-LPPM-IPB) has a crystallinity index of 92.85%, a diameter of 0.11–1.78 nm, and a thermal degradation rate of 200.50 oC. CNC (CelluForce Co.) has a density of 1.5 g/cm3, a crystallinity index of 98.95%, and diameter of 3.39–12.72 nm, a length of 44 – 108 nm, and thermal degradation of 296.96 oC. The supporting materials consist of maleic
anhydride polypropylene (MAPP)(BYK Chemie GmbH),
antioxidant (AO) (BASF Schweiz AG, Switzerland), and mineral oil (MO) (Arkema France).
2.2. Methodology
Figure 1 The Stages
of the synthesis of antistatic bio-nano composites
2.2. Antistatic
Bio-nano Composites Characterization
Infrared
spectrum analyzed using Fourier
transform IR (FT-IR) Thermo Fisher Scientific Nicolet iS5 spectrophotometer
with cleaning pump and wavelengths 300 cm-1 to 4000 cm-1,
128 accumulated scans, resolution 4 cm-1, in ATR and transmittance
module. Thermal properties analysis used differential scanning calorimetry
(DSC) TA Instruments, New Castle, UK model Q200. Dynamic DSC scans were conducted in the temperature range from 23 to 400 °C at a
heating rate of 10 °C/min. The crystallization and melting behaviors were
recorded in a nitrogen atmosphere, at the range mass used of 21.8 to 29.0 mg.
where W is the corrected energy absorbed by breaking the specimen (J), is the thickness of the specimen (mm), and is the width of the specimen (mm).
where is flexure strength
(N/mm2/MPa), is flexure
modulus (N/mm2/MPa),
where
3.1. Morphology
Analysis
Figure 2
Morphological analysis of antistatic bio-nano composites using SEM
Based on the results of the SEM analysis on the cross-section, M-DAG
agglomeration was found on the PP matrix because it shows a morphological form
of M-DAG, while the CNC is not visible. This indicates a chemical reaction or
physical interaction between M-DAG and the PP matrix. Physical interaction
occurs when the polar groups (palmitate) are oriented to the PP matrix. In
contrast, the polar groups (glyceryl) are oriented away from the PP matrix
towards the antistatic bio-nano composite surface, and it is possible to interact physically with CNC. Another
possibility is that the polar group (glyceryl) reacts with the O group of the
maleic anhydride during the synthesis process. In addition, there was no
fibrillation on the surface of the antistatic
bio-nano composites cross-section. This implies that
no significant plastic deformation occurred in the antistatic bio-nano composites layer during fracture, although the CNC concentration is increased. According
to Shojaeiarani, Bajwa, and
Chanda (2021), the
rheological properties of CNC-filled polymer melts depend on factors such as
the degree of CNC particle dispersion, their orientation state, and aspect
ratio. Consequently, the rheological properties of the antistatic bio-nano
composites in our study were also influenced by these factors. In particular,
the homogeneous dispersion of CNC particles led to more effective load transfer
to the reinforcements.
3.2. Degree of Crystallinity and
Particle Size Analysis
Crystallinity analysis was carried out on AS-BNC using XRD analysis to observe changes in the crystal structure as a function of chemical treatment and to measure the degree of crystallinity (CrI) using the deconvolution method with a ratio between the area corresponding to the crystal peak and the total area (Park et al., 2010). Figure 3A shows the XRD diffractogram of AS-BNC. XRD diffractogram patterns were recorded using Cu K irradiation, = 1.5418 A.
The results of the XRD diffractogram analysis of AS-BNC showed with seven peaks with peak heights ranging from 87.09 – 1000 I rel of AS-BNC-5, 84.61 – 1000 I rel of AS-BNC-0.5, and 90.91 – 1000 I rel of AS-BNC-2.5 (Table 2). The XRD diffractogram pattern showed peaks representing the diffraction structure of glycerol monostearate and cellulose I. The diffraction characteristic of glycerol monostearate can be observed in the peak range between 5o to 30o (Yusuf et al., 2013). The diffraction characteristics of cellulose I could be observed around peaks at 15o (001), 22.5o (002), and 34o (040) (Park et al., 2010), Couret et al. (2017) said the peaks at 15o (1-10), 17o (110), 21o(102/012), 23o (200), and 34o (004) represent the diffraction structure of cellulose I.
The XRD diffractogram of PP showed the pattern, which has seven peaks with peak heights ranging from 85.41 – 1000 I rel. According to Guerra, Wan, and McNally (2019), the XRD patterns showed the most intense peaks for PP at = 16.5o (100), 19.2o (300), 20o (040), 22o (130), and for the GNPs at = 32o (002).
The identification results of the
particle size, degree of crystallinity, and percent of amorphous components in
AS-BNC showed the average particle
sizes between 15.84 to 16.01
nm, 89.81
to 91.10 % degree of
crystallinity, and 8.90 to 10.19
% amorphous components (Table
2). The
particle size of AS-BNC was increased with the increase in CNC concentration.
The degree of crystallinity of AS-BNC-2.5 is higher than that of pure PP.
3.3. Infrared Spectrum Analysis
The FTIR spectrum of each treatment (Figure 3B) was very similar to the characteristic peaks of PP and depicted distinct
tri peaks for the antistatic bio-nano composites compared to the pure PP. The
first distinct absorption peak ranging from 1200 cm-1 to 1000 cm-1
were primarily assigned to C–O–C bond, C–C bending, and ring structures (with
typical sharpening at 1071.50 cm-1 with %T of 85.14% and 1166.98 cm-1
with %T of 69.93% (AS-BNC-0), 1080.18 cm-1 with %T of 83.03% and
1166.02 cm-1 with %T of 64.77% (AS-BNC-0.5), and 1085.01 cm-1
with %T of 74.86% and 1166.02 cm-1 with %T of 51.02% (AS-BNC-2.5),
which correspond to typical cellulose and glycerol compound (Al-Haik et al. 2020). In the band around 1080 cm-1, the %T decreased as the
concentration of CNC added to the PP matrix increased, while the %T in Pure PP
was lower than each of the AS-BNC, which was 70.79%. This indicated the
presence of an increasing C–O–C bond due to ring deformation of maleic acid compounds
or ring widening and C–O stretching. The same phenomenon occurred in the band
around 1166 cm-1, which showed a decrease of %T with increasing
concentration of CNC addition to the PP matrix, but lower than Pure PP (68.11%),
except for AS-BNC-0. This indicated the presence of an increasing C–C bending.
The second distinct band is related to the wavelength near 2900 cm-1. This broadband was assigned to stretching vibration of C–H hydroxyl groups asymmetric stretching of cellulose and glycerol. In this band, % transmittance (T) decreased as the concentration of CNC increased, while %T in pure PP was higher than AS-BNC. The presence of these bands confirmed the interaction of CNC and the distribution of M-DAG in the PP matrix. Due to this interfacial adhesion, the overall mechanical properties were enhanced for the AS-BNC. Hobuss et al. (2020) determined the asymmetric and symmetric C–H stretching mode of the fatty acid chain methylene group at 2922 cm-1 and 2853 cm-1. The third distinct peak was 3315 cm-1, which is related to O–H [ (O–H)] stretching, a characteristic of hydroxyl groups. This band showed that %T decreased as the concentration of CNC increased, while %T of pure PP was higher than AS-BNC. Hobuss et al. (2020) set the O–H stretch at 3360 cm-1.
The
bands at 1377.23 cm-1 and 1457.28 cm-1 in all treatments
were characteristics of PP (Fang et al., 2012).
The absorption bands at 1738.90 cm-1 (Pure PP), 1742.76 cm-1
(AS-BNC-0), 1738.90 cm-1 (AS-BNC-0.5), and 1742.76 cm-1
(AS-BNC-2.5) were observed, which were assigned to the absorption of carbonyl
groups (C=O) of maleic anhydride
(MA) (Rahman, Hassan, and Heidarian, 2018; Zhou et al., 2013). Finally, the spectrum on the peak 1166.02 cm-1 indicated
the C-C bending, which was the backbone of PP (Fang et al., 2012).
The
infrared (IR) spectrum of the antistatic bio-nano composites revealed several
characteristic peaks. Specifically, peaks between 3300 cm-1 and 3250
cm-1 were assigned to O–H stretching modes, while those between 3000
cm-1 and 2750 cm-1 corresponded to the stretching modes
of CH, CH2, and CH3 groups. The peak at 1750 cm-1
was indicative of carbonyl stretching (C = O).
The peaks between 1500 cm-1 to 1250 cm-1 were
characteristic of the deformation of the CH2 and CH3
groups. The peaks between 1250 cm-1 to 1150 cm-1 were
referred as the C–O and C–C bonds. The peak at 1100 cm-1 was also
characteristic of the strain of the C–O bond and ester group (–C–O–C–). The
"wag" vibration and asymmetric angular deformation of CH and CH2
groups were found at 750 cm-1 (Hobuss et
al., 2020).
3.4. Thermal Properties and Melt
Flow Index Analysis
Based on the identification results of the thermal
properties (Figure 3C and Table 3),
it was found that the melting temperature
of AS-BNC-0 was 176.54oC
higher than that of pure PP 175.44 oC.
The melting temperature of AS-BNC-0.5
and AS-BNC-2.5 were 171.70oC and
174.38oC,
respectively, slightly lower than that of pure PP. However,
when compared between AS-BNC-0.5 and AS-BNC-2.5, AS-BNC-2.5 had a higher
melting temperature than AS-BNC-0.5. This indicated an opportunity for
increasing melting temperature with an increase in CNC concentration. According to Al-Haik et
al. (2020) and Hejna et al. (2017),
the melting temperature of bio-nano composites with the addition of 2%, 4%, and
5% CNC on the PP matrix showed a higher value when compared to pure PP and the
addition of 4% CNC had the higher melting temperature than 3% CNC. According to
Yousefian and Rodrigue (2016),
the distribution of CNC particles in the polymer matrix greatly influenced the
thermal properties of the resulting bio-nano composites.
Thermal stability of AS-BNC-0, AS-BNC-0.5,
and AS-BNC-2.5 was 472.07 oC, 470.25oC, 475.15oC respectively, higher
than pure PP (468.27 oC).
Therefore, the addition of 2.5%
CNC and 2% M-DAG to
the PP matrix can increase the thermal degradation of the resulting antistatic
bio-nano composites. The higher the degree of crystallinity of the antistatic
bio-nano composites, the higher the thermal stability, except for the
antistatic bio-nano composites treated with AS-BNC-0. According to Al-Haik et al. (2020),
the thermal stability of bio-nano composites with the addition of 1%, 2%, and 3
% CNC on the PP matrix showed a higher value when compared to pure PP and the
addition of 4% and 5 % CNC. It showed that the addition
of 3 % CNC has the highest thermal
stability. CNC particles were thought to increase the thermal
resistance of AS-BNC by inhibiting the diffusion of volatile decomposition
products or by forming a charred CNC surface that
dissipates heat by absorbing it in the inorganic phase (Thomas et al., 2018; Ng et al.,
2017).
In addition, the presence of M-DAG can inhibit the thermal degradation of
AS-BNC. The reduced thermal resistance in AS-BNC with 2% M-DAG and 0.5% CNC may have been due to the non-uniform dispersion of the CNC particles (Ng et al.,
2017).
Figure 3
XRD diffractogram (A), FT-IR spectra (B), and DSC thermogram of AS-BNC
3.5. Mechanical Properties Analysis
Analysis of mechanical properties of AS-BNC includes
tensile properties (Table 4) and flexural, impact strength, and surface
resistivity properties (Table 5). The best AS-BNC was AS-BNC-2.5
with 11.071 MPa flexural modulus, 30.376 MPa tensile strength, 23.796% tensile
elongation, 1.659 J/m2 impact strength, and 1010 – 1012 /sq resistivity, which was higher than that of pure PP. However, AS-BNC-2.5 has 207.244 MPa modulus,
32.092 MPa highest tensile strength, 29.120% tensile
elongation, and 90.6950 Pa
flexural strength, which was smaller than pure PP. The higher the degree of
crystallinity of the antistatic bio-nano composites, the higher the tensile
strength, but It’s not for other mechanical properties of antistatic bio-nano
composites. According to Shojaeiarani, Bajwa, and
Chanda (2021), CNC has a large surface-to-volume ratio, high tensile strength (10 GPa), high
stiffness (110–130 GPa), and high flexibility. Incorporating CNC in composites influenced their mechanical properties by improving
filler-matrix compatibility and forming a filler network. Using CNC as a
reinforcement agent improved the stress-transfer efficiency of the composites
and the overall mechanical properties Shojaeiarani, Bajwa, and Chanda (2021).
Based on the results of the electrical resistivity of AS-BNC, the values were in the range of antistatic values of 1010 – 1012 /sq. In addition, the surface resistivity value of AS-BNC was in the static dissipative value range of 106 – 1012 /sq, which has the potential to be used for electrostatic discharge prevention or the presence of a sudden electric current caused by an electric short circuit, a dielectric fault, or the contact between two electrically charged objects (Pang et al., 2014). In general, the polymer matrix is highly insulated. Therefore, the presence of electrically conductive nanomaterials with a large aspect ratio dispersed in small quantities can drastically increase the electrical conductivity to a level that can support used for electrostatic discharge protection (Kumar et al., 2019).
Table 4 Tensile properties of antistatic
bio-nano composites
Based on the results
of mechanical properties, several test sub-parameters showed different
tendencies compared to the results of the thermal properties test, except for
yield strength and impact strength. This can occur due to uneven distribution
of CNC particles in the PP matrix, excessive MAPP concentration, and
preparation of the tensile test sample. It has been reported that using high
concentrations of MAPP will provide many opportunities for the reinforcement
material to bond to the polymer matrix. However, when MAPP completely covered
the surface of the PP matrix, it did not produce better adhesion and decreased
the mechanical properties of composites (Hassanabadi, Alemdar,
and Rodrigue, 2015). Therefore, it was necessary to optimize the use of MAPP on CNC concentration.
The impact strength of AS-BNC decreased with increasing CNC concentration because the rigid nanoparticles increased the interfacial area between the matrix and fiber and aided the possibility of crack initiation and propagation (Yousefian and Rodrigue 2016; 2015). However, in the AS-BNC-2.5, the impact strength was still higher than Pure PP. This result indicated the presence of antistatic properties of M-DAG in line with resistivity analysis which was in the static dissipative range of 106 – 1012 /sq (Pang et al., 2014).
The yield strength of AS-BNC increased with increasing CNC
concentration. According to Al-Haik et al. (2020), adding CNC to the PP matrix leads to a slight increase in yield
strength (8%). However, the concentration of CNC above 2% indicated yield
strength which tends to decrease to lower than Pure PP due to poor dispersion
of CNC in the PP matrix and the weak chemical bonding between CNC and PP
matrix.
Based on the results of SEM analysis on the cross-section, it was found that M-DAG and CNC agglomerations were found on the PP matrix. The XRD diffractogram pattern of AS-BNC showed peaks representing the diffraction structure of glycerol monostearate and cellulose I. It was observed that the particle size increased with an increase in CNC concentration in AS-BNC. The degree of crystallinity of AS-BNC-2.5 was higher than pure PP. The FTIR spectrum of each treatment was very similar to the characteristic peaks of PP and showed three distinct peaks of AS-BNC compared to pure PP. The melting temperature of AS-BNC-0 was 176.54oC, which was higher than pure PP at 175.44 oC. The thermal stability of AS-BNC-0, AS-BNC-0.5, and AS-BNC-2.5 was 472.07oC, 470.25oC, and 475.15oC, respectively, which was higher than that of pure PP at 468.27oC. Furthermore, the MFI of AS-BNC-0 was lower than pure PP, while AS-BNC-0.5 and AS-BNC-2.5 had a higher MFI value than pure PP, with an increase of 32.0% and 43.9%, respectively. The mechanical properties results showed that the best formulation was AS-BNC-2.5 with 11.071 MPa flexural modulus, 30.376 MPa yield strength, 23.796% yield elongation, 1.659 J/m2 impact strength, which was higher than pure PP, and antistatic at 1010 – 1012 /sq resistivity. Further identification is needed to improve the characteristics of antistatic bio-nano composites. Optimization of the conditions for the synthesis of antistatic bio-nano composites so that it does not easily cause clogging of the twin-screw extruder and injection molding machine and modification of the molding machine is needed to minimize the blockages during the synthesis of antistatic bio-nano composites.
Much appreciation goes to Palm Oil Plantation Fund
Management Agency (POPFMA/BPDPKS) – Ministry of Finance RI for supporting this
research trough Grant Research Sawit (GRS K18) No. PENG-1/DPKS.4/2018.
Adriana,
Jalal, R., Thamrin, Wirjosentono, B., Gea, S., 2014. Mechanical
Properties of Nanocrystal Cellulose Reinforced Polystyrene with Glycerol
Monostearic as Antistatic Agent. International Journal of ChemTech Research,
Volume 6(4), pp. 2421–2426
Al-Haik, M.Y., Aldajah, S.,
Siddique, W., Kabir, M.M., Haik, Y., 2020. Mechanical and Thermal
Characterization of Polypropylene-reinforced Nanocrystalline Cellulose
Nano-composites. Journal of Thermoplastic Composite Materials, Volume
20(10), pp. 1–12
Bhatnagar, N., Asija, N., 2016.
Durability of High-performance Ballistic Composites. In: Lightweight
Ballistic Composites: Military and Law-Enforcement Applications, Bhatnagar,
A. (Eds.), Woodhead Publishing, Inc. Sawston, Cambridge, United Kingdom, pp.
231–283
Brinkmann, A., Chen, M.,
Couillard, M., Jakubek, Z.J., Leng, T., Johnston, L.J., 2016. Correlating
Cellulose Nanocrystal Particle Size and Surface Area. Langmuir, Volume 32(24),
pp. 6105?6114
Chakraborty, B.C., Ratna, D.,
2020. Experimental Techniques and Instruments for Vibration Damping. In: Polymers
for Vibration Damping Applications, Chakraborty, B.C., Ratna, D. (Eds.),
Elsevier Inc., Amsterdam, The Netherlands, pp. 281–325
Clemons, C., Rick, R., 2020.
Preparation of Cellulose Nanocrystal-polyoropylene Masterbatches by
Water-assisted Thermokinetic Mixing. In: ANTEC 2020: The Virtual
Edition, pp. 1–6
Couret, L., Irle, M., Belloncle,
C., Cathala, B., 2017. Extraction and Characterization of Cellulose
Nanocrystals from Post-consumer Wood Fiberboard Waste. Cellulose, Volume
24(5), pp. 1–13
Fang, J., Zhang, L., Sutton, D.,
Wang, X., Lin, T., 2012. Needleless Melt-electrospinning of Polypropylene
Nanofibres. Journal of Nanomaterials, Volume 2012, pp. 1–9
Guerra, V., Wan, C., McNally, T.,
2019. Nucleation of the ?-polymorph in Composites of Poly(Propylene) and
Graphene Nanoplatelets. Journal of Composites Science, Volume 3(38), pp.
1–11
Gwon, J.G., Cho, H.J., Lee, D.,
Choi, D.H., Lee, S., Wu, Q., Lee, S.Y., 2018. Physicochemical and Mechanical
Properties of Polypropylene-cellulose Nanocrystal Nanocomposites: Effects of
Manufacturing Process and Chemical Grafting. BioResources, Volume 13(1),
pp. 1619–1636
Hassanabadi, H.M., Alemdar, A.,
Rodrigue, D., 2015. Polypropylene Reinforced with Nanocrystalline Cellulose:
Coupling Agent Optimization. Journal of Applied Polymer Science, Volume
132(34), p. 42438
Hejna, A., Kirpluks, M., Kosmela,
P., Cabulis, U., Haponiuk, J., Piszczyk, ?., 2017. The Influence of Crude
Glycerol and Castor Oil-based Polyol on the Structure and Performance of Rigid
Polyurethane-polyisocyanurate Foams. Industrial Crops and Products,
Volume 95(1), pp. 113–125
Hobuss, C.B., Da Silva, F.A., Dos
Santos, M.A.Z., De Pereira, C.M.P., Schulz, G.A.S., Bianchini, D., 2020.
Synthesis and Characterization of Monoacylglycerols through Glycerolysis of
Ethyl Esters Derived from Linseed Oil by Green Processes. RSC Advances,
Volume 10, pp. 2327–2336
Jain, R., Tiwari, A., 2015.
Biosynthesis of Planet Friendly Bioplastics using Renewable Carbon Source. Journal
of Environmental Health Science and Engineering, Volume 13(11), pp. 1–5
Kafy, A., Kim, H.C., Zhai, L.,
Kim, J.W., Hai, L. Van, Kang, T.J., Kim, J., 2017. Cellulose Long Fibers
Fabricated from Cellulose Nanofibers and Its Strong and Tough Characteristics. Scientific
Reports, Volume 7, p. 17683
Kumar, P., Narayan Maiti, U.,
Sikdar, A., Kumar Das, T., Kumar, A., Sudarsan, V., 2019. Recent Advances in
Polymer and Polymer Composites for Electromagnetic Interference Shielding:
Review and Future Prospects. Polymer Reviews, Volume 59(4), pp. 687–738
Maryniak,
W.A., Uehara, T., Noras, M.A., 2003. Surface Resistivity and Surface
Resistance Measurements. Trek Application Note, Volume 2003(1005), pp.
195–209
McKeen, L.W., 2014. Introduction
to Plastics, Polymers, and Their Properties. In: The Effect of Temperature
and other Factors on Plastics and Elastomers, McKeen, L.W. (Eds.), William
Andrew, Inc. Norwich, New York, pp. 1–45
Moser, C., Henriksson, G.,
Lindström, M.E., 2016. Specific Surface Area Increase During Cellulose
Nanofiber Manufacturing Related to Energy Input. BioResources, Volume
11(3), pp. 7124–7132
Ng, H.M., Sin, L.T., Bee, S.T.,
Tee, T.T., Rahmat, A.R., 2017. Review of Nanocellulose Polymer Composite
Characteristics and Challenges. Polymer – Plastics Technology and
Engineering, Volume 56(7), pp. 687–731
Pang, H., Xu, L., Yan, D.X., Li,
Z.M., 2014. Conductive Polymer Composites with Segregated Structures. Progress
in Polymer Science, Volume 39(11), pp. 1908–1933
Park, S., Baker, J.O., Himmel,
M.E., Parilla, P.A., Johnson, D.K., 2010. Cellulose Crystallinity Index:
Measurement Techniques and Their Impact on Interpreting Cellulase Performance. Biotechnology
for Biofuels, Volume 3(1), pp. 1–10
Purnomo,
H., Baskoro, H., Muslim, F., 2021. Stress and Strain Behavior of
Confined Lightweight Concrete using Sand Coated Polypropylene Coarse Aggregate.
International Journal of Technology, Volume 12(6), pp. 1261–1272
Rahman,
N.A., Hassan, A., Heidarian, J., 2018. Effect of
Compatibiliser on the Properties of Polypropylene/Glass Fibre/Nanoclay
Composites. Polimeros, Volume 28(2), pp. 1–9
Rochardjo, H.S.B., Fatkhurrohman,
Kusumaatmaja, A., Yudhanto, F., 2021. Fabrication of Nanofiltration Membrane
based on Polyvinyl Alcohol Nanofibers Reinforced with Cellulose Nanocrystal
using Electrospinning Techniques. International Journal of Technology,
Volume 12(2), pp. 329–338
Sabaruddin, F.A., Md-Tahir, P.,
Lee, S.H., 2019. Mechanical Properties of PP/Kenaf Core Nanocomposites Made
from Nanocrystalline Cellulose as An Additive. Journal of Reinforced
Plastics and Composites, Volume 38(2), pp. 88–95
Saleh, T.A., 2021. Structural
Characterization of Hybrid Materials. In: Polymer Hybrid Materials and
Nanocomposites, Saleh, T.A. (Eds.), William Andrew, Inc. Norwich, New York,
pp. 213-240
Salsabila, S., Setyaningsih, D.,
Jannah, Q.R., Muna, N., 2021. Formulation of Mono-diacylglycerol from Palm
Fatty Acid Distillate and Glycerol as Antistatic Agents on Plastics. In:
IOP Conference Series: Earth and Environmental Science, Volume 749, p. 012069
Sandri, G., Bonferoni, M.C.,
Rossi, S., Ferrari, F., Aguzzi, C., Viseras, C., Caramella, C., 2016. Clay
Minerals for Tissue Regeneration, Repair, and Engineering. In: Wound
Healing Biomaterials, Ågren, M.S. (Eds.), Woodhead Publishing, Inc. Sawston,
Cambridge, United Kingdom, pp. 385–402
Sarfat, M.S., Setyaningsih, D.,
Fahma, F., Indrasti, N.S., Sudirman, 2022. Characterization of
Mono-diacylglycerols, Cellulose Nanocrystals, Polypropylene, and Supporting
Materials as raw Materials for Synthesis of Antistatic Bionanocomposites.
In: IOP Conference Series: Earth and Environmental Science, Volume 1034,
pp. 012009
Setyaningsih, D., Suwarna, M.A.,
Muna, N., 2020. The effect of Solvent Type and Temperature on
Mono-diacylglycerol Purification. In: IOP Conference Series: Earth and
Environmental Science, Volume 460, p. 012037
Setyaningsih, D., Warsiki, E.,
Ulfa, S.F., Muna, N., 2020b. The Effect of Sodium Carbonate and Saccharides on
Mono-diacylglycerol (M-DAG) Purification. In: IOP Conference Series: Earth and
Environmental Science, Volume 460, pp. 012038
Shieddieque, A.D., Mardiyati,
Suratman, R., Widyanto, B., 2021. Preparation and Characterization of
Sansevieria Trifasciata Fiber/High-Impact Polypropylene and Sansevieria
Trifasciata Fiber/Vinyl Ester Biocomposites for Automotive Applications. International
Journal of Technology, Volume 12(3), pp. 549–560
Shojaeiarani, J., Bajwa, D.S.,
Chanda, S., 2021. Cellulose Nanocrystal Based Composites: A Review. Composites
Part C: Open Access, Volume 5, p. 100164
Shubhra, Q.T.H., Alam, A.K.M.M.,
Quaiyyum, M.A., 2013. Mechanical Properties of Polypropylene Composites: A
Review. Journal of Thermoplastic Composite Materials, Volume 26, pp.
362–391
Tang, J., Sisler, J.,
Grishkewich, N., Tam, K.C., 2017. Functionalization of Cellulose Nanocrystals
for Advanced Applications. Journal of Colloid and Interface Science,
Volume 494, pp. 397–409
Thomas, B., Raj, M.C., Athira,
B.K., Rubiyah, H.M., Joy, J., Moores, A., Drisko, G.L., Sanchez, C., 2018.
Nanocellulose, a Versatile Green Platform: From Biosources to Materials and
Their Applications. Chemical Reviews, Volume 118(24), pp. 11575–11625
Yousefian, H., Rodrigue, D.,
2015. Nano-crystalline Cellulose, Chemical Blowing Agent, and Mold Temperature
Effect on Morphological, Physical/Mechanical Properties of Polypropylene.
Journal of Applied Polymer Science, Volume 132(47), pp. 428–445
Yousefian, H., Rodrigue, D.,
2016. Effect of Nanocrystalline Cellulose on Morphological, Thermal, and
Mechanical Properties of Nylon 6 Composites. Polymer Composites, Volume
37, pp. 1473–1479
Yusuf, M., Khan, M., Khan, R.A.,
Ahmed, B., 2013. Preparation, Characterization, In Vivo and Biochemical
Evaluation of Brain Targeted Piperine Solid Lipid Nanoparticles in an
Experimentally Induced Alzheimer’s Disease Model. Journal of Drug Targeting,
Volume 21(3), pp. 300–311
Zhou, X., Yu, Y., Lin, Q., Chen,
L., 2013. Effects of Maleic Anhydride-grafted Polypropylene (MAPP) on the
Physico-mechanical Properties and Rheological Behavior of Bamboo
Powder-polypropylene Foamed Composites. BioResources, Volume 8(4), pp.
6263–6279