Next Article in Journal
The Use of Modified Fe3O4 Particles to Recover Polyphenolic Compounds for the Valorisation of Olive Mill Wastewater from Slovenian Istria
Previous Article in Journal
Applications of Nano Hydroxyapatite as Adsorbents: A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Degradation of 4-Tert-Butylphenol in Water Using Mono-Doped (M1: Mo, W) and Co-Doped (M2-M1: Cu, Co, Zn) Titania Catalysts

by
Saule Mergenbayeva
1,
Alisher Kumarov
1,
Timur Sh. Atabaev
2,
Evroula Hapeshi
3,
John Vakros
4,
Dionissios Mantzavinos
4 and
Stavros G. Poulopoulos
1,*
1
Department of Chemical and Materials Engineering, School of Engineering and Digital Sciences, Nazarbayev University, 53 Kabanbay Batyr Ave., Nur-Sultan 010000, Kazakhstan
2
Department of Chemistry, School of Sciences and Humanities, Nazarbayev University, 53 Kabanbay Batyr Ave., Nur-Sultan 010000, Kazakhstan
3
Department of Life and Health Sciences, School of Sciences and Engineering, University of Nicosia, 2417 Nicosia, Cyprus
4
Department of Chemical Engineering, University of Patras, Caratheodory 1, University Campus, GR-26504 Patras, Greece
*
Author to whom correspondence should be addressed.
Nanomaterials 2022, 12(14), 2326; https://doi.org/10.3390/nano12142326
Submission received: 11 June 2022 / Revised: 30 June 2022 / Accepted: 4 July 2022 / Published: 6 July 2022

Abstract

:
Mono-doped (Mo-TiO2 and W-TiO2) and co-doped TiO2 (Co-Mo-TiO2, Co-W-TiO2, Cu-Mo-TiO2, Cu-W-TiO2, Zn-Mo-TiO2, and Zn-W-TiO2) catalysts were synthesized by simple impregnation methods and tested for the photocatalytic degradation of 4-tert-butylphenol in water under UV (365 nm) light irradiation. The catalysts were characterized with various analytical methods. X-ray diffraction (XRD), Raman, Diffuse reflectance (DR) spectroscopies, Scanning electron microscopy (SEM), Transmission electron microscopy (TEM), and Energy dispersive spectroscopy (EDS) were applied to investigate the structure, optical properties, morphology, and elemental composition of the prepared catalysts. The XRD patterns revealed the presence of peaks corresponding to the WO3 in W-TiO2, Co-W-TiO2, Cu-W-TiO2, and Zn-W-TiO2. The co-doping of Cu and Mo to the TiO2 lattice was evidenced by the shift of XRD planes towards higher 2θ values, confirming the lattice distortion. Elemental mapping images confirmed the successful impregnation and uniform distribution of metal particles on the TiO2 surface. Compared to undoped TiO2, Mo-TiO2 and W-TiO2 exhibited a lower energy gap. Further incorporation of Mo-TiO2 with Co or Cu introduced slight changes in energy gap and light absorption characteristics, particularly visible light absorption. In addition, photoluminescence (PL) showed that Cu-Mo-TiO2 has a weaker PL intensity than undoped TiO2. Thus, Cu-Mo-TiO2 showed better catalytic activity than pure TiO2, achieving complete degradation of 4-tert-butylphenol under UV light irradiation after 60 min. The application of Cu-Mo-TiO2 under solar light conditions was also tested, and 70% of 4-tert-butylphenol degradation was achieved within 150 min.

1. Introduction

Industrialization on a large scale, along with urbanization and population growth, results in the development of vast volumes of wastewater with different pollutants (inorganic and organic). Numerous organic pollutants found in wastewater are hazardous and may pose a threat to the aquatic environment and living beings [1]. These pollutants include endocrine disrupting chemicals (EDC), pharmaceuticals, and personal care products (PPCPs) [2]. The presence of EDCs in water sources has become one of the major environmental issues [3]. Even at a low exposure level, they may cause the disruption of endocrine and reproductive systems [4]. 4-tert-Butylphenol (4-t-BP) is a synthetic EDC that has been widely utilized in the manufacture of polycarbonate, phenolic, and epoxy resins and, thus, is commonly detected in seas, rivers, sediments, and landfill leachate [5,6,7]. As a typical EDC, 4-t-BP was found to have poor biological degradability and high estrogenic activity [8,9]. Due to the persistence [10,11] and adverse effects of 4-t-BP on aquatic life [12] and living creatures, economically viable and sustainable technology is highly needed to eliminate 4-t-BP from water.
Several approaches have been investigated for the removal of 4-t-BP, including biological processes and advanced oxidation processes (AOPs) [11,13,14]. In contrast with the inefficiency of and relatively long time required by biological degradation, AOPs have received a lot of interest for their ability to remove such persistent pollutants by turning them into carbon dioxide and water [15,16,17]. AOPs are based on the generation of highly reactive radicals, such as hydroxyl radicals (•OH), that can easily react with organic compounds [18,19,20]. Heterogeneous photocatalysis is an AOP that has been successfully employed to remove different organic pollutants. The process is considered to be promising mainly due to its low cost and mild operating conditions, namely, ambient pressure and room temperature [21,22].
TiO2-based photocatalysts continue to be one of the most investigated materials, owing to their great photocatalytic activity, chemical stability, and availability [23,24]. TiO2 may be found in three different crystallographic forms, which are anatase, rutile, and brookite [25,26]. TiO2 in the P25 form is a mixture of anatase and rutile and is one of the most powerful photocatalytic materials. However, the large energy band gap of TiO2 (about 3.0–3.2 eV) prohibits its application under visible light. To extend its light absorption property to the visible light region, mono-doping and co-doping of TiO2 with various cationic and anionic impurities can substantially reduce the band gap energy and thereby improve photocatalytic efficiency in the visible light region [27,28].
The incorporation of transition metals with higher oxidation states, such as Mo6+ and W6+, into the TiO2 lattice has shown great promise due to the broadening of the spectral response and the ability to carry out visible light photocatalysis [29,30,31]. Both Mo6+ and W6+ have similar ionic radii to Ti4+; therefore, they are easy to introduce into the TiO2 lattice [32]. Additionally, the presence of Mo6+ has been proved to be beneficial for the formation of Ti3+ defect sites and suppressing charge carrier recombination, thanks to the redox potential of Mo6+/Mo5+ (vs. NHE), which is 0.4 V [33]. On the other hand, modification of TiO2 with W6+ may increase the surface acidity of the catalyst, leading to the absorption of more hydroxyl groups and pollutant molecules. Avilés-García et al. investigated the effect of Mo- and W- dopants on the performance of TiO2 [34]. The results demonstrated Mo-TiO2 and W-TiO2 to have better activity for 4-chlorophenol degradation than TiO2, attributed to the high surface area and enhanced charge separation [35].
In addition to Mo and W, modifying TiO2 with transition metals such as Cu, Co, and Zn could enhance the photocatalytic activity for the degradation of different organic pollutants in the UV–visible region by changing the physicochemical properties of TiO2 [36,37,38].
In this work, mono-doped (Mo-TiO2 and W-TiO2) and co-doped (Co-Mo-TiO2, Zn-Mo-TiO2, Co-W-TiO2, Cu-Mo-TiO2, Cu-W-TiO2, and Zn-W-TiO2) catalysts were synthesized through simple impregnation methods. The as-prepared samples were characterized by means of SEM, TEM, XRD, Raman, and UV–VIS DR spectroscopies to study their morphology, textural properties, crystal structure, and optical properties. Notably, this is the first time that ternary systems of TiO2 have been synthesized using two transition metals. Their photocatalytic activity was analyzed using 4-t-BP degradation under near-visible light (365 nm) and solar light irradiation.

2. Materials and Methods

2.1. Materials

4-t-BP (99%), applied as a target pollutant; titanium (IV) oxide (TiO2-P25, nanopowder, 21 nm primary particle size, ≥99.5%), used as the base photocatalyst; ammonium metatungstate ((NH4)6H2W12O40 xH2O analytical grade, CAS number 12333-11-8); ammonium heptamolybdate ((NH4)6Mo72O24 4H2O analytical grade, CAS number 12054-85-2), cobalt nitrate (Co(NO3)2·6 H2O analytical grade, CAS number 10026-22-9), copper nitrate (Cu(NO3)2·3 H2O analytical grade, CAS number 10031-43-3), and zinc nitrate (Zn(NO3)2·6 H2O analytical grade, CAS number 10196-18-6), used for the deposition of the dopant metal; and NH4OH solution (28–30% ACS reagent, CAS Number 1336-21-6) were purchased from Sigma Aldrich (Saint Louis, MO, USA). The ultrapure water used in all experiments was obtained by means of a Direct-Q 3UV (Millipore, Darmstadt, Germany) water purification system. All chemical reagents were used without further purification.

2.2. Preparation of Photocatalysts

In this work, two base photocatalysts, namely, Mo/TiO2 and W/TiO2, were prepared using commercial TiO2 (Degussa P25, Saint Louis, MO, USA) as a support with the wet impregnation method, and the active species were W(VI) or Mo(VI) oxoanions using ammonium metatungstate and ammonium heptamolybdate, respectively. A proper amount of TiO2 (3.0 g) was suspended in the solution containing the required amount of Mo or W oxo species (2.5 × 10–4 moles Mo or W) for coverage with Mo or W of 1 at/nm2. The pH of the suspension was around 5 for the Mo solution, while, for the W solution, it was raised to 10 using the 28% NH4OH solution in order to depolymerize the polytungstate species and increase solubility. The suspension was place in a rotary evaporator and left under rotation for 90 min at 45 °C in order to maximize the amount of adsorption. Afterwards, a vacuum was applied and the water evaporated, followed by drying at 105 °C for 2 h and calcination at 400 °C for 5 h. The two base catalysts were then used to prepare ternary systems by dry impregnation. The third cation deposited was either Co(II), Cu(II), or Zn(II), using the corresponding nitrate salts, with the surface concentration of the M(II) ion set to 0.5 at/nm2 (4.15 × 10–5 moles M(II) ions for 1 g of M1/TiO2). After impregnation, the samples were dried and calcined under the same conditions as for the base catalysts.

2.3. Characterization

The phase composition of all of the prepared catalysts was evaluated by X-ray diffraction (XRD, Rigaku SmartLab automated multifunctional X-ray Diffractometer, Tokyo, Japan) using Cu Kα radiation in the scanning range of 10–80°. The Scherrer equation (1) [39] was used to determine the average size of the TiO2 nanoparticles:
D = 0.9 λ β cos θ
where D is the crystallite size of the catalyst, λ is the X-ray wavelength (1.54060 Å), β is the full width at half maximum of the diffraction peak, and θ is the diffraction angle.
Raman spectra were recorded using a Raman spectrometer (Horiba, LabRam HR evolution, Kyoto, Japan) with a 532 nm laser excitation. A transmission electron microscope (TEM, JEM-2100 from Jeol Ltd., Japan) and a scanning electron microscope (SEM, Carl Zeiss Auriga Cross Beam 540) equipped with an energy-dispersive spectroscopy (EDS) system were applied to perform surface morphology measurements and to analyze the elemental composition of the catalysts. The optical properties of the mono- and co-doped TiO2 nanoparticles were investigated by means of diffuse reflectance spectroscopy (DRS, Varian Cary 3, Palo Alto, CA, USA). The recombination behaviors of charge carriers for Cu-Mo-TiO2 were obtained via photoluminescence (PL) emission analysis performed on a fluorescence spectrophotometer (F-7000, Hitachi, Tokyo, Japan). The specific surface area (SSA) of the catalysts was determined from N2 adsorption isotherms in liquid N2 temperature in a Tristar 3000 porosimeter (Micromeritics, Norcross, GA 30093-2901, USA) with the BET method.

2.4. Experimental Procedure

The photodegradation of 4-t-BP was carried out using a photochemical reactor operated in batch mode (Lanphan industry, Zhengzhou City, Henan Province, China) under UV irradiation (365 nm). In a typical experiment, 10 mg of photocatalyst was added to 100 mL of 15 ppm 4-t-BP solution. Prior to irradiation, the solution was stirred for 90 min in the dark to achieve an adsorption–desorption equilibrium between catalyst and pollutant. At a regular time interval (every 30 min), the sample was taken out and filtered through a 0.22 µm Millex syringe filter to remove the photocatalyst for further analysis.
The same procedure was applied to test the photocatalytic activity of Cu-Mo-TiO2 under simulated solar light irradiation (LCS-100 solar simulator, Oriel, Newport, Darmstadt, Germany) within 150 min.
The concentration of 4-t-BP was measured by a high-performance liquid chromatography instrument (HPLC, Agilent 1290 Infinity II, Santa Clara, CA, USA) equipped with a SB-C8 column (2.1 mm × 100 mm, 1.8 µm). The mobile phase composition was methanol and ultrapure water (50:50, v/v), which were mixed to compose the mobile phase.

3. Results and Discussion

3.1. Characterization of Photocatalysts

The SSA of the pure TiO2 (P25) was measured to be equal to 54 m2g−1. The Mo-TiO2 catalysts maintained the SSA value (53 m2g−1) after the deposition of Mo species, while a second impregnation with either Co, Zn, or Cu resulted in an almost unchanged SSA (52 m2g−1) for all of the ternary Mo-TiO2 systems. This was expected, since the loading of the second metal ion is low, while the deposition of Mo species occurs mainly with adsorption or interfacial deposition [40].
On the other hand, the deposition of W oxo species decreased the SSA value to 47 m2g−1. The decrease in SSA value can be due to the lower contribution of adsorption in the deposition of W species in contrast with the Mo deposition. This is caused by the higher solution pH, which does not favor adsorption [41]. The deposition of the second metal ion had no influence on the SSA value (45 m2g−1).
XRD analysis was performed to investigate the crystal structures of the mono- and co-doped TiO2 catalysts, and the results are shown in Figure 1, Figure 2 and Figure 3. The XRD patterns of all prepared catalysts were similar to that of pure TiO2, and a series of diffraction peaks corresponding to the anatase phase of TiO2 (ICDD Card No. 01-070-8501) can be recognized with the planes of (101), (004), (200), (105), (211), (204), (116), (220), and (215) at a degree of 2θ = 24.2°, 36.72°, 46.98°, 52.77°, 53.91°, 61.7°, 67.6°, and 73.92°, respectively. On the other hand, the peaks attributed to the rutile phase of TiO2 (ICDD Card No. 01-087-0920) were detected at around 2θ = 27.1° and 62°. Since the ionic radii of doped transition metals (Mo6+, W6+, Cu2+, Co2+, and Zn2+) are close to that of Ti4+ [42,43,44], minimum changes occurred in the original structure of TiO2.
The structure of the catalyst did not change drastically after the deposition of metals. No typical peaks were detected, which verified the impregnation of Mo-dopant into TiO2 (Figure 1) as well as the subsequent introduction of Cu, Co, and Zn particles into Mo-TiO2 (Figure 2). This finding could be attributed to the high dispersion of metal particles on the surface of TiO2 [45]. Interestingly, a slight shift of the intense TiO2 (101) peak towards a higher angle (from 24.2° to 24.6°) was observed only in Mo-Cu-TiO2, suggesting the existence of some disorders in the anatase crystal lattice [46,47].
Unlike Mo-doping, the introduction of W– into TiO2 formed new peaks at 2θ = 22.65° and 32.75° assigned to the WO3 phase (Figure 3). This is in accordance with the low decrease in SSA value for the W-TiO2 catalyst. The deposition of the second metal ion did not significantly alter the XRD pattern.
The average crystallite size of all prepared catalysts was calculated using Scherrer’s equation, and the results are listed in Table 1. The values confirmed the well-dispersed Mo phase and the existence of the nanoparticles on the prepared catalysts.
Raman spectroscopy was used to obtain more information about the mono- and co-doped TiO2 nanoparticles, and the spectra in the range of 100–800 cm−1 are depicted in Figure 4 and Figure 5. Accordingly, the peaks located at 142 cm−1, 192 cm−1, 394 cm−1, 513 cm−1, and 634 cm−1 matched well with the anatase phase,, while 268 cm−1 and 803 cm−1 confirm the presence of rutile phase. Any peaks corresponding to the doped transition metals could not be detected, although the shift of the main peak at 142 cm−1 towards a greater wavelength was observed for Cu-Mo-TiO2 and W-doped catalysts (W-TiO2, Cu-W-TiO2, Co-W-TiO2, and Zn-W-TiO2). The results of Raman spectroscopy related to the alternation in structure are in good agreement with experimental X-ray findings [48].
The surface morphology of the obtained catalysts was studied by both SEM and TEM analysis, depicted in Figure 6, Figure 7, Figure 8 and Figure 9. As can be seen from SEM images (Figure 6 and Figure 7), all prepared catalysts were found to be relatively spherical in shape, with particle sizes between 23 nm and 35 nm, like pure TiO2. These observations confirm the fact that the introduction of metals did not significantly affect the morphology of TiO2.
Notably, close-up TEM images (Figure 8 and Figure 9) reveal lattice spacing values of 0.37–0.41 nm that correspond to the [101] plane of TiO2 anatase. Overall, the results obtained from SEM and TEM characterizations (average particle size, crystal structure) are in good agreement with XRD and Raman findings.
In addition, EDS analysis was employed to investigate the elemental composition of the prepared catalysts. Although the doped metals were not visible as separate particles in TEM micrographs, EDS mappings (Figure 10, Figure 11 and Figure 12) revealed the presence and homogeneous allocation of impregnated metals throughout the surface of TiO2.
The optical absorption properties of the prepared catalysts were revealed by DRS, as presented in Figure 13, Figure 14 and Figure 15. Obviously, pure TiO2 absorbed below 350 nm, while the incorporation of transition metals (Mo and W) into TiO2 induced the enhancement of the absorption capacity of near-UV light (350–450 nm). Compared with the energy gap of about 3.09 eV of undoped TiO2, the energy gap decreased to 2.92 eV and 2.87 eV after doping with Mo- and W-. A possible reason is the interaction between Mo or W with TiO2 [49,50,51]. The origin of these interactions is the formation of M1-O-Ti bonds (M1: Mo or W) and the charge transfer from Ti to Mo. These charge transfer phenomena are common in systems where an oxidic support is covered by a transition metal oxide, as in our case [40,41,52,53,54].
As can be seen, these interactions were rather higher in the case of Mo-TiO2, since the F(R), an analogue to absorption, was more intense for this sample, while the surface coverage seemed to be a little smaller in the case of W-TiO2, as the F(R) was higher in the UV region. This is in accordance with the XRD results, where crystallites of WO3 were detected. Both binary systems absorb less in the UV region than bare TiO2.
Concerning the M2-Mo-TiO2 samples, no significant differences could be observed (Figure 13). The coverage of TiO2 was higher, while, in the case of Co-Mo-TiO2, the adsorption in the near-UV region was higher, suggesting more intense interactions with the Co phase. Absorption in the visible region was small for the samples Co-Mo-TiO2 and Cu-Mo-TiO2, although the black color of the corresponding bulk oxides was due to the small quantity of the Co and Cu phases. This may suggest that the above oxides were rather well-dispersed on the surface of the catalyst.
The M2-W-TiO2 samples had similar behavior. Only the Cu-W-TiO2 sample had smaller absorption in the UV region, suggesting that the coverage of TiO2 was higher in this case.
As was discussed, the doping of Mo-TiO2 and W- TiO2 with Co, Cu, or Zn caused the formation of a more intense peak centered at about 400 nm (Figure 14 and Figure 15) and a slight reduction in the energy gap. The energy gap (Figure 16) for Zn-Mo-TiO2, Cu-Mo-TiO2, and Co-Mo-TiO2 were estimated to be 2.85 eV, 2.82 eV, and 2.72 eV, respectively. Additionally, as for the Co-W-TiO2, Zn-W-TiO2, and Cu-W-TiO2, the Eg values were 2.87 eV, 2.86 eV, and 2.85 eV, respectively.

3.2. Adsorption and Photocatalytic Degradation of 4-t-BP

The adsorption and photocatalytic degradation of 4-t-BP using mono- and co-doped TiO2 catalysts were evaluated under dark conditions for 90 min and UV/solar light irradiation, respectively. The adsorption performance of each catalyst was identified through the determination of adsorption capacity q (mg/g) by Equation (2):
q = C 0 C e V m catalyst
where C0 and Ce represent the initial and equilibrium concentrations (mg/L) of 4-t-BP in the solution, V (L) is the volume of the 4-t-BP solution, and m catalyst is the mass of the catalyst.
As shown in Figure 17, the amount of 4-t-BP adsorbed increased more than two-fold after doping TiO2 with Mo- or W-. At the equilibria, the adsorption capacities of Mo-TiO2 and W-TiO2 were found to be 63 mg/g for both catalysts. The enhanced adsorption capacities could be ascribed to changes in strong electrical aspects between the 4-t-BP and the doped catalyst [55]. At this point, it should be noted that the deposition of W or Mo phase increases the acidity of the surface. In a recent paper [56] about the W-TiO2 system, it was found that the addition of W oxo species resulted in lower point of zero values, although it was less acidic than the correspondence value for mixed oxides and changed the acid–base properties. The electron transfer between well-dispersed W phase and the TiO2 surface increases the surface electron density, which enhances the surface basicity of TiO2.
Further addition of Co to Mo-TiO2 had a negative impact on the adsorption performance, while the incorporation of Cu or Co metal ions slightly improved the adsorption of 4-t-BP (Figure 18). Similar results were obtained for doped W-TiO2 (Figure 19). Among all of the synthesized catalysts, Zn-doped materials exhibited the highest 4-t-BP adsorption capacity, while doping with Co had a detrimental effect on the adsorption capacity of both binary systems. Doping with Co increases the interactions between Co and Mo or W and, as a result, decreases the interactions of Mo and W oxo species with the titania surface.
It was reported [57] that surface hydroxyl groups play an important role in the surface properties of a material. These groups often have Brønsted acidity, and, therefore, they play an important role in adsorption or in photocatalytic reactions. For the Mo-TiO2 system, the interactions between Mo phase and TiO2 generate hydroxyl groups. These groups can interact with the second metal ion and immobilize it on the binary system surface.
The deposition of Co2+ ions on either Mo-TiO2 or W-TiO2 shifts the absorption to higher wavelengths, evidence that the Co species can be adsorbed onto the surface –OH groups, diminishing the adsorption sites for 4-t-BP. This is expected, since it is well-known that CoMo catalysts are very stable and active, especially in hydrotreatment. On the other hand, this can significantly alter the photocatalytic properties of the ternary system.
In the presence of UV light, only 50% of 4-t-BP can be photodegraded in 120 min without the application of any catalyst (Figure 20). All prepared catalysts exhibited significant photodegradation when the light was on. Although the impregnation of TiO2 with W and Mo metals led to an improved adsorption of 4-t-BP, the photocatalytic activity of pure TiO2 was higher. This may be due to the higher absorption of TiO2 in the UV region, as was determined by DRS measurements. The results are in agreement with previous reports. For example, it was reported that the presence of various transition metals was not beneficial for the oxidation ability of the solid [58] and that only W-TiO2 had a positive effect on its activity. Generally, the doped catalysts exhibit recombination rates significantly higher than that of the support, which results in lower oxidation ability. Additionally, Mo deposition can have some positive effects on the activity if the deposition is not surface but subsurface [59]. In all cases, the loading of the oxoanion is crucial for the performance of the photocatalyst. Higher loadings result in lower degradation activity.
The introduction of Cu, Zn, and Co metals into Mo-TiO2 and W-TiO2 resulted in different photocatalytic performances of the catalyst. The improved adsorption properties of Mo-TiO2 and W-TiO2 after doping with Cu and Zn facilitated a faster degradation of 4-t-BP (Figure 21 and Figure 22). More specifically, the incorporation of Cu into both Mo-TiO2 and W-TiO2 was favorable, where a slight 4-t-BP degradation increase was observed for Cu-Mo-TiO2 compared with pure TiO2. This is likely due to structural changes induced by the presence of Mo and Cu, which was evidenced by the high adsorption capacity and reduced energy gap coupled with the extended light absorption in the visible region.
On the other hand, Cu-Mo-TiO2 showed a relatively lower PL intensity than that of pure TiO2 (Figure 23). This observation indicates a better charge separation, which could promote the photocatalytic performance of the catalyst.
Accordingly, the photocatalytic activity of Cu-Mo-TiO2 was investigated towards 4-t-BP degradation under solar light irradiation, and its performance was compared with that of mono-doped Mo-TiO2 and W-TiO2 catalysts (Figure 24). It was observed that the application of solar light required more time to achieve decent degradation for all tested catalysts. In 150 min of solar light exposure, about 70% of 4-t-BP could be degraded using the Cu-Mo-TiO2/solar system. Although the Cu-Mo-TiO2 catalyst exhibited better degradation efficiency, the difference was negligible in comparison to the Mo-TiO2 and W-TiO2 catalysts.
In Table 2, the results of this work are compared with previously reported ones.

4. Conclusions

Mono- and co-doped TiO2 nanoparticles with similar morphology were synthesized by simple preparation methods. The catalyst characterization evidenced that the incorporation of transition metals (Mo, W, Cu, Co, and Zn) led to homogeneous distribution of metal particles over the TiO2 surface and reduced the energy gap, which led to optical properties different from those of TiO2. Specifically, impregnation of Cu into Mo-TiO2 led to an increase in light absorption, particularly visible light. The catalysts were further investigated for the adsorption and photocatalytic degradation of 4-t-BP by means of UV (365 nm). Doping with transition metals increased the adsorption capacity of TiO2. The prepared Cu-Mo-TiO2 exhibited higher catalytic activity towards degradation of 4-t-BP than that of pure TiO2, probably due to the synergistic effect of visible light absorption, improved adsorption capacity, and suppressed electron-hole pair recombination. Complete and about 70% 4-t-BP degradation could be achieved within 60 min and 150 min using UV (365 nm) and solar light exposure, respectively.

Author Contributions

Conceptualization, S.G.P. and D.M.; methodology, J.V. and E.H.; investigation, S.M., A.K. and J.V.; resources, S.G.P., T.S.A. and D.M.; writing—original draft preparation, S.M. and A.K.; writing—review and editing, S.G.P., T.S.A., E.H., J.V. and D.M.; supervision, T.S.A. and S.G.P.; project administration, S.G.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Nazarbayev University project “Cost-Effective Photocatalysts for the Treatment of Wastewaters containing Emerging Pollutants”, Faculty Development Competitive Research Grants Program for 2020–2022, Grant Number 240919FD3932, awarded to S.G. Poulopoulos.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The technical support of the Core Facilities of Nazarbayev University is greatly acknowledged.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Saravanan, A.; Kumar, P.S.; Jeevanantham, S.; Anubha, M.; Jayashree, S. Degradation of Toxic Agrochemicals and Pharmaceutical Pollutants: Effective and Alternative Approaches toward Photocatalysis. Environ. Pollut. 2022, 298, 118844. [Google Scholar] [CrossRef] [PubMed]
  2. Su, C.; Cui, Y.; Liu, D.; Zhang, H.; Baninla, Y. Endocrine Disrupting Compounds, Pharmaceuticals and Personal Care Products in the Aquatic Environment of China: Which Chemicals Are the Prioritized Ones? Sci. Total Environ. 2020, 720, 137652. [Google Scholar] [CrossRef] [PubMed]
  3. Jinrong, L.I.; Ruixin, G.U.O.; Yanhua, L.I.U.; Jianqiu, C. Occurrence and risk assessment of five typical environmental endocrine disruptors. Environ. Chem. 2020, 10, 2637–2653. [Google Scholar] [CrossRef]
  4. Bilal, M.; Barceló, D.; Iqbal, H.M.N. Occurrence, Environmental Fate, Ecological Issues, and Redefining of Endocrine Disruptive Estrogens in Water Resources. Sci. Total Environ. 2021, 800, 149635. [Google Scholar] [CrossRef]
  5. Toyama, T.; Momotani, N.; Ogata, Y.; Miyamori, Y.; Inoue, D.; Sei, K.; Mori, K.; Kikuchi, S.; Ike, M. Isolation and Characterization of 4-Tert-Butylphenol-Utilizing Sphingobium Fuliginis Strains from Phragmites Australis Rhizosphere Sediment. Appl. Environ. Microbiol. 2010, 76, 6733–6740. [Google Scholar] [CrossRef] [Green Version]
  6. Villarreal-Morales, R.; Hinojosa-Reyes, L.; Hernández-Ramírez, A.; Ruíz-Ruíz, E.; Treviño, M.d.L.; Guzmán-Mar, J.L. Automated SPE-HPLC-UV Methodology for the on-Line Determination of Plasticisers in Wastewater Samples. Int. J. Environ. Anal. Chem. 2020, 102, 1683–1696. [Google Scholar] [CrossRef]
  7. Bell, A.M.; Keltsch, N.; Schweyen, P.; Reifferscheid, G.; Ternes, T.; Buchinger, S. UV Aged Epoxy Coatings–Ecotoxicological Effects and Released Compounds. Water Res. X 2021, 12, 100105. [Google Scholar] [CrossRef]
  8. Sun, H.; Xu, X.-L.; Qu, J.-H.; Hong, X.; Wang, Y.-B.; Xu, L.-C.; Wang, X.-R. 4-Alkylphenols and Related Chemicals Show Similar Effect on the Function of Human and Rat Estrogen Receptor α in Reporter Gene Assay. Chemosphere 2008, 71, 582–588. [Google Scholar] [CrossRef]
  9. Liu, Z.-H.; Yin, H.; Dang, Z. Do Estrogenic Compounds in Drinking Water Migrating from Plastic Pipe Distribution System Pose Adverse Effects to Human? An Analysis of Scientific Literature. Environ. Sci. Pollut. Res. Int. 2017, 24, 2126–2134. [Google Scholar] [CrossRef]
  10. Zheng, Q.; Wu, N.; Qu, R.; Albasher, G.; Cao, W.; Li, B.; Alsultan, N.; Wang, Z. Kinetics and Reaction Pathways for the Transformation of 4-Tert-Butylphenol by Ferrate(VI). J. Hazard. Mater. 2021, 401, 123405. [Google Scholar] [CrossRef]
  11. Ogata, Y.; Toyama, T.; Yu, N.; Wang, X.; Sei, K.; Ike, M. Occurrence of 4-Tert-Butylphenol (4-t-BP) Biodegradation in an Aquatic Sample Caused by the Presence of Spirodela Polyrrhiza and Isolation of a 4-t-BP-Utilizing Bacterium. Biodegradation 2013, 24, 191–202. [Google Scholar] [CrossRef] [PubMed]
  12. Barse, A.V.; Chakrabarti, T.; Ghosh, T.; Pal, A.; Jadhao, S. One-Tenth Dose of LC50 of 4-Tert-Butylphenol Causes Endocrine Disruption and Metabolic Changes in Cyprinus Carpio. Pestic. Biochem. Physiol. 2006, 86, 172–179. [Google Scholar] [CrossRef]
  13. He, G.; Xing, C.; Xiao, X.; Hu, R.; Zuo, X.; Nan, J. Facile Synthesis of Flower-like Bi12O17Cl2/β-Bi2O3 Composites with Enhanced Visible Light Photocatalytic Performance for the Degradation of 4-Tert-Butylphenol. Appl. Catal. B Environ. 2015, 170, 1–9. [Google Scholar] [CrossRef]
  14. Wu, Y.; Zhu, X.; Chen, H.; Dong, W.; Zhao, J. Photodegradation of 4-Tert-Butylphenol in Aqueous Solution by UV-C, UV/H2O2 and UV/S2O82− System. J. Environ. Sci. Health Part A 2016, 51, 440–445. [Google Scholar] [CrossRef]
  15. Makhatova, A.; Ulykbanova, G.; Sadyk, S.; Sarsenbay, K.; Atabaev, T.; Inglezakis, V.; Poulopoulos, S. Degradation and Mineralization of 4-Tert-Butylphenol in Water Using Fe-Doped TiO2 Catalysts. Sci. Rep. 2019, 9, 19284. [Google Scholar] [CrossRef]
  16. Kanafin, Y.N.; Makhatova, A.; Zarikas, V.; Arkhangelsky, E.; Poulopoulos, S.G. Photo-Fenton-Like Treatment of Municipal Wastewater. Catalysts 2021, 11, 1206. [Google Scholar] [CrossRef]
  17. Mergenbayeva, S.; Ashir, A.; Yergali, B.; Ulykbanova, G.; Poulopoulos, S. Modified TiO2 for Photocatalytic Removal of Organic Pollutants in Water. IOP Conf. Ser. Earth Environ. Sci. 2021, 899, 012068. [Google Scholar] [CrossRef]
  18. Kilic, M.Y.; Abdelraheem, W.H.; He, X.; Kestioglu, K.; Dionysiou, D.D. Photochemical Treatment of Tyrosol, a Model Phenolic Compound Present in Olive Mill Wastewater, by Hydroxyl and Sulfate Radical-Based Advanced Oxidation Processes (AOPs). J. Hazard. Mater. 2019, 367, 734–742. [Google Scholar] [CrossRef]
  19. Mei, Q.; Cao, H.; Han, D.; Li, M.; Yao, S.; Xie, J.; Zhan, J.; Zhang, Q.; Wang, W.; He, M. Theoretical Insight into the Degradation of P-Nitrophenol by OH Radicals Synergized with Other Active Oxidants in Aqueous Solution. J. Hazard. Mater. 2020, 389, 121901. [Google Scholar] [CrossRef]
  20. Kanafin, Y.; Satayeva, A.; Arkhangelsky, E.; Poulopoulos, S. Treatment of a Biological Effluent Containing Metronidazole. Chem. Eng. Trans. 2021, 86, 595–600. [Google Scholar] [CrossRef]
  21. Cheng, Y.; He, L.; Xia, G.; Ren, C.; Wang, Z. Nanostructured G-C3N4/AgI Composites Assembled by AgI Nanoparticles-Decorated g-C3N4 Nanosheets for Effective and Mild Photooxidation Reaction. New J. Chem. 2019, 43, 14841–14852. [Google Scholar] [CrossRef]
  22. Muñoz-Batista, M.J.; Luque, R. Heterogeneous Photocatalysis. ChemEngineering 2021, 5, 26. [Google Scholar] [CrossRef]
  23. Shaban, M.; Ashraf, A.M.; Abukhadra, M.R. TiO2 Nanoribbons/Carbon Nanotubes Composite with Enhanced Photocatalytic Activity. Fabrication, Characterization, and Application. Sci. Rep. 2018, 8, 781. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Govardhana Reddy, P.V.; Rajendra Prasad Reddy, B.; Venkata Krishna Reddy, M.; Raghava Reddy, K.; Shetti, N.P.; Saleh, T.A.; Aminabhavi, T.M. A Review on Multicomponent Reactions Catalysed by Zero-Dimensional/One-Dimensional Titanium Dioxide (TiO2) Nanomaterials: Promising Green Methodologies in Organic Chemistry. J. Environ. Manag. 2021, 279, 111603. [Google Scholar] [CrossRef] [PubMed]
  25. Liu, H.; Wang, X.; Wu, D.; Ji, S. Morphology-Controlled Synthesis of Microencapsulated Phase Change Materials with TiO2 Shell for Thermal Energy Harvesting and Temperature Regulation. Energy 2019, 172, 599–617. [Google Scholar] [CrossRef]
  26. Singh, M.K.; Mehata, M.S. Phase-Dependent Optical and Photocatalytic Performance of Synthesized Titanium Dioxide (TiO2) Nanoparticles. Optik 2019, 193, 163011. [Google Scholar] [CrossRef]
  27. Ioannidou, E.; Frontistis, Z.; Antonopoulou, M.; Venieri, D.; Konstantinou, I.; Kondarides, D.I.; Mantzavinos, D. Solar Photocatalytic Degradation of Sulfamethoxazole over Tungsten–Modified TiO2. Chem. Eng. J. 2017, 318, 143–152. [Google Scholar] [CrossRef]
  28. Petala, A.; Frontistis, Z.; Antonopoulou, M.; Konstantinou, I.; Kondarides, D.I.; Mantzavinos, D. Kinetics of Ethyl Paraben Degradation by Simulated Solar Radiation in the Presence of N-Doped TiO2 Catalysts. Water Res. 2015, 81, 157–166. [Google Scholar] [CrossRef]
  29. Gomathi Devi, L.; Narasimha Murthy, B.; Girish Kumar, S. Heterogeneous Photo Catalytic Degradation of Anionic and Cationic Dyes over TiO2 and TiO2 Doped with Mo6+ Ions under Solar Light: Correlation of Dye Structure and Its Adsorptive Tendency on the Degradation Rate. Chemosphere 2009, 76, 1163–1166. [Google Scholar] [CrossRef]
  30. Devi, L.G.; Kumar, S.G.; Murthy, B.N.; Kottam, N. Influence of Mn2+ and Mo6+ Dopants on the Phase Transformations of TiO2 Lattice and Its Photo Catalytic Activity under Solar Illumination. Catal. Commun. 2009, 10, 794–798. [Google Scholar] [CrossRef]
  31. Liu, X.; Shi, Y.; Dong, Y.; Li, H.; Xia, Y.; Wang, H. A Facile Solvothermal Approach for the Synthesis of Novel W-Doped TiO2 Nanoparticles/Reduced Graphene Oxide Composites with Enhanced Photodegradation Performance under Visible Light Irradiation. New J. Chem. 2017, 41, 13382–13390. [Google Scholar] [CrossRef]
  32. Khlyustova, A.; Sirotkin, N.; Kusova, T.; Kraev, A.; Titov, V.; Agafonov, A. Doped TiO2: Effect of Doping Element on the Photocatalytic Activity. Mater. Adv. 2020, 1, 1193–1201. [Google Scholar] [CrossRef]
  33. Du, Y.K.; Gan, Y.Q.; Yang, P.; Zhao, F.; Hua, N.P.; Jiang, L. Improvement in the Heat-Induced Hydrophilicity of TiO2 Thin Films by Doping Mo(VI) Ions. Thin Solid Film. 2005, 491, 133–136. [Google Scholar] [CrossRef]
  34. Grabowska, E.; Sobczak, J.W.; Gazda, M.; Zaleska, A. Surface Properties and Visible Light Activity of W-TiO2 Photocatalysts Prepared by Surface Impregnation and Sol–Gel Method. Appl. Catal. B Environ. 2012, 117–118, 351–359. [Google Scholar] [CrossRef]
  35. Avilés-García, O.; Espino-Valencia, J.; Romero-Romero, R.; Rico-Cerda, J.L.; Arroyo-Albiter, M.; Solís-Casados, D.A.; Natividad-Rangel, R. Enhanced Photocatalytic Activity of Titania by Co-Doping with Mo and W. Catalysts 2018, 8, 631. [Google Scholar] [CrossRef] [Green Version]
  36. Chen, Q.; Ji, F.; Liu, T.; Yan, P.; Guan, W.; Xu, X. Synergistic Effect of Bifunctional Co–TiO2 Catalyst on Degradation of Rhodamine B: Fenton-Photo Hybrid Process. Chem. Eng. J. 2013, 229, 57–65. [Google Scholar] [CrossRef]
  37. Xie, X.; Li, S.; Zhang, H.; Wang, Z.; Huang, H. Promoting Charge Separation of Biochar-Based Zn-TiO2/PBC in the Presence of ZnO for Efficient Sulfamethoxazole Photodegradation under Visible Light Irradiation. Sci. Total Environ. 2019, 659, 529–539. [Google Scholar] [CrossRef]
  38. Guan, H.; Zhou, X.; Wen, W.; Jin, B.; Li, J.; Zhang, S. Efficient and Robust Cu/TiO2 Nanorod Photocatalysts for Simultaneous Removal of Cr(VI) and Methylene Blue under Solar Light. J. Chin. Chem. Soc. 2018, 65, 706–713. [Google Scholar] [CrossRef]
  39. Patterson, A.L. The Scherrer Formula for X-Ray Particle Size Determination. Phys. Rev. 1939, 56, 978–982. [Google Scholar] [CrossRef]
  40. Panagiotou, G.D.; Petsi, T.; Bourikas, K.; Kalampounias, A.G.; Boghosian, S.; Kordulis, C.; Lycourghiotis, A. Interfacial Impregnation Chemistry in the Synthesis of Molybdenum Catalysts Supported on Titania. J. Phys. Chem. C 2010, 114, 11868–11879. [Google Scholar] [CrossRef]
  41. Panagiotou, G.D.; Petsi, T.; Bourikas, K.; Kordulis, C.; Lycourghiotis, A. The Interfacial Chemistry of the Impregnation Step Involved in the Preparation of Tungsten(VI) Supported Titania Catalysts. J. Catal. 2009, 262, 266–279. [Google Scholar] [CrossRef]
  42. Loan, T.T.; Long, N.N. Effect of Co2+ Doping on Raman Spectra and the Phase Transformation of TiO2:Co2+ Nanowires. J. Phys. Chem. Solids 2019, 124, 336–342. [Google Scholar] [CrossRef]
  43. Loan, T.T.; Huong, V.H.; Tham, V.T.; Long, N.N. Effect of Zinc Doping on the Bandgap and Photoluminescence of Zn2+-Doped TiO2 Nanowires. Phys. B Condens. Matter 2018, 532, 210–215. [Google Scholar] [CrossRef]
  44. De Los Santos, D.M.; Chahid, S.; Alcántara, R.; Navas, J.; Aguilar, T.; Gallardo, J.J.; Gómez-Villarejo, R.; Carrillo-Berdugo, I.; Fernández-Lorenzo, C. MoS2/Cu/TiO2 Nanoparticles: Synthesis, Characterization and Effect on Photocatalytic Decomposition of Methylene Blue in Water under Visible Light. Water Sci. Technol. 2017, 2017, 184–193. [Google Scholar] [CrossRef] [PubMed]
  45. Chahid, S.; Alcantara, R.; los Santos, D.M. Isotherm Analysis for Removal of Organic Pollutants Using Synthesized Mo/Cu/Co-Doped TiO2 Nanostrucrured. In Proceedings of the 2019 5th International Conference on Optimization and Applications (ICOA), Kenitra, Morocco, 25–26 April 2019; pp. 1–9. [Google Scholar] [CrossRef]
  46. Kumaravel, V.; Rhatigan, S.; Mathew, S.; Michel, M.C.; Bartlett, J.; Nolan, M.; Hinder, S.J.; Gascó, A.; Ruiz-Palomar, C.; Hermosilla, D.; et al. Mo Doped TiO2: Impact on Oxygen Vacancies, Anatase Phase Stability and Photocatalytic Activity. J. Phys. Mater. 2020, 3, 025008. [Google Scholar] [CrossRef]
  47. Bashiri, R.; Mohamed, N.M.; Kait, C.F.; Sufian, S. Hydrogen Production from Water Photosplitting Using Cu/TiO2 Nanoparticles: Effect of Hydrolysis Rate and Reaction Medium. Int. J. Hydrog. Energy 2015, 40, 6021–6037. [Google Scholar] [CrossRef]
  48. Popovic, Z.; Dohčević-Mitrović, Z.; Scepanovic, M.; Grujić-Brojčin, M.; Aškrabić, S. Raman Scattering on Nanomaterials and Nanostructures. Ann. Phys. 2011, 523, 62–74. [Google Scholar] [CrossRef]
  49. Bhattacharyya, K.; Majeed, J.; Dey, K.K.; Ayyub, P.; Tyagi, A.K.; Bharadwaj, S.R. Effect of Mo-Incorporation in the TiO2 Lattice: A Mechanistic Basis for Photocatalytic Dye Degradation. J. Phys. Chem. C 2014, 118, 15946–15962. [Google Scholar] [CrossRef]
  50. Esposito, S.; Ditaranto, N.; Dell’Agli, G.; Nasi, R.; Rivolo, P.; Bonelli, B. Effective Inclusion of Sizable Amounts of Mo within TiO2 Nanoparticles Can Be Obtained by Reverse Micelle Sol–Gel Synthesis. ACS Omega 2021, 6, 5379–5388. [Google Scholar] [CrossRef]
  51. Cui, M.; Pan, S.; Tang, Z.; Chen, X.; Qiao, X.; Xu, Q. Physiochemical Properties of N-n Heterostructured TiO2/Mo-TiO2 Composites and Their Photocatalytic Degradation of Gaseous Toluene. Chem. Speciat. Bioavailab. 2017, 29, 60–69. [Google Scholar] [CrossRef] [Green Version]
  52. Ataloglou, T.; Bourikas, K.; Vakros, J.; Kordulis, C.; Lycourghiotis, A. Kinetics of Adsorption of the Cobalt Ions on the “Electrolytic Solution/γ-Alumina” Interface. J. Phys. Chem. B 2005, 109, 4599–4607. [Google Scholar] [CrossRef] [PubMed]
  53. Stamatis, N.; Goundani, K.; Vakros, J.; Bourikas, K.; Kordulis, C. Influence of Composition and Preparation Method on the Activity of MnOx/Al2O3 Catalysts for the Reduction of Benzaldehyde with Ethanol. Appl. Catal. A Gen. 2007, 325, 322–327. [Google Scholar] [CrossRef]
  54. Vakros, J. The Influence of Preparation Method on the Physicochemical Characteristics and Catalytic Activity of Co/TiO2 Catalysts. Catalysts 2020, 10, 88. [Google Scholar] [CrossRef] [Green Version]
  55. Palanivel, M.; Lokeshkumar, E.; Amruthaluru, S.K.; Govardhanan, B.; Mahalingam, A.; Rameshbabu, N. Visible Light Photocatalytic Activity of Metal (Mo/V/W) Doped Porous TiO2 Coating Fabricated on Cp-Ti by Plasma Electrolytic Oxidation. J. Alloy. Compd. 2020, 825, 154092. [Google Scholar] [CrossRef]
  56. Tsatsos, S.; Vakros, J.; Ladas, S.; Verykios, X.E.; Kyriakou, G. The Interplay between Acid-Base Properties and Fermi Level Pinning of a Nano Dispersed Tungsten Oxide-Titania Catalytic System. J. Colloid Interface Sci. 2022, 614, 666–676. [Google Scholar] [CrossRef]
  57. Jada, N.; Jothiramalingam, S.; Sakthivel, R.; Sethi, D.; Mohapatra, P. Synergistic Effect of MoO3/TiO2 towards Discrete and Simultaneous Photocatalytic Degradation of E. Coli and Methylene Blue in Water. Bull. Mater. Sci. 2021, 44, 167. [Google Scholar] [CrossRef]
  58. Di Paola, A.; Marcì, G.; Palmisano, L.; Schiavello, M.; Uosaki, K.; Ikeda, S.; Ohtani, B. Preparation of Polycrystalline TiO2 Photocatalysts Impregnated with Various Transition Metal Ions:  Characterization and Photocatalytic Activity for the Degradation of 4-Nitrophenol. J. Phys. Chem. B 2002, 106, 637–645. [Google Scholar] [CrossRef] [Green Version]
  59. Yang, Y.; Li, X.; Chen, J.; Wang, L. Effect of Doping Mode on the Photocatalytic Activities of Mo/TiO2. J. Photochem. Photobiol. A Chem. 2004, 163, 517–522. [Google Scholar] [CrossRef]
  60. Mergenbayeva, S.; Sh. Atabaev, T.; Poulopoulos, S.G. Ti2O3/TiO2-Assisted Solar Photocatalytic Degradation of 4-Tert-Butylphenol in Water. Catalysts 2021, 11, 1379. [Google Scholar] [CrossRef]
  61. Xiao, X.; Xing, C.; He, G.; Zuo, X.; Nan, J.; Wang, L. Solvothermal Synthesis of Novel Hierarchical Bi4O5I2 Nanoflakes with Highly Visible Light Photocatalytic Performance for the Degradation of 4-Tert-Butylphenol. Appl. Catal. B Environ. 2014, 148–149, 154–163. [Google Scholar] [CrossRef]
  62. Mergenbayeva, S.; Poulopoulos, S.G. Comparative Study on UV-AOPs for Efficient Continuous Flow Removal of 4-Tert-Butylphenol. Processes 2022, 10, 8. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of Mo-TiO2 and W-TiO2.
Figure 1. XRD patterns of Mo-TiO2 and W-TiO2.
Nanomaterials 12 02326 g001
Figure 2. XRD patterns of Mo-TiO2, Co-Mo-TiO2, Cu-Mo-TiO2, and Zn-Mo-TiO2.
Figure 2. XRD patterns of Mo-TiO2, Co-Mo-TiO2, Cu-Mo-TiO2, and Zn-Mo-TiO2.
Nanomaterials 12 02326 g002
Figure 3. XRD patterns of W-TiO2, Co-W-TiO2, Cu-W-TiO2, and Zn-W-TiO2.
Figure 3. XRD patterns of W-TiO2, Co-W-TiO2, Cu-W-TiO2, and Zn-W-TiO2.
Nanomaterials 12 02326 g003
Figure 4. Raman spectra of TiO2, Mo− TiO2, Co− Mo−TiO2, Cu− Mo−TiO2, and Zn− Mo−TiO2.
Figure 4. Raman spectra of TiO2, Mo− TiO2, Co− Mo−TiO2, Cu− Mo−TiO2, and Zn− Mo−TiO2.
Nanomaterials 12 02326 g004
Figure 5. Raman spectra of TiO2, W−TiO2, Co−W−TiO2, Cu−W−TiO2, and Zn−W−TiO2.
Figure 5. Raman spectra of TiO2, W−TiO2, Co−W−TiO2, Cu−W−TiO2, and Zn−W−TiO2.
Nanomaterials 12 02326 g005
Figure 6. SEM images of (a) Mo-TiO2, (b) Co-Mo-TiO2, (c) Cu-Mo-TiO2, (d) Zn-Mo-TiO2, and (e) TiO2.
Figure 6. SEM images of (a) Mo-TiO2, (b) Co-Mo-TiO2, (c) Cu-Mo-TiO2, (d) Zn-Mo-TiO2, and (e) TiO2.
Nanomaterials 12 02326 g006
Figure 7. SEM images of (a) W-TiO2, (b) Co-W-TiO2, (c) Cu-W-TiO2, (d) Zn-W-TiO2, and (e) TiO2.
Figure 7. SEM images of (a) W-TiO2, (b) Co-W-TiO2, (c) Cu-W-TiO2, (d) Zn-W-TiO2, and (e) TiO2.
Nanomaterials 12 02326 g007
Figure 8. TEM analysis of (a) Mo-TiO2, (b) Co-Mo-TiO2, (c) Cu-Mo-TiO2, (d) Zn-Mo-TiO2, and (e) TiO2.
Figure 8. TEM analysis of (a) Mo-TiO2, (b) Co-Mo-TiO2, (c) Cu-Mo-TiO2, (d) Zn-Mo-TiO2, and (e) TiO2.
Nanomaterials 12 02326 g008
Figure 9. TEM analysis of (a) W-TiO2, (b) Co-W-TiO2, (c) Cu-W-TiO2, (d) Zn-W-TiO2, and (e) TiO2.
Figure 9. TEM analysis of (a) W-TiO2, (b) Co-W-TiO2, (c) Cu-W-TiO2, (d) Zn-W-TiO2, and (e) TiO2.
Nanomaterials 12 02326 g009
Figure 10. EDS mapping images of (a) Mo-TiO2 and (b) W-TiO2.
Figure 10. EDS mapping images of (a) Mo-TiO2 and (b) W-TiO2.
Nanomaterials 12 02326 g010
Figure 11. EDS mapping images of (a) Mo-Cu-TiO2, (b) Mo-Co-TiO2, and (c) Mo-Zn-TiO2.
Figure 11. EDS mapping images of (a) Mo-Cu-TiO2, (b) Mo-Co-TiO2, and (c) Mo-Zn-TiO2.
Nanomaterials 12 02326 g011
Figure 12. EDS mapping images of (a) W-Cu-TiO2, (b) W-Co-TiO2, and (c) W-Zn-TiO2.
Figure 12. EDS mapping images of (a) W-Cu-TiO2, (b) W-Co-TiO2, and (c) W-Zn-TiO2.
Nanomaterials 12 02326 g012
Figure 13. DR spectra of TiO2, Mo-TiO2, and W-TiO2.
Figure 13. DR spectra of TiO2, Mo-TiO2, and W-TiO2.
Nanomaterials 12 02326 g013
Figure 14. DR spectra of Mo-TiO2, Co-Mo-TiO2, Cu-Mo-TiO2, and Zn-Mo-TiO2.
Figure 14. DR spectra of Mo-TiO2, Co-Mo-TiO2, Cu-Mo-TiO2, and Zn-Mo-TiO2.
Nanomaterials 12 02326 g014
Figure 15. DR spectra of W-TiO2, Co-W-TiO2, Cu-W-TiO2, and Zn-W-TiO2.
Figure 15. DR spectra of W-TiO2, Co-W-TiO2, Cu-W-TiO2, and Zn-W-TiO2.
Nanomaterials 12 02326 g015
Figure 16. Tauc plot of all catalysts.
Figure 16. Tauc plot of all catalysts.
Nanomaterials 12 02326 g016
Figure 17. The adsorption capacities of TiO2, Mo-TiO2, and W-TiO2.
Figure 17. The adsorption capacities of TiO2, Mo-TiO2, and W-TiO2.
Nanomaterials 12 02326 g017
Figure 18. The adsorption capacities of Mo-TiO2, Co-Mo-TiO2, Cu-Mo-TiO2, and Zn-Mo-TiO2.
Figure 18. The adsorption capacities of Mo-TiO2, Co-Mo-TiO2, Cu-Mo-TiO2, and Zn-Mo-TiO2.
Nanomaterials 12 02326 g018
Figure 19. The adsorption capacities of W-TiO2, Co-W-TiO2, Cu-W-TiO2, and Zn-W-TiO2.
Figure 19. The adsorption capacities of W-TiO2, Co-W-TiO2, Cu-W-TiO2, and Zn-W-TiO2.
Nanomaterials 12 02326 g019
Figure 20. The photocatalytic activity of TiO2, Mo-TiO2, and W-TiO2 under UV irradiation.
Figure 20. The photocatalytic activity of TiO2, Mo-TiO2, and W-TiO2 under UV irradiation.
Nanomaterials 12 02326 g020
Figure 21. The photocatalytic activity of Mo-TiO2, Co-Mo-TiO2, Cu-Mo-TiO2, and Zn-Mo-TiO2 under UV irradiation.
Figure 21. The photocatalytic activity of Mo-TiO2, Co-Mo-TiO2, Cu-Mo-TiO2, and Zn-Mo-TiO2 under UV irradiation.
Nanomaterials 12 02326 g021
Figure 22. The photocatalytic activity of W-TiO2, Co-W-TiO2, Cu-W-TiO2, and Zn-W-TiO2 under UV irradiation.
Figure 22. The photocatalytic activity of W-TiO2, Co-W-TiO2, Cu-W-TiO2, and Zn-W-TiO2 under UV irradiation.
Nanomaterials 12 02326 g022
Figure 23. PL spectra of TiO2 and Cu-Mo-TiO2.
Figure 23. PL spectra of TiO2 and Cu-Mo-TiO2.
Nanomaterials 12 02326 g023
Figure 24. The photocatalytic activity of Cu-Mo-TiO2 under solar light irradiation.
Figure 24. The photocatalytic activity of Cu-Mo-TiO2 under solar light irradiation.
Nanomaterials 12 02326 g024
Table 1. Mean particle diameter and crystallinity of the photocatalysts.
Table 1. Mean particle diameter and crystallinity of the photocatalysts.
NoPhotocatalystMean (Å)% Crystallinity
1TiO221077.2
2Mo-TiO219981.2
3Co-Mo-TiO224382.1
4Cu-Mo-TiO225177.4
5Zn-Mo-TiO224077.7
6W-TiO225177.4
7Co-W-TiO225072.2
8Cu-W-TiO224173.1
9Zn-W-TiO222468.2
Table 2. Photocatalytic degradation of 4-t-BP by different materials.
Table 2. Photocatalytic degradation of 4-t-BP by different materials.
PhotocatalystsCatalyst
Concentration (mg/L)
4-t-BP
Concentration (mg/L)
Degradation
Efficiency (%)
Treatment Time (min) Light Source and Operation ModeReference
Cu-Mo-TiO21001570150Solar, batchPresent work
Cu-Mo-TiO21001510060UV (365 nm), batchPresent work
Ti2O3/TiO2-650200589.8150Solar, batch[60]
Bi4O5I2100060100 90Visible, batch[61]
Bi12O17Cl2/β-Bi2O3100060100120Visible, batch[13]
0.5% Fe/TiO21000309360UV (254 nm)[15]
4% Fe/TiO2200308760UV (254 nm), continuous flow[62]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mergenbayeva, S.; Kumarov, A.; Atabaev, T.S.; Hapeshi, E.; Vakros, J.; Mantzavinos, D.; Poulopoulos, S.G. Degradation of 4-Tert-Butylphenol in Water Using Mono-Doped (M1: Mo, W) and Co-Doped (M2-M1: Cu, Co, Zn) Titania Catalysts. Nanomaterials 2022, 12, 2326. https://doi.org/10.3390/nano12142326

AMA Style

Mergenbayeva S, Kumarov A, Atabaev TS, Hapeshi E, Vakros J, Mantzavinos D, Poulopoulos SG. Degradation of 4-Tert-Butylphenol in Water Using Mono-Doped (M1: Mo, W) and Co-Doped (M2-M1: Cu, Co, Zn) Titania Catalysts. Nanomaterials. 2022; 12(14):2326. https://doi.org/10.3390/nano12142326

Chicago/Turabian Style

Mergenbayeva, Saule, Alisher Kumarov, Timur Sh. Atabaev, Evroula Hapeshi, John Vakros, Dionissios Mantzavinos, and Stavros G. Poulopoulos. 2022. "Degradation of 4-Tert-Butylphenol in Water Using Mono-Doped (M1: Mo, W) and Co-Doped (M2-M1: Cu, Co, Zn) Titania Catalysts" Nanomaterials 12, no. 14: 2326. https://doi.org/10.3390/nano12142326

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop