Phân Hủy Kháng Sinh Oxytetracycline Bằng Chất Xúc Tác Quang TiO2 Tái Tạo Bề Mặt
Phân Hủy Kháng Sinh Oxytetracycline Bằng Chất Xúc Tác Quang TiO2 Tái Tạo Bề Mặt
A R T I C L E I N F O A B S T R A C T
Keywords: In this study, we successfully developed a photocatalyst that can respond to visible light and exhibit excellent
Oxytetracycline activity by doping terbium and copper simultaneously into TiO2 powder using the liquid phase plasma method.
Tb/Cu doped TiO2 photocatalyst When Tb/Cu doped TiO2 photocatalyst (TCTP) was used as a visible light source, the degradation rate increased
Liquid-phase plasma: Oxygen vacancies
by up to three times compared with that of bare TiO2. In the case of co-doped TiO2, an apparent synergistic effect
Photocatalyst reusability
of co-doping was confirmed compared to the single-component doping of Tb or Cu. The absorption edge moved
to the visible light region owing to simultaneous doping with Tb and Cu, and the greatest bandgap energy
reduction was achieved. In addition, Tb and Cu doping induced more oxygen vacancies, reducing the electron-
hole recombination rate by forming a charge imbalance. This is believed to improve photocatalytic activity by
adsorbing more oxygen molecules and inducing the formation of active radicals. The optimal Tb-Cu doping ratio
in this study was confirmed to be 1:1, and it can be used as a potential photocatalyst that can be activated by all
light sources, including UV and visible light.
1. Introduction extensively studied because they can decompose and remove various
types of organic chemical pollutants, are non-toxic and harmless to the
Various types of antibiotics are used to prevent and treat diseases in natural environment, and have high economic efficiency. Although TiO2
humans and animals. These antibiotics cannot be completely absorbed photocatalysts have various applications in pollutant destruction, their
or decomposed in the body and are excreted through urine and feces into efficiency is limited by two factors. First, owing to its high bandgap
aquatic ecosystems such as surface water, groundwater, and wastewater energy (3.1 eV for P-25), photocatalysis can be efficiently applied only in
[1,2]. Tetracycline antibiotics are widely used worldwide because they the UV region of the solar spectrum. Second, the recombination of
are effective against both Gram-positive and Gram-negative bacteria. electron-hole pairs occurs during the photocatalysis of TiO2 [4]. To
Different types of tetracycline (TC) and oxytetracycline (OTC) exist. overcome these limitations, the light absorption area is expanded, and
Most antibiotics are non-degradable; therefore, complete decomposition the efficiency is promoted through the optical modification of TiO2.
through wastewater treatment facilities, such as physicochemical or Various studies have been conducted, including dye sensitization,
biological treatments, is difficult. Additionally, only 4.3 to 72 % of TC combining carbonaceous materials, and doping with metals, non-metals,
and less than 50 % of OTC are processed, and the rest is discharged. metals/non-metals, and rare-earth elements [4].
Tetracycline antibiotics are detected in various places at concentrations In particular, methods for doping TiO2 with dopants are being
of approximately 0.1 to 460 µg/L [3]. Antibiotics that remain in aquatic studied in diverse ways through various changes to the dopant. Several
ecosystems owing to their non-degradable properties are potentially methods for TiO2 doping include the sol–gel method, electrophoretic
harmful to the ecosystem and humans, as they contribute to occurrence deposition, sputtering deposition, microemulsion calcination, and me
of resistant bacteria and toxicity. Therefore, various advanced oxidation chanical mixing [5]. These methods have disadvantages such as high
processes (AOPs) have been studied and applied for the treatment of manufacturing costs, long reaction times, high-temperature treatments,
antibiotics. toxic organic solvent use, and volatile toxic substances [4,6]. The liquid
Among them, the method using TiO2 photocatalysts has been phase plasma (LPP) process is a recently introduced method to address
* Corresponding author.
E-mail address: [email protected] (S.-C. Jung).
https://doi.org/10.1016/j.apsusc.2024.162244
Received 9 September 2024; Received in revised form 17 November 2024; Accepted 27 December 2024
Available online 28 December 2024
0169-4332/© 2024 Elsevier B.V. All rights are reserved, including those for text and data mining, AI training, and similar technologies.
C.-S. You et al. Applied Surface Science 687 (2025) 162244
Table 1 method. Copper and terbium precursors were copper chloride dihydrate
Various initial precursor concentrations and LPP process conditions for the (CuCl2⋅2H2O, Junsei Chemical Co.) and terbium (III) chloride hexahy
production of Tb/Cu-doped TiO2 photocatalyst. drate (TbCl3⋅6H2O, Sigma Aldrich Co.), which were prepared by dis
Samples Initial precursor concentration LPP process condition solving in deionized water (Daejung Chemicals & Metals Co., Ltd.)
(mM) without any additional treatment and were used as aqueous reactants.
Tb(Cl)3 Cu(Cl)2 The photocatalytic activity of the modified TiO2 photocatalysts was
Bare TiO2 0 0 • Power evaluated for the decomposition characteristics of oxytetracycline hy
Tb2.0-TiO2 1.5 0 • voltage (250 V) drochloride (OTC, C22H24N2O9⋅HCl, Tokyo Chemical Industry Co.).
Tb1.5Cu0.5-TiO2 1 0.5 • frequency (30 kHz)
Tb1.0Cu1.0-TiO2 0.75 0.75 • pulse width (5 μs))
Tb0.5Cu1.5-TiO2 0.5 1 • LPP reaction time: 1 hr 2.2. Method
Cu2.0-TiO2 0 1.5
Fig. 1. X-ray diffraction pattern of bare TiO2 and Tb/Cu doped TiO2 photocatalys.
2
C.-S. You et al. Applied Surface Science 687 (2025) 162244
Table 2 anatase (JCPDS card no. 21–1272) and rutile (JCPDS card no. 21–1276)
Crystalline Size of Tb/Cu- doped TiO2. (110) peak which was observed [11,12]. No additional impurity peaks
Sample DA(101) (nm) DR(110) (nm) due to the dopant were observed in the XRD patterns of the prepared
TCTPs, indicating that Tb and Cu occupied major positions in the TiO2
Bare TiO2 18.89 24.52
Tb2.0-TiO2 19.11 25.07 lattice [13]. As a result of the enlarged analysis of the anatase crystal
Tb1.5Cu0.5-TiO2 19.49 25.18 (101 peak) and the rutile crystal (110 peak), bare TiO2 was 25.26◦ and
Tb1.0Cu1.0-TiO2 19.69 25.50 27.37◦ at 2θ. Meanwhile, Tb1.0Cu1.0-doped TiO2 prepared through the
Tb0.5Cu1.5-TiO2 19.92 25.48 LPP process moved 0.04◦ and 0.05◦ to the right to 25.30◦ and 27.47◦ .
Cu2.0-TiO2 21.36 25.24
The diffraction peaks are believed to be displaced by stress or strain
effects [14]. Using the Scherrer equation from the XRD patterns, the
centrifugation and a syringe filter. Samples of the reactant solution from average crystal sizes related to the anatase (1 0 1) and rutile (1 1 0) peaks
which the TCTP powder was removed were analyzed at the maximum were calculated using the following equation:
absorption wavelength of OTC (355 nm) using a UV–vis spectrometer to Kλ
evaluate the decomposition efficiency. The reactive species generated Crystallite size, D = (1)
βcosθ
during the photodegradation process were analyzed by optical emission
spectroscopy (OES) and isopropyl alcohol (IPA, 500 mM), benzoquinone where λ is 1.5406 (Cu-source), β is FWHM (full width at half maxima,
(BQ, 2 mM), ethylenediaminetetraacetic acid (EDTA-2Na, 100 mM), and measured in radians), K is the shape coefficient of 0.94, and θ is the
methanol (Me-OH, 100 mM) were used as scavengers to investigate the Braggs angle of the diffraction peak [14,15]. Crystallite size calculation
radicals. results are shown in Table 2, which confirmed that the crystal size
increased owing to doping. In particular, when comparing the size of DA
3. Results and discussion (101) with bare TiO2, Tb2.0-TiO2, and Cu2.0-TiO2, the crystallite size
increased by 0.22 nm for Tb doping and 2.48 nm for Cu doping. Thus, Cu
3.1. Characterization of TCTP doping had a considerable effect on the crystallite size.
Fig. 2. (a) The actual image of Tb1.0Cu1.0-doped TiO2 and the mapping image of each element and the image observed with a field emission energy filtering
transmission electron microscope(b-g).
Table 3
Chemical composition of Bare TiO2 and Tb1.0Cu1.0-doped TiO2 prepared by LPP method.
Sample Initial precursor Concentration (mM) Titanium Oxygen Terbium Copper
Tb(Cl)3 Cu(Cl)2 Wt. % At. % Wt. % At. % Wt. % At. % Wt. % At. %
Bare TiO2 0 0 58.66 32.16 41.34 67.84 0.00 0.00 0.00 0.00
Tb1.0Cu1.0-doped TiO2 1.0 1.0 56.91 31.37 41.27 68.11 0.94 0.16 0.87 0.36
3
C.-S. You et al. Applied Surface Science 687 (2025) 162244
Fig. 3. Photoluminescence spectrum of bare TiO2 and Tb/Cu doped TiO2 photocatalys.
the reaction solution that dissociated from the precursor by the LPP plane (ICDD: 00–001-0562), 3.24 Å TiO2 rutile (1 1 0) plane (ICDD:
reaction were evenly doped on the TiO2 surface. Here, the numbers after 00–001-1292), 2.51 Å CuO( − 1 1 1) plane (ICDD: 00–001-1117), 2.19 Å
Tb and Cu represent the mM concentrations of the precursor when Tb2O3 (422) plane (ICDD: 00–023-1418), 1.84 Å TbO2 (220) plane
producing TiO2 via the LPP process. Table 3 shows the chemical (ICDD: 01–075-0209), and 1.74 Å Cu2O (211) plane (ICDD: 03–065-
composition of Tb1.0Cu1.0-doped TiO2 prepared at a concentration of 1 3288) were confirmed.
mM Tb and 1 mM Cu, as analyzed using Energy Dispersive X-ray Spec
troscopy (EDS). The atomic percentage ratio was doped with Tb 0.16 % 3.1.3. Photoluminescence spectra analysis
and Cu 0.36 %, which are believed to be doped on the TiO2 surface, as Tb Fig. 3 shows the photoluminescence (PL) spectra of bare TiO2 and
and Cu ions are reduced by chemically active species generated during TCTPs measured at an excitation wavelength (350 nm) within the
the LPP process [16]. measurement range of 380–630 nm. Photoluminescence emission is
For a detailed structural analysis of Tb1.0Cu1.0-doped TiO2, the FE- derived from the recombination of holes and excited electrons, and the
EF-TEM (Field Emission Energy Filtering Transmission Electron Micro PL spectrum shows the separation of photoinduced pairs. Therefore, it
scope) measurement results are shown in Fig. 2(b)–(f); (b) and (c) show can be used to identify the charges separated by the dopant (Tb and Cu)
TEM images, and (d), (e), and (f) show enlarged lattice images. (d) ions doped into TiO2 [17,18]. When comparing bare TiO2 and TCTP in
shows the lattice spacings of the TiO2 anatase (1 0 1) plane (ICDD the PL spectrum shown in Fig. 3, the intensity of bare TiO2 was higher
00–001-0562), (e) the (422) plane (ICDD 00–023-1418) of Tb2O3, and and that of TCTP was significantly reduced. The spectrum of the Cu-
(f) the ( − 1 1 1) plane (ICDD 00–001-1117) of CuO, which were 0.352, doped TiO2 changed only in intensity, and a new emission peak was
0.219, and 0.251, respectively. The Selected Area Electron Diffraction not generated in the same way as in the results of previous Cu-doped
(SAED) pattern analysis results to confirm the Tb and Cu planes that TiO2 manufacturing studies [18]. In the PL spectra of Tb-doped TiO2
were not identified in the XRD pattern are shown in Fig. 2(g). By and TCTP, new emission peaks were observed at 490, 544, 587, and 622
comparing the observed discontinuous concentric diffraction ring with nm, which are believed to be due to Tb doping. This is due to the typical
d values for each plane of the ICDD card, 3.52 Å TiO2 anatase (1 0 1) radiative electronic transitions of doped Tb3+ and are the peaks
Fig. 4. (a) UV–Vis absorbance spectra, (b) band gap energy of bare TiO2 and Tb/Cu doped TiO2 photocatalys.
4
C.-S. You et al. Applied Surface Science 687 (2025) 162244
Fig. 5. XPS spectrum of Tb1.0Cu1.0-doped TiO2 prepared using the liquid phase plasma method.
generated during the electric dipole transitions from 5D4 to 7F6, 7F5, 7F4, photocatalyst utilization under visible light.
and 7F3 [19]. The peak appearing at 390 nm is a peak due to the emission
of the bandgap transition corresponding to the bandgap energy of TiO2 3.1.5. X-ray photoelectron spectroscopy analysis
anatase. Oxygen vacancies typically cause electron trapping dynamics, X-ray photoelectron spectroscopy (XPS) characterization was per
resulting in a decrease in peak intensity or an unobservable state. In formed to clarify the chemical composition and chemical bonding state
TCTPs, the intensity at 390 nm was reduced or not observed, which is of TCTP prepared via the LPP process. Fig. 5 shows the analysis results
thought to be due to increased oxygen vacancies [20]. As observed from for the bare TiO2 and Tb1.0Cu1.0-doped TiO2. Fig. 5(a) shows the Ti2p
the TEM-SAED results, the decrease in the intensity of TCTP was due to region spectrum, with the main peak of Ti4+ oxidation state at 458.6
the formation of a charge imbalance in the combination of Ti-O-Tb and (Ti2p3/2) eV and 464.3 (Ti2p1/2) eV for bare-TiO2, and the peak of 457.6
Ti-O-Cu, which were uniformly distributed on the TiO2 surface and in (Ti2p3/2) eV for Ti3+ oxidation state; Ti-C peak was detected at 459.8 eV.
the lattice gap. To rectify this, hydroxyl groups are adsorbed on the TiO2 In the Ti2p region spectrum of Tb1.0Cu1.0-doped TiO2, two major peaks
surface, facilitating oxidation. Thus, the photogenerated electrons and corresponding to the Ti4+ oxidation state of Ti2p3/2 and Ti2p1/2
holes effectively improve the separation state or suppress recombina appeared at binding energies of 458.9 and 464.7 eV, respectively, and
tion, thereby improving the photocatalyst performance [21,22]. the Ti3+ state at 457.9 and 463.6 eV peaks were detected. In particular,
the Ti3+ peak, which was not detected for bare TiO2, was detected at
3.1.4. UV–vis diffuse reflectance spectra analysis 463.6 eV [25,26]. The energy separation of Ti2p1/2 and Ti2p3/2 in the
Fig. 4 shows the UV-DRS analysis results of bare TiO2 and TCTPs. Ti4 + oxidation state was 5.7 eV, which agrees with the P-25 peak re
Fig. 4(a) shows the UV–vis diffuse reflectance absorption spectra, indi ported in previous studies [27–29]. Comparing the positions of the
cating that the absorption of TCTPs from 400 to 700 nm was evidently detected peaks of bare-TiO2 and Tb1.0Cu1.0-doped TiO2, the positions
higher than that of bare TiO2. This indicates that doping creates a strong of the peaks shifted by approximately 0.3 eV, possibly because some Ti
absorption ability in the visible and UV light regions, expanding the light ions were replaced by dopant ions in the lattice [25].
response range to the visible light region [21]. Fig. 4(b) shows the Fig. 5(b) shows the O1s region spectrum. In bare TiO2, only two
bandgap energy (BGE) calculated using the Kubelka–Munk equation. peaks were detected: 530.3 eV of the Ti-O bond, which is lattice oxygen,
Bare TiO2 showed a value of 3.10 eV, which is similar to that reported in and 532.1 eV, a surface contamination peak such as OH. Tb1.0Cu1.0-
previous studies [23,24]. The BGE of TCTP was Tb2.0-doped TiO2 (2.89 doped TiO2 prepared by the LPP process was additionally confirmed to
eV), Tb1.5Cu0.5-doped TiO2 (2.95 eV), Tb1.0Cu1.0-doped TiO2, have a new peak, a chemisorbed oxygen peak of 533.8 eV [30]. Because
Tb0.5Cu1.5-doped TiO2 (2.86 eV), and Cu2.0-doped TiO2 (2.93 eV). of doping, the Ti-O-Tb bond appeared at the lattice oxygen peak, and the
Compared with bare TiO2, the BGE decreased in all cases owing to Ti-O-Cu bond appeared at the same peak as the surface contamination
doping. A maximum reduction of 0.24 eV was observed compared to peak. Meanwhile, as in the Ti2p spectrum, a peak shift of 0.2 to 0.4 eV
bare TiO2. This result shows that Tb and Cu doping reduces the distance was observed due to doping [31,32].
between the valence and conduction bands, thereby improving Fig. 5(c) shows the high-resolution XPS spectrum of the Tb3d region,
5
C.-S. You et al. Applied Surface Science 687 (2025) 162244
where Tb2O3, Tb3d5/2 (1242.7 eV), and Tb3d3/2 (1274.0 and 1277.4
eV), were observed. Additionally, a peak corresponding to Tb3+ in the
TbO1.5 state was observed. Furthermore, the Tb4+-related peaks by the
TbO2 and TbO1.82 states were confirmed at 1244.6 (Tb3d5/2) and 1281.7
(Tb3d3/2) eV [33–35]. These results were confirmed by TEM analysis
and suggested that the doped Tb exists in one or more oxidation states:
Tb3+ and Tb4+.
As shown in Fig. 5(d), Cu2P3/2 was observed at a binding energy of
933.0 eV, and Cu2P1/2 at a binding energy of 952.7 eV. These peaks
were caused by Cu+, indicating that Cu existed as Cu2O but not as CuO
[36]. These values are consistent with the Cu2O values reported in
previous studies [37,38]. Although two forms of Cu2O and CuO were
observed in the TEM SAED pattern, only the peak of Cu2O was observed
in the XPS analysis. This is considered to be detected only in the form of
Cu+, which exists in a relatively high form owing to the low peak height
resulting from the small amount of Cu doped on the TiO2 surface [39].
Colón reported that Cu+ has a better effect on photocatalytic activity
than Cu2+ and acts as a photoactive species that avoids the recombi
nation of electron holes [40]. These results are expected to positively
affect the photocatalytic activity of TCTP doped in Cu+ through the LPP
process.
Fig. 6. Changes in the ratio of the peak areas of Ti4+, Ti3+, Ti-C and oxygen Fig. 6 shows the change in the ratio of peak areas of Ti4+, Ti3+, Ti-C,
vacancies in Ti2p, O1s during the doping process through the LPP process.
and oxygen vacancies in Ti2p and O1s during the doping process via the
LPP process. By comparing three types of TiO2, namely bare TiO2, Tb2.0-
doped TiO2, and Tb1.0Cu1.0-doped TiO2, the effects of doping and co-
Fig. 7. Oxytetracycline decomposition behavior of bare TiO2 and Tb/Cu doped TiO2 under UV LED light and blue LED light.
6
C.-S. You et al. Applied Surface Science 687 (2025) 162244
Fig. 8. (a) cycling runs in the photocatalytic degradation of oxytetracycline under UV light irradiation, (b) XRD patterns of Tb1.0Cu1.0-TiO2 before and after pho
tocatalytic reaction.
doping were also compared. The ratio of Ti4+ oxidation state decreased results of the decomposition reaction of OTC using TCTPs and bare TiO2
from 90.4 % to 71.4 % and 81.5 %, respectively, by 19.0 % and 8.9 %, as photocatalysts are shown in Fig. 7(a) and (b) under UV LED light
but it accounted for most of the oxidation states. This indicates that Ti conditions and Fig. 7(c) and (d) under blue LED light conditions. During
ions in the TiO2 lattice mainly exist in the form of Ti4+ even when the a 90 min reaction time under UV LED light conditions, 95.0 % of bare
doping process proceeds through the LPP process. The decrease in the TiO2, 87.8 % of Tb2.0-doped TiO2, 91.0 % of Tb1.5Cu0.5-doped TiO2,
area of Ti4+ and the increase in the area ratio of Ti3+ by 4.4 and 2.4 95.3 % of Tb1.0Cu1.0-doped TiO2, 82.7 % of Tb0.5Cu1.5-doped TiO2, and
times, respectively, indicate that after the LPP process, Ti ions were 80.0 % of Cu2.0-doped TiO2 decomposed. The decomposition reaction
substituted and replaced by Tb or Cu ions, or a mixed oxide structure rate constant K during the first 30 min was 3.22 × 10-2 min− 1 for bare
with an oxidation state of Ti3+ was formed by the formation of Ti-O-Tb TiO2, 4.98 × 10-2 min− 1 for Tb2.0-doped TiO2, 5.12 × 10-2 min− 1 for
structure in the TiO2 lattice, or the formation of Ti2O3 [25]. In the Tb1.5Cu0.5-doped TiO2, 5.41 × 10-2 min− 1 for Tb1.0Cu1.0-doped TiO2,
change in the peak area ratio of oxygen vacancies, the area of lattice 3.88 × 10-2 min− 1 for Tb0.5Cu1.5-doped TiO2, and 3.44 × 10-2 min− 1 for
oxygen was significantly reduced from 92.4 % in the doping process Cu2.0-doped TiO2. The decomposition reaction rate constant, K,
using the LPP process to 80.4 % for Tb2.0-doped TiO2, which was a 12 % increased from a minimum of 1.07 times to a maximum of 1.68 times
decrease, and to 66.3 % for Tb1.0Cu1.0-doped TiO2, which was a 26 % through doping, possibly because the photocatalytic performance
decrease. It can be seen that the areas of OH peak and chemisorbed improved owing to the effects of Tb and Cu doping, consistent with the
oxygen peak increased and were newly formed. In addition, an increase results predicted from the PL analysis. Under blue LED light conditions,
in surface oxygen species was confirmed, which is thought to be because during the 90 min decomposition reaction time, bare TiO2 was reduced
the dopants, Tb and Cu, substituted Ti ions or formed oxide structures by 14.6 %, Tb2.0-doped TiO2 by 37.5 %, Tb1.5Cu0.5-doped TiO2 by 40.6
such as Cu2O, CuO, TbO2, and Tb2O3. In particular, in the case of co- %, Tb1.0Cu1.0-doped TiO2 by 43.6 %, Tb0.5Cu1.5-doped TiO2 by 36.4 %,
doping, the decrease was 2.2 times larger, suggesting that more O va and Cu2.0-doped TiO2 by 20.6 %. The decomposition reaction could be
cancies can be formed through co-doping than through single-element activated through doping, and when Tb1.0Cu1.0-doped TiO2 was used,
doping. By confirming the change in peak area of the XPS spectrum, it the decomposition rate increased up to 3 times compared to that of bare
was confirmed that the band gap reduction and red shift in Tb and Cu TiO2. The decomposition reaction rate constant K for 90 min also
doping using the LPP process were due to the O vacancy and Ti3 + improved 3.5 times from 1.81 × 10-3 min− 1 to 6.38 × 10-3 min− 1.
formation [41]. The decomposition rate of OTC by TCTP manufactured with LPP was
considerably improved compared to that of bare TiO2 under all UV and
blue light sources. The improvement in the decomposition rate of TCTP
3.2. Photocatalytic activity of TCTP is due to the weakening of the PL emission compared to bare-TiO2 due to
the Tb and Cu dopants doped on the TiO2 surface, as confirmed by the PL
This study performed a degradation experiment to evaluate the analysis. Therefore, as the separation of electron holes increases and
photocatalytic activity of bare TiO2 and TCTPs for the antibiotic recombination is suppressed, the generation of oxidants such as hy
oxytetracycline hydrochloride (OTC). The calculation of the decompo droxyl radicals can be increased, and the BGE can be reduced, thereby
sition reaction rate was assumed to follow a pseudo-first-order decom strengthening the photoelectric effect [42]. For co-doped TiO2, an
position reaction, and the decomposition reaction results during the first apparent synergistic effect of co-doping was confirmed compared to the
30 min of the reaction under UV light (λmax = 375 nm) conditions and single-component doping of Tb or Cu. Cu-doped TiO2 showed a signif
the first 90 min of the reaction under blue light (λmax = 465 nm) icant decrease in the PL emission peak, while the decrease in the BGE
conditions were applied to the calculation. was the smallest. For Tb doping, the PL emission peak was reduced to a
Fig. 7 shows the results of the OTC degradation experiment using smaller extent than that of Cu, while the bandgap energy was further
TCTPs and bare TiO2 prepared by the LPP process. To test the photo reduced. For TCTP, the absorption edge moved to the visible light region
degradation resistance of OTC, its decomposition efficiency was inves owing to the effects of Tb and Cu doping, and the largest BGE reduction
tigated during light irradiation for 90 min under blue and UV LED light. was achieved. Additionally, as shown in TEM and XPS analyses, Tb and
Consequently, 2.9 % decomposed under UV LED light (Fig. 7(a)), and Cu doping replace the Ti4+ oxidation state with Tb3+ or Cu+ oxidation
only 0.7 % decomposed under blue LED light (Fig. 7(c)). Therefore, the states, which causes oxygen vacancies and charge imbalances. It is
antibiotic OTC barely decomposed after irradiation with light alone. The
7
C.-S. You et al. Applied Surface Science 687 (2025) 162244
Fig. 9. (a) OES spectrum in each solution condition, (b) change in intensity of radicals for change in TCTPs.
8
C.-S. You et al. Applied Surface Science 687 (2025) 162244
Fig. 11. Predicted pathway for the photocatalytic decomposition reaction of oxytetracycline using Tb1.0Cu1.0-doped TiO2.
best. 147, 139, 111, 83, and 61, which are ultimately decomposed into CO2,
Using Scavenger, the results of indirectly confirming the active H2O, and N-minerals.
species that has the greatest effect on OTC photodecomposition among
the active species are shown in Fig. 10. Isopropyl alcohol (IPA), ben 4. Conclusion
zoquinone (BQ), EDTA-Na, and methanol (Me-OH) were used as scav
engers. IPA was added to remove hydroxyl radicals (•OH), BQ to remove A photocatalyst capable of responding to visible light and exhibiting
superoxide radicals (•O–2), EDTA-Na to remove holes (h+), and Me-OH to excellent activity was manufactured by doping terbium and copper
remove electrons (e-), and then photodecomposition was performed to simultaneously with TiO2 powder using the liquid phase plasma method.
confirm their respective effects [48,49]. As shown in the results in XRD analysis revealed that the crystallite size increased owing to
Fig. 10, it was confirmed that hole (h+) had the greatest effect on the doping, and Cu doping had a greater effect on the crystallite size. TEM
decomposition of OTC using TiO2 photocatalyst and was the main active analysis showed that Tb and Cu were uniformly doped on the TiO2
species. Changes due to Tb and Cu doping were confirmed by the change surface, and SAED pattern analysis confirmed the Tb2O3, TbO2, CuO,
in the ratio of OTC decomposition reaction rate constants when using and Cu2O planes. In addition, it was assumed that a Ti-O-Tb structure
bare TiO2 and Tb1.0Cu1.0-doped TiO2 according to each scavenger in was formed within the TiO2 lattice and that a mixed oxide structure with
jection. When ISO was used as an •OH scavenger, the decomposition the oxidation state of Ti3+ was formed by the formation of Ti2O3. Cu-
reaction rate constant of bare TiO2 decreased by 5.4 % compared to doped TiO2 exhibited a significant decrease in the PL emission peak,
before addition, whereas it decreased by 26.9 % when Tb1.0Cu1.0-doped whereas the decrease in the bandgap energy was the smallest. For Tb-
TiO2 was used. Through this, it was confirmed that when doping is doping, the PL emission peak was reduced to a smaller extent than
performed, •OH is formed in greater amounts than bare TiO2, thereby that of Cu, whereas the BGE was further reduced. It was assumed that
improving the decomposition efficiency. the band gap reduction and red shift owing to the simultaneous doping
of Tb and Cu by LPP were due to O vacancies and Ti3+ formation. When
3.5. Degradation pathways Tb1.0Cu1.0-doped TiO2 prepared by doping with LPP was used as a
visible light source, the decomposition rate increased up to three times
To better understand the degradation process of OTC using TCTP compared to that of bare TiO2. The improvement in the decomposition
photocatalyst and to detect intermediates and investigate the reaction rate of TCTP increased the generation of oxidants, such as hydroxyl
pathways generated during the degradation process, LC-MS/MS analysis radicals, by increasing the separation of electron holes and suppressing
was performed, and the expected intermediates and degradation path recombination by simultaneous doping with Tb and Cu, which was
ways are shown in Table S3 and Fig. 11. The decomposition experiment strengthened by reducing the band gap energy. TCTP stability and
was conducted by injecting 0.1 g of Tb1.0Cu1.0-doped TiO2 into a 5 ppm reusability tests confirmed that TCTP is a reusable and stable photo
OTC solution and conducting photodecomposition under UV irradiation. catalytic material.
The decomposition pathway of OTC can be largely divided into two
types as shown in Fig. 11, and the main decomposition processes are CRediT authorship contribution statement
estimated to be hydroxylation, ring-opening reaction, demethylation,
and dehydration. Pathway I is a process in which the double bond at Chan-Seo You: Writing – original draft, Investigation. Jun-Young
C11a-C12 positions reacts with ⋅OH radicals through hydroxylation to Noh: Investigation, Data curation. Yunju Choi: Formal analysis. Sang-
generate a primary intermediate with m/z = 477, then the acylamino Chul Jung: Writing – review & editing, Supervision.
group, carbinol groups, and methyl groups at C2 position are lost to
generate m/z 340, and m/z 279 is generated through the removal of Declaration of competing interest
methyl groups at C6 position and the opening of benzene rings [50,51].
Pathway II forms m/z 358 through dehydration and demethylation of The authors declare that they have no known competing financial
C6, m/z 227 through deamination and opening of benzene rings, and m/ interests or personal relationships that could have appeared to influence
z 208 through removal of methyl groups in an additional reaction [52]. the work reported in this paper.
Subsequent additional reactions include ring-opening reaction and
oxidation by active species such as ⋅OH to generate substances with m/z
9
C.-S. You et al. Applied Surface Science 687 (2025) 162244
Acknowledgment [19] V.S. Smitha, P. Saju, U.S. Hareesh, G. Swapankumar, K.G.K. Warrier, Optical
Properties of Rare-Earth Doped TiO2 Nanocomposites and Coatings; A Sol-Gel
Strategy towards Multi–functionality, ChemistrySelect. 1 (2016) 2140–2147,
This work was supported by the National Research Foundation of https://doi.org/10.1002/slct.201600606.
Korea (NRF) grant funded by the Korean Government (MSIT) [20] S. Chan, Y. Xiao, Y. Wang, Z. Hu, H. Zhao, W. Xie, A Facile Approach to Prepare
(2021R1A2C1006315). Black TiO2 with Oxygen Vacancy for Enhancing Photocatalytic Activity,
Nanomaterials. 8 (2018) 245, https://doi.org/10.3390/nano8040245.
[21] J.-J. Li, S.-C. Cai, Z. Xu, X. Chen, J. Chen, H.-P. Jia, J. Chen, Solvothermal syntheses
Appendix A. Supplementary data of Bi and Zn co-doped TiO2 with enhanced electron-hole separation and efficient
photodegradation of gaseous toluene under visible-light, J. Hazard. Mater. 325
(2017) 261–270, https://doi.org/10.1016/j.jhazmat.2016.12.004.
Supplementary data to this article can be found online at https://doi. [22] M.S. Lassoued, A. Lassoued, S. Ammar, A. Gadri, A.B. Salah, S. García-Granda,
org/10.1016/j.apsusc.2024.162244. Synthesis and characterization of Co-doped nano-TiO2 through co-precipitation
method for photocatalytic activity, J. Mater. Sci. Mater. Electron. 29 (2018)
8914–8922, https://doi.org/10.1007/s10854-018-8910-x.
Data availability [23] A. Amorós-Pérez, L. Cano-Casanova, A. Castillo-Deltell, M. Lillo-Ródenas, M.
C. Román-Martínez, TiO2 Modification with Transition Metallic Species (Cr Co, Ni,
Data will be made available on request. and Cu) for Photocatalytic Abatement of Acetic Acid in Liquid Phase and Propene
in Gas Phase, Materials. 12 (2019) 40, https://doi.org/10.3390/ma12010040.
[24] P.M. Leukkunen, E. Rani, A.A.S. Devi, H. Singh, G. King, M. Alatalo, W. Cao,
References M. Huttula, Synergistic effect of Ni–Ag–rutile TiO2 ternary nanocomposite for
efficient visible-light-driven photocatalytic activity, RSC Adv. 10 (2020)
36930–36940, https://doi.org/10.1039/d0ra07078e.
[1] D.J. Lapworth, N. Baran, M.E. Stuart, R.S. Ward, Emerging organic contaminants in
[25] B. Bharti, S. Kumar, H.-N. Lee, R. Kumar, Formation of oxygen vacancies and Ti3+
groundwater: A review of sources, fate and occurrence, Environ. Pollut. 163 (2012)
state in TiO2 thin film and enhanced optical properties by air plasma treatment, Sci
287–303, https://doi.org/10.1016/j.envpol.2011.12.034.
Rep. 6 (2016) 32355, https://doi.org/10.1038/srep32355.
[2] K. Kümmerer, A. Henninger, Promoting resistance by the emission of antibiotics
[26] W. Xie, R. Li, Q. Xu, Enhanced photocatalytic activity of Se-doped TiO2 under
from hospitals and households into effluent, Clin. Microbiol. Infect. 9 (2003)
visible light irradiation, Sci Rep. 8 (2018) 8752, https://doi.org/10.1038/s41598-
1203–1214, https://doi.org/10.1111/j.1469-0691.2003.00739.x.
018-27135-4.
[3] J.A. Park, M. Pineda, M.-L. Peyot, V. Yargeau, Degradation of oxytetracycline and
[27] S. Praserthdam, M. Rittiruam, K. Maungthong, T. Saelee, S. Somdee,
doxycycline by ozonation: Degradation pathways and toxicity assessment, Sci.
P. Praserthdam, Performance controlled via surface oxygen-vacancy in Ti-based
Total Environ. 856 (2023) 159076, https://doi.org/10.1016/j.
oxide catalyst during methyl oleate epoxidation, Sci Rep. 10 (2020) 18952,
scitotenv.2022.159076.
https://doi.org/10.1038/s41598-020-76094-2.
[4] G. Varshney, S.R. Kanel, D.M. Kempisty, V. Varshney, A. Agrawal, E. Sahle-
[28] P.N. Didwal, P.R. Chikate, P.K. Bankar, M.A. More, R.S. Devan, Intense field
Demessie, R.S. Varma, M.N. Nadagouda, Nanoscale TiO2 films and their
electron emission source designed from large area array of dense rutile TiO2
application in remediation of organic pollutants, Coord. Chem. Rev. 306 (2016)
nanopillars, J. Mater. Sci. Mater. Electron. 30 (2019) 2935–2941, https://doi.org/
43–63, https://doi.org/10.1016/j.ccr.2015.06.011.
10.1007/s10854-018-00570-9.
[5] A. Farzaneh, M. Javidani, M.D. Esrafili, O. Mermer, Optical and photocatalytic
[29] M. Kwoka, V. Galstyan, E. Comini, J. Szuber, Pure and Highly Nb-Doped Titanium
characteristics of Al and Cu doped TiO2: Experimental assessments and DFT
Dioxide Nanotubular Arrays: Characterization of Local Surface Properties,
calculations, J. Phys. Chem. Solids. 161 (2022) 110404, https://doi.org/10.1016/j.
Nanomaterials. 7 (2017) 456, https://doi.org/10.3390/nano7120456.
jpcs.2021.110404.
[30] X. Zhu, G. Wen, H. Liu, S. Han, S. Chen, Q. Kong, W. Feng, One-step hydrothermal
[6] E.M. Modan, A.G. Plăiasu, Advantages and Disadvantages of Chemical Methods in
synthesis and characterization of Cu-doped TiO2 nanoparticles/nanobucks/
the Elaboration of Nanomaterials, MMS. 43 (2020) 53, https://doi.org/10.35219/
nanorods with enhanced photocatalytic performance under simulated solar light,
mms.2020.1.08.
J. Mater. Sci. Mater. Electron. 30 (2019) 13826–13834, https://doi.org/10.1007/
[7] H. Lee, I.-S. Park, H.-J. Bang, Y.-K. Park, H. Kim, H.-H. Ha, B.-J. Kim, S.-C. Jung,
s10854-019-01766-3.
Fabrication of Gd-La codoped TiO2 composite via a liquid phase plasma method
[31] Z. Wang, Y. Song, X. Cai, J. Zhang, T. Tang, S. Wen, Rapid preparation of terbium-
and its application as visible-light photocatalysts, Appl. Surf. Sci. 471 (2019)
doped titanium dioxide nanoparticles and their enhanced photocatalytic
893–899, https://doi.org/10.1016/j.apsusc.2018.11.249.
performance, R. Soc. Open Sci. 6 (2019) 191077, https://doi.org/10.1098/
[8] S.J. Ki, Y.-K. Park, J.-S. Kim, W.-J. Lee, H. Lee, S.-C. Jung, Facile preparation of
rsos.191077.
tungsten oxide doped TiO2 photocatalysts using liquid phase plasma process for
[32] S. Guo, J. Liu, S. Qiu, W. Liu, Y. Wang, N. Wu, J. Guo, Z. Guo, Porous ternary TiO2/
enhanced degradation of diethyl phthalate, Chem. Eng. J. 377 (2019) 120087,
MnTiO3@C hybrid microspheres as anode materials with enhanced
https://doi.org/10.1016/j.cej.2018.10.024.
electrochemical performances, J. Mater. Chem. A. 3 (2015) 23895–23904, https://
[9] S.-C. Kim, Y.-K. Park, S.-C. Jung, Recent applications of the liquid phase plasma
doi.org/10.1039/c5ta06437f.
process, Korean J. Chem. Eng. 38 (2021) 885–898, https://doi.org/10.1007/
[33] I.M. Nagpure, S.S. Pitale, E. Coetsee, O.M. Ntwaeaborwa, J.J. Terblans, H.C. Swart,
s11814-020-0739-3.
Low voltage electron induced cathodoluminescence degradation and surface
[10] S. Bingham, W.A. Daoud, Recent advances in making nano-sized TiO2 visible-light
characterization of Sr3(PO4)2: Tb phosphor, Appl. Surf. Sci. 257 (2011)
active through rare-earth metal doping, J. Mater. Chem. 21 (2011) 2041–2050,
10147–10155, https://doi.org/10.1016/j.apsusc.2011.07.008.
https://doi.org/10.1039/C0JM02271C.
[34] A.K. Bedyal, V. Kumar, H.C. Swart, A potential green emitting citrate gel
[11] M. Shaban, J. Poostforooshan, A.P. Weber, Surface-initiated polymerization on
synthesized NaSrBO3:Tb3+ phosphor for display application, Physica B. 535
unmodified inorganic semiconductor nanoparticles via surfactant-free aerosol-
(2018) 189–193, https://doi.org/10.1016/j.physb.2017.07.034.
based synthesis toward core–shell nanohybrids with a tunable shell thickness,
[35] M. Balaguer, C.-Y. Yoo, H.J.M. Bouwmeester, J.M. Serra, Bulk transport and
J. Mater. Chem. A. 5 (2017) 18651–18663, https://doi.org/10.1039/c7ta04985d.
oxygen surface exchange of the mixed ionic–electronic conductor Ce1− xTbxO2− δ
[12] V. Štengl, S. Bakardjieva, N. Murafa, Preparation and photocatalytic activity of rare
(x = 0.1, 0.2, 0.5), J. Mater. Chem. A. 1 (2013) 10234-10242. Doi: 10.1039/
earth doped TiO2 nanoparticles, Mater. Chem. Phys. 114 (2009) 217–226, https://
c3ta11610g.
doi.org/10.1016/j.matchemphys.2008.09.025.
[36] Y. Li, W. Zhang, X. Shen, P. Peng, L. Xiong, Y. Yu, Octahedral Cu2O-modified TiO2
[13] T. Raguram, K.S. Rajni, Synthesis and characterisation of Cu - Doped TiO2
nanotube arrays for efficient photocatalytic reduction of CO2, Chin. J. Catal. 36
nanoparticles for DSSC and photocatalytic applications, Int. J. Hydrogen Energy.
(2015) 2229–2236, https://doi.org/10.1016/s1872-2067(15)60991-3.
47 (2022) 4674–4689, https://doi.org/10.1016/j.ijhydene.2021.11.113.
[37] Z. Chen, X. Zhu, J. Xiong, Z. Wen, G. Cheng, A p–n Junction by Coupling Amine-
[14] L.G. Devi, B.N. Murthy, S.G. Kumar, Photocatalytic activity of TiO2 doped with
Enriched Brookite–TiO2 Nanorods with CuxS Nanoparticles for Improved
Zn2+ and V5+ transition metal ions: Influence of crystallite size and dopant
Photocatalytic CO2 Reduction, Materials 16 (2023) 960, https://doi.org/10.3390/
electronic configuration on photocatalytic activity, Mater. Sci. Eng. B. 166 (2010)
ma16030960.
1–6, https://doi.org/10.1016/j.mseb.2009.09.008.
[38] Z. Cheng, L. Xu, X. Wang, X. Wu, F. Huang, Y. Jing, H. Su, The effect of B-site ions
[15] M.S. Hossain, S. Ahmed, Easy and green synthesis of TiO2 (Anatase and Rutile):
on crystal structure evolution and microwave dielectric properties of gillespite-
Estimation of crystallite size using Scherrer equation, Williamson-Hall Plot,
type SrCu0.95B0.05(B2+: Cu Co, Mn, Ni, Mg, Zn)Si4O10, Ceram. Int. 49 (2023)
Monshi-Scherrer Model, Size-Strain Plot, Halder- Wagner Model, Results Mater. 20
36800–36806, https://doi.org/10.1016/j.ceramint.2023.09.009.
(2023) 100492, https://doi.org/10.1016/j.rinma.2023.100492.
[39] C. Boruban, E.N. Esenturk, Activated carbon-supported CuO nanoparticles: a
[16] M.K. Mun, W.O. Lee, J.W. Park, D.S. Kim, G.Y. Yeom, D.W. Kim, Nanoparticles
hybrid material for carbon dioxide adsorption, J. Nanopart. Res. 20 (2018) 59,
Synthesis and Modification using Solution Plasma Process, Appl. Sci. Converg.
https://doi.org/10.1007/s11051-018-4139-0.
Technol. 26 (2017) 164–173, https://doi.org/10.5757/ASCT.2017.26.6.164.
[40] G. Colón, M. Maicu, M.C. Hidalgo, J.A. Navío, Cu-doped TiO2 systems with
[17] S. Wang, L. Bai, X. Ao, Preparation and photocatalytic application of a S, Nd double
improved photocatalytic activity, Appl. Catal. B. 67 (2006) 41–51, https://doi.org/
doped nano-TiO2 photocatalyst, RSC Adv. 8 (2018) 36745–36753, https://doi.org/
10.1016/j.apcatb.2006.03.019.
10.1039/c8ra06778c.
[41] A. Martinez-Oviedo, S.K. Ray, B. Joshi, S.W. Lee, Enhancement of NOx photo-
[18] X. Zhu, Q. Zhou, Y. Xia, J. Wang, H. Chen, Q. Xu, J. Liu, W. Feng, S. Chen,
oxidation by Fe- and Cu-doped blue TiO2, Environ. Sci. Pollut. Res. 27 (2020)
Preparation and characterization of Cu-doped TiO2 nanomaterials with anatase/
26702–26713, https://doi.org/10.1007/s11356-020-09078-4.
rutile/brookite triphasic structure and their photocatalytic activity, J Mater Sci:
[42] J. Xu, Y. Ao, D. Fu, C. Yuan, A simple route for the preparation of Eu, N-codoped
Mater Electron. 32 (2021) 21511–21524, https://doi.org/10.1007/s10854-021-
TiO2 nanoparticles with enhanced visible light-induced photocatalytic activity,
06660-5.
10
C.-S. You et al. Applied Surface Science 687 (2025) 162244
J. Colloid Interface Sci. 328 (2008) 447–451, https://doi.org/10.1016/j. [48] A. Das, M.K. Adak, N. Mahata, B. Biswas, Wastewater treatment with the advent of
jcis.2008.08.053. TiO2 endowed photocatalysts and their reaction kinetics with scavenger effect,
[43] S. Zhang, Y. Xu, W. Zhang, P. Cao, Synthesis, characterization, and photocatalytic J. Mol. Liq. 338 (2021) 116479, https://doi.org/10.1016/j.molliq.2021.116479.
performance of Cu/Y co-doped TiO2 nanoparticles, Mater. Chem. Phys. 277 (2022) [49] H. Wang, N. Zhang, G. cheng, H. Guo, Z. Shen, L. Yang, Y. Zhao, A. Alsaedi,
125558, https://doi.org/10.1016/j.matchemphys.2021.125558. T. Hayat, X. Wang, Preparing a photocatalytic Fe doped TiO2/rGO for enhanced
[44] Z.M. El-Bahy, A.A. Ismail, R.M. Mohamed, Enhancement of titania by doping rare bisphenol A and its analogues degradation in water sample, Appl. Surf. Sci. 505
earth for photodegradation of organic dye (Direct Blue), J. Hazard. Mater. 166 (2020) 144640, https://doi.org/10.1016/j.apsusc.2019.144640.
(2009) 138–143, https://doi.org/10.1016/j.jhazmat.2008.11.022. [50] J. Zhou, S. Wang, Enhanced removal of oxytetracycline from aquatic solution using
[45] M.Y. Naz, S. Shukrullah, S.U. Rehman, A.A. Al-Arainy, R. Meer, Optical MnOx@Fe3O4 bimetallic nanoparticle coated powdered activated carbon:
characterization of non-thermal plasma jet energy carriers for effective catalytic synergism of adsorption and chemical autocatalytic oxidation processes†, Environ.
processing of industrial wastewaters, Sci. Rep. 11 (2021) 2896, https://doi.org/ Sci.: Nano. 10 (2023) 3171–3183, https://doi.org/10.1039/D3EN00465A.
10.1038/s41598-021-82019-4. [51] D. Liu, M. Li, X. Li, F. Ren, P. Sun, L. Zhou, Core-shell Zn/Co MOFs derived Co3O4/
[46] Z.-Q. Chen, Z.-X. Yin, G.-Q. Xia, L.-L. Hong, Y.-L. Hu, M.-H. Liu, X.-W. Hu, A. CNTs as an efficient magnetic heterogeneous catalyst for persulfate activation and
A. Kudryavtsev, Pulsed microwave-driven argon plasma jet with distinctive plume oxytetracycline degradation, Chem. Eng. J. 387 (2020) 124008, https://doi.org/
patterns resonantly excited by surface plasmon polaritons, Chin. Phys. B. 24 (2015) 10.1016/j.cej.2019.124008.
025203, https://doi.org/10.1088/1674-1056/24/2/025203. [52] Y. Gao, W. Yang, F. Wang, Y. Li, S. Cui, X. Liao, J. Yang, Photocatalytic degradation
[47] J.O. Vitoriano, R.S. Pessoa, A.A.M. Filho, J. Amorim Filho, C. Alves-Junior, of oxytetracycline by UiO-66 doped three-dimensional flower-like MoS2
Titanium Oxynitriding by Plasma-Assisted Thermochemical Treatments Using a heterojunction: DFT, degradation pathways, mechanism, J. Taiwan Inst. Chem.
Competitive Atmosphere of H2–N2-O2, Mater. Res. 26 (2023) 20230272, https:// Eng. 152 (2023) 105160, https://doi.org/10.1016/j.jtice.2023.105160.
doi.org/10.1590/1980-5373-MR-2023-0272.
11