Next Article in Journal
Comparative Analyses of Complete Chloroplast Genomes and Karyotypes of Allotetraploid Iris koreana and Its Putative Diploid Parental Species (Iris Series Chinenses, Iridaceae)
Next Article in Special Issue
The Impact of Backbone Fluorination and Side-Chain Position in Thiophene-Benzothiadiazole-Based Hole-Transport Materials on the Performance and Stability of Perovskite Solar Cells
Previous Article in Journal
Unveiling Mycoviromes Using Fungal Transcriptomes
Previous Article in Special Issue
Asymmetric Non-Fullerene Small Molecule Acceptor with Unidirectional Non-Fused π-Bridge and Extended Terminal Group for High-Efficiency Organic Solar Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Nb/N Co-Doped Layered Perovskite Sr2TiO4: Preparation and Enhanced Photocatalytic Degradation Tetracycline under Visible Light

1
College of Materials Science and Engineering, North China University of Science and Technology, Tangshan 063210, China
2
Hebei Province Laboratory of Inorganic Nonmetallic Materials, North China University of Science and Technology, Tangshan 063210, China
3
Hebei (Tangshan) Ceramic Industry Technology Research Institute, Tangshan 063007, China
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(18), 10927; https://doi.org/10.3390/ijms231810927
Submission received: 18 August 2022 / Revised: 11 September 2022 / Accepted: 16 September 2022 / Published: 18 September 2022

Abstract

:
Sr2TiO4 is a promising photocatalyst for antibiotic degradation in wastewater. The photocatalytic performance of pristine Sr2TiO4 is limited to its wide bandgap, especially under visible light. Doping is an effective strategy to enhance photocatalytic performance. In this work, Nb/N co-doped layered perovskite Sr2TiO4 (Sr2TiO4:N,Nb) with varying percentages (0–5 at%) of Nb were synthesized by sol-gel and calcination. Nb/N co-doping slightly expanded the unit cell of Sr2TiO4. Their photocatalytic performance towards antibiotic (tetracycline) was studied under visible light (λ > 420 nm). When Nb/(Nb + Ti) was 2 at%, Sr2TiO4:N,Nb(2%) shows optimal photocatalytic performance with the 99% degradation after 60 min visible light irradiation, which is higher than pristine Sr2TiO4 (40%). The enhancement in photocatalytic performance is attributed to improving light absorption, and photo-generated charges separation derived from Nb/N co-doping. Sr2TiO4:N,Nb(2%) shows good stability after five cycles photocatalytic degradation reaction. The capture experiments confirm that superoxide radical is the leading active species during the photocatalytic degradation process. Therefore, the Nb/N co-doping in this work could be used as an efficient strategy for perovskite-type semiconductor to realize visible light driving for wastewater treatment.

Graphical Abstract

1. Introduction

Tetracycline (TC) is a widely used broad-spectrum antibiotic in medical treatment, animal husbandry, and aquaculture [1]. The accumulation of antibiotics in the environment leads to drug-resistant bacteria, which could bring serious threats to the ecological environment and human health [2,3,4,5]. Tetracycline is not only hard to self-degrade in the natural environment [3], but also difficult to be eliminated by conventional techniques.
Photocatalytic technology [6,7,8,9] utilizes a photocatalyst to generate light-generated holes and electrons under irradiation. Holes or electrons react with H2O or O2 to generate superoxide radicals and hydroxyl radicals with strong oxidation capacity, thereby thoroughly oxidizing and degrading organic pollutants [10] in wastewater. The core of photocatalytic oxidation technology are photocatalysts. Among many photocatalytic materials, strontium titanate has attracted wide attention because of its advantages, such as being inexpensive, pollution-free, and light corrosion resistant [11]. However, strontium titanate also has the disadvantages of having a large bandgap and high photo-generated carrier recombination rate. Layered perovskite Sr2TiO4 exhibits higher photocatalytic performance than perovskite SrTiO3, due to the particularly layered crystal structure and typical 2D charge transportation properties [12,13].
The wide bandgap of Sr2TiO4 exhibits photocatalytic activity only under ultraviolet light, which gravely affects further enhancement of photocatalytic performance. Ion doping can reduce the bandgap [14] of the perovskite material [11], including single metal doping, nonmetal doping, and co-doping. For instance, by introducing Cr [15], Ag [16], F [17], chalcogens [18], La/N [13], Cr/F [19], La/Rh [20], and La/Fe [12], the light response range of Sr2TiO4 is extended. Properties of Sr2TiO4 doping with different ions was shown in Table 1. Among them, co-doping has gained more interest, because co-doping is expected to maintain the charge-balance without forming oxygen vacancies.
Among various doping elements for oxide semiconductors, doping the nitrogen at the O site can narrow the band gap through the hybridization N 2p with the O 2p [21]. Meanwhile, N-doped oxide semiconductors are often accompanied by oxygen vacancies [22] due to charge compensation. Too high of a concentration of oxygen vacancies will act as a charge carrier recombination center, which is detrimental to photocatalytic activity [16,23,24]. The preparation methods of Sr2TiO4 mainly includes conventional solid-state reactions, the molten-salt method, and sol-gel method. Among them, the sol-gel synthesis process can achieve molecular-level doping.
This study chooses N-doped Sr2TiO4, where N3− replaces O2− to regulate the band structure. In order to control the concentration of oxygen vacancies, we replace Ti4+ (r = 0.061 nm) ions with Nb5+ (r = 0.064 nm) to balance the negative charge introduced by the unequal substitution of N3− and O2− Nb/N co-doped Sr2TiO4, which perhaps obtains charge-balanced Sr2TiO4:N,Nb and avoids too high of a concentration of oxygen vacancies. The amount of the dopant should also be optimized. Hence, we perform an investigation on Nb/N co-doped layered perovskite Sr2TiO4 for photocatalytic degradation tetracycline by varying Nb5+ doping content. To the authors’ knowledge, although there have been several studies on the doping modification of Sr2TiO4 [12,13,15,16,17,19,20] for photocatalytic hydrogen production, there is no literature report on the preparation and photocatalytic degradation performance of Nb/N co-doped Sr2TiO4. In this work, Sr2TiO4:N,Nb(2%) shows the best photocatalytic degradation performance towards tetracycline under visible light. The superior photocatalytic performance can be attributed to N-doping, narrowing the bandgap and Nb-doping compensating charge imbalance. This work affords a new insight to realizing visible-light-driven perovskite-type semiconductors.

2. Results and Discussion

2.1. Materials Characterization

As shown in Figure 1, the crystal structure of Sr2TiO4, Sr2TiO4:N and Sr2TiO4:N,Nb with different Nb-doping content were investigated by X-ray diffraction (XRD). There was only one phase of Sr2TiO4 (JCPDS NO:39-1471) without preferred growth orientation in the XRD patterns. Characteristic peaks of 2θ at 23.9, 28.3, 31.4, 32.6, 43.0, 43.7, 46.7, 55.0, 57.3, 65.4 and 68.2° were indexed to (101), (004), (103), (110), (006), (114), (200), (116), (213), (206) and (220) planes of Sr2TiO4 (JCPDS NO:39-1471), respectively.
After N-doping, the peak of (110) shown by Sr2TiO4:N in XRD shifted to a lower angle, meaning larger d-spacing than that of Sr2TiO4, and this larger spacing proved that N-doped Sr2TiO4 successfully. This is due to the N3− radius (0.146 nm) being slightly larger than the O2− radius (0.140 nm); when N3− enters the Sr2TiO4 crystal lattice, the crystal lattice expands. According to Bragg’s equation 2dsinθ = λ, with the increase of d-spacing, the diffraction angle 2θ shifts to a lower angle. Similar to Sr2TiO4:N, 2θ of Sr2TiO4:N,N shifted to a lower direction gradually with the increase of Nb-doping content, as shown in Figure 1. Because the Nb5+ radius (0.064 nm) and N3− radius (0.146 nm) are slightly larger than the Ti4+ radius (0.061 nm) and O2− radius (0.140 nm), respectively, the more the Nb-doping content, the more the lattice expansion [25,26], and the more diffraction angle offset. According to the XRD patterns, it can be concluded that both Nb5+ and N3− were doped into the lattice of Sr2TiO4 successfully. The average crystallite sizes were calculated by Scherrer’s equation [14] of the Sr2TiO4 (103) XRD reflection and were shown in Table 2.
The morphological characterization of Sr2TiO4 and Sr2TiO4:N,Nb(2%) were shown in Figure 2. After being calcined at 1000 °C for 4 h, both pristine Sr2TiO4 and Sr2TiO4:N,Nb(2%) were composed of irregular particles ranging from tens of nanometers to 1 micron.
TEM elemental mapping of Sr2TiO4 was performed and the results were shown in Figure 3a–d. It verified the existence of Sr, Ti, O, and these elements were distributed evenly. TEM elemental mapping of Sr2TiO4:N,Nb(2%) was also performed and the results were shown in Figure 3f–j. It confirmed the existence of Sr, Ti, O, Nb, and N, and these elements were evenly distributed as well.
Figure 4a shows the absorbance of Sr2TiO4, Sr2TiO4:N, and Sr2TiO4:N,Nb. The UV-Vis absorption spectra of the pristine and doped Sr2TiO4 showed a significant difference. The absorption edge of pristine Sr2TiO4 is located in the ultraviolet region. However, the absorption edge of N doping and Nb/N co-doping Sr2TiO4 were significantly red and shifted into the visible light region, and Sr2TiO4:N,Nb(2%) showed the strongest absorption capacity of visible light. According to the previous literature, N doping can reduce the band gap through hybridization of N 2p with the O 2p, raising the valence band maximum. The energy level of Nb5+ is similar to titanium 3d orbital energy, so Nb can replace Ti and mix Nb 4d with Ti 3d orbitals without lowering the conduction band minimum [27]. The same observation has been seen in other metal oxides doped with nitrogen, such as N-doped SrTiO3 [28], N-doped NaLaTiO4 [29], N-doped Sr2TiO4 [29], Nb/N co-doped TiO2 [27], and so on. The bandgap Eg deduced from the Tauc plots in Figure 4b were 3.48 eV, 3.02 eV, and 2.95 eV for pristine Sr2TiO4, Sr2TiO4:N, and Sr2TiO4:N,Nb(2%), respectively. Pristine Sr2TiO4 had a wide bandgap (3.48 eV), which was consistent with the previous literature [13]. N doping and Nb/N co-doping can reduce the bandgap of Sr2TiO4. Similar phenomena have also been observed in other materials involving N doping [13,24,30].
The chemical state and chemical composition on the surfaces of the pristine and doped Sr2TiO4 were investigated by XPS. The complete measurement scan in Figure 5a showed the presence of Sr, Ti, O, N (except pristine Sr2TiO4), and Nb (only for Sr2TiO4:N,Nb(2%)). Figure 5b shows the fine spectra of Ti. The presence of Ti4+ was validated by peaks at 458 and 464 eV, which were attributable to Ti 2p3/2 and Ti 2p1/2 [15,20]. The Ti 2p signal was gradually weakened by simultaneous Nb/N doping, confirming the substitution of Ti with Nb. A similar phenomenon was also observed in La/Fe co-doped Sr2TiO4 [12]. The peaks located around 529 eV and 531 eV belonged to the lattice oxygen and surface OH groups [31], respectively (Figure 5c). The signals of lattice oxygen gradually diminished along with N doping, meaning the substitution of lattice oxygen with N3 for Sr2TiO4:N and Sr2TiO4:N,Nb(2%), which was consistent with XRD results(Figure 1). Due to the strong signal of surface OH, it can be inferred that all samples are hydrophilic [13]. The peak around 398 eV was assigned to lattice N3− as shown in Figure 5d. Along with Nb/N co-doping, this signal of N 1s was enhanced, which demonstrated the introduction of Nb promoting N doping. The 3d5/2 and 3d5/2 of Nb 3d electrons were detected at binding energies of 207 eV and 209 eV [32], indicating the existence of Nb5+. It was worth noticing that binding-energy shifted towards lower binding energy after Nb/N co-doping, which suggests Nb5+ doping based on N3− doping was hole-doping. A similar shift was observed in SrCuO2 [33].
The charge carrier behaviors of the pristine Sr2TiO4 and Sr2TiO4:N,Nb with different Nb-doping content were examined by PL with an excitation wavelength of 325 nm at room temperature. As shown in Figure 6, the emission intensity of Nb/N co-doped Sr2TiO4 with different Nb doping content were lower than that of pristine Sr2TiO4, implying that the charge separation capability was enhanced significantly, especially for Sr2TiO4:N,Nb(2%). The negative charge (O vacancy) introduced by the unequal replacement of N3− and O2− can be balanced by Nb-doping. Therefore, Nb/N co-doping Sr2TiO4 can reduce the photoelectron-hole recombination probability, which was beneficial to the photocatalytic reaction. Nb5+, as a charge balancing agent, can compensate for a charge imbalance caused by N substitution of O, Nb Ti + N O , just as La played a charge balancing role in La/N co-doped Sr2TiO4 [13], La Sr + N O .

2.2. Photoelectrochemical Properties of Sr2TiO4:N,Nb

To further confirm the effect of Nb doping on photo-carriers separation and migration of Sr2TiO4:N,Nb(2%), photochemical measurements were performed. As shown in Figure 7a, the transient-state photocurrent densities of Sr2TiO4, Sr2TiO4:N, and Sr2TiO4:N,Nb(2%) remains stable under chopped light conditions for AM 1.5 illumination. An anodic photocurrent is clearly seen for all investigated samples, indicating that they are n-type semiconductors. The photocurrent density of Sr2TiO4, Sr2TiO4:N, and Sr2TiO4:N,Nb(2%) are 0.034 μA·cm−2, 0.199 μA·cm−2, and 0.274·cm−2 at 1.23 VRHE, respectively. The photocurrent density of Sr2TiO4:N,Nb(2%) was 8 times that of Sr2TiO4 and 1.37 times that of Sr2TiO4:N, implying the charge-separated significantly more efficiently by the introduction of Nb5+ based on N3−.
Figure 7b shows the electrochemical impedance spectroscopy (EIS) of Sr2TiO4, Sr2TiO4:N, and Sr2TiO4:N,Nb(2%) and the equivalent circuit, which can further verify the charge transport and transfer. It is well known that the smaller resistance radius indicates the higher separation and migration of electrons and holes. Compared with Sr2TiO4, Sr2TiO4:N, and Sr2TiO4:N,Nb(2%) had a smaller resistance radius, indicating more efficient separation of photocarriers. The result was consistent with the PL spectra and transient photocurrent response. The enhancement of the photocurrent density was attributed to the sufficient light absorption and effective photocarriers separation by introducing Nb5+ based on N3− doping.
The specific surface area was tested by an N2 adsorption-desorption isotherm at 77 K. As shown in Figure 8, both Sr2TiO4 and Sr2TiO4:N,Nb(2%) had typical IV isotherms and H3 hysteresis loops. H3 type hysteresis loops can be observed virtually on adsorbents with a lamellar structure. At high specific pressure, the curves showed an obvious hysteresis loop, which indicated that there was a mesoporous structure in the samples. The average pore diameter was determined using the Barrett–Joyner–Halenda (BJH) methods. BJT desorption average pore diameter (4 V/A) of Sr2TiO4 and Sr2TiO4:N,Nb(2%) were 38.24 nm and 23.22 nm, respectively. The specific surface areas of Sr2TiO4 and Sr2TiO4:N,Nb(2%) were 1.213 m2/g and 6.718 m2/g, respectively, which were very low. Specific surface area was not the main reason for the improvement of the photocatalytic degradation performance of Sr2TiO4:N,Nb(2%).

2.3. Photocatalytic Degradation of TC

As shown in Figure 9a, the adsorption capacity of photocatalysts towards TC after stirring for 30 min in the dark shows that Sr2TiO4:N,Nb(2%) exhibited the best adsorption property among them. However, the adsorption capacity was low, which was consistent with BET results. The photocatalytic properties of Sr2TiO4, Sr2TiO4:N, and Sr2TiO4:N,Nb(2%) were investigated by photocatalytic degradation of TC with a 1 g·L−1 photocatalyst under visible light. A blank experiment without a photocatalyst was also carried out to exclude the impact of photolysis. The self-degradation rate was 3.6% after 60 min due to negligible photooxygenation reactions in the presence of dissolved oxygen (Figure 9b) [34,35]. As shown in Figure 9b, the photocatalytic performance of all Nb/N co-doping Sr2TiO4 exhibited a higher degradation rate than prime Sr2TiO4 under the same conditions. The photocatalytic degradation rate of Sr2TiO4:N,Nb increased as Nb-doping content increased from 0% to 2%. When the content of Nb exceeded 2%, the degradation rate of Sr2TiO4:N,Nb decreased, which may be due to the introduction of new defects caused by excessive Nb, ig. Nb Ti + Ti Ti . Sr2TiO4:N,Nb(2%) exhibited the best photocatalytic degradation rate (99%), which was higher than that of prime Sr2TiO4 (40%) and Sr2TiO4:N (94%). On the one hand, Nb/N co-doping improved the light absorption in the visible-light region. On the other hand, the recombination probability of Sr2TiO4:N,Nb decreased, especially Sr2TiO4:N,Nb(2%). For comparison, the degradation rates of Sr2TiO4 towards different organic pollutants are shown in Table 3. To the authors’ knowledge, the photocatalytic degradation performance of N-doped Sr2TiO4 and Nb/N co-doped Sr2TiO4 has not been reported.
The photocatalytic degradation and fitting results of the TC kinetics followed pseudo-first-order kinetics, as shown in Figure 9c. Sr2TiO4:N,Nb(2%) showed a maximum apparent reaction rate constant of 0.0655 min−1 which was nearly 24.9 times that of Sr2TiO4 (0.00263 min−1) and just 1.25 times that of Sr2TiO4:N (0.0522 min−1). This indicated that N doping had the main effect whereas adding Nb yielded only a slight increase to photocatalytic degradation rates. The maximum apparent reaction rate constant obtained by Sr2TiO4:N,Nb(2%) was attributed to an appropriate Nb/N co-doping concentration.
To examine the stability and reusability of Sr2TiO4:N,Nb(2%), five cycles of photodegradation experiments were carried out. After each cycle, the photocatalyst was centrifuged (at 9000 rpm) and washed several times with deionized water for several times and re-placed in a deionized water solution containing fresh TC. As shown in Figure 9d, the degradation rate of Sr2TiO4:N,Nb(2%) from 99% to 90% after five repeated cycles without an obvious decrease, indicating that Sr2TiO4:N,Nb(2%) exhibited relatively good stability and repeatability. The decrease of degradation performance could be ascribed to the mass loss in the centrifugation and washing process. Sr2TiO4:N,Nb(2%) shows good potential in photocatalytic degradation of antibiotic pollutants.
To further understand the key active species in the photocatalytic process of TC degradation, active species trapping experiments were performed, as shown in Figure 10. Isopropyl alcohol (IPA, 0.01 M), triethanolamine (TEOA, 0.01 M), p-benzoquinone (BQ, 0.005 M), and AgNO3 (0.01M) were added as scavengers for hydroxyl radical ( OH ), hole (h+), superoxide radical ( O 2 ), and photogenerated electrons (e) to TC solution at the presence of Sr2TiO4:N,Nb(2%), respectively. When the IPA, TEOA, and AgNO3 was added, the photodegradation rate of TC only slightly decreased. However, when BQ was added, the degradation rate of TC reduced greatly, indicating that superoxide radical ( O 2 ) was the main active species during the photocatalytic degradation of TC.

2.4. Improvement Mechanism of Photocatalytic Performance

The effect of N doping and Nb/N co-doping on the valence bands of Sr2TiO4 was evaluated by XPS valence band (VB) spectra. As shown in Figure 11, valence band potential (EVB, XPS) measured by XPS valence band spectra of Sr2TiO4, Sr2TiO4:N, and Sr2TiO4:N,Nb(2%) were 2.86, 2.66, and 2.03 eV, respectively. EVB vs. SHE can be deduced according to the formula: EVB, NHE = φ + EVB, XPS − 4.5 (φ is the work function of the instrument: 4.59 eV, 4.5 eV vs. vacuum level is 0 V vs. SHE) [39,40,41,42]. Thus, EVB, NHE of Sr2TiO4, Sr2TiO4:N, and Sr2TiO4:N,Nb(2%) were calculated to be 2.95 eV, 2.75 eV, and 2.12 eV, respectively. The conduction band (VB) minimum of Sr2TiO4, Sr2TiO4:N, and Sr2TiO4:N,Nb(2%) are calculated to be −0.53 eV, −0.27 eV, and −0.83 eV(vs. SHE) according to these valance band and bandgap values (Figure 4b) mentioned above. The conduction band minimum of Sr2TiO4 reported in the previous literature varies widely and ranges from −0.254 eV to −0.87 eV [16,17,20,41], and the conduction band minimum of N-doped Sr2TiO4 and Nb/N co-doped Sr2TiO4 have not been reported.
According to our experimental results of XPS valence band spectra (Figure 11) and UV-Vis spectra (Figure 4b), schematic band structures of Sr2TiO4, Sr2TiO4:N, and Sr2TiO4:N,Nb(2%) were illustrated in Figure 12. The pure Sr2TiO4 had a wide bandgap. In contrast, Sr2TiO4:N displayed visible light response derived from N orbital. The state of the oxygen vacancies (accompanied by N-doping) was located below the conduction band minimum, which was consistent with previous literature about SrTiO3 co-doped with N and La [21]. Regarding Sr2TiO4:N,Nb(2%), Nb/N co-doping uplifted the valence band furthermore. As we know, the more negative the valance band, the higher reduction ability of photo-generated electrons to generate superoxide radicals ( O 2 ). The conduction band potential of Sr2TiO4:N,Nb(2%) was more negative than Sr2TiO4, Sr2TiO4:N, and O 2 / O 2 (−0.28 eV), so electrons can more easily transfer to the oxygen molecules, producing O 2 . The uplift of conduction band after cationic doping had also been observed in Ag doping Sr2TiO4 [16].
Hydroxyl radical ( OH ) may be difficult to produce by the photogenerated holes, directly, because the valence band position (+2.12 eV) in Sr2TiO4:N,Nb(2%) was higher than that of OH / H 2 O (+2.27 eV) and slightly more positive than that of OH / OH (1.99 eV). Therefore, O 2 is the predominant active species in the degradation of TC, which coincided with the previous result (Figure 10). A similar situation was also discovered in Bi2WO6 [43,44].

3. Materials and Methods

3.1. Chemicals

The chemicals in this work were obtained from Shanghai Maclin Biochemical Technology Co., Ltd. (Shanghai, China) and Tianjin Yongda Chemical Reagent Co., Ltd. (Tianjin, China), without any further purification.

3.2. Materials Synthesis

3.2.1. Synthesis of Sr2TiO4

Firstly, 0.01 mol tetrabutyl titanate was dissolved in 50 mL of absolute ethanol, and 0.07 mol citric acid was injected as a complex agent (solution A). Secondly, 0.02 mol Sr(NO3)2 was dissolved in deionized water (solution B). Solution B was added dropwise to solution A to gain a transparent mixed solution. The mixed solution was heated and stirred in a water bath at 60 °C to form a homogeneous transparent sol. After aging for 24 h, the sol was dried at 120 °C for several hours to form a dry gel. Finally, the obtained gel was ground and calcined at 1000 °C for 4 h (heating rate 10 °C·min−1) to obtain Sr2TiO4 powder.

3.2.2. Synthesis of N-Doped Sr2TiO4

The preparation process of N-doped Sr2TiO4 was as follows: After mixing urea and Sr2TiO4 uniformly (m(urea):m(Sr2TiO4) = 1.5:1), the mixture was calcinated at 400 °C for 2 h, cooled naturally, and grinded. N-doped Sr2TiO4 was labeled as Sr2TiO4:N.

3.2.3. Synthesis of Nb/N Co-Doped Sr2TiO4

The preparation process of Nb/N co-doped Sr2TiO4 was as follows: Nb-doped Sr2TiO4 was prepared first, which was the same as that of Sr2TiO4, except for the addition of NbCl5 in solution B (Nb/(Nb + Ti), which was 1 at%, 2 at%, 3 at%, 4 at%, 5 at%. Nb/N co-doped Sr2TiO4 was prepared the same as that of N-doped Sr2TiO4, except that urea was mixed with Nb-doped Sr2TiO4. Nb/N co-doped Sr2TiO4 with different Nb doping content were denoted as Sr2TiO4:N,Nb(1%), Sr2TiO4:N,Nb(2%), Sr2TiO4:N,Nb(3%), Sr2TiO4:N,Nb(4%), and Sr2TiO4:N,Nb(5%), respectively.

3.3. Characterization

The crystal structure analysis was carried out by a Japanese D/MAX2500PC X-ray diffractometer employing Cu Kα radiation; the surface morphology was investigated by a Japan JEM-2010 transmission electron microscope (TEM); a specific surface and pore size analyzer (3H-2000PM1, BeiShiDe Instrument Technology Co., Ltd., Beijing, China) was used to calculate the specific surface area (BET) and pore size of the sample; the optical properties were investigated by an Ultraviolet-Visible spectrophotometer (Lambda750, PerkinElmer, Norwalk, CT, USA); the photoluminescence (PL, SR830) with the excitation at 325 nm was characterized using a fluorescence spectrometer; the chemical composition and valence band spectrum were detected by X-ray photoelectron spectroscopy (Thermo Scientific, Waltham, MA, USA, ESCALAB 250XI).

3.4. Photoelectrochemical Measurements

Photocurrent density-time (I-t) and electrochemical impedance spectroscopy (EIS) were characterized on an electrochemical workstation (CHI660D, Shanghai Chenhua Instrument, Shanghai, China) through a three-electrode system. Three-electrode system was set up using 0.5 M Na2SO4 solution, Ag/AgCl as reference electrode, Pt foil as counter electrode, and photoelectrodes fabricated by these sample powders on FTO glass as working electrodes. During the measurements of photocurrent density, simulated sunlight was irradiated with an intensity of 100 mW·cm−2 (AM1.5G filter) under chopped light at a bias of 1.23 V vs. RHE. The EIS was measured with a frequency from 100 kHz to 0.01 Hz with an AC amplitude of 5 mV.

3.5. Photocatalytic Degradation of TC

The photocatalytic performances of the samples were assessed by photocatalytic degradation towards TC. In total, 50 mg Sr2TiO4 or 50 mg Sr2TiO4:N,Nb was dispersed in 50 mL of TC solution at a concentration of 20 mg·L−1. Before irradiation, the mixtures were continuously stirred in the dark for 30 min to reach adsorption-desorption equilibrium. Then, a photocatalytic reaction was performed under visible light using a 5 W LED lamp (white light, λ > 420 nm, 250 mW/cm2). Extract 3 mL of TC solution every 10 min and centrifuge at 9000 rpm for 5 min. The TC degradation rate was calculated as follows:
Degradation   rate = C 0 C t C 0 × 100 % = A 0 A t A 0 × 100 %
where C0 (A0) and Ct (At) represent the TC’s concentration (absorbance at 357 nm) of initial and after min irradiation, respectively.
In order to ascertain the active species during the photocatalytic degradation process, the active species capture experiment was carried out, similar to the photocatalytic degradation of TC experiments, except that TC solution was replaced by TC and scavenger.

4. Conclusions

In summary, Nb/N co-doped layered perovskite Sr2TiO4 (Sr2TiO4:N,Nb) with varying percentages (0–5 at%) of Nb were successfully synthesized by sol-gel and calcination. Nb/N could co-dope into Sr2TiO4 with a slight unit cell expansion, which was confirmed by X-ray diffraction. Nb/N co-doped Sr2TiO4 showed better photocatalytic performance for tetracycline photocatalytic degradation than pristine Sr2TiO4 under visible light, especially Sr2TiO4:N,Nb(2%). Sr2TiO4:N,Nb(2%) showed optimal photocatalytic performance with the 99% degradation after 60 min visible light irradiation, which was higher than pristine Sr2TiO4 (40%). Nb/N co-doping enhanced the photocatalytic activity by broadening light response to the visible region and reducing the photogenerated carrier recombination. Nb/N co-doping had a slight effect on morphology and the average grain size of Sr2TiO4. Sr2TiO4:N,Nb(2%) showed good stability and recyclability after five cycles photocatalytic degradation reaction. The superoxide radical ( O 2 ) was the leading contributor to tetracycline degradation. Nb/N co-doping strategy in this work affords insight to realizing visible-light-driven perovskite-type semiconductors for wastewater treatment.

Author Contributions

Conceptualization, J.W. and Y.Z.; methodology, Y.Z. and X.Z.; validation, P.L.; data curation, J.W. and P.L. writing—original draft preparation, J.W. and P.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Natural Science Foundation of Hebei Province (No. E2021209002), and the Tangshan Science and Technology Bureau (No. 21130211D).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mohammad, A.; Khan, M.E.; Cho, M.H.; Yoon, T. Adsorption promoted visible-light-induced photocatalytic degradation of antibiotic tetracycline by tin oxide/cerium oxide nanocomposite. Appl. Surf. Sci. 2021, 565, 150337. [Google Scholar] [CrossRef]
  2. Watkinson, A.J.; Murby, E.J.; Costanzo, S.D. Removal of antibiotics in conventional and advanced wastewater treatment: Implications for environmental discharge and wastewater recycling. Water Res. 2007, 41, 4164–4176. [Google Scholar] [CrossRef] [PubMed]
  3. Huang, L.S.; Bao, D.Y.; Li, J.H.; Jiang, X.Q.; Sun, X.S. Construction of Au modified direct Z-scheme g-C3N4/defective ZnO heterostructure with stable high-performance for tetracycline degradation. Appl. Surf. Sci. 2021, 555, 149696. [Google Scholar] [CrossRef]
  4. Li, Y.H.; Lai, Z.; Huang, Z.J.; Wang, H.Y.; Zhao, C.X.; Ruan, G.H.; Du, F.Y. Fabrication of BiOBr/MoS2/graphene oxide composites for efficient adsorption and photocatalytic removal of tetracycline antibiotics. Appl. Surf. Sci. 2021, 550, 149342. [Google Scholar] [CrossRef]
  5. Guo, Z.W.; Zheng, J.; Li, B.R.; Da, Z.L.; Meng, M.J. Fabrication of mixed matrix membranes blending with the TiO2/Bi3O4Cl 2D/2D heterojunction for photocatalytic degradation of tetracycline. Appl. Surf. Sci. 2022, 574, 151549. [Google Scholar] [CrossRef]
  6. Dong, X.; Xu, J.; Kong, C.; Zeng, X.; Wang, J.; Zhao, Y.; Zhang, W. Synthesis of β-FeOOH/TiO2/SiO2 by melting phase separation-hydrothermal method to improve photocatalytic performance. Ceram. Int. 2021, 47, 32303–32309. [Google Scholar] [CrossRef]
  7. Xu, M.; Yang, J.; Sun, C.; Cui, Y.; Liu, L.; Zhao, H.; Liang, B. Facile assembly of BiVO4/protonated g-C3N4/AgI with a novel dual Z-scheme mechanism for visible-light photocatalytic degradation of Rhodamine B. J. Mater. Sci. 2021, 56, 1328–1346. [Google Scholar] [CrossRef]
  8. Bae, H.S.; Manikandan, V.; Hwang, J.H.; Seo, Y.S.; Chung, H.S.; Ryu, H.I.; Chae, W.S.; Cho, M.; Ekambe, P.S.; Jang, J.S. Photocatalytic degradation of organic pollutants and inactivation of pathogens under visible light via CoOx surface-modified Rh/Sb-doped SrTiO3 nanocube. J. Mater. Sci. 2021, 56, 17235–17253. [Google Scholar] [CrossRef]
  9. Fan, Q.; Wang, T.; Fan, W.; Xu, L. Recyclable visible-light photocatalytic composite materials based on tubular Au/TiO2/SiO2 ternary nanocomposites for removal of organic pollutants from water. Compos. Commun. 2022, 32, 101154. [Google Scholar] [CrossRef]
  10. Liu, Q.; Fan, Z.; Yi, X.; Chen, S.; Li, B.; Luo, W. Porous polyimide/carbon quantum dots/ZnS quantum dots material aerogel for efficient visible-light photocatalytic degradation over oxytetracycline. React. Funct. Polym. 2022, 178, 105330. [Google Scholar] [CrossRef]
  11. Irshad, M.; Ain, Q.; Zaman, M.; Aslam, M.Z.; Kousar, N.; Asim, M.; Rafique, M.; Siraj, K.; Tabish, A.N.; Usman, M.; et al. Photocatalysis and perovskite oxide-based materials: A remedy for a clean and sustainable future. RSC Adv. 2022, 12, 7009–7039. [Google Scholar] [CrossRef] [PubMed]
  12. Zhang, H.; Ni, S.; Mi, Y.; Xu, X. Ruddlesden-Popper compound Sr2TiO4 co-doped with La and Fe for efficient photocatalytic hydrogen production. J. Catal. 2018, 359, 112–121. [Google Scholar] [CrossRef]
  13. Sun, X.; Mi, Y.; Jiao, F.; Xu, X. Activating layered perovskite compound Sr2TiO4 via La/N codoping for visible light photocatalytic water splitting. ACS Catal. 2018, 8, 3209–3221. [Google Scholar] [CrossRef]
  14. Mouchaal, Y.; Enesca, A.; Mihoreanu, C.; Khelil, A.; Duta, A. Tuning the opto-electrical properties of SnO2 thin films by Ag+1 and In+3 co-doping. Mater. Sci. Eng. B 2015, 199, 22. [Google Scholar] [CrossRef]
  15. Sun, X.; Xie, Y.; Wu, F.; Chen, H.; Lv, M.; Ni, S.; Liu, G.; Xu, X. Photocatalytic hydrogen production over chromium doped layered perovskite Sr2TiO4. Inorg. Chem. 2015, 54, 7445–7453. [Google Scholar] [CrossRef]
  16. Xiao, H.B.; Liu, P.Y.; Wang, W.; Ran, R.; Zhou, W.; Shao, Z.P. Enhancing the photocatalytic activity of Ruddlesden-Popper Sr2TiO4 for hydrogen evolution through synergistic silver doping and moderate reducing pretreatment. Mater. Today Energy 2022, 23, 100899. [Google Scholar] [CrossRef]
  17. Han, X.; Liu, P.Y.; Ran, R.; Wang, W.; Zhou, W.; Shao, Z.P. Non-metal fluorine doping in Ruddlesden-Popper perovskite oxide enables high-efficiency photocatalytic water splitting for hydrogen production. Mater. Today Energy 2022, 23, 100896. [Google Scholar] [CrossRef]
  18. Ziati, M.; Bekkioui, N.; Ez-Zahraouy, H. Ruddlesden-Popper compound Sr2TiO4 doped with chalcogens for optoelectronic applications: Insights from first-principle calculations. Chem. Phys. 2021, 548, 111221. [Google Scholar] [CrossRef]
  19. Yu, J.X.; Xu, X.X. Fluorination over Cr doped layered perovskite Sr2TiO4 for efficient photocatalytic hydrogen production under visible light illumination. J. Energy Chem. 2020, 51, 30–38. [Google Scholar] [CrossRef]
  20. Sun, X.; Xu, X. Efficient photocatalytic hydrogen production over La/Rh co-doped Ruddlesden-Popper compound Sr2TiO4. Appl. Catal. B-Environ. 2017, 210, 149–159. [Google Scholar] [CrossRef]
  21. Miyauchi, M.; Takashio, M.; Tobimatsu, H. Photocatalytic activity of SrTiO3 codoped with nitrogen and lanthanum under visible light illumination. Langmuir 2004, 20, 232–236. [Google Scholar] [CrossRef] [PubMed]
  22. Rumaiz, A.K.; Woicik, J.C.; Cockayne, E.; Lin, H.Y.; Jaffari, G.H.; Shah, S.I. Oxygen vacancies in N doped anatase TiO2: Experiment and first-principles calculations. Appl. Phys. Lett. 2009, 95, 262111. [Google Scholar] [CrossRef]
  23. Corby, S.; Francàs, L.; Kafizas, A.; Durrant, J.R. Determining the role of oxygen vacancies in the photoelectrocatalytic performance of WO3 for water oxidation. Chem. Sci. 2020, 11, 2907–2914. [Google Scholar] [CrossRef] [PubMed]
  24. Irie, H.; Watanabe, Y.; Hashimoto, K. Nitrogen-concentration dependence on photocatalytic activity of TiO2-xNx powders. J. Phys. Chem. B 2003, 107, 5483–5486. [Google Scholar] [CrossRef]
  25. Irshad, M.; Idrees, R.; Siraj, K.; Shakir, I.; Rafique, M.; Ain, Q.; Raza, R. Electrochemical evaluation of mixed ionic electronic perovskite cathode LaNi1-xCoxO3-δ for IT-SOFC synthesized by high temperature decomposition. Int. J. Hydrogen Energy 2021, 46, 10448–10456. [Google Scholar] [CrossRef]
  26. Irshad, M.; Ain, Q.; Siraj, K.; Raza, R.; Tabish, A.N.; Rafique, M.; Idrees, R.; Khan, F.; Majeed, S.; Ahsan, M. Evaluation of BaZr0.8X0.2 (X= Y, Gd, Sm) proton conducting electrolytes sintered at low temperature for IT-SOFC synthesized by cost effective combustion method. J. Alloys Compd. 2020, 815, 152389. [Google Scholar] [CrossRef]
  27. Cottineau, T.; Béalu, N.; Gross, P.A.; Pronkin, S.N.; Keller, N.; Savinova, E.R.; Keller, V. One step synthesis of niobium doped titania nanotube arrays to form (N,Nb) co-doped TiO2 with high visible light photoelectrochemical activity. J. Mater. Chem. A 2013, 1, 2151–2160. [Google Scholar] [CrossRef]
  28. Zou, F.; Jiang, Z.; Qin, X.; Zhao, Y.; Jiang, L.; Zhi, J.; Xiao, T.; Edwards, P.P. Template-free synthesis of mesoporous N-doped SrTiO3 perovskite with high visible-light-driven photocatalytic activity. Chem. Commun. 2012, 48, 8514–8516. [Google Scholar] [CrossRef]
  29. Xu, X.; Wang, R.; Sun, X.; Lv, M.; Ni, S. Layered perovskite compound NaLaTiO4 modified by nitrogen doping as a visible light active photocatalyst for water splitting. ACS Catal. 2020, 10, 9889–9898. [Google Scholar] [CrossRef]
  30. Atkinson, I.; Parvulescu, V.; Pandele Cusu, J.; Anghel, E.M.; Voicescu, M.; Culita, D.; Somacescu, S.; Munteanu, C.; Šćepanović, M.; Popovic, Z.V.; et al. Influence of preparation method and nitrogen (N) doping on properties and photo-catalytic activity of mesoporous SrTiO3. J. Photoch. Photobiol. A 2019, 368, 41–51. [Google Scholar] [CrossRef]
  31. Sun, X.; Wang, S.; Shen, C.; Xu, X. Efficient photocatalytic hydrogen production over Rh-doped inverse spinel Zn2TiO4. ChemCatChem 2016, 8, 2289–2295. [Google Scholar] [CrossRef]
  32. Nachaithong, T.; Moontragoon, P.; Chanlek, N.; Thongbai, P. Fe3+/Nb5+co-doped rutile–TiO2 nanocrystalline powders prepared by a combustion process: Preparation and characterization and their giant dielectric response. RSC Adv. 2020, 10, 24784–24794. [Google Scholar] [CrossRef] [PubMed]
  33. Terada, N.; Iyo, A.; Sekita, Y.; Ishibashi, S.; Ihara, H. In-situ characterization of doping-effect on electronic structure of epitaxial films of infinite layer SrCuO2. Czech. J. Phys. 1996, 46, 2683–2684. [Google Scholar] [CrossRef]
  34. Salma, A.; Thoroeboveleth, S.; Schmidt, T.C.; Tuerk, J. Dependence of transformation product formation on pH during photolytic and photocatalytic degradation of ciprofloxacin. J. Hazard. Mater. 2016, 313, 49–59. [Google Scholar] [CrossRef] [PubMed]
  35. Jia, J.; Liu, D.; Tian, J.; Wang, W.; Ni, J.; Wang, X. Visible-light-excited humic acid for peroxymonosulfate activation to degrade bisphenol A. Chem. Eng. J. 2020, 400, 125853. [Google Scholar] [CrossRef]
  36. Sorkh-Kaman-Zadeh, A.; Dashtbozorg, A. Facile chemical synthesis of nanosize structure of Sr2TiO4 for degradation of toxic dyes from aqueous solution. J. Mol. Liq. 2016, 223, 921–926. [Google Scholar] [CrossRef]
  37. Hu, C.; Chen, T.S.; Huang, H.X. Heterojunction of n-type Sr2TiO4 with p-type Bi5O7I with enhanced photocatalytic activity under irradiation of simulated sunlight. Appl. Surf. Sci. 2017, 426, 536–544. [Google Scholar] [CrossRef]
  38. Zhang, W.J.; Sun, X.; Yang, B. Sol-gel preparation of La doped strontium titanate. Asian J. Chem. 2013, 25, 3263–3266. [Google Scholar] [CrossRef]
  39. Li, X.; Kang, B.; Dong, F.; Zhang, Z.; Luo, X.; Han, L.; Huang, J.; Feng, Z.; Chen, Z.; Xu, J.; et al. Enhanced photocatalytic degradation and H2/H2O2 production performance of S-pCN/WO2.72 S-scheme heterojunction with appropriate surface oxygen vacancies. Nano Energy 2021, 81, 105671. [Google Scholar] [CrossRef]
  40. Yu, H.; Shi, R.; Zhao, Y.; Bian, T.; Zhao, Y.; Zhou, C.; Waterhouse, G.I.N.; Wu, L.Z.; Tung, C.H.; Zhang, T. Alkali-assisted synthesis of nitrogen deficient graphitic carbon nitride with tunable band structures for efficient visible-light-driven hydrogen evolution. Adv. Mater. 2017, 29, 1605148. [Google Scholar] [CrossRef]
  41. Kwak, B.S.; Do, J.Y.; Park, N.K.; Kang, M. Surface modification of layered perovskite Sr2TiO4 for improved CO2 photoreduction with H2O to CH4. Sci. Rep. 2017, 7, 16370. [Google Scholar] [CrossRef] [PubMed]
  42. Huang, Y.; Wang, Y.; Bi, Y.; Jin, J.; Ehsan, M.F.; Fu, M.; He, T. Preparation of 2D hydroxyl-rich carbon nitride nanosheets for photocatalytic reduction of CO2. RSC Adv. 2015, 5, 33254–33261. [Google Scholar] [CrossRef]
  43. Huang, Y.; Kang, S.; Yang, Y.; Qin, H.; Ni, Z.; Yang, S.; Li, X. Facile synthesis of Bi/Bi2WO6 nanocomposite with enhanced photocatalytic activity under visible light. Appl. Catal. B-Environ. 2016, 196, 89–99. [Google Scholar] [CrossRef]
  44. Fu, H.; Pan, C.; Yao, W.; Zhu, Y. Visible-light-induced degradation of rhodamine B by nanosized Bi2WO6. J. Phys. Chem. B 2005, 109, 22432–22439. [Google Scholar] [CrossRef] [PubMed]
Figure 1. XRD patterns and partial enlarged view of Sr2TiO4, Sr2TiO4:N, and Sr2TiO4:N,Nb.
Figure 1. XRD patterns and partial enlarged view of Sr2TiO4, Sr2TiO4:N, and Sr2TiO4:N,Nb.
Ijms 23 10927 g001
Figure 2. TEM images of (a) Sr2TiO4; (b) Sr2TiO4:N,Nb(2%).
Figure 2. TEM images of (a) Sr2TiO4; (b) Sr2TiO4:N,Nb(2%).
Ijms 23 10927 g002
Figure 3. (a) TEM image of Sr2TiO4; (bd) Elemental mapping of Sr, Ti, and O; (e) TEM image of Sr2TiO4:N,Nb(2%); (fj) Elemental mapping of Sr, Ti, O, Nb, and N.
Figure 3. (a) TEM image of Sr2TiO4; (bd) Elemental mapping of Sr, Ti, and O; (e) TEM image of Sr2TiO4:N,Nb(2%); (fj) Elemental mapping of Sr, Ti, O, Nb, and N.
Ijms 23 10927 g003
Figure 4. UV-Vis spectra (a) and Tauc plots (b) of Sr2TiO4, Sr2TiO4:N, and Sr2TiO4:N,Nb(2%).
Figure 4. UV-Vis spectra (a) and Tauc plots (b) of Sr2TiO4, Sr2TiO4:N, and Sr2TiO4:N,Nb(2%).
Ijms 23 10927 g004
Figure 5. XPS spectra of the pristine and doped Sr2TiO4: (a) full survey spectra, the peak circled in red belongs to Nb 3d; (b) Ti 2p; (c) O 1s; (d) N 1s; (e) Nb 3d.
Figure 5. XPS spectra of the pristine and doped Sr2TiO4: (a) full survey spectra, the peak circled in red belongs to Nb 3d; (b) Ti 2p; (c) O 1s; (d) N 1s; (e) Nb 3d.
Ijms 23 10927 g005
Figure 6. Photoluminescence (PL) spectra of Sr2TiO4 and Sr2TiO4:N,Nb.
Figure 6. Photoluminescence (PL) spectra of Sr2TiO4 and Sr2TiO4:N,Nb.
Ijms 23 10927 g006
Figure 7. Photocurrent response (a) and EIS diagram (b) of Sr2TiO4, Sr2TiO4:N, and Sr2TiO4:N,Nb(2%), inset shows the equivalent circuit.
Figure 7. Photocurrent response (a) and EIS diagram (b) of Sr2TiO4, Sr2TiO4:N, and Sr2TiO4:N,Nb(2%), inset shows the equivalent circuit.
Ijms 23 10927 g007
Figure 8. N2 adsorption/desorption isotherms Sr2TiO4 (a) and Sr2TiO4:N,Nb(2%) (b).
Figure 8. N2 adsorption/desorption isotherms Sr2TiO4 (a) and Sr2TiO4:N,Nb(2%) (b).
Ijms 23 10927 g008
Figure 9. (a) Adsorption of different photocatalysts towards TC; (b) Self-degradation rate and degradation rate with different photocatalysts of tetracycline under visible light; (c) pseudo-first order kinetics curves of the photocatalytic degradation of TC; (d) degradation rate of TC after repeated cycles by Sr2TiO4:N,Nb(2%).
Figure 9. (a) Adsorption of different photocatalysts towards TC; (b) Self-degradation rate and degradation rate with different photocatalysts of tetracycline under visible light; (c) pseudo-first order kinetics curves of the photocatalytic degradation of TC; (d) degradation rate of TC after repeated cycles by Sr2TiO4:N,Nb(2%).
Ijms 23 10927 g009
Figure 10. Photocatalytic degradation rate of TC for Sr2TiO4:N,Nb(2%) in the presence of different scavengers.
Figure 10. Photocatalytic degradation rate of TC for Sr2TiO4:N,Nb(2%) in the presence of different scavengers.
Ijms 23 10927 g010
Figure 11. XPS valence band spectra of Sr2TiO4, Sr2TiO4:N, and Sr2TiO4:N,Nb(2%).
Figure 11. XPS valence band spectra of Sr2TiO4, Sr2TiO4:N, and Sr2TiO4:N,Nb(2%).
Ijms 23 10927 g011
Figure 12. Schematic band structures of Sr2TiO4, Sr2TiO4:N, and Sr2TiO4:N,Nb(2%). The potential of O 2 / O 2 , OH / OH , and OH / H 2 O are shown in the diagram.
Figure 12. Schematic band structures of Sr2TiO4, Sr2TiO4:N, and Sr2TiO4:N,Nb(2%). The potential of O 2 / O 2 , OH / OH , and OH / H 2 O are shown in the diagram.
Ijms 23 10927 g012
Table 1. Properties of Sr2TiO4 doping with differentiation.
Table 1. Properties of Sr2TiO4 doping with differentiation.
Doping ElementsA-Site/B-Site or O-Site DopingDoping ContentEg (eV)References
CrB5 at%~1.5[15]
AgA2.5 at%3.05[16]
FO3 at%3.20[17]
Chalcogens (S, Se, Te)O--0~2.99[18]
La/NA/O10 at%/-2.2[13]
Cr/FB/O5 at%/40 at %~2.5[19]
La/RhA/B1.5 at%/3 at %2.43[20]
La/FeA/B1.5 at%/3 at %~2.75[12]
Table 2. Average crystallite size of Sr2TiO4, Sr2TiO4:N, and Sr2TiO4:N,Nb.
Table 2. Average crystallite size of Sr2TiO4, Sr2TiO4:N, and Sr2TiO4:N,Nb.
Sr2TiO4Sr2TiO4:NSr2TiO4:N,Nb(1%)Sr2TiO4:N,Nb(2%)Sr2TiO4:N,Nb(3%)Sr2TiO4:N,Nb(4%)Sr2TiO4:N,Nb(5%)
Average crystallite size43 nm65 nm64 nm55 nm68 nm71 nm61 nm
Table 3. Photocatalytic properties of Sr2TiO4.
Table 3. Photocatalytic properties of Sr2TiO4.
SamplesDoping
Elements
Organic PollutantsDegradation RatesIrradiation SystemReferences
Sr2TiO4——methylene orange32% (120 min)UV lamp (400 W Osram lamps)[36]
Sr2TiO4——methylene orange3.2% (180 min)Xenon lamp (300 W without filter)[37]
Sr2TiO4——methylene orange78% (120 min)UV lamp (λ = 253.7 nm, 2200 mW·cm−2)[38]
Sr2TiO4Lamethylene orange90.5% (120 min)UV lamp (λ = 253.7 nm, 2200 mW·cm−2)[38]
Sr2TiO4——tetracycline40% (60 min)LED lamp (λ > 420 nm)This work
Sr2TiO4Nb/Ntetracycline99% (60 min)LED lamp (λ > 420 nm)This work
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wang, J.; Li, P.; Zhao, Y.; Zeng, X. Nb/N Co-Doped Layered Perovskite Sr2TiO4: Preparation and Enhanced Photocatalytic Degradation Tetracycline under Visible Light. Int. J. Mol. Sci. 2022, 23, 10927. https://doi.org/10.3390/ijms231810927

AMA Style

Wang J, Li P, Zhao Y, Zeng X. Nb/N Co-Doped Layered Perovskite Sr2TiO4: Preparation and Enhanced Photocatalytic Degradation Tetracycline under Visible Light. International Journal of Molecular Sciences. 2022; 23(18):10927. https://doi.org/10.3390/ijms231810927

Chicago/Turabian Style

Wang, Jiansheng, Pengwei Li, Yingna Zhao, and Xiongfeng Zeng. 2022. "Nb/N Co-Doped Layered Perovskite Sr2TiO4: Preparation and Enhanced Photocatalytic Degradation Tetracycline under Visible Light" International Journal of Molecular Sciences 23, no. 18: 10927. https://doi.org/10.3390/ijms231810927

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