Research article

Influence of chlorotrifluoroethylene on crystal structure and polymer dynamics of poly(vinylidenefluoride-co-chlorotrifluoroethylene) antibacterial copolymers

  • Received: 25 July 2022 Revised: 01 January 2023 Accepted: 12 January 2023 Published: 18 January 2023
  • The effect of chlorotrifluoroethylene (CTFE) on dynamic relaxations of poly(vinylidenefluoride-co-chlorotrifluoroethylene) films (P(VDF-CTFE)) containing 0, 10, 15 and 20% of CTFE was investigated via broadband dielectric spectroscopy (DRS) and dynamic mechanical analysis (DMA). The interpretation was accompanied by the crystal structure obtained from Fourier transform infrared spectroscopy, wide-angle X-ray diffraction, small-angle X-ray scattering and differential scanning calorimetry. Increment of CTFE contents caused reducing the degree of crystallinity but did not impact the long period, lamellar thickness, and amorphous layer thickness. Four dynamic processes were clearly observed in DRS spectra for the neat poly(vinylidene fluoride) and P(VDF-CTFE) which were attributed to the local motion of amorphous chains (β), the segmental relaxation of amorphous chains (α1), the local conformational rearrangement of the TGTGʹ conformation (α2) and the process associated with Maxwell–Wagner–Sillars interfacial polarization (αMWS). The extra relaxation was observed for P(VDF-CTFE), which was more likely associated to the molecular motion of CTFE chain segments (αc), correspondent with DMA results. These PVDF and P(VDF-CTFE) conducted as self-antibacterial materials.

    Citation: Wisatre Kongcharoensuntorn, Pornpen Atorngitjawat. Influence of chlorotrifluoroethylene on crystal structure and polymer dynamics of poly(vinylidenefluoride-co-chlorotrifluoroethylene) antibacterial copolymers[J]. AIMS Materials Science, 2023, 10(1): 164-181. doi: 10.3934/matersci.2023009

    Related Papers:

    [1] Dorel Feldman . Poly(vinyl alcohol) nanocomposites: Recent contributions to engineering and medicine. AIMS Materials Science, 2021, 8(1): 119-129. doi: 10.3934/matersci.2021008
    [2] Gaurav Gupta, Varun Danke, Tamoor Babur, Mario Beiner . Morphology orientation of comb-like polymers with rigid backbones under the influence of shear fields. AIMS Materials Science, 2017, 4(4): 970-981. doi: 10.3934/matersci.2017.4.970
    [3] Giovanna Di Pasquale, Salvatore Graziani, Chiara Gugliuzzo, Antonino Pollicino . Ionic polymer-metal composites (IPMCs) and ionic polymer-polymer composites (IP2Cs): Effects of electrode on mechanical, thermal and electromechanical behaviour. AIMS Materials Science, 2017, 4(5): 1062-1077. doi: 10.3934/matersci.2017.5.1062
    [4] Seno Jose, Jyotishkumar Parameswaranpillai, Bejoy Francis, Abi Santhosh Aprem, Sabu Thomas . Thermal degradation and crystallization characteristics of multiphase polymer systems with and without compatibilizer. AIMS Materials Science, 2016, 3(3): 1177-1198. doi: 10.3934/matersci.2016.3.1177
    [5] Gérrard Eddy Jai Poinern, Nurshahidah Ali, Xuan Le, Derek Fawcett . Nano-hardness and elastic modulus of anodic aluminium oxide based Poly (2-hydroxyethylmethacrylate) composite membranes. AIMS Materials Science, 2014, 1(3): 159-173. doi: 10.3934/matersci.2014.3.159
    [6] Harikrishnan Pulikkalparambil, Jyotishkumar Parameswaranpillai, Jinu Jacob George, Krittirash Yorseng, Suchart Siengchin . Physical and thermo-mechanical properties of bionano reinforced poly(butylene adipate-co-terephthalate), hemp/CNF/Ag-NPs composites. AIMS Materials Science, 2017, 4(3): 814-831. doi: 10.3934/matersci.2017.3.814
    [7] Chia-Hao Hsieh, Wei-Chi Chen, Sheng-Hsiung Yang, Yu-Chiang Chao, Hsiao-Chin Lee, Chia-Ling Chiang, Ching-Yi Lin . A simple route to linear and hyperbranched polythiophenes containing diketopyrrolopyrrole linking groups with improved conversion efficiency. AIMS Materials Science, 2017, 4(4): 878-893. doi: 10.3934/matersci.2017.4.878
    [8] Wisawat Keaswejjareansuk, Xiang Wang, Richard D. Sisson, Jianyu Liang . Electrospinning process control for fiber-structured poly(Bisphenol A-co-Epichlorohydrin) membrane. AIMS Materials Science, 2020, 7(2): 130-143. doi: 10.3934/matersci.2020.2.130
    [9] Chaitra Srikanth, Gattumane Motappa Madhu, Hemanth Bhamidipati, Siddarth Srinivas . The effect of CdO–ZnO nanoparticles addition on structural, electrical and mechanical properties of PVA films. AIMS Materials Science, 2019, 6(6): 1107-1123. doi: 10.3934/matersci.2019.6.1107
    [10] Laura J. Weiser, Erik E. Santiso . Molecular modeling studies of peptoid polymers. AIMS Materials Science, 2017, 4(5): 1029-1051. doi: 10.3934/matersci.2017.5.1029
  • The effect of chlorotrifluoroethylene (CTFE) on dynamic relaxations of poly(vinylidenefluoride-co-chlorotrifluoroethylene) films (P(VDF-CTFE)) containing 0, 10, 15 and 20% of CTFE was investigated via broadband dielectric spectroscopy (DRS) and dynamic mechanical analysis (DMA). The interpretation was accompanied by the crystal structure obtained from Fourier transform infrared spectroscopy, wide-angle X-ray diffraction, small-angle X-ray scattering and differential scanning calorimetry. Increment of CTFE contents caused reducing the degree of crystallinity but did not impact the long period, lamellar thickness, and amorphous layer thickness. Four dynamic processes were clearly observed in DRS spectra for the neat poly(vinylidene fluoride) and P(VDF-CTFE) which were attributed to the local motion of amorphous chains (β), the segmental relaxation of amorphous chains (α1), the local conformational rearrangement of the TGTGʹ conformation (α2) and the process associated with Maxwell–Wagner–Sillars interfacial polarization (αMWS). The extra relaxation was observed for P(VDF-CTFE), which was more likely associated to the molecular motion of CTFE chain segments (αc), correspondent with DMA results. These PVDF and P(VDF-CTFE) conducted as self-antibacterial materials.



    Poly(vinylidene fluoride) (PVDF) and its copolymers are smart polymers and well known as ferroelectric materials. They are potential for the energy storage capacitors [1,2,3,4]. These polymers have a high dielectric constant, high dielectric breakdown strength and high electrical energy density (>20 J/cm3) [5]. To reduce the cooperative ferroelectricity in the PVDF crystals, various amounts of comonomers such as chlorotrifluoroethylene (CTFE), hexafluoropropylene (HFP) and trifluoroethylene (TrFE) are added at the beginning of the polymerization and randomly incorporated in PVDF [4,5,6,7,8]. It has been reported that both poly(vinylidene fluoride-co-chlorotrifluoroethylene) (P(VDF-CTFE)) and poly(vinylidene fluoride-co-hexafluoropropylene) (P(VDF-HFP)) provide a high energy density of ~25 J/cm3 and break down field > 600 MV/m [5,6], which achieved by reducing a crystalline size of the α-phase [9]. P(VDF-CTFE) shows a dramatic increase of the energy density with the electric field without saturation until reaching the breakdown field, which is different from a neat PVDF [10]. P(VDF-CTFE) random copolymers may have semicrystalline or amorphous structure dependent on CTFE contents. P(VDF-CTFE) containing a small amount of CTFE displays a hexagonal or monoclinic crystalline structure and becomes amorphous when CTFE is higher than 30% [11]. According to the beneficial property of P(VDF-CTFE), the main objective of this work is to investigate the effect of CTFE contents on dynamic relaxations of P(VDF-CTFE) copolymers as a function of temperature and frequency. This investigation would help to understand crystal structure and property relationships of these smart polymers. In addition, other than chemical, physical and ferroelectric properties of fluoropolymers, the antibacterial property has been previously found in polymers containing fluorine such as poly(ethylene terephthalate) coated with 2, 3, 5, 6-tetrafluoro-p-phenylenedimethanol and polytetrafluoroethylene [12,13]. This is due to fluorine atoms in fluoropolymers hindering the bacterial adhesion. However, there are a few investigations of antibacterial activity of commercial fluoropolymers have been reported. Therefore, in this research, a self–antibacterial property of P(VDF-CTFE) films, containing both fluorine and chlorine atoms was preliminarily studied, compared to neat PVDF. This preliminary result would further help to elucidate about the novel antibacterial polymers in nature.

    In this current paper, the crystal structure and thermal property of PVDF and P(VDF-CTFE) were examined using Fourier transform infrared spectroscopy (FTIR), wide-angle X-ray diffraction (WAXD), small-angle X-ray scattering (SAXS) and differential scanning calorimetry (DSC). The dynamic relaxations were investigated using dynamic mechanical analysis (DMA) and broadband dielectric relaxation spectroscopy (DRS). These findings set the stage for understanding the crystal structure, properties, and dynamic behaviors of antibacterial PVDF containing CTFE.

    Solef® 1008 PVDF and Solef® 31008, 31508 and 32008 P(VDF-CTFE) copolymers were obtained and prepared in blown films. The blown films were produced by Brabender single screw extruder with extrusion and die temperature of 250 ℃. P(VDF-CTFE) 31008, 31508 and 32008 blown films containing 10, 15, and 20% of CTFE are referred to as PCTFE10, PCTFE15 and PCTFE20, respectively.

    To verify the effect of CTFE on antibacterial property of PVDF and P(VDF-CTFE) films, the antibacterial activity test was performed. Briefly Escherichia coli (E.coli) and Staphylococcus aureus (S. aureus) bacteria were cultured in Nutrient agar. Then 3–5 colonies were suspended into Mueller Hinton Broth (MHB) and incubated at 37 ℃ for 3 h. Each culture was diluted with phosphate buffer solution to 1.5 × 108 colony forming unit per milliliter (CFU/mL). An amount of 100 µL of diluted suspension was transferred to 3 mL of MHB. Each sterile disc (6 mm in diameter) was transferred to the prepared E. coli and S. aureus suspension. After incubation at 37 ℃ for 24 h, the bacterial suspension was serially diluted with phosphate buffer solution. The bacteria in the diluted suspension were counted by a spread plate method. All tests were analyzed in triplicate. The effectiveness of the antibacterial activity (EAA) was calculated according to Eq 1 [14]:

    EAA(%)=N0NsN0×100 (1)

    where N0 and NS refer to the number of CFU/mL for blank (MHB) and the sample respectively.

    Attenuated total reflectance-Fourier transform Infrared (ATR-FTIR) spectra were acquired using a Thermo Scientific Nicolet 6700 spectrometer, signal averaging 200 scans at a resolution of 2 cm−1.

    Wide-angle X-ray diffraction (WAXD) patterns were collected on a Rigaku DMAX/rapid micro diffractometer in transmission mode using a copper point focused source (λ = 0.154 nm) at 50 kV and 40 mA.

    Degree of crystallinity and thermal characteristics of PVDF and P(VDF-CTFE) films were performed with a Seiko differential scanning calorimetry (DSC) in the temperature range from −90 to 200 ℃, at a heating rate of 10 ℃/min.

    Small-angle X-ray Scattering (SAXS) profiles were collected on a Rigaku SAXS instrument with a CuKα radiation source (λ = 0.154 nm) and a two-dimensional multi-wire detector with the sample-to-detector distance of 1.5 m. The diameter of the incident X-ray beam at the sample position is approximately 0.8 mm. The diameter of the sensitive area of the detector is 120 mm. Absolute scattering intensities were calculated by comparing the spectra of samples to that of a precalibrated polyethylene (S-2907) secondary standard. Data were azimuthally averaged to yield a one-dimensional profile of the absolute intensity I(q) versus scattering vector q (q = (4π/λ)sinθ, where 2θ is the scattering angle). Lorentz-corrected SAXS intensity (Iq2) versus q will be plotted to gain the details of crystalline microstructures. The one-dimensional correlation function K(r) versus the correlation distance r will calculated based on Eq 2 [15]:

    K(r)=12π20q2I(q)cos(qr)dq (2)

    The lamellar thickness, amorphous layer thickness and the linear degree of crystallinity in the lamellar stack were determined from the one-dimensional correlation function base on the first intercept of K(r) with the abscissa B and the long period Lp (Lp = B/xcxa, where xc, xa are the crystalline and amorphous volume fraction). The Lp obtained from the first maximum of the one-dimensional correlation function.

    Dynamic mechanical properties were determined in the tension mode using a TA-Q800 dynamic mechanical analysis (DMA), at frequency of 1 Hz and a heating rate of 5 oC/min. The storage (E') and loss moduli (E") were determined as a function of the temperatures range from −70 to 160 ℃. The rectangular films of 17 × 6.5 mm and 0.03–0.09 mm thick were used.

    Broadband dielectric relaxation spectroscopy (DRS) measurements were performed on a Novocontrol GmbH Concept 40 broadband dielectric spectrometer in the frequency domain range from 0.01 Hz–1 MHz in the temperature range from −100 to 160 ℃. Temperature stability was controlled within ±0.2 ℃. Sample films were sandwiched between brass electrodes of 10 mm in diameter.

    The verified antibacterial property of PVDF and P(VDF-CTFE) is provided in Table 1. The antibacterial efficiency of MHB (blank), PVDF and P(VDF-CTFE) copolymers against E. coli and S. aureus was reported as the effectiveness of the antibacterial activity (% EAA) and the average number of survival bacteria colonies (CFU/mL). PVDF and all P(VDF-CTFE) displayed the antibacterial efficiency against both E. coli and S. aureus and the effectiveness increases with increasing CTFE contents. P(VDF-CTFE) samples designated the better effectiveness against S. aureus than E. coli, compared to neat PVDF. PCTFE20 yielded the antibacterial effectiveness up to 73.01 and 70.21% for S. aureus and E. coli, respectively. In addition, the findings presented that the EAA of PVDF is higher than 50% for both S. aureus and E. coli. This indicates that PVDF itself has rather a good antibacterial property. The finding also shows that introducing chlorine atoms on polymer chains helps improving the antimicrobial activity. This might be due to chlorine atoms of CTFE enhancing antifouling performances by precluding the protein accumulating on polymer surfaces. The mechanisms how to kill bacteria is still unclear. However, from our preliminary experiments it is more likely that polymers containing fluorine and chlorine atoms act as anti-biofilms and then reject the adhesion of bacteria [12,13,16]. This preliminary investigation indicated that polymers containing both chlorine and fluorine atoms are more efficiency to prevent the adhesion of bacteria on the surface [17] than that having only fluorine atoms.

    Table 1.  Antibacterial activity of PVDF and P(VDF-CTFE) copolymers.
    Sample
    E. coli

    S. aureus
    Log CFU/mL EAA (%) Log CFU/mL EAA (%)
    MHB 11.01 blank 11.56 blank
    PVDF 5.10 53.68 5.58 51.73
    PCTFE10 3.90 64.58 4.00 65.40
    PCTFE15 3.67 66.67 3.75 67.73
    PCTFE20 3.37 70.21 3.12 73.01

     | Show Table
    DownLoad: CSV

    FTIR spectra was used to confirm the state of crystalline phases of PVDF and P(VDF-CTFE). Figure 1 displays the comparative FTIR spectra of PVDF and P(VDF-CTFE) in the region from 550 to 1500 cm−1. All films exhibit strong absorption bands at ~614 and 763 cm−1, which have been assigned to CF2 bending in the α-phase PVDF. The peaks at ~795 and 854 cm−1 correspond to CH2 in-plane bending. The absorption bands at ~975, 1148, and 1210 cm−1 is attributed respectively to the CH2 out-of-plane bending, CF2 symmetric stretching, and CF2 asymmetric stretching. These bands have been found to be the characteristic of the vibration absorption of the TGTGʹ conformation, anti-parallel orientation of C–F bonds (nonpolar α-phase) [11,18,19,20,21]. P(VDF-CTFE) samples exhibit a small peak at 840 cm−1, which correspond to the vibrational frequencies of an all-trans conformation, two neighboring chains parallel to each other called polar β-phase [22].

    Figure 1.  FTIR spectra for PVDF and P(VDF-CTFE) films in the region from 550–1500 cm−1.

    Figure 2 displays WAXD pattern of PVDF compared to P(VDF-CTFE) films at 2θ region from 10º to 50º. Owing to the previous investigation [23], the prominent peaks were assigned to the peaks of the α-crystalline phase. Neat PVDF shows the prominent peaks of α (100), α (020), α (110), α (120), α (130) and α (200) [19]. Compared to neat PVDF, P(VDF-CTFE) exhibits three prominent peaks of α (100), α (110) and α (200). The crystal plane of (020), (120) and (130) disappeared for P(VDF-CTFE) due to the CTFE molecules disrupt the molecular chain packing of neat PVDF, which confirms by the degree of crystallinity shown in Table 2. Unlike FTIR results, no β-crystalline phase was observed for P(VDF-CTFE) copolymers. This is probably due to a tiny β-crystalline phase was formed.

    Figure 2.  WAXD patterns of PVDF and P(VDF-CTFE) films.
    Table 2.  Degree of crystallinity for PVDF and P(VDF-CTFE) films.
    Sample WAXD DSC
    Xcw (±5%) Xc (±3%)
    PVDF 0.36 0.53
    PCTFE10 0.30 0.27
    PCTFE15 0.27 0.22
    PCTFE20 0.23 0.16

     | Show Table
    DownLoad: CSV

    Diffraction peaks were deconvoluted with PeakFit® software by fitting with a Gaussian curve, using a linear baseline. The degree of crystallinity (Xcw) was obtained from the fraction of the area under the crystalline peaks to the total area of the amorphous halo and crystalline [24]. As results given in Table 2, the Xcw decreases with increasing CTFE contents, which is in good agreement with that observed from DSC technique, illustrating how the bulky CTFE disrupts crystal formation.

    As seen in Figure 3a, the melting temperature (Tm), a small endothermic transition (Tcd) and glass transition temperature (Tg) are commonly observed for all PVDF and P(VDF-CTFE) samples. Upon cooling, the polymers show sharp exotherms associated with the recrystallization temperature at temperatures lower than Tm (Figure 3b). The inset in Figure 3a displays the glass transition temperature of the samples, which about −37, −36, −26 and −20 ℃, respectively. The transition around 50–60 ℃, denoted as Tcd, has been previously assigned to the conformation disorder (Tcd) transition in the PVDF crystals [19], which is clearly observed for P(VDF-CTFE) compared to neat PVDF. Although this peak has been assigned to the Curie transition (Tc), ferroelectric-paraelectric transition (TF-P), or the transformation of polar (β, γ, and δ) phases to nonpolar α-phases [15], it is not the case for our investigation. To prove this case, a stretched PVDF film (ST-PVDF), containing almost completely polar β-crystalline phases and an unstretched PVDF film, containing only nonpolar α-crystalline phases (see inset in Figure 3c) were prepared and tested. As seen in Figure 3c, this transition remained the same enthalpy for both stretched and unstretched PVDF films. Therefore, this peak cannot be the transformation of polar to nonpolar phases and the strong transition was observed for all P(VDF-CTFE) copolymers, which contain tiny β-crystalline phases compared to the neat PVDF. In addition, this transition cannot be a Curie temperature, since Tc was previously reported that it relates to mole fractions of VF2 in PVDF copolymers so Tc of neat PVDF was expected to near 200 ℃ [25]. In addition, Tc of P(VDF-CTFE) should be higher than 50 ℃ and suppose to reduce with increasing CTFE contents. Therefore, from our finding, we proposed that this transition (Tcd) is attributed to the conformation disorder in the nonpolar α-crystalline phases.

    Figure 3.  DSC thermograms of (a) PVDF and P(VDF-CTFE) copolymers (heating), (b) PVDF and P(VDF-CTFE) copolymers (cooling) and (c) unstretched (PVDF) and stretched PVDF (ST-PVDF) films. The inset is FTIR spectra of an unstretched (PVDF) and stretched PVDF (ST-PVDF) films.

    To confirm the degree of crystallinity obtained from WAXD, the degree of crystallinity was estimated from the enthalpy of Tm peaks (in the temperature range of 100−190 ℃) divided by the ideal melting enthalpy of 100% crystallinity PVDF ( = 104.5 J/g) [19,26]. Compared with WAXD result, similar trend was observed; the degree of crystallinity (Xc) decreases as the amount of CTFE increases (see Table 2).

    Figure 4 shows the Lorentz-corrected SAXS intensity (Iq2) vs. q and the one-dimensional correlation function for PVDF and P(VDF-CTFE) samples. As seen in Figure 4, there is no significant difference between P(VDF-CTFE) and PVDF films. The long period (Lp), lamellar thickness (lc = xcLp), amorphous layer thickness (la = xaLp) and the linear degree of crystallinity in the lamellar stack (XL = lc/(lc + la)) were determined, which are about 89, 69, 26 and 0.73 Å, respectively. The findings suggest that the amount of CTFE does not significantly impact on the value of the long period, lamellar thickness, amorphous layer thickness and linear degree of crystallinity of neat PVDF. This verified that for neat PVDF and P(VDF-CTFE) copolymers with different amounts of CTFE 10–20% (PCTFE10, PCTFE15 and PCTFE20), CTFE molecules do not affect the overall shape of copolymers but to an extent alter the internal arrangements of the PVDF chain. As a result, PVDF and P(VDF-CTFE) have very similar SAXS profiles and different WAXD patterns in some 2θ positions.

    Figure 4.  (a) Lorentz-corrected SAXS profiles and (b) one-dimensional correlation function estimated from SAXS patterns for PVDF and P(VDF-CTFE) copolymers.

    The plots of storage (E') and loss moduli (E") vs. temperatures for PVDF and P(VDF-CTFE) samples are shown in Figure 5. At temperatures above the Tg (about –37 ℃), two relaxations were clearly observed for PVDF (α1 and α2) and P(VDF-CTFE) copolymers (α1 and αc).

    Figure 5.  Temperature dependence of (a) storage and (b) loss moduli for PVDF and P(VDF-CTFE) copolymers at 1 Hz.

    To clarify the process above Tg, tan δ (tan δ = E"/E') was plotted as a function of temperatures (Figure 6). Relative to E", three relaxations (α1, αc and α2) can be seen for P(VDF-CTFE) polymers. Previously, the α1 and α2 have been assigned to the segmental motion of amorphous chains and the local conformational rearrangement of the α-crystalline phase, respectively [19]. The αc transition was only observed for P(VDF-CTFE) copolymers, therefore the αc transitions are more likely associated to the motions within regions of CTFE [27]. This transition will be further elucidated along with DRS studies. Interestingly the loss magnitude of the αc was much higher when the CTFE content increased, PCTFE20 > PCTFE15 > PCTFE10. This tendency follows a typical motion of amorphous chain segments, which indicates that amorphous chains are free from crystal restraints. PCTFE20 requires the lowest energy to move because it contains the lowest degree of crystallinity.

    Figure 6.  tan δ as a function of temperatures for PVDF and P(VDF-CTFE) copolymers at 1 Hz.

    The dielectric constant ε' and dielectric loss ε" were collected from DRS measurement. At high temperature with low frequency, the conduction loss created from the motion of the impurity ions obscures some dipole relaxations. The first derivative of ε' (ε"der=π2ε(f)lnf) has been used to elucidate the relaxation of the conduction-free dielectric loss, helpful for identifying a dipole relaxation of materials with high conduction losses. In addition, previously the derivative loss (ε"der) has been shown to partially resolve overlapping relaxations [28]. Figure 7 shows the temperature dependence of ε"der at a frequency of 1 Hz of the neat PVDF and P(VDF-CTFE) copolymers. Four relaxations were clearly observed for PVDF (β, α1, α2 and αMWS) and five relaxations were observed for the P(VDF-CTFE) (β, α1, αc, α2 and αMWS). Both PVDF and P(VDF-CTFE) copolymers exhibit the β and α1 relaxations at the same temperature range below and above Tg, respectively. At temperature above Tg, PVDF exhibits the α2 relaxation and a Maxwell–Wagner–Sillars (MWS) interfacial polarization (αMWS) between the amorphous and crystalline phases, while the P(VDF-CTFE) displays an extra relaxation, referred to as the αc relaxation.

    Figure 7.  Temperature dependence of ε"der at 1 Hz for PVDF and P(VDF-CTFE) copolymers.

    To deep investigation, the dielectric relaxation spectroscopy data processing was performed. The frequency dependence of ε" at various temperatures was plotted and fit with the appropriate empirical Havriliak-Negami equation as shown in Eq 3 [29]:

    ε"=nr=1ImΔεr(1+(iωτHNr)mr)nr+σ0(ωε0)s (3)

    The characteristic relaxation time (τHN) and dielectric relaxation strength (Δε) were obtained by fitting with Eq 3. The frequency maximum was calculated from Eq 4 [30]:

    fmax=12πτHN[sin(mπ2+2n)sin(mnπ2+2n)]1/m (4)

    where ω is the angular frequency. m and n are the shape parameters. σ0 is the dc conductivity (S/cm), ε0 is the permittivity of free space and the parameter s relates to the type of conduction process.

    PVDF and P(VDF-CTFE) copolymers clearly exhibit four strong relaxations, denoted as β, α1, α2, and αMWS as temperatures increase. The β relaxations were observed below Tg. This relaxation appeared within a temperature range from −100 to −50 ℃. Above −50 ℃ this relaxation was ceased and taken over by the α1 relaxation (see Figure 7). The α2 relaxations were observed from −10 to 160 oC and the αMWS begins to occur at ~100 oC.

    The temperature dependence of all relaxations, except α1 process, follows the Arrhenius law, ƒmax = ƒ0exp[–(Ea/RT)] as shown in Figure 8. The α1 process has the temperature dependence following the Vogel-Fulcher-Tammann (VFT) relation, ƒmax = ƒ0exp[–DT0/(T–T0)] [31].

    Figure 8.  Temperature dependence of the relaxation times for the β, α1, α2 and αMWS processes for PVDF and P(VDF-CTFE) copolymers.

    Below Tg, the β process exhibits an Arrhenius behavior with the activation energy range of 43–57 kJ/mol as given in Table 3, which is previously assigned to the local relaxation, arising from rotational motions of amorphous chains. The second relaxation appears above Tg but lower than Tcd, denoted as the α1 process, and followed the nonlinear VFT behavior. Therefore, the α1 relaxation is assigned to the segmental motion of amorphous chains, which agrees well with previous investigations [19]. The α2 relaxation that is slower than the α1 process exhibits Arrhenius temperature dependence which represents the typical local relaxation behavior with the activation energy range of 73–91 kJ/mol (see Table 3). Since this relaxation was observed at the temperature below Tcd and follows Arrhenius equation, it is more likely related to a local non-cooperative relaxation. Furthermore, as given in Table 4 the Δε (α2) is twice higher than the Δε (β) but less than the Δε (α1). According to the evidence from the structure and thermal properties in previous sections and our previous works, this relaxation relates to the local conformational rearrangement of the nonpolar α-crystalline phases of VDF [19]. It has been reported that the conformation disorder could change between TGTGʹ and GTG'T conformation [32,33]. In addition, this relaxation was clearly observed although CTFE is incorporated, which has no significant difference of the activation energy with increasing CTFE contents. According to FTIR results, PVDF completely contains the nonpolar α-crystalline phase, while P(VDF-CTFE) contains slightly polar β-crystalline phases. This confirms that the α2 process involves with only the nonpolar α-crystalline phases.

    Table 3.  Activation energy of the β, α2, and αMWS processes and the strength parameter of the α1 process for PVDF and P(VDF-CTFE) copolymers.
    Sample Ea (β) (kJ/mol) D (α1) Ea (α2) (kJ/mol) Ea (αMWS) (kJ/mol)
    PVDF
    PCTFE10
    PCTFE15
    PCTFE20
    56.8
    43.2
    43.0
    52.9
    3.2
    3.8
    4.4
    6.5
    86.7
    72.7
    84.9
    90.7
    141.8
    101.2
    129.8
    109.7

     | Show Table
    DownLoad: CSV
    Table 4.  Dielectric relaxation strength of the β, α1, and α2 processes at selected temperatures for PVDF and P(VDF-CTFE) copolymers.
    Sample ɛ ) (T = −80 oC) ɛ (α1) (T = 10 oC) ɛ (α2) (T = 40 oC)
    PVDF
    PCTFE10
    PCTFE15
    PCTFE20
    0.5
    0.6
    0.8
    1.3
    1.7
    2.6
    3.6
    6.7
    0.8
    1.6
    1.1
    2.1

     | Show Table
    DownLoad: CSV

    As seen in Figure 7, the αc process was observed for P(VDF-CTFE) copolymers but absent for neat PVDF, consistent to DMA results. Unfortunately, this process is merging with the α1 relaxation, and results to the broadening of the α1 relaxation for PCTFE10 and PCTFE15 but slightly excluded for PCTFE20. However, the αc process of PCTFE20 merges with the α1 relaxation when frequency is higher than 100 Hz. From the author's knowledge, this relaxation is first time observed. However, unfortunately we are unable to elucidate more information of this relaxation because this weak relaxation exhibits at the frequency closed to that of the segmental relaxation. Based on DMA results, this relaxation is clearly observed for all P(VDF-CTFE) samples with higher loss modulus magnitudes when CTFE increases. This behavior is more likely reflected to the segmental relaxation [34]. Therefore, we proposed that the αc process is the molecular motion of CTFE chain segments, which is accompanied by VDF segments.

    The αMWS process shows the Arrhenius temperature dependence with the high activation energy (Ea > 100 kJ/mol). Due to a high crystallinity in PVDF and P(VDF-CTFE) copolymers, the moving of free ions from amorphous phases is blocked by the crystalline phases and causes a strong increase in ε' at low frequency and high temperatures. Therefore, the αMWS process was proposed to a manifestation of MWS interfacial polarization, arising in multiphase systems. The sharp increase of ε" at highest temperatures and low frequency is due to electrode polarization (EP), contributed from the accumulation of mobile ions at an electrode interface [19].

    The strength parameter (D) was obtained from the VFT behavior of the α1 relaxation, which represents the topology of potential energy governing the segmental reconfigurations of polymers, reflecting the nature of bond interactions and the configurational entropy available to the system [19,30]. Typically, D was used to express the fragility of polymers, higher D corresponding to lower fragility and higher the configurational entropy available to the system. As seen in Table 3, the fragility decreases with increasing CTFE contents because of lower degree of crystallinity (see Table 2) and greater entropy of the system. Comparisons of dielectric relaxation strength (Δε) of β, α1, and α2 processes, at selected temperatures, are provided in Table 4. The Δε of β, α1, and α2 relaxations increases with rising CTFE content. An increment of Δε is mainly influenced by an increase of the dipole moment of CTFE structure.

    Neat PVDF had completely nonpolar α-crystalline phases, while P(VDF-CTFE) copolymers contained slightly polar β-crystalline phases. Three transition temperatures were observed for all samples, which are the glass transition temperature (Tg), the melting temperature (Tm), and the conformation disorder in nonpolar α-crystalline phases (Tcd). There were no significant changes on the long period, lamellar thickness, amorphous layer thickness and linear degree of crystallinity when CTFE was copolymerized with VDF up to 20%. The degrees of crystallinity decreased with increasing CTFE contents. At temperatures above Tg, three relaxations (α1, α2 and αc) were clearly observed by DMA for P(VDF-CTFE) copolymers. Four relaxation processes were clearly observed in DRS spectra for the neat PVDF which are attributed to the local motion of amorphous chains (β), the segmental relaxation of amorphous chains (α1), the local conformational rearrangement of the TGTGʹ conformation (α2) and the process arising from Maxwell–Wagner–Sillars interfacial polarization (αMWS). Other than four relaxations, the extra relaxation was observed for P(VDF-CTFE) copolymers, which was more likely associated to the molecular motion of CTFE chain segments (αc). P(VDF-CTFE) copolymers are effective on inhibiting viable growth of both S. aureus and E. coli up to 73.01 and 70.21%, respectively. The antibacterial effectiveness increased with increasing CTFE contents for both S. aureus and E. coli. Chlorine atoms existing in fluoropolymers enhanced an antibacterial ability.

    The author would like to special thank Prof. James Runt, The Penn State University, USA for all supports. Authors thanks the Faculty of Science, Burapha University, Thailand.

    The authors declare no conflict of interest.



    [1] Bruno A (2009) From vinylidene fluoride (VDF) to the applications of VDF-containing polymers and copolymers: Recent developments and future trends. Chem Rev 109: 6632–6686. https://doi.org/10.1021/cr800187m doi: 10.1021/cr800187m
    [2] Chu BJ, Zhou X, Ren KL, et al. (2006) A dielectric polymer with high electric energy density and fast discharge speed. Science 313: 334–336. https://doi.org/10.1126/science.1127798 doi: 10.1126/science.1127798
    [3] Tjong SC (2018) Polyvinylidene fluoride: a versatile polymer for biomedical, electronic, energy and environmental applications. EXPRESS Polym Lett 12: 395–395. https://doi.org/10.3144/expresspolymlett.2018.33 doi: 10.3144/expresspolymlett.2018.33
    [4] Hamdi O, Mighri F, Denis R (2018) Piezoelectric cellular polymer films: Fabrication, properties and applications. AIMS Mater Sci 5: 845–869. https://doi.org/10.3934/MATERSCI.2018.5.845 doi: 10.3934/MATERSCI.2018.5.845
    [5] Wang Y, Zhou X, Chen Q, et al. (2010) Recent development of high energy density polymers for dielectric capacitors. IEEE T Dielect El In 17: 1036–1042. https://doi.org/10.1109/TDEI.2010.5539672 doi: 10.1109/TDEI.2010.5539672
    [6] Zhou X, Chu B, Neese B, et al. (2007) Electrical energy density and discharge characteristics of a poly(vinylidene fluoride chlorotrifluoroethylene) copolymer. IEEE T Dielect El In 14: 1133–1138. https://doi.org/10.1109/TDEI.2007.4339472 doi: 10.1109/TDEI.2007.4339472
    [7] Lee JH, Lee B, Won JW, et al. (2017) Synthesis of novel telechelic fluoropolyols based on vinylidene fluoride/hexafluoropropylene copolymers by iodine transfer polymerization. Macromol Res 25: 1028–1034. https://doi.org/10.1007/s13233-017-5137-2 doi: 10.1007/s13233-017-5137-2
    [8] Ren X, Meng N, Yan H, et al. (2019) Remarkably enhanced polarisability and breakdown strength in PVDF-based interactive polymer blends for advanced energy storage applications. Polymer 168: 246–254. https://doi.org/10.1016/j.polymer.2019.02.054 doi: 10.1016/j.polymer.2019.02.054
    [9] Guan F, Pan J, Wang J, et al. (2010) Crystal orientation effect on electric energy storage in poly(vinylidene fluoride-co-hexafluoropropylene) copolymers. Macromolecules 43: 384–392. https://doi.org/10.1021/ma901921h doi: 10.1021/ma901921h
    [10] Ranjan V, Yu L, Nardelli MB, et al. (2007) Phase equilibria in high energy density PVDF-based polymers. Phys Rev Lett 99: 047801. https://doi.org/10.1103/PhysRevLett.99.047801 doi: 10.1103/PhysRevLett.99.047801
    [11] Kalfoglou NK, Williams HL (1973) Mechanical relaxations of poly(vinylidene fluoride) and some of its copolymers. J Appl Polym Sci 17: 3367–3373. https://doi.org/10.1002/app.1973.070171111 doi: 10.1002/app.1973.070171111
    [12] Bao Q, Nishimura N, Kamata H, et al. (2017) Antibacterial and anti-biofilm efficacy of fluoropolymer coating by a 2, 3, 5, 6-tetrafluoro-p-phenylenedimethanol structure. Colloid Surface B 151: 363–371. https://doi.org/10.1016/j.colsurfb.2016.12.020 doi: 10.1016/j.colsurfb.2016.12.020
    [13] Gyo M, Nikaido T, Okada K, et al. (2008) Surface response of fluorine polymer-incorporated resin composites to cariogenic biofilm adherence. Appl Environ Microbiol 74: 1428–1435. https://doi.org//10.1128/AEM.02039-07
    [14] Sedlarik V, Galya T, Sedlarikova J, et al. (2010) The effect of preparation temperature on the mechanical and antibacterial properties of poly(vinyl alcohol)/silver nitrate films. Polym Degrad Stab 95: 399–404. https://doi.org/10.1016/j.polymdegradstab.2009.11.017 doi: 10.1016/j.polymdegradstab.2009.11.017
    [15] Linares A, Nogales A, Sanz A, et al. (2010) Restricted dynamics in oriented semicrystalline polymers: poly(vinylidene fluoride). Phys Rev E 82: 031802. https://doi.org/10.1103/PhysRevE.82.031802 doi: 10.1103/PhysRevE.82.031802
    [16] Xia WW, Xia MM, Feng X, et al (2018) Surface modification of poly(vinylidene fluoride) ultrafiltration membranes with chitosan for anti-fouling and antibacterial performance. Macromol Res 26: 1225–1232. https://doi.org/10.1007/s13233-019-7019-2 doi: 10.1007/s13233-019-7019-2
    [17] Popelka A, Novak I, Lehocky M, et al. (2015) Antibacterial treatment of LDPE with halogen derivatives via cold plasma. Express Polym Lett 9: 402–411. https://doi.org/10.3144/expresspolymlett.2015.39 doi: 10.3144/expresspolymlett.2015.39
    [18] Nakhmanson SM, Korlacki R, Johnston JT, et al. (2010) Vibrational properties of ferroelectric β-vinylidene fluoride polymers and oligomers. Phys Rev B 81: 174120. https://doi.org/10.1103/PhysRevB.81.174120 doi: 10.1103/PhysRevB.81.174120
    [19] Atorngitjawat P (2017) Effects of processing conditions and crystallization on dynamic relaxations in semicrystalline poly(vinylidene fluoride) films. Macromol Res 25: 391–399. https://doi.org/10.1007/s13233-017-5060-6 doi: 10.1007/s13233-017-5060-6
    [20] Bargainab F, Thuauc D, Panined P, et al. (2019) Thermal behavior of poly(VDF-ter-TrFE-ter-CTFE) copolymers: Influence of CTFE termonomer on the crystal-crystal transitions. Polymer 161: 64–77. https://doi.org/10.1016/j.polymer.2018.11.064 doi: 10.1016/j.polymer.2018.11.064
    [21] Barrau S, Ferri A, Costa AD (2018) Nanoscale investigations of α- and γ crystal phases in PVDF-based nanocomposites. ACS Appl Mater Interfaces 10: 13092−13099. https://doi.org/10.1021/acsami.8b02172 doi: 10.1021/acsami.8b02172
    [22] Buonomenna MG, Macchi P, Davoli M, et al. (2007) Poly(vinylidene fluoride) membranes by phase inversion: the role the casting and coagulation conditions play in their morphology, crystalline structure and properties. Eur Polym J 43: 1557–1572. https://doi.org/10.1016/j.eurpolymj.2006.12.033 doi: 10.1016/j.eurpolymj.2006.12.033
    [23] Naegele D, Yoon DY, Broadhurst MG (1978) Formation of a new crystal form (αp) of poly(viny1idene fluoride) under electric field. Macromolecules 11: 1297–1298. https://doi.org/10.1021/ma60066a051 doi: 10.1021/ma60066a051
    [24] Dai R, Huang M, Ma L, et al. (2020) Study on crystal structure and phase transitions of polyamide 12 via wide-angle X-ray diffraction with variable temperature. Adv Compos Hybrid Ma 3: 522–529. https://doi.org/10.1007/s42114-020-00192-y doi: 10.1007/s42114-020-00192-y
    [25] Thomas DG (1988) Structure, morphology and models of polymer ferroelectrics, In: Wang TT, Herbert JM, Glass AM, The Applications of Ferroelectric Polymers, Glasgow: Blackie, 53. https://doi.org/10.1002/actp.1989.010400310
    [26] Sharma M, Quamara JK, Gaur A (2018) Behaviour of multiphase PVDF in (1–x)PVDF/(x)BaTiO3 nanocomposite films: structural, optical, dielectric and ferroelectric properties. J Mater Sci Mater Electron 29: 10875–10884. https://doi.org/10.1007/s10854-018-9163-4 doi: 10.1007/s10854-018-9163-4
    [27] Li H, Tan K, Hao Z, et al. (2011) Thermal characterization of a series of poly(vinylidenefluoride-chlorotrifluoroethylene-trifluoroethylene) terpolymer films. J Therm Anal Calorim 105: 357–364. https://doi.org/10.1007/s10973-011-1427-7 doi: 10.1007/s10973-011-1427-7
    [28] Atorngitjawat P, Pipatpanyanugoon K, Aree T (2014) Structure and dielectric relaxations of antibacterial sulfonated polystyrene and silver nanocomposites. Polym Adv Technol 25: 1027–1033. https://doi.org/10.1002/pat.3347 doi: 10.1002/pat.3347
    [29] Wubbenhorst M, van Turnhout J (2002) Analysis of complex dielectric spectra I: one-dimensional derivative techniques and three-dimensional modelling. J Non-Cryst Solids 305: 40–49. https://doi.org/10.1016/S0022-3093(02)01086-4 doi: 10.1016/S0022-3093(02)01086-4
    [30] Masser KA, Runt J (2010) Dynamics of polymer blends of a strongly inter-associating homopolymer with poly(vinyl methyl ether) and poly(2-vinyl pyridine). Macromolecules 43: 6414–6421. https://doi.org/10.1021/ma1011396 doi: 10.1021/ma1011396
    [31] Frenzel F, Borchert P, Anton AM, et al. (2019) Charge transport and glassy dynamics in polymeric ionic liquids as reflected by their inter- and intramolecular interactions. Soft Matter 15: 1605–1618. https://doi.org/10.1039/c8sm02135j doi: 10.1039/c8sm02135j
    [32] Yasuhiro T, Hiroyuki T (1980) Formation mechanism of Kink bands in modification Ⅱ of poly(vinylidene fluoride): evidence for flip-flop motion between TGTG¯ and TG¯TG conformations. Macromolecules 13: 1316–1317. https://doi.org/10.1021/ma60077a056 doi: 10.1021/ma60077a056
    [33] Morton RE, Balik CM (1990) Modeling of the epitaxial crystallization of poly(vinylidene fluoride): T2, TGTGʹ and T3GT3Gʹ chain conformations on (111) calcium fluoride. Macromolecules 23: 680–682. https://doi.org/10.1021/ma00204a051 doi: 10.1021/ma00204a051
    [34] Murayama T (1978) Dynamic Mechanical Analysis of Polymeric material, New York: Elsevier Scientific Publishing Company, 60–65.
  • Reader Comments
  • © 2023 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(1558) PDF downloads(56) Cited by(0)

Figures and Tables

Figures(8)  /  Tables(4)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog