You are on page 1of 10

Sustainable Chemistry and Pharmacy 19 (2021) 100367

Contents lists available at ScienceDirect

Sustainable Chemistry and Pharmacy


journal homepage: http://www.elsevier.com/locate/scp

Activation of potassium peroxymonosulfate for rhodamine B photocatalytic


degradation over visible-light-driven conjugated polyvinyl chloride/Bi2O3
hybrid structure
Huu-Huan Pham a, b, Sheng-Jie You c, Ya-Fen Wang c, Minh Thi Cao d, Van-Viet Pham a, b, *
a
Faculty of Materials Science and Technology, University of Science, VNU-HCM, 227 Nguyen Van Cu Street, District 5, Ho Chi Minh City, 700000, Viet Nam
b
Vietnam National University – Ho Chi Minh City, Thu Duc District, Ho Chi Minh City, 700000, Viet Nam
c
Department of Environmental Engineering, Chung Yuan Christian University, Chung-Li 320, Taiwan
d
Ho Chi Minh City University of Technology (HUTECH), 475A Dien Bien Phu Street, Binh Thanh District, Ho Chi Minh City, 700000, Viet Nam

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

Keywords: In this study, a conjugated polyvinyl chloride is supported hydrothermally synthesized Bi2O3 semiconductor
Conjugated polymer (cPVC/Bi2O3) by a simple, and straightforward method at low temperature. Results show that 2% cPVC/Bi2O3/
Hybrid materials PMS/Vis system being the best photocatalytic ability of RhB degradation (200 mg L− 1) with a complete
Photocatalysis
degradation after 150 min under visible light irradiation. The photocatalytic ability of cPVC/Bi2O3/PMS system
Polyvinyl chloride
Visible light
for RhB degradation has also compared to previous publications. Moreover, the trapping test results are com­
bined with the measurement of electron spin resonance (ESR) spectroscopy indicated that • SO−4 , and • OH are
primary parameters of the photocatalytic RhB degradation.

1. Introduction activate PMS, including both inorganic and organic semiconductors


such as TiO2 (Golshan et al., 2018), ZnO, SnO2 (Bang et al., 2018), and
Advanced oxidation processes (AOPs) have been considered a pri­ g-C3N4 (Wang et al., 2019). Besides the individual semiconductors, the
ority, a high performance, and a green technology to remove contami­ coupling of semiconductors or the combination of a semiconductor with
nants in wastewater. (Elmolla and Chaudhuri, 2010; Méndez-Díaz et al., metal nanoparticles such as Co, Ni, Mn, Ag, etc. have also been excitedly
2010; Benitez et al., 2011; Babuponnusami and Muthukumar, 2012; studied (Lin et al., 2016; Nekouei et al., 2018). However, the limitations
Jung et al., 2012; Karci et al., 2012). AOPs are defined as oxidation of inorganic materials synthesized processes are intricately synthesized
processes involving the generation of hydroxyl radicals (• OH) (Truong procedures, using many toxic precursors, and expensive chemicals that
et al., 2019) and sulfate radicals (• SO−4 ) (Ha et al., 2020); therein, the lead to the problematic application in the industry scale (Leonard et al.,
redox potential of the anion sulfate (• SO−4 ) is much higher than that of 2002; Irmawati et al., 2004). Among inorganic semiconductors, Bi2O3

OHradicals, leading to extend the long life span and strong oxidation shows a high visible-light photocatalytic activity due to its narrow
(Deng and Zhao, 2015). It is indicated that the use of ⋅ SO−4 is preferred bandgap (about 2–3.9 eV at room temperature). Also, the cost-effective
than that of • OHin wastewater treatment recently (Chen et al., 2007). and straightforward synthesized processes, nontoxic nature, and the
There are many routes to activate • SO−4 from peroxydisulfate or perox­ easily control of the morphology, and crystal phase are its remarkable
ymonosulfate (PMS), for instance, thermal, photochemical, radiolytic, properties. However, the rapid recombination of photoinduced
photocatalysis, or redox decomposition. (Zhao et al., 2010). Therein, the electron-hole pairs stills a drawback of Bi2O3 as a photocatalyst. To
activation of PMS over visible-light-driven photocatalysis is an efficient enhance the photocatalytic activity of Bi2O3, the combination of Bi2O3
approach compared to other methods in reducing the energy con­ with semiconductors such as TiO2, ZnO, BiOCl, conductive polymers,
sumption and the dosage of catalysts. (Chi et al., 2015; Hu et al., 2017; etc. is often mentioned (Balachandran and Swaminathan, 2012;
Lin et al., 2017). Gómez-Cerezo et al., 2014; Priya et al., 2016).
Recently, there are many photocatalysts that are used to efficiently As we have known, polyvinyl chloride (PVC) is a useful polymer

* Corresponding author. Faculty of Materials Science and Technology, University of Science, VNU-HCM, 227 Nguyen Van Cu Street, District 5, Ho Chi Minh City,
700000, Viet Nam.
E-mail address: pvviet@hcmus.edu.vn (V.-V. Pham).

https://doi.org/10.1016/j.scp.2020.100367
Received 16 September 2020; Received in revised form 16 December 2020; Accepted 20 December 2020
Available online 9 January 2021
2352-5541/© 2020 Elsevier B.V. All rights reserved.
H.-H. Pham et al. Sustainable Chemistry and Pharmacy 19 (2021) 100367

which is widely used in many applications such as the plastic industry, 2.3. Characterization of materials
household appliances, construction materials, etc.(Bidoki and Witt­
linger, 2010)., PVC can be easily transferred to conjugated PVC (cPVC) Properties of materials, including the morphology, crystal structure,
as a conductive polymer by a dehydrochlorination reaction, leading to functional groups, and chemical bonds, are measured by FESEM images,
the formation of a new partial conjugated polymer chain (Michell, 1985; XRD patterns, FTIR spectra. The XRD pattern is measured as 2θ in the
Rogestedt and Hjertberg, 1993; Owen et al., 1994; Starnes and Ge, range of 10o–70◦ to match the crystal structure and crystal size of ma­
2004). According to the best of our knowledge, this is the first-time terials, the scanning rate of 1.25◦ /min, and the step size of 0.02◦ , using
combination of Bi2O3 and cPVC was employed to activate PMS under samples XRD pattern (Bruker D8, Advance 5005). Subsequently, for
visible light for RhB degradation. Moreover, using cPVC is a potential FTIR analysis, KBr was served as the reference sample; its weight
approach to utilize PVC waste and reduce the dispose of plastic waste selected was 1% of the sample weight. The vibrations of the molecules
into the environment. are expressed in the range of 480–4000 cm− 1, with a step size of 1 cm− 1,
In this study, the cPVC/Bi2O3 composites are synthesized by a simple using the FTIR spectrum (JASCO-4700). To determine the band energy
method at low temperature. The characterizations of cPVC/Bi2O3 structure of material, UV–Vis DRS and Mott-Schottky plot were selected
composites are investigated in detail by X-ray diffraction (XRD) pattern, for implementation. Therein, the absorption of light was investigated
Fourier-transform infrared spectroscopy (FTIR), field emission scanning when the wavelengths ran from 200 to 800 nm, using the DRS spectrum
electron microscopy (FESEM), UV–Vis diffuse reflectance spectroscopy (JASCO V-550), and electrochemical measurements were performed by
(UV–Vis DRS), Raman spectroscopy and electrochemical characteriza­ an electrochemical workstation (Biologic SP - 240). To convert the po­
tions. The PMS activation over cPVC/Bi2O3 hybrid materials under tential vs. Ag/AgCl (3 M KCl) (EAgCl (V)) to NHE (ENHE (V)) at 25 ◦ C
visible light irradiation is studied based on rhodamine B (RhB) degra­ (Eqn. (1)):
dation, effect of the photocatalysts, the PMS concentration, the recycle 0
ENHE = EAg/AgCl + 0.21V (1)
ability. Finally, trapping experiments and ESR studies are inspected to
determine the primary parameters affect degradation reactions. Also,
where ENHE is the potential of a normal hydrolysis electrode, and
the ability to completely degrade RhB in cPVC/Bi2O3 system extremely
high concentration is shown. E0Ag/AgCl is the initial potential of the Ag/AgCl electrode.
For Raman analysis, samples were placed in the analytical position.
2. Experimental Then, bond fluctuations are recorded in 1000–4000 cm− 1 region using
Raman scattering spectroscopy (HORIBA) with laser excitation source at
2.1. Materials 532 nm. Finally, ESR signals were investigated by Bruker EMX Plus X-
Band spectrophotometer, generated during the photocatalytic process
Commercial polyvinyl chloride (PVC), ethylene glycol with DMPO-based spin-trapped reagents. The source of irradiation
(EC2nH4n+2On+1 Merck, 99%), tetrahydrofuran (THF, 99.0%), Bismuth xenon short-arc lamp (Labguide® 150 W Mercury arc lamp, λ > 400
nitrate pentahydrate (Bi(NO3)3⋅5H2O, 99%), hydrochloric acid (HCl, nm). Typical test conditions: temperature 298 K, central field 3483 G,
37% concentration), sodium hydroxide (NaOH, Merck, 99%), potassium sweep width 200 G, frequency 9755 GHz, modulation amplitude 1.00 G,
iodide (KI, 99.99%), isopropyl alcohol (IPA; C3H8O, 99.99%), sodium microwave power 20 mW, sample concentration 400 ppm, 1000 ppm
sulfate anhydrous (Na2SO4, Merck, 99%), potassium dichromate DMPO.
(K2Cr2O7, 99.99%), tert-butanol (TBA; (CH3)3COH, 99.99%), Deionized
water (DI), and a rhodamine B (RhB) dye was used as a pollutant organic
dye indicator to evaluate the photocatalytic activity of cPVC/Bi2O3 2.4. Evaluation of the photocatalytic activity of materials
materials.
Initially, the catalysts were taken with a dosage of 0.33 g L− 1 of 20
2.2. Synthesis of materials mg L− 1 RhB dye solution. After that, 0.5 mM PMS was added to this
solution and stirred magnetically in the dark for 1.5 h to achieve
Bismuth oxide (Bi2O3) was synthesized based on the previous study adsorption/desorption. The mixture was illuminated under visible light
(Yan et al., 2014). Firstly, 2.91 g Bi(NO3)3⋅5H2O powder was dissolved from simulated solar light (Osram Ultra-Vitalux 300 W) using UV cutting
in 30 mL glycerol and 30 mL absolute ethanol solution. The mixture was glass. At intervals of 30 min, the suspension was withdrawn and
then treated with vigorous stirring and ultrasonic dispersion to form a centrifuged rapidly to extract the solution. The absorption spectra of the
homogeneous and transparent solution. The solution as mentioned solutions were measured using a UV–Vis spectrophotometer (U2910,
above was transferred into a teflon lined stainless autoclave, and this HITACHI, Japan). The photocatalytic efficiency of cPVC/Bi2O3 com­
system was heated at 160 ◦ C for 3 h. Secondly, the autoclave was posites and their photocatalytic reaction rates were calculated as Eqn.
naturally cooled to room temperature and filtered the precipitate with (2) and Eqn. (3):
deionized water and absolute ethanol several times, then dry at 80 ◦ C for C0 − Ct
2 h. Finally, the residue was calcined at 350 ◦ C, with a heating rate of η= × 100% (2)
C0
2 ◦ C/min for 2 h to form Bi2O3 materials.
For the cPVC, according to the optimized parameters from our pre­ Ct
= e− kt
(3)
vious study (Bui et al., 2020). First, PVC was slowly added to a solution C0
containing 4 g of NaOH and 100 mL of ethylene glycol solution. Second,
the mixture was stirred for 60 min. Third, it was poured into a round where η is the photocatalytic degradation efficiency of RhB (%), C0 is the
neck three-bottomed flask, then stirred magnetically, and was purged in initial concentration of RhB, Ct is the concentration of RhB at time t, and
nitrogen gas at 190 ◦ C for 3 h. The solution was then thoroughly washed k is the reaction rate calculated from the Langmuir-Hinshelwood (L – H)
with DI water and dried at 60 ◦ C for 24 h to obtain cPVC. model (min− 1).
For the cPVC/Bi2O3 composites, the different amounts of cPVC and The trapping test experiments were carried out to determine the
1.0 g of Bi2O3 were mixed in 10 mL of THF solution to create different primary aspects of the RhB photocatalytic degradation with scavengers
ratios of cPVC and Bi2O3 (1:100, 2:100, 3:100, 5:100 and 7:100). including KI, K2Cr2O7, TBA were used as trapping agents h+ , e− ,
Finally, the cPVC/Bi2O3 composites were magnetically stirred for 30 •
OHradical, respectively. Meanwhile, IPA is a scavenger for trapping
min and ultrasonicated for 30 min. Then, the composites were dried at both • SO−4 , and • OHradicals. The concentration of the scavengers was 1
60 ◦ C for 60 min. mM to the initial photocatalyst. The experimental conditions for these

2
H.-H. Pham et al. Sustainable Chemistry and Pharmacy 19 (2021) 100367

experiments are similar to the photocatalysis test. The stability of cat­ The results demonstrated that the as-prepared materials are not the
alysts was investigated by the evaluation of the photocatalytic RhB secondary phase. It can be concluded that the successful synthesis of
degradation over materials after 4 cycles and catalysts were analyzed by α-phase is the most stable phase of both Bi2O3 and the cPVC/Bi2O3
the FTIR spectrum. composites.
Raman spectroscopy was added to provide information on the exis­
3. Results and discussion tence of cPVC and Bi2O3 (Fig. 2). The Raman spectroscopy analysis of
PVC shows that there are 4 positions of Raman signals including C–Cl
Fig. 1 (a) exhibits a broadband of the wavenumber at around bonds (636 and 694 cm− 1) (Solodovnichenko et al., 2016a) and C–H
3400–3500 cm− 1, associating to the stretching vibrations of hydrogen- (1430 and 2918 cm− 1) (Solodovnichenko et al., 2016b). After the
bonded surface water molecules and hydroxyl groups. The vibration dichlorination process, these four peaks have a significant lower in­
peaks located at 1637 and 879 cm− 1 wavenumber that were strength­ tensity, which clearly shows a decrease in the existence of C–Cl and C–H
ened to confirm the formation of = C–H and C–H bonds, respectively. bonds in the sample. In addition, major peaks appeared at 1123 and
The vibration peaks at 1386-1287 cm− 1 region, and 1459 cm− 1 repre­ 1513 cm− 1, corresponding to the elongation of variation (C–C) and
sent the C–O prolonged vibration, and the C–H bending vibration, (C–
– C) of polyene, respectively (Ellahi et al., 1995; Zhang et al., 2016;
respectively, indicating the presence of the organic solvents as ethanol Bui et al., 2020). These results concluded that the cPVC has been suc­
and glycerol in the product (Xiong et al., 2008; Wang et al., 2009). In cessfully synthesized from chlorination of PVC. For the Bi2O3 and 2%
addition, the vibration peak at 844 cm− 1 is due to the symmetrical cPVC/Bi2O3, the Raman peaks observed at 319 and 452 cm− 1 are due to
elongation of the Bi–O–Bi bonding of the Bi2O3 (Irmawati et al., 2004; the presence of Bi2O3 (In et al., 2011; Wang et al., 2012, 2014). For the
He et al., 2011) and the regions at 700–400 cm− 1 wavenumber represent 2% cPVC/Bi2O3 sample, characteristic peaks of C–C and C– – C at 1123
oxygen-metal vibrations that could be assigned to the existence of Bi2O3 and 1513 cm− 1 in cPVC and distinct peaks of Bi2O3 also exist in this
in the product (Wang et al., 2015). XRD patterns of materials (Fig. 1 (b)) composite.
provide the crystallinity, phase existence, and a confirmation of the The surface morphology of the samples was observed by FESEM
formation’s materials in the FTIR analysis. Results show that the typical images that are exhibited in Fig. 3. First, the morphology of cPVC is
diffraction peak of cPVC at 2θ = 25◦ is appeared and existed when cPVC aggregated particles with the particle size from 500 nm to 1 μm (Fig. 3(a
and Bi2O3 are combined in the composite materials. Besides, Fig. 1 (b) and b)) while Bi2O3 has the nanosheets structure that is formed over­
shows that there are typical diffraction peaks at 2θ = 25.86, 27.07, lapping and alternating sheets (Fig. 3(c and d)). For combining the two
27.54, 28.17, 33.19, 35.13, 37.17, 42.44, 46.22, 48.88, 53.25, 54.94, materials, cPVC and Bi2O3 together, the existence of cPVC particles in
57.16, and 58.11◦ , corresponding to the diffraction peaks of the (002), the composite sample is presented clearly in Fig. 3 (e). However, the
(111), (120), (012), (121), (022), (112), (122), (311), (113), (321), particle boundary of cPVC particles is no longer clear. The presence of
(241), (222), and (024) planes of the α monoclinic phase of Bi2O3 ma­ Bi2O3 is confirmed in Fig. 3 (f) that the nanosheet shape is observed
terials, respectively (JCPDS No. 76–1730). In addition, the typical clearly at high magnify.
diffraction peaks at 2θ = 27.64◦ and 30.4◦ , corresponding to the Moreover, elements mapping was performed to express the
diffraction of (310) and (222) planes of the β-Bi2O3, respectively (JCPDS elemental surface distribution in the cPVC/Bi2O3 composite. Fig. 4
No.74–1375). Also, the XRD patterns of cPVC/Bi2O3 composites existed shows the presence of the elements, including Bi, O, and C in the 2%
all typical peaks that characterized to Bi2O3 phase and cPVC materials. cPVC/Bi2O3 composite showing the coexistence of cPVC and Bi2O3 in

Fig. 1. FTIR spectra (a) and XRD patterns (b) of the cPVC/Bi2O3 composites.

3
H.-H. Pham et al. Sustainable Chemistry and Pharmacy 19 (2021) 100367

Fig. 2. Raman spectra of PVC, cPVC, Bi2O3, and 2% cPVC/ Bi2O3 composite.

the product after the synthesis process. may oxidize H2O and OH− to • OHradicals. However, the CB potential
The optical property of materials are measured through UV–vis DRS level is more favorable than the O2/• O−2 (– 0.33 V vs. NHE) potential that
to verify their optical characterization and the optical bandgap. Fig. 5 to partially produce O2 (Li et al., 2016). These results will be discussed
(a) shows a typical UV–vis DRS of cPVC, Bi2O3, and 2% cPVC/Bi2O3 more in the photocatalytic activity and photocatalytic mechanism of
composite. It can be seen that the absorption edge of cPVC at 650 nm cPVC/Bi2O3 composite.
range of wavelength while the absorption edge of Bi2O3 and 2% cPVC/ Firstly, we conduct the optimization of the content of cPVC for the
Bi2O3 composite is located at wavelength 470 nm and 475 nm, respec­ RhB photocatalytic degradation ability under visible light irradiation.
tively. The combination of cPVC and Bi2O3 did not significantly change Fig. 6 (a) shows the RhB degradation ability of cPVC/Bi2O3 composites
the shape and position of the absorption spectrum of the composite with different contents of cPVC. In the without catalyst case, RhB is
sample compared to pure Bi2O3. Also, the Tauc-plots (Eqn. (4)) (Tauc almost not degraded by visible light irradiation (blank sample) and this
et al., 1966) is calculated to determine the optical band gap of cPVC, has also occurred similarly for pure cPVC. The presence of cPVC in the
Bi2O3, and 2% cPVC/Bi2O3 composite with the values 2.0 eV, 2.4 eV, cPVC/Bi2O3 composites has improved significantly of the RhB photo­
and 2.3 eV, respectively (Fig. 5 (b)). This result indicated the excellent catalytic degradation. Detailly, the RhB photocatalytic degradation ef­
visible light response of 2% cPVC/Bi2O3 composite. ficiency achieved 56% and 59%, corresponding to 1% cPVC/Bi2O3
( )2 composite and the 2% cPVC/Bi2O3 composite. However, when the
αhν = A hν − Eg (4) content of cPVC increases from 3% wt to 7% wt, the RhB photocatalytic
degradation efficiency of these composites decreased significantly, even
where α, h, ν, A, and Eg are the absorption coefficient, Plank’s constant,
they are lower than that of the pure Bi2O3. This can be explained that the
frequency of the light, a continuous, and direct bandgap energy,
RhB decomposition occurred mainly on both the surface of Bi2O3 and
respectively.
cPVC. Besides, in our previous mention, cPVC contributes as a mem­
Determining the band structure of cPVC/Bi2O3 composite, Mott-
brane helping to accelerate the charge transfer (Bui et al., 2020).
Schottky plot is measured to determine the conduction band (CB) of
Increasing the content of cPVC will lead to covering the surface of the
Bi2O3, the lowest unoccupied molecular orbital (LUMO) of cPVC, and
Bi2O3 and reduce the efficiency of RhB decomposition. The L - H model
the CB of 2% cPVC/Bi2O3 composite (Fig. 5 (c)). The CB potential of
is used to fit the reaction kinetics of materials and provided for a more
Bi2O3 and the LUMO level of cPVC are 0.2 eV and 0 eV, respectively. The
quantitative comparison in Fig. 6 (b). Under visible light irradiation, 2%
CB potential position of 2% cPVC/Bi2O3 composite is at – 0.18 eV, and
cPVC/Bi2O3 composite is the highest reaction rate (k = 1.21 × 10− 4
its bandgap measured by DRS spectrum, which can illustrate in Fig. 5
min− 1), equaling 1.3 times, and 12.4 times over pure Bi2O3, and pure
(d). The valence band (VB) and the highest occupied molecule orbital
PVC, respectively. The order of the reaction rates was consistent with the
(HOMO) levels of Bi2O3 and cPVC are 2.63 and 2.02 eV, respectively.
results of the photocatalytic performance of cPVC/Bi2O3 composites that
The VB potential is near H2O/• OH(2.7 V vs. NHE (Li et al., 2016) that
were mentioned above. Also, the 2% cPVC/Bi2O3 composite has also

4
H.-H. Pham et al. Sustainable Chemistry and Pharmacy 19 (2021) 100367

Fig. 3. FESEM images of cPVC (a–b), Bi2O3 (c–d), and 2% cPVC/Bi2O3 composite (e–f).

shown the highest quantum efficiency with a 12 × 10− 4% valuation in concentration, the formation of the reactive radicals such as ⋅ SO−4
the weight of catalyst samples (Dcat = 0.33 g L− 1). The enhancement of and. OHis too much that will occur the reactions with HSO−5 to produce
the RhB photocatalytic degradation of 2% cPVC/Bi2O3 composites is the less oxidative ⋅ SO−5 radicals (Zhang et al., 2013) as Eqn. (5).
explained by the photocurrent response results in Fig. 6 (c). Results show / /
that the photocurrent response of cPVC, Bi2O3, and cPVC/Bi2O3 com­

SO−4 . OH + HSO−5 →⋅ SO−5 + SO2−4 + H+ /OH − (5)
posites indicated the stability and fast of the composite sample with 5 Or ⋅ SO−4 self-combination forming S2 O2−
8 (Eqn. (6)).
cycles and each extended cycle for 30 s of the on-off status. The
photocurrent density of 2% cPVC/Bi2O3 composite is much higher than ⋅
SO−4 ​ + ⋅ SO−4 →S2 O2−8 (6)
that of cPVC, Bi2O3, and other composites materials, which is consistent
Fig. 7 (b) shows that the highest photocatalytic reaction rate was
with the photocatalytic results of RhB decomposition in Fig. 6 (a) and
achieved with 3 mM PMS under visible light irradiation (k = 6.68 ×
(b). Therefore, we can conclude that the 2% cPVC/Bi2O3 material
10− 4 min− 1). In addition, the shape of the UV–Vis absorption spectra of
composite performance the best optical response among the rest of the
RhB is maintained during 150 min irradiation with the catalyst presence
composites.
(Fig. 7 (c)). This demonstrated the RhB complete degradation by 2%
The PMS concentration is changed from 0.1 to 5 mM to investigate its
cPVC/Bi2O3/PMS/Vis instead of forming intermediate complexes.
effect on the RhB degradation (200 mg L− 1) over 2% cPVC/Bi2O3
To determine the stability and reusability of the 2% cPVC/Bi2O3/
composite under visible light presented in Fig. 7. It can see clearly that
PMS/Vis system, the photocatalytic activity of RhB decomposition was
the RhB degradation ability of 2% cPVC/Bi2O3 composite increases
tested 4 times and shown in Fig. 8 (a). The RhB degradation ability of 2%
significantly when PMS concentration increases from 0.1 mM to 3 mM
cPVC/Bi2O3/PMS/Vis system after 4 cycles achieve 77% for 180 min
(Fig. 7 (a)). After that, when PMS concentration increase from 4 to 5
under visible light irradiation. This result indicated a good photo­
mM, the RhB degradation decreases gradually. The decrease of the RhB
catalytic efficiency of 2% cPVC/Bi2O3/PMS/Vis system. Moreover, the
degradation ability of 2% cPVC/Bi2O3 composite at high PMS concen­
FTIR spectra in Fig. 8 (b), the FTIR spectra show 2% cPVC/Bi2O3
trations (4 and 5 mM) can be explained as due to at high PMS

5
H.-H. Pham et al. Sustainable Chemistry and Pharmacy 19 (2021) 100367

Fig. 4. FESEM–EDS elemental mapping images of 2% cPVC/Bi2O3 (a–b), and elemental mapping with Bi (c), O (d), and C (e).

Fig. 5. UV–Vis DRS (a) of cPVC, Bi2O3, and 2% cPVC/Bi2O3 composite and their Tauc plots (b), Mott–Schottky plots (c), and the energy band alignment of ma­
terials (d).

composite before and after 4 cycles of the visible-light irradiation of the cm− 1, representing to Bi–O–Bi of Bi2O3 were not significantly different.
photocatalysis. The typical vibration peaks of Bi2O3 and cPVC did not This demonstrated that cPVC/Bi2O3 composite is a stable photocatalyst
change significantly. Detailly, the wavenumber at 1637 cm− 1, corre­ consistent with the color change in Fig. 8 (b). Therein, the typical yellow
sponding to the = C–H bond of cPVC, and the vibration peak at 844 color of the 2% cPVC/Bi2O3 composite was not almost changed

6
H.-H. Pham et al. Sustainable Chemistry and Pharmacy 19 (2021) 100367

Fig. 6. Photocatalytic activity of materials under visible light without the assistance of PMS (a), the fitted data for the reaction kinetic (b) with photocatalytic
reaction rate and quantum efficiency of cPVC, Bi2O3 and cPVC/Bi2O3 composites (c) photocurrent responses of cPVC, Bi2O3 and cPVC/Bi2O3 composites in Na2SO4
(0.5 M) aqueous solution. [RhB] = 20 mg L− 1; Dcat = 0.33 g L− 1.

compared to its original color. and holes (Eqn. (7)). The photogenerated electrons transfer from VB to
Fig. 9 (a) shows the effects of four different radical scavengers in the CB of Bi2O3 and they will also occur a transition from the HOMO level to
2% cPVC/Bi2O3/PMS/Vis system. With the addition of IPA, the RhB LUMO level of cPVC. Because the difference of the potential levels of CB
decomposition was thoroughly decreased, demonstrating that the• SO−4 of Bi2O3 and LUMO energy level position of cPVC, some electrons will
and • OHradicals are the primary factors for RhB degradation over cPVC/ migrate from LUMO of cPVC to the CB of Bi2O3. Herein, these electrons
Bi2O3/PMS/Vis system because IPA is a scavenger for trapping both will active PMS to generate • SO−4 or react dissolved oxygen molecules to

SO−4 and • OHradicals. Meanwhile, Fig. 9 (a) shows that with the form the • O−2 radicals (see Eqn. (8) and Eqn. (9)). Similarly, the holes
addition of KI and K2Cr2O7, trapping e− and h+ in the reaction process, transfer from the VB of Bi2O3 to HOMO level of cPVC and directly react
the RhB decomposition are less than the addition of IPA. This shows the with the hydroxyl anion to produce • OH radicals (Eqn. (11)), which will
role of h+ and e- in forming • SO−4 radicals (via Eqns. (8), (11) and (13)). oxidize pollutants due to its strong oxidizing capacity. Clearly, the
To compare the role of the • SO−4 and • OH, the TBA scavenger is added to photogenerated holes contribute in two processes, including oxidizing
investigate the RhB degradation. It is found that TBA scavenger RhB directly and forming • SO−5 radicals via their reaction with PMS
(• OHtrapping) is insignificantly decreased the RhB degradation effi­ (Eqn. (12)). Besides, • SO−4 anion radicals can form by other reaction
ciency, indicating the primary role of • SO−4 in decomposing RhB over pathways (Eqn. (14–16)) (Das, 2001) (Choe et al., 2020). The formation
cPVC/Bi2O3/PMS/Vis system. To clarify the presence of the active of • O−2 , • OH, and • SO−4 radicals is consistent with presence of their
radicals in the cPVC/Bi2O3/PMS system, the ESR analysis is an efficient signals in the ESR result (Fig. 9 (b)). As a result, • O−2 -, • OH, and
method to evaluate that. We conduct the ESR measurement to identify •
SO−4 contributed to excellent catalytic degradation activity in
the radicals such as superoxide radicals in methanol, hydroxyl and sul­ cPVC/Bi2O3/PMS with the aid of visible light (Eqn. (17)).
fate anion radicals in water (Fig. 9 (b)). Results show that the signals of

OH, SO−4 •, and • O−2 presences are barely clear at dark condition for the cPVC/ Bi2O3 + hν → cPVC/ Bi2O3 (e– + h+) (7)
both DMPO-H2O and DMPO-methanol. However, the photogenerated e + −
HSO−5 •
→ SO−4 + OH −
(8)
electrons will be formed and migrated to the CB of Bi2O3 and LUMO
level of cPVC while the photogenerated holes in the VB of Bi2O3 will e− + O2 →• O−2 (9)
transfer to HOMO level of cPVC. In the next step, the electrons will
active PMS to form sulfate anion radicals (Eqn. (8)). In addition, the •
O−2 + H2 O→• O H + H + (10)
photogenerated holes can also directly oxidize H2O, OH− to produce

OH radicals (Eqn. (11)). The • OHradicals also will contribute for the h+ + OH − →• O H + H + (11)
formation of • SO−4 (Eqn. (13)).
Based on the above results, the photocatalytic reaction mechanism h+ + HSO−5 →• SO−5 + H + (12)
for cPVC/Bi2O3/PMS/Vis system is proposed and presented in Fig. 10.
The excitation of visible light will form the photogenerated electrons HSO−5 + • O H→• SO−4 + OH − (13)

7
H.-H. Pham et al. Sustainable Chemistry and Pharmacy 19 (2021) 100367

Fig. 7. RhB degradation ability of 2% cPVC/Bi2O3/PMS/Vis system versus different PMS concentrations (a), the fitted data for the reaction kinetic rate (b), and
UV–Vis spectra of RhB solution with the presence of 2% cPVC/Bi2O3 composite under visible light. [RhB] = 200 mg L− 1, Dcat = 0.33 g L− 1.

Fig. 8. Photocatalytic ability of 2% cPVC/Bi2O3/PMS/Vis throughout cycle tests (a) and the FTIR spectra of a 2% cPVC/Bi2O3 composite before and after 4 cycles.

cPVC and Bi2O3 enhanced the visible–light photoactivity and created the

SO−4 + OH − →SO2−4 + • O H (14)
preferred redox potentials for the RhB photocatalytic degradation. The

SO−5 + • SO−5 →2• SO−4 + O2 (15) cPVC/Bi2O3/PMS/Vis system with 3 mM PMS degraded completely 200
mg L− 1 RhB solution after 150 min. Also, cPVC/Bi2O3/PMS/Vis system
expressed the stability and reusability after 4 cycles of the photocatalytic
SO2−4 + • O H + H + →• SO−4 + H2 O (16)
reaction.
/• /•

OH O−2 SO−4 + RhB→Degradation Products (17)
Author contributions

4. Conclusion
Pham Huu Huan: Methodology, Visualization, Writing-Original
Draft.
In this study, we have successfully synthesized cPVC/Bi2O3 com­
Cao Minh Thi: Validation, Data curation, Investigation.
posite by a friendly and straightforward method. The combination of
Thi Minh Cao: Resources, Formal analysis, Supervision.

8
H.-H. Pham et al. Sustainable Chemistry and Pharmacy 19 (2021) 100367

Fig. 9. The trapping test of 2% cPVC/Bi2O3/PMS/Vis (a), DMOP-ESR spectra in 2% cPVC/Bi2O3/PMS/Vis system of DMPO-• SO−4 and DMPO-• OH, and DMPO-
O2 (b).
• −

Fig. 10. The proposed mechanism of the photocatalytic reaction for cPVC/Bi2O3/PMS/Vis system.

Sheng-Jie You: Resources, Formal analysis. References


Ya-Fen Wang: Resources, Formal analysis.
Pham Van Viet: Chemical Conceptualization, Supervision, Funding Babuponnusami, A., Muthukumar, K., 2012. Advanced oxidation of phenol: a
comparison between Fenton, electro-Fenton, sono-electro-Fenton and photo-electro-
acquisition, Project administration. Fenton processes. Chem. Eng. J. 183, 1–9.
Balachandran, S., Swaminathan, M., 2012. Facile fabrication of heterostructured
Declaration of competing interest Bi2O3–ZnO photocatalyst and its enhanced photocatalytic activity. J. Phys. Chem. C
116, 26306–26312.
Bang, J.H., Choi, M.S., Mirzaei, A., Kwon, Y.J., Kim, S.S., Kim, T.W., Kim, H.W., 2018.
The authors declare that they have no known competing financial Selective NO2 sensor based on Bi2O3 branched SnO2 nanowires. Sensor. Actuator. B
interests or personal relationships that could have appeared to influence Chem. 274, 356–369.
Benitez, F.J., Acero, J.L., Real, F.J., Roldan, G., Casas, F., 2011. Comparison of different
the work reported in this paper. chemical oxidation treatments for the removal of selected pharmaceuticals in water
matrices. Chem. Eng. J. 168, 1149–1156.
Acknowledgements Bidoki, S.M., Wittlinger, R., 2010. Environmental and economical acceptance of
polyvinyl chloride (PVC) coating agents. J. Clean. Prod. 18, 219–225.
Bui, D.P., Pham, H., Cao, T., Pham, V., 2020. Preparation of conjugated polyvinyl
The authors would like to acknowledge the support from the Viet­ chloride/TiO2 nanotubes for Rhodamine B photocatalytic degradation under visible
nam National University Ho Chi Minh City and the CM Thi Laboratory. light. J. Appl. Chem. Biotechnol. 2707–2714.
This research is funded by Vietnam National University Ho Chi Minh Chen, X., Qiao, X., Wang, D., Lin, J., Chen, J., 2007. Kinetics of oxidative decolorization
and mineralization of Acid Orange 7 by dark and photoassisted Co2+-catalyzed
City (VNU-HCM) under grant number 562-2020-18-05. peroxymonosulfate system. Chemosphere 67, 802–808.

9
H.-H. Pham et al. Sustainable Chemistry and Pharmacy 19 (2021) 100367

Chi, F., Song, B., Yang, B., Lv, Y., Ran, S., Huo, Q., 2015. Activation of Nekouei, F., Nekouei, S., Noorizadeh, H., 2018. Enhanced adsorption and catalytic
peroxymonosulfate by BiFeO3 microspheres under visible light irradiation for oxidation of ciprofloxacin by an Ag/AgCl@N-doped activated carbon composite.
decomposition of organic pollutants. RSC Adv. 5, 67412–67417. J. Phys. Chem. Solid. 114, 36–44.
Choe, Y.J., Kim, J.-S., Kim, H., Kim, J., 2020. Open Ni site coupled with SO2- 4 Owen, E.D., Shah, M., Everall, N.J., Twigg, M.V., 1994. Raman spectroscopic study of the
functionality to prompt the radical interconversion of OH ↔ SO4− exploited to interaction of iodine with polyene sequences derived from the phase-transfer-
decompose refractory pollutants. Chem. Eng. J. 400. catalyzed dehydrochlorination of poly(vinyl chloride). Macromolecules 27,
Das, T.N., 2001. Reactivity and role of SO5•-radical in aqueous medium chain oxidation 3436–3438.
of sulfite to sulfate and atmospheric sulfuric acid generation. J. Phys. Chem. 105, Priya, B., Raizada, P., Singh, N., Thakur, P., Singh, P., 2016. Adsorptional photocatalytic
9142–9155. mineralization of oxytetracycline and ampicillin antibiotics using Bi2O3/BiOCl
Deng, Y., Zhao, R., 2015. Advanced oxidation processes (AOPs) in wastewater treatment. supported on graphene sand composite and chitosan. J. Colloid Interface Sci. 479,
Current Pollution Reports 1, 167–176. 271–283.
Ellahi, S., Hester, R.E., Williams, K.P.J., 1995. Waveguide resonance Raman Rogestedt, M., Hjertberg, T., 1993. Structure and degradation of commercial poly(vinyl
spectroscopy of degraded PVC. Spectrochim. Acta Mol. Biomol. Spectrosc. 51, chloride) obtained at different temperatures. Macromolecules 26, 60–64.
549–553. Solodovnichenko, V., Polyboyarov, V., Zhdanok, A., Arbuzov, A., Zapevalova, E.,
Elmolla, E.S., Chaudhuri, M., 2010. Comparison of different advanced oxidation Kryazhev, Y.G., Likholobov, V.J.P.e., 2016a. Synthesis of carbon materials by the
processes for treatment of antibiotic aqueous solution. Desalination 256, 43–47. short-term mechanochemical activation of polyvinyl chloride, 152, 747–752.
Golshan, M., Kakavandi, B., Ahmadi, M., Azizi, M., 2018. Photocatalytic activation of Solodovnichenko, V.S., Polyboyarov, V.A., Zhdanok, A.A., Arbuzov, A.B., Zapevalova, E.
peroxymonosulfate by TiO2 anchored on cupper ferrite (TiO2@CuFe2O4) into 2,4-D S., Kryazhev, Y.G., Likholobov, V.A., 2016b. Synthesis of carbon materials by the
degradation: process feasibility, mechanism and pathway. J. Hazard Mater. 359, short-term mechanochemical activation of polyvinyl chloride. Procedia Engineering
325–337. 152, 747–752.
Gómez-Cerezo, M.N., Muñoz-Batista, M.J., Tudela, D., Fernández-García, M., Starnes, W.H., Ge, X., 2004. Mechanism of autocatalysis in the thermal
Kubacka, A., 2014. Composite Bi2O3–TiO2 catalysts for toluene photo-degradation: dehydrochlorination of poly(vinyl chloride). Macromolecules 37, 352–359.
ultraviolet and visible light performances. Appl. Catal. B Environ. 156–157, Tauc, J., Grigorovici, R., Vancu, A., 1966. Optical properties and electronic structure of
307–313. amorphous germanium. Phys. Status Solidi 15, 627–637.
Ha, L.P.P., Huy, T.H., Huan, P.H., Thu, N.T.M., Thi, C.M., Van Viet, P., Tran, H.V., 2020. Truong, T.K., Van Doan, T., Tran, H.H., Van Le, H., Lam, V.Q., Tran, H.N., Cao, T.M., Van
Peroxymonosulfate activation on a hybrid material of conjugated PVC and TiO2 Pham, V., 2019. Effect of Cr doping on visible-light-driven photocatalytic activity of
nanotubes for enhancing degradation of rhodamine B under visible light. Adv. ZnO nanoparticles. J. Electron. Mater. 48, 7378–7388.
Polym. Technol. 2020, 1–9. Wang, C., Shao, C., Wang, L., Zhang, L., Li, X., Liu, Y., 2009. Electrospinning preparation,
He, F., Wang, J., Deng, D., 2011. Effect of Bi2O3 on structure and wetting studies of characterization and photocatalytic properties of Bi2O3 nanofibers. J. Colloid
Bi2O3–ZnO–B2O3 glasses. J. Alloys Compd. 509, 6332–6336. Interface Sci. 333, 242–248.
Hu, L., Deng, G., Lu, W., Lu, Y., Zhang, Y., 2017. Peroxymonosulfate activation by Wang, D., Sun, H., Luo, Q., Yang, X., Yin, R., 2014. An efficient visible-light photocatalyst
Mn3O4/metal-organic framework for degradation of refractory aqueous organic prepared from g-C3N4 and polyvinyl chloride. Appl. Catal. B Environ. 156–157,
pollutant rhodamine B. Chin. J. Catal. 38, 1360–1372. 323–330.
Irmawati, R., Noorfarizan Nasriah, M.N., Taufiq-Yap, Y.H., Abdul Hamid, S.B., 2004. Wang, L., Guo, X., Chen, Y., Ai, S., Ding, H., 2019. Cobalt-doped g-C3N4 as a
Characterization of bismuth oxide catalysts prepared from bismuth trinitrate heterogeneous catalyst for photo-assisted activation of peroxymonosulfate for the
pentahydrate: influence of bismuth concentration. Catal. Today 93–95, 701–709. degradation of organic contaminants. Appl. Surf. Sci. 467–468, 954–962.
Jung, Y.J., Kim, W.G., Yoon, Y., Kang, J.W., Hong, Y.M., Kim, H.W., 2012. Removal of Wang, S., Guan, Y., Wang, L., Zhao, W., He, H., Xiao, J., Yang, S., Sun, C., 2015.
amoxicillin by UV and UV/H2O2 processes. Sci. Total Environ. 420, 160–167. Fabrication of a novel bifunctional material of BiOI/Ag3VO4 with high
Karci, A., Arslan-Alaton, I., Olmez-Hanci, T., Bekbölet, M., 2012. Transformation of 2,4- adsorption–photocatalysis for efficient treatment of dye wastewater. Appl. Catal. B
dichlorophenol by H2O2/UV-C, Fenton and photo-Fenton processes: oxidation Environ. 168–169, 448–457.
products and toxicity evolution. J. Photochem. Photobiol. Chem. 230, 65–73. Wang, Y., He, Y., Li, T., Cai, J., Luo, M., Zhao, L., 2012. Photocatalytic degradation of
Leonard, N.M., Wieland, L.C., Mohan, R.S., 2002. Applications of bismuth(III) methylene blue on CaBi6O10/Bi2O3 composites under visible light. Chem. Eng. J.
compounds in organic synthesis. Tetrahedron 58, 8373–8397. 189–190, 473–481.
Li, X., Yu, J., Jaroniec, M., 2016. Hierarchical photocatalysts. Chem. Soc. Rev. 45, Xiong, Y., Wu, M., Ye, J., Chen, Q., 2008. Synthesis and luminescence properties of hand-
2603–2636. like α-Bi2O3 microcrystals. Mater. Lett. 62, 1165–1168.
Lin, K.-Y.A., Chen, B.-J., Chen, C.-K., 2016. Evaluating Prussian blue analogues MII3[MIII Yan, Y., Zhou, Z., Cheng, Y., Qiu, L., Gao, C., Zhou, J., 2014. Template-free fabrication of
(CN)6]2 (MII = Co, Cu, Fe, Mn, Ni; MIII = Co, Fe) as activators for α- and β-Bi2O3 hollow spheres and their visible light photocatalytic activity for
peroxymonosulfate in water. RSC Adv. 6, 92923–92933. water purification. J. Alloys Compd. 605, 102–108.
Lin, K.-Y.A., Chen, Y.-C., Lin, Y.-F., 2017. LaMO3 perovskites (M=Co, Cu, Fe and Ni) as Zhang, T., Zhu, H., Croue, J.P., 2013. Production of sulfate radical from
heterogeneous catalysts for activating peroxymonosulfate in water. Chem. Eng. Sci. peroxymonosulfate induced by a magnetically separable CuFe2O4 spinel in water:
160, 96–105. efficiency, stability, and mechanism. Environ. Sci. Technol. 47, 2784–2791.
Méndez-Díaz, J., Sánchez-Polo, M., Rivera-Utrilla, J., Canonica, S., von Gunten, U., 2010. Zhang, Y., Zhang, F., Yang, Z., Xue, H., Dionysiou, D.D., 2016. Development of a new
Advanced oxidation of the surfactant SDBS by means of hydroxyl and sulphate efficient visible-light-driven photocatalyst from SnS2 and polyvinyl chloride.
radicals. Chem. Eng. J. 163, 300–306. J. Catal. 344, 692–700.
Michell, E.W.J., 1985. True stabilization: the behaviour of lead compounds against the Zhao, J., Zhang, Y., Quan, X., Chen, S., 2010. Enhanced oxidation of 4-chlorophenol
thermal decomposition of polyvinyl chloride. J. Mater. Sci. 20, 3816–3830. using sulfate radicals generated from zero-valent iron and peroxydisulfate at
ambient temperature. Separ. Purif. Technol. 71, 302–307.

10

You might also like