You are on page 1of 9

The Journal of Supercritical Fluids 184 (2022) 105555

Contents lists available at ScienceDirect

The Journal of Supercritical Fluids


journal homepage: www.elsevier.com/locate/supflu

Structure disorder observation of fluoropolymers composed of vinylidene


fluoride and tetrafluoroethylene in supercritical CO2 using time-resolved
small- and wide-angle X-ray scattering
Kohei Yamanoi a, b, 1, Satoshi Shibuta c, 1, Atsushi Shiro a, Masao Noumi b,
Melvin John F. Empizo a, Marilou Cadatal-Raduban a, d, Nobuhiko Sarukura a,
Keiko Nishikawa e, f, Takeshi Morita e, *
a
Institute of Laser Engineering, Osaka University, 2-6 Yamadaoka, Suita, Osaka 565-0871, Japan
b
Technology and Innovation Center, Daikin Industries, LTD., 1-1, Nishi-Hitotsuya, Settu, Osaka 566-8585, Japan
c
Department of Physics, College of Humanities and Sciences, Nihon University, 3-25-40, Sakurajosui, Setagaya, Tokyo 156-8550, Japan
d
School of Natural Sciences, Massey University, Albany, Auckland 0632, New Zealand
e
Department of Chemistry, Graduate School of Science, Chiba University, Chiba 263-8522, Japan
f
Toyota Physical & Chemical Research Institute, Nagakute, Aichi 480-1192, Japan

H I G H L I G H T S G R A P H I C A L A B S T R A C T

• Vinylidene fluoride and tetrafluoro­


ethylene copolymers were synthesized.
• SC CO2 was applied as an SC fluid Compression molding
for WAXS
Process of structure disorder
stimulus to synthesized fluoropolymers. copolymer with
VdF:TFE = 5:5
• Time-resolved mesoscopic structural
changes in fluorocopolymers were
investigated. Melt molding copolymer with
• Little change in fluoropolymer crystal­ for time-resolved SAXS VdF:TFE = 8:2

linity was observed after SC CO2


application. 0 400 800 1200
• Higher vinylidene fluoride content SC CO 2 application time / s

enhanced tolerance to SC CO2 stimulus. Higher VdF ratio Tolerance Enhancement

A R T I C L E I N F O A B S T R A C T

Keywords: Fluoropolymers have a growing range of applications in many industries. Neat fluoropolymers and fluorinated
Fluoropolymer copolymers were prepared via the suspension polymerization method with entire mass ratio of vinylidene
Supercritical CO2 fluoride (VdF) and tetrafluoroethylene (TFE). Supercritical CO2 was applied as a fluid stimulus to the fluo­
Small-angle X-ray scattering
ropolymers. An increase in structural disorder was investigated using time-resolved small-angle X-ray scattering
Wide-angle X-ray scattering
(SAXS). The SAXS change induced by the stimulus proceeded more slowly with increasing VdF ratio of the
Vinylidene fluoride
Tetrafluoroethylene copolymer, suggesting that the higher VdF content contributes to the tolerance of microvoid formation and the
corresponding onset of breakdown in the polymer material. The degree of crystallinity of the fluoropolymers was
evaluated using wide-angle X-ray scattering, showing the degree of crystallinity was similar before and after the

Abbreviations: VdF, vinylidene fluoride; TFE, tetrafluoroethylene; PVdF, poly(vinylidene fluoride); PTFE, poly(tetrafluoroethylene); polyVdF–TFE, poly(vinyli­
dene fluoride-co-tetrafluoroethylene); SC, supercritical; CO2, carbon dioxide; SAXS, small-angle X-ray scattering; WAXS, wide-angle X-ray scattering; XRD, X-ray
diffraction; SWAXS, SAXS/WAXS.
* Corresponding author.
E-mail address: moritat@faculty.chiba-u.jp (T. Morita).
1
KY and SS contributed equally

https://doi.org/10.1016/j.supflu.2022.105555
Received 4 January 2022; Received in revised form 14 February 2022; Accepted 16 February 2022
Available online 25 February 2022
0896-8446/© 2022 Elsevier B.V. All rights reserved.
K. Yamanoi et al. The Journal of Supercritical Fluids 184 (2022) 105555

stimulus. Time-resolved and in-situ SAXS/WAXS measurements were carried out on the copolymer with a ratio of
VdF:TFE = 8:2 as the copolymer having the highest VdF content studied here.

1. Introduction process applications, such as alternative chemical extraction media, heat


exchangers, fine particle synthesis, sterilization, and surface cleaning
Fluoropolymers have increasing applications in many industries. The [18–27]. Furthermore, it is well known that the SC CO2 has the higher
fluoropolymers composed of vinylidene fluoride (VdF) and tetrafluoro­ chemical activity than nonpolar SC fluids due to the quadrupole moment
ethylene (TFE) show excellent properties in chemical resistance, me­ of the CO2 molecule. In this study, an SC CO2 treatment was applied as
chanical strength, hydrophobicity, thermal stability, low surface an SC fluid stimulus to fluoropolymers. The fluoropolymers of PVdF,
energies, low coefficients of friction, and low dielectric constants [1–4]. PTFE, and copolymers composed of both VdF and TFE were synthesized
In the viewpoint of chemical bonding, these excellent properties are through the suspension polymerization method with entire range of
predominantly caused by the electronic structure of the fluorine (F) polymerization ratios. The stimulus-induced structural changes of the
atom and the stable carbon–fluorine (C–F) covalent bonding. Further­ copolymers were investigated using time-resolved SAXS measurements.
more, the unique intramolecular and intermolecular interactions be­ The wide-angle X-ray scattering (WAXS) method was utilized to analyze
tween the polymers’ fluorinated segments and the main chains of the changes in the crystallinity of the fluoropolymers induced by the stim­
polymers contribute to the advantageous properties of the fluoropol­ ulus. Furthermore, the time-resolved and in-situ SAXS/WAXS (SWAXS)
ymers composed of VdF and TFE [5]. measurements were carried out on the copolymer with a ratio of VdF:
Considering their structure and composition, fluoropolymers are TFE = 8:2, which was selected as the copolymer having the highest VdF
classified as partially fluorinated polymers or perfluoropolymers. content studied here.
Partially fluorinated polymers contain F and hydrogen (H) atoms in their
chemical structures, whereas perfluoropolymers only contain F atoms. 2. Experimental
For example, poly(vinylidene fluoride) (PVdF) is a highly processable
and mechanically strong partially fluorinated polymer, whereas poly 2.1. Material preparation
(tetrafluoroethylene) (PTFE) is a chemically robust and heat-resistant
perfluoropolymer. A copolymer composed of VdF and TFE is thus ex­ PVdF, PTFE, and the copolymer poly(vinylidene fluoride-co-
pected to exhibit the desirable properties of high chemical resistance, tetrafluoroethylene), abbreviated as poly(VdF–TFE), with different
thermal stability, and mechanical strength, combining the characteris­ mass ratios (w/w) of VdF (–CH2–CF2–) and TFE (–CF2–CF2–) were pro­
tics of both the partially fluorinated polymer PVdF and the per­ vided by Daikin Industries, Ltd. The polymers were synthesized using
fluoropolymer PTFE. the suspension polymerization method; this procedure is similar to that
PVdF and the derived fluoropolymers have been already used as reported in Ref. [28]. The ratios used for the poly(VdF–TFE) synthesis
membrane fibers for applications related to water distillation [6–8]. were set at VdF:TFE = 8:2, 6:4, 5:5, 4:6, and 3:7. Considering the
Among their different applications, the copolymers composed of VdF structure control in the synthesis, the prepared poly(VdF–TFE) was
and TFE are valuable for their uses related to the packing and linings of classified as a random copolymer.
high-pressure vessels. This is because the Young modulus of PVdF is one To obtain a sample that was appropriate for the WAXS measure­
order of magnitude larger than that of PTFE, whereas the PTFE is rela­ ments, the synthesized polymer powder was formed into pellets (see
tively rubbery [9,10]. Thus, the Young’s moduli of these fluoropolymers Fig. 1a) via the compression molding method. The polymer powder was
have been systematically used to evaluate their pressure tolerance under placed in cylinders under pressurizing at 10–60 MPa and heating at
various conditions. Although PVdF and PTFE both appear to endure temperatures from 200◦ to 380◦ C with adjusting the conditions to pre­
high-pressure environments, their breakdown begins at the microscopic vent void formation in the material and to maintain sample uniformity.
level and manifests as distortions, as well as the irreversible stretching or For the SAXS, ultra-SAXS (U-SAXS), and SWAXS measurements, the
compression of the molecular chains in the fluoropolymers. Conse­ synthesized polymer powder was formed into a rod of 3 mm in diameter
quently, the ability to directly observe the microscopic morphology of
the material is essential in evaluating the onset of fluoropolymer dam­
age. The stretching and compression of molecular bonds during the
application of high-pressure carbon gases have been observed in several
polymers using attenuated total reflection infrared (ATR-IR) spectros­
copy [11]. Thurecht et al. utilized high-pressure nuclear magnetic
resonance (NMR) to investigate the effects of swelling on the linear
low-density polyethylene under supercritical (SC) carbon dioxide (CO2)
conditions [12].
The small-angle scattering (SAS) of neutrons and X-rays is one of the
most effective methods to study the structure on the mesoscopic length
scale. Marigo et al. have investigated the lamellar structure that consists
of both amorphous and crystal phases in the perfluorinated copolymers
of tetrafluoroethylene using the small-angle X-ray scattering (SAXS)
[13]. The SAXS method has been used to clarify the stressed or stretched
bulk PVdF along with PVdF fibers and sheets, showing deformation of
the crystalline and amorphous phases in PVdF [14–16]. Tang et al. have
analyzed the deformation of PTFE during stretching and its
radiation-induced morphological transformation [17]. However, the
structural changes on the mesoscopic length scale induced by stimuli Fig. 1. (a) Sample pellet prepared by compression molding for WAXS mea­
surements. (b) Sample rod of 3 mm in diameter prepared by melt molding for
using the SAS method have not been reported for copolymers composed
time-resolved SAXS, U-SAXS, and SWAXS measurements. Disk of 0.6 mm in
of both VdF and TFE.
thickness indicated by red arrow was used for the SWAXS measurements. L-
The SC CO2 is widely applied to various green and sustainable shaped tool is an M3-hexagonal wrench included for scale.

2
K. Yamanoi et al. The Journal of Supercritical Fluids 184 (2022) 105555

(Fig. 1b) via the melt molding method. The polymer powder was melted two-dimensional semiconductor detectors of PILATUS 1 M (Dectris) and
by heat treatment for the molding process and then solidified by cooling. PILATUS 100 K (Dectris) for the SAXS and WAXS measurements,
The polymer was passed through 3-mm-diameter stainless steel die with respectively. The distance from the sample to the detectors for the SAXS
a heat treatment at 300 ◦ C. and WAXS measurements was set to be 1446 and 254 mm, respectively.
For the SAXS measurements, the exposure time and interval were set at
2.2. Combined SAXS and U-SAXS 10 and 120 s, respectively. The exposure time and interval for the WAXS
measurements were set at 60 and 120 s, respectively; these values were
The combined SAXS and U-SAXS measurements were performed used to take into account the difference in scattering intensity between
using the BL19B2 station at SPring-8 of the Japan Synchrotron Radiation the SAXS and WAXS measurements. The prepared rod-shaped sample
Research Institute (JASRI). The scattering signals were acquired using a was cut to form a disk with a thickness of 0.6 mm for the SWAXS
two-dimensional semiconductor detector (PILATUS 2 M, Dectris). The measurement using a fine cutter, as shown in Fig. 1b.
distance from the sample to the detector was set at 3107 and 41,000 mm
for the SAXS and U-SAXS measurements, respectively. Considering the 2.5. Sample holder
long path length of U-SAXS measurements, the X-ray energy was set at a
high value, 18.0 keV, (λ = 0.0689 nm, λ: wavelength). The exposure An SC fluid sample holder was used for the time-resolved SAXS and
time was adjusted to 60 and 300 s for the SAXS and U-SAXS measure­ SWAXS measurements. The sample holder was made of the titanium
ments, respectively. The observable q-region was evaluated to be alloy (Ti–6Al–4V). A pair of single crystal diamond disks of 5.0 mm in
0.0580–2.91 nm− 1 for the SAXS and 0.00440–0.200 nm− 1 for the U- diameter and 0.70 mm in thickness (type Ib, Sumitomo Electric In­
SAXS, where q is the scattering parameter defined by 4πsinθ/λ (2θ: dustries, Ltd.) were used as the X-ray scattering window. The diamond
scattering angle). The absorption correction was examined using the disks were sealed to the cell body by applying stress from the outside to
absorption factor, μl (μ: linear absorption coefficient, l : sample thick­ the inside using a hand-made gold O-ring. The sealing system permits
ness), determined using the direct-beam monitoring apparatus implan­ the high-pressure experiments under keeping the sample thickness
ted in the beamstop. The background intensity was subtracted by constant. A K-type thermocouple (class I, Chino Co.) and a compact
strain gage (PGL-A, Kyowa Electronic Instruments Co., Ltd.) were used
background measurement using the relation of I = Iobserved ∙exp⁡(μl ) −
to measure the temperature and pressure, respectively. The measure­
Ibackground , where Iobserved is the experimentally observed intensity with
ment system is suitable for SAXS experiments at temperatures of up to
the sample polymer and Ibackground is the experimentally measured
350 ◦ C and pressures of up to 40 MPa, including in high-temperature SC
background intensity. For the SAXS measurements in this study, the
conditions.
correction of absorption and background was performed in the same
manner. The SAXS and U-SAXS signals were normalized using the cali­
2.6. WAXS
bration factor determined from the standard glassy carbon
measurements.
The WAXS measurements were performed using a Rigaku SmartLab
X-ray diffractometer. The X-rays of copper (Cu) Kα radiation
2.3. Time-resolved SAXS (λ = 0.1542 nm) was used as the probe under a 40 kV tube voltage and a
40 mA tube current. The scanning conditions were as follows: step size,
The time-resolved SAXS measurements were carried out on the poly 0.02◦ ; step speed, 10◦ /min; scattering angle 2θ range, from 5◦ to 60◦ ;
(VdF–TFE) with ratios of VdF:TFE = 5:5 and 8:2 using the BL19B2 sta­ and scanning resolution, 0.0002◦ . For 4 h, the polymer samples were
tion at SPring-8 of the JASRI. To effectively extract the scattering signal subjected to an SC CO2 environment at a temperature and pressure of
from the supercritical fluid and polymer system, the X-ray energy was 60 ◦ C and 10 MPa, respectively. The degree of crystallinity for each
set to be the value of 18.0 keV, (λ = 0.0689 nm); this value is higher polymer was determined using the profile fitting software provided by
than that used in a typical SAXS experiment. The scattering signals were Smartlab Studio II before and after the SC fluid treatment. The observed
acquired using a two-dimensional semiconductor detector (PILATUS WAXS signals were deconvoluted into broad curves (amorphous phase
2 M, Dectris). The sample-to-detector distance was set to 3107 mm. The contribution) and sharp diffraction patterns (crystal part contribution).
acquisition interval and exposure time were set to 20 and 10 s, respec­ The degree of crystallinity was calculated from the ratio of the inte­
tively. During the measurements, the SC CO2 treatment was kept at a grated values in each region.
constant temperature and pressure of 60 ◦ C and 10 MPa, respectively.
The SC CO2 was loaded from the outside of the beam line hutch using a 3. Results and discussion
high-pressure apparatus specifically designed for this study; this allowed
for the time-resolved signals to be acquired from the initial state (at time 3.1. Mesoscopic structure by SAXS and U-SAXS
zero) after the stimulus began. The prepared rod-shaped polymer sample
was cut to form a disk with a thickness of 1.00 ± 0.02 mm for the time- The fundamental information on the polymer structure on the
resolved SAXS measurements (0.98 and 1.02 mm in thickness for VdF: mesoscopic scale was obtained using the combined SAXS and U-SAXS
TFE = 5:5 and 8:2, respectively) using a fine cutter. The observable q- signals. Fig. 2 shows the SAXS and U-SAXS profiles of the copolymers
region was determined to be 0.0580–2.91 nm− 1. with ratios of VdF:TFE = 3:7, 4:6, 5:5, 6:4, and 8:2. As shown in Fig. 2,
the first peak positions gradually shifted to the higher q-region as the
2.4. SWAXS VdF ratio increased. The peak positions for VdF:TFE = 3:7 and 8:2 were
observed around q = 0.1 nm− 1 (63 nm in real space) and q = 0.3 nm− 1
The time-resolved and in-situ SWAXS signals were acquired using the (21 nm in real space), respectively. It is suggested here that the periodic
SWAXS apparatus set at the BL-6A station of the Photon Factory (PF) of structure of the copolymers on the mesoscopic scale, and possibly its
the High Energy Accelerator Research Organization (KEK), Tsukuba, lamella morphology, becomes more compact and tightened as the VdF
Japan [29]. During the SWAXS measurements, the X-ray absorption ratio becomes higher in the copolymer.
factors of the samples were measured simultaneously using an in-situ
beam monitoring apparatus with a silicon photodiode device [29,30]. 3.2. Mesoscopic structural change by time-resolved SAXS
The wavelength of the X-rays used was set to λ = 0.150 nm. The tem­
perature and pressure were kept constant at 60 ◦ C and 10 MPa, As described previously, PVdF has an advantage in mechanical
respectively. The scattering signals were measured using the strength, compared with PTFE. Furthermore, SC CO2 has the higher

3
K. Yamanoi et al. The Journal of Supercritical Fluids 184 (2022) 105555

1.6
10
10

1.2
5 VdF:TFE
10
VdF:TFE = 5:5
8:2
720 s
0.8 600 s
480 s
360 s
6:4 240 s
0
10 180 s
120 s
5:5 60 s
0.4 40 s
20 s
4:6 0s

10 3:7
10 10 10
0 0
scattering parameter q (nm ) 0 0.2 0.4 0.6
q (nm )
Fig. 2. Combined SAXS and U-SAXS profiles for poly(VdF–TFE) with ratios of
VdF:TFE = 8:2, 6:4, 5:5, 4:6, and 3:7. Fig. 3. Time-resolved SAXS intensities for poly(VdF–TFE) with ratio of VdF:
TFE = 5:5 during SC CO2 stimulus from its initial state (blue circle) to that
observed after a stimulus of 720 s in duration (red circle).
chemical activity compared with nonpolar SC fluids due to the quad­
rupole moment of the CO2 molecule. This study was focused on
analyzing poly(VdF–TFE) with a relatively high VdF proportion (not
neat PVdF) due to the possibility that it would exhibit the desirable
property of resilience to the SC CO2 exposure. The ratio of VdF:TFE VdF:TFE = 8:2
= 8:2 had the highest VdF content of the poly(VdF–TFE) materials 0.12
studied here. As a comparison, poly(VdF–TFE) with the ratio of a VdF:
TFE = 5:5 was also investigated to establish the possible VdF content 720 s
600 s
dependence on the tolerance to the SC CO2 treatment. 480 s
Fig. 3 shows the SAXS change for poly(VdF–TFE) with a ratio of VdF: 360 s
TFE = 5:5 during the SC CO2 stimulus. The SAXS intensities were 240 s
significantly increased by the stimulus. The SAXS increase was observ­ 180 s
0.08 120 s
able only in the relatively smaller-angle region. The peak position 60 s
observed in the SAXS measurements gradually shifted toward the low-q 40 s
region with increase in the application time; this indicates that the 20 s
formation of microvoids, structural disordering of the periodic structure, 0s
and swelling induced by the SC CO2 led to structural expansion on the
mesoscopic scale, as has been previously reported in the literature [10, 0.04
31–33]. The SAXS increase became slower and was almost saturated
around 720 s after the stimulus started, although the saturation time will
depend on experimental conditions, such as sample thickness, pressure,
and temperature.
Fig. 4 shows the SAXS change for the poly(VdF–TFE) with a ratio of
VdF:TFE = 8:2 during the SC CO2 stimulus. The aspects of SAXS change 0
for the poly(VdF–TFE) with a ratio of VdF:TFE = 8:2 were analogous to
those of the 5:5 poly(VdF–TFE); namely the significant increase in SAXS
0 0.2 0.4 0.6
intensities, the low-q shift, and the smaller change in the higher-q re­ q (nm )
gion. However, the SAXS increase at the lower q-region around q
Fig. 4. Time-resolved SAXS intensities for poly(VdF–TFE) with ratio of VdF:
= 0.1 nm− 1 was observed as much smaller than that for the 5:5 poly
TFE = 8:2 during SC CO2 stimulus from its initial state (blue circle) to that
(VdF–TFE). The rate of SAXS increase remained constant at 720 s after observed after a stimulus of 720 s in duration (red circle).
the stimulus started, indicating much slower change in the 8:2 poly
(VdF–TFE) compared with the 5:5 poly(VdF–TFE).
proceeded even after 1200 s of SC CO2 treatment, whereas the increase
Fig. 5 shows the changes in the SAXS intensities for poly(VdF–TFE)
in the case of the 5:5 poly(VdF–TFE) was saturated after around 600 s of
with a ratio of VdF:TFE = 5:5 and 8:2 as a function of the SC CO2
treatment. On the basis of the SAXS and combined SAXS/U-SAXS results,
application time. As shown in Fig. 5, the evaluation of the intensity in­
it is suggested that the characteristics such as the shortened-periodic
crease for the 8:2 poly(VdF–TFE) was much slower than that of the 5:5
morphology caused by the higher VdF content is one of the origins of
poly(VdF–TFE). The SAXS increase for the 8:2 poly(VdF–TFE) gradually
tolerance enhancement to SC CO2 stimulus.

4
K. Yamanoi et al. The Journal of Supercritical Fluids 184 (2022) 105555

stimulus. In contrast with PVdF fibers that change their structure from
4
α-phase to β-phase under the application of stress [14–16], the crystal
VdF:TFE = 5:5
peak intensity change structure of PVdF was maintained in the same manner after the SC CO2
stimulus. The XRD for PTFE was observed at 18.0 and 31.6◦ , which were
3 assigned to the Miller indices of (100) and (110), respectively [34]. As
shown in Fig. 6, transitions in the crystal structure of PTFE, PVdF, and
poly(VdF–TFE) were not detected after the SC CO2 stimulus was applied.
VdF:TFE = 8:2
2 3.4. XRD peak position change

The XRD pattern for the 8:2 poly(VdF–TFE) exhibited the peak
1 around 20◦ , which can be associated with the XRD peak for (110) of
PVdF at 20.1◦ . As the TFE ratio increases, the peak position observed for
0 400 800 1200 the 8:2 poly(VdF–TFE) at around 20◦ gradually shifted toward the XRD
SC CO 2 application time (s) for (100) of PTFE at 18.0◦ ; the lower angle shift indicates the expansion
of the dominant crystal structure. Fig. 7 shows the dominant peak po­
Fig. 5. Changes in SAXS intensities at peak position for poly(VdF–TFE) with sition for the polymers including the change before and after the SC CO2
ratios of VdF:TFE = 5:5 and 8:2 as a function of SC CO2 application time. Initial stimulus. As shown in Fig. 7, the 8:2 poly(VdF–TFE) showed the XRD at
position at 0 s was normalized to 1 on the vertical axis. the higher angle region, even though the polymer is a copolymer
composed of both VdF and TFE. This result indicates that the copolymer
3.3. Crystal structure by WAXS with the higher VdF ratio has specifically tightened crystal structure.
The peak positions were seen to be maintained even after the stimulus
Fig. 6 shows the WAXS signals for PVdF, PTFE, and poly(VdF–TFE) for the homopolymers and copolymers.
with the ratios of VdF:TFE = 3:7, 4:6, 5:5, 6:4, and 8:2. The signals were
acquired for the polymers before and after the SC CO2 treatment stim­
3.5. Degrees of crystallinity
ulus as static measurements. The X-ray diffraction (XRD) for the PVdF
was observed at 17.8, 18.4, 20.1, 25.6, 36.0, and 37.2◦ , which were
The enhancement of the XRD height corresponds to further ordering
assigned to be the Miller indices of (100), (020), (110), (120), (200), and
the crystal structure. Using the WAXS, the degrees of crystallinity were
(210)/(040) in an α-phase PVdF, respectively. As shown in Fig. 6, the
determined for the polymers studied here before and after the SC CO2
WAXS signals for PVdF remained in the same manner after the SC CO2
stimulus. The evaluation procedure of the degree of crystallinity for the

Fig. 6. WAXS for PVdF, PTFE, and poly(VdF–TFE) with ratios of VdF:TFE = 3:7, 4:6, 5:5, 6:4, and 8:2 before and after SC CO2 treatment stimulus. Curves labeled
“before” and “after” indicate before and after SC CO2 application, respectively.

5
K. Yamanoi et al. The Journal of Supercritical Fluids 184 (2022) 105555

Fig. 7. Dependence of VdF:TFE ratios on dominant XRD peak positions before


and after SC CO2 treatment stimulus. Labels of “before” and “after” indicate
before and after SC CO2 treatment, respectively.

typical ratios is shown in Fig. 8. The procedure for all of the polymers
studied here is shown in Fig. S1. The degree of crystallinity and differ­
ence between before and after the SC CO2 stimulus are listed in Table 1.
The WAXS signals were deconvoluted into broad curves (representing
the amorphous phase contribution) and sharp diffraction patterns
(representing the crystal phase contribution). The degree of crystallinity
was calculated from the ratio of the integrals of the amorphous phase
and crystal phase contributions. Although the degrees of crystallinity
were similar before and after the stimulus, a slight increase after the
stimulus was identified in all the polymers studied here. Considering the
enhancement of the crystallinity induced by the SC CO2 treatment, it is
suggested that the SAXS increase shown in Figs. 3 and 4 is related to the
enhancement of structural disorder in the amorphous phase area. On the
basis of the present SAXS and WAXS investigations presented here, it is
proposed that a key feature for the enhanced tolerance is considered as
structural shortening on both the mesoscopic and microscopic length
scales due to polymerization with the higher VdF ratio in the copolymer.
For semicrystalline polymers, little penetration by the SC CO2 fluid is
observed in the crystalline phase, whereas the degree of swelling in the
amorphous phase part is relatively dominant. High-pressure NMR
studies indicate that the high-pressure condition affects the chain
mobility in the amorphous region of semicrystalline polymers [12,
37–39]. The present results depicting the difference in the observed
structural changes between the amorphous phase and crystalline part
are consistent with the literature knowledge.

3.6. Time-resolved and in-situ SWAXS

In this study, the SAXS change induced by the SC CO2 stimulus was
investigated using time-resolved measurements. The static WAXS mea­
surements were conducted using a laboratory X-ray diffractometer with
a goniometer and reflective optics to permit the quantitative determi­
nation of the degree of crystallinity of a polymer. The SAXS and WAXS
signals discussed above were not acquired as simultaneous observations.
Time-resolved and in-situ SWAXS investigation was examined for the 8:2
poly(VdF–TFE), which was selected as a typical polymer possibly having
the highest tolerance property developed in the present stage.
Fig. 9 shows the SWAXS signals under SC CO2 stimulus of 60 ◦ C and
10 MPa. As shown in Fig. 9a, the significant increase in SAXS intensities
and the gradual low-q shift were observed; the changes in SAXS were the
same manner as the observation shown in Fig. 4. The observed change in
the SAXS was faster than that shown in Fig. 4 due to difference in the Fig. 8. Evaluation procedure of degree of crystallinity at typical VdF/TFE ra­
sample thickness. The equilibrium time is comparable to study focused tios. WAXS intensities were deconvoluted into broad curves (amorphous phase
on the swelling kinetics of semicrystalline polymer under a SC CO2 contribution) and sharp diffraction patterns (crystal phase part). Black solid
condition with the different pressure and temperature by in-situ optical lines, black dotted lines and red dotted lines represent experimental data,
observation [40]. As shown in Fig. 9b, the peak position in WAXS deconvoluted fitting, and total fitted results, respectively. Labels of “before”
and “after” represent before and after SC CO2 application, respectively.
remained around the same q-value. This is evidence that the crystalline
part is stable even when subjected to the SC CO2 treatment stimulus. As
also shown in Fig. 9b, a decrease in the WAXS intensities around q

6
K. Yamanoi et al. The Journal of Supercritical Fluids 184 (2022) 105555

Table 1
Degrees of crystallinity for PVdF, PTFE, and poly(VdF–TFE). Labels of “before”
(a) SAXS
and “after” represent before and after SC CO2 application, respectively. 80 80
VdF:TFE (w/w) Degree of crystallinity (%) 12 min
10 min
Before After Difference (After – Before) 60 min
10:0 (PVdF) 42.4 43.0 0.6 8 min
8:2 36.4 39.5 3.1
60

I(q) (arb. units)


6:4 44.6 46.5 1.9 60 6 min
5:5 54.1 56.1 2.0
4 min
4:6 62.0 63.5 1.5
3:7 65.8 68.4 2.6 40
0:10 (PTFE) 48.9 49.2 0.3 2 min

= 12 nm–1 was observed under application of the SC CO2 condition. The 40 0 min
20 after CO2
deconvolution analysis for the 8:2 poly(VdF–TFE) shown in Fig. 8 sug­
initial state
gests that the shoulder peak around q = 12 nm–1 could be generated by
the amorphous phase contribution. Therefore, the profile change in­
dicates an enhancement of the degree of crystallinity in the polymer. 0
The time-resolved WAXS observations are consistent with the results 20 0 0.2 0.4 0.6
obtained via static measurements, which indicates an increase in the q (nm )
0 min
degree of crystallinity by 3.1% as listed in Table 1.
Fig. 10 shows the change in the mesoscopic periodic length, Lmeso­
scale, as a function of the SC CO2 application time. The periodic length
was estimated from the q-values of the SAXS peak position using the
relation Lmesoscale = 2π/q. The peak position was carefully determined 0
via a fitting analysis using the Gaussian function. Although the satura­ 0 1 2 3 4
tion time for the 8:2 poly(VdF–TFE) was not identified in Figs. 4 and 5, q (nm )
the expansion and saturation process were clearly observed in the time-
resolved signals due to thinner sample used (0.6 mm in thickness). As
shown in Fig. 10, the value of Lmesoscale changed from 25.8 nm at 0 min
to 27.1 nm after a 10-min-application time. The result indicated a 5.0%
b) WAXS
expansion of the periodic structure on the mesoscopic scale induced by 10000
SC CO2 application. After the saturation time of ~10 min, the change in
the periodic length plateaued even while subject to a continuing SC CO2 initial state
exposure. The significant expansion and saturation of the change were
similarly observed for the periodic length change in the 5:5 poly
(VdF–TFE) discussed in Fig. 3.
8000

4. Conclusions SC CO 2 application
from 0 min
Neat fluoropolymers and fluorinated copolymers were prepared 6000
to 12 min
through the suspension polymerization method with the mass ratios of
VdF:TFE = 10:0 (PVdF), 8:2, 6:4, 5:5, 4:6, 3:7, and 0:10 (PTFE). The red : 60 min
fundamental structural information on the copolymers on the meso­
scopic scale was obtained using the combined SAXS and U-SAXS mea­
4000
surements. The SAXS and U-SAXS signals suggested that the periodic
structure on the mesoscopic scale become more shortened with
increasing the VdF ratio in the copolymers. An SC CO2 at 60 ◦ C and
10 MPa was applied as an SC fluid stimulus to the fluoropolymers. The
process of structural disorder on the mesoscopic length scale induced by 2000
the stimulus was measured by the time-resolved SAXS technique. The
SAXS change proceeded more slowly with increasing VdF ratio in the
copolymers. The degrees of crystallinity of the fluoropolymers studied
here were evaluated using the WAXS and peak deconvolution analyses.
0
After the SC CO2 stimulus was applied, the degree of crystallinity and 10 12 14 16 18
peak position in the diffraction patterns were analogous to those before
the stimulus. The time-resolved and in-situ SWAXS investigation was q (nm )
examined for the 8:2 poly(VdF–TFE). The SWAXS signals confirmed the
results of the SAXS and WAXS investigations. The SWAXS result indi­ Fig. 9. Time-resolved and in-situ SWAXS signals for poly(VdF–TFE) with ratio
of VdF:TFE = 8:2 under SC CO2 of 60 ◦ C and 10 MPa. (a) SAXS measurements.
cated 5.0% expansion of the mesoscopic periodic structure by SC CO2
Inset shows SAXS change in early stages of SC CO2 application. Dotted red line
application. On the basis of the present SAXS, WAXS, and SWAXS in­
represents peak position at initial state. (b) WAXS measurements. Shoulder
vestigations, it is proposed that the specific properties such as shortened- peak around q = 12 nm–1 highlighted by green arrow was assigned as being
periodic morphology on the mesoscopic scale and compact crystal from amorphous phase in poly(VdF–TFE) with ratio of VdF:TFE = 8:2, ac­
structure by polymerization with the higher VdF ratio are the origins of cording to deconvolution for 8:2 copolymer shown in Fig. 8.
tolerance enhancement to SC CO2 stimulus.

7
K. Yamanoi et al. The Journal of Supercritical Fluids 184 (2022) 105555

28 online version at doi:10.1016/j.supflu.2022.105555.

saturation References
Lmesoscale (nm)

27 [1] S. Ebnesajjad, P.R. Khaladkar. Fluoropolymer Applications in the Chemical


expansion Processing Industries, 2nd edition, William Andrew Publishing, 2017. ISBN:
9780815517290.
5% expansion [2] J.C. Barbosa, J.P. Dias, S. Lanceros-Méndez, C.M. Costa, Recent advances in poly
by SC CO 2 (vinylidene fluoride) and its copolymers for lithium-ion battery separators,
26 Membranes 8 (2018) 45, https://doi.org/10.3390/membranes8030045.
[3] Z. Cui, E. Drioli, Y.M. Lee, Recent progress in fluoropolymers for membranes, Prog.
start CO2 injection Polym. Sci. 39 (2014) 164–198, https://doi.org/10.1016/j.
progpolymsci.2013.07.008.
25 [4] G. Eberle, H. Schmidt, W. Eisenmenger, Piezoelectric polymer electrets, IEEE
0 2 4 6 8 10 12 14 16 Trans. Dielect. Electr. Insul., IEEE Trans. 3 (1996) 624–646, https://doi.org/
10.1109/94.544185.
application time (min) [5] H. Teng, Overview of the development of the fluoropolymer industry, Appl. Sci. 2
(2012) 496–512, https://doi.org/10.3390/app2020496.
[6] S. Atchariyawut, C. Feng, R. Wang, R. Jiraratananon, D.T. Liang, Effect of
Fig. 10. Change in mesoscopic periodic length, Lmesoscale, for the 8:2 poly
membrane structure on mass-transfer in the membrane gas–liquid contacting
(VdF–TFE) as a function of SC CO2 application time. Values of Lmesoscale were process using microporous PVDF hollow fibers, J. Membr. Sci. 285 (2006)
calculated from the time-resolved SAXS profiles obtained from SWAXS mea­ 272–281, https://doi.org/10.1016/j.memsci.2006.08.029.
surement shown in Fig. 9. Result at 0 min includes CO2 fluid injection process [7] J. Kong, K. Li, Oil removal from oil-in-water emulsions using PVDF membranes,
due to experimental operation. Sep. Purif. Technol. 16 (1999) 83–93, https://doi.org/10.1016/S1383-5866(98)
00114-2.
[8] M.M. Teoh, T.S. Chung, Membrane distillation with hydrophobic macrovoid-free
Koga and coworkers found that the swelling of polymer materials PVDF–PTFE hollow fiber membranes, Sep. Purif. Technol. 66 (2009) 229–236,
https://doi.org/10.1016/j.seppur.2009.01.005.
and the porosity in polymer induced by SC treatment depend on the
[9] B.J. Briscoe, O. Lorge, P. Dang, Hardnesses of poly(vinylidene) fluoride and
density fluctuation (inhomogeneity of molecular distribution) of SC CO2 polyamide in high-pressure gas ambience, Philos. Mag. A 82 (2002) 2081–2091,
[32,33]. The time-resolved investigation presented here should be per­ https://doi.org/10.1080/01418610208235718.
formed in more detail under different conditions including temperature, [10] B. Bonavoglia, G. Storti, M. Morbidelli, A. Rajendran, M. Mazzotti, Sorption and
swelling of semicrystalline polymers in supercritical CO2, J. Polym. Sci. B Polym.
pressure, and density fluctuation. Phys. 44 (2006) 1531–1546, https://doi.org/10.1002/polb.20799.
[11] N.M.B. Flichy, S.G. Kazarian, C.J. Lawrence, B.J. Briscoe, An ATR-IR study of poly
CRediT authorship contribution statement (dimethylsiloxane) under high-pressure carbon dioxide: Simultaneous
measurement of sorption and swelling, J. Phys. Chem. B 106 (2002) 754–759,
https://doi.org/10.1021/jp012597q.
Kohei Yamanoi: Investigation, Satoshi Shibuta: Investigation, [12] K.J. Thurecht, D.J.T. Hill, A.K. Whittaker, Equilibrium swelling measurements of
Atsushi Shiro: Investigation, Resources, Masao Noumi: Resources, network and semicrystalline polymers in supercritical carbon dioxide using high-
pressure NMR, Macromolecules 38 (2005) 3731–3737, https://doi.org/10.1021/
Project administration, Investigation, Melvin John F. Empizo: Inves­ ma0503108.
tigation, Marilou Cadatal-Raduban: Investigation, Nobuhiko Sar­ [13] A. Marigo, C. Marega, R. Zannetti, G. Ajroldi, Lamellar morphology by small-angle
ukura: Supervision, Funding acquisition, Keiko Nishikawa: x-ray scattering measurements in some perfluorinated copolymers of
tetrafluoroethylene, Macromolecules 29 (1996) 2197–2200, https://doi.org/
Supervision, Takeshi Morita: Team, Conceptualization, Data curation, 10.1021/ma950809.
Investigation, Writing- Reviewing and Editing, Project administration, [14] H. Guo, Y. Zhang, F. Xue, Z. Cai, Y. Shang, J. Li, Y. Chen, Z. Wu, S. Jiang, In-situ
Funding acquisition. synchrotron SAXS and WAXS investigations on deformation and α-β transformation
of uniaxial stretched poly(vinylidene fluoride), CrystEngComm 15 (2013)
1597–1606, https://doi.org/10.1039/C2CE26578H.
Declaration of Competing Interest [15] J. Wu, J.M. Schultz, F. Yeh, B.S. Hsiao, B. Chu, In-situ simultaneous synchrotron
small- and wide-angle x-ray scattering measurement of poly(vinylidene fluoride)
fibers under deformation, Macromolecules 33 (2000) 1765–1777, https://doi.org/
The authors declare that they have no known competing financial 10.1021/ma990896w.
interests or personal relationships that could have appeared to influence [16] D.L. Chinaglia, R. Gregório Jr., D.R. Vollet, Structural modifications in stretch-
induced crystallization in PVDF films as measured by small-angle x-ray scattering,
the work reported in this paper.
J. Appl. Polym. Sci. 125 (2012) 527–535, https://doi.org/10.1002/app.35684.
[17] Z. Tang, M. Wang, F. Tian, L. Xu, G. Wu, Crystal size shrinking in radiation-induced
Acknowledgments crosslinking of polytetrafluoroethylene: synchrotron small angle x-ray scattering
and scanning electron microscopy analysis, Eur. Polym. J. 59 (2014) 156–160,
https://doi.org/10.1016/j.eurpolymj.2014.07.013.
We express thanks to SPring-8 at JASRI for the opportunity to [18] D. Ghonasgi, S. Gupta, K.M. Dooley, F.C. Knopf, Supercritical CO2 extraction of
perform the measurements of time-resolved SAXS and combined SAXS/ organic contaminants from aqueous streams, AIChE J. 37 (1991) 944–950, https://
U-SAXS. TM and MN would like to express appreciation to Dr. Keiichi doi.org/10.1002/aic.690370617.
[19] M. Raventós, S. Duarte, R. Alarcón, Application and possibilities of supercritical
Osaka at JASRI for his helpful support for the time-resolved SAXS and U- CO2 extraction in food processing industry: an overview, Food Sci. Technol. Int 8
SAXS measurements. We are grateful to PF at KEK for providing the (2002) 269–284, https://doi.org/10.1106/108201302029451.
opportunity to perform the SWAXS measurement. TM expresses appre­ [20] Y. Ahn, S.J. Bae, M. Kim, S.K. Cho, S. Baik, J.I. Lee, J.E. Cha, Review of
supercritical CO2 power cycle technology and current status of research and
ciation to Prof. Nobutaka Shimizu and the stuff members at PF for their development, Nucl. Eng. Technol. 47 (2015) 647–661, https://doi.org/10.1016/j.
helpful support for the SWAXS experiment. The authors are grateful to net.2015.06.009.
Mr. Y. Yong, Mr. T. Taniguchi, Ms. V. C. Agulto, Mr. Y. Minami, Dr. Y. [21] C.K. Ho, B.D. Iverson, Review of high-temperature central receiver designs for
concentrating solar power, Renew. Sustain. Energy Rev. 29 (2014) 835–846,
Abe, Dr. M. V. Luong, Dr. J. L. F. Gabayno, and Prof. T. Shimizu for their https://doi.org/10.1016/j.rser.2013.08.099.
helpful support. This work was financially supported by Daikin In­ [22] B.D. Iverson, T.M. Conboy, J.J. Pasch, A.M. Kruizenga, Supercritical CO2 Brayton
dustries, LTD. NS and KY express thanks to a JSPS Grants-in-Aid for cycles for solar-thermal energy, Appl. Energy 111 (2013) 957–970, https://doi.
org/10.1016/j.apenergy.2013.06.020.
Scientific Research (KAKENHI, No. 18K04727), Creation of Research [23] M. Perrut, Sterilization and virus inactivation by supercritical fluids (a review),
Platforms and Photon Beam Platform. The authors thank Mr. Jason Hart J. Supercrit. Fluids 66 (2012) 359–371, https://doi.org/10.1016/j.
for a careful reading of this paper. supflu.2011.07.007.
[24] A. White, D. Burns, T.W. Christensen, Effective terminal sterilization using
supercritical carbon dioxide, J. Biotechnol. 123 (2006) 504–515, https://doi.org/
Appendix A. Supporting information 10.1016/j.jbiotec.2005.12.033.

Supplementary data associated with this article can be found in the

8
K. Yamanoi et al. The Journal of Supercritical Fluids 184 (2022) 105555

[25] Y. Hakuta, H. Hayashi, K. Arai, Fine particle formation using supercritical fluids, [32] T. Koga, Y.S. Seo, Y. Zhang, K. Shin, K. Kusano, K. Nishikawa, M.H. Rafailovich, J.
Curr. Opin. Solid State Mater. Sci. 7 (2003) 341–351, https://doi.org/10.1016/j. C. Sokolov, B. Chu, D. Peiffer, R. Occhiogrosso, S.K. Satija, Density-fluctuation-
cossms.2003.12.005. induced swelling of polymer thin films in carbon dioxide, Phys. Rev. Lett. 89
[26] F. Cansell, C. Aymonier, A. Loppinet-Serani, Review on materials science and (2002), 125506, https://doi.org/10.1103/PhysRevLett.89.125506.
supercritical fluids, Curr. Opin. Solid State Mater. Sci. 7 (2003) 331–340, https:// [33] T. Koga, E. Akashige, A. Reinstein, M. Bronner, Y.S. Seo, K. Shin, M.H. Rafailovich,
doi.org/10.1016/j.cossms.2004.01.003. J.C. Sokolov, B. Chu, S.K. Satija, The effect of density fluctuations in supercritical
[27] A.I. Cooper, Polymer synthesis and processing using supercritical carbon dioxide, fluids: new science and technology for polymer thin films, Physica B 357 (2005)
J. Mater. Chem. 10 (2000) 207–234, https://doi.org/10.1039/A906486I. 73–79, https://doi.org/10.1016/j.physb.2004.11.025.
[28] L. Li, E.B. Twum, X. Li, E.F. McCord, P.A. Fox, D.F. Lyons, P.L. Rinaldi, 2D NMR [34] X. Li, F. Tian, Z. Tang, C. Yang, X. Li, F. Bian, J. Wang, Study on structural
characterization of sequence distributions in the backbone of poly(vinylidene evolution of polytetrafluoroethylene irradiated by electron beam under stretching
fluoride-co-tetrafluoroethylene), Macromolecules 45 (2012) 9682–9696, https:// using SAXS/WAXS, Eur. Polym. J. 83 (2016) 35–41, https://doi.org/10.1016/j.
doi.org/10.1021/ma3020307. eurpolymj.2016.07.010.
[29] H. Takagi, N. Igarashi, T. Mori, S. Saijo, H. Ohta, T. Nagatani, T. Kosuge, [37] M. de Langen, H. Luigjes, K.O. Prins, High pressure NMR study of chain dynamics
N. Shimizu, Upgrade of small angle x-ray scattering beamline BL-6A at the photon in the orthorhombic phase of polyethylene, Polymer 41 (2000) 1183–1191,
factory, AIP Conf. Proc. 1741 (2016), 030018, https://doi.org/10.1063/ https://doi.org/10.1016/S0032-3861(99)00249-9.
1.4952841. [38] A.G.S. Hollander, K.O. Prins, NMR study of chain motion in atactic polypropylene
[30] T. Morita, Y. Tanaka, K. Ito, Y. Takahashi, K. Nishikawa, Apparatus for the at high pressure, Int. J. Thermophys. 22 (2001) 357–375, https://doi.org/
simultaneous measurement of the x-ray absorption factor developed for a small- 10.1023/A:1010762528795.
angle X-ray scattering beamline, J. Appl. Crystallogr. 40 (2007) 791–795, https:// [39] A.S. Kulik, K.O. Prins, Deuteron nuclear magnetic resonance study of high-pressure
doi.org/10.1107/S0021889807022248. effects on the molecular dynamics in amorphous polyethylene, Polymer 35 (1994)
[31] J.J. Watkins, T.J. McCarthy, Polymerization in supercritical fluid-swollen 2307–2314, https://doi.org/10.1016/0032-3861(94)90766-8.
polymers: a new route to polymer blends, Macromolecules 27 (1994) 4845–4847, [40] J.R. Royer, J.M. DeSimone, S.A. Khan, Carbon dioxide-induced swelling of poly
https://doi.org/10.1021/ma00095a031. (dimethylsiloxane), Macromolecules 32 (1999) 8965–8973, https://doi.org/
10.1021/ma9904518.

You might also like