1
1
1
Feba Anna John, Paulose Thomas, George Jacob & Ajith James Jose
To cite this article: Feba Anna John, Paulose Thomas, George Jacob & Ajith James Jose
(2024) Effective interactions and dielectric study of polymer nanocomposite thin films
of polyvinylidene fluoride and graphene oxide, Composite Interfaces, 31:1, 77-97, DOI:
10.1080/09276440.2023.2232979
CONTACT Ajith James Jose [email protected] Post Graduate and Research Department of Chemistry,
Changanassery, Kerala 686101, India
Supplemental data for this article can be accessed online at https://doi.org/10.1080/09276440.2023.2232979.
© 2023 Informa UK Limited, trading as Taylor & Francis Group
78 F. A. JOHN ET AL.
1. Introduction
Polymer-nanocomposites conquered various domains, viz. electronics, optoelectronics,
medicine, communication, food, textiles, etc., with their extraordinary features. The
universe has pieces of evidence to substantiate how the blend of polymer science and
nanoscience creates a significant breakthrough in research.
Polymers with symmetrical structures and true covalent bonds always function as
insulators, whereas polymers with polarized covalent bonds exhibit high permittivity,
thereby a high dielectric constant [1]. Polyvinylidene fluoride (PVDF) is a versatile semi-
crystalline, thermoplastic fluoropolymer with an array of properties such as high chemi
cal resistance, mechanical strength, piezoelectric and pyroelectric properties, good pro
cessability, adhesive nature, environment friendly, and so on [2]. The physical and
electrical properties of PVDF depend on several factors, viz., molecular weights, chain
configurations, crystalline form, and defects in the chain.
The semi-crystalline PVDF polymer may exist in five different phases, namely α, β, γ,
δ, and ε of which α is the most stable polymorphic phase [3–5]. These five polymorphs
may exist in different chain conformations as follows: all-trans (TTTT) planar zigzag for
β phase, Trans-Gauche-Trans-Gauche’ (TGTG`) for α and ε phases, Trans-Trans-Trans-
Gauche’ (TTTG) for γ and δ phases [6–10]. Polymorphism is crucial for polymers to
govern their dielectric constant, ferroelectric properties, polarity, and membrane perfor
mance (fouling and wetting) [11,12]. Among the five polymorphs of PVDF crystals, two
are non-polar (α and ε), and the other three are polar (β, γ, and δ). The β polymorph of
PVDF is the most desirable phase with the highest mechanical property [12].
The fate of dielectric properties of the partially fluorinated PVDF polymer lies mainly
on the carbon backbone chain (C–C bonds) and the additional bonds with fluorine
(C-F and CF2 bonds) and hydrogen (C-H). The β polymorph of PVDF polymer is more
significant for its considerable dipole moment owing to its all-trans conformation [13].
Previous studies suggest that the dielectric properties of PVDF-based nanocomposites
can be intensified by combine them with carbonaceous fillers [13,14].
Graphene oxide (GO) is a derivative of graphene where atoms are arranged two-
dimensionally in a hexagonal lattice. GO is obtained by the process of oxidation in
presence of a suitable oxidizing agent. The functional group -OH is located on the sp2
carbon basal plane. In addition, it is easily dispersible in some other organic solvents and
matrices. The pliability of GO suits it for many fields: biomedicine, electronics, energy
storage materials, composite materials, nanotechnology, etc. [15–18]. The enhancement
of dielectric permittivity of PVA polymer with the inclusion GO polymer was reported
[19]. Mensah e al. [20 explained how the dielectric properties of the blend Acrylonitrile-
Butadiene Rubber/Ethylene-Propylene-Diene-Monome reinforced with GO and r-GO.
GO captivated researchers in various fields and explored a plenty of applications
revealing the exceptional properties emerged after the combination of these two. The
improvisation in the tensile strength, Young’s modulus, and thermal stability of PVDF
with GO nanofiller was reported in previous papers [10,14,21]. The effective nucleation
of β phase and the improved dielectric properties of PVDF/GO nanocomposites are
reported in different studies [10]. Xu et al. [22] synthesized PVDF/r-GO polymer
nanocomposite through solution compounding followed by melt compounding method
and explained the dielectric properties of the composite. Recently, Hang Song et al. [23]
COMPOSITE INTERFACES 79
synthesised PVDF/GO nanocomposite using electro spun method and examined its
piezoelectric properties. The development of PVDF-based innovative energy storage
materials has been growing rapidly due to its exceptional properties [24,25].
The major works which deals with dielectric properties or charge-carrying applica
tions of PVDF/GO polymer nanocomposite-adopted conventional melt processing
method assisted by solid state shear milling technology [26], electrospun method [23],
spin coating method [27], high shear solution mixing followed by compression moulding
method [28], melt compounding method [22], phase inversion method [29] and many
more. However, the simple solution casting synthesis route to develop PVDF/GO-based
dielectric material is not much explored [10,23,27,30,31]. Therefore, this work intended
to develop the PVDF/GO nanocomposite thin films by solution casting method. In
addition, the electrical and dielectric properties of PVDF/GO nanocomposites are
described according to the interactions between the polar phases of the matrix and the
nanofiller. Field Emission Scanning Electron Microscopic analysis (FESEM),
Thermogravimetric analysis (TGA), X-Ray diffraction analysis (XRD), Confocal
Raman Microscopy (CRM), Attenuated Total Reflectance-Fourier Transform Infrared
spectroscopy (ATR-FTIR), UV–Visible absorption Spectroscopy, and LCR meter are the
tools used to investigate the morphology, thermal stability, chemical environment,
conductivity, and dielectric properties of the PVDF/GO polymer nanocomposite. The
outcome of this study expectantly leads to the development of low-cost, lightweight, and
flexible dielectric material to compete with the materials in the present scenario.
PG2, PG3, PG4, and PG5, respectively, according to the weight percentages 0.0, 0.05, 0.1,
0.5, and 1 of GO. Figure 1
Figure 1. Schematic diagram representing the fabrication of PVDF/GO polymer nanocomposite thin
films through solution casting method.
COMPOSITE INTERFACES 81
GO [33]. Figure 2(b) displayed the porous structure of pure PVDF film [34]. Figure 2
(c-f) shows that the pores in the pure PVDF decreased with the increase in the
concentration of GO [35]. The oxygen-containing groups of GO might be interacted
with the fluorine atoms of PVDF and encapsulated into the porous polymer matrix
[10,36].
Figure 3 represents the comparative thermogravimetric study of PVDF/GO nano
composite of different nanofiller concentrations. The maximum weight loss that
occurred within the temperature range of 410°C−500°C [37]. The pure PVDF started
to degrade thermally from 365°C owing to the chain scission of the polymer chain, but
the GO fabricated polymer retained almost 100% by weight at this temperature point.
This implies the downtrend in the degradation rate via the dispersion of GO nanofillers
into the PVDF matrix, and it is supported by the barrier effects associated with the
carbonaceous nanofillers. Substantial degradation of the reinforced PVDF with GO was
detected only after 426°C declared the acquired thermal stability of PVDF with the
nanofiller [31].
Furthermore, a significant depletion in the % weight loss was noticed with the addition
of the filler (GO nanoparticles) in the first-degree thermal degradation itself (Figure 4). It
has also been observed that the final residue of pure PVDF after the thermal degradation
is 17.84 wt.%, and the aggregation is increased by the factor 8.4 and 10.24 for 0.05 wt.%
and 0.1 wt.% of GO, respectively. This indicates that the interaction between the polymer
Figure 2. FESEM images of GO (a), PG1 (b), PG2 (c), PG3 (d), PG4 (e), PG5 (f).
82 F. A. JOHN ET AL.
Figure 3. TGA plot of wt.% Vs T ºC of Pure PVDF (PG1) and PVDF/GO (PG2, PG3, PG4, PG5).
Figure 4. TGA plot weight loss % Vs TºC of Pure PVDF (PG1) and PVDF/GO (PG2, PG3, PG4, PG5).
COMPOSITE INTERFACES 83
matrix and the nanofiller is probably between the oxygen-containing groups of GO and
the fluorine atoms of PVDF (Figure 11) [38]. It is noticed that after 0.1 wt.% addition, the
trend is detoured, and the aggregation is increased by a factor of 6.93 for the next two
additions of GO (0.5 and 1 wt.%) to the matrix. This might be due to the nanofiller
aggregation and bead formation [14].
Figure 5 shows the X-ray diffractograms of GO, pure PVDF, and the PVDF/GO
nanocomposites. A typical diffractogram corresponding to GO with intense peak at
10.7º is portrayed in the inset of Figure 5 [39]. PG1 gives an intense diffraction peak at
20.7º and a weak peak at 41.2º implying the polar crystalline β-phase of the polymer
[1,29,40]. As the wt.% of GO increased the peaks corresponding to β-phase get enhanced.
The diffractogram for PG4 and PG5 displayed a new peak at 26.6º corresponding to the
α- phase of PVDF [29,40]. The absence of characteristic peaks at ~2θ = 10º corresponds
to graphene oxide in the diffractograms of polymer nanocomposites, pointing to the
interaction between GO and PVDF [41].
Figure 6 illustrated the FTIR spectrum of the nanofiller GO. The peak at 1225 cm−1,
and 1014 cm−1 represented C-O-C stretching vibrations and C-O vibrations, respectively,
[42–44]. The C=O stretching vibrations belongs to carboxyl and carbonyl functional
groups are denoted by the peak at 1718 cm−1. The peaks at 1589 cm−1 and 1385 cm−1
represented the C=C stretching vibrations and the hydroxyl group deformation for GO,
respectively, [22].
Figure 7 represents the ATR-FTIR analysis of the polymer nanocomposite (PVDF/
GO) and the polymer matrix within the wavelength range of 950–1500 cm−1. The bands
at 839 cm−1, 1072 cm−1, and 1404 cm−1 are related to β phase (TTTT), and the bands at
1170 cm−1 and 1234 cm−1 are related to γ phase (TTTG). The significant narrowing at
839 cm−1 and 1234 cm−1, the exclusive bands of β (TTTT), and γ (TTTG) phases,
respectively, by the increase in the concentration of GO nanofiller remarks improvement
in the fraction of polar phases of PVDF [26,29,45–49]. Figure 8
Figure 5. X-ray diffractogram of Pure PVDF (PG1) and PVDF/GO (PG2, PG3, PG4, PG5).
84 F. A. JOHN ET AL.
Figure 7. ATR-FTIR transmittance spectra of Pure PVDF (PG1) and PVDF/GO (PG2, PG3, PG4, PG5).
The Raman spectral analysis is backing up the findings of FTIR and XRD analyses.
The typical graphitic G-band and the distorted D-band can be obtained from Figure 9
[50]. The G-band at 1594 cm−1 and the D-band at 1349 cm−1 represent the in-plane
vibrations and the out-of-plane vibrations of the carbonaceous bonds accordingly. Here,
COMPOSITE INTERFACES 85
Figure 8. Structures of (a) non-polar α-phase (TGTG), (b) polar β-phase (TTTT) and (c) polar γ-phase
(TTTG) of PVDF.
the sp2 hybridized carbons swaps to sp3 hybridized carbons, and lattice distortion
occurred [51]. The distortion happened due to the inversion of foreign adatoms, such
as oxygen, hydroxyl groups, and carboxyl groups by the oxidation of some carbons of the
sp2 network, evidencing the GO [52]. Besides the G-band and D-band, it is important to
consider the 2D bands of graphene-based materials. Here, the Raman spectrum of GO
exhibits two 2D peaks, one at 2682 cm−1 represents monolayer and the other at 2921
cm−1 represents multilayer [53].
Figure 10 depicts the Raman spectra corresponding to PVDF/GO nanocompo
sites. The pure PVDF polymer matrix displayed mainly three Raman peaks at 840
cm−1, 1430 cm−1, and 2975 cm−1. The peak observed at 840 cm−1 is the out-of-
phase combination of CH2 rocking and CF2 stretching vibrations designated β -
PVDF polymorphism [54]. The scissoring and wagging modes of vibrations of –
CH2 have been indicated by the edge at 1430 cm−1 representing the polar β, and γ
phases [55]. Furthermore, the symmetric vibrational mode of sp 3 hybridized
H-C-H bonds in PVDF is defined by the peak at 2975 cm−1 assigning the β
phase [49,55]. PG2, PG3, PG4, and PG5 manifest the introduction of GO to the
PVDF matrix by displaying the shift of the G-band from 1594 to 1598 and
D-band from 1349 to 1341 [49,56]
The ID/IG ratio values corresponds to PVDF/GO composites were decreased with the
increase in the concentration of GO owing to the effective interaction between the
nanofiller and the polymer matrix [26]. It is observed that the intensity of the peak
associated with β crystalline phase at 2975 cm−1 is improved continuously for the first
two additions of GO (PG2 and PG3) and then depleted for PG4 and PG5 this also
implying the interaction between the matrix and the nanofiller. Again, in PG4 and PG5
Figure 10. Raman spectra of Pure PVDF (PG1) and PVDF/GO (PG2, PG3, PG4, PG5).
COMPOSITE INTERFACES 87
there obtained a peak around 794 cm−1, which is considered as the typical peak for α
phase [57], this observation concedes the findings of the XRD results mentioned above.
When consolidating all the findings of the above analyses, it is found that the OH
groups of Graphene oxide interact and develop hydrophilic bonds with the electrophilic
fluorine atoms of PVDF (Figure 11), helping to sustain β polymorphism of the matrix
[25,45]. The enhancement in the β phase up to the optimum concentration of the
nanofiller (0.1 wt. % of GO) is demonstrated by the XRD, FTIR, and Raman investiga
tions. These interactions strain the polar bonds on the matrix and restrict its rotations;
consequently, the polar β bonds remain in all trans-polar phase configurations
[27,46,56]. The polar GO, too, contributed to the polarity of the composites. The
improved polarity of PVDF by the nucleation of GO is advantageous for its dielectric
properties [24,46].
Figure 11. Hydrophilic interaction between the C-F bonds of PVDF and hydrogen supplying bonds of GO.
88 F. A. JOHN ET AL.
Figure 13. Comparison of UV-visible absorption curves of Pure PVDF (PG1) and PVDF/GO (PG2, PG3,
PG4, PG5).
COMPOSITE INTERFACES 89
315 nm indicating n-π* transition assigning the C=O bonds of GO [58]. The absorbance
curves inferthat its intensity rises with the GO nanofiller. This shift in absorbance curves
of PVDF/GO composites indicate the variation in the energy gap [59].
The band gap energy indicates how far the valence band is situated from the
conduction band. Usually, the energy gap of dielectrics lies in 3 eV or more, i.e.,
the valance electrons are restricted to cross the forbidden zone. All the same, the
dielectrics loaded with some nanofillers exhibit desirable conducting proper
ties [60].
The band gap energy of the solution cast samples is determined using Tauc
relation expressing the dependence of the absorption coefficient (α) on the photon
energy (hυ).
where, ‘h’ Plank’s constant, ‘υ’ the photon’s frequency, ‘Eg ’ band gap energy, ‘B’ is
a constant called band tailing parameter.
The γ factor depends on the nature of the electron transition and is equal to ½ or 2 for
direct and indirect transition band gaps, respectively.
Figure shows the direct band gap energy determination by plotting½αhυ�2 on the
Y-axis and energy ðeVÞ on the X-axis [61–64].
As per the Tauc plot (Figure 14), the energy gap (Eg ) for pure PVDF is 5.12 eV, and it
is depleted in the presence of nanofiller for all the additions of GO, which indicates
a slight improvement in the conducting properties of the PVDF matrix [46]. It is noticed
that the band gap energy corresponding to the higher concentrations (for PG4 and PG5)
are higher than that of the PG3 (0.1 wt.% GO), which indicates deformations in the
polymer matrix.
Figure 15 depicts the variations in the dielectric constant (ɛr) values of the PVDF
matrix with the inclusion of GO at different concentrations. The dielectric constant was
determined using the following relationship;
where ‘C’ is the Capacitance, ‘d’ is the thickness of the film, ‘ε0’ is the permittivity
of free space with a value of 8.854 × 10−12 Fm−1, and ‘A’ is the area of the
sample [65].
It is observed that the dielectric constant for all the samples is higher in lower
frequencies, and it increases with the frequency and becomes constant for higher
frequencies. This phenomenon can be best explained with the MaxwellWagner Sillar
(MWS) theory stating the accumulation of charge at the interfaces of two materials with
different relaxation times or interfacial polarization (IP). At lower frequencies, the
induced and permanent dipoles get enough time to align in proper orientation and
accumulated the charges; therefore, this region shows high dielectric constant. As the
frequency increased, the diploes could not get enough time to align properly and showed
less dielectric constant than previously; this phenomenon is known as polarization
relaxation. For higher frequency ranges, the dielectric constant becomes frequency
90 F. A. JOHN ET AL.
Figure 14. Tauc – plot of Pure PVDF (PG1) and PVDF/GO (PG2, PG3, PG4, PG5).
independent so the values become almost constant at this range, this process is termed
micro-capacitor structure model [41,65].
The dielectric constant for the PVDF/GO polymer nanocomposites in various composi
tions are higher than that of pure PVDF is due to the increases in the interfacial polariza
tion, which implies the enhancement of all-trans (TTTT) PVDF β polar polymorphism
[46]. The same trend has been observed in the plot between tangent loss (Tan δ) and log
f (Figure 16). If the dielectric constant expresses the electrical polarization, the tangent loss
COMPOSITE INTERFACES 91
Figure 15. Plot of dielectric constant Vs Frequency of Pure PVDF (PG1) and PVDF/GO (PG2, PG3, PG4,
PG5).
portrays the energy dissipation of the materials [66]. The tangent loss has been obtained
from the real (ε”) and imaginary (ε’) parts of the dielectric constant using the equation
00
Tanδ ¼ εε0 . . . . (3) [31,65].
The upgradation of AC conductivity of PVDF polymer matrix with the inclusion of
GO nanofiller is demonstrated in Figure 17. It was calculated using the following
relationship;
where ‘ω’ is the angular frequency and ‘f’ is the angular displacement per unit of
time [65].
The conductivity measurements for each film increased initially with the frequency
and stay almost constant on further increase in frequency. The conductivity of the
polymer nanocomposite is increased with the addition of GO, and the maximum con
ductivity was detected at the PVDF/GO 0.1 wt.% composition. The improved electrical
conductivity of the polymer nanocomposite is the ability of the GO nanofiller to provide
percolative pathways for electron transfer [21,31]. These observations can be related to
the dielectric property demonstrations for the PVDF/GO nanocomposite films.
Abbasipour et al. [14] stated that the improvisation in polarity of PVDF polymer by
instituting GO nanoparticles leads to an appreciable hike in the dielectric properties of PVDF.
92 F. A. JOHN ET AL.
Figure 16. Plot for Tangent loss Vs Frequency of Pure PVDF (PG1) and PVDF/GO (PG2, PG3, PG4, PG5).
Figure 17. Electrical conductivity of Pure PVDF (PG1) and PVDF/GO (PG2, PG3, PG4, PG5).
The hike in the dielectric constant, the tangent loss, and AC conductivity, as well as the
reduction in the band gap energy of the PVDF/GO nanocomposite films confirm the
intercalation of GO into the PVDF matrix due to the interaction between the electrophilic
COMPOSITE INTERFACES 93
fluorine atom and functional groups of GO. This results in the improvement of the fraction of
polar phases and provides electric pathways in the polymer matrix [31,66].
However, it is observed that the band gap energy, dielectric constant, tangent loss, and
AC conductivity belong to 0.5 and 1 wt.% addition of GO is lower than that of 0.1 wt.%
GO. Here, the highest polarity attained by 0.1 wt.% GO is declined by the higher
concentration of GO; consequently, the accumulation of polar bonds failed. Thus, it is
inferred that some of the β phase portions in PG3 converted to the non-polar α phase, as
evidenced by the Raman shift from 840 cm−1 to ~794 cm−1 reported in this work. These
anomalies corresponding to PG4 and PG5 might be due to the bead formation or the
structural defects within the PVDF polymer chain by incorporating GO above the
optimal concentration of nanofiller [10,14,66].
Acknowledgements
This research did not receive any specific grant from funding agencies in the public, commercial,
or not-for-profit sectors. The authors acknowledge DST, SAIF, and MGU, Kottayam, for provid
ing technical assistance for FESEM and Confocal Raman microscopic analysis.
Disclosure statement
No potential conflict of interest was reported by the author(s).
References
[1] Xia W, Zhang Z. PVDF‐based dielectric polymers and their applications in electronic
materials. IET Nanodielectrics. 2018;1(1):17–31. doi: 10.1049/iet-nde.2018.0001
94 F. A. JOHN ET AL.
[2] Halder KK, Tomar M, Sachdev VK, et al. To study the effect of MWCNT incorporated into
PVDF-Graphite composites for EMI shielding applications. Mater Today Proc. 2018;5
(7):15348–15353. doi: 10.1016/j.matpr.2018.05.016
[3] Begum S, Ullah H, Ahmed I, et al. Investigation of morphology, crystallinity, thermal
stability, piezoelectricity and conductivity of PVDF nanocomposites reinforced with
epoxy functionalized MWCNTs. Compos Sci Technol. 2021;211:108841. doi: 10.1016/j.
compscitech.2021.108841
[4] Martins P, Costa CM, Botelho G, et al. Dielectric and magnetic properties of ferrite/poly
(vinylidene fluoride) nanocomposites. Mater Chem Phys. 2012;131(3):698–705. doi: 10.
1016/j.matchemphys.2011.10.037
[5] Bayramol DV, Shah T, Soin N, et al. Preparation and Characterization of PVDF-Based
Nanocomposites Synthesis of Poly(vinylidene fluoride). PVDF). Synthesis Techniques for
Polymer Nanocomposites; 2014. pp. 131–144. doi: 10.1002/9783527670307.ch6
[6] Sun LL, Li B, Zhang ZG, et al. Achieving very high fraction of β-crystal PVDF and PVDF/CNF
composites and their effect on AC conductivity and microstructure through a stretching
process. Eur Polym J. 2010;46(11):2112–2119. doi: 10.1016/j.eurpolymj.2010.09.003
[7] Kabir E, Khatun M, Nasrin L, et al. Pure β-phase formation in polyvinylidene fluoride
(PVDF)-carbon nanotube composites. J Phys D Appl Phys. 2017;50(16):163002. doi: 10.
1088/1361-6463/aa5f85
[8] Choolaei M, Goodarzi V, Khonakdar HA, et al. Influence of graphene oxide on crystal
lization behavior and chain folding surface free energy of poly(vinylidenefluoride-co-
hexafluoropropylene). Macromol Chem Phys. 2017;218(19):1700103–1700112. doi: 10.
1002/macp.201700103
[9] Kumar S, Supriya S, Kar M. Enhancement of dielectric constant in polymer-ceramic
nanocomposite for flexible electronics and energy storage applications. Compos Sci
Technol. 2018;157:48–56. doi: 10.1016/j.compscitech.2018.01.025
[10] Achaby ME, Arrakhiz FZ, Vaudreuil S, et al. Piezoelectric β-polymorph formation and
properties enhancement in graphene oxide – PVDF nanocomposite films. Applied Surface
Science. 2012;258(19):7668–7677. doi: 10.1016/j.apsusc.2012.04.118
[11] Song H Fabrication and characterisation of electrospun polyvinylidene fluoride (PVDF)
nanocomposites for energy harvesting applications. Brunel University London. 2016;0–147.
[12] Lee JE, Eom Y, Shin YE, et al. Effect of interfacial interaction on the conformational
variation of poly(vinylidene fluoride) (PVDF) chains in PVDF/Graphene Oxide (GO)
nanocomposite fibers and corresponding mechanical properties. ACS Appl Mater
Interfaces. 2019;11(14):13665–13675. doi: 10.1021/acsami.8b22586
[13] He ZZ, Yu X, Yang JH, et al. Largely enhanced dielectric properties of poly(vinylidene
fluoride) composites achieved by adding polypyrrole-decorated graphene oxide.
Composites Part A: Applied Science And Manufacturing. 2018;104:89–100. doi: 10.1016/j.
compositesa.2017.10.029
[14] Abbasipour M, Khajavi R, Yousefi AA, et al. The piezoelectric response of electrospun PVDF
nanofibers with graphene oxide, graphene, and halloysite nanofillers: a comparative study.
J Mater Sci Mater Electron. 2017;28(21):15942–15952. doi: 10.1007/s10854-017-7491-4
[15] Kumar N, Yadav N, Amarnath N, et al. Integrative natural medicine inspired graphene
nanovehicle-benzoxazine derivatives as potent therapy for cancer. Mol Cell Biochem.
2019;454(1–2):123–138. doi: 10.1007/s11010-018-3458-x
[16] Perrozzi F, Prezioso S, Ottaviano L. Graphene oxide: From fundamentals to applications.
J Phys Condens Matter. 2015;27(1):13002. doi: 10.1088/0953-8984/27/1/013002
[17] Jagiełło J, Chlanda A, Baran M, et al. Synthesis and characterization of graphene oxide and
reduced graphene oxide composites with inorganic nanoparticles for biomedical
applications. Nanomaterials. 2020;10(9):1–19. doi: 10.3390/nano10091846
[18] Ray SC. 2015. Micro and nano technologies applications of graphene and graphene-oxide
based nanomaterials. Norwich, NY: William Andrew publishing; 2015. Chapter 2,
Application and uses of graphene oxide and reduced graphene oxide: 39–55. doi: 10.1016/
B978-0-323-37521-4.00002-9.
COMPOSITE INTERFACES 95
[37] de Jesus Silva AJ, Contreras MM, Nascimento CR, et al. Kinetics of thermal degradation and
lifetime study of poly (vinylidene fluoride) (PVDF) subjected to bioethanol fuel accelerated
aging. Heliyon. 2020;6(7):e04573. doi: 10.1016/j.heliyon.2020.e04573
[38] Castilla-Cortázar I, Vidaurre A, Marí B, et al. Morphology, crystallinity, and molecular
weight of poly(ε-caprolactone)/graphene oxide hybrids. Polymers. 2019;11(7):1099. doi: 10.
3390/polym11071099
[39] Bansal K, Singh J, Dhaliwal AS. Synthesis and characterization of graphene oxide and its
reduction with different reducing agents. IOP Conf Ser Mater Sci Eng. 2022;1225(1):012050.
doi: 10.1088/1757-899X/1225/1/012050
[40] Nishiyama T, Sumihara T, Sasaki Y, et al. Crystalline structure control of poly (vinylidene
fluoride) films with the antisolvent addition method. Polym J. 2016;48(10):1035–1038.
doi: 10.1038/pj.2016.62
[41] Islam A, Khan AN, Shakir MF, et al. Strengthening of β polymorph in PVDF/FLG and PVDF/
GO nanocomposites. Mater Res Express. 2019;7(1):015017. doi: 10.1088/2053-1591/ab5f82
[42] Khalili D. Graphene oxide: a promising carbocatalyst for the regioselective thiocyanation of
aromatic amines, phenols, anisols and enolizable ketones by hydrogen peroxide/KSCN in
water. New J Chem. 2016;40(3):2547–2553. doi: 10.1039/C5NJ02314A
[43] Manoratne CH, Rosa SRD, XRD-HTA KI, et al. FTIR and SEM interpretation of reduced
graphene oxide synthesized from high purity vein graphite. Mat Sci Res India. 2017;14
(1):19–30. doi: 10.13005/msri/140104
[44] Rattana CS, Witit-Anun N, Witit-Anun N, et al. Preparation and characterization of graphene
oxide nanosheets. Procedia Engineering. 2012;32:759–764. doi: 10.1016/j.proeng.2012.02.009
[45] An N, Liu S, Fang C, et al. Preparation and properties of β -phase graphene oxide/PVDF
composite films. J Appl Polym Sci. 2015;132(10):41577. doi: 10.1002/app.41577
[46] Arshad AN, Wahid MHM, Rusop M, et al. Dielectric and structural properties of poly
(vinylidene fluoride) (PVDF) and poly (vinylidene fluoride-trifluoroethylene)
(PVDF-TrFE) filled with magnesium oxide nanofillers. J Nanomater. 2019;2019:1–12. doi:
10.1155/2019/5961563
[47] Cai X, Lei T, Sun D, et al. A critical analysis of the α, β and γ phases in poly(vinylidene
fluoride) using FTIR. RSC Adv. 2017;7(25):15382–15389. doi: 10.1039/C7RA01267E
[48] Issa A, Al-Maadeed M, Luyt A, et al. Physico-mechanical, dielectric, and piezoelectric
properties of PVDF electrospun mats containing silver nanoparticles. C. 2017;3(4):30.
doi: 10.3390/c3040030
[49] Kaspar P, Sobola D, Částková K, et al. Characterization of polyvinylidene fluoride (PVDF)
electrospun fibers doped by carbon flakes. Polymers. 2020;12(12):2766. doi: 10.3390/
polym12122766
[50] Badi N, Khasim S, Roy AS. Micro-Raman spectroscopy and effective conductivity studies of
graphene nanoplatelets/polyaniline composites. J Mater Sci Mater Electron. 2016;27
(6):6249–6257. doi: 10.1007/s10854-016-4556-8
[51] Vieira O, Ribeiro RS, Pedrosa M, et al. Nitrogen-doped reduced graphene oxide – PVDF
nanocomposite membrane for persulfate activation and degradation of water organic
micropollutants. Chem Eng J. 2020;402:126117. DOI:10.1016/j.cej.2020.126117
[52] Tran DT, Nguyen VN. rGO/Persulfate metal-free catalytic system for the degradation of
tetracycline: effect of reaction parameters. Mater Res Express. 2020;7(7):075501. doi: 10.
1088/2053-1591/ab9e47
[53] TNABTA M, Tan SJ, Foo KL, et al. Properties of polyaniline/graphene oxide (PANI/GO)
composites: effect of GO loading. Polym Bull. 2021;78(9):4835–4847. doi: 10.1007/s00289-
020-03334-w
[54] Peleš A, Aleksić O, Pavlović VP, et al. Structural and electrical properties of ferroelectric
poly(vinylidene fluoride) and mechanically activated ZnO nanoparticle composite films.
Phys Scr. 2018;93(10):105801. doi: 10.1088/1402-4896/aad749
[55] Barnakov YA, Paul O, Joaquim A, et al. Light intensity-induced phase transitions in
graphene oxide doped polyvinylidene fluoride. Opt Mater Express. 2018;8(9):2579.
doi: 10.1364/OME.8.002579
COMPOSITE INTERFACES 97
[56] Elashmawi IS, Gaabour LH. Raman, morphology and electrical behavior of nanocomposites
based on PEO/PVDF with multi-walled carbon nanotubes. Results Phys. 2015;5:105–110.
doi: 10.1016/j.rinp.2015.04.005
[57] Ma J, Zhang Q, Lin K, et al. Piezoelectric and optoelectronic properties of electrospinning
hybrid PVDF and ZnO nanofibers. Mater Res Express. 2018;5(3):035057. doi: 10.1088/2053-
1591/aab747
[58] Cave M, Falkner KC, McClain C. Zakim and Boyers hepatology. US: Elsevier health;2012.
Chapter 27, Occupational and environmental hepatotoxicity; pp. 476–492. doi: 10.1016/
B978-1-4377-0881-3.00027-9
[59] Picollo M, Aceto M, Vitorino T. UV-Vis spectroscopy. Phys Sci Rev. 2019;4(4):20180008.
doi: 10.1515/psr-2018-0008
[60] Kumar P, Narayan Maiti U, Sikdar A, et al. Recent advances in polymer and polymer
composites for electromagnetic interference shielding: review and future prospects. Polym
Rev. 2019;59(4):687–738. Taylor and Francis Inc. Routledge and Dove Medical Press
imprints. doi: 10.1080/15583724.2019.1625058
[61] Aldulaimi NR, Al-Bermany E. Tuning the bandgap and absorption behaviour of the
newly-fabricated ultrahigh molecular weight polyethylene oxide- polyvinyl alcohol/
Graphene oxide hybrid nanocomposites. Polym Compos. 2022;30:096739112211121.
doi: 10.1177/09673911221112196
[62] Dhanaraj A, Das K, Keller JM. A study of the optical band gap energy and Urbach energy of
fullerene (C60) doped PMMA nanocomposites. AIP Conf Proc. 2020;2270:110040.
[63] Makuła P, Pacia M, Macyk W. How to correctly determine the band gap energy of modified
semiconductor photocatalysts based on UV–Vis spectra. J Phys Chem Lett. 2018;9
(23):6814–6817. doi: 10.1021/acs.jpclett.8b02892
[64] Vandana M, Devendrappa H, Padova PD, et al. Polymer nanocomposite graphene quantum
dots for high-efficiency ultraviolet photodetector. Nanomaterials. 2022;12(18):3175. doi: 10.
3390/nano12183175
[65] Ishaq S, Kanwal F, Atiq S, et al. Dielectric properties of graphene/titania/polyvinylidene fluoride
(G/TiO2/PVDF) nanocomposites. Materials. 2020;13(1):205. doi: 10.3390/ma13010205
[66] Acharya S, Gopinath CS, Alegaonkar P, et al. Enhanced microwave absorption property of
Reduced Graphene Oxide (RGO)–Strontium Hexaferrite (SF)/Poly (Vinylidene) Fluoride
(PVDF). Diam Relat Mater. 2018;89:28–34. doi: 10.1016/j.diamond.2018.07.024