A series of Bi3+-doped TiO2 photocatalysts has been prepared via the propylene oxide (PO) assisted sol-gel method. The effect of Bi3+ doping on structural surface morphology and optical properties of the as-prepared photocatalysts was characterized using UV-Visible (UV-Vis) diffuse reflectance spectroscopy, X-ray diffraction, scanning electron microscopy, energy-dispersive X-ray spectroscopy, Brunauer-Emmett-Teller for determination of the specific surface area and porosity, and X-ray photoelectron spectroscopy. Increasing the Bi3+ doping percentage up to 10 mole percent, resulted in all as-prepared photocatalysts exhibiting pure anatase phase upon calcination at 400 °C for 3 hours. A red shift in optical band gap measurements was observed with increasing Bi3+ ion percent doping, which led to extension of the photocatalysts' activity to the visible region. The enhanced photocatalytic activity for removal of the pharmaceutical compound acetaminophen under UV-Vis light irradiation was demonstrated by comparing bismuth doped as-prepared photocatalysts with pure TiO2 photocatalysts prepared under the same conditions. Based on experimental conditions, the highest activity was achieved using 10 mole percent Bi3+-doped photocatalyst, where over a period of 4 hours more than 98% acetaminophen removal was achieved.

Pharmaceutical pollutants tend to be discharged into the aquatic environment from several sources that include: manufacturing facilities (Pérez et al. 2017), disposal of various consumer products of a chemical nature (Douziech et al. 2018), and hospital waste (Verlicchi et al. 2010; Zhang et al. 2010). The presence of these pollutants has negatively impacted both humans and aquatic species (Halpern et al. 2008; Kostich et al. 2014), raising concerns about the effect of the increased level of such pharmaceutical residues on human health and the environment (Jones et al. 2005). Acetaminophen, also known as paracetamol, is the most widely used pharmaceutical in the treatment of fever and headache, and consequently is the most common pharmaceutical residue detected in wastewater (Petrie et al. 2015). The main concern associated with the excessive use of acetaminophen is how easily it accumulates in the aquatic environment due to its characteristics that include high solubility and hydrophilicity (Granberg & Rasmuson 1999). The treatment of such pharmaceutical residue using common chemical treatment processes, e.g. chemical oxidation, results in the production of secondary pollutants resulting from the chemical reaction (Wu et al. 2012). Hence, other more environmentally friendly treatment methods need to be explored.

One such method for the environmental remediation of pollutants is the process of photocatalysis, which involves the use of a semiconductor photocatalyst to carry out redox reactions on its surface, speeding up the rate of chemical reactions (Wu et al. 1999). When light with photons equal to or greater than the band gap of the semiconductor is absorbed, photo-generated electrons and holes migrate to the surface of the semiconductor where redox reactions take place, by which an effective photocatalyst is one in which the charge couple redox potential is within the photocatalyst's band gap range (Ibhadon & Fitzpatrick 2013). For the conduction band, the bottom energy level regulates the reducing potential of the photoelectrons, whereas the valence band top energy level regulates the oxidizing potential of the photo-generated gaps (Ola & Maroto-Valer 2015). Hence, a suitable semiconductor photocatalyst should have appropriate band gap energy. Other characteristics of an ideal photocatalyst include ease of production (Li et al. 2016), stability (Meng et al. 2011; Huang et al. 2013), cost effectiveness (Zhao & Liu 2008), safety for humans and the environment (Lee et al. 2010), easy activation using solar light (Asahi et al. 2001), as well as the capability to effectively catalyze reactions (Qu & Duan 2013). This process of photocatalysis has been widely implemented in pollution control for both water and air, wherein it has the capacity to remove waste and degrade toxic substances into non-toxic forms (Chong et al. 2010; Luengas et al. 2015).

Different photocatalysts have been reported in the literature, including: GaAs, PbS, CdS, ZnO and TiO2 (Zhang et al. 2009; Wang et al. 2013; Etacheri et al. 2015). Among these photocatalysts, GaAs, PbS, and CdS suffer from instability in aqueous media and are toxic (Gupta & Tripathi 2012). As for ZnO, it exhibits lack of stability when dissolved in water where it produces Zn(OH)2 on the surface of the particles, inactivating the catalyst with time (Daneshvar et al. 2007). On the other hand, the characteristics of TiO2 make it an ideal photocatalyst, as it is cheap (Macwan et al. 2011), non-toxic (Choi 2006) and photo-stable in solution (Christy et al. 2009). Other characteristics of TiO2 that are essential for a photocatalyst include high surface area and high porosity, which result in an improved reaction rate because of the improved level of interaction of the reactants with the active sites (Yu et al. 2003). As a result, the use of TiO2 as a photocatalyst has been widely explored in the literature as displaying promise in environmental remediation applications (Pelaez et al. 2012).

TiO2 has three main phases: anatase, rutile and brookite. The anatase phase is the most active form, used in photocatalysis applications (Zhang et al. 2000). However, a disadvantage of this photocatalyst is its lack of visible light absorption due to its large band gap energy of 3.2 eV, thus limiting its application under visible light irradiation (Yin et al. 2009; Landmann et al. 2012). To overcome this drawback, different methods have been reported in the literature to extend its spectral response to visible light, including metal (Fiszka Borzyszkowska et al. 2016) and non-metal (Burda et al. 2003) doping, coupling of TiO2 with other semiconductor materials (Liu et al. 2007a), fabrication of nanocomposite photocatalysts with different morphologies (Liu et al. 2007b; Wang et al. 2007; Zhong et al. 2014; Chen et al. 2017), and dye sensitization (Park et al. 2000). Among these methods, doping has been found to be an effective way to extend the light absorption to the visible region by introducing additional energy levels between the valence band and conduction band of the photocatalyst acting as a trap for the charge carriers, facilitating their separation from the bands. This in turn allows more charge carriers to successfully diffuse to the surface of the photocatalyst (Kim et al. 2005; Yang et al. 2010).

Bismuth based composite oxide semiconductors such as bismuth titanate (Feng Yao et al. 2003) show enhanced photocatalytic activity under visible light irradiation, which can be attributed to their narrow band gaps compared to TiO2. In such a case, band gap reduction is attributed to the hybridized valence band of Bi+3 6S lying above the O 2p and the Ti4+ 3d level just below the Bi3+ 6p in the conduction band. As a result, band gap energy reduction is assumed to be due to the excitation of a 6S electron of Bi3+ into the 3d of Ti4+. This behavior has been observed in other bismuth-based materials such as bismuth vanadate (Huang et al. 2014), bismuth molybdate (Shimodaira et al. 2006) and bismuth tungstate (Zhang & Zhu 2012). In these semiconductors, the hybrid valence bands formed between O 2p and Bi3+ 6S levels and conduction band between V; 3d, Mo; 4d and W; 5d orbitals respectively and the Bi3+ 6p (Walsh et al. 2009). Accordingly, the absorption extends to a longer wavelength due to the excitation of a 6S electron of Bi3+ into the d orbital representing the metal ion. Thus, incorporating Bi3+ ion into the TiO2 band structure is anticipated to have a similar effect, shifting light absorption to the longer wavelengths, i.e. visible region. This in turn induces high photocatalytic activity under visible light irradiation of the bismuth doped photocatalysts.

In this work, we present a facile propylene oxide (PO) assisted sol-gel preparation of TiO2 and bismuth-doped TiO2. The performance of the as-prepared photocatalysts was evaluated by the removal of acetaminophen from an aqueous solution over a period of 4 hours under UV and visible light irradiation.

Materials

Titanium (IV) n-butoxide (Ti(OnBu)4, 97%), Bi(NO3)3.5H2O, 2-propanol (i-PrOH, 99.7%), hydrochloric acid (HCl, 37%), acetylacetone (acac, 99.3%) and propylene oxide (PO, 99%) were purchased from Sigma Aldrich and used as received. All chemicals were used without further purification. Solutions were prepared using doubly distilled water passed through a Milli-Q apparatus.

Preparation of the photocatalysts

Preparation of TiO2

The TiO2 photocatalyst was prepared through mixing Ti(OnBu)4 solution and the hydrolysis solution that were prepared as indicated below.

Ti(OnBu)4 solution: The TiO2 photocatalyst was synthesized using Ti(OnBu)4, 2-propanol (i-PrOH) used as solvent, acetylacetone (acac) as the complexing agent and PO as the gelation promoter using the sol-gel method at room temperature. In a typical synthesis, Ti(OnBu)4 was dissolved in i-PrOH in a i-PrOH:Ti(OnBu)4 v/v ratio of 5:1 where acac was added in a molar ratio Ti(OnBu)4: acac of 1:2. The prepared solution was kept under constant stirring for 2 hours followed by the addition of PO where the PO: Ti(OnBu)4 molar ratio was 10:1.

The hydrolysis solution: The hydrolysis solution was made by mixing distilled water and i-PrOH (i-PrOH:H2O, 1:20) where the H2O:Ti(OnBu)4 molar ratio was 10:1. The hydrolysis reaction was initiated through the addition of Ti(OnBu)4 solution to the hydrolysis solution drop wise under vigorous stirring. The pH of the final solution was adjusted to pH = 2 through the addition of 37% HCl. This reaction mixture was stirred for an additional period of 3 hours and then aged for a few days for the formation of the gel. After drying, powder was air-calcinated at 400 °C for 3 hours.

Preparation of Bi3+-doped TiO2 photocatalysts

For the preparation of the Bi3+-doped TiO2 photocatalysts, the same procedure as that for the preparation of TiO2 photocatalyst was implemented, where molar amounts of Bi were added. Samples of 1%, 3%, 5%, and 10% of Bi-doped TiO2 were prepared, respectively, where Bi was first dissolved in a i-PrOH:H2O v/v ratio of 1:5 followed by the addition of a few drops of 37% HCl. The Bi3+ solution was then added to the Ti(OnBu)4 solution prior to the PO addition. Finally, hydrolysis solution was added dropwise under vigorous stirring. All as-prepared Bi-doped photocatalysts were air-calcinated at 400 °C for 3 hours.

Characterization of the photocatalysts

UV-Visible diffuse reflectance spectroscopy (UV-Vis DRS)

UV-Vis diffuse reflectance spectroscopy measurements were obtained using a Shimadzu UV-3600 UV-Vis spectrophotometer from 200 to 800 nm where the baseline was corrected using BaSO4 as a reference standard. The band gap energies were calculated using the results from the spectra and the application of the Tauc Plot method (Tauc 1970). The band gap was calculated considering that these photocatalysts were indirect semiconductors and using Equation (1), where α, Eg, h, ν, A, n represent the absorption coefficient, band gap, Planck's constant, frequency of light, a constant, and the number characterizing transition for indirect semiconductor materials (n = 1/2 for TiO2), respectively (Zhang et al. 2014).
(1)

According to the above equation, through the construction of a plot of (αhν)2 versus (eV), the band gap energy Eg can be obtained by extrapolating a line through the steep linear part of the curve to the hν axis as shown in Figure 1, whereas the band gap values are tabulated in Table 1.

Table 1

Physical structure parameters of as-prepared photocatalysts

Catalyst (% Bi doping)Band gap energy (eV)SBET (m2 g−1)Pore size (nm)Vpores (cm3 g−1)Crystalline size (nm)Apparent rate constant (h−1)Lattice constant (Å)
EDS (Atom %)
a = bcTiOBi
3.08 125.36 7.17 0.138 73 0.58 3.795 9.786 30.18 69.82 
2.91 33.78 7.74 0.0595 105 0.63 3.793 9.754 30.61 69.12 0.27 
2.87 39.27 7.92 0.139 112 0.73 3.796 9.773 27.27 71.78 0.95 
2.84 42.86 11.87 0.120 125 0.72 3.791 9.642 29.93 68.75 1.31 
10 2.80 60.00 11.29 0.181 131 0.97 3.776 9.345 30.25 68.19 1.56 
Catalyst (% Bi doping)Band gap energy (eV)SBET (m2 g−1)Pore size (nm)Vpores (cm3 g−1)Crystalline size (nm)Apparent rate constant (h−1)Lattice constant (Å)
EDS (Atom %)
a = bcTiOBi
3.08 125.36 7.17 0.138 73 0.58 3.795 9.786 30.18 69.82 
2.91 33.78 7.74 0.0595 105 0.63 3.793 9.754 30.61 69.12 0.27 
2.87 39.27 7.92 0.139 112 0.73 3.796 9.773 27.27 71.78 0.95 
2.84 42.86 11.87 0.120 125 0.72 3.791 9.642 29.93 68.75 1.31 
10 2.80 60.00 11.29 0.181 131 0.97 3.776 9.345 30.25 68.19 1.56 
Figure 1

UV-Vis diffuse reflectance spectra of pure (a) and different % Bi3+ doped TiO2 as-prepared photocatalysts (b, c).

Figure 1

UV-Vis diffuse reflectance spectra of pure (a) and different % Bi3+ doped TiO2 as-prepared photocatalysts (b, c).

Close modal

X-ray diffraction (XRD)

X-ray diffraction characterization was performed using a Shimadzu-6100 powder XRD diffractometer with Cu-Kα radiation. The working voltage and current of the X-ray tube were 40 kV and 40 mA respectively, where λ = 1.542 Å. The diffraction data were collected in the 2θ angle range of 20–80 degrees at a rate of 2 degrees/min. Figure 2 displays the XRD results for all the photocatalysts used in this study.

Figure 2

XRD chart of the as-prepared % Bi doped TiO2 photocatalysts compared with pure TiO2.

Figure 2

XRD chart of the as-prepared % Bi doped TiO2 photocatalysts compared with pure TiO2.

Close modal

Scanning electron microscopy (SEM) and energy-dispersive X-ray spectroscopy (EDS)

Scanning electron microscope images and energy-dispersive X-ray spectroscopy were obtained using JEOL JSM–6010LA. After film preparation on ITO-coated glass, samples were rinsed and allowed to dry. Images and data were then collected at accelerating voltage of 20 kV as presented in Figures 3 and 4 and Table 1.

Specific surface area and porosity

Surface area and porosity were characterized using N2 adsorption at 77 K using a Quantachrome Autosorb-1 volumetric gas sorption instrument. Before measurements, samples were degassed at 150 °C for 1 hour. Brunauer-Emmett-Teller (BET) theory was used to calculate surface area, and pore size distributions were determined by the Barett-Joyner-Halenda (BJH) model based on the desorption branch of the N2 isotherms. Specific surface area and porosity data analysis are presented in Table 1.

X-ray photoelectron spectroscopy (XPS)

X-ray photoelectron spectroscopy measurements of all as-prepared photocatalysts were performed using the Kratos Axis Ultra DLD spectrometer (Kratos Analytical Ltd, Manchester, UK) with Al Kα1 X-ray source. The energy of an X-ray photon of 1.486 keV with pass energy of 160 eV was used for the survey spectrum and 20 eV for narrow scans. All spectra were collected at a 54° take-off angle and analyzed area of (700 × 300 μm) using the combination of electrostatic and magnetic lens (hybrid mode). The C 1 s peak from adventitious hydrocarbon at 284.8 eV was used as an energy reference to correct for charging. Surface charging effects were minimized using a charge balance operating at 3.6 V and 1.8 V maintained as filament bias. During XPS analysis, sample charging was neutralized using an electron flood gun. All spectra were recorded under ultra-high vacuum conditions below 5 × 10−10 mbar. All data processing was carried out using the Casa XPS software (Casa Software Ltd) package. For calculation of the elemental composition of the prepared photocatalysts, the determined areas under the peaks were normalized using the Scofield sensitivity factors. For data evaluation, the background of the Auger photoelectron peaks was subtracted by applying a Shirley-type of background and the data were curve-resolved using an 80% Gaussian/20% Lorentzian sum.

Photocatalytic activities

The photocatalytic evaluation of Bi3+-doped TiO2 as-prepared photocatalysts in the removal of acetaminophen in aqueous solution was performed using a Luzchem Photoreactor (Mod. LZC-1, Luzchem Research Inc., ON, CAN) equipped with two UVA lamps centered at ∼350 nm, two UVB lamps centered at ∼300 nm with a peak of 313 nm, two UVC (254 nm) germicidal lamps, and seven cool white fluorescent tubes. Temperature was kept constant at 25 °C for all the reactions. In a typical photocatalytic run, powdered photocatalyst in the amount of 0.1 gL−1 was suspended in aqueous solution of 10−4 M acetaminophen in a 100-mL quartz beaker. The suspension was magnetically stirred in the dark over 30 min to achieve a complete adsorption/desorption equilibrium and then it was irradiated with a UV-Vis light source. The light intensity was measured using a digital light meter (TES-1330A) and it was 11,460 lux (44 W/cm2). Aliquots of the reaction medium were periodically sampled and filtered using a polytetrafluoroethylene (PTFE) membrane filter (Millipore, 0.22 μm) prior to analysis. The acetaminophen concentration was measured employing a spectrophotometric method by using a double beam UV-Vis SPECORD 210 PLUS spectrophotometer and following the decrease in the absorbance of acetaminophen at 250 nm over time definite intervals.

Characterization results

Band gap UV-Vis DRS

Tauc plots of the UV-Vis DRS spectra of the pure TiO2 and bismuth-doped TiO2 as-prepared photocatalysts are presented in Figure 1. Pure TiO2 exhibited a direct allowed band gap of 3.08 eV. On the other hand, all bismuth doped as-prepared photocatalysts showed quite similar light absorption shifts towards the visible region of the spectrum, with the lowest band gap of 2.8 eV for the 10% Bi3+-doped TiO2 photocatalyst, i.e. a red shift of approximately 40 nm from that of pure TiO2. This red shift observed for all doped as-prepared photocatalysts can be attributed to the presence of bismuth ion within the TiO2 lattice and lack of presence as a secondary phase as indicated solely by the XRD spectra presented in Figure 2.

XRD

Figure 2 displays the XRD patterns of the as-prepared pure and bismuth-doped TiO2 photocatalysts with varying amounts of Bi3+ dopant. From the figure, it can be observed that the peaks for the doped TiO2 samples coincide with those of the pure TiO2 in which all the patterns displayed indicate the presence of the pure anatase TiO2 phase with no peaks corresponding to another phase or newly formed lattice structure. Furthermore, the fact that there was no observed shift in the diffraction angles of the XRD peaks indicates that doped Bi3+ ions did not cause any changes in the phase composition to the TiO2 lattice structure at the studied doping level. The fact that Bi3+ species were incorporated into the TiO2 lattice without changing its phase does not necessarily rule out the possibility that bismuth may present as a separate phase onto the surface of TiO2, as in the case of Bi2O3 as claimed by Wang et al. (2008). Similar results have been reported by Yang et al. in the work performed on TiO2 photocatalysts prepared through coupling with Bi2O3 and doping with Si (Yang et al. 2014). However, in this study, the XRD peaks of the as-prepared photocatalysts corresponding to Bi2O3 were not found even when samples were calcinated at 700 °C. Therefore, the absence of this compound is not conclusive, since it is believed that presence of minor amounts of Bi2O3 cannot be detected by XRD (Ji et al. 2009).

Using the XRD data, the average crystalline sizes of the as-prepared photocatalysts can be calculated by applying the Debye-Scherrer formula on the anatase (101) diffraction peaks (Patterson 1939).
(2)
where D is the crystalline size, λ the wavelength of X-ray radiation (0.1541 nm), K a constant 0.89, and β the peak width (in radians) at half maximum height after the subtraction of the equipment broadening, 2θ = 25.4 for anatase phase of TiO2. From the XRD results, the intensity of this characterized peak increased whilst becoming narrower as the mol % of Bi3+ doping was increased. In general, the crystal size of the as-prepared photocatalyst increased with increase in mol % of Bi dopant. The results of the crystalline size in nm are presented in Table 1.
The lattice constants of pure and Bi3+-doped as-prepared TiO2 photocatalysts were calculated using the formula,
(3)
where d is the interplanar spacing in XRD pattern, a and c are the lattice constants in angstrom (Å) and h, k and l are the Miller indices. The calculated values of lattice constant (a = b, and c) for the pure and Bi3+-doped TiO2 as-prepared photocatalysts are presented in Table 1. The values are in good agreement with the literature values of a = b = 3.785 Å and c = 9.504 Å for pure TiO2 anatase nanoparticles reported by Milićević et al. (2017). Moreover, these values were found to decrease with increasing concentration of Bi3+ ions in the as-prepared TiO2 anatase photocatalysts. This can be explained based on the variation in the ionic radius of Bi3+ (0.74 Å) and Ti4+ (0.68 Å) in the as-prepared pure and Bi3+-doped TiO2 photocatalysts, thereby affecting the unit cell and thus resulting in the smaller lattice constants (Li et al. 2004; Štengl & Bakardjieva 2010).

SEM

SEM images of pure and Bi3+-doped as-prepared photocatalysts are depicted in Figure 3; the pure and bismuth-doped TiO2 photocatalyst particles seem to be irregular in shape. These particles displayed a rock-like morphology, where the sizes of the particles of the doped samples seemed to be larger than the un-doped sample under the same magnification. This can be attributed to the presence of particle agglomeration. Similar agglomeration results were reported by Aware & Jadhav (2016) in the work performed on the doping of TiO2 nanoparticles with zinc using the same method used in the current study, the sol-gel method.

Figure 3

SEM 20 (μm) images of different % Bi-doped TiO2 photocatalysts. (a) 0%, (b) 1%, (c) 3%, (d) 5% and (e) 10%.

Figure 3

SEM 20 (μm) images of different % Bi-doped TiO2 photocatalysts. (a) 0%, (b) 1%, (c) 3%, (d) 5% and (e) 10%.

Close modal

EDS analysis

EDS analysis was used to investigate the quantitative chemical composition for the constituent elements of pure and bismuth-doped TiO2 as-prepared photocatalysts. Figure 4(a)–4(e) show the binding energy (BE) (keV) in the x-axis versus intensity of X-rays (counts) in the y-axis of the EDS analysis spectrum of as-prepared Bi3+-doped TiO2 photocatalysts. Based on the EDS spectra, the existence of all corresponding elements, i.e. the peaks of Bi, Ti and O, respectively, present in the as-prepared photocatalysts were confirmed. A carbon peak is also present in the EDS spectra, which is due to the use of carbon tape during analysis. Moreover, the increased carbon content in the EDS spectra may also be attributed to possible contamination due to the hydrocarbon used in the preparation method. Hence, it is tempting to say that the presence of carbon does not dramatically affect as-prepared photocatalysts' performance, since it is present in all photocatalysts in almost similar amounts. No additional impurity peaks were detected, which confirms phase purity of the as-prepared photocatalysts, as further supported by XRD and XPS analysis.

Figure 4

EDS spectrum of Bi-doped TiO2. (a) 0%, (b) 1%, (c) 3%, (d) 5% and (e) 10% samples, insert spectrum shows the elemental composition of corresponding EDS spectrum.

Figure 4

EDS spectrum of Bi-doped TiO2. (a) 0%, (b) 1%, (c) 3%, (d) 5% and (e) 10% samples, insert spectrum shows the elemental composition of corresponding EDS spectrum.

Close modal

N2 adsorption-desorption analysis

The N2 adsorption-desorption isotherms for the pure and Bi3+-doped TiO2 as-prepared photocatalysts are presented in Figure 5, with the surface area, pore size and volume of the pores as presented in Table 1. All as-prepared photocatalysts were characterized by type IV isotherms, which indicates the presence of mesoporous material. The surface areas for the 1%, 3%, 5% and 10% Bi-doped photocatalysts were reported at 33.78, 38.27, 42.86 and 60 m2/g respectively, compared to the pure TiO2 photocatalyst's surface area of 125.3 m2/g. According to Table 1, the change caused by Bi3+ doping is clear in the 10 mol % Bi-doped photocatalyst exhibiting the highest surface area (60 m2/g). Moreover, as Bi3+ dopant increased, the pore size became larger and the surface area was reduced to a large extent compared with pure TiO2.

Figure 5

BET isotherms of un-doped and % Bi-doped TiO2 photocatalysts.

Figure 5

BET isotherms of un-doped and % Bi-doped TiO2 photocatalysts.

Close modal

XPS analysis

The elemental composition and valence state of the as-prepared photocatalysts were analyzed by XPS study. The XPS results are presented in Figure 6 and Table 2. The elements Ti, Bi and O in the as-prepared photocatalysts were confirmed in the wide spectrum. It also shows carbon; the high C 1 s content is due to the hydrocarbon contamination introduced from the laboratory environment as shown from EDS analysis. Thus, the observed C 1 s peak was used as an energy reference for determining the peak positions of core level spectra.

Table 2

Data obtained from wide XPS spectra for Ti1-xBixO2 (x = 0, x = 0.01, x = 0.03, x = 0.05, x = 0.10)

SamplePeakPosition (eV)Area% Atomic Conc.a
x = 0 O 1s 528.00 1,860.09 45.21 
 C 1s 283.00 1,541.59 37.47 
 Ti 2p 456.50 712.881 17.33 
x = 0.01 Bi 4f 162.50 0.559001 0.01 
 Ti 2p 456.50 544.509 13.45 
 O 1s 528.00 1,640.05 40.52 
 C 1s 283.00 1,862.03 46.01 
x = 0.03 Bi 4f 157.50 114.038 2.84 
 Ti 2p 457.00 754.015 18.76 
 O 1s 528.00 1,841.17 45.81 
 C 1s 283.50 1,310.07 32.59 
x = 0.05 O 1s 528.00 1,688.17 44.56 
 C 1s 283.00 1,249.55 32.98 
 Ti 2p 456.50 711.049 18.77 
 Bi 4f 157.00 139.736 3.69 
x = 0.10 O 1s 528.00 1,488.07 39.73 
 C 1s 283.00 1,630.95 43.55 
 Ti 2p 456.50 547.426 14.62 
 Bi 4f 157.00 78.8011 2.10 
SamplePeakPosition (eV)Area% Atomic Conc.a
x = 0 O 1s 528.00 1,860.09 45.21 
 C 1s 283.00 1,541.59 37.47 
 Ti 2p 456.50 712.881 17.33 
x = 0.01 Bi 4f 162.50 0.559001 0.01 
 Ti 2p 456.50 544.509 13.45 
 O 1s 528.00 1,640.05 40.52 
 C 1s 283.00 1,862.03 46.01 
x = 0.03 Bi 4f 157.50 114.038 2.84 
 Ti 2p 457.00 754.015 18.76 
 O 1s 528.00 1,841.17 45.81 
 C 1s 283.50 1,310.07 32.59 
x = 0.05 O 1s 528.00 1,688.17 44.56 
 C 1s 283.00 1,249.55 32.98 
 Ti 2p 456.50 711.049 18.77 
 Bi 4f 157.00 139.736 3.69 
x = 0.10 O 1s 528.00 1,488.07 39.73 
 C 1s 283.00 1,630.95 43.55 
 Ti 2p 456.50 547.426 14.62 
 Bi 4f 157.00 78.8011 2.10 

aPercentage contribution of each peak to the total number of counts in Ti 2p, Bi 4f, O 1 s and C 1 s peak.

Figure 6

XPS wide spectrum of synthesized Ti1-xBixO2 (a) x = 0, (b) x = 0.01, (c) x = 0.03, (d) x = 0.05, (e) x = 0.10 samples.

Figure 6

XPS wide spectrum of synthesized Ti1-xBixO2 (a) x = 0, (b) x = 0.01, (c) x = 0.03, (d) x = 0.05, (e) x = 0.10 samples.

Close modal

For more information about the valence state of titanium and bismuth ions – Ti 2p and Bi 4f – present in the as-prepared photocatalysts, high-resolution spectra were analyzed with respect to Bi3+ doping level as shown in Figures 7 and 8. Table 3 displays the fitted peak position corresponding to the BE values of Ti 2p spectra using the Lorentzian-Gaussian curve fitting method. For the pristine TiO2, Ti 2p peak is distinguishable as two peaks at 462.32 eV for 2p3/2 and 456.59 eV for 2p1/2, which correspond to the characteristics peak of Ti4+ ion (Bapna et al. 2011). After doping bismuth into the TiO2, the deconvolution of these peaks shifted towards lower binding energies of about 0.17 eV and peak width increased, indicating a small contribution of Ti3+ ion present in the sample. The shifts of these peaks are consistent with what was previously reported by Liu et al. (2012). These results may indicate an oxygen deficiency in TiO2 lattice (Hamdy et al. 2012). It is conceivable that bismuth cations are substituted for Ti ion in the Ti1-xBixO2 matrix, because of the decreased Ti 2p peak area in comparison to the pure counterpart.

Table 3

Data obtained from high resolution Ti 2p XPS spectra for Ti1-xBixO2 (x = 0, x = 0.01, x = 0.03, x = 0.05, x = 0.10) samples

SampleTi 2p3/2Ti 2p1/2zΔE Ti 2p
BEx (eV)FWHMy (eV)Area (%)BExFWHMy (eV)Area (%)
x = 0 458.65 1.45 69.34 464.32 2.13 30.66 5.67 
x = 0.01 458.59 1.54 67.06 464.24 2.38 32.94 5.65 
x = 0.03 458.47 2.12 63.63 464.21 3.41 36.37 5.74 
x = 0.05 458.41 1.58 62.90 464.16 2.61 37.10 5.75 
x = 0.10 458.29 1.59 40.10 464.11 2.61 59.90 5.82 
SampleTi 2p3/2Ti 2p1/2zΔE Ti 2p
BEx (eV)FWHMy (eV)Area (%)BExFWHMy (eV)Area (%)
x = 0 458.65 1.45 69.34 464.32 2.13 30.66 5.67 
x = 0.01 458.59 1.54 67.06 464.24 2.38 32.94 5.65 
x = 0.03 458.47 2.12 63.63 464.21 3.41 36.37 5.74 
x = 0.05 458.41 1.58 62.90 464.16 2.61 37.10 5.75 
x = 0.10 458.29 1.59 40.10 464.11 2.61 59.90 5.82 

xBinding energies, yfull width at half maximum and zΔE Ti 2p is the binding energy difference between Ti 2p3/2 and Ti 2p1/2 peaks.

Figure 7

XPS wide spectrum of synthesized Ti1-xBixO2 (a) x = 0, (b) x = 0.01, (c) x = 0.03, (d) x = 0.05, (e) x = 0.10 samples.

Figure 7

XPS wide spectrum of synthesized Ti1-xBixO2 (a) x = 0, (b) x = 0.01, (c) x = 0.03, (d) x = 0.05, (e) x = 0.10 samples.

Close modal
Figure 8

XPS wide spectrum of synthesized Ti1-xBixO2 (a) x = 0.01, (b) x = 0.03, (c) x = 0.05, (d) x = 0.10 samples.

Figure 8

XPS wide spectrum of synthesized Ti1-xBixO2 (a) x = 0.01, (b) x = 0.03, (c) x = 0.05, (d) x = 0.10 samples.

Close modal

It is seen from Figure 8 that the Bi 4f peak is split into two peaks located at ∼159.06 eV and ∼164.41 eV corresponding to Bi 4f5/2 and Bi 4f7/2 peaks, as fitted by the Gaussian-Lorentzian product function after Shirley-type background subtraction. Both are mainly assigned to Bi-O bonds. Spin orbital splitting energy of the Bi 4f doublet is equal to ∼5.35 eV, which is a slight increase of about 0.15 eV with increase in the bismuth doping content x in the sample. Furthermore, each of the Bi 4f5/2 and Bi 4f7/2 peaks could be fitted by two sub-peaks. The sub-peaks located at 159.19 and 158.24 eV are assigned to Bi 4f7/2-O, whereas 166.5 and 164.62 eV are of Bi 4f5/2-O bonds for Bi3+ (Mekki et al. 1996). Table 4 displays the fitted Bi 4f XPS peak of the samples.

Table 4

Fitting parameters of deconvoluted Bi 4f XPS spectra for Ti1-xBixO2 (x = 0.01, x = 0.03, x = 0.05, x = 0.10) samples

SampleBi 4f7/2
ΔE1Bi 4f5/2
ΔE2
Bi 4f7/2(1)
Bi 4f7/2(2)
Bi 4f5/2(1)
Bi 4f5/2(2)
B.ExFWyAzB.ExFWyAzB.ExFWyAzB.ExFWyAz
x= 0.01 159.19 1.36 176.107 158.24 1.91 60.914 0.95 166.50 1.45 164.954 164.62 1.81 35.0319 1.88 
x= 0.03 160.00 3.79 1,115.16 158.35 1.57 641.652 1.65 166.54 2.76 584.831 164.79 1.73 511.807 1.75 
x= 0.05 159.21 1.62 1,999.34 158.20 1.72 71.3654 1.01 166.62 1.58 1,271.75 164.09 1.35 232.819 2.53 
x= 0.10 159.18 1.49 1,022.83 158.48 1.74 99.6585 0.7 166.47 1.53 810.443 164.53 1.43 57.468 1.94 
SampleBi 4f7/2
ΔE1Bi 4f5/2
ΔE2
Bi 4f7/2(1)
Bi 4f7/2(2)
Bi 4f5/2(1)
Bi 4f5/2(2)
B.ExFWyAzB.ExFWyAzB.ExFWyAzB.ExFWyAz
x= 0.01 159.19 1.36 176.107 158.24 1.91 60.914 0.95 166.50 1.45 164.954 164.62 1.81 35.0319 1.88 
x= 0.03 160.00 3.79 1,115.16 158.35 1.57 641.652 1.65 166.54 2.76 584.831 164.79 1.73 511.807 1.75 
x= 0.05 159.21 1.62 1,999.34 158.20 1.72 71.3654 1.01 166.62 1.58 1,271.75 164.09 1.35 232.819 2.53 
x= 0.10 159.18 1.49 1,022.83 158.48 1.74 99.6585 0.7 166.47 1.53 810.443 164.53 1.43 57.468 1.94 

xBinding energies, yfull width at half maximum, zcorresponding area of the peaks, and ΔE1 is the binding energy between Bi 4f7/2(1) and Bi 4f7/2(2) whereas ΔE2 is the binding energy between Bi 4f5/2(1) and Bi 4f5/2(2).

Figure 9 and Table 5 show the O 1 s XPS peak of the Ti1-xBixO2 sample. The O 1 s XPS spectra are well fitted into two peaks situated at 529.88 eV and 531.57 eV. The BE at 529.88 eV is attributed to the contribution of crystal lattice oxygen, whereas the BE at 531.57 eV is called a shoulder peak, which arises due to chemisorbed oxygen species on the surface of samples (Xu et al. 2008). Based on the above results, XPS observations confirm that no additional/impurity phase was formed, which further supports the EDS and XRD results.

Table 5

Data obtained from high resolution O 1 s XPS spectra for Ti1-xBixO2 (x = 0, x = 0.01, x = 0.03, x = 0.05, x = 0.10) samples

SampleO 1 s (1) (eV)O 1 s (2) (eV)z ΔO 1 s
BExFWHMyArea (%)BExFWHMyArea (%)
x = 0 529.89 1.53 4,430.84 531.58 2.24 1,113.13 1.69 
x = 0.01 529.63 1.48 3,229.49 531.65 2.87 1,640.25 2.02 
x = 0.03 529.79 1.51 4,049.25 531.37 2.48 1,442.86 1.58 
x = 0.05 529.58 1.59 4,101.56 531.26 2.12 927.752 1.68 
x = 0.10 529.84 1.49 3,038.75 531.74 2.78 1,376.96 1.9 
SampleO 1 s (1) (eV)O 1 s (2) (eV)z ΔO 1 s
BExFWHMyArea (%)BExFWHMyArea (%)
x = 0 529.89 1.53 4,430.84 531.58 2.24 1,113.13 1.69 
x = 0.01 529.63 1.48 3,229.49 531.65 2.87 1,640.25 2.02 
x = 0.03 529.79 1.51 4,049.25 531.37 2.48 1,442.86 1.58 
x = 0.05 529.58 1.59 4,101.56 531.26 2.12 927.752 1.68 
x = 0.10 529.84 1.49 3,038.75 531.74 2.78 1,376.96 1.9 

xBinding energies, yfull width at half maximum and zΔO 1 s is the difference in binding energies between O 1 s (1) and O 1 s (2) peaks in (eV).

Figure 9

Deconvolution O 1 s XPS spectra of synthesized Ti1-xBixO2 (a) x = 0, (b) x = 0.01, (c) x = 0.03, (d) x = 0.05, (e) x = 0.10 samples.

Figure 9

Deconvolution O 1 s XPS spectra of synthesized Ti1-xBixO2 (a) x = 0, (b) x = 0.01, (c) x = 0.03, (d) x = 0.05, (e) x = 0.10 samples.

Close modal

The photocatalytic activity of the as-prepared pure and Bi3+-doped photocatalysts was evaluated in terms of their rate of acetaminophen removal from aqueous solution at pH = 5. The results indicated a direct relation between the percentages of Bi3+-doped TiO2 and the removal rate of the pharmaceutical as presented in Figures 1013. Upon comparing the photocatalytic removal of acetaminophen, 10 mol % Bi3+-doped photocatalyst's performance exceeded that of the pure as-prepared TiO2. Figure 11 presents the graph of C/Co vs. time by which the variation of the concentration of acetaminophen is observed for the different photocatalysts under UV-Vis light with irradiation time up to 4 hours. Although the surface areas of as-prepared Bi-doped photocatalysts are significantly lower than that of pure TiO2, their enhanced photocatalytic activities might be due to their suitable band gaps absorption that occurs in the visible light region.

Figure 10

UV-Vis removal of acetaminophen from aqueous solutions of different % Bi3+-doped TiO2 photocatalysts. (a) 0%, (b) 1%, (c) 3%, (d) 5% and (e) 10%.

Figure 10

UV-Vis removal of acetaminophen from aqueous solutions of different % Bi3+-doped TiO2 photocatalysts. (a) 0%, (b) 1%, (c) 3%, (d) 5% and (e) 10%.

Close modal
Figure 11

Variation of the concentration of acetaminophen over time under UV-Vis irradiation in the presence of different % Bi3+ dopants.

Figure 11

Variation of the concentration of acetaminophen over time under UV-Vis irradiation in the presence of different % Bi3+ dopants.

Close modal
Figure 12

Kinetic plot of −ln(C0/C) over time for the disappearance of acetaminophen for the different % Bi3+-doped TiO2.

Figure 12

Kinetic plot of −ln(C0/C) over time for the disappearance of acetaminophen for the different % Bi3+-doped TiO2.

Close modal
Figure 13

Rate constants for kinetic studies of different % Bi3+-doped photocatalysts.

Figure 13

Rate constants for kinetic studies of different % Bi3+-doped photocatalysts.

Close modal

The photocatalytic removal of acetaminophen follows pseudo-first-order reaction kinetics for the concentration range of the present study (Dalida et al. 2014). As a result, the variation in -ln(Co/C) as a function of irradiation time is reported in Figure 12, by which the rate constants were calculated in Figure 13 and listed in Table 1. The observed photocatalytic activity of Bi3+-doped as-prepared photocatalysts follows the order: 10 > 5 > 3 > 1> pure TiO2, where enhanced removal of acetaminophen from solution under UV-Vis light irradiation is mainly ascribed to the Bi3+ doping. The doping with Bi3+ ion into the lattice of TiO2 reduces the band gap energy. Therefore, the excitation energy is extended to the visible region. In our studied doping range, results from photocatalytic experiments are found to be in agreement with that of UV-Visible diffuse reflectance spectra analysis. Generally, the increase in mol % of Bi3+ doping led to a red shift increase, thereby resulting in higher photocatalytic activity. Moreover, as evident from the enhanced photocatalytic performance of Bi3+-doped photocatalysts in comparison to the un-doped photocatalysts, the presence of Bi3+ ions in the TiO2 lattice significantly influences photoactivity by reducing charge carrier recombination rates during the irradiation process.

In conclusion, mesoporous Bi3+-doped anatase TiO2 photocatalysts have been prepared using the PO assisted sol-gel method. The characterization results displayed suggest that the Bi3+ ion conceivably caused substitutional changes to the crystal lattice without altering its anatase phase. All doped as-prepared photocatalysts were found to have enhanced photocatalytic activities towards the removal of acetaminophen in aqueous solution under UV-Vis solar light irradiation compared to pure TiO2. The enhanced photocatalytic activity indicated that Bi3+ doping into the TiO2 enhanced the visible light absorption, as compared to the pure TiO2. The rate of the photocatalytic reactions was found to depend mainly on the percent doping of Bi3+ ion, with the highest activity achieved for the highest doping percent of 10% added within the range of this study.

This research project was financially supported by the United Arab Emirates University ‘SURE Plus’ (grant no. 31S292, Ahmed Alzamly). The authors of this paper would like to acknowledge Mr Bassam Al-Hindawi for his assistance and guidance in the performance of BET analysis as well as Ms Afra Gharib Al-Blooshi for her cooperation with obtaining the SEM images.

Asahi
R.
,
Morikawa
T.
,
Ohwaki
T.
,
Aoki
K.
&
Taga
Y
, .
2001
Visible-light photocatalysis in nitrogen-doped titanium oxides
.
Sci
.
293
(
5528
),
269
271
.
Bapna
K.
,
Phase
D.
&
Choudhary
R
, .
2011
Study of valence band structure of Fe doped anatase TiO2 thin films
.
J. Appl. Phys
.
110
(
4
),
043910
.
Burda
C.
,
Lou
Y.
,
Chen
X.
,
Samia
A.
,
Stout
J.
&
Gole
J
, .
2003
Enhanced nitrogen doping in TiO2 nanoparticles
.
NANO Lett
.
3
(
8
),
1049
1051
.
Chong
M.
,
Jin
B.
,
Chow
C.
&
Saint
C
, .
2010
Recent developments in photocatalytic water treatment technology: a review
.
Water Res
.
44
(
10
),
2997
3027
.
Christy
P.
,
Jothi
N.
,
Melikechi
N.
&
Sagayaraj
P
, .
2009
Synthesis, structural and optical properties of well dispersed anatase TiO2 nanoparticles by non-hydrothermal method
.
Cryst. Res. Technol
.
44
(
5
),
484
488
.
Dalida
M.
,
Amer
K.
,
Su
C.
&
Lu
M
, .
2014
Photocatalytic degradation of acetaminophen in modified TiO2 under visible irradiation
.
Environ. Sci. Pollut. Res.
21
(
2
),
1208
1216
.
Daneshvar
N.
,
Aber
S.
,
Sayed Dorraji
M.
,
Khataee
A.
&
Rasoulifard
M
, .
2007
Photocatalytic degradation of the insecticide diazinon in the presence of prepared nanocrystalline ZnO powders under irradiation of UV-C light
.
Sep. Purif. Technol
.
58
(
1
),
91
98
.
Douziech
M.
,
van Zelm
R.
,
Oldenkamp
R.
,
Franco
A.
,
Hendriks
A.
,
King
H.
&
Huijbregts
M
, .
2018
Estimation of chemical emissions from down-the-drain consumer products using consumer survey data at a country and wastewater treatment plant level
.
Chemosphere
193
(
3–4
)
32
41
.
Etacheri
V.
,
Di Valentin
C.
,
Schneider
J.
,
Bahnemann
D.
&
Pillai
S
, .
2015
Visible-light activation of TiO2 photocatalysts: advances in theory and experiments
.
J. Photochem. Photobiol. C: Photochem. Rev.
25
,
1
29
.
Feng Yao
W.
,
Wang
H.
,
Hong Xu
X.
,
Feng Cheng
X.
,
Huang
J.
,
Xia Shang
S.
,
Na Yang
X.
&
Wang
M.
2003
Photocatalytic property of bismuth titanate Bi12TiO20 crystals
.
Appl. Catal. A Gen
.
243
(
1
),
185
190
.
Fiszka Borzyszkowska
A.
,
Pieczyńska
A.
,
Ofiarska
A.
,
Nikiforow
K.
,
Stepnowski
P.
&
Siedlecka
E.
2016
Bi-B-TiO2-based photocatalytic decomposition of cytostatic drugs under simulated sunlight treatments
.
Sep. Purif. Technol
.
169
,
113
120
.
Granberg
R.
&
Rasmuson
Å
, .
1999
Solubility of paracetamol in pure solvents
.
J. Chem. Eng. Data
44
(
6
),
1391
1395
.
Halpern
B.
,
Walbridge
S.
,
Selkoe
K.
,
Kappel
C.
,
Micheli
F.
,
D'Agrosa
C.
,
Bruno
J.
,
Casey
K.
,
Ebert
C.
,
Fox
H.
,
Fujita
R.
,
Heinemann
D.
,
Lenihan
H.
,
Madin
E.
,
Perry
M.
,
Selig
E.
,
Spalding
M.
,
Steneck
R.
&
Watson
R
, .
2008
A global map of human impact on marine ecosystems
.
Sci
.
319
(
5865
),
948
952
.
Huang
S.
,
Lin
Y.
,
Yang
J.
,
Li
X.
,
Zhang
J.
,
Yu
J.
,
Shi
H.
,
Wang
W.
&
Yu
Y
, .
2013
Enhanced photocatalytic activity and stability of semiconductor by Ag doping and simultaneous deposition: the case of CdS
.
RSC Adv
.
3
(
43
),
20782
.
Ibhadon
A.
&
Fitzpatrick
P
, .
2013
Heterogeneous photocatalysis: recent advances and applications
.
Catal
.
3
(
1
),
189
218
.
Ji
T.
,
Yang
F.
,
Lv
Y.
,
Zhou
J.
&
Sun
J
, .
2009
Synthesis and visible-light photocatalytic activity of Bi-doped TiO2 nanobelts
.
Mater. Lett
.
63
(
23
),
2044
2046
.
Jones
O.
,
Lester
J.
&
Voulvoulis
N
, .
2005
Pharmaceuticals: a threat to drinking water?
Trends in Biotechnol
.
23
(
4
),
163
167
.
Kim
S.
,
Hwang
S.
&
Choi
W
, .
2005
Visible light active platinum-ion-doped TiO2 photocatalyst
.
J. Phys. Chem. B
109
(
51
),
24260
24267
.
Landmann
M.
,
Rauls
E.
&
Schmidt
W
, .
2012
The electronic structure and optical response of rutile, anatase and brookite TiO2
.
J. Phys. Condens. Matter
.
24
(
19
),
195503
.
Liu
F.
,
Lu
L.
,
Xiao
P.
,
He
H.
,
Qiao
L.
&
Zhang
Y
, .
2012
Effect of oxygen vacancies on photocatalytic efficiency of TiO2 nanotubes aggregation
.
Bull. Korean Chem. Soc
.
33
(
7
),
2255
2259
.
Luengas
A.
,
Barona
A.
,
Hort
C.
,
Gallastegui
G.
,
Platel
V.
&
Elias
A
, .
2015
A review of indoor air treatment technologies
.
Rev. Env. Sci. Biotechnol.
14
(
3
),
499
522
.
Macwan
D.
,
Dave
P.
&
Chaturvedi
S
, .
2011
A review on nano-TiO2 sol–gel type syntheses and its applications
.
J. Mater. Sci
.
46
(
11
),
3669
3686
.
Mekki
A.
,
Holland
D.
,
McConville
C.
&
Salim
M
, .
1996
An XPS study of iron sodium silicate glass surfaces
.
J. Non-Cryst. Solids
.
208
(
3
),
267
276
.
Meng
S.
,
Li
D.
,
Sun
M.
,
Li
W.
,
Wang
J.
,
Chen
J.
,
Fu
X.
&
Xiao
G
, .
2011
Sonochemical synthesis, characterization and photocatalytic properties of a novel cube-shaped CaSn(OH)6
.
Catal. Commun
.
12
(
11
),
972
975
.
Milićević
B.
,
Đorđević
V.
,
Vuković
K.
,
Dražić
G.
&
Dramićanin
M.
2017
Effects of Li+ co-doping on properties of Eu3+ activated TiO2 anatase nanoparticles
.
Optical Materials
72
,
316
322
.
Ola
O.
&
Maroto-Valer
M
, .
2015
Review of material design and reactor engineering on TiO2 photocatalysis for CO2 reduction
.
J. Photochem. Photobiol. C Photochem. Rev.
24
,
16
42
.
Park
N.
,
van de Lagemaat
J.
&
Frank
A
, .
2000
Comparison of dye-sensitized rutile- and anatase-based TiO2 solar cells
.
J. Phys. Chem. B
104
(
38
),
8989
8994
.
Pelaez
M.
,
Nolan
N.
,
Pillai
S.
,
Seery
M.
,
Falaras
P.
,
Kontos
A.
,
Dunlop
P.
,
Hamilton
J.
,
Byrne
J.
,
O'Shea
K.
,
Entezari
M.
&
Dionysiou
D
, .
2012
A review on the visible light active titanium dioxide photocatalysts for environmental applications
.
Catal. B Environ.
125
,
331
349
.
Shimodaira
Y.
,
Kato
H.
,
Kobayashi
H.
&
Kudo
A
, .
2006
Photophysical properties and photocatalytic activities of bismuth molybdates under visible light irradiation
.
J. Phys. Chem. B
110
(
36
),
17790
17797
.
Walsh
A.
,
Yan
Y.
,
Huda
M.
,
Al-Jassim
M.
&
Wei
S
, .
2009
Band edge electronic structure of BiVO4: elucidating the role of the Bi s and V d orbitals
.
Chem. Mater
.
21
(
3
),
547
551
.
Wang
Y.
,
Zhang
L.
,
Deng
K.
,
Chen
X.
&
Zou
Z
, .
2007
Low temperature synthesis and photocatalytic activity of rutile TiO2 nanorod superstructures
.
J. Phys. Chem. C
111
(
6
),
2709
2714
.
Wu
S.
,
Zhang
L.
&
Chen
J
, .
2012
Paracetamol in the environment and its degradation by microorganisms
.
Appl. Microbiol. Biotechnol
.
96
(
4
),
875
884
.
Xu
N.
,
Liu
L.
,
Sun
X.
,
Liu
X.
,
Han
D.
,
Wang
Y.
,
Han
R.
,
Kang
J.
&
Yu
B
, .
2008
Characteristics and mechanism of conduction/set process in TiN∕ZnO∕Pt resistance switching random-access memories
.
Appl. Phys. Lett
.
92
(
23
),
232112
.
Yang
G.
,
Jiang
Z.
,
Shi
H.
,
Xiao
T.
&
Yan
Z
, .
2010
Preparation of highly visible-light active N-doped TiO2 photocatalyst
.
J. Mater. Chem.
20
(
25
),
5301
.
Yang
J.
,
Wang
X.
,
Dai
J.
&
Li
J
, .
2014
Efficient visible-light-driven photocatalytic degradation with Bi2O3 coupling silica doped TiO2
.
Ind. Eng. Chem. Res
.
53
(
32
),
12575
12586
.
Zhong
J.
,
Li
J.
,
Zeng
J.
,
He
X.
,
Huang
S.
,
Jiang
W.
&
Li
M
, .
2014
Enhanced photocatalytic activity of In2O3-decorated TiO2
.
Appl. Phys. A
115
(
4
),
1231
1238
.
This is an Open Access article distributed under the terms of the Creative Commons Attribution Licence (CC BY-NC-ND 4.0), which permits copying and redistribution for non-commercial purposes with no derivatives, provided the original work is properly cited (http://creativecommons.org/licenses/by-nc-nd/4.0/).