You are on page 1of 14

Progress in Natural Science: Materials International 32 (2022) 314–327

H O S T E D BY Contents lists available at ScienceDirect

Progress in Natural Science: Materials International


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

Surface characterization of polyimide and polyethylene terephthalate


membranes toward plasma and UV treatments
Thi Sinh Vo a, *, Tran Thi Bich Chau Vo b
a
School of Mechanical Engineering, Sungkyunkwan University, Suwon, 16419, South Korea
b
Department of Industrial Management, Can Tho University, Can Tho, Viet Nam

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

Keywords: This study is to bring an overview on various methods regarding the surface treatment performed on different
Plasma treatment polymeric membranes in relation to a deference between the treatment mechanism and their chemical structure.
UV treatment Herein, plasma and UV surface treatments (PST and UST) were utilized to treat the surfaces of two commercial
Polymeric membrane
membranes, i.e., aromatic polyimide and polyethylene terephthalate for various durations using a discharge
Aromatic polyimide
Polyethylene terephthalate
electrode and an UV LED curing system, respectively. Based on the effects of PST and UST processes on the surface
of these membranes, their surface properties have been characterized to surface morphology and hydrophilicity
(or wettability), at same time the changes have probably occurred on their surface during the PST and UST
processes. As a result, their hydrophilic behavior was significantly enhanced after conducting these treatment
processes based on the investigation of water contact angle. Moreover, Fourier-transform infrared and Raman
spectroscopy manifested the changes in the chemical bonds, interactions and orientations of molecules, which
confirmed the apparition and insertion of polar functional groups after the surface treatments, and accordingly,
their wettability was enhanced significantly. Also, optical microscopy images showed direct impactions on their
surface and the evidence of slight surface aging/damage during the treatment processes based on the presentation
of the strange streaks on the treated surfaces. In particular, the peel forces exhibited an increase comparing to the
initial membranes, which is characteristically for enhancing their adhesion behavior. Hence, these results are
needful for improving the surface properties of a material, and obtaining better adhesion behavior on the
membrane-like materials and recycling polymer materials.

1. Introduction disordering the bulk properties of materials [9,10]. Specifically, plasma


surface treatment (PST) is considered as a complex method for control-
As known, polymeric material system has been widely utilized in ling and modifying the surface properties of materials, which exhibits
various practical applications and research fields [1–8]. Usually, an also a chemically active media to induce chemical reactions on interface
important factor is required to possible interactions of these polymeric surfaces of various materials, i.e., ionized molecular and atomic, photons,
materials generated at the interface surface, at same time these polymeric radicals, etc., meaning a formation of oxygenated chemical groups on the
materials often contain low surface energies that can be inert with surface to probably allow higher surface chemical reactivity [12–14].
chemical activities on the surface of materials corresponding to invalid of Besides, an atmospheric pressure plasma treatment is concerned signif-
polar and actively functional groups. Thus, they are truly needful to icantly, owing to the lower cost and simplified operation [15]. Behaviors
undergo the corresponding surface treatments to improve the adhesive of hydrophilicity (or wettability), adhesion, and surface morphology are
properties for interacting with other materials [9–11]. Among various able to improve and enhance effectively through the use of PST method
surface treatment methods, plasma and ultraviolet (UV) treatments are on the material surface [16,17]. Specially, its outstanding advantage
commonly used to improve the surface properties of polymeric materials, regards to a chemical effect, an example for nylon 6,6 membrane with the
i.e., physical and chemical surface changes in polymers. Actually, the use of PST method, their adhesion and surface properties markedly
polymeric materials exposed to plasma or UV technologies is considered improved through a conversion of carbonyl and hydrocarbon groups into
to be the approaches of reaching relative surface treatments without carboxylic groups [18], at same time that this surface treatment does not

* Corresponding author.
E-mail addresses: vtsinh92@skku.edu (T.S. Vo), vttbchau@ctu.edu.vn (T.T.B.C. Vo).

https://doi.org/10.1016/j.pnsc.2022.03.010
Received 29 September 2021; Received in revised form 19 March 2022; Accepted 29 March 2022
Available online 12 April 2022
1002-0071/© 2022 Chinese Materials Research Society. Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license (http://
creativecommons.org/licenses/by-nc-nd/4.0/).
T.S. Vo, T.T.B.C. Vo Progress in Natural Science: Materials International 32 (2022) 314–327

Fig. 1. Chemical structures and images of PI and PET.

almost impact to the mechanical properties during the surface treatment exposed directly under the discharge electrode and UV LED curing sys-
for a material [19,20]. For UV surface treatment (UST), it is considered tem to be treated their surfaces becoming better hydrophilicity (or
that the treatment probably breaks up molecular bonds and interactions wettability), adhesion, and surface behaviors, owing to physical and
on the surface of materials through high-energy UV photons; concomi- chemical surface changes in polymers. Different treating and recovering
tantly, an UV oxidation process can form several possible functional durations are also investigated on the analysis based on the water contact
groups in relation to hydroxyl, carboxyl, or carbonyl groups that en- angle (WCA) measurement, Fourier-transform infrared (FTIR) and
hances hydrophilicity and adhesion behaviors presented on the treated Raman spectroscopy, and optical microscopy (OM) images, which are
surface of materials [21,22]. A general aim in the use of both PST and probably the support to confirm the hydrophilicity and surface properties
UST are able to induce physical and chemical surface changes in poly- of the polymeric membranes after the surface treatments, and a 90 peel
mers through several concurrent processes, i.e., etching, grafting, poly- test is applied to examine their adhesion behaviors.
merization, cross-linking, usually without modifying their original bulk
qualities [23]. In these approaches, low-pressure plasma and UV treat- 2. Materials and methods
ments are able to be proposed to modify many surface properties of
polymers, including adhesion, friction, penetrability, wettability, 2.1. Materials
dye-ability, or biocompatibility, which adapt to specific applications
[23]. Thereby, these approaches are favorable for opening up of a wide The commercially available polyimide (PI) (0.075 mm  750  750
spectrum of surface polymers with desired compositions. mm2) and polyethylene terephthalate (PET) (0.125 mm  600  600
Nowadays, polymeric materials were applied widely in multiple mm2) membranes (see Fig. 1) were purchased from Good-fellow Com-
modern industries; however, almost all polymeric materials show the pany. A 3M scotch brand tape (5413K) was offered by a Korean company.
surfaces with low hydrophilicity that can significantly impact the
wettability, biocompatibility, printability, and adhesion of the materials. 2.2. Plasma and UV surface treatments
Hence, polymer materials, especially for polymeric membranes, are truly
necessary to undergo surface treatments to reach better hydrophilicity, Both PI (0.075 mm  150  150 mm2) and PET membranes (0.125
adhesion, and surface activities [24–26]. In particular, aromatic poly- mm  150  150 mm2) were treated by an atmospheric plasma generator
imide (PI) and polyethylene terephthalate (PET) are polymeric mem- (FEMTO SCIENCE, plasma power ¼ 150 W, gas pressure ¼ 1.5 torr) and
branes utilized widely in several different applications [27]. PI an oven UV instrument (MS TECH, UV power ¼ 5000 mW/cm2, λ ¼ 365
membrane contains imide monomers pertaining to a high performance nm). The distance between the membrane surface and the electrode, and
plastic membrane, which is frequency applied to other rugged organic UV LED curing system was 100 mm and 30 mm, respectively. All mem-
materials, i.e., displays, fuel cells, etc., owing to its availably high branes were surveyed with different tested times (t ¼ 0; 2; 5; 10 min) by
heat-resistance [28]. Whereas, PET, a polyester, which is a linear ther- plasma and UV instruments. Then, the surfaces of treated polymeric
moplastic polymer, has a vast applicability based on numerous of membranes were recovered corresponding to the above treated times (t
modification processes [27]. Overall, these polymeric membranes were ¼ 2; 5; 10 min). The experimental systems of PST and UST utilized in this
different in the chemical structure and the surface characterization, i.e., study are shown in Fig. 2.
amine, hydroxyl, carboxyl, and aldehyde groups, which enable the polar
surface behaviors happening on the surface of both polymeric mem-
2.3. Water contract angle measurement
branes after the PST and UST processes [16,17]. Plasma and UV treat-
ments are powerful techniques to modify only the polymer surface.
In order to determine the hydrophobic–hydrophilic behavior of the
Simultaneously, the main aim of this work is to probably give an over-
polymeric membranes after PST and UST, a drop technique was applied
view of PST and UST approaches, which all were performed on the sur-
to investigate the water contact angle (WCA) [SEO Phoenix MT(M)] by
face of PI and PET membranes to grant a relation of the treatment
using a distilled water drop (volume ¼ ~5 μL; average values, n ¼ 3). The
mechanism and the corresponding changes in the chemical structure of
WCA values were calculated by analyzing the captured drop images
treated membranes. In another hand, these polymeric membranes will be
basing on the ImageJ® program.

315
T.S. Vo, T.T.B.C. Vo Progress in Natural Science: Materials International 32 (2022) 314–327

Fig. 2. Images and schematic diagrams of PST (A–C) and UST (D–F) systems.

2.4. FTIR and Raman analysis of tensile test (a utilized size ¼ 10  50 mm2) was performed by a uni-
versal tensile machine (UTM model 5565, UK) with 250 N of load cell
The chemical characterization of the treated membrane surfaces was and 10 mm⋅min1 of pulling rate (average values, n ¼ 3).
analyzed by a Fourier-transform infrared (FTIR) spectrophotometer
(Nicolet 380, Ietled Co.) and a Raman spectrum (XperRam200) with a 2.6. SEM and OM measurements
laser wavelength of 532 nm. It was recorded at the wavenumber region of
4000–500 cm1 and the Raman Shift region of 3200–100 cm1 corre- The initial morphological structures of PI and PET membranes (i.e.,
sponding to FTIR and Raman spectrums. top-surface and cross-section) were captured from scanning electron
microscopy (SEM) (FESEM JSM-7600F) at different magnifications and
2.5. Peel strength and tensile analysis positions. Besides, in order to observe the change in the surface
morphology of the treated membranes, their morphology surfaces were
A 90 peel test was carried out on a Peel Adhesion Strength Tester investigated before and after PST and UST using an optical microscopy
instrument with a 50 N of universal load cell and 10 mm⋅min1 of pulling (OM) (Olympus SZX12) and i-Solution™ IMTcamCCD Digital Camera.
speed. The membrane was tightly fixed on a platform, and the tape was
firmly attached in the upper clamp that would pull the tape up at 90 , 3. Results and discussion
while the under platform moved to the right side to could be maintained
with the 90 of angle. With increasing a steady rate of the upper clamp, 3.1. WCA and surface morphology of polymeric membranes
the peeling force could be recorded through peeling the tape off of the
membrane (average values, n ¼ 3). Concomitantly, the stress-strain curve The morphological characterization of a membrane-like material is

316
T.S. Vo, T.T.B.C. Vo Progress in Natural Science: Materials International 32 (2022) 314–327

Fig. 3. SEM images of initial PI (A–B) and PET (C–D) membranes: top-surface (A, C) and cross-section (B, D). SEM images of PI (E–F) and PET (G–H) membranes with
PST (E, G) and UST (F, H) (t ¼ 5 min).

one of important factors to acquire desired evaluations depending on appear on their corresponding top-surface and cross-section regions.
different studies' purposes. Herein, the morphological structure of Thereby, these initial membranes were the smooth and flat surfaces. The
membranes has been detected by SEM images (Fig. 3), indicating that the surface treatment is considered as an expressing step so that it is able to
top-surface (Fig. 3 A, C) and cross-section (Fig. 3 B, D) morphologies of successfully coat the other materials on the treated substrates conducted
the initial membranes (i.e., untreated membranes, t ¼ 0 min) were in various applications, owing to physical and chemical surface changes
relatively smooth and non-rough surfaces, and the pores did also not in polymers. To better understand the changes of the PST- and UST-

317
T.S. Vo, T.T.B.C. Vo Progress in Natural Science: Materials International 32 (2022) 314–327

Fig. 4. WCA images of PI and PET membranes after PST–Recovery (A) and UST–Recovery (B) processes with different tested times (t ¼ 0; 2; 5; 10 min).

treated membranes, their corresponding topography has been also which could be favorable for significantly enhancing wettability and
characterized by SEM images (Fig. 3E–H). It showed that the strange adhesion behaviors of the treated PI membrane.
streaks appeared on the surfaces of these treated membranes (t ¼ 5 min) The surface treatment is known to successfully coat the other mate-
comparing to the initial membranes (yellow arrows were illustrated in rials on the treated substrates conducted in various applications. For the
the SEM images), and their surfaces became rougher than that of un- hydrophobic–hydrophilic behavior of the polymeric membranes after
treated membranes. As such, these resulted in the evidence of slight PST and UST processes, a drop technique was applied to investigate the
surface aging/damage during the PST and UST processes, especially for WCA by using a distilled water drop. The WCA values were calculated
the PST. Similarly, this phenomenon has been also clearly demonstrated through analyzing the captured drop images. Herein, we will firstly
in relevantly previous researches regarding to surface modifications of assess the different PST and UST treated durations (t ¼ 2; 5; 10 min), as
treated polymers [29–32]. As seen in Fig. 3E–H, the strange streaks on an investigation for possible changes on these polymer surfaces, and then
the PST-treated membranes (Fig. 3 E, G) appeared more densely than the surfaces of treated membranes were recovered corresponding to the
those of the UST-treated ones (Fig. 3 F, H), revealing that the effect of above treated durations. In theory, the WCA value (θ) is less than 90
atmospheric plasma treatment with the electric direction was more indicating a hydrophilic surface of a material. Besides the smooth and flat
effective comparing to the UV LED field. Besides, the PST- and surface morphologies of the initial membranes (Fig. 3 A, C), the surfaces
UST-treated PI membranes (Fig. 3E and F) were observed with the presented the WCA of 56.7  3.7 and 61.4  4.3 corresponding to the
rougher surfaces comparing to the treated PET ones (Fig. 3G and H), initial PI and PET membranes (t ¼ 0 min) (Figs. 4 and 5); therefore, the

318
T.S. Vo, T.T.B.C. Vo Progress in Natural Science: Materials International 32 (2022) 314–327

Fig. 5. WCA values of PI and PET membranes after PST (A) – Recovery (B) and UST (C) –Recovery (D) processes with different tested times (t ¼ 0; 2; 5; 10 min). WCA
images of PI and PET membranes after PST–Recovery–PST process (E) (t ¼ 5 min). Schematic diagram of hydrophobic–hydrophilic behavior of membrane before and
after PST/UST–Recovery (F).

wettability of PI membrane was better than that of PET one, due to the (Fig. 5F). Actually, the modified surfaces by the atmospheric PST process
available hydrophilic nature of each membrane. After performing the based on several factors, especially for the average energy occurring in
PST and UST processes, the WCA values of both the treated PI and PET free electrons, which was in charge to make certain for changes in
membranes exhibited a decrease corresponding to the increasing treated chemical bonds, interactions and orientations of molecules in the mem-
durations, and which reached the lower WCA values comparing to the branes being effective more than the UST process, by using UV LED field.
initial membranes, especially in the PST process [Fig. 4 (A, B) and Fig. 5 The wettability is considered as an especial trend of the adhesive
(A, C)]. As such, the wettability improved significantly on the surface- behavior of polymeric materials; concomitantly, it also relates to an
treated membranes (i.e., PST-treated PI membrane: 73.9%/2 min, important relationship with surface energy. It means that the surface
84.7%/5 min, 100%/10 min; and PST-treated PET membrane: 38.6%/2 energy is increased according to the wettability (or hydrophilicity) [29].
min, 49.5%/5 min, 50.8%/10 min; and UST-treated PI membrane: Additionally, Figs. 4A and 5A depicted the changes in the WCA for the
15.2%/2 min, 30.3%/5 min, 1.6%/10 min; and UST-treated PET mem- PST-treated polymeric membranes, resulting that the WCA for a polar
brane: 6.0%/2 min, 2.4%/5 min, 0.2%/10 min). Moreover, the medium (water) reduced significantly after 5 min of PST, and the
wettability of PST-treated membranes was greater than that in the UST remaining membranes unchanged much beyond 5 min (i.e., PI mem-
process, maybe due to the contribution of polar functional groups rep- brane: 14.8  3.3 /2 min, 8.7  3.5 /5 min, 0 /10 min; and PET
resented on the surface of PST- and UST-treated membranes [33,34] membrane: 37.7  3.9 /2 min, 31.0  4.5 /5 min, 30.2  3.6 /10

319
T.S. Vo, T.T.B.C. Vo Progress in Natural Science: Materials International 32 (2022) 314–327

Fig. 6. OM images of PI (A–C) and PET (D–F) membranes before (A, D) and after PST (B, E) and UST (C, F) (t ¼ 5 min). Schematic diagram of surface morphology
behavior of membrane before and after PST/UST (G).

min). Whereas, the change in the WCA values of UST-treated membranes 0.9%/10 min; and recovered PET membrane after UST: 8.0%/2 min,
was not according to the above-suggested regulations, such as an increase 0.6%/5 min, 3.9%/10 min), meaning that the wettability of these
in the treated PI membranes and an almost constant in the treated PET recovered membranes was less not much than this of treated membranes.
membranes occurring between the treatment durations of 5 and 10 min; Obviously, the recovery ability of the polar component involved to the
in contrast, a decrease in the treated PI membranes and a rise in the unstable polar functional groups on the surface, which were able to be
treated PET membranes happened between the treatment durations of 2 oxidized and formed peroxides associating to the surrounding oxygen
and 5 min (i.e., PI membrane: 48.1  3.9 /2 min, 39.5  4.5 /5 min, and moisture [13], meaning that this decrease was attributed to an in-
55.8  3.6 /10 min; and PET membrane: 57.7  3.3 /2 min, 62.9  crease in the molecular orientations leading to reducing the possibly
3.5 /5 min, 62.3  3.0 /10 min) (Figs. 4B and 5C). Overall, besides 5 dispersive interactions [13], which was suitable and clearly analyzed by
min of the treatments, 10 min was also able to use successful treating on Shaw et al. [13]. Overall, the wettability of these polymeric membranes
the surface of polymeric membranes. Although there were differences in occurring in the PST and recovery processes was still greater than that in
the changes of each surface treatment method, a tested time of 5 min is the UST and recovery process, meaning that the effect of hydrophilic
able to be selected as an optimal time for examining essential modifi- surface was happened better during the PST process. The hydrophilic
cations on the surface of polymeric membranes to reduce the process groups were organized onto the surface of polymeric membranes after
time. PST and UST processes; in particular, these functional groups were able
In the recovery processes [Fig. 4 (A, B) and Fig. 5 (B, D)], comparing to be oxidized or oriented after the recovered time, based on the increase
to the initial membranes, the WCA values still display a decrease with the in the WCA values. Nonetheless, there were still some rational steps being
increasing recovered durations, as well as a significant improvement in missed (i.e., uniformity, mobility, hydrophobicity, orientation, polarity
the wettability occurring on the membrane surfaces, excepting for the and interaction) [35,36], the recovery ability of the hydrophobicity
recovered PET membrane after UST (i.e., recovered PI membrane after happening on the UST-treated surfaces was mainly attributed to the
PST: 72.8%/2 min, 55.7%/5 min, 73.5%/10 min; and recovered PET mobility and orientation of the polar functional groups [13]. To further
membrane after PST: 40.7%/2 min, 48.5%/5 min, 49.2%/10 min; and understand, the recovered membranes after the PST process were utilized
recovered PI membrane after UST: 14.8%/2 min, 22.5%/5 min, 2.5%/10 for treating once more time (t ¼ 5 min) on the surface of these recovered
min; and recovered PET membrane after UST: 1.5%/2 min, 3.1%/5 membranes (Fig. 5E). As a result, the WCA values of polymeric mem-
min, 3.7%/10 min). Besides, according to a comparison of the corre- branes after the second PST process were almost equal with the ones after
sponding treated-recovered membranes, the WCA values showed a slight the first PST process (i.e., 9.0  3.9 /PI and 31.8  4.3 /PET), sug-
increase, excepted for the 5 min - recovered PI membrane after PST (i.e., gesting that their wettability was still well maintained, might be due to
recovered PI membrane after PST: 4.1%/2 min, 65.5%/5 min; and the repeated physical (i.e., aging or damage) and chemical (i.e., chemical
recovered PET membrane after PST: 3.4%/2 min, 1.9%/5 min, 3.3%/10 bonds, interactions and orientations of molecules) surface changes of in
min; and recovered PI membrane after UST: 0.4%/2 min, 11.1%/5 min, the membranes.

320
T.S. Vo, T.T.B.C. Vo Progress in Natural Science: Materials International 32 (2022) 314–327

Fig. 7. FTIR (A, B) and Raman (C, D) spectrums of PI (A, C) and PET (B, D) membranes before and after PST and UST.

Moreover, Fig. 6 (A-F) showed the OM images of the membranes’ actual, a heterogeneous discharge was happened with numerous current
surface before and after the PST and UST processes (t ¼ 0 min and 5 min). pulses at atmospheric pressure that was able to create an asymmetric
Resulting that the whole of corresponding views of the treated mem- erosion on the surface of the polymeric membranes leading to eroding
branes (t ¼ 5 min) indicated evenly strange streaks with different colors and modifying the surface morphology, at same time that it could also
on the surfaces comparing to the initial membranes (t ¼ 0 min), as induce the differences in changes of chemical bonds, interactions and
noticed by yellow dashed arrows on the OM images. In another word, the orientations of molecules occurring on the surface of polymeric mem-
characterization by OM images displayed direct impactions on the initial branes (Fig. 6G). These were obviously pointed in the increase in the
surface of these membranes and the evidence of slight surface aging/ wettability of polymeric membranes after the PST process comparing to
damage during the PST and UST processes [37], i.e., physical and the UST one, at same time that the wettability enhancement of treated
chemical surface changes in polymers, which was also appropriate with membranes was also agreed in relevant research of Toufik et al. [29].
the abovementioned SEM results. Indeed, Novak et al. [38] investigated
that the crystallinity degree of isotactic polypropylene had an influence
on the aging of PST-treated polymeric materials, indicating that 3.2. FTIR and Raman spectroscopy of polymeric membranes
restricted mobility (crystalline polymer) decreased the aging rate effect
by reducing the surface polymer chains rearrangement. Herein, the In general, it is truly necessary to determine possible changes on the
number of these streaks were appeared greater on the PST-treated surface of polymeric membranes during the PST and UST processes, as
membrane surfaces comparing to the UST-treated one, meaning that well as their mentioned chemical characterization. In addition to the
the effect of modified surface was occurred better during the PST process, abovementioned morphological characterization of the polymeric
which was also agreed in the abovementioned changes of WCA values. In membranes with the treated surfaces, Fig. 7A and B showed the FTIR
spectra of polymeric membranes before and after the PST and UST

321
T.S. Vo, T.T.B.C. Vo Progress in Natural Science: Materials International 32 (2022) 314–327

Table 1 Table 3
FTIR and Raman characteristics and the corresponding chemical bonds of initial FTIR and Raman characteristics and the corresponding chemical bonds of PST-
PI and PET membranes (t ¼ 0 min). and UST-treated PET membrane (t ¼ 5 min).
FTIR, cm1 Chemical bonds of FTIR, cm1 Chemical bonds of PET FTIR, cm1 PST-treated PET FTIR, cm1 UST-treated PET
PI
3510–3250 Broader O–H 3510–3250 Broader O–H
3520–3180 CH2 (ethyl 3510–3250 O–H 3065 C–H (phenyl ring) 3065 C–H (phenyl ring)
groups) 2970 CH2 (ethyl groups) 2973 CH2 (ethyl groups)
2931–2895 C¼O (imide I 3065 C–H (phenyl ring) 2909 H–C¼O 2909 H–C¼O
linkage) 1721 C¼O (ester linkage) 1719 C¼O (ester linkage)
1705 C–H (ethyl groups) 2972 CH2 (ethyl groups) 1567 C–H (phenyl ring) 1577 C–H (phenyl ring)
1219 C–O (backbone) 2909 H–C¼O 1477 C–C (phenyl ring) 1505 C–C (phenyl ring)
1494–1347 C–N (imide II/III 1720 C¼O (ester linkage) 1361–1230 C–O (ester linkage) 1350–1227 C–O (ester linkage)
1186–899 linkages) 1151–900 C–H (ethyl groups) 1150–889 C–H (ethyl groups)
729 C–C (backbone) 1598 C–H (phenyl ring) 720 Sharper and larger 720 Sharper and larger
N–H 1508 C–C (phenyl ring) C–C (backbone) C–C (backbone)
1359–1228 C–O (ester linkage)
Raman, cm¡1 PST-treated PET Raman, cm¡1 UST-treated PET
1152–905 C–H (ethyl groups)
720 C–C (backbone) 3083 Benzene ring 3083 Benzene ring
skeleton skeleton
Raman, cm¡1 Chemical bonds of Raman, Chemical bonds of PET
1730 C¼O (ester linkage) 1730 C¼O (ester linkage)
PI cm¡1
1511 C¼C 1511 C¼C
1785 C¼O (imide I) 3083 Benzene ring skeleton 731–1351 C–O (ester linkage) 731–1351 C–O (ester linkage)
1509 C¼C 1730 C¼O (ester linkage) 1612 C–C (phenyl ring) 1612 C–C (phenyl ring)
1398 C–N–C axial (imide 1511 C¼C 1178–1186 C–C (phenyl ring) 1178–1186 C–C (phenyl ring)
II) <900 C–H 1089 Crystallinity degree of
1260 C–O–C backbone 731–1351 C–O (ester linkage) trans structure
1125 C–N–C transverse 1612 C–C (phenyl ring) 883 (weak) – –O–CH2CH2–O– <900 C–H
(imide III) 994 (strong) (gauche and trans)
865 Diamine ring 1178–1186 C–C (phenyl ring) 883 (weak) – –O–CH2CH2–O–
breathing 994 (strong) (gauche and trans)
739–826 C–H <900 C–H
640 C–O–C backbone 883–994 –O–CH2CH2–O– (gauche
orientations and trans) membrane relating to the C–H (phenyl ring) and H–C¼O vibrations,
1621–750 Dianhydride part respectively [39]. The typical peaks of the PI membranes were observed
at 1705 cm1, 1494–1347 cm1 and 1186–899 cm1 corresponding to
the C¼O (imide I linkage), C–O (backbone) and C–N (imide II and III
Table 2 linkages) stretching vibrations, while those of the PET membrane were
FTIR and Raman characteristics and the corresponding chemical bonds of PST- suggested at 1720 cm1, 1508 cm1 and 1359–1228 cm1 relating to the
and UST-treated PI membrane (t ¼ 5 min). C¼O (ester linkage), C–C (phenyl ring) and C–O (ester linkage) stretching
FTIR, cm1 PST-treated PI FTIR, cm1 UST-treated PI vibrations [39]. Concomitantly, the peaks at 729 cm1 (PI) and 720 cm1
3520–3180 Broader N–H 3520–3180 Broader N–H (PET) contributed to the C–C (backbone) bending vibration. Moreover,
2941–2900 CH2 (ethyl groups) 2952–2905 CH2 (ethyl groups) the peaks of C–H (ethyl groups) bending vibration were displayed at
1689 C¼O (imide I linkage) 1678 C¼O (imide I linkage) 1219 cm1 (PI) and 1152–905 cm1 (PET), at same time those of C–H
1720 C¼O (asymmetric 1718 C¼O (asymmetric
bending vibration containing in phenyl ring were observed at 1598 cm1
tension) tension)
1220–1249 C–H 1211–1240 C–H
in the FTIR spectrum of initial PET membrane [39]. In particular, the
1514–1359 C–O (backbone) 1526–1366 C–O (backbone) broad peak at 3520–3180 cm1 (PI) and 3510–3250 cm1 (PET) was
1186–899 Broader and larger C–N 1186–899 Broader and larger C–N characteristic to the N–H and O–H stretching vibrations, respectively
(imide II/III linkages) (imide II/III linkages) [39].
729 Sharper and larger 729 Sharper and larger
For the PST- and UST-treated PI membrane (t ¼ 5 min) (Fig. 7A and
C–C(backbone) C–C (backbone)
Table 2), it displayed the characteristic peak corresponding to the C¼O
Raman, PST-treated PI Raman, UST-treated PI
stretching vibration at 1689 cm1 (PST) and 1678 cm1 (UST), at same
cm¡1 cm¡1
time these treated PI membranes indicated another peak at 1720 cm1
1785 Broader C¼O (imide I) 1785 Sharper C¼O (imide I) (PST) and 1718 cm1 (UST) that also regarded to the C¼O vibration
1509 C¼C 1509 C¼C
1415 Broader C–N–C (imide 1411 C–N–C (imide II)
(asymmetric tension vibration) at lower intensities. Besides, the peaks
II) relating to the C–O stretching vibration moved to higher wavenumber
1260 C–O–C backbone 1260 C–O–C backbone regions (1514–1359 cm1/PST and 1526–1366 cm1/UST) and higher
1139 C–N–C (imide III) 1130 C–N–C (imide III) intensities comparing to the initial PI membrane (1494–1347 cm1). For
865 Diamine ring breathing 865 Diamine ring
the PST- and UST-treated PET membrane (t ¼ 5 min) (Fig. 7B and
breathing
739–826 C–H 739–826 C–H Table 3), the peaks of C¼O and C–O stretching vibrations almost un-
640 C–O–C backbone 640 C–O–C backbone changed much about the wavenumber regions, i.e., the C¼O stretching
orientations orientations vibration was observed at 1721 cm1 (PST) and 1719 cm1 (UST), while
610 Benzene deformation 1629 (strong) Dianhydride part the C–O vibration was showed at 1361–1230 cm1 (PST) and
750(weak)
1350–1227 cm1 (UST), but their intensities were higher than those of
750 (weak) Dianhydride part

initial PET membrane. As such, these suggested the small surface changes
processes (t ¼ 5 min) to investigate the chemical characterization of their of polymeric membranes occurring in the wavenumber region of
treated surfaces. For the FTIR spectrums of initial PI and PET membranes carboxyl groups during the PST and UST processes [39], especially in the
(t ¼ 0 min) (Fig. 7A and B and Table 1), all exhibited the characteristic PST process. Simultaneously, the other changes of N–H (PI) and O–H
peaks at 2931–2895 cm1 (PI) and 2972 cm1 (PET) corresponding to (PET) stretching vibrations could be observed in the FTIR spectra of
the –CH2 (ethyl groups) stretching vibrations, especially for the peaks at treated polymeric membranes occurred in the very broad wavenumber
3065 cm1 and 2909 cm1 appeared in the FTIR spectrum of initial PET regions of 3520–3180 cm1 (PI) and 3510–3250 cm1 (PET) and higher

322
T.S. Vo, T.T.B.C. Vo Progress in Natural Science: Materials International 32 (2022) 314–327

intensities (Fig. 7A and B). As abovementioned hydrophilic behaviors, benzene ring was at <900 cm1 [45], and the C¼C vibration related to
this investigation manifested a considerable quantity of polar functional the peak at 1511 cm1. The notable peaks at 883 cm1 (weak) and 994
groups on the surface of treated polymeric membranes. Besides, the peak cm1 (strong) were attributed to the gauche and trans structures in the
of C–H beading vibration (1219 cm1) appeared in the treated PI ethylene glycol segments (–O–CH2CH2–O–) regarding to the orientation
membranes (1220 cm1/PST and 1211 cm1/UST), another peak of C–H and crystallinity degree of the PET membrane [43,46]; moreover, the
vibration was also presented at 1249 cm1 and 1240 cm1 corresponding weak peak at 3083 cm1 was yielded on the benzene ring skeleton
to the PST- and UST-treated PI membranes (Fig. 7A and Table 2). vibration.
Whereas, the peaks of C–H bending vibration (1152–905 cm1) were There were some changes of chemical bonds, interactions and ori-
larger and sharper, at same time they moved slightly to lower wave- entations of molecules in the Raman spectrums of PST- and UST-treated
number regions (1151–900 cm1/PST and 1150–889 cm1/UST) in the polymeric membranes (t ¼ 5 min) (Fig. 7C and D). Notably, all intensities
treated PET membranes (Fig. 7B and Table 3). Furthermore, the treated lowered after the PST and UST processes, maybe due to the chain length
polymeric membranes have been found the C–C (backbone) bending reduction regarding to the chain breakage, i.e., the C–H and C–C groups
vibration (729 cm1/PI and 720 cm1/PET) at very sharp and large would be fractured by the yielded free radicals on the membranes' sur-
peaks comparing to the initial polymeric membranes (Fig. 7A and B) faces leading to combining with the species (PST and UST) and producing
[39]. Notably, the intensities of C–H bending (1598 cm1) and C–C the polar groups on these treated surfaces, at same time the molecular
stretching (1508 cm1) vibrations containing in phenyl ring decreased at chains and planes were also rearranged and reoriented [47]. Conse-
lower wavenumber regions occurring in the treated PET membranes quently, these peaks were also moved to the lower Raman shifts' regions
(Fig. 7B and Table 3), while the C–N stretching vibrations (1186–899 comparing with the initial membranes. The peak of mono-substituted
cm1) corresponding to the imide II and III linkages of treated PI mem- benzene deformation (610 cm1) in the PI membrane spectrum
branes reached at the broader and larger peaks (Fig. 7A and Table 2). As a (Fig. 7C and Table 2) also changed for the intensity and the region of
result, there were the obvious changes occurring on the surface of Raman shift after the PST process; simultaneously, the characteristic
polymeric membranes during the PST and UST processes. Thereby, the peaks of the initial PI membrane at 1621 cm1, 1398 cm1 and 1125
bending and stretching vibrations were well determined consisting of cm1 shifted to other regions of Raman shift in the PST- and UST-treated
their relationships in the possible changes on the treated polymeric PI membrane (i.e., PST: those at 1632 cm1, 1415 cm1 and 1139 cm1;
surfaces, but researches in the low wavenumber range were uncommon and UST: those at 1629 cm1, 1411 cm1 and 1130 cm1). Thereby,
[40]. these differences manifested that the aromatic imide ring and C–N–C
As known, there are some different principles between Raman and groups (imide II and III) were much affected through these treatment
FTIR spectra. Specifically, the FTIR spectroscopy bases on the changes of processes, especially for the PST process [48]. In another word, the
molecular dipole moment, while the Raman spectrum involves to the dipole C¼O moment containing in the aromatic imide ring could vibrate
polarizations-induced molecules. In fact, both the FTIR and Raman much more in the electric direction comparing to the UV LED field, which
spectrums support mutually on the molecular structures. Indeed, the led to moving the molecular segments and orientating the dipole in the
peaks of stretching/bending vibrations can be strong in the Raman cases of the corresponding vectors. On the other hand, this moment could
spectra, and which cannot be specifically observed in the FTIR spectrum. also vibrate according to the opposite direction in the aromatic imide
One of the most notable features of Raman spectroscopy is regarded to ring. As such, the changes of chemical bonds, interactions and orienta-
the photonic probe that can be seen as a nondestructive detection tions of molecules, specifically in the aromatic imide ring, which would
approach. Previously, both the FTIR and Raman spectrums utilized occur and be much impacted leading to the Raman shifts moving to larger
models of variable calibration for morphology characterization of poly- regions compared to other groups on a basic of repeated vibrations in the
meric materials, but the molecular orientations in a polymeric material two directions [48]. Furthermore, it could be also found in the rela-
regarding its available composition can considerably affect to the con- tionship of amide-imide linkages and the orientations of imide linkages
tained band shapes and intensities. Herein, the Raman spectrum was also occurred in the PI membrane after the PST and UST processes, especially
used to evaluate the nature of chemical bonds, interactions and orien- in the PST process (Fig. 7C) [49]. Basically, the peak at 1621 cm1 (ar-
tations of molecules in the polymeric membranes. Specific Raman spec- omatic imide ring) was selected as a reference peak, and which could
trums of initial and treated polymeric membranes’ surfaces were allow manifesting the orientations of various availably functional groups
displayed in Fig. 7C and D. For the Raman spectra of initial PI membrane regarding to the imide phenyl ring, i.e., the C–O–C backbone (1260
(t ¼ 0 min) (Fig. 7C and Table 1), the characteristic peaks were exhibited cm1) orientation was corresponded to the aromatic imide ring and the
at 1785 cm1, 1398 cm1 and 1125 cm1 corresponding to the stretching C¼O group (1785 cm1, imide I). As a result, the intensity of C–O–C
vibrations of imide I (C¼O stretching vibration), imide II (C–N–C axial backbone lowered and the C¼O group was more stretched during the PST
stretching vibration) and imide III (C–N–C transverse stretching vibra- and UST processes, which could be due to the effects of the radiation and
tion) [41,42]. Besides, the C–O–C backbone orientations involved to the the alternating electric field respectively, especially for the C¼O
aromatic imide ring at 1260 cm1 (stretching vibration) and 640 cm1 stretching occurred in the PST process [49,50]. This phenomenon could
(bending), at same time that those of aromatic imide ring in di-anhydride probably induce the possible chemical bond fractures, the formation of
part were indicated at 1621 cm1 (strong) and 750 cm1 (weak) [41,42]. free radicals and the PI molecular degradation on the basic of PST process
The peaks at 1509 cm1, 865 cm1 and 739–826 cm1 belonged to the [50]; instead of the reorientations and reversion of the chemical bonds in
C¼C, diamine ring breathing and C–H vibrations, respectively. For the the PI molecules basing on the UST process. Also, this was occurred
initial PET membrane spectra (t ¼ 0 min) (Fig. 7D and Table 1), it similarly with the C–N–C backbone orientation (1398 cm1/imide II and
indicated that the main peak intensities were larger, which could be due 1125 cm1/imide III), which led to stretching the C–N–C vibrations more
to the PET molecules included the high Raman scattering intensity during the PST process [50]. For the PST- and UST-treated PET mem-
benzene ring. The typical peaks were displayed at 1730 cm1 (sharp and brane (Fig. 7D and Table 3), in addition to the abovementioned main
large peak) and 731–1351 cm1 regarding to the C¼O (ester linkage) and chain length reductions, the molecular chain rearrangements and plane
C–O (ester linkage) stretching vibrations, while the peaks of C–C (phenyl reorientations occurred during the PST and UST processes, the phenyl
ring) vibrations were seen at 1178–1186 cm1 (anti-symmetric stretch- ring contributions to the Raman spectra were sited at 1612 cm1 (C–C
ing vibration) and 1612 cm1 [43,44]. Especially, the bending vibration stretching vibrations), 1178–1186 cm1 (anti-symmetric stretching vi-
peak of C–H with isolated adjacent hydrogen and hydrogen bond on bration) and 1511 cm1 (C¼C vibration), which were almost unchanged

323
T.S. Vo, T.T.B.C. Vo Progress in Natural Science: Materials International 32 (2022) 314–327

Fig. 8. Schematic diagram of membrane under a 90 peel test (A). Peel Force-Length curves of PI (B, D) and PET (C, E) membranes before and after PST (B, C) and UST
(D, E).

much during the PST and UST processes that could be due to possible abovementioned hydrophilic behaviors, more specially for the apparition
strength of phenyl ring [43,44,46,51]. This approach was similar with and insertion of polar functional groups (C¼O and –OH) after the surface
the carbonyl stretching peaks (C¼O and C–O groups, ester linkages), treatments.
which did not also appear on the changes of intensity and Raman shift
regions [40]. Notably, the peak at 994 cm1 (trans) was observed clearly 3.3. Peel strength behavior of polymeric membranes
more than that at 883 cm1 (gauche) (–O–CH2CH2–O– segments) in both
the PST- and UST-treated PET membrane relating to the amorphous For a surface-treated membrane, the adhesive behavior is an impor-
phase of the treated PET membrane, main due to the reorientation and tant factor to assess performances of the surface modification. The most
rearrangement of these segments under the effects of electric and UV LED usual way is used to investigate the bonding and interaction happened
fields, especially in the PST process [43,46]. However, the new peak between the surfaces (or layers), named “peel test”. Tape will be manu-
appeared at 1089 cm1 in the UST-treated PET membrane, which was a ally applied to the testing surface, which is then quickly removed. In this
probe of crystallinity degree and a conformation of the trans structure work, a 90 peel test was conducted to evaluate the adhesive perfor-
[52]. Obviously, there were the changes occurred on the membranes’ mances on the treated surface of these membranes, as described in
surface during the PST and UST processes, which were based on the Fig. 8A. For the initial membranes (t ¼ 0 min), the tape was peeled with a
chemical bonds, interactions and orientations of molecules through the calculated average force of 2.99  0.35 N and 2.79  0.31 N, while a
FTIR and Raman spectroscopy, and it can significantly support to the maximum force obtained at 3.09  0.31 N and 2.89  0.33 N

324
T.S. Vo, T.T.B.C. Vo Progress in Natural Science: Materials International 32 (2022) 314–327

Table 4 calculated average force almost unchanged significantly). Generally,


The peel force values of all samples. these peel forces of initial PI membrane were greater than those of initial
Samples Maximum Force Average Force Calculated Average PET one (Table 4), maybe due to the hydrophilic behavior of the mem-
(N) (N) Force (N) brane surface and the chemical bonds/interactions between the tape and
PST-treated PI – 0 3.09  0.31 3.00  0.32 2.99  0.35 these membranes [53–55]. After performing the PST and UST processes,
min the peel forces exhibited an increase comparing to the initial membranes,
PST-treated PI – 2 3.28  0.33 3.01  0.34 3.07  0.34 especially in the PST process (Fig. 8B–E). The adhesive behavior
min mentioned in this study was characterized by a delamination of the tape
PST-treated PI – 5 3.67  0.35 3.21  0.39 3.25  0.38
min
from these membranes without any tearing, which was demonstrative in
PST-treated PI – 3.56  0.40 3.34  0.29 3.39  0.32 reversible bonds and interactions (Fig. 9B and C). Specifically, it
10 min described three stages taking place during the 90 peel process, such as:
PST-treated PET – 2.89  0.33 2.78  0.31 2.79  0.31 (1) the tape would begin peeling from the surface of polymeric mem-
0 min branes, (2) which continued then partial peel from this surface consisting
PST-treated PET – 3.11  0.36 2.99  0.38 2.97  0.35 of extra elongation and bending of the tape layer. Resulting that the local
2 min peel angle was able to be variable leading to more complexing and so-
PST-treated PET – 3.28  0.38 2.98  0.37 2.94  0.32
5 min
phisticating in this stage, and then (3) the tape was completely peeled
PST-treated PET – 3.44  0.34 3.20  0.32 3.22  0.38 from this surface [53–55]. Actually, the case of low peel forces shows that
10 min the chemical bonding and interactions do not happen or were weak in
UST-treated PI – 0 3.09  0.31 3.00  0.32 3.05  0.33 order to probably couple the tape onto the polymeric membranes. Based
min on a comparison of the initial and treated membranes, the adhesive
UST-treated PI – 2 3.07  0.42 3.11  0.36 3.16  0.37 behavior favored the treated specimens more, including the tape onto the
min PST- and UST-treated membranes.
UST-treated PI – 5 3.23  0.41 3.13  0.38 3.20  0.35
min
In the case of increasing treated durations (t ¼ 2; 5; 10 min), the tape
UST-treated PI – 3.28  0.39 3.20  0.39 3.24  0.42 was still delaminated from the treated membranes during testing,
10 min meaning that the maximum and calculated average forces significantly
UST-treated PET – 2.89  0.33 2.78  0.31 2.82  0.33 improved with the treated durations (Fig. 8B–E). Besides, the adhesive
0 min behavior favored the PST-treated specimens more through their high
UST-treated PET – 3.31  0.39 3.09  0.38 3.02  0.32 peel forces (Table 4). The adhesion behavior of both PST- and UST-
2 min treated processes was happened well through the delamination of tape
UST-treated PET – 1.98  0.32 3.10  0.37 3.09  0.35
5 min
onto these membrane surfaces that was main due to the reversible bonds
UST-treated PET – 2.72  0.33 3.18  0.32 3.15  0.39 and interactions [53–55]; however, changing the treatment method from
10 min the UST to PST process improved significantly the regarding peel forces
(Table 4). Similarly, the PST of low-density polyethylene and PET also
resulted in adhesion enhancement by up to a factor of two to ten [56].
corresponding to the initial PI and PET membrane (Fig. 8B–E and
These adhesion enhancement measurements were also associated to
Table 4) (the maximum and calculated average peel forces of each
relevantly previous researches of Shenton et al. [56,57], where the
specimen were determined as in Fig. 9A, the average force and the

Fig. 9. Peel Force-Length profile of a membrane under 90 peel test (A, B). Schematic diagram of membrane under a 90 peel test with various peeling steps (C).
Stress-Strain curve of initial membranes (D).

325
T.S. Vo, T.T.B.C. Vo Progress in Natural Science: Materials International 32 (2022) 314–327

surface energy of polymers could be altered and surface oxygen content References
increased by exposure to the approaches of these surface treatments. In
our opinion, the effect of membrane surface and in the gap of the treated [1] T.S. Vo, T.T.B.C. Vo, J. Eng. Sci. Tech. Rev. 13 (2020) 110–116.
[2] T.S. Vo, T.T.B.C. Vo, T.T. Tien, N.T. Sinh, J. Turkish. Chem. Soc. Sec. A: Chemistry 8
surfaces could form a possibly binding matrix to enhance the adhesion (2021) 519–526.
occurring between the tape and the membrane surface by forming the [3] T.S. Vo, T.T.B.C. Vo, N.D. Pham, T.N.H. Lai, J. Turkish. Chem. Soc. Sec. A:
possible chemical bonds and interactions, at same time that this surface Chemistry 8 (2021) 787–802.
[4] T.S. Vo, J. Turkish. Chem. Soc. Sec. A: Chemistry 8 (2021) 1121–1136.
treatment cannot reach sufficient bond strength between the treated [5] T.S. Vo, T.T.T.N. Vo, T.T.B. Chau, J. Eng. Sci. Tech. Rev. 14 (2021) 21–36.
membranes and the tape layer in the case of lower treated durations [55], [6] T.S. Vo, T.T.B.C. Vo, T.S. Nguyen, T.T. Tien, J. Turkish. Chem. Soc. Sec. A:
the discharge electrode and UV LED were not enough time to be probably Chemistry 8 (2021) 1045–1056.
[7] T.S. Vo, T.T.B.C. Vo, T.S. Nguyen, N.D. Pham, Prog. Nat. Sci.: Mater. Int. 31 (2021)
conducted for achieving polarity reversal interface on the treated mem- 664–671.
brane surface. Concomitantly, the tensile stress-strain curve of each [8] T.S. Vo, T.T.B.C. Vo, T.T. Tran, N.D. Pham, Prog. Nat. Sci.: Mater. Int. (2021),
initial polymeric membrane was presented in Fig. 9D, indicating that the https://doi.org/10.1016/j.pnsc.2021.10.001.
[9] B. Mutel, M. Bigan, H. Vezin, Appl. Surf. Sci. 239 (2004) 25–35.
PET membrane was stiffer than the PI membrane relating to the stress at
[10] M. Mavadat, M. Ghasemzadeh-Barvarz, S. Turgeon, C. Duchesne, G. Laroche,
break and Young's modulus of PET membrane higher than those of PI Langmuir 29 (2013) 15859–15867.
membrane. Overall, the peel forces of the treated membranes exhibited [11] F. Arefi-Khonsari, M. Tatoulian, F. Bretagnol, O. Bouloussa, F. Rondelez, Surf.
an increase comparing to the initial membranes, characteristically for Coating. Technol. 200 (2005) 14–20.
[12] C. Wang, M. Du, J. Lv, Q. Zhou, Y. Ren, G. Liu, D. Gao, L. Jin, Appl. Surf. Sci. 349
enhancing their adhesion behavior. Moreover, under the conditions of (2015) 333–342.
this study, the enhancement of adhesion behavior was readily achievable [13] D. Shaw, A. West, J. Bredin, E. Wagenaars, Plasma Sources Sci. Technol. 25 (2016),
by the PST and UST methods, more especially for the PST. Although this 065018.
[14] K. Terpiłowski, Int. J. Polymer. Sci. (2017) 2017.
corresponding enhancement by the UST was not as large as that [15] E.E. Kunhardt, IEEE Trans. Plasma Sci. 28 (2000) 189–200.
achievable by the PST, it is still significant as in many cases only small [16] Z. Gao, J. Sun, S. Peng, L. Yao, Y. Qiu, Appl. Surf. Sci. 256 (2009) 1496–1501.
changes in adhesion are necessary. Also, we hope that by process opti- [17] L. Zhu, W. Teng, H. Xu, Y. Liu, Q. Jiang, C. Wang, Y. Qiu, Surf. Coating. Technol.
202 (2008) 1966–1974.
mization, the degree of adhesion enhancement for the UST can be [18] H. Lee, I. Ohsawa, J. Takahashi, Appl. Surf. Sci. 328 (2015) 241–246.
improved. [19] C. Mandolfino, E. Lertora, S. Genna, C. Leone, C. Gambaro, Procedia. Crip. 33
(2015) 458–463.
[20] N. Perkas, G. Amirian, S. Dubinsky, S. Gazit, A. Gedanken, J. Appl. Polym. Sci. 104
4. Conclusion (2007) 1423–1430.
[21] C.H. Thatcher, B.R. Adams, Chem. Eng. Sci. 230 (2021) 116204.
In summary, the various approaches of the PST and UST were applied [22] G. Scarselli, D. Quan, N. Murphy, B. Deegan, D. Dowling, A. Ivankovic, Appl.
Compos. Mater. 28 (2021) 71–89.
successfully on the PI and PET membranes, indicating that the changes
[23] N. Inagaki, S. Tasaka, K. Ishii, J. Appl. Polym. Sci. 48 (1993) 1433–1440.
occurred on the surface of these membranes. Their hydrophilic behavior [24] A.H. Poulsson, R.G. Richards, Surface Modification Techniques of
significantly improved through the WCA investigation, as well as there Polyetheretherketone, Including Plasma Surface Treatment, PEEK Biomaterials
were the changes in the chemical bonds, interactions and orientations of Handbook, Elsevier, 2012, pp. 145–161.
[25] J. Wang, C. Pan, N. Huang, H. Sun, P. Yang, Y. Leng, J. Chen, G. Wan, P.K. Chu,
molecules based on the FTIR and Raman spectroscopy. In another word, Surf. Coating. Technol. 196 (2005) 307–311.
these spectrums confirmed the apparition and insertion of polar func- [26] B. Fillon, Micromanufacturing Engineering and Technology, William Andrew
tional groups (C¼O and –OH) after the surface treatments, accordingly, Applied Science Publishers, USA, 2010, pp. 241–263.
[27] J. Charles, G.R. Ramkumaar, S. Azhagiri, S. Gunasekaran, E-J. Chem. 6 (2009)
their wettability enhanced significantly. Besides, the SEM and OM im- 23–33.
ages also displayed the direct impactions on these membranes' surface [28] W.W. Wright, M. Hallden-Abberton, Polyimides, Ullmann's encyclopedia of
and the evidence of slight surface aging/damage by the strange streaks’ industrial chemistry. doi: https://doi.org/10.1002/14356007.a21_253.
[29] M. Toufik, A. Mas, V. Shkinev, A. Nechaev, A. Elharfi, F. Schue, Eur. Polym. J. 38
presentation on the PST- and UST-treated surfaces. Moreover, their peel (2002) 203–209.
strength behavior exhibited an increase comparing to the initial mem- [30] G. Poletti, F. Orsini, A. Raffaele-Addamo, C. Riccardi, E. Selli, Appl. Surf. Sci. 219
branes, characteristically for enhancing their adhesion behavior. A (2003) 311–316.
[31] M. Rodríguez, E. Vazquez-Velez, H. Martinez, A. Torres, J. Nuclear. Physics. Mater.
comparison between the PST and UST methods, the performance of PST Radia. Appl. 8 (2021) 191–196.
reached better than that of UST, owing to special characteristics of PST [32] O. Flores, B. Campillo, F. Castillo, H. Martínez, J. Colín, J. Nuclear. Physics. Mater.
system. Besides, the physical and chemical surface changes in the PI Radia. Appl. 8 (2021) 129–134.
[33] W. Viratyaporn, R.L. Lehman, J. Therm. Anal. Calorim. 103 (2011) 267–273.
membrane (an aromatic thermoplastic and thermosetting polymer) were
[34] T. Zhang, F.D. Blum, J. Colloid Interface Sci. 504 (2017) 111–114.
more effective than those of PET one (a thermoplastic polymer) that was [35] W.W. Hu, Y.T. Hsu, Y.C. Cheng, C. Li, R.C. Ruaan, C.C. Chien, C.A. Chung,
attributed to its availably chemical features of structure. These results C.W. Tsao, Mat. Sci. Eng. C-Mater. 37 (2014) 28–36.
were needful for enhancing surface characterization, and reaching better [36] J. Bacharouche, H. Haidara, P. Kunemann, M.F. Vallat, V. Roucoules, Sensor.
Actuat. Phys. 197 (2013) 25–29.
adhesion behavior on the membrane-like materials and recycling poly- [37] V. Takke, N. Behary, A. Perwuelz, C. Campagne, J. Appl. Polym. Sci. 114 (2009)
mer materials. Concomitantly, this work also gave an overview on 348–357.
various approaches of the surface treatment performed on different [38] I. Novak, S. Florian, J. Mater. Sci. 39 (2004) 2033–2036.
[39] K. Krishnan, R. Krishnan, Raman and infrared spectra of ethylene glycol, in:
polymeric materials. Proceedings of the Indian Academy of Sciences-Section A, Springer, 1966,
pp. 111–122.
Authors’ contributions [40] B.J. Bulkin, M. Lewin, F.J. DeBlase, Macromolecules 18 (1985) 2587–2594.
[41] H. Ishida, S.T. Wellinghoff, E. Baer, J.L. Koenig, Macromolecules 13 (1980)
826–834.
T.S. Vo: Conceptualization, Methodology, Formal analysis, Investi- [42] H. Ishida, M. Huang, Spectrochim. Acta Mol. Biomol. Spectrosc. 51 (1995)
gation, Data curation, Visualization, Writing – original draft & review & 319–331.
[43] J. Rodriguez-Cabello, L. Quintanilla, J. Pastor, J. Raman Spectrosc. 25 (1994)
editing. T.T.B.C. Vo: Writing – review & editing. 335–344.
[44] R.J. Young, W.-Y. Yeh, Polymer 35 (1994) 3844–3847.
Declaration of competing interest [45] C. Zhu, N. Tong, L. Song, G. Zhang, Investigation of Raman Spectra of Polyethylene
Terephthalate, International Symposium on Photonics and Optoelectronics 2015,
International Society for Optics and Photonics, 2015, p. 96560E.
The authors declare that they have no known competing financial [46] C. Lesko, J. Rabolt, R. Ikeda, B. Chase, A. Kennedy, J. Mol. Struct. 521 (2000)
interests or personal relationships that could have appeared to influence 127–136.
the work reported in this paper. [47] J. Nakamatsu, L.F. Delgado-Aparicio, R. Da Silva, F. Soberon, J. Adhes. Sci. Technol.
13 (1999) 753–761.

326
T.S. Vo, T.T.B.C. Vo Progress in Natural Science: Materials International 32 (2022) 314–327

[48] P. Samyn, G. Schoukens, Surface and Interface Analysis: an International Journal [52] P. Colomban, J.H. Ramirez, R. Paquin, A. Marcellan, A. Bunsell, Eng. Fract. Mech.
devoted to the development and application of techniques for the analysis of 73 (2006) 2463–2475.
surfaces, Interface. Thin. Flims. 40 (2008) 853–857. [53] M. Katz, R.J. Theis, IEEE Electr. Insul. Mag. 13 (1997) 24–30.
[49] P. Samyn, P. De Baets, J. Van Craenenbroeck, F. Verpoort, G. Schoukens, J. Appl. [54] D. Fabiani, G. Montanari, IEEE Electr. Insul. Mag. 17 (2001) 24–33.
Polym. Sci. 101 (2006) 1407–1425. [55] J.A. Cella, Polym. Degrad. Stabil. 36 (1992) 99–110.
[50] K.C. Kao, J. Appl. Phys. 55 (1984) 752–755. [56] M.J. Shenton, M.C. Lovell-Hoare, G.C. Stevens, J. Phys. D Appl. Phys. 34 (2001)
[51] A. Melveger, J. Polym. Sci. 2 Polym. Phys. 10 (1972) 317–322. 2754–2760.
[57] M.J. Shenton, G.C. Stevens, J. Phys. D Appl. Phys. 34 (2001) 2761–2768.

327

You might also like